id
stringlengths
27
33
source
stringclasses
1 value
format
stringclasses
1 value
text
stringlengths
13
1.81M
warning/0001/quant-ph0001093.html
ar5iv
text
# Consistent Quantum Realism: A Reply to Bassi and Ghirardi ## 1 Introduction The first paper on the consistent histories (CH) interpretation of quantum theory was published in the Journal of Statistical Physics in 1984 . In the years since then this approach, sometimes called “decoherent histories”, has been refined and extended in several books and papers, of which some of the more significant are . It provides a realistic picture of the atomic realm without the need to invoke quantum measurement as a fundamental principle, and for this reason it can resolve the “measurement problem” (there are actually two measurement problems, see ) which has long beset attempts to place the foundations of quantum theory on a sound basis, and which probably cannot be dealt with consistently by traditional methods . Because it resolves various quantum paradoxes using an analysis based upon the mathematics of Hilbert space, the CH approach removes any need to look for alternatives to standard quantum theory, such as those found in the hidden-variables approach of de Broglie and Bohm , or in the spontaneous reduction ideas of Ghirardi, Rimini, Weber, and Pearle . There have, to be sure, been a number of criticisms of the CH approach, and these have proven helpful in constructing improved versions of the formalism, and better expositions of its physical interpretation. The most significant of these criticisms are discussed in , and reasons are given why they do not invalidate CH quantum theory. This reference provides the material needed to counter various claims, such as in , that the CH approach is logically inconsistent or unsound. In this connection it is worth pointing out that even one of its severest critics has admitted that the CH approach is logically consistent when its rules are properly followed . One of these rules, known as the single framework (or single family, single logic, or single set) rule, plays a central role in the CH approach, as has been repeatedly emphasized in various publications . Despite the extensive discussion of this rule in the CH literature, accompanied by numerous applications to specific problems, it is still sometimes misunderstood, as in some recent work by Bassi and Ghirardi . In particular, these authors have claimed, in an article appearing in this Journal, that the CH interpretation of quantum theory when interpreted in a realistic way using some reasonable assumptions leads to contradictions in the sense of violating a result of Bell , and Kochen and Specker . On the face of it this seems rather surprising. The Bell-Kochen-Specker result shows that a certain type of hidden-variables approach to quantum theory can lead to a contradiction because it makes assumptions incompatible with the structure of Hilbert space. On the other hand, the CH interpretation has been explicitly constructed to take account of the structure of Hilbert space, and does not rely on hidden variables in any way. Closer examination shows that the Bassi and Ghirardi argument violates the single-framework rule, and thus the claimed contradiction with Bell-Kochen-Specker is not a consequence of the principles of the CH approach, but is instead due to Bassi and Ghirardi’s having rejected those principles. This was pointed out in in response to , but since is considerably longer and also somewhat clearer than , raises the issue in a somewhat different way, and has appeared in a different journal, a separate reply to it seems appropriate. The present article, in order to be self-contained, contains a certain amount of overlap with . Since the arguments in (as in ) deal entirely with the Hilbert space description of a system at a single time, most of the formal machinery of the CH approach—histories, consistency, and assignment of probabilities by use of the time-dependent Schrödinger equation—is not needed for the following discussion. The essential point we wish to make is that a quantum Hilbert space differs in crucial respects from a classical phase space, and this mathematical difference must be reflected in any valid physical interpretation of quantum theory. Importing “intuitively obvious” ideas of classical physics into quantum mechanics without paying adequate attention to the mathematical structure of the latter, in direct contradiction to the rules of the CH interpretation, is what has given rise to the contradiction noted by Bassi and Ghirardi, as we shall show. Before dealing with the main issue, we need to indicate the connection between truth functionals—our name for the homomorphisms denoted by $`h`$ in —and certain elementary concepts from standard probability theory. This is done in Sec. 2, and the quantum counterparts of truth functionals and probability concepts are taken up in Sec. 3, along with the single-framework rule. In Sec. 4 we show that in their argument Bassi and Ghirardi have mistakenly replaced the single-framework rule with what we call the every-framework principle, which is not only not the same as the single-framework rule, but stands in direct contradiction to it; for this reason their argument has basically nothing to do with CH quantum theory. Section 5 responds to some other less-important issues in , and Sec. 6 has a brief conclusion. ## 2 Sample Spaces and Truth Functionals A basic concept in elementary probability theory is that of a sample space. According to Feller , the possible outcomes of an idealized experiment correspond to precisely one and only one point of the sample space. If a coin is tossed, the sample space consists of two possibilities, H and T; if a die is rolled, there are six points in the sample space. Before the idealized experiment is carried out one does not, in general, know what the outcome will be, but when it has taken place, one and only one of the outcomes actually occurs, i.e., is the true result of the experiment. The books on probability theory with which I am familiar do not seem to employ the terms “true” and “false”, but the way in which they define a sample space justifies the association of “true” with the sample point that represents the actual outcome, and “false” with all the others. In addition, Feller distinguishes simple events, the elements of the sample space, from compound events which are associated with some subset of the elements of the sample space. A compound event is “true” if it contains the point of the sample space which actually occurs, and is “false” otherwise. Rather than the terminology of ordinary probability theory, uses what I call a truth functional: a homomorphism (there denoted by $`h`$) from a Boolean algebra of events to the set $`\{0,1\}`$, also thought of as a Boolean algebra. In light of the preceding remarks, the connection of a sample space, and its corresponding event algebra, to a system of truth functionals can be explained in the following way. Let $`\mathrm{SS}`$ be the sample space, and $`𝒫`$ be some subset of $`\mathrm{SS}`$, thus a compound event in Feller’s terminology. The indicator $`P`$ of $`𝒫`$ is a function on $`\mathrm{SS}`$ taking the value 1 at all points which lie in $`𝒫`$ and 0 at all other points of $`\mathrm{SS}`$. The usual Boolean algebra of subsets of $`\mathrm{SS}`$ is then isomorphic to a Boolean algebra $``$ of indicator functions in which the greatest element is the function $`I`$, equal to 1 at all points of $`\mathrm{SS}`$; the least element is $`0`$, equal to 0 everywhere; the complement of an indicator $`P`$ is $`IP`$; the join $`PQ`$ of two indicators is their product $`PQ`$; and the meet $`PQ`$ is the indicator $`P+QPQ`$. A truth functional $`\theta `$ is then a function which assigns to each indicator in the algebra $``$ either the value 1 (true) or 0 (false) in a way which satisfies the following three conditions: $$\theta (I)=1,\theta (IP)=1\theta (P),\theta (PQ)=\theta (P)\theta (Q).$$ (1) It is not hard to show that any such function is necessarily of the form $$\theta _q(P)=P(q)=\{\begin{array}{cc}1\hfill & \text{if }q𝒫\text{,}\hfill \\ 0\hfill & \text{if }q𝒫\text{,}\hfill \end{array}$$ (2) where $`q`$ is some point in the sample space $`\mathrm{SS}`$. One should think of $`\theta _q`$ as the the truth functional appropriate for the case in which the sample point $`q`$ actually occurs, or is true, since it then assigns the value 1 (true) to every compound event which contains $`q`$, and 0 (false) to the ones which do not contain $`q`$. If the sample space is discrete, one can think of $`\theta _q(P)`$ in probabilistic terms as the conditional probability of $`P`$ given $`q`$, assuming the probability of $`q`$ is greater than zero, so that the conditional probability is defined. It is in this sense, among others, that one can say that “true” is associated with a (conditional) probability of 1, and “false” with probability 0, in a probabilistic theory. This approach can be employed in classical statistical mechanics in the following way. Let $`\gamma `$ be a representative point of the phase space $`\mathrm{\Gamma }`$. A physical property $`P`$ of the system corresponds to the subset $`𝒫`$ of $`\mathrm{\Gamma }`$ consisting of those points $`\gamma `$ for which this property is true. The corresponding indicator $`P(\gamma )`$ is 1 whenever $`\gamma `$ is in $`𝒫`$, and 0 otherwise. For example, if $`P`$ is the property that the total energy of a one-dimensional harmonic oscillator is less than some constant $`E_0`$, $`𝒫`$ is the region inside an appropriate ellipse in the $`x,p`$ plane ($`x`$ the position, $`p`$ the momentum), and $`P(\gamma )`$ is 1 for $`\gamma `$ inside and 0 for $`\gamma `$ outside this ellipse. Now consider a coarse graining of the phase space into a collection $`𝒟`$ of $`N`$ non-overlapping regions or “cells”. We can write the identity indicator $`I`$ (equal to 1 for all $`\gamma `$) in the form $$I=\underset{j=1}{\overset{N}{}}D_j,$$ (3) where $`D_j`$ is the indicator corresponding to the $`j`$’th cell. Since the cells do not overlap it follows that $$D_jD_k=\delta _{jk}D_j,$$ (4) consistent with the obvious fact that $`I^2=I`$. The set of $`2^N`$ indicators which can be written as $$P=\underset{j=1}{\overset{N}{}}\pi _jD_j,$$ (5) with $`\pi _j`$ is either 0 or 1, form a Boolean algebra $``$ using the definitions of complement, meet, and join introduced earlier. A truth functional $`\theta `$ is a function on $``$ taking the values 0 or 1 in a way which satisfies (1), so it has the form $$\theta _k(P)=\{\begin{array}{cc}1\hfill & \text{if }PD_k=D_k\text{,}\hfill \\ 0\hfill & \text{if }PD_k=0\text{,}\hfill \end{array}$$ (6) where $`D_k`$ is one of the elements of (3). Note that the collection $`𝒟`$ of cells constitutes a sample space, because in any given “experiment” the phase point $`\gamma `$ representing the system will be in one and only one of the cells. The truth functional $`\theta _k`$ corresponds to the case in which the phase point $`\gamma `$ is somewhere in the cell $`𝒟_k`$; it assigns the value 1 to all collections of cells whose union contains the phase point, and 0 to all others. One can again interpret $`\theta _k(P)`$ as a conditional probability, assuming that the probability assigned to $`𝒟_k`$ is positive. Notice that it is because we are assuming that $`P`$ is of the form (5) that the product $`PD_k`$ must have one of the two forms on the right side of (6): no property of the form (5) can include part but not all of some cell $`D_k`$. Consequently, (6) defines a truth functional for indicators belonging to this particular algebra $``$, but not for all possible properties; in this sense a truth functional is relative to a particular coarse graining $`𝒟`$, or its Boolean algebra $``$. However, in classical mechanics it is possible to construct a universal truth functional which is not limited to a single Boolean algebra, but which will assign 0 or 1 to any indicator on the classical phase space in a manner which satisfies (1). To do this, choose some point $`\gamma _q`$ in $`\mathrm{\Gamma }`$, and let $$\theta _q(P)=P(\gamma _q).$$ (7) That is, $`\theta _q`$ assigns the value 1 to any property which contains the point $`\gamma _q`$, and 0 to any property which does not contain this point, in agreement with how one would normally understand “true” in a case in which the state of the system is correctly described by the phase point $`\gamma _q`$. ## 3 Quantum Truth Functionals and the Single Family Rule The quantum counterpart of a classical phase space is a Hilbert space $``$. For our purposes it suffices to consider cases in which $``$ is of finite dimension, thus avoiding the mathematical complications of infinite-dimensional spaces. Following von Neumann , we associate a quantum property, the counterpart of a set of points in the classical phase space, with a linear subspace $`𝒫`$ of $``$, or the corresponding orthogonal projection operator or projector $`P`$ onto this subspace. If $`I`$ is the identity operator, the negation of a property $`P`$ corresponds to the projector $`IP`$, and the conjunction $`PQ`$ of two properties corresponds to the projector $`PQ`$ in the case in which $`P`$ and $`Q`$ commute with each other. If $`PQQP`$, then neither $`PQ`$ nor $`QP`$ is a projector, so there is no obvious way to define a property corresponding to the conjunction, an issue to which we shall return. The quantum counterpart of a coarse graining of a classical phase space is a decomposition $`𝒟`$ of the identity, a collection of mutually orthogonal projectors $`\{D_j\}`$ satisfying (4) whose sum is the identity, as in (3). This decomposition gives rise to a set of projectors of the form (5), all of which commute with each other, and which form a Boolean algebra $``$ analogous to the algebra of classical indicator functions. One can define a quantum truth functional $`\theta `$ on the elements of $``$ in the manner indicated previously: it assigns to every projector $`P`$ in $``$ a value 0 or 1 in a way which satisfies the three conditions in (1). Once again, any truth functional of this type can be written in the form (6) for some $`k`$, and thus there is a one-to-one correspondence between truth functionals and the elements of $`𝒟`$, which one can think of as the quantum version of a sample space. The CH approach to quantum theory is “realistic” in the sense that it treats the members of a particular decomposition of the identity, a quantum sample space, as mutually exclusive possibilities, one and only one of which occurs, or is true, for a particular physical system at a particular instant of time, in precisely the same sense as in classical statistical mechanics. The difference between quantum and classical physics emerges not when one considers a single quantum sample space, but when one asks about the relationship between two or more different sample spaces. Here quantum theory is very different from classical physics because the product of two quantum projectors $`P`$ and $`Q`$ on the same Hilbert space can depend upon the order, and when $`PQ`$ is unequal to $`QP`$, neither of these products is a projector. By contrast, the product of two indicators on the same classical phase space is always an indicator, since multiplication is commutative. For example, for a spin-half particle with components of angular momentum $`S_x`$, $`S_y`$, and $`S_z`$ (in units of $`\mathrm{}`$), the projector for the property $`S_x=+1/2`$, which is $`\frac{1}{2}I+S_x`$, does not commute with $`\frac{1}{2}I+S_z`$, the projector for $`S_z=+1/2`$. Consequently, a key question in quantum theory, with no counterpart in classical physics, is: How can one make sense out of the conjunction of two quantum properties, such as $`S_x=+1/2`$ AND $`S_z=+1/2`$, when the corresponding projectors do not commute with each other? The answer of the consistent historian is that one cannot make sense of $`S_x=+1/2`$ AND $`S_z=+1/2`$; it is a meaningless statement in the sense that (CH) quantum theory assigns it no meaning. There are no hidden variables, and thus there is a one-to-one correspondence between quantum properties and subspaces of the Hilbert space in CH quantum theory. Since every one-dimensional subspace of the two-dimensional Hilbert space $``$ of a spin-half particle corresponds to a spin in a particular direction, there is no subspace left over which could plausibly represent the property $`S_x=+1/2`$ AND $`S_z=+1/2`$. To be sure, one might assign to it the zero element of $``$, a zero-dimensional subspace corresponding to the property which is always false (analogous to the classical indicator which is 0 everywhere). This, in fact, was the proposal, for this particular situation, of Birkhoff and von Neumann in their discussion of quantum logic . It is important to notice the difference between their approach and the one used in CH. A proposition which is meaningful but false is very different from a meaningless proposition: the negation of a false proposition is a true proposition, whereas the negation of a meaningless proposition is equally meaningless. The Birkhoff and von Neumann approach requires, as they themselves pointed out, a modification of the ordinary rules of propositional logic, whereas the CH approach does not.<sup>1</sup><sup>1</sup>1For further remarks on some of these issues, see Sec. IV A of . However, in CH quantum theory it then becomes necessary to exclude meaningless talk from meaningful discussions, a task which is not altogether trivial. Generalizing from this example, the CH approach requires that a meaningful probabilistic description of a single single quantum system at a particular time must employ a single framework: a single Boolean algebra of commuting projectors generated, in the sense of (5), from a specific decomposition of the identity or quantum sample space. To be sure, many alternative descriptions can be constructed using different decompositions of the identity; the single-framework rule is certainly not intended to restrain the imagination of theoretical physicists! However, combining results from different sample spaces into a single description is forbidden by the single-framework rule, apart from the following exception. Two frameworks involving properties of a single system at a single time (we are ignoring genuine histories, for which the rules are more complex) are said to be compatible provided the two Boolean algebras are parts of a single, larger Boolean algebra of commuting projectors. This is true if and only if every projector belonging to one of the original algebras commutes with every projector belonging to the other algebra, which in turn is the same as requiring that the projectors from the two decompositions of the identity, or sample spaces, commute with one another. A larger collection of frameworks is compatible if all pairs are compatible, and frameworks are said to be mutually incompatible if they are not compatible. Descriptions based upon two or more compatible frameworks can always be combined by using the single Boolean algebra which contains all of the different (mutually commuting) algebras, and thus one is still employing a single framework, corresponding to a single decomposition of the identity, in accordance with the single-framework rule. The single-framework rule is not at all unreasonable from the perspective of elementary probability theory, where problems are generally set up using a single sample space. Thus if a coin is to be tossed ten times in a row, the statistical properties are worked out not by constructing ten sample spaces, but by using a single sample space containing $`2^{10}`$ points. The single-framework rule is also perfectly compatible with classical statistical mechanics, for if one uses two or more coarse grainings of the phase space, the results can always be combined by means of a single coarse graining which uses a collection of cells generated by intersections of other cells in an obvious way. Thus ordinary probabilistic arguments and classical statistical mechanics satisfy the single-framework rule, albeit in a somewhat trivial sense. As already noted, the single-framework rule as applied to Boolean algebras of properties refers to a single system at a single instant of time. Given two nominally identical systems, there is no reason why one cannot use one framework for the first and a different framework for the second. For instance, in the case of two spin-half particles, $`S_x=+1/2`$ could be a correct description of one of them at the same time that $`S_z=+1/2`$ is a correct description of the other. Similarly, the same particle may have $`S_x=+1/2`$ at an earlier and $`S_z=+1/2`$ at a later time. Conversely, when incompatible frameworks turn up in some discussion of a quantum system, it is best to think of them as referring to different systems, or to a single system at different times, or perhaps simply as tentative or hypothetical proposals without any suggestion that they should be taken in a realistic sense. (Ascribing properties to a single system at more than one time requires the use of a history, and this requires additional considerations which lie outside the scope of the present discussion.) In any application of probability theory, precisely one of the elements of the sample space is thought of as existing, or “true” in any realization of an ideal experiment. In this sense the notion of “truth” in a probabilistic theory is necessarily connected with, and thus depends upon the sample space or framework one is considering. In classical physics one can forget about this dependence, because if more than one framework is under consideration, in the case of a single system at a single time, they can always be combined into a single framework. This is reflected in the fact that one can always define a universal truth functional for a classical phase space, as noted in Sec. 2. Because of the possibility that frameworks can be incompatible, the framework dependence of “true” is not at all trivial in quantum physics; indeed, one might say that this is one of the main ways in which the mathematical structure of quantum theory forces one to adopt a different kind of physical interpretation from what one is used to in classical mechanics. In particular, in quantum theory there is no universal truth functional $`\theta _q`$ which can be used to assign values of 0 and 1 to all projectors in a way which agrees with the three conditions in (1). In a certain sense this is immediately obvious for any Hilbert space of dimension 2 or more, since in such a space there will always be projectors $`P`$ and $`Q`$ which do not commute with each other. In such a case $`PQ`$ is not a projector, and the third condition in (1) is not even defined, much less satisfied. One might hope to get around this problem by modifying the third condition and only requiring that it hold in cases in which $`P`$ and $`Q`$ commute with each other. This, however, gains very little, for the results of Bell and of Kochen and Specker referred to earlier demonstrate that even such a “modified” universal truth functional does not exist for a Hilbert space of dimension 3 or more. (A simple example due to Mermin, showing the impossibility of a universal truth functional in a Hilbert space of dimension 4, is discussed in .) The absence of a universal truth functional causes no difficulties for CH quantum theory because of the single-framework rule, which prevents the comparison of incompatible frameworks. The situation is different for the alternative principle proposed by Bassi and Ghirardi, which they have somehow managed to confuse with the single-framework rule, and which will be taken up next. ## 4 The Every-Framework Principle of Bassi and Ghirardi In Bassi and Ghirardi introduce, in the discussion leading up to and including their (6.1), what I shall call the “every-framework principle”, which in the notation of the present paper can be stated in the following way. Consider a quantum Hilbert space, and let $`\{𝒟_f\}`$ be the different possible decompositions of the identity, where $`f`$ is a label which takes on uncountably many values. For example, for a spin-half particle, $`f`$ will run over all directions in space $`w`$, as long as $`+w`$ is identified with $`w`$, since each decomposition of the identity corresponds to a sample space with just the two points $`S_w=\pm 1/2`$. Corresponding to $`𝒟_f`$ there is a corresponding Boolean algebra $`_f`$ of projectors of the form (5). Given a projector $`P`$, we define $$(P)=\{f:P_f\}$$ (8) to be the collection of labels such that $`P`$ is a member of the Boolean algebra $`_f`$. The every-framework principle asserts that there is a collection of truth functionals $`\{\theta _f\}`$, one for each decomposition of the identity, with the following property: if $`P`$ is any projector in the Hilbert space, the value of $`\theta _f(P)`$ is the same for all $`f`$ in $`(P)`$. That is to say, $`P`$ is assigned precisely the same truth value, 0 or 1, by all members of the collection $`\{\theta _f\}`$ for which $`\theta _f(P)`$ is actually defined. The every-framework principle has a certain intuitive appeal when one is thinking of a single system at a single time. It is actually correct for classical statistical mechanics, where a property $`P`$ is true as long as the representative phase point $`\gamma `$ is inside the corresponding set $`𝒫`$, and false otherwise. Hence to construct a collection of truth functionals, associated with a collection of coarse grainings, satisfying the every-framework principle, one simply chooses some representative point $`\gamma _q`$ in the phase space, and for a coarse graining $`𝒟_f`$ lets the indicator for the cell which contains $`\gamma _q`$ play the role of the special $`D_k`$ in (6). Or, to put the matter in a slightly different way, one simply lets $`\theta _f`$ be the restriction to $`_f`$ of the universal truth functional defined in (7). Given this result, one is not surprised to learn that the quantum-mechanical version of the every-framework principle implies the existence of a universal truth functional $`\theta _u`$ of the modified form discussed towards the end of Sec. 3. For each $`P`$, one sets $`\theta _u(P)`$ equal to the common value specified by the every-framework principle, noting that every $`P`$ is contained in at least one decomposition of the identity, that consisting of $`P`$ and $`IP`$. Since when restricted to the Boolean algebra $`_f`$ the functional $`\theta _u`$ is identical to $`\theta _f`$, it is at once evident that the first two requirements in (1) will always be satisfied, while the third will be satisfied in cases in which $`P`$ and $`Q`$ commute, since there is then at least one Boolean algebra $`_g`$ which contains both of them, and $`\theta _u`$ when restricted to $`_g`$ is the same as the corresponding truth functional $`\theta _g`$. But, as we have already noted, the nonexistence of a universal truth functional has been proven mathematically for any Hilbert space of dimension greater than two. Consequently, the every-framework principle is in clear contradiction with the principles of quantum mechanics. The proofs of this fact given in are correct but superfluous; they simply repeat what is already well known to people who work in the foundations of quantum theory. What is the relationship of the every-framework principle and the single-framework rule? They are mutually contradictory, for fairly obvious reasons. The every-framework principle requires us to compare the results of truth functionals, or equivalently sample spaces, associated with different and in general incompatible frameworks, in precisely the manner forbidden by the single-framework rule. According to the latter, such a comparison makes no sense in the case of incompatible frameworks. (Note that it is only with the help of incompatible frameworks that one can reach a contradiction using the Bell-Kochen-Specker approach, so incompatible frameworks are essential to the argument in .) To be sure, two different frameworks might refer to two different systems (or the same system at two different times), but in that case there is no reason whatsoever to expect that a particular property has the same truth value for the two systems, and thus no motivation for invoking the every-framework principle. One must admit that the every-framework principle has a certain intuitive appeal: how could the truth value of some physical property possible depend upon the sample space in which it is embedded? Surely if it is true it is really true, apart from anything one can say about the rest of the world, and if it is false it is false! This appeal is seductive because it focuses attention on the physical property rather than on the sample space. When, however, one pays attention to the latter, things appear in a quite different light. Let us consider as an example two incompatible quantum sample spaces $$\mathrm{SS}_1=\{A,B,C\}.\mathrm{SS}_2=\{A,D,E\},$$ (9) where $`A`$, $`B`$, and $`C`$ are three projectors which add up to $`I`$, and likewise $`A+D+E=I`$. However, neither $`B`$ nor $`C`$ commutes with either $`D`$ or $`E`$. Let us employ the every-framework principle, and suppose that $`A`$ is false in both $`\mathrm{SS}_1`$ and $`\mathrm{SS}_2`$. Because $`\mathrm{SS}_1`$ is a sample space, this means that either $`B`$ or $`C`$ is true, and because $`\mathrm{SS}_2`$ is a sample space, either $`D`$ or $`E`$ must be true. Suppose for the sake of argument that it is $`B`$ which is true in $`\mathrm{SS}_1`$ and $`D`$ which is true in $`\mathrm{SS}_2`$. Then if we insist that $`\mathrm{SS}_1`$ and $`\mathrm{SS}_2`$ apply to the same system at the same time, this means that two properties represented by non-commuting projectors are simultaneously true. For example, they could be $`S_x=+1/2`$ and $`S_z=+1/2`$ for a spin-half particle. This is hard to reconcile with the Hilbert space structure of ordinary quantum mechanics, as pointed out in Sec. 3, and in CH quantum theory it is forbidden by the single-framework rule. Thus we see that when $`A`$ is false, the every-framework principle has certain implications which, when brought to light, make it much less appealing. The case in which $`A`$ is true in both $`\mathrm{SS}_1`$ and $`\mathrm{SS}_2`$ also leads to unsatisfactory results. Because $`\mathrm{SS}_1`$ and $`\mathrm{SS}_2`$ are sample spaces, the truth of $`A`$ means that all four properties $`B`$, $`C`$, $`D`$, and $`E`$ are false. One might be tempted to suppose that the falsity of two incompatible properties is unproblematical—after all, who cares about things which do not occur? The trouble is that when $`B`$ is false, its negation $`\stackrel{~}{B}=IB`$ is true. Furthermore, if two projectors $`B`$ and $`E`$ do not commute, the same is true of their negations $`\stackrel{~}{B}`$ and $`\stackrel{~}{E}=IE`$. Thus using the every-framework principle once again leads to the conclusion that two properties represented by non-commuting operators are simultaneously true. In summary, whatever may be its initial intuitive appeal, much of the allure of the every-framework principle vanishes when one realizes what it really means. Let us look at this example from a slightly different perspective, by introducing a third sample space $$\mathrm{SS}_0=\{A,\stackrel{~}{A}=IA\}$$ (10) containing only $`A`$ and its negation. This is obviously the smallest sample space in which “$`A`$ is true” and “$`A`$ is false” make sense. The sample space $`\mathrm{SS}_0`$ is compatible with both $`\mathrm{SS}_1`$ and with $`\mathrm{SS}_2`$, each of which represents a refinement of $`\mathrm{SS}_0`$. The rules of quantum reasoning employed in allow one to deduce that if $`A`$ is true/false in $`\mathrm{SS}_0`$, then it is also true/false in $`\mathrm{SS}_1`$, and the same deduction is possible going from $`\mathrm{SS}_0`$ to $`\mathrm{SS}_2`$. However, the single-framework rule prevents combining these results in a single description: one cannot employ both $`\mathrm{SS}_1`$ and $`\mathrm{SS}_2`$ for the same system at the same time, for the reasons indicated previously. This means that at least some of the intuitive appeal which seems to lie behind the every-framework principle, the notion that the truth of some property should not depend upon what else is going on in the world, is supported in CH quantum theory. And this can be done in a consistent way without leading to any logical contradictions precisely because the CH approach employs the single-framework rule rather than the every-framework principle. The single-framework rule and the every-framework principle are, thus, completely incompatible with each other, whether one regards them either from a purely formal perspective—the former forbids combinations which the latter allows—or in terms of their intuitive significance. Hence one can only regard with astonishment the claim of Bassi and Ghirardi, found in the very next sentence after their (6.1), that the every-framework principle constitutes the “only reasonable way” to interpret the single-framework rule! It is hard to imagine a more serious misunderstanding of a rule that has been stated over and over again in the literature on CH quantum theory, and illustrated by means of numerous examples. Discussions and criticisms of the single-framework rule can be a valuable component of the scientific enterprise. But to introduce a new principle which is not only different from, but directly contrary to the single-framework rule, and then claim that the former is the only reasonable way to interpret the latter does nothing but cause confusion. ## 5 Some Other Issues Aside from the every-framework principle, there are some other points in (and also ) which merit at least a brief response. $``$In Secs. 4 and 9 of , Bassi and Ghirardi assert that what I call MQS (macroscopic quantum superposition) states—for example, Schrödinger’s infamous cat—are physically unacceptable, and fault the CH approach for not providing some criterion for excluding them. In response, it will help to use an analogy from classical physics, while remembering that any classical analogy can only go part way in helping us understand quantum phenomena. A coin spontaneously rising a centimeter above a table on which it is sitting at rest is physically unacceptable in the sense that such a violation of the second law of thermodynamics is never observed to occur, despite the fact that nothing in the laws of classical mechanics excludes such a possibility. We understand why we never observe such things by using statistical mechanics, which assigns an extremely small probability to such an event. That is, we have a scientific understanding of why violations of the second law are not observed, despite the fact that the laws of classical (and also quantum) mechanics permit such possibilities. The quantum Hilbert space certainly contains MQS states, because it is, by definition, a linear vector space which includes superpositions of any of its elements. However, MQS states are incompatible, in the quantum sense, with the quasi-classical framework(s) needed to describe our ordinary experience with macroscopic objects. Thus the single-framework rule tells us that it makes no sense trying to include MQS states in descriptions of the everyday world of human experience. Conversely, if a quantum description employs the MQS state that is (formally) a linear superposition of a live and dead cat, it makes no sense, according to the single-framework rule, to think of the whatever-it-is as somehow involving a cat, for the properties typically used to identify a cat will, in quantum theory, be represented by projectors which do not commute with the MQS state, and are therefore of no use for discussing the meaning of such a state. In this sense, at least, CH quantum theory does provide criteria for excluding MQS states from certain types of quantum descriptions. Quantum physicists who refuse to employ the single-framework rule must, of course, find some other means of disposing of, or perhaps peacefully coexisting with MQS states. It is also worth noting that the reason quantum superpositions states of this sort cannot be detected in the laboratory, even for microscopic objects, as long as they contain a substantial number of atoms, is by now reasonably well understood in terms of the process of decoherence, a topic which has been treated from the CH perspective by Omnès . (Decoherence is much like classical irreversibility, making the jumping coin an even better analogy.) To summarize the situation, CH quantum theory certainly permits descriptions using MQS states, but at the same time provides an explanation as to why they are neither needed nor particularly useful for a science of the macroscopic world. $``$In a not unrelated point, Bassi and Ghirardi suggest that one may be able to get around the difficulties they have encountered by employing their every-framework rule by making a drastic reduction in the set of consistent families which can be considered to be physically significant. In response, there is nothing wrong with these authors announcing a direction for their future research, as long as they make it plain that the motivation for it comes not from any problem involving the CH approach, but rather from the disagreement between their every-framework principle (itself completely contrary to the CH single-framework rule) and standard quantum theory. No proposal for using a restricted class of families in the manner they propose has thus far turned out to be very useful for quantum interpretation, but no doubt those who consider this a worthwhile approach will continue the search. $``$Bassi and Ghirardi take the position, both in Sec. 6 of and in , that the only alternative to their every-framework principle in which any property $`P`$ has the precisely the same truth value in every framework which contains it, is to suppose that in certain of frameworks it is true and in other frameworks it is false, something they consider unacceptable. The response to this is contained in the material in Sec. 4 above, but it may be worthwhile making it quite explicit. From the CH perspective, using two incompatible frameworks to describe the same system at the same time is not meaningful—this is precisely the point of the single-family rule. Since meaningless truth values are meaningless, there is no reason to be concerned about whether they agree or disagree. Alternatively, one can suppose that two incompatible frameworks do not refer to the same system at the same time. In that case, there is no a priori reason to expect the truth values for a particular property to be the same, and so no reason to be worried if they are different. $``$In Bassi and Ghirardi assert that in the previous literature on consistent histories the single-framework rule was not explained well enough or clearly enough so as to obviously exclude their every-framework principle. It is quite true that the language of truth functionals was not employed by consistent historians (so far as I am aware) prior to the recent . Previous work used the standard language of elementary probability theory, with its sample spaces and event algebras, and assumed the usual association between probability theory and reality, as pointed out, for example, in Sec. 7.2.3 of . A basic understanding of how sample spaces function in ordinary probability theory is all that one really needs in order to understand the CH approach and the significance of the single-framework rule, including the fact that it is quite contrary to the every-framework principle. The use of truth functionals, while it may be advantageous for some purposes, is not actually needed. $``$At the beginning of Sec. 7.1 in , Bassi and Ghirardi, in a footnote, issue a challenge to me and an anonymous referee to identify which of their four precisely formulated (in their opinion) assumptions are inconsistent with the CH single-framework rule, and accept the consequences of this identification. In fact these four assumptions are not precisely formulated, as was pointed out in . Writing in response to that, Bassi and Ghirardi have themselves identified their assumption (c) as the one which is incompatible with the single-framework rule, and I see no reason to dispute this. That rejecting their (c)—the every-framework principle—leads to dire consequences is not true, as should be clear from the discussion in Sec. 4 above. Instead, it allows a sensible discussion of quantum properties using consistent quantum principles. ## 6 Conclusion In Bassi and Ghirardi have, in essence, substituted their every-framework principle for the single-framework rule of CH quantum theory, and then concluded, correctly, that the every-framework principle makes no sense in the quantum world. Their only mistake is in supposing that the every-framework principle has something to do with CH quantum theory, whereas in fact the two are directly contrary to each other. While this error is easily spotted by someone who is familiar with CH methods, it is nonetheless regrettable that others less familiar with them have, once again, been given the mistaken impression that there is something logically unsound, or at least suspicious, about CH quantum theory. To be sure, Bassi and Ghirardi and other critics of CH perform a valuable function in looking for flaws in this approach. Their failure (at least thus far) to find anything wrong with CH, while at the same time demonstrating that the various alternatives that they propose posses serious flaws, adds to one’s confidence that the CH approach does, in fact, provide a satisfactory realistic interpretation of quantum theory. ## Acknowledgments The author is indebted to T. A. Brun and O. Cohen for reading and commenting on the manuscript. The research described here was supported by the National Science Foundation Grant PHY 99-00755.
warning/0001/hep-th0001125.html
ar5iv
text
# 1 Introduction ## 1 Introduction An essential ingredient for the proper formulation of systems consisting of strings and (D-) branes is the Born-Infeld (BI) action. The pioneering works derive it from the condition that the beta-function for the string vanishes, when the string is embedded in an external gauge field. From the extensive literature on the application of this action in brane theory and its vast number of consequences we only refer to some recent work which is also based upon older results . In our present note we address some delicate point in the derivation of the BI action which is related to the existence of a nontrivial boundary. The use of the heat kernel technique appears to be a necessity in this context. However, in order to obtain a well-posed spectral problem it turns out that also a new rule for Euclidean continuation has to be introduced. We also rely heavily on the seminal paper of O. Alvarez who already a long time ago strongly emphazised the consistent (gauge-invariant) treatment of the boundary. In our present case these techniques are extended by the introduction of an (abelian) gauge field. ## 2 The sigma model To simplify the discussion we represent the string by a sigma model action with Euclidean metric both on the world sheet and in space-time $$S=\frac{1}{2\pi \alpha ^{}}\left[\frac{1}{2}_{}d^2z\sqrt{h}h^{ab}_aX_\mu _bX^\mu +_{}𝑑\tau A_\mu _\tau X^\mu \right]$$ (1) The string tension $`2\pi \alpha ^{}`$ which is usually written as a factor in front of the boundary term has been absorbed in $`A_\mu `$ for convenience. We suppose that the metric $`h_{ab}`$ is flat, but the boundary $``$ can be of an arbitrary shape and may contain several connected components. We ignore non-trivial background fields in the closed string sector ($`G_{\mu \nu }=\delta _{\mu \nu }`$,) and do not include a dilaton interaction and the $`B`$-field. The generalization is straightforward for the case when these fields are taken into account. Let $`N^a`$ be an inward pointing unit normal to the boundary. We choose the coordinate system in such way that on $``$ the vectors $`N_adz^a,d\tau `$ form an orthonormal pair. It should be emphazised that the boundary term in (1) is real in contrast to the Euclidean actions of . There the relative factor of $`i`$ originated from the continuation of the volume element $`\sqrt{h}i\sqrt{h}`$ in the first term. We suggest to also rotate $`A_\mu `$ to $`iA_\mu `$ during continuation to the Euclidean space so that the factor $`i`$ is compensated. Actually, it does not matter which particular continuation is chosen as long as one gets a meaningful theory in Euclidean space so that a proper continuation back to the physical Minkowski space is possible. As will be demonstrated below, our rule of continuation has the crucial advantage to provide a well-posed spectral problem in Euclidean space, as opposed to the usual approach when the factor $`i`$ enters the boundary conditions for real fields and thus makes the boundary value problem ill-defined. The fact that parity odd fields can get an imaginary factor in the Euclidean space has been observed long ago in the context of chiral theories. It is easy to see that the field $`A_\mu `$ is of odd parity from the world volume point of view. Indeed, the last term in (1) can be rewritten as $`_{}_a(\epsilon ^{ab}A_\mu _bX^\mu )`$. Here $`A_\mu `$ couples to the parity-odd quantity $`\epsilon ^{ab}`$. Whenever it is possible to compare our results to those of previous related papers where this mathematical subtlety was just ignored, they are compatible after the replacement $`AiA`$. Let us expand the action (1) around an arbitrary background $`\overline{X}`$, $`X=\overline{X}+\xi `$. To calculate the one-loop (in the field theory sense) effective action we need only the part which is quadratic in $`\xi `$: $`S_2={\displaystyle \frac{1}{2\pi \alpha ^{}}}[{\displaystyle \frac{1}{2}}{\displaystyle _{}}d^2z\sqrt{h}h^{ab}_a\xi _\mu _b\xi ^\mu +`$ $`+{\displaystyle \frac{1}{2}}{\displaystyle _{}}(F_{\nu \mu }\xi ^\nu \dot{\xi }^\mu +\dot{\overline{X}}^\mu _\nu F_{\rho \mu }\xi ^\nu \xi ^\rho )]`$ (2) Clearly the one-loop effective action is $$W=\frac{1}{2}\mathrm{log}det(\mathrm{\Delta }\delta _{\mu \nu }),$$ (3) with the scalar Laplacian $`\mathrm{\Delta }`$ and with the boundary condition $$(_N\delta _{\mu \nu }+F_{\mu \nu }_\tau +\dot{\overline{X}}^\rho (_\nu F_{\mu \rho }))\xi ^\nu |_{}=0,$$ (4) where $`_N=N^a_a`$ is the derivative with respect to the inward pointing unit normal vector N. It is useful to rewrite (4) by adding and subtracting a term with $`\mu \nu `$ as $$\xi |_{}=\left(_N+\frac{1}{2}(_\tau \mathrm{\Gamma }+\mathrm{\Gamma }_\tau )+S\right)\xi |_{}=0,$$ (5) where $$\mathrm{\Gamma }_{\mu \nu }=F_{\mu \nu },S_{\mu \nu }=\frac{1}{2}\dot{\overline{X}}^\rho (_\nu F_{\mu \rho }+_\mu F_{\nu \rho }).$$ (6) Thus the boundary condition (5) also ensures the hermiticity of the Laplace operator. Indeed, $$_{}\left(\xi _{}^{}{}_{\mu }{}^{}\mathrm{\Delta }\xi ^\mu \xi _\mu \mathrm{\Delta }\xi _{}^{}{}_{}{}^{\mu }\right)=_{}\left(\xi _{}^{}{}_{\mu }{}^{}_N\xi ^\mu +\xi _\mu _N\xi _{}^{}{}_{}{}^{\mu }\right)=0,$$ (7) if $`\xi `$ and $`\xi ^{}`$ satisfy (5), when the hermiticity of $`\frac{1}{2}(_\tau \mathrm{\Gamma }+\mathrm{\Gamma }_\tau )+S`$ is used. The boundary condition (5) contains tangential derivatives and belongs to the Gilkey-Smith class<sup>1</sup><sup>1</sup>1Sometimes the boundary conditions with tangential derivatives are called “mixed” . In the theory of the heat equation asymptotics the name “mixed” is reserved for a completely different type of the boundary conditions . . ## 3 The heat kernel expansion In this section we collect some basic information on the heat kernel expansion for Gilkey–Smith boundary conditions. The spectral geometry of such boundary conditions was first studied in . The increased interest for that heat kernel expansion with these boundary conditions in recent years was motivated primarily by one-loop calculations in quantum cosmology . Consider an operator of the Laplace type $`D=(^a_a+E)`$ acting in a smooth vector bundle over a smooth Riemannian manifold $``$ of dimension $`m`$. $``$ denotes a covariant derivative with respect to a certain connection on $``$. $`E`$ is an endomorphism (a matrix-valued function). Let the boundary operator be $$=_N+\frac{1}{2}(\mathrm{\Gamma }^i\widehat{}_i+\widehat{}_i\mathrm{\Gamma }^i)+S,$$ (8) where $`\mathrm{\Gamma }^i`$ and $`S`$ are some matrices depending on the coordinates on the boundary, $`i=1,\mathrm{},m1`$. The covariant derivative $`\widehat{}`$ contains both the standard Levi-Civita connection of the boundary and the restriction of the bundle connection to the boundary. The operator (8) defines the boundary condition $`\varphi |_M=0`$. The Laplace operator $`D`$ is symmetric if the $`\mathrm{\Gamma }^i`$’s are antihermitian and $`S`$ is hermitian. In our case the matrices (6) satisfy this requirement. If $`f`$ is a smooth function on $``$, there is an asymptotic series as $`t+0`$ of the form $$\text{Tr}(f\mathrm{exp}(tD))=\underset{n0}{}t^{m/2+n}a_n(f,D,),$$ (9) where $`n=0,\frac{1}{2},1,\mathrm{}`$. The functional trace in the space of square integrable functions is denoted by $`\mathrm{Tr}`$. The heat kernel (9) is well defined only if the boundary value problem is strongly elliptic. A criterion for strong ellipticity for the boundary conditions (8) has been proved recently in . Roughly speaking, it requires the absolute values of the eigenvalues of $`\mathrm{\Gamma }^i`$ to be smaller than $`1`$. In the present context the points where the strong ellipticity is lost correspond to critical values of the electric field, which were discussed for Minkowski signature in . The bulk terms in the heat kernel expansion do not depend on $`\mathrm{\Gamma }`$. The leading boundary contributions to the heat kernel can be taken directly from refs. $`a_{1/2}(f,P,)={\displaystyle \frac{1}{(4\pi )^{(m1)/2}}}{\displaystyle _{}}\text{tr}(\gamma f),`$ $`a_1(f,P,)={\displaystyle \frac{1}{(4\pi )^{m/2}}}{\displaystyle _{}}\text{tr}\left(f(b_0k+b_2S+\sigma _1k_{ij}\mathrm{\Gamma }^i\mathrm{\Gamma }^j)+b_1_Nf\right),`$ (10) where $`k_{ij}`$ is the extrinsic curvature of the boundary. Here tr is the ordinary matrix trace. The functions $`b_0,b_1,b_2,\gamma ,\sigma _1`$ are $`\gamma ={\displaystyle \frac{1}{4}}\left[{\displaystyle \frac{2}{\sqrt{1+\mathrm{\Gamma }^2}}}1\right],`$ $`b_1={\displaystyle \frac{1}{\sqrt{\mathrm{\Gamma }^2}}}\text{Artanh}(\sqrt{\mathrm{\Gamma }^2}){\displaystyle \frac{1}{2}},`$ $`b_2={\displaystyle \frac{2}{1+\mathrm{\Gamma }^2}},`$ $`b_0+\sigma _1\mathrm{\Gamma }^2={\displaystyle \frac{1}{3}},`$ (11) In two dimensions the functions $`b_0`$ and $`\sigma _1`$ enter the heat kernel expansion only in the combination $`b_0+\sigma _1\mathrm{\Gamma }^2`$ due the identity $`k_{ij}\mathrm{\Gamma }^i\mathrm{\Gamma }^j=k\mathrm{\Gamma }^2`$, valid on the one-dimensional boundary. In this paper we are not discussing the contribution of the zero modes to the path integral. Therefore, strictly speaking the equations (10) and (11) are valid only up to some “global” effects. ## 4 Beta function and the conformal anomaly Next we make use of the $`\zeta `$-function regularization . The zeta function of an elliptic operator $`D`$ is defined as $$\zeta _D(s)=\text{Tr}(D^s).$$ (12) In term of the zeta function (12) the effective action (3) reads $$W=\frac{1}{2s}\zeta _D(0)\frac{1}{2}\zeta _D^{}(0),$$ (13) where the prime denotes differentiation with respect to $`s`$. By using the well known relation between the zeta function and the heat kernel coefficients $`\zeta _D(0)=a_1(1,D,)`$ the divergent part of the effective action at $`s0`$ may be written as $$W_{\text{div}}=\frac{1}{2s}\frac{1}{4\pi }_{}𝑑\tau \left[\dot{\overline{X}}^\rho (_\nu F_{\mu \rho }+_\mu F_{\nu \rho })(1+F^2)_{\nu \mu }^1+\frac{1}{3}k\delta _\nu ^\nu \right].$$ (14) The first term on the right hand side of (14) can be represented as $`(1/2\pi )_{}𝑑\tau G_\mu \dot{\overline{X}}^\mu `$. From this we can read off the beta function (in the notations of ) $$\beta _\mu ^A(_\rho F_{\nu \mu })(1+F^2)_{\nu \rho }^1.$$ (15) After collecting the factors of $`i`$ which originate from our prescription for Euclidean continuation it gives the same equation of motion as in and, therefore, reproduces the variation of the BI action for the $`A_\mu `$. It should stressed that to derive this result we neither supposed a special geometry of the world sheet, nor had to neglect higher derivatives of $`F_{\mu \nu }`$. The second term under the integral in (14), which is proportional to the dimension of the target space $`\delta _\nu ^\nu `$, does not depend on $`F`$ and is always present in the theory of open strings . Since we had assumed that the scalar curvature of the two-manifold $``$ is zero, this term can be expressed in terms of the Euler characteristic of $``$: $`2\pi \chi ()=_{}𝑑\tau k`$. We next turn to the conformal anomaly. An infinitesimal conformal transformation $`\delta h_{ab}=(\delta k)h_{ab}`$ with a local parameter $`\delta k`$ produces the trace of the (effective) energy momentum tensor $$\delta W_{\text{ren}}=\frac{1}{2}_{}d^2z\sqrt{h}\delta h^{ab}T_{ab}=\frac{1}{2}_{}d^2z\sqrt{h}\delta k(x)T_a^a(x),$$ (16) where the $`W_{\text{ren}}`$ is the second (finite) term in (13). It is quite important that both the Laplace operator and the boundary operator transform covariantly under the metric rescaling $`\mathrm{\Delta }(1k+\mathrm{})\mathrm{\Delta },`$ (17) $`(1{\displaystyle \frac{k}{2}}+\mathrm{}).`$ (18) The second property (18) follows from our assumption that $`N_adz^a,d\tau `$ are two orthonormal vectors – which we used to derive the equation (4). It guarantees that the functional space defined by (4) is invariant under the conformal transformations. With the definition of a generalized $`\zeta `$-function $$\zeta (s|\delta k,D)=\text{Tr}(\delta kD^s)$$ (19) the variation in (16) can be identified with $$\delta W_{\text{ren}}=\frac{1}{2}\zeta (0|\delta k,D),$$ (20) where we used $`\delta \zeta _{D_k}(s)=s\text{Tr}(D^s\delta k)`$. Combining (20) and (16) we obtain $$\zeta (0|\delta k,D)=d^2z\sqrt{h}\delta k(z)T_a^a(x).$$ (21) By a Mellin transformation one can show that $`\zeta (0|\delta k,D)=a_1(\delta k,D,)`$. The (smeared) conformal anomaly reads: $`{\displaystyle _{}}\sqrt{h}d^2zf(z)T_a^a(z)={\displaystyle \frac{1}{4\pi }}{\displaystyle _{}}d\tau [f(\tau )({\displaystyle \frac{1}{3}}k\delta _\nu ^\nu 2\dot{\overline{X}}^\rho (_\nu F_{\mu \rho })(1+F^2)_{\nu \mu }^1)`$ $`+(_Nf)((F^2)_{\mu \nu }^{1/2}\text{Artanh}(\sqrt{F^2})_{\nu \mu }{\displaystyle \frac{1}{2}}\delta _\mu ^\mu )].`$ (22) In the limit $`F0`$ the conformal anomaly (22) coincides with the standard expression . The second term on the first line appears quite naturally and is a manifestation of the BI action. The second line contains a somewhat unusual contribution. To the best of our knowledge it has not been obtained before. To perform a full-scale analysis of this term one should include the dilaton field in the “bare” action (1). We postpone this to a future more detailed publication. In any case, the last term in (22) suggests a very interesting interplay between the gauge sector and the conformal sector of string theory. ## 5 Conclusions By a careful interpretation of the transition to Euclidean space we are able to reformulate the problem of the string in the presence of a nontrivial boundary where the string is coupled to an abelian background field. As a consequence of the effect that our operator, appearing in the boundary condition, is hermitian we arrive at a well-posed elliptic problem in the sense of the heat kernel technique where known formulas from that field can be applied directly. Our central result is the one for the conformal anomaly at the boundary, Eq. (22). Beside the standard term leading to the BI action for the gauge field from the vanishing of the beta-function and a term proportional to the extrinsic curvature at the boundary (which was known for a long time ) we find a new contribution which depends on the gauge field $`F_{\mu \nu }`$. It should be stressed that our whole argument (in contrast to previous ones for the BI action) contains no restriction on the vanishing of derivatives for $`F_{\mu \nu }`$ . Also no special geometry of the world volume need be assumed. This means that our calculations are valid for an arbitrary number of string loops. Of course, higher loop corrections (in the sense of quantum field theory) will contain higher derivatives of $`F_{\mu \nu }`$ . We see a rather wide range of applications of our present result, the most obvious extension being the presence of other fields which, however, should not present any new technical difficulties. Also the indication for a further contribution to the anomaly at the intersection of branes possibly could shed new light upon the problems encountered for interacting strings and branes . There could be even a relation to the very recent work on the noncommutative geometry approach for strings and branes . ## Acknowledgments We are grateful to A. Andrianov, I. Avramidi, I. Bandos and G. Esposito for discussions and/or correspondence. This work has been supported by the Fonds zur Förderung der wissenschaftlichen Forschung project P-12.815-TPH. One of the authors (D.V.) thanks the Alexander von Humboldt Foundation and the Russian Foundation for Fundamental Research, grant 97-01-01186, for the support. ## Note added After this paper has been completed and posted on the net we were informed by Professor Osborn that our expression for the conformal anomaly (22) (that is our main result) is contained in the Appendix of his paper . We find it however quite striking that such important result is not widely recognised in modern literature on strings and branes. Therefore, we decided to leave our preprint on the net and add this short note.
warning/0001/hep-ph0001048.html
ar5iv
text
# 1 Introduction ## 1 Introduction One of the primary tasks for the forthcoming hadronic and leptonic colliders is a detailed study of the top quark properties, in particular, the measurements of the top couplings to gauge fields. Special interest to such measurements is based on the huge difference of the top quark mass and all other fermion masses, providing enhanced expectations for a signal of new physics at the top mass scale . Among the top couplings to other particles the $`Wtb`$ coupling plays a crucial role because it is responsible for practically all top quark decays. Therefore the spacetime structure of the $`Wtb`$ vertex defines the top total width and the characteristics of its decay products. There are two general possibilities to probe and measure directly the $`Wtb`$ vertex structure in collider experiments, either from top pair production processes or from reactions of single top production. The rate of single top production processes is directly proportional to the $`Wtb`$ coupling, and thus it is potentially very sensitive to the $`Wtb`$ structure. This was indeed demonstrated in high energy $`\gamma e`$ collisions as well as for the upgraded Tevatron and the LHC . However, the rate of single top production is usually less than the top pair production rate, in both the lepton and hadron colliders. On the other hand, the reaction $`e^+e^{}t\overline{t}W^+bW^{}\overline{b}`$ includes the $`Wtb`$ coupling only in the subsequent top decays, with the $`t(\overline{t})`$ on-shell decay rate given, apart from small finite width corrections, by the top decay branching fraction to $`Wb`$, which is close to 100%. Consequently, the total rate depends only negligibly on the $`Wtb`$ vertex structure and more sensitive observables, like $`C`$ and $`P`$ asymmetries, top polarization and spin correlations, have to be analysed. The paper starts with the analysis of the process $`e^+e^{}t\overline{t}W^+bW^{}\overline{b}`$ in the infinitely small width approximation, including anomalous couplings in the $`Wtb`$ vertex. The narrow-width approximation enables qualitative interpretations of precise calculations presented later in this study. In Sect. 4 we perform precise tree-level computations in the Standard Model (SM) and in the generalization with the effective $`Wtb`$ vertex. Asymmetries, energy distributions and spin-spin correlations are studied, including the option of electron beam polarization. In Sect. 5 the bounds of the anomalous coupling parameter space, within those no distinction from the SM is possible, are presented and compared with the corresponding limits obtained from single top production processes at NLC in the $`\gamma e`$ mode and LHC in the $`pp`$ mode. ## 2 Effective $`Wtb`$ lagrangian and the anomalous couplings $`f_{2L}`$, $`f_{2R}`$ In the effective lagrangian approach seven gauge invariant and $`CP`$ parity conserving operators of dimension six contribute to the $`Wtb`$ vertex with four independent formfactors. In our analysis we use the effective lagrangian in the unitary gauge as given in $``$ $`={\displaystyle \frac{g}{\sqrt{2}}}[`$ $`W_\mu ^{}\overline{b}(\gamma _\mu f_{1L}P_{}+\gamma _\mu f_{1R}P_+)t`$ (1) $`{\displaystyle \frac{1}{2M_W}}W_{\mu \nu }\overline{b}\sigma ^{\mu \nu }(f_{2R}P_{}+f_{2L}P_+)t]+\mathrm{h}.\mathrm{c}.`$ where $`W_{\mu \nu }=D_\mu W_\nu D_\nu W_\mu ,D_\mu =_\mu ieA_\mu ,P_\pm =1/2(1\pm \gamma _5)`$ and $`\sigma ^{\mu \nu }=i/2(\gamma _\mu \gamma _\nu \gamma _\nu \gamma _\mu )`$. In the SM, the coupling $`f_{1L}`$ is equal to one and the other three couplings, $`f_{1R}`$, $`f_{2L}`$ and $`f_{2R}`$, are equal to zero. The possible (V+A) coupling $`f_{1R}`$ is severely constraint to zero by the CLEO $`bs\gamma `$ data on a level which is stronger than expected even at high energy $`\gamma e`$ colliders. So in the following, we set $`f_{1R}=`$0 and $`f_{1L}=`$1 due to the fact that the (V-A) coupling is as in the SM with the coupling $`V_{tb}`$ very close to unity, as required by present data . This leaves us to perform the analysis only for the two ’magnetic’ anomalous couplings $`f_{2L}`$ and $`f_{2R}`$. The couplings $`f_{2L}`$ and $`f_{2R}`$ are related to the effective couplings $`C_{tW\mathrm{\Phi }}`$ and $`C_{bW\mathrm{\Phi }}`$ in the general effective lagrangian by $`f_{2L(R)}={\displaystyle \frac{C_{t(b)W\mathrm{\Phi }}}{\mathrm{\Lambda }^2}}{\displaystyle \frac{v\sqrt{2}m_W}{g}}`$ (2) where $`\mathrm{\Lambda }`$ is the scale of new physics. Natural values for couplings $`|f_{2L(R)}|`$ are of the order $`\sqrt{m_bm_t}/v`$0.1 . Unitarity limit from $`t\overline{t}`$ scattering at the scale $`\mathrm{\Lambda }=`$1 TeV gives the restriction $`|C_{tW\mathrm{\Phi }}|`$13.5 , or $`|f_{2L(R)}|`$0.65. Expected upgraded Tevatron limits on $`|C_{tW\mathrm{\Phi }}|/(\mathrm{\Lambda }/TeV)^2`$ are $``$2.6 , so the corresponding upper bounds on $`|f_{2L(R)}|`$ are of the order of 0.1-0.2 . In all our calculations which follow the Feynman rules in the momentum space corresponding to the effective lagrangian (1) were implemented in the program package CompHEP . ## 3 Parity violating observables in the <br>top decay It is straightforward to demonstrate by direct calculation that, as mentioned in the introduction, the total rate of the process $`e^+e^{}t\mu ^{}\overline{\nu }_\mu \overline{b}`$ is weakly dependent on the anomalous couplings $`f_{2L}`$ and $`f_{2R}`$. For instance, if $`(f_{2L},f_{2R})`$= (-0.6, 0), the total cross section at $`\sqrt{s}`$= 500 GeV equals 62.7 fb, while the SM value is 63.0 fb. The effect of non-zero $`f_{2L,R}`$ couplings in the amplitude is largely compensated by the increase of the top quark width ($`\mathrm{\Gamma }_{top}`$=1.60 GeV in the standard case and 4.35 GeV at $`(f_{2L},f_{2R})`$=(-0.6, 0)). Hence, the observation of nonstandard interactions is only possible in variables which are sensitive to the effective lagrangian terms (1). It is however a priori not evident which variables provide sufficiently high sensitivity to anomalous $`Wtb`$ operators, so that we are prompted to look, as a first example, for the forward-backward asymmetry of top decay products which is the ratio of integrated single differential distributions. ### 3.1 Forward-backward asymmetry in the infinitely small width approximation In the usual approach to the reaction $`e^+e^{}t\overline{t}`$ 6 fermions, the final state topology is calculated in the approximation of infinitely small top and $`W`$ widths $$\frac{1}{(q^2m^2)^2+m^2\mathrm{\Gamma }^2}=\frac{1}{m\mathrm{\Gamma }}\delta (q^2m^2).$$ (3) Representations of the general expression for distributions in the $`W^+W^{}b\overline{b}`$ final state in terms of the unpolarized $`t\overline{t}`$ cross section $`\mathrm{\Sigma }_{unpol}`$, factorized top-antitop branching ratios, polarization functions $`P`$, $`\overline{P}`$ of the $`t`$, $`\overline{t}`$ and the $`t\overline{t}`$ spin-spin correlation function $`Q`$ can be found in , see also . They can be obtained from the convolution of the $`t\overline{t}`$ production amplitude with the amplitude density matrices of the $`tW^+b`$ and $`\overline{t}W^{}b`$ decays. Following the notations of one gets $`{\displaystyle \frac{d^4\sigma (e^+e^{}t\overline{t}W^+bW^{}\overline{b})}{d\mathrm{cos}\mathrm{\Theta }d\mathrm{cos}\theta d\phi d\mathrm{cos}\theta ^{}d\phi ^{}}}`$ $`={\displaystyle \frac{3\alpha ^2\beta }{32\pi s}}Br(tW^+b)Br(\overline{t}W^{}\overline{b})\mathrm{\Sigma }(\mathrm{\Theta },\theta ,\phi ,\theta ^{},\phi ^{})`$ $`,`$ where $`\mathrm{\Theta }`$ is the top production angle, $`\beta =\sqrt{14m_t^2/s}`$ and $`\mathrm{\Sigma }(\theta ,\phi ,\theta ^{},\phi ^{})`$ $`=`$ $`\mathrm{\Sigma }_{unpol}+kP\mathrm{cos}\theta +\overline{k}\overline{P}\mathrm{cos}\theta ^{}+\mathrm{cos}\theta \mathrm{cos}\theta ^{}k\overline{k}Q`$ $`+(\phi ,\phi ^{}dependentterms).`$ The angles $`\theta ,\phi `$/$`\theta ^{},\phi ^{}`$ define the $`W`$ momentum direction in the rest frame of the top/antitop. The definitions of these angles can be found in the Appendix. In the following, integrations over the azimutal angles $`\phi ,\phi ^{}`$ will be always carried out, with the result that $`\phi ,\phi ^{}`$ dependent terms are equal to zero. The variables $`k`$ and $`\overline{k}`$ are the polarization degree of the top and antitop decay amplitudes. The expressions for $`\mathrm{\Sigma }_{unpol}`$, $`P`$, $`\overline{P}`$ and $`Q`$ in terms of the helicity amplitudes $`\sigma ;h_th_{\overline{t}}`$ for $`t\overline{t}`$ production have the form $`\mathrm{\Sigma }_{unpol}={\displaystyle \frac{1}{4}}{\displaystyle }d\mathrm{cos}\mathrm{\Theta }{\displaystyle \underset{\sigma =\pm }{}}[|\sigma ;++|^2+|\sigma ;+|^2+|\sigma ;+|^2+|\sigma ;|^2]`$ (6) $`P={\displaystyle \frac{1}{4}}{\displaystyle }d\mathrm{cos}\mathrm{\Theta }{\displaystyle \underset{\sigma =\pm }{}}[|\sigma ;++|^2+|\sigma ;+|^2|\sigma ;+|^2|\sigma ;|^2]`$ (7) $`\overline{P}={\displaystyle \frac{1}{4}}{\displaystyle }d\mathrm{cos}\mathrm{\Theta }{\displaystyle \underset{\sigma =\pm }{}}[|\sigma ;++|^2|\sigma ;+|^2+|\sigma ;+|^2|\sigma ;|^2]`$ (8) $`Q={\displaystyle \frac{1}{4}}{\displaystyle }d\mathrm{cos}\mathrm{\Theta }{\displaystyle \underset{\sigma =\pm }{}}[|\sigma ;++|^2|\sigma ;+|^2|\sigma ;+|^2+|\sigma ;|^2]`$ (9) where (see, for instance, ) $`\pm `$ $`=`$ $`(v_L\beta a_L)(1\pm \mathrm{cos}\mathrm{\Theta })`$ (10) $``$ $`=`$ $`\pm {\displaystyle \frac{2m_t}{\sqrt{s}}}v_L\mathrm{sin}\mathrm{\Theta }`$ (11) $`+\pm `$ $`=`$ $`\pm (v_R\beta a_R)(1\mathrm{cos}\mathrm{\Theta })`$ (12) $`+`$ $`=`$ $`\pm {\displaystyle \frac{2m_t}{\sqrt{s}}}v_R\mathrm{sin}\mathrm{\Theta },`$ (13) and $`v_{L,R}`$ and $`a_{L,R}`$ are the standard vector and axial couplings of the $`\gamma `$ and $`Z`$ to the electron and top quark currents. Numerical values of $`P`$, $`\overline{P}`$ and $`Q`$, in units of $`\mathrm{\Sigma }_{unpol}`$, at $`\sqrt{s}`$= 500 GeV and integrated over $`\mathrm{\Theta }`$ are $$\mathrm{\Sigma }_{unpol}:P:\overline{P}:Q=\mathrm{\hspace{0.33em}1}:0.18:\mathrm{\hspace{0.33em}0.18}:0.63$$ (14) Thus, the spin-spin correlation term $`Q`$ in eq.(5) is expected to be significant; it is found to be about four times larger than the polarization function $`P`$. The ratio $`Q/P`$ depends weakly on $`\sqrt{s}`$ in the range from 360 to about 1000 GeV, so that for a $`\mathrm{\Sigma }_{unpol}`$ variation in this energy range by approximately a factor of two to three, our analysis is not critically dependent on $`\sqrt{s}`$. Throughout the paper we have chosen $`\sqrt{s}=`$ 500 GeV. The polarization degrees $`k`$ and $`\overline{k}`$ of the $`t`$ and $`\overline{t}`$ decay amplitudes, summed over the $`W`$ helicity states, are defined by the structure of the $`Wtb`$ vertex. If the spin quantization axis is collinear to the top momentum, the $`tW^+b`$ amplitude polarization density matrix in the rest frame of the top has the form (, see details in the Appendix) $`{\displaystyle \frac{1}{2}}\left(\begin{array}{cc}1+k\mathrm{cos}\theta & k\mathrm{sin}\theta e^{i\phi }\\ k\mathrm{sin}\theta e^{i\phi }& 1k\mathrm{cos}\theta \end{array}\right)`$ (17) The explicit expression for the polarization degree $`k`$ for the $`tW^+b`$ decay can be obtained in models with the general effective lagrangian (1) by means of the eight helicity amplitudes of the top decay defined in the Appendix. In the case $`f_{1L}`$=1, $`f_{1R}`$=0 we get $`k={\displaystyle \frac{(\frac{m_t}{m_W}+f_{2L})^22(1+\frac{m_t}{m_W}f_{2L})^2(12(\frac{m_t}{m_W})^2)f_{2R}^2}{(\frac{m_t}{m_W}+f_{2L})^2+2(1+\frac{m_t}{m_W}f_{2L})^2+(1+2(\frac{m_t}{m_W})^2)f_{2R}^2}}`$ (18) The expressions (4), (5), (16) provide the basis for a qualitative understanding of the results from exact matrix element Monte Carlo calculations, when nonzero bottom quark mass and finite top quark and $`W`$-boson widths are accounted for. It follows from (5) that natural integrated angular observables are the $`b`$-quark and the lepton forward-backward asymmetries, measured in the rest frame of the top. It is straightforward to show that these asymmetries have the form $$A_{FB}=\frac{\sigma (\theta <90^{})\sigma (\theta >90^{})}{\sigma (\theta <90^{})+\sigma (\theta >90^{})}=\frac{k}{2}\frac{P}{\mathrm{\Sigma }_{unpol}}$$ (19) For the $`b`$-quark, the polarization degree $`k`$ in the SM equals 0.41 (see the Appendix) and the ratio $`P/\mathrm{\Sigma }_{unpol}`$ (the degree of longitudinal top quark polarization integrated over $`\mathrm{\Theta }`$) is equal to 0.18 <sup>1</sup><sup>1</sup>1The P dependence on the top production angle can be found in . at $`\sqrt{s}=`$ 500 GeV. Hence, $`A_{FB}^b`$ in the infinitely small width approximation equals 3.6%, while for the lepton from $`W`$ decay, with $`k`$=1 in the SM, $`A_{FB}^l=`$-9.0%. ### 3.2 Infinitely small $`W`$ width approximation in the top quark anomalous decay The effective lagrangian terms of the $`Wtb`$ vertex can significantly change the top quark two-body and three-body decay widths if the anomalous couplings $`f_{2L}`$, $`f_{2R}`$ are sufficiently large. Whether however finite $`W`$ width corrections substantially obscure effects of anomalous couplings demands to investigate computations done within approximation (3) and to reveal its relation to exact Breit-Wigner propagator calculations. In order to quantify this question we performed an explicit symbolic calculation of the factorized branching ratios in formula (4). The result for the two-body top decay width can be obtained from the helicity amplitudes (30)-(37): $`\mathrm{\Gamma }_2(tW^+b)={\displaystyle \frac{G_Fm_t^3}{8\sqrt{2}\pi }}(1r^2)^2[1+2r^2+6f_{2L}r+(f_{2L}^2+f_{2R}^2)(2+r^2)]`$ (20) where $`r=m_W/m_t`$. The three-body top decay width, after integration of the symbolic expression over the Dalitz plot, is given by $`\mathrm{\Gamma }_3(te^+\nu _eb)=`$ $`{\displaystyle \frac{G_F^2m_t^3m_W^2}{96\pi ^3}}[F_1{\displaystyle \frac{m_W}{\mathrm{\Gamma }_W}}(\pi \mathrm{arctan}{\displaystyle \frac{m_t^2\mathrm{\Gamma }_W}{m_W(m_t^2m_W^2\mathrm{\Gamma }_W^2)}})`$ $`+F_2\mathrm{log}{\displaystyle \frac{m_W^2(m_W^2+\mathrm{\Gamma }_W^2)}{(m_t^2m_W^2)^2+m_W^2\mathrm{\Gamma }_W^2}}+F_3]`$ where $`F_1`$ $`=`$ $`13r^4+2r^6+3r^2\gamma ^26r^4\gamma ^2`$ $`6f_{2L}(2r^3r^5\gamma 2r\gamma ^2+3r^3\gamma ^2)`$ $`+(f_{2L}^2+f_{2R}^2)(23r^2+r^6+3\gamma ^26r^4\gamma ^2+r^2\gamma ^4)`$ $`F_2`$ $`=`$ $`3r^43r^6+r^4\gamma ^23f_{2L}(r4r^3+3r^5r^3\gamma ^2)`$ $`+(f_{2L}^2+f_{2R}^2)(1+3r^22r^6+2r^4\gamma ^2)`$ $`F_3`$ $`=`$ $`3rf_{2L}(34r^2)+(f_{2L}^2+f_{2R}^2)({\displaystyle \frac{1}{3}}+r^2+3r^4r^2\gamma ^2)`$ and $`\gamma =\mathrm{\Gamma }_W/m_t`$. If we set $`f_{2L}=f_{2R}=0`$ and use the approximation $`F_1=13r^4+2r^6`$ and neglect $`\mathrm{\Gamma }_W/m_t`$ power terms, we obtain by comparing (18) and (19) the explicit narrow width factorization in the SM case: $`\mathrm{\Gamma }_3(te^+\nu _eb)={\displaystyle \frac{G_F^2m_t^3m_W^3}{96\pi ^3}}{\displaystyle \frac{\pi }{\mathrm{\Gamma }_W}}(13r^4+2r^6)=`$ $`={\displaystyle \frac{G_Fm_t^3}{8\sqrt{2}\pi }}(1r^2)^2(1+2r^2){\displaystyle \frac{1}{\mathrm{\Gamma }_W}}{\displaystyle \frac{G_Fm_W^3}{6\sqrt{2}\pi }}=\mathrm{\Gamma }_2(tW^+b)Br(W^+e^+\nu _e)`$ Since the $`W`$ branching factorisation (20) is in general not valid, its violation by $`f_{2L,R}\mathrm{\Gamma }_W/m_t`$ and $`f_{2L,R}m_W/m_t`$ power terms is only weak provided modulus of $`f_{2L}`$ and $`f_{2R}`$ are around or less than 1. More details about the precision of the factorization approximation can be obtained from Fig.1, where the ratio $`\mathrm{\Gamma }_3/(\mathrm{\Gamma }_2Br(W^+e^+\nu _e))`$ as a function of the anomalous couplings $`f_{2L}`$ and $`f_{2R}`$ is shown. Clearly, the accuracy of $`\mathrm{\Gamma }_3`$ within the infinitely small $`W`$ width approximation is convincing in the range considered for $`f_{2L}`$ and $`f_{2R}`$; deviations are expected to be of the order 1% or less. Thus, calculations done within the approximation (3) imply small corrections which are less important than e.g. interferences between the signal diagrams (see below). In general, however, careful investigations are appropriate when anomalous top quark decay calculations are carried out within the infinitely small $`W`$ width approximation. ## 4 Tree-level results for $`e^+e^{}t\overline{t}tl\overline{\nu }_l\overline{b}`$ If precise measurements of top decay products are envisaged, it is demanding to know the SM predictions with very high accuracy. The program package CompHEP which performs analytic calculations of the matrix element squared, generates an optimized FORTRAN code and generates an event flow, overcomes the shortcomings due to infinitely small width and zero fermion mass approximations. Furthermore, it allows to include all diagrams of the irreducible background and their interferences. In the case of the signal process, $`e^+e^{}t\overline{t}tl\overline{\nu }_l\overline{b}`$, only two diagrams and their interference exist. If the anomalous couplings $`f_{2L}`$ and $`f_{2R}`$ are allowed to contribute to the lagrangian, the corresponding Feynman rules implemented in CompHEP can be found in the Appendix of the second ref. in . ### 4.1 Forward-backward asymmetries It follows from eqs.(4) and (5) that observables of experimental interest are the distributions in $`\theta `$, $`\theta ^{}`$ for the $`b`$-quark and lepton in the top rest frame, or in the $`e^+e^{}`$ center-of-mass system (c.m.s.) for a more general discussion. From these distributions the forward-backward asymmetry (17) can be easily calculated in the SM and in models extended by anomalous couplings. In a first step we compare precise calculations with results obtained from the infinitely small width approximation (3), within the SM ($`f_{2L}=f_{2R}=`$0). Numerical values for the $`b`$-quark and lepton asymmetries, $`A_{FB}^b`$ and $`A_{FB}^l`$, calculated by means of CompHEP, are shown in Table 1, in the top rest frame as well as in the $`e^+e^{}`$ c.m.s. If compared with the top rest frame asymmetries obtained within the narrow width approximation of sect.3.1, one notices a 15% difference for the $`b`$-quark, whereas for the lepton the difference is negligible. Thus, already this example demonstrates the importance of precise calculations which include interference terms, finite width and non-zero mass contributions. The SM forward-backward $`b`$-quark asymmetries (Table 1) are significantly larger in the $`e^+e^{}`$ c.m.s. than in the top rest frame, while for the lepton such differences are less evident. This observation can be understood by recalling that the $`t`$ ($`\overline{t}`$) is produced mainly in the $`e^{}`$ ($`e^+`$) direction with left (right) helicity and, in the top decay, the lepton ($`b`$ quark) is emitted preferrably in the direction of (in the opposite direction to) the top spin. It is also worth to mention that irreducible background, which might remain after any $`t\overline{t}`$ selection procedure, should be carefully accounted for. CompHEP calculation shows that if the electron being the lepton in the final state (with 18 contributing diagrams in total) forward electrons from $`t`$-channel photon exchange alter significantly $`A_{FB}^e`$ compared to only signal diagrams calculations. When we allow for anomalous $`Wtb`$ couplings, the asymmetry $`A_{FB}^b`$, measured in the top rest frame, is shown in Fig. 2 in the narrow width approximation and for exact calculations. The qualitative behaviour of both asymmetries as a function of $`f_{2L}`$ and $`f_{2R}`$ is very similar; only close inspections reveal significant differences. Furthermore, $`A_{FB}^b`$ depends stronger on $`f_{2L}`$ than on $`f_{2R}`$, as can be better seen in Fig.3a, where also two standard exclusion contour plots are shown for 100 fb<sup>-1</sup> and 500 fb<sup>-1</sup> integrated luminosities. This greater sensitivity is directly connected to the stronger influence of the linear $`f_{2L}`$ term in (16) than that of the quadratic $`f_{2R}`$ term, which in turn is an inherent property of the helicity amplitudes of anomalous top quark decays, as outlined in the Appendix. The impact of the anomalous couplings to the lepton forward-backward asymmetry $`A_{FB}^l`$ is less important both in the $`e^+e^{}`$ c.m.s. and the top rest frame (see Fig.3b), and being only indirect due to the presence of the standard left current $`W`$ boson decay. However, in contrast to the two-fold ambiguity of the b-quark asymmetry (Fig. 3a), $`A_{FB}^l`$ is unique in the sense that for a given $`f_{2R}`$ value only one $`f_{2L}`$ range (shaded) exists, in which no distinction from the SM (within $`2\sigma `$) is possible. Table 1 contains some numerical examples of $`A_{FB}^{b/l}`$ for several $`f_{2L}`$ and $`f_{2R}`$ values at $`\sqrt{s}=`$500 GeV, measured in both reference frames discussed so far. Clearly, the largest forward-backward asymmetry is obtained for the $`b`$ quark if measured in the $`e^+e^{}`$ c.m.s. One should however remember that $`e^+e^{}`$ c.m.s. asymmetries are a superposition of production and decay asymmetries, while asymmetries reconstructed in the top rest frame can be considered as a ’pure’ effect. Left electron beam polarization not only increases the $`t\overline{t}`$ production rate by a factor of about three ($`\sigma _{tot}=`$176 fb for 100% left polarized electrons at $`\sqrt{s}=`$500 GeV) but also enhances $`A_{FB}^{b/l}`$ by a factor of 2-3 in the top rest frame. At the same time, the $`f_{2L}`$ sensitivity of $`A_{FB}^b`$ increases most significantly if measured in the $`e^+e^{}`$ c.m.s. (see the examples in Table 1). Besides the study of the $`\theta `$ and $`\theta ^{}`$ decay angular distributions, more sophisticated angular observables were proposed to study parity violating effects: (1) the angle between the lepton momentum in the $`W`$ rest frame and the momentum of the top in the $`e^+e^{}`$ c.m.s. and (2) the angle between the top production plane and the production plane of the $`b`$-quark (or the lepton) in the $`e^+e^{}`$ c.m.s.: $$\mathrm{cos}\theta _{tb}=\frac{([\mathrm{𝐤𝐭}][\mathrm{𝐤𝐛}])}{|[\mathrm{𝐤𝐭}]||[\mathrm{𝐤𝐛}]|},$$ (23) where $`𝐤`$ is a unit length vector in the $`e^{}`$ direction. A similar variable was proposed in to measure the transverse quark polarization. In both angular distributions very large asymmetries (up to 95%) exist. However, their sensitivity to the anomalous $`Wtb`$ couplings $`f_{2L}`$ and $`f_{2R}`$ is very small and is, in good approximation, independent on the electron polarization. ### 4.2 Energy spectrum asymmetry Besides angular distributions, energy spectra of the top decay products may also possess high sensitivity to anomalous couplings. In this section we study the asymmetry of the lepton energy spectrum defined in the top rest frame using the dimensionless variable $`x_\mu =2E_\mu /m_{top}`$: $`A_E^\mu ={\displaystyle \frac{\sigma (x_\mu <0.5)\sigma (x_\mu >0.5)}{\sigma (x_\mu <0.5)+\sigma (x_\mu >0.5)}}`$ (24) Fig.3c shows the results for $`A_E^\mu `$ from the process $`e^+e^{}t\mu \overline{\nu }_\mu \overline{b}`$ as a function of $`f_{2L}`$ and $`f_{2R}`$. As can be seen, $`A_E^\mu `$ is significantly more sensitive to $`f_{2L}`$ than its forward-backward asymmetry $`A_{FB}^\mu `$, and this result is independent whether $`A_E^\mu `$ is measured in the top rest frame or in the $`e^+e^{}`$ c.m.s. Alike to $`A_{FB}^b`$, the sensitivity to $`f_{2R}`$ is somewhat less pronounced than to $`f_{2L}`$ and the ambiguity exists also. The $`b`$-quark energy spectrum in the top rest frame has a resonance peak at $`x_b=1(m_W/m_t)^2`$, resulting to an energy asymmetry insensitive to anomalous couplings. If the neutrino is used as the analyser (by means of the missing energy technique), its energy asymmetry is slightly less sensitive to $`f_{2L}`$ and $`f_{2R}`$ than the lepton energy asymmetry. Whether the neutrino is at all usable for precise measurements, requires however detailed experimental studies including full event simulation and reconstruction. As clearly visible from Figs.3a-c, only the combination of forward-backward and energy asymmetry measurements results to an allowed region (Fig.4, the sum of the grey and dark areas) much smaller than that for each measurement alone and ensures a significant improvement of the sensitivity on anomalous couplings, with limits for $`f_{2L}`$ and $`f_{2R}`$ sensible luminosity dependent. ### 4.3 Spin-spin asymmetries The spin correlations and the spin-spin asymmetries, which are related to each other, are in general double differential distributions where one of the variables is integrated over a certain kinematical region. As already mentioned in sect. 3, the spin correlation term $`k\overline{k}Q`$ in (5) is comparable to the polarization term $`kP`$ and therefore spin-spin correlations, although suppressed by the additional power of $`k`$, are expected to be not small. For instance, the forward-backward asymmetry of the $`b`$-quark measured in the top rest frame, under the condition that the $`\overline{b}`$-quark is observed only in the forward hemisphere in the $`\overline{t}`$ rest frame, can be derived from (5) as $$A_s^b=\frac{k}{2}\frac{P}{\mathrm{\Sigma }_{unpol}}(1+\overline{k}\frac{Q}{2P}).$$ (25) For simplicity, the term $`\overline{k}\overline{P}/\mathrm{\Sigma }`$1 is omitted here. At e.g. $`\sqrt{s}=`$500 GeV, we find $`A_s^b=`$0.062 within the SM, which is two times larger than $`A_{FB}^b=`$0.036, when no restriction is imposed on the $`\overline{t}`$ side. If anomalous couplings are allowed to contribute, the dependence of the spin-spin asymmetry, $`A_s^b`$, on $`f_{2L},f_{2R}`$ is shown in Fig.5, within the narrow width approximation. Clearly, improved constraints on $`f_{2L},f_{2R}`$ can be obtained from $`A_s^b`$ compared to the unrestricted forward-backward asymmetries as discussed in sect. 4.1. However, since $`A_s^b`$ is calculated in the infinitely small width approximation <sup>2</sup><sup>2</sup>2For reasons of unsufficient computer memory, CompHEP $`26`$ calculations are only possible for the SM $`Wtb`$ vertex, with the result $`A_s^b=`$0.049 at $`\sqrt{s}=`$500 GeV., the reliability of the results needs more careful investigation. We expect some further enhancement of parity-violating effects by using polarized beams. If e.g. 100% left polarized electrons collide with unpolarized positrons, forward $`t`$ (backward $`\overline{t}`$) quarks (w.r.t. the $`e^{}`$ direction) are mainly produced in the helicity configuration L (R), while backscattered $`t`$ (forward $`\overline{t}`$) are produced in the helicity configuration R (L) (their production angular behaviour can be found in ). As a consequence, the $`t`$ and $`\overline{t}`$ decay products in the reaction $`e^+e^{}t\overline{t}(e^{}\nu _e\overline{b})(u\overline{d}b)`$ are expected to be strongly correlated, in so far as the $`e^{}`$ and the $`\overline{d}`$ are produced mainly in the top spin direction, while the u- and b-quarks prefer production in the opposite direction . Using CompHEP to calculate the exact $`26`$ SM amplitudes, we obtain the electron decay angular distribution in the top rest frame under the condition that the $`\overline{d}`$ decay angle in the antitop rest frame is less than 90, for the top produced either forward (Fig. 6a) or backward (Fig. 6b) in the $`e^+e^{}`$ c.m.s. The spin-spin asymmetries in these two cases are -0.258 and -0.096, respectively, demonstrating strong sensitivity to parity-violating effects when beam polarization is available. QCD corretions to the spin correlations in $`t\overline{t}`$ production are expected to be in general small , but their inclusion is recommended in searches for nonstandard interactions. ## 5 Conclusions The total rate of the reaction $`e^+e^{}t\overline{t}`$ 6 fermions at NLC energies is negligibly affected by the anomalous lagrangian terms (1). Hence, it is straightforward to use single and double differential distributions of the top/antitop decay products to eventually observe effects due to anomalous $`Wtb`$ operators, and as larger their sensitivity as stronger limits on anomalous couplings can be imposed. In this paper we investigate forward-backward asymmetries for the b-quark and the lepton in the top rest frame or in the $`e^+e^{}`$ c.m.s., the energy asymmetry of the lepton in the top rest frame and the spin-spin asymmetry in the $`t`$/$`\overline{t}`$ decay. Precise tree-level Monte Carlo calculations for the signal diagrams and their interference were performed and compared in the case of forward-backward asymmetries with the symbolic expressions obtained in the infinitely small width and zero fermion mass approximation. We realized that in general careful investigations are appropriate when such approximations are intended to be used in analyses of multiparticle final state topologies. Concerning the sensitivity of the observables considered in this study we found that (a) $`A_{FB}^b`$, $`A_{FB}^l`$ and $`A_E^l`$ have stronger sensitivity to $`f_{2L}`$ than to $`f_{2R}`$, as seen in Fig.3 <sup>3</sup><sup>3</sup>3 For $`f_{1R}=`$0, the helicity amplitudes (30)-(37) have linear and quadratic terms in $`f_{2L}`$ and only quadratic terms in $`f_{2R}`$.; (b) the sensitivity of the forward-backward $`b`$-quark asymmetry (Fig.3a) is larger than the sensitivity of the lepton forward-backward asymmetry (Fig.3b), which is somewhat degraded due to the subsequent $`W`$ decay; (c) it is important to note that $`A_{FB}^l`$ resolves the ambiguity observed in $`A_{FB}^b`$ and $`A_E^l`$; (d) the lepton energy asymmetry has the largest sensitivity on $`f_{2L}`$ and $`f_{2R}`$ (Fig.3c). In summary, it turns out that particle orientations seem to be less sensitive to anomalous $`Wtb`$ operators than particle energies. As indicated by the 2$`\sigma `$ exclusion contour plots in Figs. 3a-c, no satisfactory restriction on $`f_{2L}`$ and $`f_{2R}`$ has been obtained for each variable alone. But their combined annulus (Fig. 4) allows significant improvements of the sensitivity on anomalous couplings. If in addition the spin-spin asymmetry of Fig. 5, although calculated within the narrow width approximation, is included, further restrictions on anomalous $`Wtb`$ operators are possible for 100 fb<sup>-1</sup> (dark area in Fig.4), while for 500 fb<sup>-1</sup> no improvements are observed. Thus, for the high luminosity option of the TESLA linear collider the bounds on the anomalous couplings $`f_{2L}`$ and $`f_{2R}`$, within those no distinction from the SM is possible, are \[-0.025, 0\] for $`f_{2L}`$ and $`\pm 0.20`$ for $`f_{2R}`$. These rather promising results demonstrate the reliability of the top pair production process in $`e^+e^{}`$ collisions to probe the $`Wtb`$ vertex. It is interesting to compare these limits with the expectations from single top production processes at LHC . The LHC limitations, being 2-3 times better than the possible restrictions from the upgraded Tevatron, are comparable to the $`e^+e^{}`$ LC estimates provided the LHC systematic uncertainties are controlled at a level better than about 10%. The advantage of the LHC to measure the single top production rates in the $`Wb\overline{b}`$/$`Wb\overline{b}+jet`$ channels is however somewhat degraded by relatively large uncertainties in the absolute normalization of the cross sections and the presence of reducible background not easy to control. In the clean environment of $`e^+e^{}`$ collisions, the selection of $`t\overline{t}`$ events is thought to be very reliable and further improvements in probing the $`Wtb`$ vertex can be expected if additional sensitive observables are included in the analysis and electron beam polarization is used. Whether however the superior sensitivity to the anomalous couplings $`f_{2L}`$ and $`f_{2R}`$ of a linear collider in the $`\gamma e`$ mode at high energies ($`\sqrt{s_{e\gamma }}`$1 TeV) could be achieved or even superseded, remains open for future studies. Acknowledgments E.B. and M.D. are grateful to DESY-Zeuthen for hospitality. M.D. thanks very much H.S.Song for useful discussions. The work of E.B. and M.D. was partially supported by the RFBR-DFG grant 99-02-04011, the CERN-INTAS grant and the KCFE grant (SPb). ## Appendix The helicity amplitudes for the decay $`tW^+b`$ in models with the general interaction lagrangian (1) can be found in . Our calculation follows the formalism of , where the chiral representation for the gamma matrices is used. The four component spinors can be split into two component helicity eigenstates $`\chi _\lambda (p)`$ $`u(p,\lambda )_\pm =\omega _{\pm \lambda }(p)\chi _\lambda (p)`$ (26) $`v(p,\lambda )_\pm =\pm \lambda \omega _\lambda (p)\chi _\lambda (p),`$ where $`\omega _\pm (p)=\sqrt{E\pm p}.`$ In the rest frame of the top, the helicity eigenstates of the $`b`$-quark can be written in the form $`\chi _+(p_b)=\left(\begin{array}{c}\mathrm{𝚜𝚒𝚗}\frac{\theta }{2}\\ \mathrm{𝚌𝚘𝚜}\frac{\theta }{2}e^{i\phi }\end{array}\right),\chi _{}(p_b)=\left(\begin{array}{c}\mathrm{𝚌𝚘𝚜}\frac{\theta }{2}e^{i\phi }\\ \mathrm{𝚜𝚒𝚗}\frac{\theta }{2}\end{array}\right)`$ (31) with the following component representation of $`W`$ and $`b`$ momenta in the spherical coordinate system $`p_W`$ $`=`$ $`\{E_W,|𝐩_W|\mathrm{𝚜𝚒𝚗}\theta \mathrm{𝚌𝚘𝚜}\phi ,|𝐩_W|\mathrm{𝚜𝚒𝚗}\theta \mathrm{𝚜𝚒𝚗}\phi ,|𝐩_W|\mathrm{𝚌𝚘𝚜}\theta \}`$ (32) $`p_b`$ $`=`$ $`|𝐩_b|\{1,\mathrm{𝚜𝚒𝚗}\theta \mathrm{𝚌𝚘𝚜}\phi ,\mathrm{𝚜𝚒𝚗}\theta \mathrm{𝚜𝚒𝚗}\phi ,\mathrm{𝚌𝚘𝚜}\theta \}`$ (33) The polarization vectors of the $`W`$ boson can be taken in the form $`ϵ_+`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\{0,\mathrm{𝚌𝚘𝚜}\theta \mathrm{𝚌𝚘𝚜}\phi +i\mathrm{𝚜𝚒𝚗}\phi ,\mathrm{𝚌𝚘𝚜}\theta \mathrm{𝚜𝚒𝚗}\phi i\mathrm{𝚜𝚒𝚗}\phi ,\mathrm{𝚜𝚒𝚗}\theta \}`$ (34) $`ϵ_{}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\{0,\mathrm{𝚌𝚘𝚜}\theta \mathrm{𝚌𝚘𝚜}\phi +i\mathrm{𝚜𝚒𝚗}\phi ,\mathrm{𝚌𝚘𝚜}\theta \mathrm{𝚜𝚒𝚗}\phi i\mathrm{𝚌𝚘𝚜}\phi ,\mathrm{𝚜𝚒𝚗}\theta \}`$ $`ϵ_0`$ $`=`$ $`{\displaystyle \frac{E_W}{m_W}}\{{\displaystyle \frac{|𝐩_W|}{E_W}},\mathrm{𝚜𝚒𝚗}\theta \mathrm{𝚌𝚘𝚜}\phi ,\mathrm{𝚜𝚒𝚗}\theta \mathrm{𝚜𝚒𝚗}\phi ,\mathrm{𝚌𝚘𝚜}\theta \}`$ In the symbolic calculations we always neglect the $`b`$-quark mass. The eight helicity amplitudes $`\sqrt{2E_bm_t}h_t,h_W,h_b`$ corresponding to the matrix element of the top decay $`{\displaystyle \frac{g}{\sqrt{2}}}\overline{u}(p_b)\mathrm{\Gamma }_\mu u(p_t)\epsilon _\mu ^{}(p_W)`$ with $`\mathrm{\Gamma }_\mu `$ $`=`$ $`f_{1L}\gamma _\mu (1\gamma _5)+f_{1R}\gamma _\mu (1+\gamma _5)`$ $`+{\displaystyle \frac{f_{2L}}{2m_W}}(\widehat{p}_W\gamma _\mu \gamma _\mu \widehat{p}_W)(1+\gamma _5)+{\displaystyle \frac{f_{2R}}{2m_W}}(\widehat{p}_W\gamma _\mu \gamma _\mu \widehat{p}_W)(1\gamma _5)`$ can be calculated in the rest frame of the top using (24)-(28): $`,0,`$ $`=`$ $`({\displaystyle \frac{m_t}{m_W}}f_{1L}+f_{2L})\mathrm{𝚜𝚒𝚗}{\displaystyle \frac{\theta }{2}}`$ (36) $`,,`$ $`=`$ $`\sqrt{2}(f_{1L}+{\displaystyle \frac{m_t}{m_W}}f_{2L})\mathrm{𝚌𝚘𝚜}{\displaystyle \frac{\theta }{2}}`$ (37) $`+,0,`$ $`=`$ $`({\displaystyle \frac{m_t}{m_W}}f_{1L}+f_{2L})\mathrm{𝚌𝚘𝚜}{\displaystyle \frac{\theta }{2}}e^{i\phi }`$ (38) $`+,,`$ $`=`$ $`\sqrt{2}(f_{1L}+{\displaystyle \frac{m_t}{m_W}}f_{2L})\mathrm{𝚜𝚒𝚗}{\displaystyle \frac{\theta }{2}}e^{i\phi }`$ (39) $`+,0,+`$ $`=`$ $`({\displaystyle \frac{m_t}{m_W}}f_{1R}+f_{2R})\mathrm{𝚜𝚒𝚗}{\displaystyle \frac{\theta }{2}}`$ (40) $`+,+,+`$ $`=`$ $`\sqrt{2}(f_{1R}+{\displaystyle \frac{m_t}{m_W}}f_{2R})\mathrm{𝚌𝚘𝚜}{\displaystyle \frac{\theta }{2}}`$ (41) $`,0,+`$ $`=`$ $`({\displaystyle \frac{m_t}{m_W}}f_{1R}+f_{2R})\mathrm{𝚌𝚘𝚜}{\displaystyle \frac{\theta }{2}}e^{i\phi }`$ (42) $`,+,+`$ $`=`$ $`\sqrt{2}(f_{1R}+{\displaystyle \frac{m_t}{m_W}}f_{2R})\mathrm{𝚜𝚒𝚗}{\displaystyle \frac{\theta }{2}}e^{i\phi }`$ (43) The sum of the eight helicity amplitudes squared gives the total decay width of the top (18) for the general interaction lagrangian (1). If $`f_{1R}`$=0, the width $`\mathrm{\Gamma }(tW^+b)`$ contains a linear term in $`f_{2L}`$ and quadratic terms in both $`f_{2R}`$ and $`f_{2L}`$. The eight helicity amplitudes of the antitop decay, $`\overline{t}W^{}\overline{b}`$, can be obtained from (30)-(37) by the replacements $`f_{1L}f_{1R}`$ and $`f_{2L}f_{2R}`$ (only real $`f`$ are considered). In the Standard Model (SM) with $`f_{1L}`$=1, $`f_{1R}=f_{2L}=f_{2R}`$=0 only four nonvanishing helicity amplitudes remain . If $`(\theta ,\phi )`$ are the polar and azimutal angles of the $`b`$-quark with respect to the top momentum, the helicity amplitudes of the top with spin up and spin down, $`a_1`$ and $`a_2`$, allow to define the $`tW^+b`$ amplitude polarization density matrix . This matrix is different from the polarization density matrix defined by the individual top spin function. The squared sum of (32)-(35) gives the probability of spin up top decay, while the probability of spin down top decay is given by the squared sum of (30)-(31) and (36)-(37). The polarization density matrix can be defined for the ($`a_1`$,$`a_2`$) spin function $$\rho =\left(\begin{array}{cc}|a_1|^2& a_1a_2^{}\\ a_1^{}a_2& |a_2|^2\end{array}\right),$$ (44) where the normalised SM components derived from (30)-(33) have the form $`a_1^2`$ $`=`$ $`{\displaystyle \frac{M_W^2}{m_t^2+2m_W^2}}({\displaystyle \frac{m_t^2}{m_w^2}}\mathrm{𝚌𝚘𝚜}^2{\displaystyle \frac{\theta }{2}}+2\mathrm{𝚜𝚒𝚗}^2{\displaystyle \frac{\theta }{2}})`$ $`a_2^2`$ $`=`$ $`{\displaystyle \frac{M_W^2}{m_t^2+2m_W^2}}({\displaystyle \frac{m_t^2}{m_w^2}}\mathrm{𝚜𝚒𝚗}^2{\displaystyle \frac{\theta }{2}}+2\mathrm{𝚌𝚘𝚜}^2{\displaystyle \frac{\theta }{2}})`$ $`a_1a_2^{}=a_1^{}a_2`$ $`=`$ $`{\displaystyle \frac{m_t^2}{m_W^2}}\mathrm{𝚜𝚒𝚗}^2{\displaystyle \frac{\theta }{2}}+2\mathrm{𝚌𝚘𝚜}^2{\displaystyle \frac{\theta }{2}}.`$ The amplitude polarization density matrix (38) can also be represented in the standard form $`\rho ={\displaystyle \frac{1}{2}}\left(\begin{array}{cc}1+k\mathrm{𝚌𝚘𝚜}\theta & k\mathrm{𝚜𝚒𝚗}\theta e^{i\phi }\\ k\mathrm{𝚜𝚒𝚗}\theta e^{i\phi }& 1k\mathrm{𝚌𝚘𝚜}\theta \end{array}\right)={\displaystyle \frac{I}{2}}+𝐏\widehat{𝐒}`$ (47) $`={\displaystyle \frac{1}{2}}[I+k(\mathrm{𝚜𝚒𝚗}\theta \mathrm{𝚌𝚘𝚜}\phi \sigma _1+\mathrm{𝚜𝚒𝚗}\theta \mathrm{𝚜𝚒𝚗}\phi \sigma _2+\mathrm{𝚌𝚘𝚜}\theta \sigma _3)],`$ where $`\widehat{𝐒}=\{\sigma _\mathrm{𝟏},\sigma _\mathrm{𝟐},\sigma _\mathrm{𝟑}\}`$ is the spin operator. The polarization vector $`𝐏`$ is collinear to the $`b`$-quark momentum and the absolute value of the polarization vector, also called the polarization degree, is defined by the matrix element of the $`tW^+b`$ decay. In the SM $`k={\displaystyle \frac{m_t^22m_W^2}{m_t^2+2m_W^2}}=0.41.`$ In the case of a general interaction lagrangian the polarization degree depends on $`f_{2L,R}`$ (see (16) in sect.3.1).
warning/0001/hep-th0001190.html
ar5iv
text
# 1 Introduction ## 1 Introduction S-duality has emerged as a self-equivalence structure of the IIB superstring and evidence has accumulated for its existence. Structures in string theories have in the past yielded a better understanding of aspects of supersymmetric field theories. In this work we shall explore a manifestly S-dual compatible, perturbative and non-perturbative, expansion of the ten-dimensional S-matrix for graviton scattering. This S-matrix is highly constrained by modular forms, and in agreement with known results in the low energy limit. Recent work on the effective action in the low-energy limit of the superstring has revealed the modular property of the S-duality invariant graviton scattering to several orders in derivatives , in eleven-dimensional supergravity , and for additional couplings in IIB superstring theory in . We examine the same structure at all orders in derivatives and generate these terms through the use of modular forms, on the $`SL(2,Z)`$ fundamental domain for ten and nine-dimensional theories, and on more complicated domains for lower-dimensional compactified theories. In related work there have been $`SL(2,Z)`$ based reformulations at the level of the world-sheet action which might lead to a similar description that produces higher derivative corrections compatible with S-duality of the IIB superstring as for the $`R^4`$ term . Tests of S-duality at the amplitude level require knowledge of the perturbative expansion for graviton scattering at genus greater than one, and is formidable. The S-duality of the superstring imposed on the S-matrix leads to non-trivial predictions for its structure, and predicts contributions from the genus expansion without performing string perturbation theory. Standard perturbative expansions involve expansions in the coupling constant, or in conjunction with large $`N`$; however, Feynman diagrams are typically difficult at high loop order and alternative approaches are useful. The derivative expansion orderwise must be compatible with the non-perturbative symmetry structures that exchange weak and strong coupling and must obey these invariance properties in IIB superstring theory; possibly this re-ordering avoids summation problems, because of factorial dependence in the diagram expansion, as it is non-perturbative in the coupling and should thus take into account the solitonic contributions. Such an expansion is necessarily non-perturbative from the point of view of the microscopic coupling constant in the maximally supersymmetric theory, although perturbative in the energy scale for theories without dimensional transmutation. M-theory in the eleven dimensional limit requires such an approximation, because there is no dilatonic coupling constant in this eleven dimensional corner. IIB superstring theory has IIB supergravity describing its zero-mode degrees of freedom. Compactification to lower dimensions through tori gives rise to several supergravity theories including the example of maximally extended $`N=8`$ supergravity in four dimensions . The remnant of S-duality, and U-duality in general , imposes severe constraints on the perturbative expansions of the supergravity theories in various dimensions: Primarily, we are interested in the finiteness properties of the latter theories in this regard. Recently, $`N=8`$ supergravity in four dimensions has been re-examined, and there is strong evidence that the previously thought first primitive divergence of the four-point amplitude does not occur at three-loops, but at higher order . In specifying the supergravity quantum theory from the low-energy limit of the superstring, a regulator must be chosen that is compatible with the global symmetries of the field equations; on the other hand, superstring theory points to a specific regulator, the one in which S-duality remains as a remnant on the massless degrees of freedom. At one-loop this is most easily seen in comparing the SUGRA expansion with the string and its non-perturbative structure. Further divergence nullifications beyond the known properties of N=8 supergravity require a field theory mechanism. Recent work in has shown that additional cancellations not accounted for in superspace powercounting arguments occur. In the explicit construction of the two-loop amplitudes these additional cancellations follow in the cut construction from use of on-shell supersymmetry Ward identities. Alternatively, this cancellation occurs because of S-duality, where this structure permeates to higher order and in various dimensions. In this work we shall reformulate the expansion of the IIB superstring graviton scattering using constraints imposed on automorphic functions. A uniqueness theorem regarding the functional form incorporating additional properties of the scattering elements potentially allows, given S-duality, a route to computing complete S-matrix elements without perturbative string theory. A further aspect of the modular construction in terms of Eisenstein functions implies a truncation property in perturbative supergravity defined by the toroidal compactification of IIB in ten dimensions (or $`d=11`$ supergravity). This work is organized as follows. In section 2 we examine the general structure of the low-energy limit of the graviton scattering and the constraints of S-duality in the uncompactified IIB theory. In section 3 we compare the S-matrix with definitions of the quantum effective action. In section 4 we give relevant properties of the Eisenstein series used in the construction outlined in section 2. In section 5 we analyze toroidal compactifications to lower dimensions and the U-duality structure together with the $`SL(2,Z)`$ subgroup; the emphasis in this section is on finding further truncations. In section 6 we extract implications for the graviton scattering in the field theory limit. In the last section we give conclusions and discuss extensions related to this work. ## 2 S-matrices and Constraints In Einstein frame, S-duality exchanges the coupling constant $`\tau =\chi +ie^\varphi `$ of the IIB superstring with its inverse, and more generally, under the $`SL(2,Z)`$ fractional linear transformation, $$\tau \frac{a\tau +b}{c\tau +d}.$$ (2.1) An invariant perturbative expansion for the scattering of gravitons is obtained by expanding in the string scale $`\alpha ^{}`$ at all orders in the perturbative series, rather than in the string coupling constant $`g_s=e^\varphi `$. Such an expansion is necessarily non-perturbative in form from the point of view of the string coupling constant, but there are constraints from the known structure of the perturbative series in supergravity to be compared with. The low-energy expansion of the string S-matrix is found by expanding the kinematic invariants parametrizing the scattering at small values below the string scale $`\alpha ^{}`$. We define the Mandelstam invariants relevant for the four-point function by $`s=(k_1+k_2)^2`$, $`u=(k_1+k_3)^2`$ and $`t=(k_2+k_3)^2`$. The general structure of this expansion contains polynomial terms in the kinematic invariants together with logarithmic functions, as demanded by unitarity of the massless modes. For $`s_{ij}>4/\alpha ^{}`$, the unitarity cuts for the massive modes of the IIB superstring appear after a resummation of the former terms; an infinite resummation of the higher derivative terms may produce the required unitarity cuts for the massive string states. For example, the genus one form for a unitarity threshold of the first massive mode of the superstring is $`\mathrm{ln}(1\alpha ^{}s/4)`$ and may be expanded at low-energy as an infinite series in $`s`$ via $`_{j=1}^{\mathrm{}}(\alpha ^{}s/4)^j`$. The fact that the energy scale of the kinematics is below the first massive mode of the string is crucial for preserving the manifest S-duality invariant expressions so far found in the literature. The string perturbative series is normalized with the conventions $`\kappa _{10}^2={\displaystyle \frac{1}{2}}(2\pi )^7\alpha ^4`$ (2.2) for the ten-dimensional gravitational coupling constant. Because $`\alpha ^{}`$ enters both in the coupling from the ten-dimensional field theory point of view, as well as in the parametrization of the mass levels, disentangling of the contributions to the amplitudes from the massive modes verse the massless ones of the string needs to be carried out. However, the contributions easily separate to genus zero and one in the string perturbative series. As of yet, there is no known consistent form of the S-duality compliant S-matrix for the IIB superstring in flat ten-dimensional Minkowski background. The S-duality invariance of the Einstein frame perturbative expansion demands a stringent form of the scattering, for example, of four-graviton scattering to low orders in the derivative expansion at fixed, but small, $`\alpha ^{}`$. The work of indicates that the form of the scattering of the polynomial terms up to twelve derivatives is given non-perturbatively by the series, $`S_{4point}={\displaystyle d^{10}x\sqrt{g}\left[\frac{1}{\mathrm{}^3}R^4+E_{3/2}(\tau ,\overline{\tau })R^4+E_{5/2}(\tau ,\overline{\tau })\mathrm{}^2R^4\right]},`$ (2.3) in Einstein frame. $`E_s(\tau ,\overline{\tau })`$ are non-holomorphic Eisenstein series (defined in section 4) and the IIB string coupling is $`\tau =\chi +ie^\varphi `$. We have included in the first term in (2.3) the massless bosonic exchange at tree-level. The string coupling constant is $`g_s=e^\varphi `$ and the derivatives in the first term is shorthand (2.3) for the factor $`1/stu`$. The modular construction of the graviton S-matrix is S-duality invariant, and previous attempts to generalize to all orders followed by examination of the tree-amplitude for four gravitons . However, it gives incorrect predictions for the one-loop contribution to the fourteen derivative term $`\mathrm{}^3R^4`$. The tree-amplitude for the scattering of four gravitons in a flat background in IIB theory is $`A_{4,g=0}^{\mathrm{IIB}}=64R^4{\displaystyle \frac{e^{2\varphi }}{\alpha ^3stu}}{\displaystyle \frac{\mathrm{\Gamma }(1\frac{\alpha ^{}s}{4})\mathrm{\Gamma }(1\frac{\alpha ^{}t}{4})\mathrm{\Gamma }(1\frac{\alpha ^{}u}{4})}{\mathrm{\Gamma }(1+\frac{\alpha ^{}s}{4})\mathrm{\Gamma }(1+\frac{\alpha ^{}t}{4})\mathrm{\Gamma }(1+\frac{\alpha ^{}u}{4})}},`$ (2.4) or in an alternative form, $`A_{4,g=0}^{\mathrm{IIB}}=64e^{2\varphi }{\displaystyle \frac{R^4}{\alpha ^3stu}}\mathrm{exp}\left({\displaystyle \underset{p=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{2\zeta (2p+1)}{2p+1}}\left({\displaystyle \frac{\alpha ^{}}{4}}\right)^{2p+1}\left(s^{2p+1}+t^{2p+1}+u^{2p+1}\right)\right),`$ (2.5) where the tree-level on-shell effective action found by integrating out massive string modes is contained in the second term of the expansion, i.e. $`2\zeta (3)e^{2\varphi }d^{10}x\sqrt{g}R^4`$. The $`R^4`$ factor represents the well-known tensor , the square of the Bel-Robinson tensor , which in linearized $`k`$-space form is found by contracting eight momenta with the external four polarization vectors, $$R^4=t_8^{\mu _1\mathrm{}\mu _8}t_8^{\nu _1\mathrm{}\nu _8}\underset{i=1}{\overset{4}{}}\epsilon _{\mu _i\nu _i}k_{\mu _{i+4}}^ik_{\nu _{i+4}}^i,$$ (2.6) in momentum space, or in the linear approximation (contributing only to the four-point function) of the contraction of four Weyl tensors with the tensors $`t_8`$. The conjecture in for the non-perturbative form of the S-matrix is found by replacing the Riemann-zeta functions in (2.5) with an Eisenstein series through, $$\zeta (2p+1)E_{p+1/2}(\tau ,\overline{\tau }),$$ (2.7) which takes into account the Einstein frame dependence of the string scattering. Although this substitution leads to the correct $`R^4`$ coupling, it predicts a contribution at genus one for the $`\mathrm{}^3R^4`$ term that is twice its calculated value and is not normalization dependent. It also does not address the required unitarity properties of the massless modes (i.e. supergravity). Building modular invariant functions from covariantizing through the substitution in (2.7) is not unique, and we shall modify this conjecture in this work. The one-loop amplitude for four-graviton scattering is $`A_{4,g=1}^{\mathrm{IIB}}=R^4{\displaystyle \underset{i=1}{\overset{4}{}}d^2z_i__1\frac{d^2\tau }{\tau _2^2}\underset{ij}{}|E(z_i,z_j,\tau )|^{\alpha ^{}s_{ij}}e^{\frac{2\pi }{\tau _2}k_ik_j\mathrm{Im}(z_i)\mathrm{Im}(z_j))}},`$ (2.8) with the genus one prime form, $`E(z=z_iz_j;\tau )={\displaystyle \frac{\mathrm{\Theta }\left[\genfrac{}{}{0pt}{}{\frac{1}{2}}{\frac{1}{2}}\right](z,\tau )}{\mathrm{\Theta }^{}\left[\genfrac{}{}{0pt}{}{\frac{1}{2}}{\frac{1}{2}}\right](0,\tau )}},`$ (2.9) and may be similarly expanded in the $`\alpha ^{}0`$ limit . (The denominator in (2.9) drops out due to momentum conservation.) The low-energy expansion and analytic properties of this amplitude is taken in , and the integral expansion may be directly reproduce as an infinite summation of field theory ten-dimensional box diagrams with the appropriate internal mass parameters as in figure 4. In the expansion in $`\alpha ^{}`$ is carried out to order $`\mathrm{}^2R^4`$ in the derivative (or equivalently $`\alpha ^{}`$) expansion. We shall need to disentangle the massive from the massless mode contributions in this expansion to compare with the predicted form of graviton scattering in pure IIB supergravity. The massive modes generate terms in the $`q`$ expansion of the oscillators when the power is non-zero. It is not that the fundamental integration region in (2.8) is demanded by a modular cancellation between the massive and massless modes of the superstring; modular $`SL(2,Z)`$ transformations of the punctured torus map regions in the field theory limit of the loop integration to other regions in addition to those of the massive contributions. The string, in the field theory limit, dictates a well-defined regulator. In the open-string limit suitable for Yang-Mills theory, this complication does not enter until the world-sheet corresponds to two-loops. Although straightforward at one-loop, technicalities associated with modular parameterizations and world-sheet ghosts at higher genus complicate a similar expansion at this order and will be analyzed in future work. We end this section with a clarification of the string-inspired regulator at one-loop. The fundamental domain $`_1`$ is over a restricted domain in the complex plane with a non-trivial region near the origin, $`_1=\{\tau =\tau _1+i\tau _2:\tau _1^2+\tau _2^21,|\tau _1|{\displaystyle \frac{1}{2}}\}.`$ (2.10) In field theory the box diagram with masses on the internal lines labelled by $`j`$ is given by the integral representation, $`\mathrm{I}_4(k_j,m_j)`$ $`=`$ $`{\displaystyle \frac{d^dl}{(2\pi )^d}\underset{j}{}\frac{1}{(lp_j)^2}}`$ (2.11) $`=(2\pi )^{\frac{d}{2}}{\displaystyle \underset{i=1}{\overset{4}{}}da_i\delta (1\underset{j=1}{\overset{4}{}}a_j)_0^{\mathrm{}}\frac{d\stackrel{~}{\tau }_2}{\stackrel{~}{\tau }_2^{2+d/2}}e^{\stackrel{~}{\tau }_2[f(a_i;p_i)+a_jm_j^2]}}`$ $`=_R(2\pi )^{\frac{d}{2}}\mathrm{\Lambda }^{3d/2}{\displaystyle \underset{i=1}{\overset{4}{}}da_i\delta (1\underset{j=1}{\overset{4}{}}a_j)_1^{\mathrm{}}\frac{d\tau _2}{\tau _2^{2+d/2}}e^{\tau _2[f(a_i;p_i)+a_jm_j^2]/\mathrm{\Lambda }^2}},`$ where in the last line we have implemented a Schwinger proper time regulator. The boundary on the integration is specified through the use of a regulator. In a Schwinger proper time regulator the region for $`\stackrel{~}{\tau }_2`$ is be limited to $`\stackrel{~}{\tau }_21/\mathrm{\Lambda }^2`$. In dimensional reduction (or regularization) we need to integrate over the entire region of $`\tau _20`$ with the dimensionally continued integral (in the measure). However, an alernative cutoff scheme for the supergravity amplitudes that is consistent with the underlying modular invariance of the string is to use the region delimited in (2.10), which is at one-loop found by replacing the integration regime in (2.11), $`{\displaystyle _1^{\mathrm{}}}{\displaystyle \frac{d\tau _2}{\tau _2^2}}{\displaystyle __1}{\displaystyle \frac{d^2\tau }{\tau _2^2}}.`$ (2.12) The region in the complex plane with $`\tau _21`$ is the same, because the integration over $`\tau _1`$ integrates to unity in (2.11). Its difference from the Schwinger proper time cutoff is in a finite region in the corner of moduli space. The integration region in (2.12) should be considered as an alternative regularization scheme, one that is compatible in supergravity with the non-perturbative S-duality transformations of the superstring. The parameterization of the domain of integration in the higher-genus moduli space (and references) serves as a string-inspired regulator for higher loop field theory, and gives rise to the string-inspired regulated supergravity quantum field theory. The regulated higher genus form is similar to (2.12) but with a more complicated corner region. The Schwinger proper-time regularization generalizes through the additional moduli of the superstring as at one-loop. ### 2.1 Ten Dimensions The previously conjectured form for the manifestly S-duality invariant function corresponding to the four-graviton scattering amplitude used the Eisenstein functions $`E_s(\tau ,\overline{\tau })`$ to relate the tree-level contributions to the higher genus ones by covariantizing the phase in (2.5). However, it clearly is only an approximation for the reason that it predicts the incorrect coefficient of the $`\mathrm{}^3R^4`$ term at genus one, and does not take into account the thresholds associated with the massless string modes (i.e. $`d=10`$ IIB supergravity). In this section, we give a different version that incorporates both of these features. The simplest modification is to enlarge the set of Eisenstein functions to their generalized non-holomorphic and non-modular invariant counterparts. The constraints for writing down suitable modular forms and combinations follow from : 1) the perturbative series is a power series in even powers of the string coupling $`\tau _2`$, 2) invariance under modular transformations in Einstein frame, 3) compatibility with the unitarity structure of the massless string modes, 4) the power series predicts the appropriate maximum power of $`\tau _2`$ consistent with the derivative term $`\mathrm{}^kR^4`$ term in Einstein frame at tree-level. We shall also consider only the Eisenstein series with half-integer or integer characteristics (the value of which is related to the half-integral R-charge shown in and ) and in ten dimensions those that converge (lower dimensional compactifications of supergravity have infra-red divergences in $`d6`$, and possibly the convergence properties are related to the infra-red divergences). The first condition is tight enough to rule out most commonly known modular forms, and we shall consider the set of generalized Eisenstein functions for $`SL(2,Z)`$, $`E_s^{(q,q)}(\tau ,\overline{\tau })`$, $`E_s^{(q,q)}(\tau ,\overline{\tau })={\displaystyle \underset{(p,q)(0,0)}{}}{\displaystyle \frac{\tau _2^s}{(p+q\tau )^{s+q}(p+q\overline{\tau })^{sq}}}.`$ (2.13) Their properties are discussed in section 4, together with the exceptional modular invariant function $$f=\mathrm{ln}\tau _2|\eta (\tau ,\overline{\tau })|^4$$ (2.14) which is related to the non-convergent $`E_1(\tau ,\overline{\tau })`$ series after subtracting the singularity, denoted by $`\widehat{E}_1(\tau ,\overline{\tau })`$, $`\widehat{E}_1(\tau ,\overline{\tau })=\pi \mathrm{ln}(\tau _2|\eta (\tau )|^4).`$ (2.15) The $`SL(2,Z)`$ Laplacian acting on $`f(\tau ,\overline{\tau })`$ is one, $$\tau _2^2_\tau _{\overline{\tau }}\mathrm{ln}\tau _2|\eta (\tau ,\overline{\tau })|^4=1,$$ (2.16) which in the string setting represents contributions from only one perturbative order (the form is required for genus one and for the $`\mathrm{}\mathrm{ln}\mathrm{}R^4`$ tensor). Our ansatz consists of polynomials in this combined set. The cusp forms are modular invariant functions that vanish at $`\tau _2\mathrm{}`$ and admit the expansion, $`f_{\mathrm{cusp}}(\tau ,\overline{\tau })={\displaystyle \underset{n0}{}}a_n\tau _2^{\frac{1}{2}}K_{n1/2}(2\pi |n|\tau _2)e^{2\pi in\tau _1},`$ (2.17) expressed in terms of exponentials of $`\tau _2`$. They do not contribute to the perturbative expansion of either the superstring or supergravity, and we shall not consider them or the modifications of the polynomial terms by them in any detail. The first instanton correction to the $`R^4`$ term in provides evidence that they do not contribute to this order, and a similar calculation at higher $`\alpha ^{}`$ is necessary to predict whether or not they contribute to the higher derivative terms. Explicit forms for the cusp forms on the fundamental domain $`U(1)\backslash SL(2,R)/SL(2,Z)`$ are not known . Similarly, the ones relevant for the moduli spaces of the toroidally compactified theories are not known (these spaces contain the former as a subspace). We first discuss the polynomial terms arising in the low-energy expansion, followed by the non-analytic (logarithmic) ones. The transformation from string frame to Einstein frame, $`\eta _{\mu \nu }^{(s)}=\tau _2^{\frac{1}{2}}\eta _{\mu \nu }^{(E)}`$, induces a coupling constant into the derivative expansion that is important for the $`\tau _2`$ counting. For example, the $`\mathrm{}^kR^4`$ terms in the string frame at tree-level are described in Einstein frame, $$A_{4,g=0}^{IIB,k}=N_{g=0}^k\zeta (\frac{3}{2}+\frac{k}{2})\tau _2^{3/2+k/2}\mathrm{}_{(E)}^kR_E^4,$$ (2.18) for $`k1`$, and where the $`e^{2\varphi }`$ string coupling inherit in (2.5) and the transformation of the tensors are taken into account. This power of $`\tau _2`$ is the maximum one possible in the perturbative series and successive higher genus contributions lower it by two units successively. The general form of the contribution at arbitrary $`k`$ to all genus, $$A_4^{IIB,k}=f_k(\tau ,\overline{\tau })\mathrm{}_{(E)}^kR_{(E)}^4,$$ (2.19) demands the structure of $`f_k(\tau ,\overline{\tau })`$ to have the form, $$f_k(\tau ,\overline{\tau })=\zeta (\frac{3+k}{2})\tau _2^{\frac{3}{2}+\frac{k}{2}}+a_1\tau _2^{\frac{1}{2}+\frac{k}{2}}+a_2\tau _2^{\frac{3}{2}+\frac{k}{2}}+\mathrm{}+𝒪(e^{2\pi i\tau }).$$ (2.20) The coefficients $`a_g`$ are constants that may be found by doing explicit low-energy string perturbation theory calculations up to genus $`g`$. We will see that given the set of modular forms discussed above the series $`a_k`$ receives non-vanishing values up to a maximum genus for a given $`k`$. The fact that only one tensor, namely the eight-derivative $`R^4`$ covariantized from the linear form in (2.6), appears in the description of the four-point function is consistent both with perturbative IIB superstring theory up to genus four as well as with constraints in supergravity from unitarity constructions. This property is due to $`N=8`$ supersymmetry at one and two-loops and also to higher orders : The uniqueness of this tensor structure is the primary influence of $`N=8`$ supersymmetry in constraining the S-matrix as the loop integrations are not constrained directly by supersymmetry without additional regulator specification (such as choosing one compatible with the global duality symmetries of the classical field equations of maximal supergravity). Lower derivative tensor structures appear in theories with lower supersymmetric gravitational theories . For $`k=0`$, the only function in the set considered that could describe the expansion is $`E_{3/2}^{(0,0)}`$ (modulo cusp forms); the $`R^4`$ term then receives perturbative contributions at tree-level and one-loop. Explicit calculations in supergravity indicate that, at two-loops and higher, an additional $`\mathrm{}^2`$ may explicitly be pulled out from within the loop integrations (in figure 2 we list the integral contributions, where the tensor $`R^4`$ is not displayed). Thus, this truncation is consistent with known results (at tree and one-loop the massive modes in the $`q`$ expansion explicitly are of order $`\alpha ^{}`$, hence $`\mathrm{}`$ by dimensional analysis, higher than the massless modes). The value $`k=1`$ is special because momentum conservation of the four external gravitons, through $`s+t+u=0`$, forces the on-shell ten-derivative polynomial term to vanish. At $`k=2`$, again there is only one function one may write down, namely $`E_{5/2}(\tau ,\overline{\tau })`$, and predicts a genus zero and genus two contribution but none from genus one; explicit string one-loop calculations shows the vanishing of the genus one coefficient. At values of $`k3`$ a break with the previous pattern emerges because the non-modular invariant $`E_s^{(q,q)}(\tau ,\overline{\tau })`$ for $`q0`$ may contribute. At $`k=3`$ we may introduce the form $`f_3(\tau ,\overline{\tau })=\alpha _1E_{3/2}^2(\tau ,\overline{\tau })+\alpha _2E_{3/2}^{(1,1)}(\tau ,\overline{\tau })E_{3/2}^{(1,1)}(\tau ,\overline{\tau }),`$ (2.21) (where $`E_{3/2}^{(1,1)}=E_{3/2}^{(1,1)}`$). It has the large $`\tau _2`$ perturbative expansion, $`f_3(\tau ,\overline{\tau })=4(\alpha _1+\alpha _2)\left[\zeta (3)^2\tau _2^3+\left({\displaystyle \frac{\alpha _1\alpha _2/3}{\alpha _1+\alpha _2}}\right)4\zeta (2)\zeta (3)\tau _2+\left({\displaystyle \frac{\alpha _1\alpha _2/9}{\alpha _1+\alpha _2}}\right)4\zeta ^2(2)\tau _2^1\right].`$ (2.22) The relative coefficient between $`\alpha _1`$ and $`\alpha _2`$ is chosen to agree with the genus one contribution for $`\mathrm{}^3R^4`$. Note that (2.22) differs from previous forms because of the introduction of the modular invariant contribution of the generalized Eisenstein functions. A relative value of $`\alpha _1=\frac{5}{3}\alpha _2`$ leads to agreement with the calculated genus zero and one coefficients in string theory; this combination predicts a non-vanishing coefficient at genus two and no further contributions. Combinations of these modular forms is straightforward to construct at higher derivatives, although it is not uniquely determined through direct covariantization via the substitution in (2.7). Futhermore, in contrast to the $`k3`$ cases, such combinations allow for perturbative contributions arising from modular forms that contribute at orders of genus solely above tree-level. The non-analytic terms in the low-energy expansion require such combinations. As another example, consider $`k=4`$ and $`k=5`$. In this case we may use, $`f_4(\tau ,\overline{\tau })=\alpha _1E_{7/2}(\tau ,\overline{\tau })+\alpha _2E_{3/2}(\tau ,\overline{\tau }),`$ (2.23) and, $`f_5(\tau ,\overline{\tau })`$ $`=`$ $`\beta _1E_2(\tau ,\overline{\tau })E_2(\tau ,\overline{\tau })+\beta _2E_2(\tau ,\overline{\tau })+\beta _3E_{3/2}(\tau ,\overline{\tau })E_{5/2}(\tau ,\overline{\tau })`$ (2.24) $`+\beta _4E_2^{(1,1)}(\tau ,\overline{\tau })E_2^{(1,1)}(\tau ,\overline{\tau })+\beta _5E_{3/2}^{(1,1)}(\tau ,\overline{\tau })E_{5/2}^{(1,1)}(\tau ,\overline{\tau })+\mathrm{c}.\mathrm{c}..`$ There is one condition on coefficients $`\beta _1`$, $`\beta _2`$, $`\beta _4`$ in the asymptotic expansion from the perturbative expansion, that the $`\tau _2^1`$ term is zero. The number of different combinations increases with larger $`k`$ values. Similar functions may be constructed to all orders, and the functional form at higher derivatives is one of the central results in this work. In (2.23) the different combinations generate perturbative corrections at the following genus orders tabulated in Table 1. The perturbative series of the polynomial terms for the $`\mathrm{}^kR^4`$ tensor in string theory, using the generalized Eisenstein functions in (2.13) for $`s=n/2`$ and $`n`$ integer in a polynomial fashion, receives corrections up to maximum $`g_{\mathrm{max}}=\frac{1}{2}(k+2)`$ genus in string theory for $`k`$ even and $`g_{\mathrm{max}}=\frac{1}{2}(k+1)`$ genus for $`k`$ odd, listed in Table 2. Higher genus calculations in IIB superstring theory is necessary to verify the form as well as to make agreement with the coefficients; an indirect resummation of the leading terms in $`\tau _2`$ of the derivatives must also agree with the thresholds of the massive modes of the superstring and might fix the coefficients. In the remainder of this section we examine other functions built out of Eisenstein series. Generic combinations of the generalized Eisenstein series either do not give rise to the appropriate powers of $`\tau _2`$ in accord with the string perturbative series or have values of $`s1`$ and not converge. Ratios of Eisenstein series in the large $`\tau _2`$ perturbative regime, for example, $`{\displaystyle \frac{E_3(\tau ,\overline{\tau })}{E_{\frac{3}{2}}(\tau ,\overline{\tau })}}`$ $`=`$ $`{\displaystyle \frac{a_3\tau _2^3+b_3\tau _2^2+n_1e^{2\pi \tau _2}+\mathrm{}}{a_{3/2}\tau _2^{\frac{3}{2}}+b_{3/2}\tau _2^{\frac{1}{2}}+m_1e^{2\pi \tau _2}+\mathrm{}}},`$ (2.25) either do not generically reflect the appropriate $`\tau _2`$ dependence of string perturbation theory, or else, $`{\displaystyle \frac{E_{3/2}^3(\tau ,\overline{\tau })}{E_{5/2}(\tau ,\overline{\tau })}}={\displaystyle \frac{(a_{3/2}\tau _2^{\frac{3}{2}}+b_{3/2}\tau _2^{\frac{1}{2}}+e^{2\pi i\tau }+\mathrm{})^3}{a_{5/2}\tau _2^{\frac{5}{2}}+b_{5/2}\tau _2^{\frac{3}{2}}+ce^{2\pi i\tau }+\mathrm{}}}`$ (2.26) generate at large $`\tau _2`$ perturbative $`\tau _2`$ dependence in accord with string perturbation theory and could potentially be used to describe the $`\mathrm{}^5R^4`$ term. However, the dependence of $`\tau _2`$ on the exponential terms generically does not match with expectations of the D-instanton series (which begin with a single order one factor for the first correction after expanding (2.26)). The function in (2.26) has dependence on half-integral powers of $`\tau _2`$, and its exponentials are suppressed by a power of $`\tau _2^{5/2}`$. Rigorously, we can not rule out the ratio combinations similar to that in (2.26) at all orders without further constraints. Non-half integral or integral values of $`s`$ also generically lead to dependence on $`\tau _2`$ which is not captured by perturbation theory. For example, the product $`E_{5/4}^2(\tau ,\overline{\tau })`$ $`=`$ $`\left(a\tau _2^{\frac{5}{4}}+b\tau _2^{\frac{1}{4}}+𝒪(e^{2\pi i\tau })\right)^2`$ (2.27) $`=a^2\tau _2^{\frac{5}{2}}+2ab\tau _2+b^2\tau _2^{\frac{1}{2}}+𝒪(e^{2\pi i\tau }),`$ has a $`\tau _2`$ factor which is not suppressed by a factor of $`\tau _2^2`$ relative to the first term and cannot arise in string perturbation theory. Generic products of Eisenstein series with non-half integral values of $`s`$ also have $`\tau _2`$ dependence that does not agree with the perturbative series. Any of the functions built out of the products of generalized Eisenstein series may also be used to generate further automorphic functions by acting on them with the covariantized Laplacian $`=\tau _2^2_\tau _{\overline{\tau }}`$, e.g. $`E_{3/2}^2(\tau ,\overline{\tau })`$. The polynomial action of $``$ generates the same perturbative truncation in the modular ansatz up to a maximum genus for a given $`\alpha ^{}`$ order, although the instanton corrections may be modified together with cusp forms. Their effect on the perturbative series is to adjust coefficients in the perturbative series up to the maximum genus. The action up to $`\mathrm{}^3`$ order in the derivative expansion does not introduce additional automorphic functions because the single Eisenstein functions are eigenfunctions of $``$, and their use at higher derivatives appears redundant perturbatively. In lower dimensional toroidally compactified theories, the limiting Eisenstein functions relevant for describing the $`\mathrm{}^kR^4`$ terms have less well behaved convergence and are possibly related to the infra-red divergences in $`d6`$ supergravity. The general structure of the perturbative series arising through the modular form construction required by S-duality invariance is listed in table 2. The pre-factors $`E_{3/2}(\tau ,\overline{\tau })`$ and $`E_{5/2}(\tau ,\overline{\tau })`$ of the low-energy polynomial expansion up to $`\mathrm{}^2R^4`$ satisfy Laplacian eigenvalue conditions. We are not imposing a property $`4\tau _2^2{\displaystyle \frac{}{\tau }}{\displaystyle \frac{}{\overline{\tau }}}f_k^{(i)}(\tau ,\overline{\tau })=\lambda _k^{(i)}f_k^{(i)}(\tau ,\overline{\tau }),`$ (2.28) on terms $`f_k^{(i)}`$ of $`f_k(\tau ,\overline{\tau })`$ or a similar one on the potentially covariantized phase in (2.5), which together with the asymptotic behavior in (2.20), limits only single Eisenstein series and their generalizations to being a solution (together with cusp forms) to (2.28) . The $`SL(2,Z)`$ covariantization of the leading asymptotic term in $`f_k`$ does satisfy a Laplacian condition, i.e. $`(\tau _2^{\frac{3}{2}+\frac{k}{2}})=4(\frac{3+k}{2})(\frac{1+k}{2})(\tau _2^{\frac{3}{2}+\frac{k}{2}})`$, but may be covariantized with $`SL(2,Z)`$ in different ways. For example, the simplest gives $`E_{\frac{3}{2}+\frac{k}{2}}(\tau ,\overline{\tau })`$ but also $`E_{\frac{3}{4}+\frac{k}{4}}^2(\tau ,\overline{\tau })`$ or $`|E_{\frac{3}{4}+\frac{k}{4}}^{(1,1)}(\tau ,\overline{\tau })|^2`$. A naive covariantization of the leading terms is not unique and for this reason the graviton scattering then is not going to exponentiate directly into the form of a tree-like amplitude. ### 2.2 Non-analytic terms The unitarity structure of the perturbative superstring amplitudes also requires a description in terms of modular forms in order to construct manifestly S-dual scattering. Clearly such a construction is different than the above because at tree-level there are no non-analytic contributions and the perturbative expansion involving logarithmic functions, for example $`\mathrm{ln}^L(\mathrm{})`$ and $`\mathrm{ln}\mathrm{ln}\mathrm{}\mathrm{ln}\mathrm{}`$ (up to L iteratively), begins at genus one. The form of the action containing higher orders in derivatives depends on the choice of the definition of the effective action but is uniquely defined by an S-matrix compuation in string theory. As required by unitarity, the S-matrix and the effective action will both contain logarithmic terms. At one-loop, for example, explicit expressions found from expanding the integrated four-point S-matrix element has a contribution of the form $`C_{\mathrm{log}}^{g=1}=\left(s\mathrm{ln}s+t\mathrm{ln}t+u\mathrm{ln}u\right)R^4.`$ (2.29) Higher logarithms appear at multi-loop and similar structure is required in higher-point gravitational amplitudes. Under the transformation to Einstein frame from string frame the kinematic invariants pick up factors $`s_ss_e\sqrt{\tau _2}`$. (The metric is rescaled by a square-root factor of the coupling constant in ten dimensions.) The logs potentially produce further non-analyticity in the string coupling constant, $`s\mathrm{ln}s\sqrt{\tau _2}s_e\mathrm{ln}(\sqrt{\tau _2}s_e).`$ (2.30) In Einstein frame, the coupling $`e^\varphi =\tau _2`$ does not just count loops via the the topogical coupling, $`S_{\mathrm{dil}}={\displaystyle d^2z\sqrt{g}R\varphi }=2(1g)\varphi ,`$ (2.31) but rather includes a dilatonic factor within the free superstring action. S-duality does not simply exchange weak coupling with strong coupling in the Einstein frame for individual modes of the string: In a field theory setting the the coupling constant is inverted, but also the contributions of the massive modes of the superstring get mixed differently than the massless ones within the integration because of the scale factor introduced in transforming to Einstein frame. The $`\tau _2`$ dependence in the massive modes is not removed in the space-time propagator $`\sqrt{\tau _2}s_em_i^2`$. In (2.29) because of momentum conservation, the $`\tau _2`$ dependence in the logarithm cancels out $`(s+t+u=0)`$. The unitarity construction below is independent of $`\tau _2`$ dependence in the non-analytic (i.e. logarithmic) terms at higher loops, which iteratively constructs $`SL(2,Z)`$ invariant non-analytic terms. The on-shell polynomial terms in previous sections are sufficient to determine, through supersymmetrizing the on-shell $`\mathrm{}^kR^4`$ tensor, these non-analytic terms through a unitarity construction. The imaginary part in a particular channel, for example $`s=(k_1+k_2)^2`$, of the four-graviton scattering amplitude is determined through the unitarity relation $`\mathrm{Im}_sA_4(k_i)={\displaystyle \underset{n=2}{\overset{\mathrm{}}{}}}{\displaystyle \underset{\lambda _j}{}}`$ $`{\displaystyle }`$ $`d\varphi _nA_{n,\lambda _j}[g(k_1),g(k_2);p(\stackrel{~}{k}_1),p(\stackrel{~}{k}_2),\mathrm{},p(\stackrel{~}{k}_n)]`$ (2.32) $`\times A_{n,\lambda _j}^{}[g(k_3),g(k_4);p(\stackrel{~}{k}_1),p(\stackrel{~}{k}_4),\mathrm{},p(\stackrel{~}{k}_n)],`$ with the $`n`$-body phase space integration measure given by, $`d\varphi _n={\displaystyle \underset{j=1}{\overset{n}{}}}{\displaystyle \frac{d^d\stackrel{~}{k}_j}{(2\pi )^d}}\delta ({\displaystyle \underset{i=1}{\overset{n}{}}}\stackrel{~}{k}_i+{\displaystyle \underset{j=1}{\overset{4}{}}}k_j){\displaystyle \underset{j=1}{\overset{n}{}}}\delta ^{(d)}(\stackrel{~}{k}_j^2)\mathrm{\Theta }(\stackrel{~}{k}_j^0),`$ (2.33) and where $`\lambda _j`$ denotes the quantum numbers of the physical states “$`p`$” of the intermediate lines (gravitons and their supersymmetric partners in the gravitational multiplet). The phase space integration is over the region $`\stackrel{~}{k}_j^2=0{\displaystyle \underset{i=1}{\overset{4}{}}}k_i+{\displaystyle \underset{j=1}{\overset{n}{}}}\stackrel{~}{k}_j=0,`$ (2.34) of out-going momenta. The equation in (2.32) gives an iterative construction of the non-analytic terms in the four-graviton scattering amplitude. At energies $`s4/\alpha ^{}`$ only the massless modes of the superstring contribute to the imaginary part in the s-channel. Furthermore, the on-shell supersymmetrization of the $`\mathrm{}^kR^4`$ terms (together with a similar construction of higher-point graviton scattering amplitudes) provides the necessary amplitudes $`A_n`$ that are to be inserted into (2.32); a on-shell linearized IIB superspace is known in the absence of an off-shell one which might aid in the explicit supersymmetrization to higher order. Given an S-duality compliant form of the higher derivative polynomial terms, the non-analytic ones are necessarily also invariant under $`SL(2,Z)`$ transformations through (2.32). Higher-point graviton scattering amplitudes may also be S-duality covariantized which is an ingredient in (2.32). We may expand the amplitudes in (2.32) to a total order $`m`$ which receives contributions from the respective expansions of the two amplitudes to all orders $`k`$ and $`l`$ so that $`k+l=m`$. The order $`\alpha ^m`$ non-analytic term is then determined from the polynomial ones together with the lower-order non-analytic ones. In this iterative manner, the non-analytic terms in the low-energy expansion of the gravitational S-matrix is $`SL(2,Z)`$ invariant if the polynomial terms are. ## 3 Effective action and S-matrix in derivative expansion We shall re-examine the eight-derivative $`R^4`$ term in this section to review its form in view of the S-matrix; the off-shell effective action for the IIB superstring is difficult to define because off-shell string scattering and the action for the five-form self-dual field strength off-shell are not available directly although several different on-shell effective action constructions may be given. In this section we examine primarily the four-point function, the form of which does not alter significantly at one-loop; however, at higher-point because of contributing independent high-point diagrams from boxes to $`n`$-gons ($`nd`$) there will be significant differences at genus one and higher. The on-shell effective action presented in the literature up to eight derivatives is virtually indistinguishable from the S-matrix due to supersymmetry. There are two differences between different functional form of an on-shell effective action to this derivative order. First, the term arising from massless string exchange at tree-level, i.e. in $`\alpha ^{}0`$ limit the graph in figure 3 with intermediate boson lines, is not included. Second, there is regulator dependence in the quantum one-loop contribution in the field theory setting (entirely from the massless modes in a $`q`$-expansion) that breaks the S-duality invariance if a string-inspired regulator is not used in comparing the supergravity with the string. Furthermore, the linearized $`N=8`$ (non-linear) supersymmetry will cancel any triangle or bubble sub-graph in a multi-loop graviton scattering amplitude. The string diagrams then, in the field theory limit, that have been included in the definition of the IIB effective action are identical to the ones that contribute to the four-point scattering amplitude in IIB superstring theory. Although at the four-point level the only difference between the S-matrix and the effective action in is contained in (3.35), at higher-point there are contributions of external massless trees to non-vanishing loop diagrams and then more differences associated with the external massless trees from a definition adopted in . Agreement with results in is obtained in an S-matrix calculation in superstring theory. In this remainder we examine the differences of the different definitions together with S-duality at the eight derivative order. In Einstein frame, the massless tree-exchange, identical to that in figure 3 but with intermediate bosonic modes together with the four-point vertex, gives rise to a covariantized contribution, $$A=\frac{1}{\kappa _{10}^2}d^{10}x\sqrt{g}\frac{1}{\mathrm{}^3}R^4,$$ (3.35) where the $`\mathrm{}^3`$ represents the appropriate combination of derivatives to produce a $`1/stu`$ in momentum space, and is independently invariant under the S-duality transformation. In not keeping it, the effective action is thought of as giving rise to 1PI quantum corrected vertices which must be sewn together to generate an S-matrix, i.e. illustrated as shaded circles in figure 3, and requires an off-shell generalization to construct the S-matrix. Furthermore, at one-loop there is an additional $`SL(2,Z)`$ of modular invariance and in not keeping the term in (3.35) arising from the massless sector breaks this in the string scattering perturbation theory. To this order the terms that break modular invariance do not break with S-duality, however, at higher-order this is possible. We examine a more precise definition for the S-duality invariant graviton scattering expression in the following - different definitions of the low-energy quantum effective action will produce further differences at the four-point level at order eight derivatives than just that in (3.35) and potentially break the S-duality. The S-duality invariant expressions that have been obtained so far demand that the external kinematics are below the string scale, e.g. $`s_{ij}4/\alpha ^{}`$ despite the fact that the massless and massive modes are being integrated out at the quantum level. Not keeping the term in (3.35) gives a definition similar to a combination of a one-particle irreducible for the massless quantum fields together with one-particle reducible for the massive modes. In examining the implications of S- and U-duality on the supergravity theory the regulator must be chosen in a way that is compatible with these symmetries of the classical field equations. Supersymmetry together with M-theory constraints suggest that the coefficient of the $`R^4`$ term is a function $`f`$ satisfying a Laplacian condition on the fundamental domain of $`U(1)\backslash SL(2,R)/SL(2,Z)`$: $`\tau _2^2_\tau _{\overline{\tau }}f=3/4f`$. Boundary conditions coming from the perturbative one-loop calculation need to be specified in order to find the solution. In supergravity the domain of vacua is the entire complex $`\tau `$-plane of couplings as S-duality is not a structure only in the massless sector of the string only. The general form of the tree- and one-loop contributions to the $`R^4`$ term are of the type, $`f_{\mathrm{pert}}(\tau ,\overline{\tau })=a\tau _2^{3/2}+b\tau _2^{1/2},`$ (3.36) and arise from massive mode exchange at tree-level and massless ones at one-loop (with coefficients $`a`$ and $`b`$ respectively). The $`SL(2,Z)`$ invariant completion of the former term gives a form, $`f(\tau ,\overline{\tau })=a\tau _2^{1/2}+b{\displaystyle \underset{p,q}{}}{\displaystyle \frac{\tau _2^{3/2}}{|p+q\tau |^3}}`$ (3.37) and agrees with perturbative IIB superstring theory when, $`f=2\zeta (3)\tau _2^{3/2}+4\zeta (2)\tau _2^{1/2}+𝒪(e^{2\pi i\tau }),`$ (3.38) where the exponentially suppressed terms correspond to $`k`$-multiple D-instanton corrections in the superstring theory (for example, calculated for $`k=1`$ in ). The effective actions found by distinguishing the massless modes of the superstring with different treatments changes the first coefficient $`a`$ in (3.37) to different values. It measures the regulator influence in the supergravity, that is, the cutoff $`\mathrm{\Lambda }`$ dependence or the use of dimensional regularization. Only for one value is the result in (3.37) $`SL(2,Z)`$ invariant, $`a=0`$, and that is the one that comes directly from either an S-matrix element calculation in IIB string theory or through supergravity with a regulator modelling the one that arises in the low-energy limit of the quantum superstring. In supergravity at one-loop the massless modes contribute, $`I(s,t,u)={\displaystyle \frac{d^dl}{(2\pi )^d}\frac{1}{l^2(lk_1)^2(lk_1k_2)^2(l+k_4)^2}}+ut+su,`$ (3.39) to the $`R^4`$ tensor and must be regulated because it is quadratically divergent in ten dimensions. In dimensional reduction ($`d=10ϵ`$) the result is, $`I(s,t,u)=C_ϵ\left[{\displaystyle \frac{1}{ϵ}}(s+t)+{\displaystyle \frac{1}{ϵ}}(u+s)+{\displaystyle \frac{1}{ϵ}}(t+u)\right]+(\mathrm{non}\mathrm{analytic}),`$ (3.40) and gives no contribution to $`R^4`$; dimensional reduction breaks S-duality. However, in a string-inspired regulator (2.12) the result from (3.39) is $`A_4^{[N=8]}(k_i,ϵ_i)R^4\left[\mathrm{\Lambda }^2+(s^2+t^2+u^2)\mathrm{ln}\mathrm{\Lambda }^2+(\mathrm{finite})\right].`$ (3.41) The quadratic divergence in a string-inspired regulator is roughly proportional to the inherit string scale $`\mathrm{\Lambda }^21/\alpha ^{}`$. The precise coefficient of the one-loop contribution to the $`R^4`$ term in $`d`$-dimensions depends on the regulator chosen and agrees with the string result for the $`R^4`$ term when the domain in (2.12) is taken. Scaling the coupling constants to force agreement between a general integration in (3.41) changes the instanton corrections predicted from the factor $`E_{3/2}`$ of the $`R^4`$ term. The calculation may be straightforwardly generalized to arbitrary dimensions and to its ultra-violet finite loop integration when $`d<8`$. The string-inspired regulator from the field theory point of view makes the calculated result from (3.41) agree with S-duality. Changing the regulator with the use of dimensional reduction gives a different value, $`c=0`$ in (3.38) from (3.40). The above regulators preserve (non-linear) supersymmetry to this order, but only one gives rise to an explicit $`SL(2,Z)`$ invariant expression, namely the string-inspired regulator built into the supergravity theory. S-duality is not expected to be a property of the low-energy IIB supergravity theory, as may be explicitly found to the order eight derivative $`R^4`$ term by dropping the massive mode contributions which contribute the $`\tau _2^{3/2}`$ term in (3.37). However, as a structure in the IIB superstring, it survives as a remnant in the IIB supergravity theory directly if a regulator is chosen so that the integration region in a first quantized form is chosen to mimic that of the superstring. At one-loop this corresponds to using a Schwinger proper time form of the four-point function in (2.12). There is a one parameter family of functions $`f_{\mathrm{pert}}(\tau ,\overline{\tau })`$ consistent with supersymmetry that leads to the form in (3.37), and the definition that arises from the low-energy limit of the string amplitude or in supegravity with the regulator chosen in (2.12) agrees with the S-duality invariant form when $`a=0`$ in (3.37). At two-loops the supegravity modular parameterization and regulator has been elucidated in . Explicit calculations of one-loop four-graviton scattering in IIB supergravity indicate that all the different definitions of the effective action are encoded in one term to order eight derivatives. In deducing results regarding the perturbative series of maximal supergravity directly from the superstring a multi-loop regulator must be chosen to agree with the modular properties of the IIB superstring S-matrix. ## 4 Modular Forms: $`SL(2,Z)`$ and Eisenstein Functions In this section we describe the generalized Eisenstein series on the fundamental domain of $`U(1)\backslash SL(2,R)/SL(2,Z)`$ and their relevant properties. These functions are non-holomorphic modular forms defined by $`E_s^{(q,q)}(\tau ,\overline{\tau })={\displaystyle \underset{(m,n)(0,0)}{}}{\displaystyle \frac{\tau _2^s}{(m+n\tau )^{s+q}(m+n\overline{\tau })^{sq}}},`$ (4.42) where the sum over the integral pairs $`(m,n)`$ does not include $`m=n=0`$. For $`q=0`$ we recover the non-holomorphic Eisenstein series. The series converges for $`s>1`$ on the fundamental domain ($`|\tau _1|1/2`$ and $`|\tau |1`$) and transforms under the fractional linear transformation $$\tau \tau ^{}=\frac{a\tau +b}{c\tau +d},$$ (4.43) as $`E_s^{(q,q)}(\tau ^{},\overline{\tau }^{})=\left[{\displaystyle \frac{c\tau +d}{c\overline{\tau }+d}}\right]^qE_s^{(q,q)}(\tau ,\overline{\tau }).`$ (4.44) Although the Eisenstein series are modular invariant, the generalized series for $`q0`$ transforms with a weight $`(q,q)`$. The latter may be used to find additional modular invariant functions by pairing them together as in $`E_s^{(q,q)}(\tau ,\overline{\tau })E_s^{(q,q)}(\tau ,\overline{\tau })`$ together with additional $`n`$-tuple products where the weights add up as $`_{i=1}^nq_i=0`$. The generalized Eisenstein series $`E_s^{(q,q)}(\tau ,\overline{\tau })`$ are related differentially to $`E_s(\tau ,\overline{\tau })`$ by $`E_s^{(q+1,q1)}(\tau ,\overline{\tau })=\left(i\tau _2{\displaystyle \frac{}{\tau }}+{\displaystyle \frac{q}{2}}\right)E_s^{(q,q)},`$ (4.45) and $`E_s^{(q1,q+1)}(\tau ,\overline{\tau })=\left(i\tau _2{\displaystyle \frac{}{\overline{\tau }}}i{\displaystyle \frac{q}{2}}\right)E_s^{(q,q)}.`$ (4.46) The asymptotic expansion for $`q0`$ may be obtained from $`q=0`$ via the relation in (4.45). Further, the functions satisfy the $`SL(2,Z)`$ covariantized differential relation, $$4\left(\tau _2\frac{}{\tau }+i\frac{1q}{2}\right)\left(\tau _2\frac{}{\overline{\tau }}i\frac{q}{2}\right)E_s^{(q,q)}(\tau ,\overline{\tau })=\lambda _s^{(q,q)}E_s^{(q,q)}(\tau ,\overline{\tau }),$$ (4.47) with an iteratively constructed eigenvalue from (4.45). The complete asymptotic form for large $`\tau _2`$ is found via manipulating the series through a Poisson resummation. The general form is $`E_s^{(0,0)}(\tau ,\overline{\tau })=a_s\tau _2^s+b_s\tau _2^{1s}`$ (4.48) $`+{\displaystyle \frac{2\sqrt{\tau _2}\pi ^s}{\mathrm{\Gamma }(s)}}{\displaystyle \underset{(m,n)(0,0)}{}}|{\displaystyle \frac{m}{n}}|^{s1/2}K_{s\frac{1}{2}}(2\pi \tau _2|mn|)e^{2\pi imn\tau _1},`$ (4.49) where $`K_s(x)`$ is the standard modified Bessel function with expansion for large $`x`$, $`K_r(x)=({\displaystyle \frac{\pi }{2x}})^{\frac{1}{2}}e^x\left[{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{(2x)^n}}{\displaystyle \frac{\mathrm{\Gamma }(r+n+\frac{1}{2})}{\mathrm{\Gamma }(rn+\frac{1}{2})\mathrm{\Gamma }(n+1)}}\right]`$ (4.50) and where the coefficients in (4.49) are, $`a_s=2\zeta (2s)b_s=2\sqrt{\pi }\zeta (2s1){\displaystyle \frac{\mathrm{\Gamma }(s\frac{1}{2})}{\mathrm{\Gamma }(s)}}.`$ (4.51) The asymptotic form for all $`s>1`$ contains two terms which are powers of $`\tau _2`$ together with an infinite series of exponentially suppressed terms. The former will relate to the perturbative expansion of the superstring S-matrix and the latter to a series of conjectured D-instanton corrections in the uncompactified theory. The generalized Eisenstein functions also have the same structure; however, the coefficients are different. For example, $`a_s^{(1,1)}=2s\zeta (2s)b_s^{(1,1)}=2\sqrt{\pi }(1s)\zeta (2s1){\displaystyle \frac{\mathrm{\Gamma }(s\frac{1}{2})}{\mathrm{\Gamma }(s)}}.`$ (4.52) All of these modular forms for general $`q`$-values give rise to two power suppressed terms in the large $`\tau _2`$ limit together with exponentially suppressed corrections. ## 5 U-dualities and Constraints In order to examine the same construction of the S-matrix in dimensions other then ten and compare with maximally extended supergravity theories, we shall compactify IIB superstring theory on $`M_d\times T^{10d}`$. In this section, we examine the modular form construction and find the similar truncation property to $`d=10`$ in perturbation theory. The S-duality inherited in ten dimensions is modified to the discrete U-duality group acting on all of the moduli of the toroidally compactified theory , and the low-energy theory is described by the dimensional reduction of the $`d=10`$ IIB supergravity. The non-perturbative vacuum is parameterized by the full complement of scalar fields in the lower dimensional maximally extended supergravity. In this section, we generalize the previous construction for the S-matrix of ten-dimensional string theory to the compactified cases with the use of additional automorphic functions. The vacuum state of compactified supergravity is parameterized by the values of scalars living in the symmetric space $`G(R)/H(R)`$ where $`H`$ is the maximal compact subgroup of $`G`$; we shall take the group $`G(R)=E_{p+1(p+1)}(R)`$ which contains as a subgroup $`SO(p,p,R)`$ (with the latter having maximal compact subgroup $`SO(p,R)\times SO(p,R)`$. Duality transformations generate an equivalence class of theories identified by an action of the infinite discrete group $`G(Z)`$. In compactifications $`M_d\times T^{10d}`$ there is a T-duality group $`SO(d,d,Z)`$ and a non-perturbative action of $`SL(2,Z)`$ (the coupling constant of the dilaton and axion scalar of the uncompactified IIB superstring) inherited from the uncompactified ten-dimensional IIB superstring. The full symmetry of the equations of motion and the quantum U-duality group is known to be larger and is listed in table 5 . (We will not consider the $`O(d,d,Z)`$ enlargement of the T-duality group as the elements with negative determinant exchange the IIB string with the type IIA one.) The larger U-duality group $`E_{11d(11d)}(Z)`$ ($`d9`$) contains the subgroup $`SO(10d,10d,Z)\times SL(2,Z)`$ for different $`d`$. It treats all of the scalars in the supergravity on the same footing, although the dilaton plays a special role in the compactified string in measuring the loop expansion. The general form of the symmetry and the constraints on the functional form of automorphic terms in the low-energy expansion of the S-matrix appear to fix the functional form of scattering in M-theory, defined in this case by T-dualizing the result of graviton scattering in IIB on $`S_R`$ and decompactifying the two-torus with complex structure $`\tau `$. The simplest theory from the above is uncompactified type IIB superstring theory which is parameterized by $`U(1)\backslash SL(2,R)`$. Duality transformations take the theories and maps them to equivalent theories under $`SL(2,Z)`$; the fundamental domain of $`U(1)\backslash SL(2,R)/SL(2,Z)`$ parameterizes inequivalent vacua of the ten-dimensional IIB superstring. In the compactified theories the inequivalent theories are described by vacuum expectation values of the set of moduli parameterizing the fundamental domain of $`H(R)\backslash E_{11d(11d)}(R)/E_{11d(11d)}(Z)`$. In analogy with the ten-dimensional case we consider as building blocks of the S-matrix the $`E_{11d(11d)}(Z)`$ invariant functions which satisfy the differential equations $`_{H(R)\backslash E_{11d(11d)}(R)}f_s(\varphi _j)=\lambda _{11d,s}f_s(\varphi _j),`$ (5.53) and their generalizations, in analogy with the Eisenstein series for the ten-dimensional uncompactified superstring theory. This form immediately decompactifies to higher dimensional U-duality invariant Eisenstein series differentially because the metric reduces on the U-duality subgroup. These functions relevant for the lowest derivative term, i.e. $`R^4`$, have been discussed for $`d7`$ in and for general integer dimensions in . In the latter work the unified set of perturbative and non-perturbative contributions contributions to the $`R^4`$ term was argued to be described by $`A_d={\displaystyle \frac{1}{\kappa _d^2}}{\displaystyle d^dx\sqrt{g}f_d(\varphi _j)R^4},`$ (5.54) where, $`f_d(\varphi _j)=E_{\mathrm{string};\mathrm{s}=3/2}^{E_{11d(11d)}(Z)}(\varphi _j),`$ (5.55) where $`\kappa _d^2`$ is the gravitational coupling constant in $`10d`$ dimensions and the Eisenstein function takes into account the summation over the “string multiplets”; the string multiplet is a unified representation of the full duality group $`E_{11d(11d)}(Z)`$ of the particle and membrane representations with given charges under the T-duality group and we refer the reader to for the construction of the general invariants (5.55). This function gives rise to perturbative contributions at tree-level and one-loop only which is in accord with the explicit results of maximal supergravity in $`d`$ dimensions in . It also limits consistently to reproduce decompactified dimensions as well as giving agreement with the known tree- and one-loop level contributions. The perturbative truncation property of the function in (5.55) permits contributions at only tree and one-loop level and is consistent with the tensor properties of maximally extended supergravity in different integer dimensions . The general constrained Eisenstein functions for a symmetric space $`G(R)/H(R)`$ are described in detail in , and are defined for a given representation $``$ of $`G`$ by, $`E_{R,s}^{G(Z)}(\varphi _j)={\displaystyle \underset{m\mathrm{\Lambda }_R0}{}}{\displaystyle \frac{\delta (mm)}{\left[M^2(m)\right]^s}}.`$ (5.56) The form in (5.56) is analogous to constructing modular invariant functions on the fundamental domain of $`SL(2,Z)`$ in the manner of $`g(\tau )={\displaystyle \underset{\gamma G}{}}f(\gamma \tau ),`$ (5.57) but with additional scalars involved in the mass formula in the denominator of (5.56). In (5.56), $`\varphi _j`$ denotes elements in $`G(R)/H(R)`$ and $`m`$ is a vector in the integer lattice $`\mathrm{\Lambda }_R`$ transforming in the representation $`R`$. The $`mm=0`$ condition projects onto sets of integers $`m`$ so that the symmetric tensor product $`R\times _sR`$ gives its highest irreducible multiplet which defines $`m`$ to be the most symmetric piece of the direct product. Physically, the set of integers in the lattice $`\mathrm{\Lambda }_R`$ labels the set of BPS states in the representation $`R`$ of the duality group (for example, one may consider the unified string representation of $`E_{11d(11d)}(Z)`$ or its further decompositions.) The condition $`mm=0`$ is the half BPS condition, and the mass formula contributing to the sum in the Eisenstein series for sets of states contributing to (5.56) is $`M_{\mathrm{BPS}}^2(m)=mMm.`$ (5.58) One could relax the condition $`mm=0`$ to include quarter and eigth BPS states by considering a more general definition of the the constrained Eisenstein function in (5.56). The functions $`E_R^{G(Z)}(\varphi _j)`$ are by construction invariant under the duality group when $`G(Z)=E_{11d(11d)}(Z)`$ and take values in the fundamental domain $`H(R)\backslash E_{11d(11d)}(R)/E_{11d(11d)}(Z)`$. The various discrete U-duality groups are listed in table 4. In analogy with the generalized Eisenstein functions on the $`U(1)\backslash SL(2,R)/SL(2,Z)`$ fundamental domain, a generalized series $`E_{R,s}^{(q,q),E_{11d(11d)(Z)}}(\tau ,\overline{\tau })`$ in the limit of zero moduli may be given. The asymptotic form in (5.70) may be given a weight by covariantizing as in (4.45). The coupling $`\tau _2V_{10d}^{\frac{4}{10d}}`$ is inert under $`SL(2,Z)`$ transformations. The simplest forms in the $`d`$-dimensional set are the Eisenstein functions relevant to uncompactified IIB superstring theory. The construction given in previous sections generalizes similarly, although the precise functional form is more complicated, to the toroidally compactified cases; however, the $`E_{11d(11d)}(Z)`$ Eiseinstein functions relevant to lower target dimensions have different convergence properties but the same type of perturbative truncation to a finite order in the coupling constants as is found in later sections in the small volume and null moduli limit where only the action on the $`SL(2,Z)`$ transformations is taken on the string coupling. As a consistency in the decompactification limit where the radii are taken to infinity, the successively higher-dimensional modular ansätze for the S-matrix should be retrieved. In order to generalize the S-matrix form in dimensions $`d7`$ analogous to the one in ten dimensions, we need a systematic treatment of the Eisenstein series only for the $`SL(p5)`$ groups, while in lower dimensions the exceptional groups enter as special cases with representations that may be decomposed into the $`SL(p)`$ ones or the T-duality ones. The symmetry of the supergravity field equations pertaining to the U-duality group enforce that the different states and solitonic configurations from the compactified string fall into representations of the duality group . These representations correspond to half-BPS states and are used in the construction (5.56) to find invariant functions under $`E_{11d(11d)}(Z)`$ which live on the fundamental domain parameterizing the moduli space of the compactified IIB superstring theory (listed in Table 3). As discussed in , these representations are listed in tables 4-6, and there may be degeneracy amongst the different constrained Eisenstein series between the functions constructed from the different representations at lowest order in derivatives, $`R^4`$. Consistency requires the higher-derivative terms in the low-energy limit to reduce under decompactification to the higher-dimensional forms (in integer dimensions). In the remainder of the section we confirm the $`\tau _2`$ dependence in Einstein frame of the perturbative series in different integral dimensions through comparison with the expansions of the appropriate modular forms at small coupling. The agreement of the expansions with the powers of $`\tau _2`$ is a check on the modular properties of the scattering in different dimensions. The transformation of string frame to Einstein frame is given by $`g_{\mu \nu }^{(s)}=e^{\alpha \varphi }g_{\mu \nu }^{(E)},`$ (5.59) together with $`\sqrt{g_{(s)}}=e^{\alpha \frac{d}{2}\varphi }\sqrt{g_{(E)}}R_{(s)}=e^{\alpha \varphi }R_{(E)}.`$ (5.60) The Einstein frame of the Einstein-Hilbert action derived from the string frame is found from, $`\alpha ={\displaystyle \frac{4}{d2}},`$ (5.61) or in terms of the ten-dimensional string coupling $`g_s^{(10)}`$ $`g_{\mu \nu }^{(s)}=\tau _2^{\frac{4}{d2}}g_{\mu \nu }^{(E)}s_{ij}^{(s)}=\tau _2^{\frac{4}{d2}}s_{ij}^{(E)},`$ (5.62) where we have illustrated how invariants scale also in (5.62). In (5.62) we see that dimensions $`d<6`$ is special: The Einstein frame coupling dependence increases by $`\tau _2^\beta `$ with $`\beta >1`$ for every pair of derivatives in the low-energy expansion and the transition dimension is $`d=6`$. Six dimensions is also singled out due to the presence of the four-form self-dual moduli. At weak IIB coupling, $`\tau _2\mathrm{}`$, the dimensions $`d6`$ have the feature that the tree-level terms are larger in string coupling at higher derivatives (in $`d>6`$ the terms are of smaller value), although the appropriate expansion parameter in this comparison at low-energy is $`\alpha =p_{(E)}^2\tau _2^{\frac{4}{d2}},`$ (5.63) which may give a decreasing effect if the momentum scale is such that $`\alpha <1`$. (The volume modulus is given in (5.65)). The coupling constant dependence in the low-energy S-matrix expansion has the distinguishing feature in $`d6`$, the same dimensions in which supergravity has extra finiteness properties in the Regge limit and also to higher loop orders implied by the automorphic IIB graviton scattering. Alternatively, we may transform to Einstein frame in ten dimensions and then compactify on an $`T^{10d}`$ torus, taking into account the volume dependence of the torii. The frame dependence in (5.62) forces the perturbative series in Einstein frame of the compactified IIB superstring to have tree-level contributions to $`\mathrm{}^kR^4`$ with the $`\tau _2`$ dependence, $`A_4^{\mathrm{tree},\mathrm{k}}=N_k\left[\tau _2V_{10d}^{\frac{4}{10d}}\right]^{(\frac{3+k}{d2})(\frac{10d}{2})}\tau _2^{\frac{3}{2}+\frac{k}{2}}{\displaystyle d^dx\sqrt{g_E}\mathrm{}_{(E)}^kR_{(E)}^4}`$ (5.64) $`=N_kV_{10d}^{\frac{2(3+k)}{(d2)}}\tau _2^{\frac{12+4k}{d2}}{\displaystyle d^dx\sqrt{g_E}\mathrm{}_{(E)}^kR_{(E)}^4},`$ (5.65) where the volume factor $`\tau _2V_{10d}^{\frac{4}{10d}}`$ is inert under the subgroup $`SL(2,Z)`$ transformations of the U-duality group in Einstein frame. In $`d=10`$ the perturbative series reproduces the series in (2.20) and sub-leading string genus corrections to (5.65) generate powers $`\tau _2^{2g}`$ relative to the tree-level; the coefficient is modified in the toroidal compactifications due to the volume dependence and scaling in $`d`$-dimensional compactified Einstein frame. We collect some explicit powers of $`\tau _2`$ dependence of the $`\mathrm{}^kR^4`$ terms in the low-energy expansion of the IIB superstring S-matrix on $`M_d\times T^{10d}`$ dimensions below: The $`\tau _2`$ dependence in (5.65) at tree-level in Einstein frame may be matched with the modular form expansions in $`d10`$ in accord with criteria (1) in section 2. The Eisenstein function in lower dimensions on the appropriate fundamental domain with value $`s=3/2`$ and $`s=5/2`$ gives rise to the dependence in Table 7. ### 5.1 Asymptotic Limits We give the large coupling expansions in this subsection and first examine the simplest case, namely the $`SL(n,Z)`$ Eisenstein functions to the fundamental (and anti-fundamental) representations. The $`SL(2,Z)`$ Eisenstein function to the fundamental representation is given by $`E_s(\tau ,\overline{\tau })={\displaystyle \underset{(p,q)(0,0)}{}}{\displaystyle \frac{\tau _2^s}{|p+q\tau |^s}},`$ (5.66) and for $`SL(n,Z)`$ in the fundamental representation we have the explicit analogous function, $`E_{R=d,s}^{\mathrm{SL}(\mathrm{n},\mathrm{Z})}(\varphi _j)={\displaystyle \underset{(m^1,\mathrm{},m^n)0}{}}{\displaystyle \frac{1}{\left[m^ig_{ij}(\varphi _k)m^j\right]^s}}.`$ (5.67) The anti-fundamental representation is given by (5.67) but with indices lowered/raised. The series in (5.67) is absolutely convergent only for $`s>n/2`$. Upon taking a $`d`$-dimensional torus into a direct product of one circle of radius $`R`$ times a $`T^{d1}`$ generically non-orthogonal torus together with large $`R`$, the limit of (5.67) is $`E_{R=d,s}^{\mathrm{SL}(\mathrm{d},\mathrm{Z})}(\varphi _j)=E_{R=d1,s}^{\mathrm{SL}(\mathrm{d}1,\mathrm{Z})}(\stackrel{~}{\varphi }_j)+{\displaystyle \frac{2\pi ^s\mathrm{\Gamma }(s\frac{d1}{2})\zeta (2sd+1)}{\pi ^{s\frac{d1}{2}}\mathrm{\Gamma }(s)R^{2sd+1}V_{d1}}}`$ (5.68) $`+{\displaystyle \frac{2\pi ^s}{\mathrm{\Gamma }(s)R^{2sd1}}}{\displaystyle \underset{m^a,n^b}{}}|{\displaystyle \frac{n^ag_{ab}n^b}{m^2}}|K_{s\frac{d1}{2}}\left(2\pi |m|R\sqrt{n^ag_{ab}n^b}\right).`$ (5.69) The asymptotic limit in (5.69) of $`E_{R=d,s}^{\mathrm{SL}(\mathrm{d},\mathrm{Z})}(\varphi _j)`$ gives rise to the Eisenstein series of $`SL(d1)`$ to the fundamental representation together with an infinite number of terms depending on the large radius (one power-behaved and the rest exponentially suppressed) which vanish as $`R\mathrm{}`$. Although only the fundamental representation is analyzed in (5.69) a similar structure is expected for other representations of the duality group $`U(Z)`$; the zero moduli approximation also generates this limit. This structure is consistent with the expected form of the modular properties of the S-matrix in the decompactification limit and the exponential terms in the expansion of the forms represent the various wrapping of string solitonic states in the compact directions of the compactified torii dimensions. In subsequent work we shall analyze the $`d`$-dimensional form in view of the finite radii dependence of the toroidal compactifications. The weak coupling expansion is defined by taking the dilaton IIB scalar to large vacuum expectation value (small string $`g_s`$ coupling constant). Turning off all of the vacuum values of the scalar fields except for $`\tau `$ which is acted on by the S-duality $`SL(2,Z)`$ subgroup of $`U(Z)`$, these (regulated) functions reduce to the Eisenstein series for $`E_s^{\mathrm{SL}(2,\mathrm{Z})}(\tau ,\overline{\tau })`$ together with a perturbative truncation property described in previous sections that gives rise to the same power counting in the low-energy limit of the superstring scattering elements. We can consider the special point of the moduli of the toroidally compactified theory when all moduli except for the ten-dimensional dilaton have vanishing values. In this case, the large $`\tau _2`$ behavior of the Eisenstein series in (5.53) may be constructed by solving the differential equation for the $`SL(p,Z)`$ representation decompositions into the $`SL(2,Z)`$ subgroup of the various string multiplet contributions. The functional form of the small moduli contribution is invariant under $`SL(2,Z)`$ transformations and is, $`E_s^{E_{11d(11d)}}(\tau ,\overline{\tau })=\left[\tau _2V_{10d}^{\frac{4}{10d}}\right]^{(\frac{3+k}{d2})(\frac{10d}{2})}{\displaystyle \underset{(p,q)(0,0)}{}}{\displaystyle \frac{\tau _2^s}{|p+q\tau |^{2s}}}+\mathrm{},`$ (5.70) for $`k=2s3`$, which truncates in every dimension, as it does in $`d=10`$ (for the subtracted or convergent series $`E_s`$). The form in (5.70) is due to the $`SL(2,Z)`$ subgroup of the larger U-duality group. Constructions using the generalized Eisenstein series follows similarly in the toroidal compactifications, with the difference involving the volume factor of the $`10d`$ dimensional tori. The unitarity construction described in previous sections gaurantees that the imaginary parts will also be invariant under the U-duality group, after construction of the supersymmetric extension of the $`\mathrm{}^kR^4`$ terms and integrating to find the imaginary parts. ### 5.2 D=8 Example In the remainder of this section we give an example for the polynomial terms in the low-energy expansion of the S-matrix in the compactified theories analogous to the one presented for the uncompactified IIB superstring but invariant under the larger U-duality group. We give the example of the $`D=8`$ theory which has a U-duality group in the toroidally compactified theory of $`SL(2,Z)\times SL(3,Z)`$ and a set of scalars parameterizing the fundamental domain of $`SO(2,R)\times SO(3,R)\backslash SL(2,R)\times SL(3,R)/SL(2,Z)\times SL(3,Z)`$. An explicit parameterization of the moduli space consists of the complex structure of the two-torus with metric (and volume $`V=V_2`$), $`g_{ab}={\displaystyle \frac{V}{U_2}}\left(\begin{array}{cc}1& U_1\\ U_1& |U|^2\end{array}\right)`$ (5.73) where $`U`$ spans the region $`U(1)\backslash SL(2,R)`$ and with the enhanced moduli of $`SL(3,R)/SO(3)`$. The scalar $`\tau _2V`$ is invariant under the $`SL(2,Z)`$ sub-group of $`SL(3,Z)`$ acting on the IIB scalar $`\tau `$. The moduli of $`SL(3,R)/SO(3)`$ consist of the dilatonic and axionic couplings $`\tau `$ together with those of the reduction of the anti-symmetric complex tensors $`B_{NS,R}=ϵ_{ab}B_{NS,R}^{ab}`$ grouped as $`B=B_R+\tau B_{NS}`$ into the symmetric coset matrix form with the matrix satisfying the determinant $`\mathrm{det}M=1`$ condition, $`M=(\tau _2V)^{1/3}\left(\begin{array}{ccc}\frac{1}{\tau _2}& \frac{\tau _1}{\tau _2}& \frac{(B)}{\tau _2}\\ \frac{\tau _1}{\tau _2}& \frac{|\tau |^2}{\tau _2}& \frac{(\overline{\tau }B)}{\tau _2}\\ \frac{(B)}{\tau _2}& \frac{(\overline{\tau }B)}{\tau _2}& \tau _2V^2+\frac{|B|^2}{\tau _2}\end{array}\right).`$ (5.77) The matrix $`M`$ in (5.77) together with $`U`$ parameterizes the scalar manifold $`U(1)\backslash SL(2,R)\times SL(3,Z)/SO(3,R)`$ of the classical IIB superstring theory compactified on a two-torus with metric in (5.73) and the complex structure parameterized by $`U`$. The action of the scalars and Einstein-Hilbert low-energy theory is $`S={\displaystyle \frac{1}{\kappa _8^2}}{\displaystyle d^8x\sqrt{g}\left[R\frac{U\overline{U}}{U_2^2}+\frac{1}{4}\mathrm{Tr}\left(MM^1\right)\right]},`$ (5.78) and we are primarily interested in this work in the string corrections to gravitational action. The Eisenstein series relevant to the eight-derivative term has been elucidated for the $`R^4`$ term in . The $`R^4`$ term in Einstein frame arising in the low-energy expansion is described in (5.55). In accord with the functional form in (2.3) and (2.13) we take the twelve derivative term to be $`S_{d=8}^{12}={\displaystyle d^8x\sqrt{g}E_{\mathrm{string};\mathrm{s}=5/2}^{SL(2,Z)\times SL(3,Z)}(\varphi _j)\mathrm{}^2R^4},`$ (5.79) which generalizes to arbitrary dimensions to $`S_d^{12}={\displaystyle d^dx\sqrt{g}E_{\mathrm{string};\mathrm{s}=5/2}^{E_{11d(11d)}(Z)}(\varphi _j)\mathrm{}^2R^4}.`$ (5.80) The string multiplet in $`d=8`$ decomposes under $`SL(2,R)\times SL(3,Z)`$ into the representation $`\mathrm{𝟏}\times \mathrm{𝟑}`$, and accordingly, the automorphic contribution in (5.79) breaks into, $`S_{d=8}^{12}={\displaystyle }d^8x\sqrt{g}E_{\mathrm{𝟑};5/2}^{SL(3,Z)}(\varphi _j^{(2)})]\mathrm{}^2R^4,`$ (5.81) or alternatively, $`S_{d=8}^{12}={\displaystyle d^8x\sqrt{g}\left[E_{\mathrm{𝟏};\stackrel{~}{s}_1}^{SO(2,2,Z)}(\varphi _j^{(1)})+E_{\mathrm{𝟐};\stackrel{~}{s}_2}^{SO(2,2,Z)}(\varphi _j^{(2)})\right]\mathrm{}^2R^4}.`$ (5.82) The modular weights $`s_1=3/2`$ and $`s_2=5/2`$ are determined by decomposing the form of the modular summand into the lower representations. The complex structure of the torus $`U`$ parameterizes the $`SL(2,R)/U(1)`$ portion of the T-duality group ($`SO(2,2,R)=SL(2,R)\times SL(2,R)`$) and in the decompactification limit. The $`SL(3,Z)`$ Eisenstein functions entering into (5.81) are explicitly in the fundamental representation constructed from the matrix in (5.77), $`E_{\mathrm{𝟑};𝐬}^{SL(3,Z)}(\varphi _j^{(2)})`$ $`=`$ $`{\displaystyle \underset{a,b=1,2,3}{}}{\displaystyle \underset{n^a,n^b}{}}{\displaystyle \frac{1}{\left[n^aM_{ab}n^b\right]^s}}`$ (5.84) $`={\displaystyle \underset{(n^1,n^2,n^3)(0,0,0)}{}}\tau _2^s{\displaystyle \frac{\left(\tau _2V\right)^{s/3}}{\left[|n^1+n^2\tau +n^3B|^2+(n^3\tau _2V)^2\right]^s}},`$ togther with the (subtracted) form of the order one Eisenstein funtion of the complex structure $`U`$, $`\widehat{E}_{\mathrm{𝟏};1}^{SL(2,Z)}(\varphi _j^{(1)})=\pi \mathrm{ln}\left(U|\eta (U)|^4\right),`$ (5.85) that we must also add in. The series in (5.84) is explicitly invariant under the $`SL(3,Z)`$ transformations and to leading order in $`\tau _2`$ has the expected power dependence to agree with the Einstein frame string coupling at tree-level (as it must if the compactified string theory possesses a U-duality structure). At small volume and at zero $`B`$ moduli we have, $`E_{\mathrm{𝟑};\frac{\mathrm{𝟓}}{\mathrm{𝟐}}}^{SL(3,Z)}(\varphi _j^{(2)})`$ $`=`$ $`\tau _2^{\frac{10}{3}}V^{\frac{5}{3}}{\displaystyle \underset{(m,n)(0,0)}{}}{\displaystyle \frac{1}{|m+n\tau |^5}}`$ (5.87) $`=\tau _2^{\frac{5}{6}}V^{\frac{5}{3}}\left(\tau _2^{\frac{5}{2}}\zeta (5)+a\tau _2^{\frac{3}{2}}+𝒪\left(e^{2\pi i\tau }\right)\right).`$ The perturbative series in (5.87) only receives contributions at genus zero and two in this limit. In the decompactification limit $`R_1,R_2\mathrm{}`$, the functional form in (5.79) must reconcile with the $`d=9`$ and $`d=10`$ results for the twelve-derivative term and be invariant under the U-duality group $`SL(2,Z)`$. Using (5.69), we see that the only term surviving the large R limit is the $`SL(3,Z)`$ Eisenstein function in (5.81), which correctly limits to an order $`s=5/2`$ $`SL(2,Z)`$ Eisenstein function. It would be interesting to check the finite radius dependence further. ## 6 Perturbative truncations in integral dimensions In this section we examine the implications of the perturbative truncation, that of a maximum genus for individual terms in the derivative expansion, with maximally extended supergravity theories. The modular structure of the graviton scattering amplitude indicates the perturbative truncation property in different dimensions and implies a similar behavior in supergravity. This property is present already at order $`\mathrm{}^2R^4`$ . Two options to explain the perturbative truncation implied by the modular structure are either an infinite number of independent cancellations at each genus between the massive and massless modes are nullifying the contributions to $`\mathrm{}^kR^4`$ for $`k2`$, or that the massive modes give rise to order $`\alpha ^{}`$ higher corrections compared with the supergravity modes in the low-energy limit (as at $`g=0`$ and $`g=1`$) and that IIB supergravity possesses this truncation. In order to compare the string amplitudes with the supergravity ones: the massive modes need to be distinguished from the massless ones in the $`\alpha ^{}0`$ limit at $`g3`$, the ghost dependence in the string amplitudes needs to be disentangled in the field theory limit, and regulator dependence in the supergravity theory implied by the integration region coming from the moduli space (at one-loop in (2.12)) has to be examined with regards to possible cancellations due to the region of integration. Ghost dependence in the genus two four-graviton scattering amplitude, for example, naively indicates a contribution to the $`R^4`$ term in the derivative expansion before moduli integration that is in disagreement with explicit two-loop supergravity graviton scattering ; consistency with S-duality and perturbative supergravity requires these dependencies to integrate to zero. In the IIB supergravity limit of the superstring scattering, primitive divergences in $`d=10`$ of the four-point amplitude are of the form: $`A_4^{L=1,m=0}\mathrm{\Lambda }^2R^4+\mathrm{}R^4+\mathrm{}`$ (6.88) $`A_4^{L=1,m0}\mathrm{}R^4+\mathrm{}^2R^4+\mathrm{}`$ (6.89) The $`\alpha ^{}`$ indicates an overall factor of $`\mathrm{}`$ or $`s_{ij}`$ in comparison to the massless modes after they are normalized correctly at the given order. At two-loops in field theory, an explicit twelve derivatives may be extracted from the loop integration and the amplitude has the generic tensor structure, $`A_4^{L=2,m=0}\mathrm{}^2\left(\mathrm{\Lambda }^6+\mathrm{\Lambda }^4\mathrm{}+\mathrm{\Lambda }^2\mathrm{}^2+\mathrm{}^3\right)R^4+\mathrm{}.`$ (6.90) Explicit string theory calculations at genus two may be computed in the $`\alpha ^{}0`$ limit. If the massive modes of the string contribute at an order $`\alpha ^{}`$ higher in the low-energy limit (as at order $`g=0`$ and $`g=1`$), then an additional pair of derivatives must also be extracted, $`A_4^{L=2,m0}\mathrm{}^3R^4+\mathrm{}.`$ (6.91) At tree and one-loop level the massive modes explicitly contribute an order higher in $`\alpha ^{}`$, and it is reasonable to suspect that such a property persists at higher order. However, an explicit integration of the genus two four-point function in the low-energy limit is required; picture dependence in this order complicates the analysis although unitarity might be faulted if this were not the case. At three-loops the conjectured form of the four-point supergravity amplitude in ten dimensions leads to $`A_4^{L=3,m=0}\mathrm{}^2\left(\mathrm{\Lambda }^{14}+\mathrm{\Lambda }^{12}\mathrm{}+\mathrm{}\right)R^4`$ (6.92) and if the massive modes contribute to a higher-order in $`\alpha ^{}`$ then again their contribution must be of an order higher in derivatives by dimensional analysis, $`A_4^{L=3,m0}\mathrm{}^3R^4+\mathrm{}.`$ (6.93) The contribution to $`\mathrm{}^2R^4`$ in the low-energy limit of the IIB superstring amplitude gets contributions at most from genus two from the modular form $`E_{5/2}(\tau ,\overline{\tau })`$ Eisenstein function; this requires the leading divergence in (6.92) to have a zero coefficient. The agreement of the low-energy field theory modelling of the superstring requires an infinite number of cancellations or nullifications at higher genus for every $`\mathrm{}^k`$ term in the low-energy expansion. The genus zero and two contributions to the $`\mathrm{}^2R^4`$ term in the ten-dimensional S-matrix, for example, indicates that separately every contribution (primitive ones above for example) proportional to $`\mathrm{}^2`$ at $`g>2`$ has to be zero (every perturbative term at different genus differs by powers of $`\tau _2^2`$). The same structure occurs because of the truncation $`g_{\mathrm{max}}^k=\frac{1}{2}(k+2)`$ and $`g_{\mathrm{min}}^k=\frac{1}{2}(k+1)`$, for even and odd $`k`$, for the higher derivatives. The perturbative counting arising from the Eisenstein series, i.e. $`g_{\mathrm{min}}=0`$ to $`g_{\mathrm{max}}^k`$, is most naturally explained in the field theory by an explicit extraction of an additional four derivatives at every order (as found for example at loop order one and two). In the Regge limit, where only the ladder diagrams contribute in the construction of , this property is seen to infinite loop order for the massless modes, and pure $`N=8`$ supergravity at vanishing moduli is finite in this kinematical regime to all orders in four through six dimensions. However, at three-loops it is not clear if an additional set of momenta may be extracted from the complete set of integral functions contributing to the amplitude due to possible non-planar contributions possessing only three- and higher particle cuts; the complete form of the non-planar contributions to the four-point supergravity amplitude are not explicitly known to this order and the tensor property of three and higher loops within the amplitude for the supergravity modes is an open question. Sample non-planar contributions are illustrated in Figure 5, where the internal powers of $`l^4`$ are not inserted . The tensor property, i.e. an extraction of internal loop momenta associated with the gravitational couplings in the pattern $`4(1+L)`$, implies that $`N=8`$ supergravity with all moduli tuned to zero except for the dilaton coupling is finite in four, five and six dimensions. The fact that the modular ansatz indicates such a cancellation in the different integer dimensions is suggestive of the tensor property, perhaps in a string-inspired regulator at higher loop order that preserves the S-duality structure in the supergravity quantum field theory. Such cancellations after integration could be associated with boundaries of the integration region and divergence contributions vanishing on them in different dimensions (for example, $`_{(p,q)(0,0)}1/|p+q\tau |^4`$ on the fundamental domain at genus one). A complete three-loop supergravity divergence calculation is necessary to find out if there is a cancellation in the field theory (before or after integration), or an explicit two-loop string theory calculation to disentangle picture dependence in the field theory limit and verify the $`q`$ expansion of the massive modes. ## 7 Conclusions We have explored a manifestly S-duality (and U-duality) invariant perturbative series in the derivative expansion of four-graviton scattering in IIB superstring theory compatible with the perturbative structure. Generalizations to higher-point gravitational amplitudes are straightforward. We have written the derivative expansion in terms of combinations of generalized Eisenstein series with coefficients not generally fixed by tree-level IIB graviton scattering. In ten dimensions this form predicts the absence of singularities in the moduli space of vacua which is consistent with taking into account all non-perturbative corrections. The duality structure imposes strong constraints on the perturbative series from the perturbative superstring amplitude calculations. Imposing the duality structure is an additional postulate on the S-matrix akin to unitarity in the approach of analytic S-matrix theory, and allows the exciting possibility of determining the functional form of graviton scattering in IIB superstring theory. The form of the scattering amplitude is also useful for computing finite $`\lambda =g^2N`$ effects of $`N=4`$ super Yang-Mills correlation functions through the AdS/CFT correspondence . Unitarity of the massless modes is $`SL(2,Z)`$ invariant by construction in this formulation. The thresholds associated with the massive modes of the string must be found by summing the derivative expansion. Furthermore, these thresholds at genus one and higher in IIB string perturbation theory are additional constraints not yet imposed on the full functional form of the derivative expansion in this work, and could be used to fix the coefficients of the linear combinations of the Eisenstein series. Factorization conditions and consistency of the D-instanton corrections with the coupling constant dependence of the instantons through a holographic AdS relation to $`N=4`$ super Yang-Mills also provide additional constraints. This approach, via T-duality from IIB theory on a circle, provides an avenue for pinning down in the derivative expansion the covariantized form of graviton scattering in the eleven-dimensional corner of M-theory. Furthermore, the same approach may be used to find the constraints of S-duality on the low-energy scattering obtained purely in $`N=4`$ super Yang-Mills theory, the planar form of which in the unbroken theory has been conjectured in perturbation theory in terms of integral functions for the four-gluon scattering process . Additional analysis through instanton calculus or constraints such as those listed above are necessary to determine uniqueness or possible contributions of cusp forms, for example, at higher derivatives which generate non-perturbative corrections. Further checks of S-duality at the amplitude level involve low-energy expansions of $`g2`$ genus four-graviton contributions. The modular property in this work predicts a perturbative truncation, that of a maximum genus contribution for a given derivative order, when expanded in the string coupling constant beyond the known supergravity structure at two-loops. Although the direct field theory limit of the superstring is difficult to obtain explicitly for genus greater than two, the most straightforward interpretation given the $`\alpha ^{}`$ expansion in the field theory limit is that the perturbative series of the massless modes also truncates, perhaps in one regulator that preserves the remnant of the S- and U-duality in the supergravity approximation or field equations. This feature in maximally extended supergravity theory indicates a much higher degree of finiteness in four through six dimensions than expected. Acknowledgements This work is supported in part by the U.S. Department of Energy, Division of High Energy Physics, Contract W-31-109-ENG-38. G.C. thanks Zvi Bern for numerous discussions on the finiteness properties of multi-loop maximally extended supergravity, and Jeffrey Harvey, Finn Larsen, Emil Martinec and Alan White for discussions.
warning/0001/hep-ph0001023.html
ar5iv
text
# Untitled Document A four–neutrino texture implying bimaximal flavor mixing and reduced LSND effect<sup>*</sup><sup>*</sup>*Work supported in part by the Polish KBN–Grant 2 P03B 052 16 (1999–2000). Wojciech Królikowski Institute of Theoretical Physics, Warsaw University Hoża 69, PL–00–681 Warszawa, Poland Abstract A four–neutrino effective texture is described, where a sterile neutrino mixes nearly maximally with the electron neutrino and so, is responsible for the deficit of solar $`\nu _e`$’s (according to the large–angle MSW solution or vacuum solution, of which the former is selected a posteriori). But, while maximal mixing of muon neutrino with tauon neutrino causes the deficit of atmospheric $`\nu _\mu `$’s, the original magnitude of LSND effect is reduced by as much as four orders, becoming unobservable. PACS numbers: 12.15.Ff , 14.60.Pq , 12.15.Hh . January 2000 As is well known, in addition to three active neutrinos $`\nu _e,\nu _\mu ,\nu _\tau `$, one sterile neutrino $`\nu _s`$, at least, is needed to explain in terms of neutrino oscillations three neutrino effects: the deficits of solar $`\nu _e`$’s and atmospheric $`\nu _\mu `$’s as well as the possible LSND excess of $`\nu _e`$’s in accelerator beam of $`\nu _\mu `$’s . This is a phenomenological reason for introducing sterile neutrinos. From the theoretical viewpoint, however, sterile neutrinos may exist in Nature, whether the LSND effect is real or not. In this paper, we describe a four–neutrino effective texture implying bimaximal mixing of $`\nu _e`$ with $`\nu _s`$ and $`\nu _\mu `$ with $`\nu _\tau `$, but, at the same time, only a tiny LSND effect, reduced by as much as four orders of magnitude in comparison with its original estimation. In our texture, the mass matrix for active neutrinos $`\nu _e,\nu _\mu ,\nu _\tau `$ gets the same form $`M=\left(M_{\alpha \beta }\right)(\alpha ,\beta =e,\mu ,\tau )`$ as the mass matrix for charged leptons $`e^{},\mu ^{},\tau ^{}`$ (only the values of parameters are expected to be different). In order to operate with an explicit model, we accept in both cases the ansatz $$\left(M_{\alpha \beta }\right)=\frac{1}{29}\left(\begin{array}{ccc}\mu \epsilon & 2\alpha & 0\\ 2\alpha & 4\mu (80+\epsilon )/9& 8\sqrt{3}\alpha \\ 0& 8\sqrt{3}\alpha & 24\mu (624+\epsilon )/25\end{array}\right),$$ (1) where $`\epsilon >0`$, $`\mu >0`$ and $`\alpha >0`$ are three parameters, taking different values for neutrinos and charged leptons. In the case of charged leptons, the ansatz (1) leads for $`\alpha 0`$ to the prediction $$m_\tau 1776.80\mathrm{MeV},\epsilon 0.172329,\mu 85.9924\mathrm{MeV},$$ (2) if experimental values of $`m_e`$ and $`m_\mu `$ are used as an input. In fact, the lowest perturbative calculation with respect to $`\alpha /\mu `$, when applied to the eigenvalue equation for the matrix (1), gives in particular $`m_\tau `$ $`=`$ $`{\displaystyle \frac{6}{125}}\left(351m_\mu 136m_e\right)+10.2112\left({\displaystyle \frac{\alpha }{\mu }}\right)^2\mathrm{MeV}`$ (3) $`=`$ $`\left[1776.80+10.2112\left({\displaystyle \frac{\alpha }{\mu }}\right)^2\right]\mathrm{MeV}.`$ When the experimental value $`m_\tau ^{\mathrm{exp}}=1777.05_{0.26}^{+0.29}`$ is used, Eq. (3) implies $$\left(\frac{\alpha }{\mu }\right)^2=0.024_{0.025}^{+0.028},$$ (4) what is not inconsistent with $`\alpha =0`$ (then $`M`$ becomes diagonal). Impressive agreement of the prediction for $`m_\tau `$ with the experimental $`m_\tau ^{\mathrm{exp}}`$ is our phenomenological motivation for the use of form (1) as the lepton mass matrix $`M`$. Methodologically, we consider here our form (1) of $`M`$ as a detailed ansatz, though it can be somehow theoretically supported (the interested reader may find some arguments in Appendix to Ref. ). In contrast to the charged–lepton case, where $`\alpha /\mu 1`$ (and so, $`M`$ is nearly diagonal), we conjecture in the neutrino case that $`\mu /\alpha 1`$ (and it is small enough to get $`M`$ nearly off–diagonal). The reason is that only in such a situation we can expect nearly maximal neutrino mixing, namely of $`\nu _\mu `$ with $`\nu _\tau `$ as it is preferably suggested by Super–Kamiokande experiments on the deficit of atmospheric $`\nu _\mu `$’s . Then, in order to explain potentially also the deficit of solar $`\nu _e`$’s as well as the possible LSND effect for accelerator $`\nu _\mu `$’s , we accept the popular hypothesis that in Nature there is a sterile neutrino $`\nu _s`$ which may mix with active neutrinos $`\nu _e`$, $`\nu _\mu `$, $`\nu _\tau `$, dominantly with $`\nu _e`$. To construct an effective model of four–neutrino texture, we assume that the mass matrix for neutrinos $`\nu _s`$, $`\nu _e`$, $`\nu _\mu `$, $`\nu _\tau `$ has the $`4\times 4`$ form $`M=\left(M_{\alpha \beta }\right)(\alpha ,\beta =s,e,\mu ,\tau )`$, where $$M_{ss}=0,M_{se}=\lambda M_{e\mu }=M_{es},M_{s\mu }=0=M_{\mu s},M_{s\tau }=0=M_{\tau s}$$ (5) are seven new matrix elements, while the rest of them are old, as given in Eq. (1). Here, the ratio $`\lambda M_{se}/M_{e\mu }>0`$ is a neutrino fourth free parameter. The old neutrino free parameter $`\epsilon `$ will be put zero (as seen from Eq. (2), even for charged leptons $`\epsilon `$ is small). Then, $$M_{ee}=0,M_{\mu \mu }=\frac{4}{9}\mathrm{\hspace{0.17em}80}\frac{\mu }{29},M_{\tau \tau }=\frac{24}{25}\mathrm{\hspace{0.17em}624}\frac{\mu }{29}.$$ (6) The ratios $$\xi \frac{M_{\tau \tau }}{M_{e\mu }}=299.52\frac{\mu }{\alpha },\chi \frac{M_{\mu \mu }}{M_{e\mu }}=\frac{1}{16.848}\xi $$ (7) are small, when $`\mu /\alpha 1`$ is small enough. Now, solving the eigenvalue equation for the $`4\times 4`$ matrix $`M`$ in the first perturbative order with respect to $`\xi `$, we obtain the following neutrino masses: $`m_0`$ $`=`$ $`{\displaystyle \frac{2\alpha }{29}}\left\{{\displaystyle \frac{1}{\sqrt{2}}}\left[49+\lambda ^2\sqrt{(49\lambda ^2)^2+4\lambda ^2}\right]^{1/2}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{1}{49}}\xi \right\}{\displaystyle \frac{2\alpha }{29}}\left[\sqrt{{\displaystyle \frac{48}{49}}}\lambda +{\displaystyle \frac{1}{2}}{\displaystyle \frac{1}{49}}\xi \right],`$ $`m_1`$ $`=`$ $`{\displaystyle \frac{2\alpha }{29}}\left\{{\displaystyle \frac{1}{\sqrt{2}}}\left[49+\lambda ^2\sqrt{(49\lambda ^2)^2+4\lambda ^2}\right]^{1/2}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{1}{49}}\xi \right\}{\displaystyle \frac{2\alpha }{29}}\left[\sqrt{{\displaystyle \frac{48}{49}}}\lambda +{\displaystyle \frac{1}{2}}{\displaystyle \frac{1}{49}}\xi \right],`$ $`m_2`$ $`=`$ $`{\displaystyle \frac{2\alpha }{29}}\left\{{\displaystyle \frac{1}{\sqrt{2}}}\left[49+\lambda ^2+\sqrt{(49\lambda ^2)^2+4\lambda ^2}\right]^{1/2}+{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{48}{49}}\xi +\chi \right)\right\}`$ $``$ $`{\displaystyle \frac{2\alpha }{29}}\left[7+{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{48}{49}}\xi +\chi \right)\right],`$ $`m_3`$ $`=`$ $`{\displaystyle \frac{2\alpha }{29}}\left\{{\displaystyle \frac{1}{\sqrt{2}}}\left[49+\lambda ^2+\sqrt{(49\lambda ^2)^2+4\lambda ^2}\right]^{1/2}+{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{48}{49}}\xi +\chi \right)\right\}`$ (8) $``$ $`{\displaystyle \frac{2\alpha }{29}}\left[7+{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{48}{49}}\xi +\chi \right)\right].`$ Here, the second step is valid in the linear approximation in $`\lambda `$, what requires small $`\lambda /49`$, while the former perturbative calculation with respect to $`\xi `$ works for small $`\xi /7`$. We can conclude from Eqs. (8) that $`m_3\stackrel{>}{}|m_2|m_1\stackrel{>}{}|m_0|`$. The neutrino diagonalizing $`4\times 4`$ matrix $`U=\left(U_{\alpha i}\right)(\alpha =s,e,\mu ,\tau ,i=0,\mathrm{\hspace{0.17em}1},\mathrm{\hspace{0.17em}2},\mathrm{\hspace{0.17em}3})`$, such that $`U^{}MU=\mathrm{diag}(m_0,m_1,m_2,m_3)`$, gets in the zero order with respect to $`\xi `$ and in the linear approximation in $`\lambda `$ the following form: $$\left(U_{\alpha i}\right)\left(\begin{array}{cccc}\frac{1}{\sqrt{2}}& \frac{1}{\sqrt{2}}& \frac{\lambda }{49\sqrt{2}}& \frac{\lambda }{49\sqrt{2}}\\ \frac{\sqrt{48}}{7\sqrt{2}}& \frac{\sqrt{48}}{7\sqrt{2}}& \frac{1}{7\sqrt{2}}& \frac{1}{7\sqrt{2}}\\ \frac{\lambda }{49\sqrt{2}}& \frac{\lambda }{49\sqrt{2}}& \frac{1}{\sqrt{2}}& \frac{1}{\sqrt{2}}\\ \frac{1}{7\sqrt{2}}& \frac{1}{7\sqrt{2}}& \frac{\sqrt{48}}{7\sqrt{2}}& \frac{\sqrt{48}}{7\sqrt{2}}\end{array}\right)+O(\xi /7).$$ (9) Evidently, in this case $`\xi /7`$ ought to be smaller than $`\lambda /49`$. If the charged–lepton diagonalizing $`3\times 3`$ matrix is nearly unit due to the small value (4) of $`\alpha /\mu `$, the lepton counterpart $`V=\left(V_{i\alpha }\right)`$ of the quark Cabibbo—Kobayashi—Maskawa matrix is approximately equal to $`U^{}=\left(U_{}^{}{}_{i\alpha }{}^{}\right)=\left(U_{\alpha i}^{}\right)`$. Thus, in this approximation, the fields $$\nu _i=\underset{\alpha }{}V_{i\alpha }\nu _\alpha =\underset{\alpha }{}U_{\alpha i}^{}\nu _\alpha $$ (10) describe four massive neutrinos $`\nu _i(i=0,1,2,3)`$ in terms of four flavor neutrinos $`\nu _\alpha (\alpha =s,e,\mu ,\tau )`$. Hence, $$\nu _\alpha =\underset{i}{}U_{\alpha i}\nu _i,|\nu _\alpha =\nu _\alpha ^{}|0=\underset{i}{}U_{\alpha i}^{}|\nu _i.$$ (11) Then, the neutrino oscillation probabilities on the energy shell $`E`$ read $`P\left(\nu _\alpha \nu _\beta \right)`$ $`=`$ $`|\nu _\beta |e^{iPL}|\nu _\alpha |^2`$ (12) $`=`$ $`\delta _{\beta \alpha }4{\displaystyle \underset{j>i}{}}U_{\beta j}^{}U_{\alpha j}U_{\beta i}U_{\alpha i}^{}\mathrm{sin}^2x_{ji},`$ where $`L`$ denotes the experimental baseline and $$x_{ji}=1.27\frac{\mathrm{\Delta }m_{ji}^2L}{E},\mathrm{\Delta }m_{ji}^2=m_j^2m_i^2$$ (13) with $`\mathrm{\Delta }m_{ji}`$, $`L`$ and $`E`$ expressed in eV, km and GeV, respectively. In Eq. (12) the eigenvalues of momentum operator $`P`$ are $`p_i=\sqrt{E^2m_i^2}Em_i^2/2E`$. Evidently, because of real $`M_{\alpha \beta }`$ and thus real $`U_{\alpha i}`$, the possible CP violation is here neglected. From Eqs. (12) and (9) we calculate in the zero perturbative order with respect to $`\xi `$ and linear approximation in $`\lambda `$ the following oscillation probabilities: $`P\left(\nu _e\nu _e\right)`$ $``$ $`1{\displaystyle \frac{48^2}{49^2}}\mathrm{sin}^2x_{10}{\displaystyle \frac{448}{49^2}}\mathrm{sin}^2x_{21}{\displaystyle \frac{1}{49^2}}\mathrm{sin}^2x_{32},`$ $`P\left(\nu _\mu \nu _\mu \right)`$ $``$ $`1\mathrm{sin}^2x_{32},`$ $`P\left(\nu _\mu \nu _e\right)`$ $``$ $`{\displaystyle \frac{1}{49}}\mathrm{sin}^2x_{32}.`$ (14) In the first and third formula (14) we put approximately $`\mathrm{\Delta }m_{20}^2\mathrm{\Delta }m_{30}^2\mathrm{\Delta }m_{21}^2\mathrm{\Delta }m_{31}^2`$ due to Eqs. (8) with $`\xi 0`$ (then, a linear term in $`\lambda `$ appearing in the third formula vanishes). Note from Eqs. (8) that $$\mathrm{\Delta }m_{10}^2\frac{2}{49}\sqrt{\frac{48}{49}}\left(\frac{2\alpha }{29}\right)^2\lambda \xi ,\mathrm{\Delta }m_{32}^214\left(\frac{2\alpha }{49}\right)^2\left(\frac{48}{49}\xi +\chi \right)$$ (15) for small $`\lambda /49`$ and $`\xi /7`$. Here, $`\chi =5.9354\times 10^2\xi `$. If $`1.27\mathrm{\Delta }m_{32}^2L_{\mathrm{atm}}/E_{\mathrm{atm}}=O(1)`$ and $`\mathrm{\Delta }m_{32}^2\mathrm{\Delta }m_{\mathrm{atm}}^23\times 10^3\mathrm{eV}^2`$ , the second formula (14) is able to describe oscillations of atmospheric $`\nu _\mu `$’s (dominantly into $`\nu _\tau `$’s) with maximal amplitude. Then, the second Eq. (15) gives the estimate $$\alpha ^2\xi 4.3\times 10^2\mathrm{eV}^2.$$ (16) Hence, if one assumes reasonably that $`\alpha O(1\mathrm{eV})`$, one gets $`\xi O(10^2)`$. On the other hand, if $`1.27\mathrm{\Delta }m_{10}^2L_{\mathrm{sol}}/E_{\mathrm{sol}}=O(1)`$ and $`\mathrm{\Delta }m_{10}^2\mathrm{\Delta }m_{\mathrm{sol}}^210^5\mathrm{eV}^2`$ or $`10^{10}\mathrm{eV}^2`$ , the first formula (14) can describe respectively large–angle MSW oscillations or vacuum oscillations of solar $`\nu _e`$’s (dominantly into $`\nu _s`$’s) with nearly maximal amplitude. In fact, it implies $$P\left(\nu _e\nu _e\right)1\frac{48^2}{49^2}\mathrm{sin}^2x_{10}\frac{448+1}{249^2}1\frac{48^2}{49^2}\mathrm{sin}^2x_{10}$$ (17) due to $`x_{10}x_{32}x_{21}`$. Then, from the first Eq. (15) we get the estimate $$\alpha ^2\lambda \xi 5.2\times 10^2\mathrm{eV}^2\mathrm{or}\mathrm{\hspace{0.33em}\hspace{0.33em}5.2}\times 10^7\mathrm{eV}^2,$$ (18) respectively. Thus, we find from Eqs. (16) and (18) that $$\lambda 1.2\mathrm{or}\mathrm{\hspace{0.33em}\hspace{0.33em}1.2}\times 10^5,$$ (19) respectively. This shows that the matrix element $`M_{se}`$ is comparable or small versus $`M_{e\mu }`$. Evidently, only the first option (related to large–angle MSW oscillations of solar $`\nu _e`$’s) can be compatible with the mixing matrix (9) and so, with the oscillation formulae (14) leading to the nearly maximal mixing of $`\nu _s`$ with $`\nu _e`$. In fact, only in this option, $`\xi `$ may be smaller than $`\lambda `$, as required by the form (9) of neutrino mixing matrix. In the case of Chooz experiment searching for oscillations of reactor $`\overline{\nu }_e`$’s , where it happens that $`1.27\mathrm{\Delta }m_{32}^2L_{\mathrm{Chooz}}/E_{\mathrm{Chooz}}=O(1)`$, the first formula (14) leads to $$P\left(\overline{\nu }_e\overline{\nu }_e\right)1\frac{1}{49^2}\mathrm{sin}^2x_{32}\frac{248}{49^2}1$$ (20) since $`x_{10}x_{32}x_{21}`$. This is consistent with the negative result of Chooz experiment. The third formula (14) implies te existence of $`\nu _\mu \nu _e`$ neutrino oscillations with the amplitude equal to $`1/490.02`$ and the mass–squared scale given by $`\mathrm{\Delta }m_{32}^2`$. Such an amplitude is compatible with the LSND estimation, say, $`\mathrm{sin}^22\theta _{\mathrm{LSND}}0.02`$, but the mass–squared scale $`\mathrm{\Delta }m_{32}^2`$ — being equal to the atmospheric $`\mathrm{\Delta }m_{\mathrm{atm}}^23\times 10^3\mathrm{eV}^2`$ — is smaller than the LSND estimation, say, $`\mathrm{\Delta }m_{\mathrm{LSND}}^20.5\mathrm{eV}^2`$ roughly by two orders of magnitude. In conclusion, our four–neutrino effective texture may describe correctly both deficits of solar $`\nu _e`$’s and atmospheric $`\nu _\mu `$’s. Then, it predicts the existence of a tiny LSND effect of the magnitude reduced by four orders in comparison with the original LSND estimation. It is so, because $$\mathrm{sin}^2\left(1.27\frac{\mathrm{\Delta }m_{32}^2L_{\mathrm{LSND}}}{E_{\mathrm{LSND}}}\right)\mathrm{sin}^2\left(1.27\frac{10^2\mathrm{\Delta }m_{\mathrm{LSND}}^2L_{\mathrm{LSND}}}{E_{\mathrm{LSND}}}\right)10^4$$ (21) for $`1.27\mathrm{\Delta }m_{32}^2L_{\mathrm{LSND}}/E_{\mathrm{LSND}}1`$. This reduced LSND effect would be, therefore, practically unobservable (for original $`L=L_{\mathrm{LSND}}`$ and $`E=E_{\mathrm{LSND}}`$). Obviously, the experimental problem of existence of the LSND effect, or of another realization of $`\nu _\mu \nu _e`$ neutrino oscillations, is crucial for all discussions about neutrino texture. In particular, a clear confirmation of the original LSND effect would exclude our four–neutrino effective texture. In such a case, the option of three pseudo–Dirac neutrinos might be invoked to explain all three neutrino–oscillation effects: the deficits of solar $`\nu _e`$’s and atmospheric $`\nu _\mu `$’s as well as the LSND effect (cf. e.g., Refs. and ). This option involves three natural Majorana sterile neutrinos mixing nearly maximally with three Majorana active neutrinos, and produces three pairs of light mass–neutrino states. It is in contrast to the popular see–saw option, where the natural Majorana sterile neutrinos and Majorana active neutrinos practically do not mix, and where they produce heavy and light mass–neutrino states, respectively. In the see–saw option, small masses of the latter states are conditioned by large masses of the former. References 1. Cf. e.g., C. Giunti, Talk at the ICFA/ECFA Workshop, Lyon, July 1999, hep–ph/9910336; and references therein. 2. W. Królikowski, Acta Phys. Pol. B 30, 2631 (1999); and references therein. 3. Review of Particle Physics, Eur. Phys. J. C 3, 1 (1998). 4. W. Królikowski, ”Oscillations of the mixed pseudo–Dirac neutrinos”, Nuovo Cim. A (to appear); and references therein. 5. Y. Fukuda et al. (Super–Kamiokande Collaboration), Phys. Rev. Lett. 81, 1562 (1998). 6. Cf. e.g., J.N. Bahcall, P.I. Krastev and A.Y. Smirnov, Phys. Rev. D 58, 096016 (1998); hep–ph/9905220v2. 7. C. Athanassopoulos et al. (LSND Collaboration), Phys. Rev. Lett. 75, 2650 (1995); Phys. Rev. C 54, 2685 (1996); Phys. Rev. Lett. 77, 3082 (1996); 81, 1774 (1998). 8. M. Appolonio et al. (Chooz Collaboration), Phys. Lett. B 420, 397 (1998). 9. W. Królikowski, hep–ph/9910308 (to appear in Acta Phys. Pol. B); and references therein.
warning/0001/cond-mat0001093.html
ar5iv
text
# Breaking a one-dimensional chain: fracture in 1 + 1 dimensions ## I Introduction A stretched solid breaks instantaneously when the external widening stress exceeds a critical value corresponding to the limit of mechanical stability of the system. For smaller stresses the fracture is time-delayed and assisted by thermal fluctuations. At sufficiently low temperatures there can be quantum effects, and the breaking rate becomes temperature independent. This last situation is the main subject of this paper. We consider a one-dimensional system (for example, a polymer chain or nanotube) stretched by a constant external force applied at the chain ends at zero temperature, where quantum tunneling is the only mechanism responsible for the fracture. There have been several attempts in the past to study this problem however its many-body nature was not taken into account until the pioneering work of Dyakonov, who gave qualitatively correct results in the limits of weak and large external forces. A more precise analytical approach to the problem valid for small external forces has been suggested recently by Levitov et al. We will adopt a general formalism that allows us to look at various limits from a single viewpoint. According to Feynman the tunneling rate per unit length is given by $`w|\mathrm{exp}(A/\mathrm{})|^2`$ where A is the imaginary-time Action calculated along particle trajectories connecting the ”initial” (stretched) and ”final” (broken) states of the chain, $`\mathrm{}`$ is Planck’s constant, and the summation is performed over all trajectories. In the quasiclassical limit the trajectories contributing most to the sum satisfy the stationarity condition $`\delta A=0`$. These trajectories describe how the system traverses the classically- forbidden region, going from an unbroken chain into a configuration from which even a classical chain released at rest would break irreversibly. It is convenient to combine this motion with its time reversal to form a localized path called a ”bounce”. In the quasiclassical approximation the tunneling rate is given by $`wB\mathrm{exp}(A_{bounce}/\mathrm{})`$ (1) where the amplitude B and action $`A_{bounce}`$ are calculated for the bounce (there is no factor of 2 in the exponential from the squaring because the bounce covers both the past and future). We will demonstrate below that the problem of finding bounce solutions is identical to the classical problem of finding the equilibrium crack configuration for a two-dimensional solid stretched by a constant uniaxial stress – yet another example of the correspondence between one-dimensional quantum field theory and two-dimensional statistical mechanics. In view of this equivalence the results of and are already implied by the classical works of Griffith. The present contribution to this field contains two main results. First, in the limit of a large applied force close to the limit of mechanical stability we find an exact bounce solution; simultaneously this for the first time determines the equilibrium crack configuration in the large stress classical limit. Establishing the range of applicability of this solution reveals that fluctuations inevitably come into play in the immediate vicinity of the limit of mechanical stability. Second, we show how the bounce solutions can be used to provide insight into the real time post-tunneling dynamics. ## II General formalism The imaginary-time Action describing a one-dimensional chain stretched by an external force p applied to the chain ends has the form $`A={\displaystyle 𝑑x𝑑t[\frac{\rho }{2}(\dot{u}^2+c^2u^2)pu^{}]}+{\displaystyle 𝑑t[V(h)ph]}`$ (2) where $`\rho `$ is the linear particle density, $`x`$ is the spatial coordinate (along the chain length), $`t`$ is the imaginary time coordinate, $`u(x,t)`$ is the particle displacement field, and dots and primes stand for the imaginary time (t), and spatial (x) derivatives. The first integral is over all positions and times except for a small region near $`x=0`$, where the break will appear. This separates the line into two pieces, so that for all $`t`$ the field u is discontinuous: $`u(x=+0,t)u(x=0,t)=h(t)`$; however, we expect $`h(t)`$ to be small outside a well-defined time interval. We have defined $`h(t)`$ so that the distance between the two parts of the chain is $`a+h`$, where a is the equilibrium $`(p=0)`$ interparticle spacing in the chain. The coupling between the segments of the chain in the tear region is given by the potential of cohesive forces V(h) which describes the interaction between the two ends of the broken chain for arbitrary separation – that is, it goes beyond the harmonic approximation. The properties of the function V(h) can be summarized as follows (Fig. 1). For $`|h|a`$ it obeys Hooke’s law: $`V(h)=\rho c^2h^2/2a`$, so that the two parts of the chain are joined into one elastic medium. For large negative h, the cohesive potential increases without bound, $`V(h\mathrm{})+\mathrm{}`$, reflecting the impossibility of indefinite compression. For large positive h it approaches a constant, $`V(h+\mathrm{})=2\gamma `$, which is the work done in infinitely separating the halves of the chain; $`2\gamma `$ can also be called the bond energy and the factor of 2 is introduced to indicate that there is an energy $`\gamma `$ associated with each free end of the half- infinite chain pieces. In what follows we will also need to know the properties of the cohesive force function $`G(h)=dV/dh`$ shown schematically on Fig. 2: for small h it grows linearly with $`h,G(h)=\rho c^2h/a`$, reaching a maximal value $`p_c\rho c^2\gamma /a`$ at $`h=h_ca`$, while G(h) decreases to zero as $`h+\mathrm{}`$. The parameter $`h_c`$ can be called the critical bond lengthening, while $`p_c`$ is the limit of mechanical stability of the system – indeed only for $`p<p_c`$ does $`U(h)=V(h)ph(`$the total potential energy in the external field) have a metastable minimum at $`h=h_1(`$Fig.3) corresponding to the equilibrium stretched chain. The bounce solution is shown schematically on Fig.4 using the imaginary time variable $`t`$ as a second (space-like) coordinate. The particle world lines belonging to the two ends of the breaking chain deviate significantly from each other for a time interval and then come back together. The external force p applied at the chain ends is time-independent, therefore it is shown schematically by series of arrows of the same length pointing in the spatial (x) direction. The Action (2) and the Figures can be interpreted as representing a two-dimensional ”crystal” of particle world lines subject to a uniaxial stress p, and in this interpretation the bounce of Fig. 4 is a critical crack, poised between the small perturbations that can spontaneously heal and the large disruptions that lead to fracture. We note in passing that our point of view that the crack opening is nonzero everywhere is different from that of many classical approaches, which restrict the break to a finite region completely surrounded by an elastic medium. Our treatment contains the standard approach as a special case. To see the connection to the classical fracture problem explicitly let us seek extremal paths for the Action (3) by varying $`u`$ while keeping the boundary values $`h(t)`$ fixed. The condition $`\delta A/\delta u=0`$ reduces to the Laplace equation $`d^2u/dt^2+c^2u^{\prime \prime }=0`$ (3) The solution to (3) satisfying the boundary conditions $`u(x=\pm 0,t)=\pm h(t)/2`$ has the form $`u(x,t)={\displaystyle \frac{p}{\rho c^2}}x+{\displaystyle \frac{signx}{2}}{\displaystyle _{\mathrm{}}^+\mathrm{}}{\displaystyle \frac{d\omega }{2\pi }}h(\omega )\mathrm{exp}[i\omega t(|\omega ||x|/c)]`$ (4) $`{\displaystyle \frac{p}{\rho c^2}}x+{\displaystyle \frac{x}{2\pi c}}{\displaystyle _{\mathrm{}}^+\mathrm{}}{\displaystyle \frac{dt^{}h(t^{})}{(t^{}t)^2+x^2/c^2}}`$ (5) where the first term describes the stretched chain, the second is due to the break, and the Fourier transform $`h(\omega )=h(t)\mathrm{exp}(i\omega t)𝑑t`$ has been introduced. Substituting (5) in (2), integrating over x, and ignoring overall constants we find the Action in a form that depends only on the field h $`A={\displaystyle \frac{D}{2}}{\displaystyle _{|tt^{}|t_0}}\dot{h}(t)𝑑t\mathrm{ln}(|tt^{}|/t_0)\dot{h}(t^{})𝑑t^{}+{\displaystyle 𝑑t[V(h)ph]}`$ (6) where $`D=\rho c/2\pi `$ and $`t_0a/c`$ is a cutoff introduced to prevent singularities for $`tt^{}`$. If imaginary time is viewed as a space-like variable then the Action (6) can be interpreted as a Hamiltonian describing a ”crack” in a two-dimensional ”crystal” of particle world lines. The crack can be represented by a distribution of fictitious dislocations, and then the first term of (6) describes the logarithmic interaction of these dislocations with each other with an interaction strength D, a ”dislocation density” given by $`\dot{h}`$, and ”Burgers vectors” pointing along the chain. Similarly the second integral in (6) can be thought of as describing the cohesive interaction V(h) between the sides of the ”crack” in the presence of the external opening field p. A free-energy functional similar to (6) has been proposed in and as a field-theoretical starting point of classical fracture mechanics. The configuration $`h(t)`$ for which (6) is extremal is determined by the condition $`\delta A/\delta h=0`$: $`D{\displaystyle _{\mathrm{}}^+\mathrm{}}{\displaystyle \frac{\dot{h}(t^{})}{t^{}t}}𝑑t^{}=G(h)p`$ (7) Here and below the singular integral is taken as a principal value. A model of cracks based on a singular nonlinear integral equation of the form (7) has been analyzed by Blekherman and Indenbom. The sound velocity c sets the upper limit on the value of $`|\dot{h}|`$, therefore the physical solutions to (7) should satisfy $`|\dot{h}(t)/c|1`$; at the same time the field $`h(t)`$ can take on arbitrarily large values. A bounce solution to (7) will have the following form (Figs. 3, 4). For $`t\mathrm{}`$ the function $`h(t)`$ starts at the metastable minimum of the total potential $`U(h),h=h_1`$, then makes an excursion past the unstable maximum, $`h=h_2`$, and comes back to $`h=h_1`$ as $`t+\mathrm{}`$. Substituting (7) into (6) and subtracting from the result the Action of the stretched unbroken chain $`(h=h_1)`$, we find the tunneling Action that goes into the probability (1) $`A_{bounce}={\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑t[U(h)U(h_1){\displaystyle \frac{hh_1}{2}}{\displaystyle \frac{dU(h)}{dh}}]`$ (8) where we used the result due to Blekherman and Indenbom that $`_{\mathrm{}}^+\mathrm{}𝑑t[G(h)p]=0`$ (total zero force along the tear) for the solution to (7) satisfying $`h(\pm \mathrm{})=h_1`$. For large $`|t|`$ we can use a harmonic approximation, since the function $`h(t)`$ is close to its asymptotic limit $`h_1`$. Thus we introduce $`\phi =hh_1`$ and approximate the right-hand side of (7) by $`U^{\prime \prime }(h_1)\phi `$, valid for $`|\phi |a`$. Expanding the integrand of (7) in $`t^{}/t`$ and noting that $`h(t)`$ is even in $`t`$ we find $`\phi (t\pm \mathrm{})={\displaystyle \frac{D}{t^2U^{\prime \prime }(h_1)}}{\displaystyle _{\mathrm{}}^+\mathrm{}}\phi (t^{})𝑑t^{}`$ (9) However, this is not very interesting, since within the harmonic approximation the tunneling Action (8) vanishes. Further progress can be made by looking separately at the cases of weak and strong tearing force where the tunneling Action can be computed in a controlled fashion. ## III Small-force limit, p$``$ p<sub>c</sub> In studying the weak force limit, $`pp_c`$, we can import some ideas from classical fracture mechanics. The position of the unstable maximum $`(h=h_2)`$ of the function U(h) (Fig.3) shifts to infinity as $`p0`$. However the amplitude of the bounce solution must be larger than $`h_2`$, thus implying that over some time the distance between the edges of the tear is much bigger than equilibrium interparticle spacing. Since the bounce tail (9) does not contribute to the tunneling Action (8) in quadratic order, we can take h to be nonzero only within a time interval $`L<t<L(`$Fig.4). The tunneling time 2L will be determined later. The cohesive force, G(h), is negligible for $`LtL`$ and operative beyond this interval. The latter feature implies that there is a strain outside the segment $`[L;L]`$ that takes care of the assumed $`h(t)=0`$. Then (7) simplifies to $`D{\displaystyle _L^{+L}}{\displaystyle \frac{\dot{h}(t^{})}{t^{}t}}𝑑t^{}=p`$ (10) for $`Lt`$ L. Neglect of the cohesive forces within the tunneling interval implies that the edges of the tear are force free, i.e. $`u^{}(x=\pm 0,t)=0`$; this condition applied to the representation (5) with the limits of the integration set by $`L`$ and $`L`$ indeed reproduces (10). In the same approximation the total potential $`U(h)=V(h)ph`$ can be replaced by $`U(h)=2\gamma ph`$, and (8) simplifies to $`A_{bounce}={\displaystyle _L^{+L}}𝑑t(2\gamma {\displaystyle \frac{1}{2}}ph)`$ (11) where for $`pp_c`$ we neglected $`U(h_1)+ph_1/2=p^2a/\rho c^2`$ compared to $`2\gamma `$. The solution to (10) satisfying $`h(\pm L)=0`$ is the elliptical crack of fracture mechanics : $`h(t)={\displaystyle \frac{p}{\pi D}}\sqrt{L^2t^2}{\displaystyle \frac{2p}{\rho c}}\sqrt{L^2t^2}`$ (12) Substituting this into (11) and evaluating the integral we find $`A_{bounce}(L)=4\gamma Lp^2L^2/4D`$ (13) This expression determines the Action parameterized by the time $`2L`$, which is chosen so that $`A_{bounce}(L)`$ is extremal: then the Action is stationary with respect to all variations in the path connecting broken to unbroken configurations. Eq. (13) has a maximum at $`2L_G=16\gamma D/p^28\gamma \rho c/\pi p^2`$ (14) which is identified as the tunneling time. In the context of classical fracture mechanics this is called the Griffith criterion . Combining (1), (13), and (14) we find that for $`pp_c`$ the fracture rate (per unit time and per unit length of the chain) is given by $`w(1/cL_G^2)\mathrm{exp}[A_{bounce}(L_G)/\mathrm{}](p^4/\gamma ^2D^2c)\mathrm{exp}(16\gamma ^2D/\mathrm{}p^2)`$ (15) $`(p^4/\gamma ^2\rho ^2c^3)\mathrm{exp}(8\gamma ^2\rho c/\pi \mathrm{}p^2)`$ (16) where the prefactor B from (1) has been estimated by arguing that the break can occur anywhere in two-dimensional space-imaginary time: the prefactor $`1/cL_G^2`$ is the density of independent bounces that can fill up two-dimensional space- imaginary time (to be independent, two bounces must be separated in time by more than $`L_G`$ and in space by more than $`cL_G)`$. The effective mass involved in the tunneling can be estimated from an argument parallel to that of Dyakonov. The representation (5) implies that the effect of an inhomogeneity of size $`L_G`$ in the t-direction perturbs the picture of the world lines over a distance of order $`cL_G(`$in the x- direction) containing about $`cL_G/a`$ atoms. Each of them has a mass $`m=\rho a`$, therefore the effective mass $`M`$ involved in the tunneling is found to be $`Mm(cL_G/a)m(\rho c^2\gamma /ap^2)m(p_c/p)^2`$ (17) The tunneling time $`2L_G`$ (14), the maximal width of the tear \[see Eq(12)\], $`h(0)=(p/\pi D)L_G=8\gamma /\pi p`$, and the tunneling mass (17) all diverge as $`p0`$ implying that fracture through tunneling has many-body nature. Here as in fracture mechanics this approach breaks down in the vicinity of $`t=\pm L`$ where the derivative $`\dot{h}(t)`$ diverges. There is also an associated stress singularity (which we will not discuss here) near $`t=\pm L`$. These defects can be remedied by taking proper account of the cohesive forces with the general conclusion that all the scaling dependencies predicted in the framework of the Griffith approximation are correct and that the numerical factors entering Eqs.(13)- (16) are accurate in the limit $`pp_c`$; these comments apply equally well to the quantum fracture problem. For example, the range of applicability of Eq(12) can be found by requiring that both $`|\dot{h}|c`$ and $`ha`$ with the conclusion that independently of the value of the tearing force p, (12) can be trusted outside a small region – a dozen microscopic time spans of order $`a/c`$ in the vicinity of $`t=\pm L_G`$. For $`pp_c`$ the tail of the true solution (9) becomes $`\phi (t\pm \mathrm{})={\displaystyle \frac{apL_G^2}{\rho c^2t^2}}a(p/p_c)(L_G/t)^2`$ (18) which is generally small since $`pp_c`$. The analysis given above shows that for weak breaking forces $`(pp_c)`$ the true bounce solution is approximated quite well by Eq(12) for $`|t|<L`$, and by Eq.(18) for $`|t|>L`$, outside of the vicinity of $`|t|=L=L_G`$. The determination of the size of the transient region (expected to be a dozen microscopic time scales $`a/c)`$ and the precise functional form $`h(t)`$ inside it require a detailed knowledge of the cohesive potential V(h); it should not affect the main conclusions (14)-(17). Apart from numerical and preexponential factors, the results (14) and (16) were first found by Dyakonov using heuristic arguments based on the tunneling properties of a particle of effective mass (17) passing through a triangular potential barrier. A conformal mapping technique similar to what is employed in fracture mechanics was used by Levitov et al; this reproduced Eq(12) and (13), the numerical factors entering in Eq.(14) and the exponential of Eq.(16). However, the link to classical fracture mechanics has not been previously noticed. Experimental verification of the weak tension results will be hindered by the exponentially large lifetime of the chain \[see Eq.(16)\]. Therefore the case of large tearing force (p close to $`p_c)`$, where the lifetime is shorter, seems more important from the practical standpoint. ## IV Vicinity of the limit of mechanical stability, $`\mathrm{\Delta }pp_c`$ For $`\mathrm{\Delta }p=p_cpp_c`$ the tunneling still has a collective nature, because the limit of mechanical stability, $`p=p_c`$, is formally similar to the spinodal of magnetic systems . In this limit the tunneling involves the vicinity of $`h=h_c(`$Fig.2) and the bounce amplitude is very small. Regardless of the underlying interactions the cohesive force can be approximated here as $`G(h)=p_cb(hh_c)^2`$ where $`b=\frac{1}{2}d^3V(h=h_c)/dh^3`$. The zeros of $`G(h)p=\mathrm{\Delta }pb(hh_c)^2`$ determine the positions of the metastable minimum $`(h_1)`$ and unstable maximum $`(h_2)`$ of the total potential energy function U(h) (see Fig. 3): $`h_{1,2}=h_c\sqrt{\mathrm{\Delta }p/b}`$ (19) Expressing $`G(h)p`$ in terms of $`\phi =hh_1`$ we find $`G(h)p=2\sqrt{\mathrm{\Delta }pb}\phi b\phi ^2`$ (20) Similarly the potential energy function $`U(h)=V(h)ph`$ can be written as $`U(h)=U(h_1)+\sqrt{\mathrm{\Delta }pb}\phi ^2{\displaystyle \frac{b}{3}}\phi ^3`$ (21) and Eqs. (7) and (20) give the equation for the bounce: $`D{\displaystyle _{\mathrm{}}^+\mathrm{}}{\displaystyle \frac{\dot{\phi }(t^{})dt^{}}{t^{}t}}=2\sqrt{\mathrm{\Delta }pb}\phi b\phi ^2`$ (22) Close to the limit of mechanical stability the properties of strongly stretched unbroken chain substantially deviate from those of the equilibrium one: the linear mass density $`\rho `$ is significantly reduced, and the lattice looses its stiffness at $`p=p_c`$. The latter means that both the sound velocity c, and the parameter $`D=\rho c/2\pi `$ entering (22) vanish as $`\mathrm{\Delta }p0`$. The functional dependencies can be recovered by comparing the harmonic terms of (21) and (2), $`\sqrt{\mathrm{\Delta }pb}\rho c^2/a`$, which gives us $`D(m^2b\mathrm{\Delta }p)^{1/4},ca(b\mathrm{\Delta }p/m^2)^{1/4}`$ (23) The solution to Eq.(22) is $`\phi (t)=\mathrm{\Phi }{\displaystyle \frac{\xi ^2}{t^2+\xi ^2}},`$ (24) where $`\mathrm{\Phi }=4\sqrt{\mathrm{\Delta }p/b}`$ (25) $`\xi ={\displaystyle \frac{D\pi }{2\sqrt{\mathrm{\Delta }pb}}}(m^2/b\mathrm{\Delta }p)^{1/4}`$ (26) Combining (21) and (24)-(26) we see that the general ”sum rule” (9) is indeed satisfied. The value of the bounce amplitude $`\phi (t=0)=\mathrm{\Phi }`$ is small close to $`p=p_c`$; at the same time $`h(0)=\mathrm{\Phi }+h_1=h_c+3\sqrt{\mathrm{\Delta }p/b}=h_2+2\sqrt{\mathrm{\Delta }p/b}>h_2`$, i.e. as expected for the bounce solution, the function $`h(t)`$ goes past $`h_2`$ for some time. The divergent time scale $`\xi `$ (26) can be interpreted as a tunneling time, and the effective tunneling mass $`M`$ can be estimated by replacing $`L_G`$ by $`\xi `$ in (17): $`Mm(c\xi /a)`$ m. We see that as $`\mathrm{\Delta }p0`$ the effective mass involved in tunneling is comparable to the single particle mass in agreement with Dyakonov. Substituting the potential energy function (21) into (8) with the bounce solution given by (24)-(26), and computing the integral we find for the tunneling Action $`A_{bounce}=2\pi ^2D\mathrm{\Delta }p/b(m^2/b^3)^{1/4}(\mathrm{\Delta }p)^{5/4}`$ (27) which in view of (1) implies that for $`\mathrm{\Delta }pp_c`$ the tunneling rate is given by $`w(1/c\xi ^2)\mathrm{exp}(2\pi ^2D\mathrm{\Delta }p/b\mathrm{})(b\mathrm{\Delta }p/m^2a^4)^{1/4}\mathrm{exp}[const(m^2/b^3)^{1/4}(\mathrm{\Delta }p)^{5/4}/\mathrm{}]`$ (28) where the prefactor is taken from (16), $`L_G`$ is replaced by $`\xi `$ \[Eq(26)\], and the dimensionless constant under the exponential is of order unity. Our results (27) and (28) in the limit $`\mathrm{\Delta }pp_c`$ reproduce those of Dyakonov even though he could not find the exact bounce solution (24)-(26). Even in the absence of the exact solution it can be seen that Eq(22) has a solution of the form $`\phi (t)=4\sqrt{\mathrm{\Delta }p/b}f(t/\xi )`$ where $`\xi `$ is given by (26) and f(z) is a scaling function obeying an integral equation that has no free parameters at all. This identifies $`\xi (\mathrm{\Delta }p)^{1/4}`$ with the tunneling time, and then the tunneling Action is estimated as proportional to $`(\mathrm{\Delta }p)^{3/2}(\mathrm{\Delta }p)^{1/4}=(\mathrm{\Delta }p)^{5/4}`$ in agreement with the calculation that led to (27). In view of the equivalence between the bounce solutions and equilibrium cracks of the corresponding two-dimensional classical problem, Eqs. (24)-(26) also give the equilibrium crack configuration in the limit of large stresses close to the limit of mechanical stability $`p_c`$. In this correspondence the imaginary time t becomes a spatial coordinate along the crack, D is an elastic constant \[still vanishing according to (23)\], $`\xi `$ is the effective length of the crack, and the parameter m will have a meaning of a typical binding energy divided by interparticle spacing. The rate of the crack nucleation will be given by (28) with the physical temperature $`T`$ substituting for the Planck’s constant $`\mathrm{}`$. The classical problem of finding the equilibrium crack configuration in the large stress limit close to the stability threshold was previously considered in . The conclusions of that work differ from ours for two reasons. First, it was assumed that the long-range nature of the interaction between the different parts of the crack profile through the bulk of crystal can be replaced by a finite-range interaction; as a result the equation for the profile is differential rather than integral \[see (22)\]. However, this is certainly not true for materials obeying conventional elasticity theory. Second, it was assumed that for small $`\phi `$, $`G(h)p=\mathrm{\Delta }pb\phi ^2`$, instead of our Eq(20). But this cannot be correct, because for $`\phi =0(`$metastable equilibrium) the cohesive force G(h) balances the external force p and therefore $`G(h)p`$ must vanish. ## V Levanyuk-Ginzburg criterion and critical region According to (27) the tunneling Action vanishes upon approach to the limit of mechanical stability $`p=p_c`$. However, the results of the previous Sections were based on a quasiclassical picture which is only valid for $`A_{bounce}/\mathrm{}1`$, i.e. when the exponential factor of (1) dominates the breaking rate. This condition imposes a range of validity for our results $`\mathrm{\Delta }p/p_c(b^3\mathrm{}^4/m^2p_c^5)^{1/5}`$ (29) that eliminates from consideration the immediate vicinity of the threshold of mechanical stability $`p=p_c`$. This condition is analogous to the Levanyuk-Ginzburg criterion of weakness for thermal fluctuations in the theory of critical phenomena. The results (23)-(28) were also obtained under the assumption that $`\mathrm{\Delta }p/p_c1[`$only then is the expansion in powers of $`\phi `$ used in (20) - (22) accurate\]. This will be consistent with (29) only if $`\mathrm{𝐆𝐢}(b^3\mathrm{}^4/m^2p_c^5)^{1/5}1`$ (30) The dimensionless quantity Gi is a property of a given substance and is analogous to the Ginzburg parameter of the theory of critical phenomena. In the present context the requirement $`\mathrm{𝐆𝐢}1`$ can be viewed as the condition of weakness of quantum fluctuations. The inequalities (29) and (30) set the range of applicability of the theory of Section IV. The immediate vicinity of $`p=p_c`$ corresponding to the reversed sign in the inequality (29) can be called the fluctuational region; quantum fluctuations play here a dominant role and the quasiclassical approximation is insufficient. In the case of the reversed sign in the inequality (30), $`\mathrm{𝐆𝐢}1`$, the theory of Section IV has no range of applicability at all. For the case that one of the conditions (29), (30) is broken, our analytic approach fails, but some understanding can be gained with the aid of a scaling theory constructed in analogy with the theory of critical phenomena. Assume that in the critical regime there is a single independent divergent time scale $`\xi `$, the tunneling time, and all other physical quantities can be expressed in terms of $`\xi `$. According to (21), the barrier for tunneling can be estimated as $`(\mathrm{\Delta }p)^{3/2}/b^{1/2}`$, and therefore the tunneling Action behaves as $`\xi (\mathrm{\Delta }p)^{3/2}/b^{1/2}`$. In the critical region the tunneling Action should be of the order of Planck’s constant $`\mathrm{}`$, so that $`\xi \mathrm{}b^{1/2}/(\mathrm{\Delta }p)^{3/2}`$ (31) We also expect that the tunneling mass is still of order m which implies $`ca/\xi `$ for the critical behavior of the sound velocity. These conjectures are made plausible by two mild self-consistency checks: (i) the divergence in (31) is stronger than (26)) as expected for the critical regime; (ii) the dependencies (31) and (26) match each other on the border of the critical regime defined by (29). The conditions (29) and (30) have the classical analogs that the results of Section IV are applicable to the classical fracture of solids whenever the following inequalities are satisfied $`\mathrm{\Delta }p/p_c(b^3T^4/m^2p_c^5)^{1/5}`$ (32) $`\mathrm{𝐆𝐢}(b^3T^4/m^2p_c^5)^{1/5}1`$ (33) Now the constraints (32) and (33) describe the conditions for thermal fluctuations to be sufficiently weak so that the critical crack described by (24)-(26) is relevant for the case of large applied stresses. Inside the critical region, $`\mathrm{\Delta }p/p_c(b^3T^4/m^2p_c^5)^{1/5}`$, thermal fluctuations play a dominant role and the mean-field-like treatment of Section IV is invalid. If the Ginzburg parameter (33) is not small compared to unity, the theory of Section IV does not apply. In the critical regime the analog of (31), $`\xi Tb^{1/2}/(\mathrm{\Delta }p)^{3/2},`$ (34) gives the dependence of the critical crack length $`\xi `$ on the distance to the instability threshold $`\mathrm{\Delta }p`$. ## VI Real time dynamics Coleman notes that the analytic continuation of the bounce to real time describes the evolution of the system after the tunneling takes place. Introducing the physical time $`\tau =`$ -it into (12) we find that for a small tearing force, $`pp_c`$, the separation between the pieces of the chain evolves with time as $`h(\tau )={\displaystyle \frac{2p}{\rho c}}\sqrt{L_G+\tau ^2}`$ (35) where L<sub>G</sub> is the true tunneling time (14). Equation (35) meets our physical expectations: after penetrating through the barrier at $`t=i\tau =0`$, the end of the chain starts moving classically from rest from the escape position $`h(0)=2pL_G/\rho c=8\gamma /\pi p`$ which is the turning point of the classical motion. The result (35) is also precise analytically as in getting (12) the potential energy function U(h) was approximated by $`2\gamma ph`$ which is even more accurate in the classically allowed region than under the barrier. For small times $`(\tau L),Eq`$.(35) predicts a motion with a constant acceleration $`2p/\rho cL_G`$ that can be understood as follows. Immediately after the break each point of the chain excepting a finite segment near the edge will be in mechanical equilibrium. Therefore the whole unbalanced force p is applied to the edge segment. The mass of this region is the same as the tunneling mass (17), $`M\rho cL_G`$, and its acceleration due to the external force p is of order $`p/Mp/\rho cL_G`$ in agreement with (35). This motion breaks the force balance of the region next to the moving edge segment causing the propagation of acceleration away from the tear. To understand the dynamics inside the chain we use the expression for the displacement field $`u(x,t)`$ corresponding to the solution (12): $`u(x,t)={\displaystyle \frac{p}{\rho c^2}}Re[(x+ict)^2+(cL_G)^2]^{1/2}`$ (36) Analytically continuing this to the real time $`\tau =it`$, we find $`u(x,\tau )={\displaystyle \frac{p}{2\rho c^2}}signx\left([(xc\tau )^2+(cL_G)^2]^{1/2}+[(x+c\tau )^2+(cL_G)^2]^{1/2}\right)`$ (37) This expression is a solution to the wave equation $`^2u/\tau ^2c^2^2u/x^2=0`$ in the two parts of the broken chain satisfying the conditions $`u(\pm 0,\tau )=\pm h(\tau )/2`$ and $`u^{}(\pm 0,\tau )=0`$. It describes sound waves propagating away from the break and leaving behind virtually unstrained material moving at constant speed. Not much is happening for $`\tau <L_G`$; however, for $`\tau L_G`$ the following picture emerges (by symmetry we only need to consider the $`x>0`$ segment). For $`xc\tau `$ the strain $`u^{}`$ is small and $`u(x,\tau )p\tau /\rho c`$, whereas for $`xc\tau ,u/\tau `$ is small and $`u(x,\tau )px/\rho c^2:a`$ moving unstrained region encounters a stretched immobile region at $`x=c\tau `$. In a transition region having the width of order $`cL_G`$ the matter is beginning to move to become part of the moving region. The unstrained region set into motion has mass that grows linearly with time $`[M(\tau )=mc\tau /a=\rho c\tau ]`$ and moves at the constant speed $`v=p/\rho c`$; the momentum $`M(\tau )v=p\tau `$ is growing linearly in time in response to the constant applied force p. Figure 5 shows how this soliton-like wave traveling away from the break is presented in terms of particle trajectories. The break occurred at $`x=0`$, $`\tau =0`$, and the Figure was constructed by adding $`u(x,\tau )`$ \[given by Eq.(37)\] to $`x`$ for each $`\tau `$. In the large force limit, $`\mathrm{\Delta }pp_c`$, we analytically continue the bounce solution (23) to find that the distance between the tearing edges of the chain evolves with time as $`h(\tau )=h_1+\mathrm{\Phi }{\displaystyle \frac{\xi ^2}{\xi ^2\tau ^2}}`$ (38) Similar to the case of small tearing force, $`pp_c`$, the initial stage of the motion, $`\tau \xi ,`$, described by Eq(38) can be interpreted as a constant acceleration of the tunneling mass m by the unbalanced force $`\mathrm{\Delta }p`$. However the range of applicability of (38) is rather narrow – Eqs(24)-(26) \[and thus (38)\] were obtained assuming the expansions (20)) and (21) to be valid. At the same time the distance between the tearing edges grows with time according to (38), thus implying that as time progresses, $`h(\tau )h_1`$ becomes comparable to the equilibrium interparticle spacing a and the expansions (20) and (21) cease to be accurate. Therefore (38) is valid only for $`\tau \xi `$. In the same approximation the motion in the chain can be obtained by substituting (24) in (5), evaluating the integral, and doing analytic continuation to the real time $`\tau =it`$: $`u(x,\tau )={\displaystyle \frac{p_c}{\rho c^2}}x+{\displaystyle \frac{signx}{2}}\left(h_1+\mathrm{\Phi }\xi {\displaystyle \frac{|x|/c+\xi }{[|x|/c+\xi ]^2\tau ^2}}\right)`$ (39) where the sound velocity c is given by (23). Beyond the range $`\tau \xi `$, the dynamics is expected to be described by a theory qualitatively similar to the one given above for the case $`pp_c`$ with $`p_c`$ and $`\xi `$ substituting for p and $`L_G`$. ###### Acknowledgements. Our interest to this subject was initiated by very fruitful arguments with A.Buchel and J.P.Sethna. We also thank K.Burton, C.L.Henley, H.Hui, S.L.Phoenix, and A.V.Shytov for critical remarks.
warning/0001/astro-ph0001478.html
ar5iv
text
# Conversion of neutron stars to strange stars as the central engine of 𝛾-ray bursts ## 1 Introduction There is now compelling evidence to suggest that a substantial fraction of all gamma-ray bursts (GRBs) occur at cosmological distances (red shift $`z13`$). In particular, the measured red shift $`z=3.42`$ for GRB 971214 (Kulkarni et al. 1998), and $`z1.6`$ for GRB 990123 (Kulkarni et al. 1999) implies an energy release of $`3\times 10^{53}`$erg and $`3.4\times 10^{54}`$erg respectively, in the $`\gamma `$-rays alone, assuming isotropic emission. The latter energy estimate could be substantially reduced if the energy emission is not isotropic, but displays a jet-like geometry (Dar 1998; Kulkarni et al. 1999). Models in which the burst is produced by a narrow jet are able to explain the complex temporal structure observed in many GRBs (Sari et al. 1997; 1999). In any case, a cosmological origin of GRBs leads to the conclusion of a huge energy output. Depending on the degree of burst beaming and on the efficiency of $`\gamma `$-ray production, the central engine powering these extraordinary events should be capable of releasing a total energy of a few $`10^{53}`$erg. Many cosmological models for GRBs have been proposed. Among the most popular is the merging of two neutron stars (or a neutron star and a black hole) in a binary system (Paczynski 1998). Recent results (Janka & Ruffert 1996) within this model, indicate that, even under the most favorable conditions, the energy provided by $`\nu \overline{\nu }`$ annihilation during the merger is too small by at least an order of magnitude, and more probably two or three orders of magnitude, to power typical GRBs at cosmological distances. An alternative model is the so–called ”failed supernova” (Woosley 1993), or ”hypernova” model (Paczynski 1998). In the present work, we consider the conversion of a neutron star to a strange star (hereafter NS$``$SS conversion) as a possible central engine for GRBs. In particular, we focus on the energetics of the NS$``$SS conversion, and not on the mechanism by which $`\gamma `$-rays are produced. Previous estimate of the total energy $`E^{conv}`$ released in the NS$``$SS conversion (Olinto 1987; Cheng & Dai 1996) or in the conversion of a neutron star to hybrid star (Ma & Xie 1996) gave $`E^{conv}10^{52}`$erg, too low to power GRBs at cosmological distances. These calculations did not include the various details of the neutron star and strange star structural properties, which go into the binding energy release considerations. Here, we present accurate and systematic calculations of the total energy released in the NS$``$SS conversion using different models for the equation of state (EOS) of neutron star matter (NSM) and strange quark matter (SQM). We show that the total amount of energy liberated in the conversion is in the range $`E^{conv}=`$(1—4)$`\times 10^{53}`$erg, in agreement with the energy required to power $`\gamma `$-ray burst sources at cosmological distances. The existence of strange stars (made up of degenerate u, d, and s quarks in equilibrium with respect to the weak interactions) is allowable within uncertainties inherent in our present theoretical understanding of the physics of strongly interacting matter (Bodmer 1971; Witten 1984; Farhi & Jaffe 1984). Thus, strange stars may exist in the universe, but until now, these have remained purely speculative entities. This situation changed in the last few years, thanks to the large amount of new observational data collected by the new generation of X-ray satellites. In fact, recent studies have shown that the compact objects associated with the X-ray bursters GRO J1744-28 (Cheng et al. 1998), SAX J1808.4-3658 (Li et al. 1999a), and with the X-ray pulsar Her X-1 (Dey et al. 1998) are good strange star candidates. Recently, Li et al. (1999b) have shown that the observed high– and low–frequency Quasi Periodic Oscillations (QPOs) in the atoll source 4U 1728-34 (Méndez & van der Klis 1999), are more consistent with a strange star compared to a neutron star, if the model of Osherovich & Titarchuk (1999) correctly interprets the QPO phenomena. Originally, the idea that GRBs could be powered by the conversion of a neutron star to a strange star was proposed by Alcock et al. (1986) (see also Olinto (1987)), and recently reconsidered by other authors (Cheng & Dai 1996). A similar model has been discussed by Ma and Xie (1996) for the conversion of a neutron star to a so–called hybrid star (a neutron star with a strange quark matter core). A number of different mechanisms have been proposed for the NS$``$SS conversion. All of them are based on the formation of a “seed” of SQM inside the neutron star. For example, (i) a seed of SQM enters in a NS and converts it to a SS (Olinto 1987). These seeds of SQM, according to Witten (1984), are relics of the primordial quark–hadron phase transition microseconds after the Big Bang. (ii) a seed of SQM forms in the core of a neutron star as a result of a phase transition from neutron star matter to deconfined strange quark matter (NSM$``$SQM phase transition). This could possibly happen when a neutron star (NS) is a member of a binary stellar system. The NS accretes matter from the companion star. The central density of the NS increases and it may overcome the critical density for the NSM$``$SQM phase transition. The NS is then converted to a SS. In the case of accretion induced conversion in a binary stellar system, the conversion rate has been estimated (Cheng & Dai 1996) to be in the range (3–30)$`\times 10^{10}`$ conversions per day per galaxy. This rate is consistent with the observed GRBs rate. However, once there is a seed of SQM inside a neutron star, it is possible to calculate the rate of growth (Olinto 1987; Horvath & Benvenuto 1988). The SQM front absorbs neutrons, protons, and hyperons (if present), liberating their constituent quarks. Weak equilibrium is then re-established by the weak interactions. As shown by Horvath and Benvenuto (1988), the conversion of the whole star will then occur in a very short time (detonation mode), in the range 1 ms – 1 s, which is in agreement with the typical observed duration of GRBs. A detailed simulation of the conversion process is still lacking, and only rough estimates of the total energy liberated in the conversion have been made. As we show below, the dominant contribution to $`E^{conv}`$ arises from the internal energy released in the conversion, i.e. in the NSM$``$SQM phase transition. Moreover, the gravitational mass of the star will change during the conversion process, even under the assumption that the total number of baryons in the star is conserved. The total energy released in the NS$``$SS conversion is given by the difference between the total binding energy of the strange star BE(SS) and the total binding energy of the neutron star BE(NS) $$E^{conv}=BE(SS)BE(NS)$$ (1) In the present work we assume that the baryonic mass $`M_B`$ of the compact object is conserved in the conversion process, i.e $`M_B(SS)=M_B(NS)M_B`$. Then $`E^{conv}`$ is given in terms of the difference between the gravitational mass of the NS and SS: $`E^{conv}=[M_G(NS)M_G(SS)]c^2`$. In general the total binding energy for a compact object can be written $`BE=BE_I+BE_G=(M_BM_P)c^2+(M_PM_G)c^2`$, where $`BE_I`$ and $`BE_G`$ denote the internal and gravitational binding energies respectively, and $`M_P`$ is the proper mass of the compact object defined as $$M_P=_0^R𝑑r4\pi r^2\left[1\frac{2Gm(r)}{c^2r}\right]^{1/2}\rho (r),$$ (2) $`\rho (r)`$ being the total mass–energy density and $`m(r)`$ the gravitational mass enclosed within a spherical volume of radius $`r`$. The proper mass is equal to the sum of the mass elements on the whole volume of the star; it includes the contributions of rest mass and internal energy (kinetic and interactions (other than gravitation)) of the constituents of the star. The total conversion energy can then be written as the sum of two contributions $$E^{conv}=E_I^{conv}+E_G^{conv}$$ (3) related to the internal and gravitational energy changes in the conversion. These two contributions can be written as: $$E_I^{conv}=BE_I(SS)BE_I(NS)=[M_P(NS)M_P(SS)]c^2,$$ (4) $$E_G^{conv}=BE_G(SS)BE_G(NS)=[M_P(SS)M_G(SS)M_P(NS)+M_G(NS)]c^2,$$ (5) and these can be evaluated solving the structural equations for non-rotating compact objects (Oppenheimer & Volkoff 1939). To highlight the dependence of $`E^{conv}`$ upon the present uncertainties in the microphysics, we employed different models for the EOS of both NSM and SQM. Recently, a microscopic EOS of dense stellar matter has been calculated by Baldo et al. (1997), and used to compute the structure of static (Baldo et al. 1997) as well as rapidly rotating neutron stars (Datta et al. 1998). In this model for the EOS, the neutron star core is composed of asymmetric nuclear matter in equilibrium, with respect to the weak interactions, with electrons and muons ($`\beta `$-stable matter). In particular, we consider their EOS based on the Argonne $`v`$14 nucleon-nucleon interaction implemented by nuclear three-body forces (hereafter BBB1 EOS). At the high densities expected in the core of a neutron star, additional baryonic states besides the neutron and the proton may be present, including the hyperons $`\mathrm{\Lambda }`$, $`\mathrm{\Sigma }`$, $`\mathrm{\Xi }`$, $`\mathrm{\Omega }`$, and the isospin 3/2 nucleon resonance $`\mathrm{\Delta }`$. The equation of state of this hyperonic matter is traditionally investigated in the framework of Lagrangian field theory in the mean field approximation (Glendenning 1985; Schaffner & Mishustin 1996; Prakash et al. 1997). According to this model, the onset for hyperon formation in $`\beta `$-stable–charged neutral dense matter is about 2–3 times the normal nuclear matter density ($`n_0=0.17`$ fm<sup>-3</sup>). The latter result has been confirmed by recent microscopic calculations based on the Brueckner-Hartree-Fock theory (Baldo et al. 1998). The appearance of hyperons, in general, gives a softening of the EOS with respect to the pure nucleonic case. In the present work we considered one of the EOS for hyperonic matter given in Prakash et al. (1997). For strange quark matter, we consider a simple EOS (Farhi and Jaffe 1984) based on the MIT bag model for hadrons. We begin with the case of massless non-interacting ($`\alpha _c=0`$) quarks and with a bag constant $`B=60`$ MeV/fm<sup>3</sup>: we denote the corresponding EOS as $`B60_0`$. Next we consider a finite value for the mass of the strange quark within the same MIT bag model EOS. We take $`m_s=200`$ MeV (and $`m_u=m_d=0`$, $`B=60`$ MeV/fm<sup>3</sup>, and $`\alpha _c=0`$; hereafter EOS $`B60_{200}`$). To investigate the effect of the bag constant on the energy released in the NS$``$SS conversion, we take (almost) the largest possible value of $`B`$ for which SQM is still the ground state of strongly interacting matter, according to the so–called strange matter hypothesis (Witten 1984). For massless non-interacting quarks this gives $`B=90`$ MeV/fm<sup>3</sup>; we denote the corresponding EOS as $`B90_0`$. Recently, Dey et al. (1998) derived an EOS for SQM using a different quark model with respect to the MIT bag model. The EOS by Dey et al. has asymptotic freedom built in, shows confinement at zero baryon density, deconfinement at high density, and, for an appropriate choice of the EOS parameters entering in the model, gives absolutely stable SQM according to the strange matter hypothesis. In this model, the quark interaction is described by a screened inter–quark vector potential originating from gluon exchange, and by a density-dependent scalar potential which restores the chiral symmetry at high density. The density-dependent scalar potential arises from the density dependence of the in-medium effective quark masses $`M_q`$, which, in the model by Dey et al. (1998), are taken to depend upon the baryon number density $`n_B`$ according to $`M_q=m_q+310\mathrm{M}\mathrm{e}\mathrm{V}\times sech\left(\nu \frac{n_B}{n_0}\right)`$, where $`n_0`$ is the normal nuclear matter density, $`q(=u,d,s)`$ is the flavor index, and $`\nu `$ is a parameter. The effective quark mass $`M_q(n_B)`$ goes from its constituent masses at zero density, to its current mass $`m_q`$, as $`n_B`$ goes to infinity. Here we consider two different parameterizations of the EOS by Dey et al., which correspond to a different choice for the parameter $`\nu `$. The equation of state SS1 (SS2) corresponds to $`\nu =0.333`$ ($`\nu =0.286`$). ## 2 Results and conclusions To begin with, we fix as a “standard” EOS for neutron star matter the BBB1 EOS (Baldo et al. 1997), to explore how the energy budget in the NS$``$SS conversion depends on the details of the EOS for strange quark matter. First we consider the $`B60_0`$ equation of state. The NS$``$SS conversion based on this couple of EOSs, will be referred to as the BBB1$`B60_0`$ conversion model. Similar notation will be employed according to the EOS of NSM and SQM. The total conversion energy, together with the partial contributions, is shown in the upper panel of Fig. 1. As we can see, for $`M_B`$ larger than $`1M_{}`$ (i.e. values of the baryonic mass compatible with the measured neutron star gravitational masses) the total energy released in the NS$``$SS conversion is in the range (1–3)$`\times 10^{53}`$erg, one order of magnitude larger than previous estimates (Olinto 1987; Cheng & Dai 1996; Ma & Xie 1996). Moreover, contrary to a common expectation, the gravitational conversion energy $`E_G^{conv}`$ is negative for this couple of EOSs. To make a more quantitative analysis, we consider a neutron star with a baryonic mass $`M_B=1.574M_{}`$ (see tab. 1), which has a gravitational mass $`M_G=1.409M_{}`$, a radius R(NS) = 11.0 km, and a gravitational binding energy BE<sub>G</sub>(NS) = 4.497 $`\times 10^{53}`$ erg. After conversion, the corresponding strange star has $`M_G=1.254M_{}`$, R(SS) = 10.5 km, and BE<sub>G</sub>(SS) = 3.061 $`\times 10^{53}`$ erg. The NS$``$SS conversion is energetically possible in this case, thanks to the large amount of (internal) energy liberated in the NSM$``$SQM phase transition. Similar qualitative results for the total conversion energy are obtained for other choices of the two EOSs, but as we show below the magnitude of the two partial contributions are strongly dependent on the underlying EOS for NSM and SQM. The total conversion energy for the BBB1$`B60_{200}`$ model is plotted in the lower left panel of Fig. 1. Comparing with the previous case, we notice that the strange quark mass produces a large modification of the conversion energy, which is reduced by a factor between 2–3 with respect to the case $`m_s=0`$. The bag constant $`B`$ has also a sizeable influence on the conversion energy. Increasing the value of $`B`$ reduces $`E^{conv}`$ and strongly modifies $`E_G^{conv}`$. This can be seen comparing the results for the BBB1$`B60_0`$ conversion model with those in the lower right panel of Fig. 1 for the BBB1$`B90_0`$ model. These results are a consequence of the sizeable effects of the strange quark mass and of the bag constant mainly on the internal binding energy $`BE_I(SS)`$ for strange stars (see e.g. Bombaci 1999). In fact, all strange stars configurations within the $`B60_0`$ EOS are self–bound objects (i.e. $`BE_I(SS)>0`$). Strange star configurations within the $`B90_0`$ ($`B60_{200}`$) EOS are self–bound objects up to $`M_G0.8M_{}`$ ($`M_G1.6M_{}`$), to compare with the corresponding maximum gravitational mass $`M_{max}=1.60M_{}`$ ($`M_{max}=1.75M_{}`$). The results depicted in the two upper panels of Fig. 2 have been obtained using the EOS of Dey et al. (1998) for SQM, for two different choices of the parameter $`\nu `$ which controls the rate at which chiral symmetry is restored to the quark masses at high density. For $`\nu =0.286`$ (SS2), chiral symmetry is broken up to larger densities with respect to the case $`\nu =0.333`$ (SS1). The parameter $`\nu `$ has a strong influence on the internal binding energy of the strange star. In fact, we found that strange stars within the SS2 (SS1) EOS are self–bound objects up to $`M_G0.7M_{}`$ ($`M_G1.4M_{}`$), to compare with the maximum gravitational mass $`M_{max}=1.33M_{}`$ ($`M_{max}=1.44M_{}`$). This effect is the main source for the differences in the calculated conversion energies for the two conversion models BBB1$``$SS1 and BBB1$``$SS2. The next step in our study is to consider a different neutron star matter EOS, which allows for the presence of hyperons in the neutron star core. We consider one of the EOS (hereafter Hyp) for hyperonic matter given in Prakash et al. (1997). In the two lower panels of Fig. 2 we plot the total conversion energy, together with the partial contributions, for the Hyp$`B60_0`$ and for the Hyp$``$SS1 conversion models. These results are in qualitative agreement with those reported in the previous figures. In table 1, we report the conversion energy, together with the partial contributions for the conversion of a neutron star with a gravitational mass $`M_G(NS)1.4M_{}`$ and for various conversion models. In the present work, we considered the conversion of a neutron star to a strange star as a possible energy source for $`\gamma `$-ray bursts. Our main focus was to perform an accurate calculation of the total released energy compatible with our current understanding of the microphysics of strong interacting matter. We showed that the total amount of energy liberated in the NS$``$SS conversion is in the range (1—4)$`\times 10^{53}`$erg, one order of magnitude larger than previous estimates, and in agreement with the energy required to power $`\gamma `$-ray burst sources at cosmological distances. One of the authors (I.B.) thanks Prof. J.E. Horvath for valuable discussions during the Quark Matter Conference in Torino. FIGURE CAPTIONS Fig. 1. The total energy liberated in the conversion of a neutron star to a strange star and the partial contributions from internal energy $`E_I^{conv}`$ (curves labeled “Int”) and from gravitational energy $`E_G^{conv}`$ (curves labeled “Gra”) as a function of $`M_B`$. See text for details on the equations of state for neutron star matter and strange quark matter Fig. 2. Same as in figure 1, but for different conversion models
warning/0001/hep-th0001026.html
ar5iv
text
# Poisson-Sigma Models11footnote 1Talk presented by T.Schwarzweller ## 1 Introduction The Poisson-Sigma model is a gauge theory based on a Poisson algebra, i.e. it is a non-linear extension of an ordinary gauge theory, which is based on a linear Lie algebra. The class of these models, which are based on a non-linear Lie algebra, i.e. a finite W-algebra or a Poisson algebra, are known in this context as non-linear gauge theory . The Poisson-Sigma model associates to any Poisson structure on a finite-dimensional manifold a two-dimensional field theory . Choosing different Poisson structures leads to specific models which include most of the topological and semi-topological field theories which have been of interest in recent years. These include gravity models, non-abelian gauge theories and the Wess-Zumino-Witten model. Because of the non-linearity of the algebra the Poisson-Sigma model involves in the language of gauge theories an open gauge algebra, i.e. ​the algebra closes only on-shell. In such cases the Faddeev-Popov method of path integral quantization fails. Quantization procedures which rely on the BRST symmetry of the extended action are in principle more powerful . We find that the antifield formalism of Batalin and Vilkovisky is the most effective method to get a suitable action for the path integral quantization. The path integral for the Poisson-Sigma model was first discussed in a preliminary way by Schaller and Strobl in . In a recent paper Cattaneo and Felder used the pertubation expansion of the path integral in the covariant gauge to elucidate Kontsevich’s formula for the deformation quantization of the algebra of functions on a Poisson manifold . Kummer et.al. have investigated the special case of 2d dilaton gravity and they have calculated the generating functional using BRST methods . We have investigated in a recent paper the path integral quantization of the Poisson-Sigma model in a general gauge and derived an almost closed expression for the partition function . The article is structured as follows. In Section 2 we introduce the Poisson-Sigma model and show how the model reduces to known field theories. In Section 3 we construct the Batalin-Vilkovisky action and perform the calculation of the path integral. We also show how the derived partition function for the general model reduces under certain circumstances to the more familiar Yang-Mills case . Section 4 contains some concluding remarks. ## 2 The Poisson-Sigma Model ### 2.1 Poisson manifolds and the model A Poisson manifold $`(N,P)`$ is a smooth manifold $`N`$ equipped with a Poisson structure $`P\mathrm{\Lambda }^2TN`$ . In local coordinates $`X^i`$ on $`N`$ $$P=\frac{1}{2}P^{ij}(X)_i_j,$$ (2.1) and $`P^{ij}`$ has to satisfy the condition $$P^{i[j}P^{lk]}{}_{,i}{}^{}=0,$$ (2.2) which reflects the vanishing of the corresponding Schouten-Nijenhuis bracket for $`P`$ with itself. Here the bracketed indices denote an antisymmetric sum. In the notation of Poisson brackets $$\{f(X),g(X)\}=P^{ij}(X)f_{,i}g_{,j}$$ (2.3) and the Jacobi identity follows from Eq. (2.2): $$\{f,\{g,h\}\}+\mathrm{cyclic}=0.$$ (2.4) The Poisson bracket satisfies the Leibniz derivation rule: $$\{h,fg\}=\{h,f\}g+f\{h,g\},$$ (2.5) where $`f,g,h`$ are functions on the manifold $`N`$. These facts are of an algebraic nature, and it is natural to define a Poisson algebra as an associative commutative algebra endowed with a bracket that satisfies (2.3), (2.4) and (2.5). Indeed, a smooth manifold $`N`$ is called a Poisson manifold if the algebra of smooth functions is a Poisson algebra. The splitting theorem of Weinstein states that for a regular Poisson manifold, i.e. the Poisson tensor has constant rank, there exist so-called Casimir-Darboux coordinates on the Poisson manifold $`(N,P)`$. For $`P`$ degenerate there are nonvanishing functions $`f`$ on $`N`$ whose Hamiltonian vector fields $`X_f=f_{,i}P^{ij}_j`$ vanish. These functions are called Casimir functions. Let $`\{C^I\}`$ be a maximal set of independent Casimir functions. Then $`C^I(X)=\mathrm{const}.=C^I(X_0)`$ defines a level surface through $`X_0`$ whose connected components may be identified with the symplectic leaves $`S`$ which constitute the symplectic foliation of the Poisson manifold $`(N,P)`$. According to Darboux’s theorem there are local coordinates $`X^\alpha `$ on $`S`$ such that the symplectic form $`\mathrm{\Omega }_S`$ is given by $$\mathrm{\Omega }_S=dX^1dX^2+dX^3dX^4+\mathrm{}.$$ (2.6) Together with the Casimir functions we then have a natural coordinate system $`\{X^I,X^\alpha \}`$ on $`N`$ with $`P^{IJ}=P^{I\alpha }=0`$ and $`P^{\alpha \beta }=\mathrm{constant}`$. The Poisson-Sigma model is a field theory on a closed two-dimensional world sheet $`M`$. First it involves a set of bosonic scalar fields $`X^i`$, which can be interpreted as mappings from the world sheet to a Poisson manifold, $`X^i:MN`$. In addition one needs fields $`A_i`$ which are one-forms on $`M`$ taking values in $`T^{}N`$, i.e. one-forms on the world sheet which are simultanously the pullback of sections of $`T^{}N`$ by the map $`X(x)`$, where $`x`$ denotes the coordinates of the world sheet. These fields can be seen as a non-linear extension of the gauge fields of an ordinary gauge theory. The action of the semi-topological Poisson-Sigma model is: $$𝒮_0[X,A]=\underset{M}{}\left[A_i\mathrm{d}X^i+\frac{1}{2}P^{ij}(X)A_iA_j+\mu C(X)\right],$$ (2.7) where $`\mu `$ is the volume form on $`M`$, $`C(X)`$ is a Casimir function and d denotes an exterior derivative on the world sheet $`M`$. Note that the Casmir function in the action breaks the topological nature of the theory. The action is invariant under the following symmetry transformations: $$\delta X^i=P^{ij}(X)\epsilon _j,\delta A_i=D_i^j\epsilon _j,$$ (2.8) where $`D_i^j=\delta _i^j\mathrm{d}+P^{kj}{}_{,i}{}^{}A_{k}^{}`$ is the covariant derivative on $`M`$. The equations of motion are $$D_i^jA_j+\frac{C(X)}{X^i}=0\mathrm{and}\mathrm{d}X^i+P^{ij}A_j=DX^i=0.$$ (2.9) The symmetry algebra is given by: $$[\delta (\epsilon _1),\delta (\epsilon _2)]X^i=P^{ji}(P^{mn}{}_{,j}{}^{}\epsilon _{1n}^{}\epsilon _{2m}),$$ $$[\delta (\epsilon _1),\delta (\epsilon _2)]A_i=D_i^j(P^{mn}{}_{,j}{}^{}\epsilon _{1n}^{}\epsilon _{2m})(DX^j)P^{mn}{}_{,ji}{}^{}\epsilon _{1n}^{}\epsilon _{2m}.$$ (2.10) Note that here the non-linearity of the Poisson algebra is manifested in the additonal term which is proportional to the variation of the action with respect to $`A_i`$. ### 2.2 Three Examples Now we want to show how the Poisson-Sigma model reduces to specific two-dimensional field theories. This goal can be achieved by the choice of a particular Poisson structure on the Poisson manifold, which corresponds to a choice of the target manifold. Non-degenerate Poisson structure: the first example concerns the case of the nondegenerate Poisson structure, i.e the Poisson structure P has an inverse $`\mathrm{\Omega }`$. This is exactly the symplectic 2-Form on the manifold $`N`$. It turns out that the target space is now a symplectic manifold. Note that in this case the only Casimir function is the trivial function $`C=0`$. It is possible to solve the equations of motion for the fields $`A_i`$, and one has: $$A_i=\mathrm{\Omega }_{ij}\mathrm{d}X^j.$$ (2.11) Using this equation to eliminate the fields $`A_i`$ in the action, one gets: $$𝒮_{top}=\underset{M}{}\mathrm{\Omega }_{ij}dX^i\mathrm{d}X^j.$$ (2.12) This is exactly the action of the topological Sigma model proposed by E.Witten in the Baulieu-Singer approach . Linear Poisson structure: next we consider a linear Poisson structure $`P^{ij}=c_k^{ij}X^k`$ on the three dimensional space $`^3`$. Because of the Jacobi identity the structure coefficients $`c_k^{ij}`$ define a Lie algebra structure on the dual space $`𝒢`$, of $`N`$. For this reason the linear Poisson structure is also called a Lie-Poisson structure on $`N`$. It is not hard to see that the fields $`A_i`$ reduce to the ordinary gauge fields and their covariant derivative is now the ordinary curvature of a gauge theory: $$F_i=D_i^jA_j=\mathrm{d}A_i+\frac{1}{2}c_i^{kl}A_kA_l.$$ (2.13) For a linear Poisson structure there exist two different types if Casimir functions, namely the trivial case $`C=0`$ and the quadratic Casimir $`C=_iX^iX^i`$ . For $`C=0`$ the action is given by: $$𝒮_{BF}=\underset{M}{}X^iF_i,$$ (2.14) and one sees that it is the action of a topological BF gauge theory . Choosing now the quadratic Casimir yields for the action after a short calculation: $$𝒮_{YM}=\underset{M}{}F^iF_i,$$ (2.15) where $``$ denotes the Hodge Star operator on the world sheet $`M`$. This is now the action of a 2d Yang-Mills theory. 2-dimensional Gravity: as a last example we want to see what happens if we choose a non-linear Poisson structure on $`^3`$ given by: $`P^{ij}`$ $`=ϵ^{ijk}u_k(X),`$ (2.16) $`\mathrm{with}u_a`$ $`=\eta _{ab}X^b,\mathrm{where}a,b=1,2`$ (2.17) $`\mathrm{and}u_3`$ $`=V(X^a\eta _{ab}X^b,X^3).`$ (2.18) Choosing the trivial Casimir function $`C=0`$ the action is: $$𝒮=\underset{M}{}\left(X^a𝒟_3A_a+X^3\mathrm{d}A_3+\frac{1}{2}Vϵ^{ab}A_aA_b\right).$$ (2.19) If we now identify the first two components of the field $`A_i`$ with the Zweibein $`e_a`$ and the third with the spin connection $`\omega `$ we get for the action: $$𝒮=\underset{M}{}\left(X^a𝒟_\omega e_a+X^3\mathrm{d}\omega +\frac{1}{2}Vϵ^{ab}e_ae_b\right)$$ (2.20) Hence we get two-dimensional dilaton gravity with a potential $`V`$ depending on the dilaton field $`X^3`$ . ## 3 Quantization of the Model ### 3.1 The Batalin-Vilkovisky Action We use the path integral appraoch for quantization, because we are interested in the quantum theory for different space-time topologies. To obtain a suitable action for the path integral we apply the Batalin-Vilkovisky method for the Poisson-Sigma Model. We just point out some of the essential ingredients of this approach, for a general description see for example . The first thing one has to do is to introduce ghosts $`C`$ for the symmetry transformations, and for each field $`\mathrm{\Phi }^A`$ a corresponding antifield $`\mathrm{\Phi }_A^{}`$. For the Poisson-Sigma model one has: $$\mathrm{\Phi }^A=\{A_{\mu i},X^i,C_i\}\mathrm{and}\mathrm{\Phi }_A^{}=\{A^{\mu i},X_i^{},C^i\}.$$ (3.1) For the fields and antifields there are two gradings, one is the form-degree with respect to the world sheet $`M`$ and the other is the ghost number: | $`\mathrm{gh}\mathrm{deg}`$ | 0 | 1 | 2 | | --- | --- | --- | --- | | -2 | | | $`C^i`$ | | -1 | | $`A^i`$ | $`X_i^{}`$ | | 0 | $`X^i`$ | $`A_i`$ | | | 1 | $`C_i`$ | | | On the space of the fields and antifields one defines a symplectic structure using the antibracket. For the fields $`B`$ and $`C`$ this bracket is given by: $$(B,C)=\underset{A}{}\left[\frac{B}{\mathrm{\Phi }^A}\frac{C}{\mathrm{\Phi }_A^{}}(1)^{\mathrm{deg}(\mathrm{\Phi }^A)}\frac{C}{\mathrm{\Phi }_A^{}}\frac{C}{\mathrm{\Phi }^A}\right].$$ (3.2) One also defines a Laplace operator : $$\mathrm{}C=\underset{A}{}(1)^{\mathrm{gh}(A)}\frac{^2C}{\mathrm{\Phi }^A\mathrm{\Phi }_A^{}}.$$ (3.3) The required action has to fullfill the following conditions. First, for the vanishing of the antifields the original classical action should be obtained. And second, the action must satisfy the Quantum Master Equation : $$(𝒮_{BV},𝒮_{BV})2\mathrm{}i\mathrm{}𝒮_{BV}=0.$$ (3.4) For the Poisson-Sigma model the extended action which satisfies both conditions is: $$\begin{array}{c}𝒮_{BV}=\underset{M}{}[A_i\mathrm{d}X^i+\frac{1}{2}P^{ij}(X)A_iA_j+\mu C(X)+A^iD_i^jC_j+X_i^{}P^{ji}(X)C_j\hfill \\ \hfill +\frac{1}{2}C^iP^{jk}{}_{,i}{}^{}(X)C_jC_k+\frac{1}{4}A^iA^jP^{kl}{}_{,ij}{}^{}(X)C_kC_l].\end{array}$$ (3.5) Transforming now the extended action into Casimir-Darboux coordinates one gets: $$\begin{array}{c}𝒮_{BV}=\underset{M}{}[A_I\mathrm{d}X^I+A_\alpha \mathrm{d}X^\alpha +\frac{1}{2}P^{\alpha \beta }A_\alpha A_\beta +\mu C(X^I)\hfill \\ \hfill +A^I\mathrm{d}C_I+A^\alpha \mathrm{d}C_\alpha +X_\alpha ^{}P^{\beta \alpha }C_\beta ].\end{array}$$ (3.6) Note that there are two essential simplifications. First, the terms which are quadratic in the ghosts vanish, and second the covariant derivative reduces to the normal exterior derivative. These two facts are essential for the nonpertubative calculation of the path integral. ### 3.2 Gauge fixing In the Batalin-Vilkovisky approach the gauge fixing is incorporated by the gauge fermion $`\mathrm{\Psi }`$ in the following manner. The unphysical antifields will be eliminated by the derivative of the gauge fermion with respect to the fields: $$\mathrm{\Phi }_A^{}=\frac{\mathrm{\Psi }}{\mathrm{\Phi }^A}.$$ (3.7) The ghost number of the gauge fermion has to be $`(1)`$, so that additional fields are necessary. The simplest choice is a so-called trivial pair : $`\overline{C}_i,\overline{\pi }_i`$ and the corresponding antifields. In the Casimir-Darboux coodinates the gauge fermion can be chosen to be: $$\mathrm{\Psi }=\underset{M}{}\left[\overline{C}^I\chi _I(A_I)+\overline{C}^\alpha \chi _\alpha (X^\alpha )\right].$$ (3.8) The gauge fixed action in Casimir-Darboux coordinates is given by: $$\begin{array}{c}𝒮_\mathrm{\Psi }=\underset{M}{}[A_I\mathrm{d}X^I+A_\alpha \mathrm{d}X^\alpha +\frac{1}{2}P^{\alpha \beta }A_\alpha A_\beta +\mu C(X^I)\hfill \\ \hfill +\overline{C}^J\frac{\chi _J(A_J)}{A_I}\mathrm{d}C_I+\overline{C}^\alpha \frac{\chi _\alpha (X^\alpha )}{X^\beta }P^{\gamma \beta }C_\gamma \overline{\pi }^I\chi _I(A_I)\overline{\pi }^\alpha \chi _\alpha (X^\alpha )].\end{array}$$ (3.9) Now we have arrived at an action which can be used in the path integral, since the ambiguity in the path integral which occurs because of the gauge freedom, is removed by the incorporation of the gauge fermion. ### 3.3 Calculation of the Path integral The path integral for the Poisson-Sigma model in Casimir-Darboux coordinates is $$Z=_{\mathrm{\Sigma }_\mathrm{\Psi }}𝒟X^I𝒟X^\alpha 𝒟A_I𝒟A_\alpha 𝒟C_I𝒟\overline{C}_I𝒟C_\alpha 𝒟\overline{C}_\alpha 𝒟\overline{\pi }_I𝒟\overline{\pi }_\alpha \mathrm{exp}\left(\frac{1}{\mathrm{}}𝒮_\mathrm{\Psi }\right),$$ (3.10) where we have performed the usual Wick rotation $`t=i\tau `$, so that the exponent of the path integral is now real. Integrating over the ghost and antighost fields yields the Faddeev-Popov determinants: $$\begin{array}{c}Z=_{\mathrm{\Sigma }_\mathrm{\Psi }}𝒟X^I𝒟X^\alpha 𝒟A_I𝒟A_\alpha 𝒟\overline{\pi }_I𝒟\overline{\pi }_\alpha \mathrm{det}\left(\frac{\chi _I(A_I)}{A_I}\mathrm{d}\right)_{\mathrm{\Omega }^0(M)}\mathrm{det}\left(\frac{\chi _\alpha (X^\alpha )}{X^\gamma }P^{\gamma \beta }(X^I)\right)_{\mathrm{\Omega }^0(M)}\hfill \\ \hfill \times \mathrm{exp}\left(\frac{1}{\mathrm{}}\underset{M}{}[A_I\mathrm{d}X^I+A_\alpha \mathrm{d}X^\alpha +\frac{1}{2}P^{\alpha \beta }A_\alpha A_\beta +\mu C(X^I)\overline{\pi }^I\chi _I(A)\overline{\pi }^\alpha \chi _\alpha (X^\alpha )]\right),\end{array}$$ (3.11) where the subscripts $`\mathrm{\Omega }^k(M)`$ indicate that the determinant results from an integration over k-forms on $`M`$. The integrations over $`\overline{\pi }_I`$ and $`\overline{\pi }_\alpha `$ yield $`\delta `$\- functions which implement the gauge conditions. $$\begin{array}{c}Z=_{\mathrm{\Sigma }_\mathrm{\Psi }}𝒟X^I𝒟X^\alpha 𝒟A_I𝒟A_\alpha \mathrm{det}\left(\frac{\chi _I(A_I)}{A_I}\mathrm{d}\right)_{\mathrm{\Omega }^0(M)}\mathrm{det}\left(\frac{\chi _\alpha (X^\alpha )}{X^\gamma }P^{\gamma \beta }(X^I)\right)_{\mathrm{\Omega }^0(M)}\hfill \\ \hfill \times \mathrm{exp}\left(\frac{1}{\mathrm{}}\underset{M}{}\left[A_I\mathrm{d}X^I+A_\alpha \mathrm{d}X^\alpha +\frac{1}{2}P^{\alpha \beta }A_\alpha A_\beta +\mu C(X^I)\right]\right),\end{array}$$ (3.12) where from now on the integrations extend only over the degrees of freedom which respect the gauge-fixing conditions. The integration over $`A_{\mu \alpha }`$ is gaussian, it yields $$\begin{array}{c}Z=_{\mathrm{\Sigma }_\mathrm{\Psi }}𝒟X^I𝒟X^\alpha 𝒟A_I\mathrm{det}\left(\frac{\chi _I(A^I)}{A_I}\mathrm{d}\right)_{\mathrm{\Omega }^0(M)}\mathrm{det}\left(\frac{\chi _\alpha (X^\alpha )}{X^\gamma }P^{\gamma \beta }(X^I)\right)_{\mathrm{\Omega }^0(M)}\hfill \\ \hfill \times \mathrm{det}^{1/2}\left(P^{\alpha \beta }(X^I)\right)_{\mathrm{\Omega }^1(M)}\mathrm{exp}\left(\frac{1}{\mathrm{}}\underset{M}{}\left[A_I\mathrm{d}X^I+\mathrm{\Omega }_{\alpha \beta }\mathrm{d}X^\alpha \mathrm{d}X^\beta +\mu C(X^I)\right]\right).\end{array}$$ (3.13) Besides the term in the exponent the only dependence on $`A_I`$ is in the relevant Faddeev-Popov determinant. If we choose a gauge condition linear in $`A_I`$ this determinant becomes independent of the fields, and can be absorbed into a normalization factor. The integration over $`A_I`$ then yields a $`\delta `$-function for $`\mathrm{d}X^I`$. When this $`\delta `$-function is implemented the fields $`X^I`$ become independent of the coordinates $`\{x\}`$ on $`M`$. Hence the Casimir functions are constants. The constant modes of the Casimir coordinates $`X_0^I`$ count the symplectic leaves. The path integral is now $$\begin{array}{c}Z=_{\mathrm{\Sigma }_\mathrm{\Psi }}𝑑X_0^I𝒟X^\alpha \mathrm{det}\left(\frac{\chi _\alpha (X^\alpha )}{X^\gamma }P^{\gamma \beta }(X_0^I)\right)_{\mathrm{\Omega }^0(M)}\mathrm{det}^{1/2}\left(P^{\alpha \beta }(X_0^I)\right)_{\mathrm{\Omega }^1(M)}\hfill \\ \hfill \times \mathrm{exp}\left(\frac{1}{\mathrm{}}\underset{M}{}\mathrm{\Omega }_{\alpha \beta }𝑑X^\alpha dX^\beta \right)\mathrm{exp}\left(\underset{M}{}\frac{1}{\mathrm{}}\mu C(X_0^I)\right).\end{array}$$ (3.14) The gauge-fixing of the fields $`X^\alpha `$ reduces the integral $`𝒟X^\alpha `$ to a sum over the homotopy classes of the maps: $$\begin{array}{c}Z=_{\mathrm{\Sigma }_\mathrm{\Psi }}𝑑X_0^I\underset{[MS(X_0^I)]}{}\mathrm{det}\left(\frac{\chi _\alpha (X)}{X^\gamma }P^{\gamma \beta }(X_0^I)\right)_{\mathrm{\Omega }^0(M)}\mathrm{det}^{1/2}\left(P^{\alpha \beta }(X_0^I)\right)_{\mathrm{\Omega }^1(M)}\hfill \\ \hfill \times \mathrm{exp}\left(\frac{1}{\mathrm{}}\underset{M}{}\mathrm{\Omega }_{\alpha \beta }𝑑X^\alpha dX^\beta \right)\mathrm{exp}\left(\frac{1}{\mathrm{}}\underset{M}{}\mu C(X_0^I)\right).\end{array}$$ (3.15) Since the $`C(X_0^I)`$ are independent of the coordinates on $`M`$ the last exponent simplifies to $$\mathrm{exp}\left(\frac{1}{\mathrm{}}\underset{M}{}\mu C(X_0^I)\right)=\mathrm{exp}\left(\frac{1}{\mathrm{}}A_MC(X_0^I)\right),$$ (3.16) where $`A_M`$ is the surface area of $`M`$. The form of the path integral then becomes $$\begin{array}{c}Z=_{\mathrm{\Sigma }_\mathrm{\Psi }}𝑑X_0^I\underset{[MS(X_0^I)]}{}\mathrm{det}\left(\frac{\chi _\alpha (X)}{X^\gamma }P^{\gamma \beta }(X_0^I)\right)_{\mathrm{\Omega }^0(M)}\mathrm{det}^{1/2}\left(P^{\alpha \beta }(X_0^I)\right)_{\mathrm{\Omega }^1(M)}\hfill \\ \hfill \times \mathrm{exp}\left(\frac{1}{\mathrm{}}\underset{M}{}\mathrm{\Omega }_{\alpha \beta }𝑑X^\alpha dX^\beta \right)\mathrm{exp}\left(\frac{1}{\mathrm{}}A_MC(X_0^I)\right).\end{array}$$ (3.17) Note that we have now arrived at an almost closed expression for the partition function for the Poisson-Sigma model, i.e. all the functional integrations have been performed. ### 3.4 The case of the linear Poisson structure We again consider the special case where the Poisson manifold $`N=^3`$, and the Poisson structure is linear: $`P^{ij}=c_k^{ij}X^k`$. Here we are interested in the case of the quadratic Casimir function which leads to the 2d Yang-Mills theory. The corresponding symplectic leaves are two dimensional spheres characterized, in the Casimir-Darboux coordinates, by their radius $`X_0^I`$. Weinstein has shown that the symplectic leaves of a linear Poisson structure are the co-adjoint orbits of the corresponding compact, connected Lie group G of $`𝒢`$. Because the Lie algebra has three dimensions we are restricted here to the case were the Lie group is the group SU(2). By a theorem of Kirillov these orbits can in turn be identified with the irreducible unitary representations of G . These considerations can be used to further reduce the expression for the path integral. Consider the homotopy classes of the maps $`X^\alpha :MS(X_0^I)`$. The Hopf theorem tells us that the mappings $`f,g:MS(X_0^I)`$ are homotopic if and only if the degree of the mapping $`f`$ is the same as the degree of $`g`$. This means that the sum over the homotopy classes of the maps $`[X^\alpha ]`$ can be expressed as a sum over the degrees $`n=\mathrm{deg}[X^\alpha ]`$: $$\underset{[X^\alpha ]}{}\underset{n}{}.$$ (3.18) For a map $`f:XY`$, where $`X`$ and $`Y`$ are k-dimensional oriented manifolds and $`\omega `$ a k-form on $`Y`$, the degree of the mapping is given by $$\underset{X}{}f^{}\omega =\mathrm{deg}[f]\underset{Y}{}\omega .$$ (3.19) Using this formula yields: $$\underset{M}{}\mathrm{\Omega }_{\alpha \beta }dX^\alpha \mathrm{d}X^\beta =n\underset{S}{}\mathrm{\Omega }_S(X_0^I),$$ (3.20) where $`\mathrm{\Omega }_S(X_0^I)`$ is the symplectic form on the corresponding leaf $`S`$. This gives for the partition function of Eq. (3.17) $$\begin{array}{c}Z=_{\mathrm{\Sigma }_\mathrm{\Psi }}dX_0^I\underset{n}{}\mathrm{det}\left(\frac{\chi _\alpha (X)}{X^\gamma }P^{\gamma \beta }(X_0^I)\right)_{\mathrm{\Omega }^0(M)}\mathrm{det}^{1/2}\left(P^{\alpha \beta }(X_0^I)\right)_{\mathrm{\Omega }^1(M)}\hfill \\ \hfill \times \mathrm{exp}\left(n\underset{S}{}\mathrm{\Omega }_S(X_0^I)\right)\mathrm{exp}\left(\frac{1}{\mathrm{}}A_MC(X_0^I)\right).\end{array}$$ (3.21) The sum over $`n`$ yields a periodic $`\delta `$-function: $$\begin{array}{c}Z=_{\mathrm{\Sigma }_\mathrm{\Psi }}dX_0^I\underset{n}{}\mathrm{det}\left(\frac{\chi _\alpha (X)}{X^\gamma }P^{\gamma \beta }(X_0^I)\right)_{\mathrm{\Omega }^0(M)}\mathrm{det}^{1/2}\left(P^{\alpha \beta }(X_0^I)\right)_{\mathrm{\Omega }^1(M)}\hfill \\ \hfill \times \delta \left(\underset{S}{}\mathrm{\Omega }_S(X_0^I)n\right)\mathrm{exp}\left(\frac{1}{\mathrm{}}A_MC(X_0^I)\right).\end{array}$$ (3.22) The $`\delta `$-function says that the symplectic leaves must be integral. By the identification of the leaves with the co-adjoint orbits, the orbits must also be integral. The fact that the orbits are integral reduces the number of the co-adjoint orbits to a countable set, which we label by $`𝒪(\mathrm{\Omega })`$. We now consider the two determinants in the path integral. We choose the “unitary gauge” $`\chi _\alpha (X^\alpha )=X^\alpha `$, so that $`\chi _\alpha (X)/X^\gamma =\delta _\gamma ^\alpha `$, and the two determinants have the same form. The restriction of the scalar fields to the Casimir-Darboux coordinates $`X^I`$corresponds to the restriction of the scalar fields to the invariant Cartan subalgebra considered by Blau and Thompson in , so we may adopt their argumentation concerning the powers to which the determinants occur for a manifold with Euler characteristic $`\chi (M)`$. The result is a factor $$\mathrm{det}(P^{\alpha \beta }(X_0^I))^{\chi (M)}.$$ (3.23) The determinant of a mapping equals the volume of the image of that mapping, hence the determinant $`\mathrm{det}(P^{\alpha \beta }(X_0^I))`$ corresponds to the symplectic volume of the leaf, which we denote by $`\mathrm{Vol}(\mathrm{\Omega }_S(X_0^I))`$. The path integral then takes the form: $$Z=_{\mathrm{\Sigma }_\mathrm{\Psi }}dX_0^I\underset{n}{}\mathrm{Vol}(\mathrm{\Omega }_S(X_0^I))^{\chi (M)}\delta \left(\underset{S}{}\mathrm{\Omega }_S(X_0^I)n\right)\mathrm{exp}\left(\frac{1}{\mathrm{}}A_MC(X_0^I)\right).$$ (3.24) Implementing the $`\delta `$-function by integrating over $`X_0^I`$ the sum over the mapping degrees becomes a sum over the set $`𝒪(\mathrm{\Omega })`$ of the integral orbits: $$Z=\underset{𝒪(\mathrm{\Omega })}{}\mathrm{Vol}(\mathrm{\Omega }_S(X_0^I))^{\chi (M)}\mathrm{exp}\left(\frac{1}{\mathrm{}}A_MC(X_0^I)\right).$$ (3.25) Because of the identification of the integral orbits with the irreducible unitary representations this leads to a sum over the representations. A special form of the character formula of Kirillov says that the symplectic volume of the co-adjoint orbit equals the dimension of the corresponding irreducible unitary representation. So the final form of the partition function is $$Z=\underset{\lambda }{}\mathrm{d}(\lambda )^{\chi (M)}\mathrm{exp}\left(\frac{1}{\mathrm{}}A_MC(\lambda )\right),$$ (3.26) where $`\lambda `$ denotes the irreducible unitary representation corresponding to the co-adjoint orbit, and $`\mathrm{d}(\lambda )`$ is the dimension of this representation. This is exactly the partition function for the two-dimensional Yang-Mills theory . When we omit the Casimir term in the action we get just a sum over the dimensions of the representations, which is the correct result for the BF-theory, see e.g. . ## 4 Concluding Remarks The Poisson-Sigma model is more than a unified framework for different topological and semi-topological field theories. Due to its reformulation of the degrees of freedom of the theories in terms of the coordinates of a Poisson manifold it achieves a description in terms of the natural variables of general dynamical systems. Gauge theories, which are characterized by singular Lagrangians, cannot in general be described in terms of symplectic manifolds; the foliation which is characteristic for Poisson manifolds is neccesary. Such a description of gauge theories allows one to discuss the quantization by a direct application of the techniques of deformation quantization. The connection between the Poisson-Sigma model and the deformation quantization was shown by Cattaneo and Felder by calculation of the pertubation expansion in the covariant gauge. Our nonpertubative calculation of the path integral which depends essentially on the framework of Poisson manifolds leads to an almost closed expression for the partition function. In the special case of a linear Poisson structure we were able to calculate the well known formula for the partition function of the SU(2) Yang-Mills theory with the help of fundamental facts of the representation theory of groups and algebras. We believe that further research will uncover ways of utilizing these structures more thoroughly. The techniques used here should in principle be applicable in more general situations than the particular case in Section (3.4). Finally, an understanding of the mechanisms active in the general case could help to understand the structure of gauge theories in a more fundamental way. Acknowledgement This work was supported in part (T.Schwarzweller) by the Deutsche Forschungsgesellschaft in connection with the Graduate College for Elementary Particle Physics in Dortmund.
warning/0001/math-ph0001005.html
ar5iv
text
# Quantization of Poisson algebras associated to Lie algebroids ## 1. Introduction The idea of “quantization” has evolved through a number of stages. At the beginning of this century, one meant the fact that at a microscopic scale certain physical quantities (like energy or angular momentum) assume only discrete values. Such discreteness is easily understood within the Hilbert space formalism of quantum mechanics, where self-adjoint operators may or may not have a discrete spectrum, and is no longer seen as the defining property of a quantum theory. Since about 1925, the idea has referred to the passage from a classical to a “corresponding” quantum theory. There are basically two ways to describe this passage, starting either from the Lagrangian or from the Hamiltonian version of classical mechanics. As first recognized by Feynman, the Lagrangian approach naturally leads to the path-integral formulation of quantum mechanics, and writing down a path integral may be seen as an act of quantization. All quantum-mechanical observables are constructed from so-called transition amplitudes, which correspond to the integration of certain functions with respect to the path integral. Though intuitively appealing, it is often hard to make this step rigorous. Through the nineties, many ideas of Witten originated from the use of the path integral. Most mathematically precise work on quantization is based on the Hamiltonian formulation of classical mechanics. It was initially believed that the underlying mathematics consisted of symplectic geometry, but since about 1976 it has been understood that classical mechanics should be based on the concept of a Poisson algebra \[Ki, Li\]. This is a commutative algebra with a Lie bracket that turns each element of the algebra into a derivation with respect to the commutative structure; see Definition 4.1 below. Most Poisson algebras one encounters in physics are of the form $`C^{\mathrm{}}(P)`$, where $`P`$ is a so-called Poisson manifold; the commutative algebra structure comes from pointwise multiplication, and the Lie bracket is just the Poisson bracket $`\{,\}`$. Symplectic manifolds form a special, nondegenerate case. As first recognized by Heisenberg in 1925, the quantum-mechanical observables of a given physical system should form a noncommutative algebra; the noncommutativity leads to the uncertainty relations that form the physical basis of quantum mechanics. A general, heuristic theory incorporating this idea was given in Dirac’s book \[Di\]. The correct mathematical formalism of quantum mechanics, which still stands today, is due to von Neumann \[vN\], who created the abstract theory of Hilbert spaces and self-adjoint operators for this purpose. The only (but crucial) modification to von Neumann’s formalism has been to allow other $`C^{}`$-algebras than $`B()`$ (the algebra of all bounded operators on a Hilbert space $``$) as algebras of observables. This generalization corresponds to admitting superselection rules; in the context of the quantization of finite-dimensional systems this means that one incorporates Poisson manifolds that are not symplectic in the underlying classical theory. The examples in this paper are of such a form. In the Hamiltonian formalism, quantizing a Poisson algebra $`A^0`$ of classical observables amounts to finding a “corresponding” noncommutative associative algebra $`A`$ of quantum observables, as well as a quantization map $`𝒬:A^0A`$, subject to certain conditions. Initially, practically all of quantization theory was based on a single idea of Dirac \[Di\], which he conceived in 1926 during a Sunday walk near Cambridge. Namely, in quantum mechanics the role of the Poisson bracket of the classical theory should be played by $`1/i\mathrm{}`$ times the commutator. Here $`\mathrm{}`$ is a small positive number, which in physics is a constant of nature. Hence if a classical observable $`f`$ is quantized by a quantum observable $`𝒬(f)`$, one expects that “Dirac’s condition” $$(i\mathrm{})^1[𝒬(f),𝒬(g)]=𝒬(\{f,g\})$$ holds at some fixed value of $`\mathrm{}`$. In other words, the quantization map $`𝒬`$ should be a Lie algebra homomorphism with respect to the Poisson bracket and the (rescaled) commutator. Geometric quantization \[Wo\] was an attempt to make this idea precise. Given a “prequantizable” symplectic manifold $`S`$ with associated Poisson algebra $`C^{\mathrm{}}(S)`$, this approach produces a Hilbert space $`(S)`$ (the space of $`L^2`$-sections of the prequantization line bundle $`L(S)`$ over $`S`$) and a Lie algebra homomorphism $`𝒬^{\text{pre}}`$ from $`C^{\mathrm{}}(S)`$ to a certain algebra $`𝒜(S)`$ of unbounded operators on $`(S)`$ that are densely defined on the domain of smooth sections of $`L(S)`$. The case where $`S`$ is a coadjoint orbit in the dual $`g^{}`$ of the Lie algebra $`g`$ of a Lie group $`G`$ has been studied in particular detail, for the following reason. Each element of $`Xg`$ defines a function $`\stackrel{~}{X}`$ on $`g^{}`$, and hence on $`S`$, by $`\stackrel{~}{X}(\theta )=\theta (X)`$. One has (1.1) $$\{\stackrel{~}{X},\stackrel{~}{Y}\}=\stackrel{~}{[X,Y]},$$ so that $`𝒬^{\text{pre}}`$ restricts to a Lie algebra homomorphism from $`g`$ to $`𝒜(S)`$. Assuming that this (infinitesimal) representation of $`g`$ is integrable to a unitary representation $`U`$ of $`G`$, one finds that $`U`$ tends to be reducible. In order to obtain an irreducible subrepresentation of $`U`$, one must restrict $`(S)`$. In the simplest cases one may simply project onto an irreducible subspace with orthogonal projection $`P`$ in the commutant of $`U(G)`$. For general $`fC^{\mathrm{}}(S)`$, one then has to define $`𝒬(f)=P𝒬^{\text{pre}}(f)P`$; it is clear that this modification destroys the Lie algebra homomorphism property (except on the linear span of the $`\stackrel{~}{X}`$, since $`[𝒬^{pre}(\stackrel{~}{X}),P]=0`$). A vast number of other methods has been proposed to achieve irreducibility; see \[Ki\] for a recent overview. The conclusion is that one cannot achieve irreducibility of $`U(G)`$ while preserving the Lie algebra homomorphism property on all smooth functions. In other examples than coadjoint orbits, one finds that the map $`𝒬^{\text{pre}}`$ is unsatisfactory for other (though related) reasons, the conclusion being the same: a satisfactory quantization map is only a Lie algebra homomorphism in the stated sense on some subspace of the Poisson algebra of classical observables. The way out of this dilemma emerged in the seventies, mainly as a consequence of the work of Berezin \[Be\], Vey \[Ve\], and Bayen et al. \[BFFLS\]. It is, quite simply, to require Dirac’s condition only asymptotically, that is, for $`\mathrm{}0`$. This, of course, necessitates the dramatic step of defining the quantization data $`A`$ and $`𝒬`$ for a family $`I`$ of values of $`\mathrm{}`$ that contains 0 as an accumulation point. Hence, given a Poisson algebra $`A^0`$, one now needs to find a family of algebras $`\{A_{\mathrm{}}\}_\mathrm{}I`$ and maps $`𝒬_{\mathrm{}}:A^0A_{\mathrm{}}`$, with $`𝒬_0=\mathrm{id}`$, such that $`(i\mathrm{})^1[𝒬_{\mathrm{}}(f),𝒬_{\mathrm{}}(g)]𝒬_{\mathrm{}}(\{f,g\})`$ in some sense. A method that accomplishes this is generically referred to as a deformation quantization. There are (at least) two ways to make rigorous sense of the above convergence. The oldest, introduced in the context of quantization theory in \[BFFLS\], is to define $`A_{\mathrm{}}`$ and $`𝒬_{\mathrm{}}`$ as formal power series in $`\mathrm{}`$. This method, called formal deformation quantization, remains the most popular; see \[Ko\] for a recent highlight. The second approach, which is based on choosing the $`A_{\mathrm{}}`$ to be $`C^{}`$-algebras, was introduced by Rieffel \[Ri2\]-\[Ri7\]. Here $`\mathrm{}`$ is simply a real number rather than a formal deformation parameter, and one imposes Dirac’s condition asymptotically in norm; see Definition 4.3 below. It is then mathematically natural and physically meaningful to require that the $`A_{\mathrm{}}`$ form a continuous field of $`C^{}`$-algebras over the index set $`I\mathrm{}`$. In section 4 we give a precise formulation of this approach to deformation quantization, introducing two appropriate generalizations of Rieffel’s original definition that lie at the basis of the results in the present paper. Having discussed the development of modern quantization theory in broad outline, we now turn to the class of classical systems (that is, Poisson manifolds) that are quantized in this paper. Our motivation comes from a number of directions. The first is the quantization theory of coadjoint orbits discussed above. Although the attempt to quantize individual orbits is only successful in special cases, one could try to quantize all coadjoint orbits of a given Lie group at one stroke by quantizing $`g^{}`$ (or $`C^{\mathrm{}}(g^{})`$) as a whole. To do so, one starts from the fact that $`g^{}`$ is canonically a Poisson manifold when equipped with the so-called Lie–Poisson structure that may be defined by (1.1). As first shown by Rieffel \[Ri4\] in the special case that $`G`$ is exponential and nilpotent, one may construct a quantization of $`C^{\mathrm{}}(g^{})`$ in which $`A_{\mathrm{}}=C^{}(G)`$ for all $`\mathrm{}0`$, and the maps $`𝒬_{\mathrm{}}`$ are essentially given by the pullback of the exponential map from $`g`$ to $`G`$, rescaled by $`1/\mathrm{}`$. We show that this works without any assumption on the Lie group $`G`$. The idea that coadjoint orbits in $`g^{}`$ are quantized by unitary irreducible representations of $`G`$ then re-enters through the back door, as follows. The well-known correspondence between nondegenerate representations of $`C^{}(G)`$ and unitary representations of $`G`$, preserving irreducibility, is matched by the fact that the irreducible representations of $`C^{\mathrm{}}(g^{})`$ as a Poisson algebra \[La3\] are precisely given by the coadjoint orbits of $`G`$ (or covering spaces thereof). In general, there is no correspondence through quantization between the irreducible representations of $`C^{\mathrm{}}(g^{})`$ and the irreducible representations of $`C^{}(G)`$; the correspondence through (deformation) quantization is between the algebras in question themselves. Secondly, one may quantize a particle moving on a Riemannian manifold $`M`$ (so that the pertinent Poisson manifold is the cotangent bundle $`T^{}M`$, which is symplectic) by taking $`A_{\mathrm{}}`$ to be the $`C^{}`$-algebra $`B_0(L^2(M))`$ of compact operators on $`L^2(M)`$ for all $`\mathrm{}0`$, and defining a quantization map $`𝒬_{\mathrm{}}`$ in terms of the geodesics on $`M`$. This result may be generalized by coupling the particle to an external Yang–Mills field with gauge group $`H`$; one then starts from a principal $`H`$-bundle $`𝖯`$, so that $`M=𝖯/H`$. The Poisson manifold describing the classical theory is then given by $`P=(T^{}𝖯)/H`$, which is not symplectic unless $`H`$ is discrete, and when $`H`$ is compact the quantum algebra of observables turns out to be $`B_0(L^2(𝖯))^H`$; see \[La1, La3\] for details and the generalization to noncompact $`H`$. Taking $`𝖯=H=G`$ actually reproduces the previous example, since $`(T^{}G)/Gg^{}`$ as Poisson manifolds, whereas for compact Lie groups one has $`B_0(L^2(G))^GC^{}(G)`$ (and analogously for the noncompact case). The quantization maps in question are also compatible with this specialization. Thirdly, in the early eighties Lie groupoids and locally compact groupoids started to play a role in $`C^{}`$-algebras as a result of the work of Connes and of Renault, whereas in the late eighties Lie groupoids entered symplectic and Poisson geometry, as well as quantization theory, through the idea of a symplectic groupoid (cf. \[We2\]). Against this background, it was an obvious idea that the above quantizations ought to be interpretable in terms of Lie groupoids. The key to this possibility lies in the fact the passage from a Lie groupoid to its convolution $`C^{}`$-algebra has a classical analogue. Namely, like a Lie group, a Lie groupoid $`G`$ has an associated “infinitesimal” object, its Lie algebroid \[Pr1\] $`A(G)`$, which is a vector bundle over the base space $`G_0`$ of $`G`$. The dual bundle $`A^{}(G)`$ of $`A(G)`$ admits a canonical Poisson structure \[CDW, Cou\], so that eventually one may associate a Poisson algebra $`C^{\mathrm{}}(A^{}(G))`$ to $`G`$. All threads now come together \[La1\], in that in each of the three classes of examples above the classical Poisson algebra is of the form $`C^{\mathrm{}}(A^{}(G))`$ and the quantum $`C^{}`$-algebra is of the form $`C^{}(G)`$, where $`G`$ is some Lie groupoid. This is obvious in the first example. In the second, one takes $`G`$ to be the pair groupoid $`G=M\times M`$ on $`M`$, whose Lie algebroid is the tangent bundle $`TM`$; the Poisson structure on the dual bundle $`T^{}M`$, which is nothing but the cotangent bundle of $`M`$, is just its usual canonical structure (which is symplectic). See also \[CCFGRV\]. Finally, in the third example $`G`$ is the so-called gauge groupoid of the principal bundle $`𝖯`$ \[Ma\], with associated Lie algebroid $`(T𝖯)/H`$. Based on these examples, it was conjectured in \[La2\] that one should obtain a deformation quantization of this type for any Lie groupoid $`G`$. For formal deformation quantization a version of this conjecture was proved in \[NWX\], and a proof in $`C^{}`$-algebraic deformation quantization above appeared in \[La3, La4, Ra1\]. In this paper we give a precise and complete formulation of all mathematical concepts surrounding this class of $`C^{}`$-algebraic deformation quantizations, and prove two technically distinct versions of the above conjecture, each of which has its merits. The generalization of Connes’s tangent groupoid \[Co2\] $`\widehat{G}`$ to arbitrary Lie groupoids $`G`$ \[HS, We1\] plays an important role. A significant intermediate result is that $`C^{}(\widehat{G})`$ is the $`C^{}`$-algebra of a continuous field of $`C^{}`$-algebras over $``$ with fibers $`A_0=C^{}(A(G))C_0(A^{}(G))`$ and $`A_{\mathrm{}}=C^{}(G)`$ for $`\mathrm{}0`$. The same is true for the corresponding reduced $`C^{}`$-algebras. See section 6. The sections of the paper are listed in the table of contents following the abstract. Our main results, Theorems 4.4 and 4.6, are stated at the end of section 4. The remainder of the paper develops the proofs of these theorems. ## 2. Lie groupoids and Lie algebroids This section is a brief review of our objects of study, mainly intended to establish our notation. We assume that the reader is familiar with the basic theory of locally compact groupoids (cf. \[Re1\]), which we always assume to be second countable. We denote a groupoid as a whole by $`G`$, the base is called $`G_0`$, which is seen as a subspace of the arrow space (or total space) $`G_1`$ under the inclusion map $`\iota `$. The source and range projections are called $`s`$ and $`r`$, respectively, and for $`xG_0`$ we put (2.1) $$G^x=r^1(x)G_1.$$ The space of composable arrows is $`G_2=\{(\gamma ,\gamma ^{})G_1\times G_1r(\gamma ^{})=s(\gamma )\}`$. ###### Definition 2.1. A Lie groupoid is a locally compact groupoid $`G`$ for which $`G_1`$ and $`G_0`$ are manifolds, $`s`$ and $`r`$ are surjective submersions, and multiplication and inclusion are smooth maps. It follows that $`\iota `$ is an immersion, that the inverse is a diffeomorphism, that $`G_2`$ is a closed submanifold of $`G_1\times G_1`$, that the fibers $`r^1(x)`$ and $`s^1(x)`$ are submanifolds of $`G_1`$ for all $`xG_0`$, and that all isotropy groups are Lie groups. The basic reference on Lie groupoids is Mackenzie’s book \[Ma\]; also see \[Co2, La3\]. A Lie group defines its Lie algebra; one may also study Lie algebras in their own right. The situation for Lie groupoids is similar. ###### Definition 2.2. A Lie algebroid on a manifold $`M`$ is a vector bundle $`E`$ over $`M`$, which apart from the bundle projection $`\tau :EM`$ is equipped with a vector bundle map $`\rho :ETM`$ (called the anchor), as well as with a Lie bracket $`[,]_E`$ on the space $`C^{\mathrm{}}(M,E)`$ of smooth sections of $`E`$, satisfying (2.2) $$\rho [X,Y]_E=[\rho X,\rho Y],$$ where the right-hand side is the usual commutator of vector fields on $`C^{\mathrm{}}(M,TM)`$, and (2.3) $$[X,fY]_E=f[X,Y]_E+((\rho X)f)Y$$ for all $`X,YC^{\mathrm{}}(M,E)`$ and $`fC^{\mathrm{}}(M)`$. We generally omit the suffix $`E`$ on the Lie bracket. It is part of the definition of a bundle map that the anchor is fiber-preserving and linear on each fiber. This concept is due to Pradines \[Pr2\]. The basic reference on Lie algebroids is \[Ma\]; also see \[La3\]. The simplest examples of Lie algebroids are tangent bundles $`TM`$, where $`\rho `$ is the identity, and Lie algebras, for which $`M`$ is a point. We now explain how one may associate a Lie algebroid $`A(G)`$ with a given Lie groupoid $`G`$ \[Pr1, Ma, La3\]. A left-invariant vector field $`\xi ^L`$ on $`G_1`$ is a vector field satisfying $`Tr(\xi ^L)=0`$ and $`TL_\gamma (\xi ^L)(\gamma ^{})=\xi ^L(\gamma \gamma ^{})`$ for all $`(\gamma ,\gamma ^{})G_2`$. Here $`L_\gamma :G^{s(\gamma )}G^{r(\gamma )}`$ is defined for each $`\gamma G_1`$ by (2.4) $$L_\gamma (\gamma _1)=\gamma \gamma _1.$$ Note that the second condition is well-defined because of the first one. The space of all smooth left-invariant vector fields on $`G_1`$ is denoted by $`C^{\mathrm{}}(G_1,TG_1)^L`$, which is easily shown to be a Lie algebra under the usual commutator borrowed from $`C^{\mathrm{}}(G_1,TG_1)`$. Also, a left-invariant vector field is obviously determined by its values on the unit space $`G_0`$. The tangent bundle of $`G_1`$ at the unit space has a decomposition (2.5) $$T_xG_1=T_xG_0(\mathrm{ker}Tr)_x,$$ where $`\mathrm{ker}TrTG_1`$ is the vector bundle over $`G_1`$ whose fiber $`(\mathrm{ker}Tr)_\gamma `$ above $`\gamma G_1`$ is the kernel of the derivative $`T_\gamma r:T_\gamma G_1T_{r(\gamma )}G_0`$ of the range projection $`r:G_1G_0`$. The special case $`\gamma =xG_0`$ occurs in (2.5). ###### Definition 2.3. The Lie algebroid $`A(G)`$ of a Lie groupoid $`G`$ is given by the following: 1. The vector bundle $`A(G)`$ over $`G_0`$ is the normal bundle defined by the embedding $`G_0G_1`$; accordingly, the bundle projection $`\tau :A(G)G_0`$ is given by $`s`$ or $`r`$ (which coincide on $`G_0`$). 2. Identifying the normal bundle with $`(\mathrm{ker}Tr)|_{G_0}`$ by (2.5), the anchor is given by $`\rho =Ts:\mathrm{ker}TrTG_0`$. 3. Identifying a section of the normal bundle with an element of $`C^{\mathrm{}}(G_1,TG_1)^L`$ through the previous item, the Lie bracket $`[,]_{A(G)}`$ is given by the commutator. The required equality (2.2) is automatically satisfied (as it holds for all vector fields on $`G_1`$). We leave the verification of (2.3) to the reader. It follows from this definition that a Lie algebra $`g`$ is the Lie algebroid of a Lie group $`G`$, and that the tangent bundle $`TG_0`$ is the Lie algebroid of the pair groupoid $`G_1=G_0\times G_0`$. For Lie groups one has an exponential map from the Lie algebra to the group. For manifolds $`M`$ with connection the exponential map is defined on the tangent bundle $`TM`$, which it maps into $`M`$. As indicated by Pradines \[Pr3\], these are special cases of a construction that holds for general Lie groupoids $`G`$, provided its Lie algebroid $`A(G)`$ is endowed with a connection. Following \[La3, La4\], we here present a slightly different construction that is more suitable for our application to deformation quantization; also see \[NWX\]. First note that the vector bundles $`\mathrm{ker}Tr`$ and $`s^{}A(G)`$ (over $`G_1`$) are isomorphic via the map $`TL_\gamma `$, applied fiberwise. Hence the connection on $`A(G)`$, with associated horizontal lift $`\mathrm{}^{A(G)}`$, yields a connection on $`\mathrm{ker}Tr`$ through pull-back. Going through the definitions, one finds that the associated horizontal lift $`\mathrm{}`$ of a tangent vector $`X=\dot{\gamma }=d\gamma (t)/dt_{t=0}`$ in $`T_\gamma G_1`$ to $`YT_\gamma ^rG_1`$ is (2.6) $$\mathrm{}_Y(\dot{\gamma })=\frac{d}{dt}[L_{\gamma (t)}\mathrm{}_{TL_{\gamma ^1}(Y)}^{A(G)}(s(\gamma (t)))]_{t=0},$$ which is an element of $`T_Y(\mathrm{ker}Tr)`$ (here $`\mathrm{}^{A(G)}(\mathrm{})`$ lifts a curve). Since the bundle $`\mathrm{ker}TrG_1`$ has a connection, one can define parallel transport $`XX(t)`$ on $`\mathrm{ker}Tr`$ in precisely the same way as on a tangent bundle with affine connection. That is, the flow $`X(t)`$ is the solution of (2.7) $$\dot{X}(t)=\mathrm{}_{X(t)}(X(t))$$ with initial condition $`X(0)=X`$. The projection of $`X(t)`$ to $`G_1`$ is a “geodesic” $`\gamma _X(t)`$. ###### Definition 2.4. The left exponential map $`\mathrm{Exp}^L:A(G)G_1`$ is defined by (2.8) $$\mathrm{Exp}^L(X)=\gamma _X(1)=\stackrel{~}{\tau }(X(1)),$$ where $`\stackrel{~}{\tau }:\mathrm{ker}TrG`$ is the restriction of the bundle projection $`TG_1G_1`$ to $`\mathrm{ker}Tr`$. It is assumed that the geodesic flow $`X(t)`$ on $`\mathrm{ker}Tr`$ (defined by the connection on $`\mathrm{ker}Tr`$ pulled back from the one on $`A(G)`$) is defined at $`t=1`$. If not, $`\mathrm{Exp}^L(X)`$ is undefined. There is a symmetrized version of $`\mathrm{Exp}^L`$ that plays a key role in our quantization theory. First note that for all $`XA(G)`$ for which $`\mathrm{Exp}^L(X)`$ is defined one has (2.9) $$r(\mathrm{Exp}^L(X))=\tau (X);$$ recall that $`\tau `$ is the bundle projection of the Lie algebroid. To derive this, note that $`r(\gamma _X(0))=\tau (X)`$ by construction, and $`dr(\gamma _X(t))/dt=Tr(X(t))=0`$. We combine this with the obvious $`\tau (\frac{1}{2}X)=\tau (\frac{1}{2}X)`$ to infer that $$r(\mathrm{Exp}^L(\frac{1}{2}X))=r(\mathrm{Exp}^L(\frac{1}{2}X))=s(\mathrm{Exp}^L(\frac{1}{2}X)^1).$$ Thus the (groupoid) multiplication in (2.10) below is well-defined. ###### Definition 2.5. The Weyl exponential map $`\mathrm{Exp}^W:A(G)G_1`$ is defined by (2.10) $$\mathrm{Exp}^W(X)=\mathrm{Exp}^L(\frac{1}{2}X)^1\mathrm{Exp}^L(\frac{1}{2}X).$$ The following well-known result is a form of the tubular neighbourhood theorem. ###### Proposition 2.6. The maps $`\mathrm{Exp}^L`$ and $`\mathrm{Exp}^W`$ are diffeomorphisms from a neighbourhood $`V`$ of $`G_0A(G)`$ (as the zero section) to a neighbourhood $`W`$ of $`G_0`$ in $`G_1`$, such that $`\mathrm{Exp}^L(x)=\mathrm{Exp}^W(x)=x`$ for all $`xG_0`$. Cf. \[La3, La4\] for a proof. ## 3. The $`C^{}`$-algebra of a Lie groupoid In this section we show that to every Lie groupoid $`G`$ one can associate a $`C^{}`$-algebra $`C^{}(G)`$. This $`C^{}`$-algebra was introduced by Bigonnet \[Bg\], but the idea to use half densities for an intrinsic construction of the $`C^{}`$-algebra of a Lie groupoid first appeared in Connes’s work \[Co1\] for the special case of the holonomy groupoid of a foliation. Connes also proposed a general definition in \[Co2\]; this only works if fiber isomorphisms of the type discussed below are understood. The construction we give in this paper is due to Renault \[Re4\]; it is essentially the same as that in \[La2, La3\]. Also see \[CW\]. Further details on densities as used below may be found in \[Tr, La3\]. Let $`V`$ be an $`n`$-dimensional vector space and let $`\alpha `$ be a fixed real number. A density of weight $`\alpha `$ on $`V`$ is a function $`\rho :^n(V)`$ that satisfies $$\rho (\lambda T)=\lambda ^\alpha \rho (T)$$ for every $`\lambda `$ and $`T^n(V)`$. The $`1`$-dimensional vector space of the densities of weight $`\alpha `$ on $`V`$ is denoted by $`|\mathrm{\Omega }|^\alpha (V)`$. If a density is positive at a point, then it is positive everywhere on $`^n(V)\backslash \{0\}`$. Hence one may unambiguously speak of positive densities. We have the canonical isomorphisms $`|\mathrm{\Omega }|^{\alpha +\beta }(V)|\mathrm{\Omega }|^\alpha (V)|\mathrm{\Omega }|^\beta (V)`$ and $`|\mathrm{\Omega }|^\alpha (VW)|\mathrm{\Omega }|^\alpha (V)|\mathrm{\Omega }|^\alpha (W)`$, where $`W`$ is another vector space. Denote the parallelepiped generated by a family $`v_1,\mathrm{},v_n`$ of vectors in $`V`$ by $`[v_1,\mathrm{},v_n]`$. Given a positive $`1`$-density $`\rho `$ on $`V`$, the translation invariant measure $`\mu `$ on $`V`$ that satisfies $`\mu ([v_1,\mathrm{},v_n])=\rho (v_1\mathrm{}v_n)`$ for every family $`\{v_1,\mathrm{},v_n\}`$ of vectors is said to be the measure generated by $`\rho `$. If $`\{e_1,\mathrm{},e_n\}`$ is a basis of $`V`$ and $`fL^1(V,\mu )`$, then (3.1) $$f\left(\stackrel{n}{\underset{i=1}{}}x_ie_i\right)\rho (e_1\mathrm{}e_n)dx=fd\mu .$$ We now generalize this construction to fibers in vector bundles. Let $`E`$ be a vector bundle on $`M`$, with fibers $`E_x`$ over $`xM`$. For each $`xM`$, consider $`|\mathrm{\Omega }|^\alpha (E_x)`$, the $`1`$-dimensional vector space of densities of weight $`\alpha `$ on the vector space $`E_x`$. The vector bundle over $`M`$ with fibers $`|\mathrm{\Omega }|^\alpha (E_x)`$ is denoted by $`|\mathrm{\Omega }|^\alpha (E)`$. The sections of the vector bundle $`|\mathrm{\Omega }|^\alpha (E)`$ over $`M`$ are called $`\alpha `$-densities. An $`\alpha `$-density $`\mu C^{\mathrm{}}(M,|\mathrm{\Omega }|^\alpha (E))`$ is called positive if $`\mu (x)`$ is a positive density for every $`xM`$. The function associated to $`\mu `$ in the local frame $`e=(e_1,\mathrm{},e_p)`$ of $`E`$ on $`UM`$ is the map $`\mu _eC^{\mathrm{}}(U)`$ defined by $`\mu _e(x)=\mu (x)\left(e_1(x)\mathrm{}e_p(x)\right)`$. The set of $`\alpha `$-densities on $`E`$ is a free $`1`$-dimensional module over the ring of functions defined on $`M`$. Moreover, $`C^{\mathrm{}}(M,|\mathrm{\Omega }|^\alpha (E))`$ is a free $`1`$-dimensional module over $`C^{\mathrm{}}(M)`$, and choosing a positive smooth density we can identify densities with functions. Now let $`G`$ be a Lie groupoid with Lie algebroid $`A(G)`$; recall from Definition 2.3 that $`A(G)`$ is a vector bundle over $`G_0`$. For every real number $`\alpha `$, one may associate the vector bundle $`|\mathrm{\Omega }|^\alpha (A(G))`$ to $`A(G)`$. In addition, form the vector bundle $`|\mathrm{\Omega }|^\alpha (\mathrm{ker}Tr)`$ of $`\alpha `$-densities associated to the vector bundle $`\mathrm{ker}TrTG_1`$. (defined after (2.5)). The groupoid $`G`$ acts (from the left) on the bundle $`|\mathrm{\Omega }|^\alpha (\text{ Ker}Tr)`$. To define this action we need a further bit of notation. Let $`V,W`$ be vector spaces with dim $`V`$=dim $`W`$. For every linear map $`f:VW`$, define $`f^\mathrm{\Omega }:|\mathrm{\Omega }|^\alpha (F)|\mathrm{\Omega }|^\alpha (E)`$ on $`\rho \left|\mathrm{\Omega }\right|^\alpha (F)`$, $`v_1,v_2,\mathrm{},v_nV`$ by $`(f^\mathrm{\Omega }\rho )(v_1\mathrm{}v_n)=\rho (f(v_1)\mathrm{}f(v_n))`$. Obviously, if $`\rho `$ is a positive density then $`f^\mathrm{\Omega }\rho `$ is positive, too. To the derivative $`T_{\gamma \gamma _1}L_{\gamma ^1}:T_{\gamma \gamma _1}G^{r(\gamma )}T_{\gamma _1}G^{s(\gamma )}`$ of $`L_\gamma `$ (cf. (2.4)) one associates the corresponding map between the spaces of densities $`(T_{\gamma \gamma _1}L_{\gamma ^1})^\mathrm{\Omega }:\left|\mathrm{\Omega }\right|^\alpha (T_{\gamma _1}G^{s(\gamma )})|\mathrm{\Omega }|^\alpha (T_{\gamma \gamma _1}G^{r(\gamma )})`$. For simplicity, we denote $`(T_{\gamma \gamma _1}L_{\gamma ^1})^\mathrm{\Omega }`$ by $`\gamma `$, and have an action of $`G`$ on $`|\mathrm{\Omega }|^\alpha (\mathrm{ker}Tr)`$ given by $`\gamma :|\mathrm{\Omega }|^\alpha (T_{\gamma _1}G^{s(\gamma )})|\mathrm{\Omega }|^\alpha (T_{\gamma \gamma _1}G^{r(\gamma )})`$, where $`\gamma _1G^{s(\gamma )}`$. Denote the vector bundle $`r^{}|\mathrm{\Omega }|^{1/2}(A(G))s^{}|\mathrm{\Omega }|^{1/2}(A(G))`$ over $`G`$ by $`G`$. The fiber of $`G`$ over $`\gamma G`$ is $`G_\gamma =|\mathrm{\Omega }|^{1/2}(T_{r(\gamma )}G^{r(\gamma )})|\mathrm{\Omega }|^{1/2}(T_{s(\gamma )}G^{s(\gamma )})`$. ###### Definition 3.1. The convolution algebra of a Lie groupoid $`G`$ is the space $`C_c^{\mathrm{}}(G_1,G)`$ of smooth compactly supported sections of the vector bundle $`G`$ over $`G_0`$, with product $`fg`$ of $`f,gC_c^{\mathrm{}}(G_1,G)`$ defined by (3.2) $$(fg)(\gamma )=\underset{G^{r(\gamma )}}{}(\text{id}\gamma _1)f(\gamma _1)(\gamma _1\text{id})g(\gamma _1^1\gamma ),$$ where id is the identity map, and involution given by (3.3) $$f^{}(\gamma )=\stackrel{~}{f(\gamma ^1)}.$$ Here the map $$:|\mathrm{\Omega }|^{1/2}(T_xG^x)|\mathrm{\Omega }|^{1/2}(T_yG^y)|\mathrm{\Omega }|^{1/2}(T_yG^y)|\mathrm{\Omega }|^{1/2}(T_xG^x)$$ is defined by $`a(x)b(y)\overline{b(y)a(x)}.`$ It is not difficult to show that these expressions are well defined. Let us show, for example, that the integral in (3.2) is well defined. We extend the action of $`G`$ on the bundle $`|\mathrm{\Omega }|^{1/2}(\mathrm{ker}Tr)`$ to the tensor product in an obvious way. For $`f,gC_c^{\mathrm{}}(G_1,G)`$ we see that $`(\text{id}\gamma _1)f(\gamma _1)`$ is an element of $`|\mathrm{\Omega }|^{1/2}(T_{r(\gamma _1)}G^{r(\gamma _1)})|\mathrm{\Omega }|^{1/2}(T_{\gamma _1}G^{r(\gamma _1)})`$, and that $`(\gamma _1\text{id})g(\gamma _1^1\gamma )`$ lies in $`|\mathrm{\Omega }|^{1/2}(T_{\gamma _1}G^{r(\gamma _1)})|\mathrm{\Omega }|^{1/2}(T_{s(\gamma )}G^{s(\gamma )})`$. The integrand in (3.2) is then an element of $`|\mathrm{\Omega }|^{1/2}(T_{r(\gamma )}G^{r(\gamma )})|\mathrm{\Omega }|^1(T_{\gamma _1}G^{r(\gamma _1)})|\mathrm{\Omega }|^{1/2}(T_{s(\gamma )}G^{s(\gamma )})`$, so that it may be integrated as a 1-density on the manifold $`G^{r(\gamma )}`$. Hence we obtain an element of $`|\mathrm{\Omega }|^{1/2}(T_{r(\gamma )}G^{r(\gamma )})|\mathrm{\Omega }|^{1/2}(T_{s(\gamma )}G^{s(\gamma )})`$. ###### Proposition 3.2. The map $`\lambda \lambda |_{G_0}`$ is a bijection between the left invariant sections of $`|\mathrm{\Omega }|^\alpha (\mathrm{ker}Tr)`$ and the sections of $`|\mathrm{\Omega }|^\alpha (A(G))`$. This bijection preserves smoothness as well as positivity. ###### Proof. The restriction is well defined, since $`|\mathrm{\Omega }|^\alpha (\mathrm{ker}Tr)`$ contains $`|\mathrm{\Omega }|^\alpha (A(G))`$. To prove that the map in question is bijective, it is sufficient to check that its inverse is $`\rho \lambda `$, where $`\rho |\mathrm{\Omega }|^\alpha (A(G))`$ and $`\lambda (\gamma )=\gamma \rho (s(\gamma ))`$. $`\mathrm{}`$ We now define the groupoid $`C^{}`$-algebra $`C^{}(G)`$ and its reduced counterpart $`C_r^{}(G)`$. Firstly, to endow the convolution $`^{}`$-algebra $`C_c^{\mathrm{}}(G_1,G)`$ with a $`C^{}`$-norm we consider its $`^{}`$-representations on a Hilbert space. Let $``$ be the set of all $`^{}`$-representations $`\pi (C_c^{\mathrm{}}(G_1,G))`$ for which $`f\xi ,\pi (f)\eta `$ is a Radon measure on $`C_c^{\mathrm{}}(G_1,G)`$ for each pair $`\xi ,\eta H`$. One then defines a $`C^{}`$-norm on $`C_c^{\mathrm{}}(G_1,G)`$ by (3.4) $$f=\underset{\pi }{\text{sup}}\pi (f);$$ cf. \[Re1\]. The completion of $`C_c^{\mathrm{}}(G_1,G)`$ in this norm is the $`C^{}`$-algebra $`C^{}(G)`$. Secondly, for every $`xG_0`$ we define an involutive representation $`\pi _x`$ of $`C_c^{\mathrm{}}(G_1,G)`$ on the Hilbert space $`L^2(G^x)`$ of half-densities on the manifold $`G^x`$ by (3.5) $$[\pi _x(f)\xi ](\gamma )=\underset{G^x}{}(\text{id}\gamma _1)f(\gamma _1)(\gamma _1\text{id})\xi (\gamma _1^1\gamma ),$$ for $`\gamma G^x`$, where $`\xi L^2(G^x)`$. Then $`f_r=\underset{xG_0}{\text{sup}}\pi _x(f)`$ is a norm (cf. \[Re1\]), and the completion of $`C_c^{\mathrm{}}(G_1,G)`$ in this norm is the reduced $`C^{}`$-algebra $`C_r^{}(G)`$. For an amenable groupoid $`G`$ we have $`C_r^{}(G)=C^{}(G)`$. For the proof of this result and a detailed discussion of amenable groupoids see \[AR\]. Recall the following concept (Def. I.2.2 in \[Re1\]). ###### Definition 3.3. A left Haar system on a locally compact groupoid $`G`$ is a family of measures $`(\lambda ^x)_{xG_0}`$ such that the support of $`\lambda ^x`$ is $`G^x`$, the system is left-invariant under the map $`L_\gamma :G^{s(\gamma )}G^{r(\gamma )}`$, and for each $`\varphi C_c(G_1)`$, the map (3.6) $$x\lambda (\varphi )(x)=\varphi 𝑑\lambda ^x$$ defines a continuous function on $`G_0`$. If $`G`$ is a Lie groupoid, we say that the Haar system is smooth if each such function is smooth for $`\varphi C_c^{\mathrm{}}(G_1)`$. As explained in \[Re1\], one can associate $`C^{}`$-algebras $`C^{}(G,\lambda )`$ and $`C_r^{}(G,\lambda )`$ to a locally compact groupoid $`G`$ with Haar system $`(\lambda ^x)`$. Recall that convolution and involution are given on the dense subalgebra $`C_c(G_1)`$ by (3.7) $`fg(\gamma )`$ $`=`$ $`{\displaystyle _{G^{s(\gamma )}}}f(\gamma \gamma _1)g(\gamma _1^1)𝑑\lambda ^{s(\gamma )}(\gamma _1);`$ (3.8) $`f^{}(\gamma )`$ $`=`$ $`\overline{f(\gamma ^1)}.`$ This turns $`C_c(G_1)`$ into a $`^{}`$-algebra, which we denote by $`C_c(G,\lambda )`$. Similarly, if $`G`$ is smooth one has the $`^{}`$-algebra $`C_c^{\mathrm{}}(G,\lambda )`$. The $`C^{}`$-algebras $`C^{}(G,\lambda )`$ and $`C_r^{}(G,\lambda )`$ are completions of $`C_c(G,\lambda )`$ or $`C_c^{\mathrm{}}(G,\lambda )`$ in suitable $`C^{}`$-norms. Not every locally compact groupoid admits a left Haar system; in the theory of groupoid $`C^{}`$-algebras one therefore usually postulates its existence. For Lie groupoids the situation is more favourable. ###### Proposition 3.4. Any Lie groupoid $`G`$ admits a smooth left Haar system $`(\lambda ^x)_{xG_0}`$. The associated convolution $`^{}`$-algebra $`C_c^{\mathrm{}}(G,\lambda )`$ is isomorphic to $`C_c^{\mathrm{}}(G_1,G)`$, and the associated $`C^{}`$-algebras $`C^{}(G,\lambda )`$ and $`C_r^{}(G,\lambda )`$ are isomorphic to $`C^{}(G)`$ and $`C_r^{}(G)`$, respectively. ###### Proof. Let $`\lambda `$ be a smooth positive section of $`|\mathrm{\Omega }|^1(A(G))`$. By Proposition 3.2, $`\lambda `$ can be extended to a smooth, positive, and left invariant section $`\lambda _G`$ of $`|\mathrm{\Omega }|^1(\mathrm{ker}Tr)`$. Then $`\lambda _G^{1/2}`$ is a smooth left invariant section of $`|\mathrm{\Omega }|^{1/2}(\mathrm{ker}Tr)`$, and $`\gamma \lambda _G^{1/2}(r(\gamma ))\lambda _G^{1/2}(s(\gamma ))`$ is a smooth section of $`G`$. For every section $`fC_c^{\mathrm{}}(G_1,G)`$ there exists a function $`f_\lambda C_c^{\mathrm{}}(G_1)`$ such that (3.9) $$f(\gamma )=f_\lambda (\gamma )\lambda _G^{1/2}(r(\gamma ))\lambda _G^{1/2}(s(\gamma )).$$ An easy calculation shows that for $`f,gC_c^{\mathrm{}}(G_1,G)`$ we have $$(fg)(\gamma )=(f_\lambda g_\lambda )(\gamma )\lambda _G^{1/2}(r(\gamma ))\lambda _G^{1/2}(s(\gamma )),$$ where (3.10) $$(f_\lambda g_\lambda )(\gamma )=\underset{G^{r(\gamma )}}{}f_\lambda (\gamma _1)g_\lambda (\gamma _1^1\gamma )\lambda _G(\gamma _1).$$ The right-hand side of the last equality is well defined as the integral of the $`1`$-density $`f_\lambda (\gamma _1)g_\lambda (\gamma _1^1\gamma )\lambda _G(\gamma _1)`$ over the manifold $`G^{r(\gamma )}`$. It is also easy to see that (3.11) $$f^{}(\gamma )=f_\lambda ^{}(\gamma )\lambda _G^{1/2}(r(\gamma ))\lambda _G^{1/2}(s(\gamma )),$$ where $`f_\lambda ^{}(\gamma )=\overline{f_\lambda (\gamma ^1)}`$. Let $`xG_0`$. The restriction of $`\lambda _G`$ to the submanifold $`G^x`$ is a $`1`$-density, hence it has an associated measure on $`G^x`$. As we identify the $`1`$-densities and their associated measures, we use the same notation $`\lambda ^x`$ for the restriction of $`\lambda _G`$ to $`G^x`$ and the induced measure. The equality $`\gamma \lambda ^{s(\gamma )}=\lambda ^{r(\gamma )}`$ proves that the $`\lambda ^x`$ form a left Haar system; its is easy to check that each function (3.6) is smooth. It is easily checked that, under the correspondence $`ff_\lambda `$ defined in (3.9) and the replacement of $`\lambda _G`$ by the left Haar system $`(\lambda ^x)_{xG_0}`$, the expressions (3.10) and (3.11) are transformed into (3.7) and (3.8), respectively. We leave the proof of the isomorphisms of the completions of the $`^{}`$-algebras in question to the reader. $`\mathrm{}`$ . ## 4. Strict deformation quantization As we have seen, the idea of $`C^{}`$-algebraic deformation quantization is to relate a given Poisson algebra to a family of $`C^{}`$-algebras. ###### Definition 4.1. A Poisson algebra is a complex commutative associative algebra equipped with a Lie bracket $`\{,\}`$ for the which the Leibniz rule holds; that is, one has (4.1) $$\{f,gh\}=\{f,g\}h+g\{f,h\}.$$ A Poisson manifold is a manifold $`P`$ equipped with a Lie bracket $`\{,\}`$ on $`C^{\mathrm{}}(P)`$ that together with pointwise multiplication turns $`C^{\mathrm{}}(P)`$ into a Poisson algebra. This definition is due to Lichnerowicz \[Li\] and Kirillov \[Ki\]. The Lie bracket in a Poisson algebra is usually called the Poisson bracket. For the theory of Poisson manifolds we refer to these papers, and to the textbooks \[MR, Va\]. A Poisson algebra is the classical analogue of a $`C^{}`$-algebra \[La3\]. This analogy is clear if one sees a $`C^{}`$-algebra as a non-associative version of a Poisson algebra, in which the commutative product is given by the anti-commutator $`xy+yx`$ and the Lie bracket is $`\{x,y\}=i(xyyx)`$. The Leibniz rule is then satisfied. The analogy is even better when one defines a Poisson algebra as a real algebra, and restricts the above two operations to the self-adjoint part of a $`C^{}`$-algebra. In this paper, the key example of a Poisson algebra is provided by the dual bundle $`E^{}`$ of a Lie algebroid $`E`$ over $`M`$ (cf. Definition 2.2). This Poisson structure was discovered in \[CDW, Cou\]. The corresponding bracket is most easily defined by listing special cases by which it is uniquely determined; these are (4.2) $`\{f,g\}`$ $`=`$ $`0;`$ (4.3) $`\{\stackrel{~}{X},f\}`$ $`=`$ $`(\rho X)f;`$ (4.4) $`\{\stackrel{~}{X},\stackrel{~}{Y}\}`$ $`=`$ $`\stackrel{~}{[X,Y]_{A(G)}}.`$ Here $`f,gC^{\mathrm{}}(M)`$ are regarded as functions on $`E^{}`$ in the obvious way, and $`\stackrel{~}{X}C^{\mathrm{}}(E^{})`$ is defined by a section $`XC^{\mathrm{}}(M,E)`$ through $`\stackrel{~}{X}(\theta )=\theta (X(\tau ^{}(\theta )))`$. See \[CDW\] for an intrinsic definition. In particular, a Lie groupoid $`G`$ canonically determines a Poisson algebra $`C^{\mathrm{}}(A^{}(G))`$. This is the generic Poisson algebra that we are going to quantize. Our first definition of $`C^{}`$-algebraic deformation quantization uses the concept of a continuous field of $`C^{}`$-algebras. We here state a reformulation of Dixmier’s familiar definition \[Di\] due to Kirchberg–S. Wassermann \[KW\], which is tailor made for our applications. ###### Definition 4.2. A continuous field of $`C^{}`$-algebras $`(C,\{A_t,\phi _t\}_{tT})`$ over a locally compact Hausdorff space $`T`$ consists of a $`C^{}`$-algebra $`C`$, a collection of $`C^{}`$-algebras $`\{A_t\}_{tT}`$, and a set $`\{\phi _t:CA_t\}_{tT}`$ of surjective $`^{}`$-homomorphisms, such that for all $`cC`$ 1. the function $`t\phi _t(c)`$ is in $`C_0(T)`$; 2. one has $`c=sup_{tT}\phi _t(c)`$; 3. there is an element $`fcC`$ for any $`fC_0(T)`$ for which $`\phi _t(fc)=f(t)\phi _t(c)`$ for all $`tT`$. The continuous cross-sections of the field in the sense of \[Di\] consist of those elements $`\{a_t\}_{tT}`$ of $`_{tT}A_t`$ for which there is a (necessarily unique) $`aC`$ such that $`a_t=\phi _t(a)`$ for all $`tT`$. Our first definition of $`C^{}`$-algebraic deformation quantization is now as follows \[La3, La4\]. ###### Definition 4.3. A strict deformation quantization of a Poisson manifold $`P`$ consists of 1. a dense Poisson algebra $`A^0C_0(P)`$ under the given Poisson bracket $`\{,\}`$ on $`P`$; 2. A subset $`I`$ containing $`0`$ as an accumulation point; 3. a continuous field of $`C^{}`$-algebras $`(C,\{A_{\mathrm{}},\phi _{\mathrm{}}\}_\mathrm{}I)`$, with $`A_0=C_0(P)`$; 4. a linear map $`𝒬:A^0C`$ that satisfies (with $`𝒬_{\mathrm{}}(f)\phi _{\mathrm{}}(𝒬(f))`$) (4.5) $`𝒬_0(f)`$ $`=`$ $`f;`$ (4.6) $`𝒬_{\mathrm{}}(f^{})`$ $`=`$ $`𝒬_{\mathrm{}}(f)^{},`$ for all $`fA^0`$ and $`\mathrm{}I`$ , and for all $`f,gA^0`$ satisfies Dirac’s condition (4.7) $$\underset{\mathrm{}0}{lim}(i\mathrm{})^1[𝒬_{\mathrm{}}(f),𝒬_{\mathrm{}}(g)]𝒬_{\mathrm{}}(\{f,g\})=0.$$ In view of the comment after Definition 4.2, for fixed $`fA^0`$ each family $`\{𝒬_{\mathrm{}}(f)\}_\mathrm{}I`$ is a continuous cross-section of the continuous field in question. In view of (4.5) this implies, in particular, that (4.8) $$\underset{\mathrm{}0}{lim}𝒬_{\mathrm{}}(f)𝒬_{\mathrm{}}(g)𝒬_{\mathrm{}}(fg)=0.$$ This shows that a strict deformation quantization yields asymptotic morphisms in the sense of $`E`$-theory \[Co2\]. See \[La3\] for an extensive discussion of quantization theory from the above perspective. Every good definition is the hypothesis of a theorem. Indeed, our first main result \[La3, La4\] is as follows. ###### Theorem 4.4. Let $`G`$ be a Lie groupoid, with associated $`C^{}`$-algebras $`C^{}(G)`$ and $`C_r^{}(G)`$, and Poisson manifold $`A^{}(G)`$. There exists a strict deformation quantization of $`A^{}(G)`$ for which $`I=`$, $`A_0=C_0(A^{}(G))`$, and $`A_{\mathrm{}}=C^{}(G)`$ for $`\mathrm{}0`$. In particular, there exists a continuous field of $`C^{}`$-algebras over $``$ with these fibers. The same claim holds with $`C^{}(G)`$ replaced by $`C_r^{}(G)`$. A different definition of $`C^{}`$-algebraic deformation quantization, which generalizes Rieffel’s original definition \[Ri2\], was introduced in \[Ra1\]. This definition is closer to the notion of formal deformation quantization introduced in \[BFFLS\] than Definition 4.3. ###### Definition 4.5. Let $`P`$ be a Poisson manifold. A semi-strict deformation quantization of $`P`$ consists of 1. a dense Poisson algebra $`A^0C_0(P)`$ under the given Poisson bracket $`\{,\}`$ on $`P`$; 2. A subset $`I`$ containing $`0`$ as an accumulation point; 3. For each $`hI`$, an associative product $`\times _{\mathrm{}}`$, an involution $`^{_{\mathrm{}}}`$, and a $`C^{}`$-semi-norm $`_{\mathrm{}}`$ on $`A^0`$, such that 1. For $`\mathrm{}=0`$ we recover the usual product, involution and norm of $`C_0(P)`$; 2. For each $`fA^0`$, the map $`hf_{\mathrm{}}`$ is continuous; 3. For all $`f,gA^0`$ one has (4.9) $$\underset{\mathrm{}0}{lim}(i\mathrm{})^1(f\times _{\mathrm{}}gg\times _{\mathrm{}}f)\{f,g\}_{\mathrm{}}=0.$$ A strict deformation quantization cannot necessarily be turned into a semi-strict one, since $`𝒬(A^0)𝒬(A^0)`$ is not necessarily contained in $`𝒬(A^0)`$, so that one may not be able to transfer the products on the $`A_{\mathrm{}}`$ to an $`\mathrm{}`$-dependent product on $`A^0`$. Conversely, a semi-strict deformation quantization is not necessarily strict, for the seminorms can fail to be norms. However, let $`fA^0`$ be a non-null function. Then $`f_00`$ and by the condition 3.2 in Definition 4.5 one has $`f_{\mathrm{}}0`$ for $`h`$ in a neighborhood of $`0`$. For each $`\mathrm{}I`$, $`N_{\mathrm{}}=\left\{fA^0\right|f_{\mathrm{}}=0\}`$ is a two-sided closed ideal in $`(A^0,\times _{\mathrm{}},{}_{}{}^{_{\mathrm{}}},_{\mathrm{}})`$, and $`A^0/N_{\mathrm{}}`$ is a pre-$`C^{}`$-algebra, with completion $`A_{\mathrm{}}^{}`$. Defining $`𝒬_{\mathrm{}}`$ as the canonical projection of $`A^0`$ onto $`A^0/N_{\mathrm{}}`$, we almost obtain a strict deformation quantization of $`P`$; the only difficulty is that the $`A_{\mathrm{}}^{}`$ may not form a continuous field of $`C^{}`$-algebras. The counterpart of Theorem 4.4 for strict deformation quantization is as follows \[Ra1\]. ###### Theorem 4.6. Let $`A(G)`$ be the Lie algebroid of a Lie groupoid, with associated Poisson manifold $`A^{}(G)`$. There exists a semi-strict deformation quantization of $`A^{}(G)`$ over $`I=`$. As explained above, this theorem does not follow from Theorem 4.4, though the proofs have many elements in common. We now briefly outline the organization of the proofs. Sections 5 and 6 address the continuous field of $`C^{}`$-algebras claimed to exist in Theorem 4.4, and in similar vein contain the heart of the proof of condition (b) of Definition 4.5 in Theorem 4.6. Section 7 develops the theory of the Fourier transform on vector bundles. In section 8 this theory is used to define the map $`𝒬`$ of Theorem 4.4 as well as the product $`\times _{\mathrm{}}`$, the involution $`_{\mathrm{}}`$, and the seminorm $`_{\mathrm{}}`$ of Theorem 4.6. The proofs of both theorems are then complete up to Dirac’s condition (4.7) and (4.9). Section 9 develops local techniques that give detailed information on the Poisson structure on $`A^{}(G)`$. Finally, (4.7) and (4.9) are proved in section 10. ## 5. Continuous fields of groupoids Rieffel has developed useful techniques for proving continuity of fields of $`C^{}`$-algebras occurring in examples of strict deformation quantization in the context of groups \[Ri1\]. The main result of this section, Theorem 5.5, generalizes these results to the context of groupoids. This theorem, which is an application of results of Blanchard \[Bl\], gives information on the field of $`C^{}`$-algebras associated to a continuous field of locally compact groupoids. Although the examples studied in this paper concern Lie groupoids, we prove Theorem 5.5 for the general case of locally compact groupoids. We first mention a known result. Recall that, when $`EG_0`$, $`G_E`$ is the subgroupoid of $`G`$ consisting of all $`\gamma G_1`$ for which $`r(\gamma )E`$ and $`s(\gamma )E`$. ###### Proposition 5.1. Let $`G`$ be a locally compact groupoid with left Haar system $`(\lambda ^x)_{xG_0}`$, and let $`U`$ be an open invariant subset of $`G_0`$. Write $`F=G_0\backslash U`$. 1. The following sequence is exact: (5.1) $$0C^{}(G_U)\stackrel{e}{}C^{}(G)\stackrel{i^{}}{}C^{}(G_F)0.$$ Here $`e`$ is firstly defined as map from $`C_c(G_U)`$ to $`C_c(G_1)`$ by extending a function on $`G_U`$ to one on $`G_1`$ by making it zero on the complement of $`G_U`$, and secondly extended to a map from $`C^{}(G_U)`$ to $`C^{}(G)`$ by continuity. Similarly, $`i^{}`$ is firstly defined from $`C_c(G_1)`$ to $`C_c(G_F)`$ as the pullback of the inclusion $`G_FG_1`$, and then extended by continuity. 2. If the groupoid $`G_F`$ is amenable, then the following sequence is exact: (5.2) $$0C_r^{}(G_U)\stackrel{e}{}C_r^{}(G)\stackrel{i^{}}{}C_r^{}(G_F)0.$$ This result was given by Torpe \[To\] in the case of $`C^{}`$-algebras associated to foliations, and by Hilsum and Skandalis \[HS\] in the general case; also see \[Re1\], p.102. A complete proof, based on Renault’s theorem of disintegration of representations \[Re2\], may be found in \[Ra1\]. A counterexample that shows that the second sequence of reduced $`C^{}`$-algebras is not necessarily exact for non-amenable $`G_F`$ is given in \[Re3\]. ###### Definition 5.2. A field of groupoids is a triple $`(G,T,p)`$, with $`G`$ a groupoid, $`T`$ a set, and $`p:GT`$ a surjective map such that $`p=p_0r=p_0s`$, where $`p_0=p_{|_{G_0}}`$. If $`G`$ is locally compact, $`T`$ Hausdorff, and $`p`$ continuous and open, we say that $`(G,T,p)`$ is a continuous field of locally compact groupoids. If $`G`$ is a Lie groupoid, $`T`$ a manifold, and $`p`$ a submersion, $`(G,T,p)`$ is called a smooth field of Lie groupoids. If $`(G,T,p)`$ is a field of groupoids and $`YT`$, then $`A=p_0^1(Y)`$ is an invariant subset of $`G_0`$, and $`G_A=p^1(Y)`$ is a subgroupoid of $`G`$. In the case of a continuous field of locally compact groupoids, (5.3) $$G(t)=p^1(\{t\})$$ is a closed locally compact subgroupoid of $`G`$ for every $`tT`$. In the case of a smooth field of Lie groupoids, $`G(t)`$ is a Lie groupoid for every $`tT`$. In this section, $`(G,T,p)`$ is a continuous field of locally compact groupoids over a locally compact Hausdorff space $`T`$, for which there exists a left Haar system $`(\lambda ^x)_{xG_0}`$ on $`G`$. Let $`C_0(T)`$ be the $`C^{}`$-algebra of continuous functions on $`T`$ that vanish at infinity. We define a structure of $`C_0(T)`$ left module on the space of continuous compactly supported functions $`C_c(G_1)`$ by $`(fa)(\gamma )=f(p(\gamma ))a(\gamma )`$, where $`fC_0(T)`$ and $`aC_c(G_1)`$. For every $`fC_0(T)`$ and $`a,bC_c(G_1)`$, the following equalities hold: $`f(ab)`$ $`=`$ $`(fa)b=a(fb);`$ (5.4) $`(fa)^{}`$ $`=`$ $`f^{}a^{}.`$ We also have $`C_0(T)C_c(G_1)=C_c(G_1)`$. Indeed, for every $`aC_c(G_1)`$, there is a function $`fC_c(T)`$ such that $`f=1`$ on $`p(\text{supp}a)`$, and then $`a=faC_0(T)C_c(G_1)`$. An easy rewriting of Lemma 1.13 from \[Re1\] proves ###### Lemma 5.3. For every representation $`\pi :C_c(G_1)()`$ that is continuous for the inductive limit topology there is a unique continuous representation $`\varphi :C_0(T)()`$ such that $`\pi (fa)=\varphi (f)\pi (a)`$. Using the previous lemma for $`fC_0(T)`$, $`aC_c(G_1)`$, and $`\pi `$ continuous representation of $`C_c(G_1)`$, we have $`\pi (fa)\varphi (f)\pi (a)fa`$, hence $`fafa`$. For every $`fC_0(T)`$, the map $`C_c(G_1)afaC_c(G_1)`$ has a continuous extension on $`C^{}(G)`$. This provides a $`C_0(T)`$ Banach module structure on $`C^{}(G)`$. Corollary 1.9 in \[Bl\] shows that $`C_0(T)C^{}(G)`$ is closed in $`C^{}(G)`$. But one has $`C_0(T)C^{}(G)C_0(T)C_c(G_1)=C_c(G_1)`$, and we obtain that $`C^{}(G)`$ is a nondegenerate module. For every $`a,bC^{}(G)`$ and every $`fC_0(T)`$, the conditions $`f(ab)=(fa)b=a(fb)`$ and $`(fa)^{}=f^{}a^{}`$ are easily verified, hence $`C^{}(G)`$ is a $`C_0(T)`$ $`^{}`$-algebra (cf. \[Bl\]). Similar arguments prove that $`C_r^{}(G)`$ is a $`C_0(T)`$ $`^{}`$-algebra. ###### Lemma 5.4. Let $`T`$, $`Y`$ be topological spaces, $`p:YT`$ an open onto map, and $`\phi :Y`$ an upper semicontinuous function such that for every $`tT`$ one has sup$`\left\{\phi (y)\right|p(y)=t\}<\mathrm{}`$. Then the function $`\psi :T`$ given by $`\psi (t)=\underset{p(y)=t}{sup}\phi (y)`$ is lower semi-continuous. ###### Proof. Let $`t_i`$ be a net converging in $`T`$ to $`t`$, and pick $`a`$. We show that $`\psi (t_i)a`$ implies $`\psi (t)a`$. Let $`yY`$ be such that $`p(y)=t`$. A lemma on open maps (cf. \[FD\], p.126) proves the existence of a net $`y_iY`$, for which $`p(y_i)=t_i`$ and $`y_iy`$. For every $`i`$, $`\phi (y_i)\psi (t_i)a`$. But $`\phi `$ is lower semicontinuous, so $`\phi (y)a`$ and $`\psi (t)a`$. $`\mathrm{}`$ We now come to the main result of this section. ###### Theorem 5.5. Let $`(G,T,p)`$ be a continuous field of locally compact groupoids, take $`aC_c(G_1)`$, and, for each $`tT`$, write $`a_t=a_{|_{G(t)}}`$. Then 1. The map $`ta_t_{C^{}(G(t))}`$ is upper semicontinuous. 2. The map $`ta_t_{C_r^{}(G(t))}`$ is lower semicontinuous. ###### Proof. 1. For $`tT`$, $`C_t(T)=\left\{fC_0(T)\right|f(t)=0\}`$ is a closed two sided ideal, hence $`C_t(T)C^{}(G)`$ is a closed two sided ideal in $`C^{}(G)`$. Corollary 1.9 in \[Bl\] applied to the $`C_t(T)`$ Banach module $`C^{}(G)`$ shows that $`C_t(T)C^{}(G)`$ is closed in $`C^{}(G)`$. Let $`\pi _t:C^{}(G)C^{}(G)/C_t(T)C^{}(G)`$ be the quotient map. By Lemma 1.10 from \[Bl\] one has $`\pi _t(a)=\underset{fC_0(T)}{inf}(1f+f(t))a`$; hence the map $`t\pi _t(a)`$ is upper semicontinuous as the infimum of a family of continuous functions. Denote the closed invariant subset $`p_0^1(\left\{t\right\})`$ of $`G_0`$ by $`F`$, and let $`U=G_0\backslash F`$. Obviously, $`G(t)=G_F`$. We claim that (5.5) $$i\left(C_c(G_U)\right)=C_t(T)C_c(G),$$ where $`i:C^{}(G_U)C^{}(G)`$ is as in Proposition 5.1. Let $`ai\left(C_c(G_U)\right)`$. In $`T`$ there is an open neighborhood $`V`$ of the compact set $`p(\text{supp}a)`$ such that $`tV`$. We take $`fC_c(T)`$ with $`f=1`$ on $`p(\text{supp}a)`$ and $`\text{supp}fV`$, and then $`a=faC_t(T)C_c(G_1)`$. The other inclusion is obvious. Taking the norm closure in (5.5), we obtain $`i(C^{}(G_U))=C_t(T)C^{}(G)`$. With Proposition 5.1 this shows that $`C^{}(G(t))=C^{}(G)/C_t(T)C^{}(G)`$, so $`\pi _t(a)=a_t`$, which ends the proof. 2. The left regular representation $`\pi _x^L(f):L^2(G_1,\lambda ^x)L^2(G_1,\lambda ^x)`$ of $`C_c(G_1)`$ associated to $`xG_0`$ is given by (5.6) $$\pi _x^L(f)\xi (\gamma )=_{G^x}f(\gamma \gamma _1)\xi (\gamma _1^1)𝑑\lambda ^x(\gamma _1).$$ Set $`\xi _{\mathrm{}}=\underset{xG_0}{sup}\xi _{L^2(G_1,\lambda ^x)}`$; then $$\pi _x^L(f)=sup\left\{\pi _x^L(f)\xi ,\eta \right|\xi ,\eta C_c(G_1),\xi _{\mathrm{}}<\mathrm{},\eta _{\mathrm{}}<\mathrm{}\}.$$ The map $`x\pi _x^L(f)\xi ,\eta `$ is continuous for each $`fC_c(G_1)`$ and $`\xi ,\eta C_c(G_1)`$, hence $`x\pi _x^L(f)`$ is lower semicontinuous. But $`a_t=\underset{tp^1(t)}{sup}\pi _x^L(a_t)`$, and Lemma 5.4 ends the proof. $`\mathrm{}`$ ###### Corollary 5.6. If $`(G,T,p)`$ is a continuous field of locally compact groupoids and $`aC_c(G_1)`$, then 1. The maps $`ta_t`$ and $`ta_t_r`$ are continuous at every $`tT`$ for which the groupoid $`G(t)`$ is amenable. 2. If $`G`$ is amenable, then the maps $`ta_t`$ and $`ta_t_r`$ are continuous. ###### Proof. 1. Straightforward. 2. By Proposition 5.2.3 from \[AR\], $`G`$ is amenable if and only if $`G(t)`$ is amenable for every $`tT`$. $`\mathrm{}`$ ## 6. The tangent groupoid The tangent groupoid of a manifold was introduced by Connes \[Co2\]. This idea was generalized to arbitrary Lie groupoids by Hilsum and Skandalis \[HS\] under the name normal groupoid, and by Weinstein \[We1\] under the name blowup. Connes’s construction corresponds to the pair groupoid of a manifold. We here use the name “tangent groupoid” also for the general case. ###### Definition 6.1. Let $`G`$ be a Lie groupoid with Lie algebroid $`A(G)`$. The tangent groupoid $`\widehat{G}`$ is a Lie groupoid with base $`\widehat{G}_0=[0,1]\times G_0`$, defined by the following structures. * As a set, $`\widehat{G}_1=A(G)\{(0,1]\times G_1\}`$. We write elements of $`\widehat{G}`$ as pairs $`(\mathrm{},u)`$, where $`uA(G)`$ for $`\mathrm{}=0`$ and $`uG`$ for $`\mathrm{}0`$. Thus $`A(G)`$ is identified with $`\{0\}\times A(G)`$. * As a groupoid, $`\widehat{G}=\{0\times A(G)\}\{(0,1]\times G\}`$. Here $`A(G)`$ is regarded as a Lie groupoid over $`G_0`$ with $`s=r=\tau `$ (the bundle projection of $`A(G)`$), and addition in the fibers as the groupoid multiplication. The groupoid operations in $`(0,1]\times G`$ are those in $`G`$. For example, for $`\mathrm{}0`$ one has (6.1) $$\widehat{r}(\mathrm{},u)=(\mathrm{},r(u).$$ * The smooth structure on $`\widehat{G}_1`$, making it a manifold with boundary, is as follows. To start, the open subset $`𝒪_1=(0,1]\times G_1\widehat{G}_1`$ inherits the product manifold structure. Let $`G_0VA(G)`$ and $`\iota (G_0)WG_1`$, as in Theorem 2.6. Let $`𝒪`$ be the open subset of $`[0,1]\times A(G)`$ (equipped with the product manifold structure; this is a manifold with boundary, since $`[0,1]`$ is), defined as $`𝒪=\{(\mathrm{},X)|\mathrm{}XV\}`$. Note that $`\{0\}\times A(G)𝒪`$. The map $`\psi :𝒪\widehat{G}_1`$ is defined by $`\psi (0,X)`$ $`=`$ $`(0,X);`$ (6.2) $`\psi (\mathrm{},X)`$ $`=`$ $`(\mathrm{},\mathrm{Exp}^W(\mathrm{}X)).`$ Since $`\mathrm{Exp}^W:VW`$ is a diffeomorphism (cf. Proposition 2.6) we see that $`\psi `$ is a bijection from $`𝒪`$ to $`𝒪_2=\{0\times A(G)\}\{(0,1]\times W\}`$. This defines the smooth structure on $`𝒪_2`$ in terms of the smooth structure on $`𝒪`$. Since $`𝒪_1`$ and $`𝒪_2`$ cover $`\widehat{G}_1`$, this specifies a smooth structure on $`\widehat{G}_1`$. Using $`\mathrm{Exp}^L`$ rather than $`\mathrm{Exp}^W`$ in this definition, one obtains the smooth structure defined in \[HS\], which is equivalent to the one constructed above. A proof that the change of coordinates from $`𝒪_1`$ to $`𝒪_2`$ on their intersection is smooth may be found in \[HS\]. We omit the proof that $`\widehat{G}`$ is a Lie groupoid; see \[HS\], and, for full details, \[Ra1\]. It follows that $`\widehat{G}`$ is a smooth field of Lie groupoids over $``$. ###### Proposition 6.2. Let $`G`$ be a Lie groupoid with tangent groupoid $`\widehat{G}`$. Writing $`^{}=\backslash \{0\}`$, the following sequences are exact: (6.3) $`0`$ $``$ $`C_0(^{})C^{}(G)C^{}(\widehat{G})C_0(A^{}(G))0;`$ (6.4) $`0`$ $``$ $`C_0(^{})C_r^{}(G)C_r^{}(\widehat{G})C_0(A^{}(G))0.`$ ###### Proof. Consider the open invariant subset $`U=G_0\times ^{}`$ of $`\widehat{G}_0`$, with complement $`F=G_0\backslash U`$. Then $`\widehat{G}_U=G\times ^{}`$ and $`\widehat{G}_F=A(G)\times \left\{0\right\}`$. By Proposition 5.1 we have the exact sequence $$0C^{}(G\times ^{})C^{}(\widehat{G})C^{}(A(G))0.$$ Gelfand’s theorem implies $`C^{}(A(G))=C_0(A^{}(G))`$ (also cf. section 7), and since $`C^{}(G\times ^{})`$ is the $`C^{}`$-algebra of the trivial field of the $`C^{}`$-algebras $`C^{}(G)`$ on $`^{}`$, we have $`C^{}(G\times ^{})=C_0(^{})C^{}(G)`$. The second exact sequence is a consequence of Proposition 5.1 as well, since $`A(G)`$ is commutative as a groupoid (cf. Definition 6.1), and therefore amenable \[AR\]. $`\mathrm{}`$ ###### Corollary 6.3. (cf. \[Co2\]) Let $`M`$ be a manifold. Then the tangent groupoid $`\widehat{M\times M}=(M\times M\times ^{})(TM\times \left\{0\right\})`$ is amenable, and the sequence (6.5) $$0C_0(^{})𝒦(L^2(M))C^{}(\widehat{M\times M})C_0(T^{}M)0$$ is exact. ###### Proof. By Proposition 5.2.3 in \[AR\], a groupoid bundle is amenable if and only if all its fibers are amenable. Since the Lie algebroid $`A(G)`$ is amenable as a groupoid, the amenability of $`\widehat{G}`$ is reduced to the amenability of the groupoid $`G`$. In this case, $`C^{}(\widehat{G})=C_r^{}(\widehat{G})`$, cf. \[Re1, AR\]. Now use Proposition 6.2 for the amenable groupoid $`G=M\times M`$ to finish the proof. $`\mathrm{}`$ Recall Definition 4.2, whose notation we adopt. ###### Theorem 6.4. Let $`G`$ be a Lie groupoid with tangent groupoid $`\widehat{G}`$. For $`\mathrm{}`$, define $`G(0)`$ $`=`$ $`A(G);`$ (6.6) $`G(\mathrm{})`$ $`=`$ $`G\mathrm{}0.`$ Define $`\widehat{\phi }_{\mathrm{}}:C_c^{\mathrm{}}(\widehat{G})C_c^{\mathrm{}}(G(\mathrm{}))`$ as the pullback of the inclusion $`G(\mathrm{})\widehat{G}`$ (cf. Definition 6.1); in other words, $`\widehat{\phi }_{\mathrm{}}(\widehat{f})`$ is the restriction of $`\widehat{f}C_c^{\mathrm{}}(\widehat{G})`$ to $`G(\mathrm{})`$. 1. Each $`\widehat{\phi }_{\mathrm{}}`$ may be extended by continuity to a surjective $`^{}`$-homomorphism $`\phi _{\mathrm{}}:C^{}(\widehat{G})C^{}(G(\mathrm{}))`$, and also to a surjective $`^{}`$-homomorphism $`\phi _{(r)\mathrm{}}:C_r^{}(\widehat{G})C_r^{}(G(\mathrm{}))`$. 2. The $`C^{}`$-algebras $`C=C^{}(\widehat{G})`$ and $`A_{\mathrm{}}=C^{}(G(\mathrm{}))`$, and the maps $`\phi _{\mathrm{}}`$ form a continuous field of $`C^{}`$-algebras over $`I=`$. 3. The $`C^{}`$-algebras $`C=C_r^{}(\widehat{G})`$ and $`A_{\mathrm{}}=C_r^{}(G(\mathrm{}))`$, and the maps $`\phi _{(r)\mathrm{}}`$ form a continuous field of $`C^{}`$-algebras over $`I=`$. ###### Proof. 1. Using the definition of the norms in question, one checks that each map $`\widehat{\phi }_{\mathrm{}}`$ is contractive. The $`^{}`$-homomorphism property is obvious on $`C_c^{\mathrm{}}(\widehat{G})`$ from the definition of $`\widehat{G}`$ and $`C^{}(G)`$, and extends by continuity. 2. Condition 1 in Definition 4.2 follows at $`\mathrm{}0`$ since the field of $`C^{}`$-algebras is trivial away from 0. At $`\mathrm{}=0`$ continuity follows from Corollary 5.6, since $`A(G)`$, being a commutative groupoid, is amenable \[AR\]. 3. The reduced case is proved in the same way. $`\mathrm{}`$ For a different proof of this theorem, note that each family $`\widehat{\phi }_{\mathrm{}}(\widehat{f})`$, $`\widehat{f}C_c^{\mathrm{}}(\widehat{G})`$, is stable under convolution and involution, and dense in $`C_{(r)}^{}(G)`$. One then uses Prop. 10.3.2 in \[Di\] and the equivalence between Kirchberg–S. Wassermann’s Definition 4.2 and Dixmier’s definition Def. 10.3.1 in \[Di\] of a continuous field of $`C^{}`$-algebras. Condition 10.1.2 (ii) in \[Di\] is then proved in the same way as above. A different proof of Theorem 6.4, which makes no use of the results in section 5 is possible as well \[La3, La4\]. This proof is based on the following lemma, due Lee; cf. Theorem 4 in \[Le\]. ###### Lemma 6.5. Let $`C`$ be a $`C^{}`$-algebra, and let $`\psi :\text{Prim}(C)T`$ be a continuous and open map from the primitive spectrum $`\text{Prim}(C)`$ (equipped with the Jacobson topology \[Di\]) to a locally compact Hausdorff space $`T`$. Define $`I_t=\psi ^1(t)`$; i.e., $`cI_t`$ iff $`\pi _I(c)=0`$ for all $`I\psi ^1(t)`$ (here $`\pi _I(C)`$ is the irreducible representation whose kernel is $`I`$). Note that $`I_t`$ is a closed two-sided ideal in $`C`$. Taking $`A_t=C/I_t`$ and $`\phi _t:CA_t`$ to be the canonical projection, $`(C,\{A_t,\phi _t\}_{tT})`$ is a continuous field of $`C^{}`$-algebras. For a proof see \[Le\] or \[ENN1\]. We apply this lemma with $`C=C^{}(\widehat{G})`$ and $`T=I=`$. In order to verify the assumption in the lemma, we first note that $`I_0C_0((0,1])C^{}(G)`$, as follows from a glance at the topology of $`\widehat{G}`$. Hence $`\text{Prim}(I_0)=(0,1]\times \text{Prim}(C^{}(G))`$, with the product topology. Furthermore, one has $`C^{}(\widehat{G})/I_0C^{}(A(G))C_0(A^{}(G))`$. Hence $`\text{Prim}(C^{}(\widehat{G})/I_0)A^{}(G)`$. Using this in Prop. 3.2.1 in \[Di\], with $`A=C^{}(\widehat{G})`$ and $`I`$ the ideal $`I_0`$ generated by those $`fC_c^{\mathrm{}}(\widehat{G})`$ that vanish at $`\mathrm{}=0`$, yields the decomposition (6.7) $$\text{Prim}(C^{}(\widehat{G}))A^{}(G)\{(0,1]\times \text{Prim}(C^{}(G))\},$$ in which $`A^{}(G)`$ is closed. This does not provide the full topology on $`\text{Prim}(C^{}(\widehat{G}))`$, but it is sufficient to know that $`A^{}(G)`$ is not open. If it were, $`(0,1]\times \text{Prim}(C^{}(G))`$ would be closed in $`\text{Prim}(C^{}(\widehat{G}))`$, and this possibility can be excluded by looking at the topology of $`\widehat{G}`$ and the definition of the Jacobson topology. Using (6.7), we can define a map $`\psi :\text{Prim}(C^{}(\widehat{G}))[0,1]`$ by $`\psi (I)=0`$ for all $`IA^{}(G)`$ and $`\psi (\mathrm{},I)=\mathrm{}`$ for $`\mathrm{}0`$ and $`I\text{Prim}(C^{}(G))`$. It is clear from the preceding considerations that $`\psi `$ is continuous and open. Using this in Lemma 6.5, one sees that $`I_{\mathrm{}}`$ is the ideal in $`C^{}(\widehat{G})`$ generated by those $`\widehat{f}C_c^{\mathrm{}}(\widehat{G})`$ vanish at $`\mathrm{}`$. Hence $`A_0C_0(A^{}(G))`$, as above, and $`A_{\mathrm{}}C^{}(G)`$ for $`\mathrm{}0`$. Theorem 6.4 then follows from Lemma 6.5. We now give a left Haar system for the tangent groupoid $`\widehat{G}`$. Firstly, pick a positive section $`\mu C^{\mathrm{}}(G_0,|\mathrm{\Omega }|^1(A(G)))`$, and let $`\lambda `$ be the associated smooth, positive, left invariant section of $`|\mathrm{\Omega }|^1(\mathrm{ker}Tr)`$, as in Proposition 3.2; hence $`\mu =\lambda |_{G_0}`$. Now define a section $`\widehat{\lambda }C^{\mathrm{}}(\widehat{G}_1,|\mathrm{\Omega }|^1(\mathrm{ker}T\widehat{r})`$ by $`\widehat{\lambda }(x,X,0)`$ $`=`$ $`\lambda (x);`$ (6.8) $`\widehat{\lambda }(\gamma ,\mathrm{})`$ $`=`$ $`|\mathrm{}|^p\lambda (\gamma ).`$ Here $`p`$ is the dimension of the typical fiber of $`A(G)`$; the factor $`|\mathrm{}|^p`$ is necessary in order to have a smooth system also at $`\mathrm{}=0`$, as is easily verified using the manifold structure on $`\widehat{G}`$. This section is smooth, positive and left invariant, so that it defines a left Haar system $`(\widehat{\lambda }^{(\mathrm{},x)})_{(\mathrm{},x)\widehat{G}_0}`$ for $`\widehat{G}`$ by Proposition 3.4. The $`^{}`$-algebraic structure on $`C_c^{\mathrm{}}(\widehat{G},\widehat{\lambda })`$ defined by (3.7) and (3.8) with (6.8) becomes (6.9) $`\widehat{f}\widehat{g}(0,\xi _x)`$ $`=`$ $`{\displaystyle _{\tau ^1(x)}}\widehat{f}(0,\xi \eta _x)\widehat{g}(0,\eta _x)𝑑\mu ^x(\eta _x);`$ (6.10) $`\widehat{f}\widehat{g}(\mathrm{},\gamma )`$ $`=`$ $`|\mathrm{}|^p{\displaystyle _{G^{s(\gamma )}}}\widehat{f}(\mathrm{},\gamma \gamma _1)\widehat{g}(\mathrm{},\gamma _1^1)𝑑\lambda ^{s(\gamma )}(\gamma _1);`$ (6.11) $`\widehat{f}^{}(0,\xi )`$ $`=`$ $`\overline{\widehat{f}(0,\xi )};`$ (6.12) $`\widehat{f}^{}(\mathrm{},\gamma )`$ $`=`$ $`\overline{\widehat{f}(\mathrm{},\gamma ^1)}.`$ Here $`\xi _x\tau ^1(x)`$, and $`\mu ^x`$ is the measure on $`\tau ^1(x)A(G)`$ generated by the density $`\mu (x)`$ (cf. (3.1)). Finally, $`(\lambda ^x)`$ is the left Haar system on $`G`$ defined by $`\lambda `$ according to Proposition 3.4. These formulae should be compared with (3.7) and (3.8). ## 7. The Fourier transform on vector bundles In this section we extend the notion of a Fourier transform from $`^n`$ to vector bundles, and show that it remains an isomorphism of a suitably defined Schwartz space of rapidly decreasing functions. Let $`M`$ be a manifold of dimension $`n`$, and let $`E`$ be a vector bundle over $`M`$ with fiber dimension $`p`$. Now $`E`$ is a commutative Lie groupoid if the source and range projections are both equal to the bundle projection, and groupoid multiplication is addition in each fiber. The $`C^{}`$-algebra $`C^{}(E)`$ is then commutative, too. Using the proposition on page 582 of \[FD\], it can be shown that the maximal ideal space of $`C^{}(E)`$ is $`E^{}`$. The Gelfand transform is an isomorphism between the $`C^{}`$-algebras $`C^{}(E)`$ and $`C_0(E^{})`$. Its explicit form is a fiberwise Fourier transform $`:C^{}(E)C_0(E^{})`$. On the convolution algebra $`C_c^{\mathrm{}}(E,E)`$ the Fourier transform is given, for $`\theta _xE_x^{}E^{}`$, by $$(\rho )(\theta _x)=\underset{E_x}{}e^{i\theta _x,\xi _x}\rho (\xi _x).$$ It is easier to write formulas for functions rather than densities. To do so, fix a positive 1-density $`\mu C^{\mathrm{}}(M,|\mathrm{\Omega }|^1(E))`$; cf. the proof of Proposition 3.4. For every $`xM`$, the density $`\mu (x)|\mathrm{\Omega }|^1(E_x)`$ defines a translation invariant measure $`\mu ^x`$ on $`E_x`$, and the family $`(\mu ^x)_{xM}`$ is a smooth Haar system for $`E`$. The Fourier transform is denoted by $`_\mu :C^{}(E,\mu )C_0(E^{})`$ in order to emphasize its dependence on $`\mu `$. For $`fL^1(E,d\mu ^x)`$, $`\theta _xE_x^{}`$, we then have (7.1) $$(_\mu f)(\theta _x)=f(\xi _x)e^{i\theta _x,\xi _x}𝑑\mu ^x(\xi _x).$$ We now generalize the definition given by Rieffel in the case of the trivial bundle $`M\times ^p`$ (cf. \[Ri3\]) to define a Schwartz space on $`E`$. But first we fix some notations. The variables in $`^n\times ^p`$ are denoted $`(u,v)`$, and for $`\beta ^p`$ we put $`_v^\beta F={\displaystyle \frac{^{\beta _1+\mathrm{}+\beta _p}F}{v_1^{\beta _1}\mathrm{}v_p^{\beta _p}}}`$. The Fourier transform $`_u`$ on $`^n\times ^p`$ with $`u`$ constant is given by $`(_uF)(u,w)={\displaystyle F(u,v)e^{iv,w}𝑑v}`$. Let $`(e_1,\mathrm{},e_p)`$ be a local frame of $`E`$, with dual frame $`(e_1^{},\mathrm{},e_p^{})`$ of $`E^{}`$, let $`(q,\lambda )`$ be local coordinates on $`E`$, and finally let $`(q,ϵ)`$ be local coordinates of $`E^{}`$. The expression of a function $`f`$ on $`E`$ in local coordinates is denoted by the corresponding capital letter $`F`$. Then (3.1) for $`fL^1(E,\mu ^x)`$ becomes (7.2) $$f\left(\stackrel{n}{\underset{i=1}{}}v_ie_i(x)\right)\mu _e(x)dv=fd\mu ^x.$$ ###### Definition 7.1. We say that a function $`f:E`$ is $`M`$-compactly supported when $`\pi (\text{supp}f)`$ is relatively compact in $`M`$. The set of continuous $`M`$-compactly supported functions is called $`C_{c,M}(E)`$. A function $`fC_{c,M}(E)`$ is said to be rapidly decreasing if $`\lambda ^\alpha f(q,\lambda )`$ is bounded for every $`\alpha ^p`$, where $`\lambda ^\alpha =\lambda _1^{\alpha _1}\mathrm{}\lambda _p^{\alpha _p}`$, with $`\lambda _1,\mathrm{},\lambda _p`$ the coordinates of $`\lambda `$ in the frame $`(e_1,\mathrm{},e_p)`$. This definition is independent of the frame. Indeed, let $`(e_1,\mathrm{},e_p)`$ be a frame of $`E`$ on an open subset $`U`$ of $`M`$, and suppose that $`\lambda ^\alpha f(q,\lambda )`$ is bounded for every $`\alpha ^p`$. If $`(e_1^{},\mathrm{},e_p^{})`$ is another frame of $`E`$ on $`U`$, let $`g:UGL(p,)`$ be the function given by $`e^{}=eg`$. If we write the same point of $`E`$ in the two frames, its coordinates are related by $`\lambda _i^{}=\underset{j}{}\lambda _jg_{ji}`$. To show that $`\lambda _{}^{}{}_{}{}^{\alpha }f(q,\lambda ^{})`$ is bounded, remark that expanding $`\lambda _{1}^{}{}_{}{}^{\alpha _1}\mathrm{}\lambda _{p}^{}{}_{}{}^{\alpha _p}`$ we obtain a polynomial in $`\lambda _1,\mathrm{},\lambda _p`$ with continuous functions on $`M`$ that vanishes outside the compact closure of $`\pi (\text{supp}f)`$ as coefficients. ###### Definition 7.2. We say that a function $`fC^{\mathrm{}}(E)`$ is of Schwartz type on $`E`$ if $`f`$ is $`M`$-compactly supported and $`_\lambda ^\beta f={\displaystyle \frac{^{\beta _1+\mathrm{}+\beta _p}f}{\lambda _1^{\beta _1}\mathrm{}\lambda _p^{\beta _p}}}`$ is rapidly decreasing for every $`\beta =(\beta _1,\mathrm{},\beta _p)^p`$. The set of Schwartz functions on $`E`$ is denoted by $`S(E)`$. ###### Proposition 7.3. $`S(E)`$ is a dense $`^{}`$-subalgebra of $`C^{}(E,\mu )`$ and $`S(E^{})`$ is a dense $`^{}`$-subalgebra of $`C_0(E^{})`$. ###### Proof. We show that $`S(E)`$ is closed under convolution. For $`f,g`$ in $`S(E)`$ it is obvious that $`fg`$ is $`M`$-compactly supported. Using a partition of unity argument, it easily follows that $`fgS(E)`$. It is trivial that $`S(E)`$ is closed under involution. Now $`S(E)`$ contains the smooth compactly supported functions on $`E`$, and by a result from \[FD\], page 140, $`C_c^{\mathrm{}}(E)`$ is dense in $`C^{}(E,\mu )`$; it follows that $`S(E)`$ is dense in $`C^{}(E,\mu )`$. For the second part of the proposition, $`S(E^{})`$ is obviously a $`^{}`$-subalgebra of $`C_0(E^{})`$ containing the dense subalgebra $`C_c^{\mathrm{}}(E^{})`$ of $`C_0(E^{})`$. $`\mathrm{}`$ ###### Lemma 7.4. If $`fS(E)`$, then $`_\mu fS(E^{})`$ ###### Proof. The restriction of $`f`$ to $`E_x`$ is integrable because it is rapidly decreasing, so (7.1) makes sense in this case. It is then easy to show that $`_\mu f`$ is smooth and $`M`$-compactly supported. Choosing a suitable partition of unity, we may assume that the projection onto $`M`$ of the support of $`f`$ is a subset of the domain of a chart $`\alpha :U^n`$ of $`M`$. We can also suppose that there exists a frame $`(e_1,\mathrm{},e_p)`$ of $`E`$ on $`U`$. The expression of $`f`$ in the corresponding local coordinates of $`E`$ is given by $`F:\alpha (U)\times ^p`$, $`F(u,v)=f(v_ie_i(\alpha ^1(u)))`$. It is straightforward to see that the local expression of $`_\mu f`$, denoted by $`\widehat{F}`$, satisfies $`\widehat{F}(u,w)=\mu _e(\alpha ^1(u))(_uF)(u,w)`$. The relation $`w^\alpha _w^\beta \widehat{F}(u,w)=\mu _e(\alpha ^1(u))(i)^{|\alpha |+|\beta |}_u\left(_v^\alpha (v^\beta F)\right)(u,w)`$ reduces the problem to the fact that $`_uG`$ is bounded for every $`G`$ rapidly decreasing on the trivial bundle $`^n\times ^p`$ of base $`^n`$. But $`\left|_uG(u,w)\right|{\displaystyle |F(u,v)|𝑑v}`$, and the proof is finished by the remark that on the right side we have a continuous compactly supported function in $`u`$. $`\mathrm{}`$ For a vector space $`V`$ and a nowhere vanishing element $`\rho `$ of $`|\mathrm{\Omega }|^\alpha (V)`$, let $`\rho ^{}`$ be the element of $`|\mathrm{\Omega }|^\alpha (V^{})`$ that satisfies $`\rho ^{}(v_1^{}\mathrm{}v_n^{})={\displaystyle \frac{1}{\rho (v_1\mathrm{}v_n)}}`$ for every pair of dual bases $`\{v_1,\mathrm{},v_n\}`$ on $`V`$ and $`\{v_1^{},\mathrm{},v_n^{}\}`$ on $`V^{}`$. Then $`\mu _{}(x)=\left(\mu (x)\right)^{}`$ defines a 1-density $`\mu _{}C^{\mathrm{}}(M,|\mathrm{\Omega }|^1(E^{}))`$, and hence a left Haar system $`(\mu _{}^x)_{xM}`$ on $`E^{}`$ (seen as a Lie groupoid in the same way as $`E`$). ###### Lemma 7.5. The restriction of the Fourier transform $`_\mu `$ to $`S(E)`$ is a bijection onto $`S(E^{})`$, with inverse (7.3) $$(_\mu ^1g)(\xi _x)=(2\pi )^pg(\theta _x)e^{i\theta _x,\xi _x}𝑑\mu _{}^x(\theta _x).$$ ###### Proof. Compute in local coordinates. $`\mathrm{}`$ Using these two lemmas we have ###### Proposition 7.6. $`_\mu `$ is an algebra isomorphism between $`S(E)`$ with the convolution $`^{}`$-algebra structure inherited from $`C^{}(E,\mu )`$ and $`S(E^{})`$ with the $`^{}`$-algebra structure borrowed from $`C_0(E^{})`$. In the following lemma, which is used in the proof of Proposition 9.1, we list a series of properties of the Fourier transform on $`S(E)`$. ###### Lemma 7.7. If $`f,gS(E)`$, $`\varphi S(E^{})`$ and $`aC^{\mathrm{}}(M)`$, then 1. $`\left[(a\pi )f\right]^\widehat{}=(a\pi )\widehat{f}`$; 2. $`{\displaystyle \frac{\widehat{f}}{q_j}}(\theta _x)=\widehat{{\displaystyle \frac{f}{q_j}}}(\theta _x)+{\displaystyle \frac{ln\mu _e}{q_j}}(x)\widehat{f}(\theta _x)`$; 3. $`{\displaystyle \frac{\widehat{f}}{ϵ_i}}(\theta _x)=i(\xi _if)^\widehat{}(\theta _x)`$; 4. $`{\displaystyle \frac{\stackrel{ˇ}{\varphi }}{\lambda _i}}(\xi _x)=i(w_i\varphi )^\stackrel{ˇ}{}(\xi _x)`$; 5. $`i\theta _k\widehat{f}(\theta _x)=\left({\displaystyle \frac{f}{\lambda _k}}\right)^\widehat{}(\theta _x)`$. Here $`()\widehat{}`$ is the Fourier transform $`_\mu `$ of $`()`$ and $`()\stackrel{ˇ}{}`$ is the inverse Fourier transform $`_\mu ^1`$ of $`()`$. The Schwartz functions do not form the only class of interest to us. ###### Definition 7.8. The Paley–Wiener functions on $`E^{}`$, denoted by $`C_{\text{PW}}^{\mathrm{}}(E^{})`$, consist all functions in $`S(E^{})`$ whose (inverse) Fourier transform is in $`C_c^{\mathrm{}}(E)`$. We use this definition for $`E=A(G)`$; the Poisson algebra $`A^0`$ in Theorems 4.4 and 4.6 is $`C_{\text{PW}}^{\mathrm{}}(A^{}(G))`$. Writing the Poisson bracket and the pointwise product in terms of the Fourier transform, one quickly establishes that $`A^0`$ is indeed a Poisson algebra (cf. section 9). ## 8. Weyl quantization Having constructed the continuous field of $`C^{}`$-algebras called for in Theorem 4.4 in section 6, it remains to define a Poisson algebra $`A^0C_0(A^{}(G))`$ and a map $`𝒬:A^0C^{}(\widehat{G})`$, or equivalently, a family of involutive maps $`𝒬_{\mathrm{}}:A^0C^{}(G)`$ satisfying (4.7). These maps will, in addition, provide the data of Theorem 4.6. We do so by a generalization of Weyl quantization on $`T^{}^n`$; cf. \[La3\]. We pick a positive 1-density $`\mu C^{\mathrm{}}(G_0,|\mathrm{\Omega }|^1(A(G)))`$, with associated left Haar system $`(\lambda ^x)_{xG_0}`$ on $`G`$ (see section 3), left Haar system $`(\widehat{\lambda }^{(\mathrm{},x)})_{(\mathrm{},x)\widehat{G}_0}`$ on $`\widehat{G}`$ (see section 6), and left Haar system $`(\mu ^x)_{xG_0}`$ on $`A(G)`$ (see section 7). We shall work with the concrete $`C^{}`$-algebras $`C^{}(G,\lambda )`$ and $`C^{}(\widehat{G},\widehat{\lambda })`$ rather than with their intrinsically defined versions $`C^{}(G)`$ and $`C^{}(\widehat{G})`$; see section 3. Moreover, since the argument below is the same for the reduced $`C^{}`$-algebras, we will not consider that case explicitly. Now choose some function $`\kappa C^{\mathrm{}}(A(G),)`$ with support in $`V`$ (cf. Proposition 2.6), equaling unity in some smaller tubular neighbourhood of $`G_0`$, as well as satisfying $`\kappa (\xi )=\kappa (\xi )`$ for all $`\xi A(G)`$. ###### Definition 8.1. Let $`G`$ be a Lie groupoid with Lie algebroid $`A(G)`$. We put (8.1) $$A^0=C_{\text{PW}}^{\mathrm{}}(A^{}(G))A_0=C_0(A^{}(G)).$$ For $`\mathrm{}0`$, the Weyl quantization of $`fA^0`$ is the element $`𝒬_{\mathrm{}}^W(f)C_c^{\mathrm{}}(G_1)`$, regarded as a dense subalgebra of $`C^{}(G,\lambda )`$, defined by $`𝒬_{\mathrm{}}^W(f)(\mathrm{Exp}^W(\xi ))`$ $`=`$ $`|\mathrm{}|^p\kappa (\xi )_\mu ^1f(\xi /\mathrm{})\xi V;`$ (8.2) $`𝒬_{\mathrm{}}^W(f)(\gamma )`$ $`=`$ $`0\gamma W.`$ Here the Weyl exponential $`\mathrm{Exp}^W:A(G)G`$ is defined in (2.10). This definition is possible by virtue of Proposition 2.6. (An analogous definition using $`\mathrm{Exp}^L`$ would not satisfy (4.6); such a definition would generalize the Kohn–Nirenberg calculus of pseudodifferential operators rather than the Weyl calculus.) By our choice of $`A^0`$, the operator $`𝒬_{\mathrm{}}^W(f)`$ is independent of $`\kappa `$ for small enough $`\mathrm{}`$ (depending on $`f`$). ###### Proposition 8.2. For each $`fA^0`$, the operator $`𝒬_{\mathrm{}}^W(f)`$ of Definition 8.1 satisfies $`𝒬_{\mathrm{}}^W(f)^{}=𝒬_{\mathrm{}}^W(f^{})`$, and the family $`\{𝒬_{\mathrm{}}^W(f)\}_{\mathrm{}}`$, with $`𝒬_0^W(f)=f`$, is a continuous cross-section of both continuous fields of $`C^{}`$-algebras in Theorem 6.4. ###### Proof. It is immediate from (3.8) and (2.10) that for real-valued $`fA^0`$ the operator $`𝒬_{\mathrm{}}^W(f)`$ is self-adjoint in $`C^{}(G,\lambda )`$; this implies the first claim. Take $`fC_{\text{PW}}^{\mathrm{}}(A^{}(G))`$. The function $`𝒬^W(f)`$ on $`\widehat{G}_1`$ that is defined by $`𝒬^W(f)(0,\xi )`$ $`=`$ $`_\mu ^1f(\xi );`$ $`𝒬^W(f)(\mathrm{},\mathrm{Exp}^W(\xi ))`$ $`=`$ $`\kappa (\xi )_\mu ^1f(\xi /\mathrm{})\xi V;`$ (8.3) $`𝒬^W(f)(\gamma )`$ $`=`$ $`0\gamma W`$ and is in $`C_c^{\mathrm{}}(\widehat{G}_1)`$; cf. Definition 6.1. In other words, $`𝒬^W(f)`$ is an element of $`C^{}(\widehat{G},\widehat{\lambda })`$. For $`\widehat{f}C_c^{\mathrm{}}(\widehat{G})`$, define the restriction maps $`\widehat{\phi }_0(\widehat{f}):\theta `$ $``$ $`_\mu \widehat{f}(0,\theta );`$ (8.4) $`\widehat{\phi }_{\mathrm{}}(\widehat{f}):\gamma `$ $``$ $`\widehat{f}(\mathrm{},\gamma )(\mathrm{}0).`$ Then $`\widehat{\phi }_0`$ is a surjective $`^{}`$-homomorphisms from $`C_c^{\mathrm{}}(\widehat{G}_1,\widehat{\lambda })`$ to $`C_{\text{PW}}^{\mathrm{}}(A^{}(G),\mu )`$, and each $`\widehat{\phi }_{\mathrm{}}`$ is a surjective $`^{}`$-homomorphisms from $`C_c^{\mathrm{}}(\widehat{G}_1,\widehat{\lambda })`$ to $`C_c^{\mathrm{}}(G,|\mathrm{}|^p\lambda )`$. These maps are contractive, and extend by continuity to surjective $`^{}`$-homomorphisms from $`C^{}(\widehat{G},\widehat{\lambda })`$ to $`C_0(A^{}(G))`$ and $`C^{}(G,|\mathrm{}|^p\lambda )`$, respectively. However, in Theorem 4.4 we have $`A_0=C_0(A^{}(G))`$ and $`A_{\mathrm{}}=C^{}(G,\lambda )`$ for $`\mathrm{}0`$. Hence the maps $`\phi _{\mathrm{}}`$ of Definition 4.2 should be taken as $`\phi _0(\widehat{f}):\theta `$ $``$ $`_\mu \widehat{f}(0,\theta );`$ (8.5) $`\phi _{\mathrm{}}(\widehat{f}):\gamma `$ $``$ $`|\mathrm{}|^p\widehat{f}(\mathrm{},\gamma )(\mathrm{}0).`$ These maps extend to surjective $`^{}`$-homomorphisms from $`C^{}(\widehat{G},\widehat{\lambda })`$ to $`C_0(A^{}(G))`$ and $`C^{}(G,\lambda )`$, respectively. It follows that $`A_{\mathrm{}}C^{}(\widehat{G},\widehat{\lambda })/\mathrm{ker}\phi _{\mathrm{}}`$ for all $`\mathrm{}`$, and that $`𝒬_{\mathrm{}}^W(f)=\phi _{\mathrm{}}(𝒬^W(f))`$. The proposition follows. $`\mathrm{}`$ We have now proved Theorem 4.4 up to Dirac’s condition, and turn to Theorem 4.6. We in addition need to choose the tubular neighbourhood $`W`$ of $`G_0G_1`$ so that $`\gamma \gamma ^{}W`$ whenever $`(\gamma ,\gamma ^{})G_2(W\times W)`$. We define a semi-strict deformation quantization of $`A^{}(G)`$ in the following way. For given $`fA^0=C_{\text{PW}}^{\mathrm{}}(A^{}(G))`$, we choose a cutoff function $`\chi _f`$ on $``$ that is 1 in a neighbourhood of 0 and has support well inside the set of values of $`\mathrm{}`$ for which $`𝒬^W(f)`$ as defined in (8.3) is independent of $`\kappa `$. It is clear from the choice of $`A^0`$ that this can be done. For each $`\mathrm{}^{}`$, we then have: * an involution $`f^{_{\mathrm{}}}=\overline{f}`$ (independent of $`\mathrm{}`$); * a semi-norm $`f_{\mathrm{}}=\chi _f(\mathrm{})𝒬_{\mathrm{}}^W(f)_{C^{}(G,\lambda )}`$; * a product $`\times _{\mathrm{}}`$ defined by the condition $$𝒬_{\mathrm{}}^W(f\times _{\mathrm{}}g)=\chi _f\chi _g(\mathrm{})𝒬_{\mathrm{}}^W(f)𝒬_{\mathrm{}}^W(g),$$ where $``$ is the product in $`C^{}(G,\lambda )`$. The existence of the product follows from a detailed but straightforward analysis of the support properties of $`𝒬_{\mathrm{}}^W(f)𝒬_{\mathrm{}}^W(g)`$, leading to the conclusion that, seen as a function on $`G_1`$ for fixed $`\mathrm{}`$, under the stated assumptions its support is contained inside $`W`$; see \[Ra1\]. Hence it can be pulled back using $`\mathrm{Exp}^W`$ and can subsequently be Fourier-transformed so as to yield the desired function $`f\times _{\mathrm{}}g`$. The fact that $`_{\mathrm{}}`$ is indeed an involution is a consequence of the self-adjointness property (4.6) of $`𝒬_{\mathrm{}}^W`$. The same comment applies to the $`C^{}`$-property of the seminorm. In general, these are not norms, since for a given $`f0`$ there may well exist values of $`\mathrm{}`$ such that $`\chi (\mathrm{})𝒬_{\mathrm{}}^W(f)=0`$, seen as a function on $`G(\mathrm{})\widehat{G}`$. The conditions defining a semi-strict deformation quantization, except (4.9), are now trivially satisfied as a consequence of Theorem 4.4 as proved so far. A direct proof of condition (b) in Definition 4.5 is also immediate, based on Theorem 6.4. ## 9. The local structure of the Poisson bracket In this section, taken from \[Ra1\], we express the structure of Lie algebroids $`E`$, as well as the Poisson bracket on $`C^{\mathrm{}}(E^{})`$, in local coordinates. This is of interest in itself, but in the context of quantization it is a key tool for proving Dirac’s condition (4.7) or (4.9). Recall Definition 2.2 of a Lie algebroid. Let $`\{e_1,e_2,\mathrm{},e_p\}`$ be a local frame on $`UM`$ for $`E`$, and let $`(q_1,\mathrm{},q_n,\lambda _1,\mathrm{},\lambda _p)`$ be local coordinates of $`E`$, with the $`q_i`$’s local coordinates for $`M`$ and the $`\lambda _j`$’s local coordinates for the fibers associated to the frame $`\{e_1,e_2,\mathrm{},e_p\}`$. Then, locally, the fact that $`E`$ is a Lie algebroid implies the existence of structure functions $`c_{ijk}`$ and $`a_{ij}`$ in $`C^{\mathrm{}}(U)`$ such that (9.1) $`[e_i,e_j]`$ $`=`$ $`{\displaystyle \underset{k}{}}c_{ijk}e_k;`$ (9.2) $`\rho (e_i)`$ $`=`$ $`{\displaystyle \underset{j}{}}a_{ij}{\displaystyle \frac{}{q_j}}.`$ The Poisson bracket (4.2) - (4.4) of the functions $`\varphi ,\psi C^{\mathrm{}}(E^{})`$ is then locally given by $$\{\varphi ,\psi \}=\underset{i,j}{}\rho (e_i)(q_j)(\frac{\varphi }{ϵ_i}\frac{\psi }{q_j}\frac{\varphi }{q_j}\frac{\psi }{ϵ_i})+\underset{i,j}{}\stackrel{~}{[e_i,e_j]}\frac{\varphi }{ϵ_i}\frac{\psi }{ϵ_j};$$ recall the notation explained below (4.4 ). Taking the value at the point $`(x,\alpha )E_x^{}`$ and using the structure functions of the Lie algebroid, we obtain (9.3) $`\{\varphi ,\psi \}(x,\alpha )`$ $`=`$ $`{\displaystyle \underset{i,j}{}}a_{ij}(x)\left({\displaystyle \frac{\varphi }{ϵ_i}}(x,\alpha ){\displaystyle \frac{\psi }{q_j}}(x,\alpha ){\displaystyle \frac{\varphi }{q_j}}(x,\alpha ){\displaystyle \frac{\psi }{ϵ_i}}(x,\alpha )\right)`$ $`+`$ $`{\displaystyle \underset{i,j,k}{}}c_{ijk}(x)ϵ_k(\alpha ){\displaystyle \frac{\varphi }{ϵ_i}}(x,\alpha ){\displaystyle \frac{\psi }{ϵ_j}}(x,\alpha ).`$ Fix a positive 1-density $`\mu C^{\mathrm{}}(M,|\mathrm{\Omega }|^1(E))`$. We define a Poisson bracket on the dense subalgebra $`S(E)`$ of $`C^{}(E,\mu )`$, by transporting the Poisson bracket of $`C^{\mathrm{}}(E^{})`$ using the Fourier transform $`_\mu `$. That is, (9.4) $$\{f,g\}_\mu =_\mu ^1\left(\{_\mu f,_\mu g\}\right).$$ ###### Proposition 9.1. Eq. (9.4) defines a Poisson bracket on the convolution algebra $`S(E)`$. The Poisson bracket of $`f,gS(E)`$ is explicitly given by (9.5) $`\{f,g\}_\mu (\xi _x)`$ $`=`$ $`i{\displaystyle \underset{i,j}{}}a_{ij}(x)\left(\xi _if{\displaystyle \frac{g}{q_j}}\xi _ig{\displaystyle \frac{f}{q_j}}\right)(\xi _x)`$ $``$ $`i{\displaystyle \underset{i,j}{}}a_{ij}(x){\displaystyle \frac{\mathrm{ln}\mu _e}{q_j}}(x)\left(\xi _ifg\xi _igf\right)(\xi _x)`$ $`+`$ $`i{\displaystyle \underset{i,j,k}{}}c_{ijk}(x){\displaystyle \frac{}{\lambda _k}}\left(\xi _if\xi _jg\right)(\xi _x).`$ ###### Proof. Eq. (9.3) proves that $`S(E^{})`$ is stable under the Poisson bracket $`\{,\}`$, hence (9.4) is well defined. Since the Fourier transform is an algebra isomorphism between $`S(E)`$ and $`S(E^{})`$, it can easily be shown that we have a Poisson algebra structure on $`S(E)`$. Eq. (9.5) is a consequence of Lemma 7.7 and (9.3). $`\mathrm{}`$ We now specialize to the case of relevance to us, where $`E=A(G)`$ is the Lie algebroid of a Lie groupoid $`G`$. We analyze the local structure of $`G`$ in the neighborhood of a fixed point $`x_0G_0G`$ in a convenient parametrization; also cf. \[NWX\]. The map $`r`$ is a submersion at $`x_0G_0G_1`$, hence there exists an open neighborhood $`U`$ of 0 in $`^n`$, an open neighborhood $`V`$ of 0 $`^m`$, and parametrizations $`\psi :U\times VG_1`$ and $`\phi :UG_0`$ such that (9.6) $$\begin{array}{c}1.\psi (0,0)=x_0;\hfill \\ 2.r(\psi (u,v))=\phi (u);\hfill \\ 3.\psi (U\times \{0\})=\psi (U\times V)G_0.\hfill \end{array}$$ It follows from the first two conditions that $`\phi (u)=\psi (u,0)`$. To the parametrization $`\psi `$ of the Lie groupoid $`G`$ one can associate a parametrization $`\theta :U\times ^mA(G)`$ of the neighborhood $`A(G)_{\phi (U)}`$ of the fiber $`A(G)_{x_0}`$ in $`A(G)`$, given by $`\theta (u,v)=(\phi (u),{\displaystyle \frac{\psi }{v}}(u,0)v)`$. For every $`x_0G_0`$ there exists a neighborhood $`\psi (U\times V)`$ of $`x_0`$ in $`G`$ that is diffeomorphic to the neighborhood $`\theta (U\times V)`$ of $`(x_0,0)`$ in $`A(G)`$ by $`\alpha =\psi \theta ^1`$. Moreover, $`\alpha (A(G)_x)G^x`$ for each $`x\phi (U)`$. This result can be formulated in a stronger form, based on the existence of an exponential map for a Lie groupoid; see section 2. Namely, taking $`\alpha `$ equal to the restriction of $`\mathrm{Exp}^L`$ (cf. Definition 2.4) to $`V`$, and $`\alpha _x`$ as the restriction of $`\alpha `$ to $`A(G)_xV`$, one achieves that $`\alpha (A(G)_xV)=G^xW`$ and that $`\alpha _x^{}(0)`$ is the identity of $`A(G)_x`$. The submersion $`\sigma =\phi ^1s\psi :U\times VU`$ is the local expression of the source map $`s`$ in the parametrization $`\psi `$. One has $`\sigma (u,0)=u`$, since $`\phi (\sigma (u,0))=s(\psi (u,0))=\psi (u,0)=\phi (u)`$. Similar expression may be given for the multiplication and the inversion. The following theorem gives a version of the Baker-Campbell-Hausdorff formula for Lie groupoids. This will enter the proof of Dirac’s condition. ###### Theorem 9.2. Let $`G`$ be a Lie groupoid. Then 1. $`(\psi (u,v),\psi (u_1,w))G_2`$ if and only if $`u_1=\sigma (u,v)`$. In that case, their product is of the form $`\psi (u,v)\psi (\sigma (u,v),w)=\psi (u,p(u,v,w))`$, where $`p:U\times V\times VV`$ is a smooth map which has the expansion (9.7) $$p(u,v,w)=v+w+B(u,v,w)+O_3(u,v,w),$$ with $`B`$ bilinear in $`(v,w)`$, and $`O_3(u,v,w)`$ of the order of $`(v,w)^3`$. 2. Let $`(u,v)U\times V`$ be such that $`\psi (u,v)^1\psi (U\times V)`$. Then $`\psi (u,v)^1=\psi (\sigma (u,v),w)`$, where $`w`$ satisfies $`p(u,v,w)=0`$. Moreover, $`w=v+B(u,v,v)+O_3(u,v)`$, with $`O_3(u,v)`$ of the order of $`v^3`$. ###### Proof. We merely sketch the proof. For details cf. \[Ra4\]. 1. Set $`g=\psi (u,v)`$ and $`h=\psi (u_1,w)`$. With $`s(g)=\phi (\sigma (u,v))`$ and $`r(h)=\phi (u_1)`$, we have $`(g,h)G_2`$ if and only if $`u_1=\sigma (u,v)`$. Also, $`r(gh)=\phi (u)`$ implies the existence of a unique $`p(u,v,w)V`$ such that $`\psi (u,v)\psi (\sigma (u,v),w)=\psi (u,p(u,v,w))`$. Hence we obtain a map $`p:U\times V\times VV`$ that satisfies $`p(u,0,w)=w`$ and $`p(u,v,0)=v`$. Using these equations in a Taylor expansion of $`p`$ yields (9.7). 2. Similar to 1. $`\mathrm{}`$ We can now give explicit formulae for the structure functions of the Lie algebroid $`A(G)`$. First, remark that the family $`\{e_1,e_2,\mathrm{},e_m\}`$ defined by $`e_i(\phi (u))=\theta (u,f_i)`$, $`i=\overline{1,m}`$, where $`\{f_1,f_2,\mathrm{},f_m\}`$ is the canonical basis of $`^m`$, is a frame of $`A(G)`$ on $`\phi (U)`$. Also, recall that the anchor of $`A(G)`$ is given by $`\rho =Ts`$ (cf. Definition 2.3.2), that the local coordinate functions of $`G_0`$ are $`q_j=pr_j\phi ^1`$, and that $`B_1,\mathrm{},B_m`$ are the coordinates of $`B:U\times V\times VV`$ in the base $`\{f_1,f_2,\mathrm{},f_m\}`$ of $`^m`$. Finally, recall that $`a_{ij}=\rho (e_i)(q_j)`$, and that the $`c_{ijk}`$ are given by $`[e_i,e_j]={\displaystyle c_{ijk}e_k}`$. ###### Proposition 9.3. For each $`uU`$, the structure functions of the Lie algebroid $`A(G)`$ are given by $`a_{ij}(\phi (u))={\displaystyle \frac{\sigma _j}{v_i}}(u,0)`$ and $`c_{ijk}(\phi (u))=B_k(u,f_i,f_j)B_k(u,f_j,f_i)`$. ###### Proof. Direct calculations. See \[Ra1\], \[Ra4\] for details. $`\mathrm{}`$ We now return to the Poisson bracket on $`A(G)`$. Recall Proposition 9.1, and put $`E=A(G)`$. The next proposition gives a local expression of the Poisson bracket on $`S(A(G))`$. Let $`\lambda C^{\mathrm{}}(G_0,|\mathrm{\Omega }|^1(A(G)))`$ be a positive 1-density and $`\mathrm{\Lambda }`$ the local expression in the parametrization $`\theta `$ of the function associated to $`\lambda `$ in the local frame $`\{e_1,e_2\mathrm{},e_m\}`$. ###### Proposition 9.4. For $`f,gS(A(G))`$, let $`h=\{f,g\}_\lambda `$ be the Poisson bracket of $`f`$ and $`g`$ given by (9.5). Denote the local expressions of $`f,g,h`$ in the parametrization $`\theta `$ by $`F,G,H`$, respectively. Then $`H(u,v)=i{\displaystyle F(u,w)G(u,vw)}{\displaystyle \underset{j}{}}\left[B_j(u,w,f_j)B_j(u,f_j,w)\right]\mathrm{\Lambda }(u)dw`$ $`+i{\displaystyle }{\displaystyle \underset{k}{}}\left[B_k(u,w,v)B_k(u,v,w)\right]F(u,w){\displaystyle \frac{G}{v_k}}(u,vw)\mathrm{\Lambda }(u)dw`$ $`i{\displaystyle \underset{i,j}{}}{\displaystyle \frac{\sigma _j}{v_i}}(u,0){\displaystyle \frac{\mathrm{\Lambda }}{u_j}}(u){\displaystyle w_i\left[F(u,w)G(u,vw)G(u,w)F(u,vw)\right]𝑑w}`$ $`i{\displaystyle \underset{i,j}{}}{\displaystyle \frac{\sigma _j}{v_i}}(u,0){\displaystyle w_i\left[F(u,w)\frac{G}{u_j}(u,vw)G(u,w)\frac{F}{u_j}(u,vw)\right]\mathrm{\Lambda }(u)𝑑w}.`$ ###### Proof. Replacing the structure functions in (9.5) by their expressions calculated in Proposition 9.3, we can directly identify the first two lines of (9.5) with the last two in the formula of $`H(u,v)`$. Denote the expression in the last line of (9.5) by $`l`$. We then have $`l(\theta (u,v))`$ $`=`$ $`{\displaystyle \underset{i,j,k}{}}c_{ijk}(\phi (u)){\displaystyle \frac{}{v_k}}\left({\displaystyle }w_iF(u,w)(v_jw_j)G(u,vw)\mathrm{\Lambda }(u)dw\right)`$ $`=`$ $`{\displaystyle \underset{i,j}{}}{\displaystyle (B_j(u,f_j,w_if_i)B_j(u,w_if_i,f_j))F(u,w)G(u,vw)\mathrm{\Lambda }(u)𝑑w}`$ $`+`$ $`{\displaystyle \underset{i,j,k}{}}{\displaystyle [B_k(u,(v_jw_j)f_j,w_if_i)B_k(u,w_if_i,(v_jw_j)f_j)]}`$ $``$ $`F(u,w){\displaystyle \frac{G}{v_k}}(u,vw)\mathrm{\Lambda }(u)dw`$ $`=`$ $`{\displaystyle \underset{j}{}}{\displaystyle (B_j(u,f_j,w)B_j(u,w,f_j))F(u,w)G(u,vw)\mathrm{\Lambda }(u)𝑑w}`$ $`+`$ $`{\displaystyle \underset{k}{}}{\displaystyle (B_k(u,vw,w)B_k(u,w,vw))F(u,w)\frac{G}{v_k}(u,vw)\mathrm{\Lambda }(u)𝑑w}.`$ By bilinearity one has $`B_k(u,w,vw)B_k(u,vw,w)=B_k(u,w,v)B_k(u,v,w)`$; this finishes the proof. $`\mathrm{}`$ ## 10. Proof of Dirac’s condition We now use the results of the preceding section to prove (4.7) and (4.9) in the case at hand. We start with two lemmas. Let $`\psi `$ be a parametrization of $`G`$ as in section 9, and denote the function associated to $`\lambda `$ in the local frame of $`TG`$ generated by $`\psi `$ by $`\lambda _0`$. Put $`\mathrm{\Lambda }=\lambda _0\psi `$. ###### Lemma 10.1. One has $`\lambda _0(\gamma )={\displaystyle \frac{\lambda _0(s(\gamma ))}{|detJ_{L_\gamma }(s(\gamma ))|}}`$ for every $`\gamma `$ in the image of the parametrization $`\psi `$. ###### Proof. The left invariance of $`\lambda `$ means $`\lambda (\gamma )=\gamma \lambda (s(\gamma ))`$. It only remains to write the action of $`G`$ on $`|\mathrm{\Omega }|^1(\mathrm{ker}Tr)`$ in the local parametrization $`\psi `$. We leave this to the reader. $`\mathrm{}`$ ###### Lemma 10.2. Let $`\mu _{u,v}(t)=\lambda _0(\psi (u,tv))`$. Then $`\mu _{u,v}^{}(0)=\underset{i,j}{}{\displaystyle \frac{\mathrm{\Lambda }}{u_i}}(u,0){\displaystyle \frac{\sigma _i}{v_j}}(u,0)v_j[B_1(u,v,f_1)++\mathrm{}+B_m(u,v,f_m)]\mathrm{\Lambda }(u,0).`$ ###### Proof. By Lemma 10.1, $`\mu _{u,v}(t)={\displaystyle \frac{\lambda _0(\psi (\sigma (u,tv),0))}{|detJ_{L_{\psi (u,tv)}}(\psi (\sigma (u,tv),0))|}}`$. Write $`a_{u,v}(t)`$ $`=`$ $`\lambda _0(\psi (\sigma (u,tv),0));`$ $`b_{u,v}(t)`$ $`=`$ $`|detJ_{L_{\psi (u,tv)}}(\psi (\sigma (u,tv),0))|.`$ We have $`a_{u,v}=\mathrm{\Lambda }(\sigma (u,tv),0)`$ and $`a_{u,v}^{}(0)=\underset{i,j}{}{\displaystyle \frac{\mathrm{\Lambda }}{u_i}}(u,0){\displaystyle \frac{\sigma _i}{v_j}}(u,0)v_j`$. The local expression of the multiplication map $`L_{\psi (u,tv)}:G^{\phi (\sigma (u,tv))}G^{\phi (u)}`$ is given by $`p(u,tv,)`$ and this shows, after some calculation, $$b_{u,v}(t)=\left|\begin{array}{cccc}1+tB_1(u,v,f_1)& tB_1(u,v,f_2)& \mathrm{}& tB_1(u,v,f_m)\\ tB_2(u,v,f_1)& 1+tB_2(u,v,f_2)& \mathrm{}& tB_2(u,v,f_m)\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ tB_m(u,v,f_1)& tB_m(u,v,f_2)& \mathrm{}& 1+tB_m(u,v,f_m)\end{array}\right|.$$ Taking the derivative we obtain $`b_{u,v}^{}(0)=B_1(u,v,f_1)+B_2(u,v,f_2)+\mathrm{}+B_m(u,v,f_m)`$, and to finish the proof remark that $`a_{u,v}(0)=\mathrm{\Lambda }(u,0)`$, $`b_{u,v}(0)=1`$. $`\mathrm{}`$ Now let $`(\gamma ,\mathrm{})\widehat{G}`$, with $`\mathrm{}0`$. The smooth structure in a neighborhood of $`(\gamma ,\mathrm{})`$ is given by a parametrization of the form $`\widehat{\psi }=\psi \times id:U\times \widehat{G}`$, where $`\psi :UG`$ is a parametrization of $`G`$ in a neighborhood of $`\gamma `$ and $`U`$ is an open subset of $`^{n+m}`$. Write $`\alpha =\widehat{\psi }^1:\widehat{\psi }(U\times )U\times `$, and define an element of $`T_{(\gamma ,\mathrm{})}\widehat{G}`$ by $$\frac{}{\mathrm{}}\widehat{f}|_{(\gamma ,\mathrm{})}=\frac{\widehat{f}}{\alpha _{n+m+1}}(\gamma ,\mathrm{}).$$ Here $`\widehat{f}C^{\mathrm{}}(\widehat{G})`$. This definition is independent of the parametrization $`\psi `$ of $`G`$. Thus we obtain a vector field $`{\displaystyle \frac{}{\mathrm{}}}C^{\mathrm{}}(\widehat{G},T\widehat{G})`$. The following result is the key lemma towards the proof of Dirac’s condition. ###### Lemma 10.3. Let $`\widehat{f},\widehat{g}C_c^{\mathrm{}}(\widehat{G},\widehat{\lambda })`$ (seen as a convolution algebra), and write $`\widehat{f}_0=\widehat{f}_{|_{\widehat{G}(0)}}`$, $`\widehat{g}_0=\widehat{g}_{|_{\widehat{G}(0)}}`$ for their the restrictions to the Lie algebroid $`A(G)`$. Then $$\frac{}{\mathrm{}}[\widehat{f},\widehat{g}]_{|_{\mathrm{}=0}}=i\{\widehat{f}_0,\widehat{g}_0\}_\mu .$$ ###### Proof. The proof is only sketched; for details cf. \[Ra1, Ra3\]. Using a partition of unity, one may assume that $`\widehat{f}`$ and $`\widehat{g}`$ have their support contained in the image of the parametrization $`\widehat{\psi }`$. Let $`\widehat{k}`$ be the convolution of $`\widehat{f}`$ and $`\widehat{g}`$, and let $`K`$ be the expression of $`\widehat{k}`$ in the parametrization $`\widehat{\psi }`$. After some calculation and the use of (9.7) we obtain $`K(u,v,\mathrm{})`$ $`=`$ $`{\displaystyle }F(u,w,\mathrm{})G(\sigma (u,\mathrm{}w),vw+\mathrm{}B(u,w,w)`$ $``$ $`\mathrm{}B(\sigma (u,\mathrm{}w),w,v)+O(\mathrm{}^2),\mathrm{})\mu _{u,w}(\mathrm{})dw.`$ This leads to an expression for the commutator $`\widehat{f}\widehat{g}\widehat{g}\widehat{f}`$. We then differentiate $`K`$ with respect to $`\mathrm{}`$ at $`\mathrm{}=0`$, use Lemma 10.2 and some changes of variables, and eventually recover the local expression of the Poisson bracket as given in Proposition 9.4. $`\mathrm{}`$ We are now in a position to prove Dirac’s condition (4.9) and (4.7). We only deal with the former; the latter is proved in exactly the same way. The essence of the proof is that Dirac’s condition follows from the continuity condition (b) in Definition 4.5; in the context of Definition 4.3 this condition is satisfied as a consequence of the definition of a continuous field of $`C^{}`$-algebras. Define $`\widehat{V}\widehat{A(G)}`$ as $`V`$ in Proposition 2.6, but now for the tangent groupoid $`\widehat{G}`$ rather than $`G`$. For $`f,gC_c^{\mathrm{}}(A(G))`$, let $`\phi :\widehat{V}`$ be defined by $`\phi (x,X,0)`$ $`=`$ $`0;`$ $`\phi (x,X,\mathrm{})`$ $`=`$ $`{\displaystyle \frac{(f\times _{\mathrm{}}g)(x,X)(g\times _{\mathrm{}}f)(x,X)}{\mathrm{}}}i\{f,g\}(x,X)\mathrm{}0.`$ Lemma 10.3 shows that $`\phi C^{\mathrm{}}(\widehat{V})`$, and arguments like those in the discussion of the product $`f\times _{\mathrm{}}g`$ in section 8 show that supp$`\phi `$ is compact in $`\widehat{V}`$. Write $`\widehat{\phi }=\phi \widehat{\alpha }^1C_c^{\mathrm{}}(\widehat{\alpha }(\widehat{V}))`$. One may regard $`\phi `$ as an element of $`C_c^{\mathrm{}}(A(G)\times )`$ and $`\widehat{\phi }`$ as an element of $`C_c^{\mathrm{}}(\widehat{G})`$ by extending these functions by 0 outside $`\widehat{V}`$ and $`\widehat{\alpha }(\widehat{V})`$, respectively. Remark that for every $`\mathrm{}`$ one has function $`\phi (,\mathrm{})A^0`$. Applying Theorem 6.4 and recalling the notation (6.6), where we look at the $`G(\mathrm{})`$ as subgroupoids of $`\widehat{G}`$, one then has $$\underset{\mathrm{}0}{lim}\phi (,\mathrm{})_{\mathrm{}}=\underset{\mathrm{}0}{lim}\widehat{\phi }_{|_{G(\mathrm{})}}_{C^{}(G(\mathrm{}),\lambda )}=\widehat{\phi }_{|_{G(0)}}_{C^{}(G(0),\mu )}=0.$$ This finishes the proof of Dirac’s condition. Having giving the other parts of the proof in section 8, this finishes the proof of Theorems 4.4 and 4.6. ## 11. Examples and comments Since a vast number of interesting $`C^{}`$-algebras are defined by some Lie groupoid, one obtains a large reservoir of examples of our theorems, of which in this section we merely scratch the surface. ###### Example 11.1. Quantization of the Lie-Poisson structure on a dual Lie algebra This example was already introduced in the Introduction. A Lie group is a Lie groupoid with $`G_0=e`$. A left-invariant Haar measure on $`G`$ provides a left Haar system; the ensuing convolution algebra $`C^{}(G)`$ is the usual group $`C^{}`$-algebra. The Poisson structure on the dual Lie algebra $`A^{}(G)`$ is the well-known Lie-Poisson structure \[MR, Va\]. No connection is needed to define the exponential map, and one has (11.1) $$\mathrm{Exp}^L(X)=\mathrm{Exp}^W(X)=\mathrm{Exp}(X),$$ where $`Xg`$ and $`\mathrm{Exp}:gG`$ is the usual exponential map. We obtain a semi-strict or strict deformation quantization of $`A^{}(G)`$ for any Lie group $`G`$, but the situation is particularly favourable when $`G`$ is exponential (in that $`\mathrm{Exp}`$ is a diffeomorphism). One may then omit the cutoff functions $`\kappa `$ and $`\chi `$ in section 8, and $`𝒬_{\mathrm{}}^W(f)`$ in (8.2) is simply given by (11.2) $$𝒬_{\mathrm{}}^W(f):\mathrm{Exp}(X)(2\pi \mathrm{})^n_g^{}e^{i\theta ,X/\mathrm{}}f(\theta )d^n\theta .$$ This is precisely Rieffel’s prescription \[Ri4\]; his assumption that $`G`$ be nilpotent is now seen to be unnecessary in order to obtain a strict deformation quantization. Moreover, in the exponential case our semi-strict quantization is easily shown to be strict, since the semi-norms $`_{\mathrm{}}`$ are now norms. More generally, if a Lie groupoid is diffeomorphic to its Lie algebroid then the semi-strict deformation quantization we have given is always strict; see \[Ra1\]. ###### Example 11.2. Transformation group $`C^{}`$-algebras Let a Lie group $`H`$ act smoothly on a manifold $`M`$. The transformation groupoid (or action groupoid) $`G=H\times M`$ is defined by the operations $`s(x,m)=x^1m`$ and $`r(x,m)=m`$, so that $`((x,m),(y,m^{}))G_2`$ when $`m^{}=x^1m`$. In that case, $`(x,m)(y,x^1m)=(xy,m)`$. The inclusion is $`m(e,m)`$, and the inverse is $`(x,m)^1=(x^1,x^1m)`$. Each left-invariant Haar measure $`dx`$ on $`H`$ leads to a left Haar system on $`G`$. The corresponding groupoid $`C^{}`$-algebra is the usual transformation group $`C^{}`$-algebra $`C^{}(G)=C^{}(H,M)`$, cf. \[Re1\]. The Lie algebroid $`\times M`$ (where $``$ is the Lie algebra of $`H`$) is a trivial bundle over $`M`$, with anchor $`\rho (X,m)=\xi _X(m)`$ (the fundamental vector field on $`M`$ defined by $`X`$). Identifying sections of $`\times M`$ with $``$-valued functions $`X()`$ on $`M`$, the Lie bracket on $`C^{\mathrm{}}(M,\times M)`$ is (11.3) $$[X,Y]_{\times M}(m)=[X(m),Y(m)]_{}+\xi _YX(m)\xi _XY(m).$$ The associated Poisson bracket coincides with the semi-direct product bracket defined in \[KM\]. The trivial connection on $`\times MM`$ yields the exponential maps (11.4) $`\mathrm{Exp}^L(X,m)`$ $`=`$ $`(\mathrm{Exp}(X),m);`$ (11.5) $`\mathrm{Exp}^W(X,m)`$ $`=`$ $`(\mathrm{Exp}(X),\mathrm{Exp}(\frac{1}{2}X)m).`$ The cutoff function $`\kappa `$ in (8.2) is independent of $`m`$, and coincides with the function appearing in Example 11.1. For small enough $`\mathrm{}`$, a function $`fC_{\text{PW}}^{\mathrm{}}(^{}\times M)`$ is then quantized by (11.6) $$𝒬_{\mathrm{}}^W(f):(\mathrm{Exp}(X),m)(2\pi \mathrm{})^n_{^{}}e^{i\theta ,X/\mathrm{}}f(\theta ,\mathrm{Exp}(\frac{1}{2}X)m)d^n\theta .$$ When $`H=^n`$ and $`M`$ has a $`H`$-invariant measure, the map $`f𝒬_{\mathrm{}}^W(f)`$ is equivalent to the deformation quantization considered by Rieffel \[Ri2\], who already proved that it is strict. Finally, when $`H`$ is exponential we are in the situation discussed at the end of the previous example, so that our semi-strict deformation quantization is strict. ###### Example 11.3. Weyl quantization on $`T^{}^n`$ As already remarked in the Introduction, the tangent bundle $`TM`$ of a manifold $`M`$ is the Lie algebroid of the pair groupoid $`G=M\times M`$. Hence our formalism produces a (semi) strict deformation quantization of the cotangent bundle $`T^{}M`$ of any manifold with linear connection. In order to understand Weyl quantization in the light of our formalism, we must change some signs: we add minus signs in front of the right-hand sides of (4.3) and (4.4), and change $`f(\xi /\mathrm{})`$ in (8.2) to $`f(\xi /\mathrm{})`$ (cf. \[La3\]) for the rationale behind these signs). For $`M=^n`$ with flat connection (8.2) then simply becomes (11.7) $$𝒬_{\mathrm{}}^W(f)\mathrm{\Psi }(x)=(2\pi \mathrm{})^n_{T^{}^n}e^{ip(xy)/\mathrm{}}f(p,\frac{1}{2}(x+y))\mathrm{\Psi }(y)d^npd^ny,$$ where $`\mathrm{\Psi }L^2(^n)`$; we here use the fact that $`C^{}(^n\times ^n)=B_0(L^2(^n))`$. This is precisely Weyl’s original prescription, now written in a form that emphasizes its geometric origin. For $`\frac{1}{2}(x+y)`$ is the midpoint of the geodesic from $`x`$ to $`y`$, and $`xy`$ is a tangent vector to the geodesic at this midpoint. Looked at in this way, Weyl quantization may easily be generalized to manifolds with connection \[La1, La3\], and forms a special case of what we have called Weyl quantization for general Lie groupoids in this paper. This answers Rieffel’s Question 20 in \[Ri5\]. The fact that this quantization is strict, and in particular satisfies (4.7), had already been proved by Rieffel \[Ri6\]. The associated continuous field of $`C^{}`$-algebras has fibers $`A_0=C_0(T^{}R^n)`$ and $`A_{\mathrm{}}=B_0(L^2(^n))`$ for $`\mathrm{}0`$. The $`C^{}`$-algebra $`C`$ in Definition 4.2 is $`C^{}(H_n)`$, the group algebra of the simply connected Heisenberg group on $`^n`$; also see \[ENN1\]. This is indeed the $`C^{}`$-algebra of the tangent groupoid of $`^n`$. When $`M`$ be Riemannian and exponential (in that the exponential map is a diffeomorphism between the tangent bundle $`TM`$ and the pair groupoid $`M\times M`$), our semi-strict deformation quantization is strict \[Ra1\]. As is well known, for $`M=^n`$ with flat metric the product $`\times _{\mathrm{}}`$ is then equal to the Moyal product \[Mo, BFFLS\]. For a detailed analysis of this situation in the setting of the present paper see \[CCFGRV\]. We close with two remarks. ###### Remark 11.4. There are clear connections between index theory (in the sense of Atiyah–Singer), $`C^{}`$-algebraic K-theory, and quantization; see, e.g., \[Fe, Hi, ENN2, Ro\]. Moreover, Connes \[Co2\] discovered a beautiful proof of the index theorem that is based on the tangent groupoid of a manifold. Let $`M`$ be a compact Riemannian manifold, with associated cosphere bundle $`S^{}M`$ and $`C^{}`$-algebra of bounded classical pseudodifferential operators $`\mathrm{\Psi }^0(M)`$. The interpretation of index theory in terms of the K-theory of $`C^{}`$-algebras comes from the short exact sequence of $`C^{}`$-algebras (11.8) $$0B_0(L^2(M))\mathrm{\Psi }^0(M)\stackrel{\sigma }{}C(S^{}M)0;$$ here $`\sigma `$ is the symbol map. Any short exact sequence $`0JAB0`$ leads to a connecting map $`:K^1(B)K_0(J)`$, so that for (11.8) one obtains $$:K_1(C(S^{}M))=K^1(S^{}(M))K_0(B_0)=Z.$$ The analytic index of an elliptic pseudodifferential operator $`P\mathrm{\Psi }^0(M)`$ is given by $`([\sigma (P)])Z`$, where, for $`fC(X)`$, $`[f]`$ is the class in $`K^1(X)`$ defined by $`f`$. One may prove the Atiyah–Singer index theorem from the existence of a strict deformation quantization $`𝒬_{\mathrm{}}:C_0(T^{}M)B_0(L^2(M))`$; cf. \[Hi\]. Seen in the light of Lie groupoids and quantization, its is obvious that the above data may be generalized \[MP\] by replacing the pair groupoid on $`M`$ by an arbitrary Lie groupoid $`G`$.Then $`B_0(L^2(M))`$ is replaced by $`C^{}(G)`$, the cosphere bundle $`S^{}M`$ is replaced by a cosphere bundle $`S^{}A(G)`$ in $`A^{}(G)`$, and the appropriate generalization $`\mathrm{\Psi }^0(G)`$ (with some abuse of notation) of $`\mathrm{\Psi }^0(M)`$ has recently been defined as well \[NWX, MP\]. Thus (11.8) is generalized to (11.9) $$0C^{}(G)\mathrm{\Psi }^0(M)\stackrel{\sigma }{}C_0(S^{}A(G))0;$$ see \[MP\]. On may therefore expect that our strict deformation quantization of $`A^{}(G)`$ eventually leads to an index theorem, where the index now takes values in $`K_0(C^{}(G))`$. ###### Remark 11.5. There is no analogue of Lie’s Third Theorem for Lie algebroids. Examples given in \[AM\] show the existence of Lie algebroids that are not associated to any Lie groupoid. For an arbitrary Lie algebroid $`A(G)`$, Pradines proved (cf. \[Pr4\]) that one can merely construct a local Lie groupoid $`G`$ which has $`A(G)`$ as its Lie algebroid. In this paper, we always assume that we have a given Lie groupoid $`G`$ with Lie algebroid $`A(G)`$. In order to use the ideas from this paper to quantize all Lie algebroids, we would need to extend two constructions, the $`C^{}`$-algebra and the tangent groupoid which can be associated to every Lie groupoid, to the context of local Lie groupoids. The construction of the tangent groupoid has been extended to local Lie groupoids in the context of pseudo-differential operators by Nistor, Weinstein and Xu (cf. \[NWX\]). The missing ingredient in order to extend our results to arbitrary Lie algebroids is therefore the construction of the $`C^{}`$-algebra of a local Lie groupoid. Acknowledgement. B.R. is greatly indebted to Jean Renault for many helpful discussions.
warning/0001/math0001168.html
ar5iv
text
# Hall-Littlewood vertex operators and generalized Kostka polynomials ## 1. Introduction Kostka-Folkes polynomials may be considered as coefficients of the formal power series representing the character of certain graded $`GL(n)`$-modules. These $`GL(n)`$-modules are defined by twisting the coordinate ring of the nullcone by a suitable line bundle and the definition may be generalized by twisting the coordinate ring of any nilpotent conjugacy closure in $`gl(n)`$ by a suitable vector bundle . The resulting polynomials have been called generalized Kostka polynomials. Jing defined a vertex operator that generates the Hall-Littlewood symmetric function $`Q[X;q]`$ , thereby giving an elegant symmetric function recursion for the Kostka-Foulkes polynomials. Garsia used a variant of Jing’s vertex operator to derive various new formulas for the Kostka-Foulkes polynomials. Our point of departure was the observation that the Hall-Littlewood vertex operators can be used to obtain formulas for generalized Kostka polynomials. Our treatment uses Garsia’s plethystic type formulas. One striking fact is that the $`[q,q^1]`$-linear span of $`n`$-fold compositions of components of the Hall-Littlewood vertex operators, is isomorphic to $`K_{G\times ^{}}(𝒩)`$, the $`GL(n)\times ^{}`$-equivariant $`K`$-theory of the nullcone. Under this isomorphism, an $`n`$-fold composite operator is sent to the class of the Euler characteristic of a twisted module. This fact has a generalization for all the nilpotent conjugacy class closures in $`gl(n)`$. These Grothendieck groups were studied in . We derive many explicit relations among the vertex operators, most of which can be in interpreted as relations in the Grothendieck groups which arise from certain Koszul complexes. This allows for more explicit proofs of some basis theorems for these Grothendieck groups that were proved in using geometric arguments. There is a particularly well-behaved subfamily of the generalized Kostka polynomials, namely, those that are indexed by a sequence of rectangular partitions. For this subfamily almost all of the formulas for Kostka-Foulkes polynomials have generalizations. There are combinatorial formulas involving Littlewood-Richardson tableaux , rigged configurations , and inhomogeneous paths with energy function . ## 2. Hopf algebra of symmetric functions and plethysm This section contains standard background material on symmetric functions which can be found in . The possible exception to this is the definition of the plethystic notation used here. Let $`\mathrm{\Lambda }=\mathrm{\Lambda }_𝔽`$ be the algebra of symmetric functions over a field $`𝔽`$ of characteristic zero. $`\mathrm{\Lambda }`$ may be defined as the polynomial algebra $`𝔽[p_1,p_2,\mathrm{}]`$ where the $`p_k`$ are commuting algebraically independent variables. Let $`𝒫`$ be the set of partitions. For $`\lambda =(\lambda _1,\mathrm{},\lambda _k)𝒫`$, write $`|\lambda |=_{i=1}^k\lambda _i`$ and let $`p_\lambda =p_{\lambda _1}\mathrm{}p_{\lambda _k}`$ denote the power symmetric function. Declaring $`p_k`$ to have degree $`k`$, $`\mathrm{\Lambda }`$ is endowed with a grading $`\mathrm{\Lambda }=_{n0}\mathrm{\Lambda }^n`$ where $`\mathrm{\Lambda }^n`$ is the homogeneous component of degree $`n`$. $`\mathrm{\Lambda }^n`$ has $`𝔽`$-basis $`\{p_\lambda \lambda 𝒫,|\lambda |=n\}`$. $`\mathrm{\Lambda }`$ may be realized by symmetric series. Let $`X=(x_1,x_2,\mathrm{})`$ be a countable set of commuting indeterminates. Denote by $`𝔽[[X]]`$ the $`𝔽`$-algebra of formal power series in the $`x_i`$ where each $`x_i`$ has degree 1, and $`𝔽_b[[X]]𝔽[[X]]`$ the subalgebra consisting of series whose monomials are of bounded degree. Let $`\mathrm{\Lambda }^X𝔽_b[[X]]`$ denote the $`𝔽`$-subalgebra consisting of the series that are symmetric in the variables $`x_i`$. By the universal mapping property of a polynomial algebra, there is a $`𝔽`$-algebra homomorphism $`\mathrm{\Lambda }\mathrm{\Lambda }^X`$ given by $`p_kx_1^k+x_2^k+\mathrm{}`$. This map is in fact a graded $`𝔽`$-algebra isomorphism. There is a scalar product on $`\mathrm{\Lambda }`$ for which the power symmetric functions are an orthogonal basis: $$p_\lambda ,p_\mu =\delta _{\lambda \mu }z_\lambda $$ where $`\delta _{\lambda \mu }`$ is the Kronecker delta, $`z_\lambda =_in_i(\lambda )!i^{n_i(\lambda )}`$, and $`n_i(\lambda )`$ is the number of parts of size $`i`$ in the partition $`\lambda `$. We now define notation for plethystic substitution. Let $`q𝔽`$ be a distinguished element that is transcendental over $``$. Fix an element $`E𝔽_b[[X]]`$. Define $`p_k[E]𝔽_b[[X]]`$ to be the series obtained from $`E`$ by replacing $`q`$ by $`q^k`$ and $`x_i`$ by $`x_i^k`$ for all $`i1`$. By the universal mapping property of a polynomial algebra, there is a unique $`𝔽`$-algebra homomorphism $`\mathrm{\Lambda }𝔽_b[[X]]`$ such that $`p_kp_k[E]`$. The image of $`P\mathrm{\Lambda }`$ under this map is denoted $`P[E]`$. Switching viewpoints and considering $`p_k[E]`$ for various $`E`$, we see directly from the definitions that $`Ep_k[E]`$ is a $``$-algebra homomorphism $`𝔽_b[[X]]𝔽_b[[X]]`$ and $`\mathrm{\Lambda }^X\mathrm{\Lambda }^X`$ given by $`p_k[E\pm E^{}]=p_k[E]\pm p_k[E^{}]`$ and $`p_k[EE^{}]=p_k[E]p_k[E^{}]`$ for all $`E,E^{}𝔽_b[[X]]`$. It is not an $`𝔽`$-algebra homomorphism since it changes the scalar $`q`$ to $`q^k`$. By abuse of notation the variable $`X`$ will also be used to represent the formal sum $`x_1+x_2+x_3+\mathrm{}`$. By definition $$p_k[X]=p_k[x_1+x_2+x_3+\mathrm{}]=x_1^k+x_2^k+x_3^k+\mathrm{}.$$ ###### Example 1. $`e_2=p_1^2/2p_2/2\mathrm{\Lambda }`$. For the expression $`E=x_1+x_2`$ we have $$e_2[x_1+x_2]=(x_1+x_2)^2/2(x_1^2+x_2^2)/2=x_1x_2$$ If $`E=\frac{1}{1q}`$ then $`p_1[\frac{1}{1q}]=\frac{1}{1q}`$ and $`p_2[\frac{1}{1q}]=\frac{1}{1q^2}`$, so that $$e_2\left[\frac{1}{1q}\right]=\frac{1}{2(1q)^2}\frac{1}{2(1q^2)}=\frac{q}{(1q)(1q^2)}$$ Recall that each $`E𝔽_b[[X]]`$ defines an $`𝔽`$-algebra homomorphism $`\mathrm{\Lambda }𝔽_b[[X]]`$ such that $`PP[E]`$. 1. If $`E=X`$ then $`PP[X]`$ yields the above isomorphism $`\mathrm{\Lambda }\mathrm{\Lambda }^X`$. 2. If $`E=0`$ then the $`𝔽`$-algebra homomorphism $`ϵ:\mathrm{\Lambda }𝔽`$ given by $`PP[0]`$ is the counit for $`\mathrm{\Lambda }`$; it selects the constant term. 3. If $`E=x_1+x_2+\mathrm{}+x_m`$ then $`PP[x_1+x_2+\mathrm{}+x_m]`$ is the $`𝔽`$-algebra epimorphism $`\mathrm{\Lambda }𝔽[x_1,\mathrm{},x_m]^{S_m}`$ onto the $`𝔽`$-algebra of symmetric polynomials in $`x_1`$ through $`x_m`$ where $`S_m`$ is the symmetric group on $`m`$ letters. 4. If $`E=X/(q1)`$ then the map $`P[X]P[X/(q1)]`$ is an $`𝔽`$-algebra isomorphism $`\mathrm{\Lambda }\mathrm{\Lambda }`$ with inverse map $`P[X]P[X(q1)]`$. In particular if $`\{f_\lambda \}`$ is a basis of $`\mathrm{\Lambda }`$ then so are $`\{f_\lambda [X(q1)]\}`$ and $`\{f_\lambda [X/(q1)]\}`$. We now discuss the Hopf algebra structure on $`\mathrm{\Lambda }`$. Let $`Y=y_1+y_2+\mathrm{}`$ where the $`y_i`$ are another countable set of commuting indeterminates. Let $`\mathrm{\Lambda }^{X;Y}`$ denote the $`𝔽`$-subalgebra of $`𝔽_b[[X,Y]]`$ consisting of the series that are symmetric in both the $`x_i`$ and the $`y_j`$. Then there is an isomorphism $`\mathrm{\Lambda }_𝔽\mathrm{\Lambda }\mathrm{\Lambda }^{X;Y}`$ given by $`fgf[X]g[Y]`$ for $`f,g\mathrm{\Lambda }`$. Letting $`E=X+Y`$, the map $`PP[X+Y]`$ defines an $`𝔽`$-algebra homomorphism $`\mathrm{\Delta }:\mathrm{\Lambda }\mathrm{\Lambda }\mathrm{\Lambda }`$. $`\mathrm{\Delta }`$ is the comultiplication map for $`\mathrm{\Lambda }`$. The maps $`\mathrm{\Delta }`$ and $`ϵ`$ give $`\mathrm{\Lambda }`$ the structure of a coassociative cocommutative Hopf algebra. There is a scalar product on $`\mathrm{\Lambda }\mathrm{\Lambda }`$ defined by $$f_1[X]g_1[Y],f_2[X]g_2[Y]_{\mathrm{\Lambda }^X\mathrm{\Lambda }^Y}=f_1,f_2g_1,g_2$$ for all $`f_i,g_i\mathrm{\Lambda }`$ for $`i=1,2`$ and this relation is then extended by linearity. With respect to this scalar product on $`\mathrm{\Lambda }\mathrm{\Lambda }`$, the map $`\mathrm{\Delta }`$ is adjoint to the multiplication map $`\mathrm{\Lambda }\mathrm{\Lambda }\mathrm{\Lambda }`$ given by $`fgfg`$. That is, (1) $$f,gh=f[X+Y],g[X]h[Y]$$ for all $`f,g,h\mathrm{\Lambda }`$. To deal with the Cauchy kernel it is necessary to work in a completion of $`\mathrm{\Lambda }`$. Let $`\widehat{\mathrm{\Lambda }}=𝔽[[p_1,p_2,\mathrm{}]]`$, the $`𝔽`$-algebra of formal power series in the $`p_i`$. The symmetric realization $`\widehat{\mathrm{\Lambda }}^X`$ of $`\widehat{\mathrm{\Lambda }}`$ is given by the $`𝔽`$-subalgebra of symmetric series in $`𝔽[[X]]`$. Given any element $`E𝔽_b[[X]]`$ with no constant term and $`P\widehat{\mathrm{\Lambda }}`$, the plethysm $`P[E]𝔽[[X]]`$ may be defined by substitution as before. Define the Cauchy kernel $`\mathrm{\Omega }\widehat{\mathrm{\Lambda }}`$ by $$\mathrm{\Omega }=\mathrm{exp}\left(\underset{r1}{}p_r/r\right).$$ The following formulas are consequences of the definitions: $$\begin{array}{cc}\hfill \mathrm{\Omega }[zX]& =\underset{i}{}\frac{1}{1zx_i}=\underset{k0}{}z^kh_k[X]\hfill \\ \hfill \mathrm{\Omega }[zX]& =\underset{i}{}1zx_i=\underset{k0}{}(z)^ke_k[X]\hfill \\ \hfill \mathrm{\Omega }[X+Y]& =\mathrm{\Omega }[X]\mathrm{\Omega }[Y]\hfill \\ \hfill \mathrm{\Omega }[XY]& =\mathrm{\Omega }[X]\mathrm{\Omega }[Y]=\mathrm{\Omega }[X]/\mathrm{\Omega }[Y]\hfill \\ \hfill \mathrm{\Omega }[XY]& =\underset{\lambda 𝒫}{}s_\lambda [X]s_\lambda [Y]\hfill \end{array}$$ where $`h_k`$, $`e_k`$, and $`s_\lambda `$ are the homogeneous, elementary, and Schur symmetric functions respectively. Two bases $`\{a_\lambda \}`$ and $`\{b_\lambda \}`$ of $`\mathrm{\Lambda }`$ are dual with respect to the scalar product if and only if $`_{\lambda 𝒫}a_\lambda [X]b_\lambda [Y]=\mathrm{\Omega }[XY]`$. For $`f\mathrm{\Lambda }`$, let $`f^{}`$ be the linear operator on $`\mathrm{\Lambda }`$ that is adjoint to multiplication by $`f`$ with respect to the scalar product: (2) $$f^{}(g),h=g,fh$$ for all $`g,h\mathrm{\Lambda }`$. This operator is usually referred to as “f-perp” or “skewing by f.” The skewing operators and the map $`\mathrm{\Delta }`$ (or equivalently the plethystic map $`PP[X+Y]`$) can be expressed in terms of each other. For any $`f,g,h\mathrm{\Lambda }`$, (3) $$f^{}(g),h=g[X+Y],f[Y]h[X]$$ using (2), (1), and the commutativity of multiplication in $`\mathrm{\Lambda }`$. Let $`\{a_\lambda \}`$ and $`\{b_\lambda \}`$ be dual bases of $`\mathrm{\Lambda }`$. Then for every $`P\mathrm{\Lambda }`$, (4) $$P[X+Y]=\underset{\lambda }{}(b_\lambda ^{}(P))[X]a_\lambda [Y].$$ ## 3. Definition of the operator Using plethystic substitution we define a family of linear operators on symmetric functions. Define the formal Laurent series $`H(Z^k)`$ in an ordered set of variables $`Z^k=(z_1,z_2,\mathrm{},z_k)`$ with coefficients given by operators on $`\mathrm{\Lambda }`$, which acts on $`P\mathrm{\Lambda }`$ by (5) $$H(Z^k)P[X]=P[X(1q)Z^{}]\mathrm{\Omega }[ZX]R(Z^k)$$ where $`Z^{}=_{i=1}^kz_i^1`$, $`Z=_{i=1}^kz_i`$ and $`R(Z^k)=_{1i<jk}1z_j/z_i`$. For $`v^k`$, define the operator $`H_v^q`$ by (6) $$H_v^qP[X]=H(Z^k)P[X]|_{z^v}.$$ ###### Remark 2. 1. If $`k=1`$ this is Garsia’s version of Jing’s Hall-Littlewood vertex operator . 2. At $`q=0`$ this formula reduces to plethystic notation for a repeated application of the Schur function vertex operator that is due to Bernstein \[9, p.96, # 29 (d)\], . So $`H_{(v_1)}^0H_{(v_2)}^0\mathrm{}H_{(v_k)}^0=H_v^0`$ for any $`v^k`$. If $`v`$ is a partition $`\lambda =(\lambda _1,\mathrm{},\lambda _k)`$ with at most $`k`$ parts then $`H_\lambda ^0\mathrm{\hspace{0.17em}1}=s_\lambda [X]`$. 3. When $`q=1`$ and $`\lambda `$ is a partition, this formula reduces to multiplication by the Schur function $`s_\lambda `$: $`\mathrm{\Omega }[ZX]R(Z^k)|_{z^\lambda }=s_\lambda [X].`$ 4. $`H_\lambda ^q\mathrm{\hspace{0.17em}1}=s_\lambda `$ for any partition $`\lambda =(\lambda _1,\mathrm{},\lambda _k)`$ with at most $`k`$ parts. If $`v_k<0`$ then $`H_v^q\mathrm{\hspace{0.17em}1}=0`$. This operator possesses the same shifted skew symmetry in its subscript that Schur functions have. Let $`_{}^k`$ denote the set of dominant integral weights in $`^k`$, that is, the weakly decreasing sequences in $`^k`$. Let $`𝒫^k_{}^k`$ be the set of partitions of length at most $`k`$, which are always regarded as having $`k`$ parts, some of which may be zero. ###### Proposition 3. Let $`v^k`$. Then $$H_v^q=H_{(v_1,v_2\mathrm{},v_{i+1}1,v_i+1,\mathrm{},v_k)}^q.$$ In particular if $`v_{i+1}=v_i+1`$ then $`H_v^q=0`$. ###### Proof. Let $`s_iZ^k`$ be the sequence $`Z^k`$ with $`z_i`$ and $`z_{i+1}`$ exchanged. Then $`H(s_iZ^k)=\frac{1z_i/z_{i+1}}{1z_{i+1}/z_i}H(Z^k)=\frac{z_i}{z_{i+1}}H(Z^k)`$. The result follows by taking the coefficient of $`z^v`$ on both sides of this equation. If $`v_{i+1}=v_i+1`$ then $`H_v^q=H_v^q`$ and hence must be zero. ∎ The following corollary shows that for every $`v^k`$, $`H_v^q`$ is either zero or up to sign equal to $`H_\nu ^q`$ for some $`\nu _{}^k`$. ###### Corollary 4. Let $`v^k`$, $`\sigma S_k`$, and $`\rho =\rho ^{(k)}=(k1,k2,\mathrm{},0)`$. Then (7) $$H_v^q=sign(\sigma )H_{\sigma (v+\rho )\rho }^q.$$ In particular, if $`v+\rho `$ has two equal parts then $`H_v^q=0`$. Otherwise there is a unique $`\sigma S_k`$ such that $`\sigma (v+\rho )`$ is strictly decreasing, so that $`\sigma (v+\rho )\rho _{}^k`$. ## 4. Another formula for the vertex operator We derive another formula for the operator $`H_\nu ^q`$, which will be used to prove a strong linear independence property of the operators. To do this, it is convenient to introduce some notation for the irreducible rational characters of $`GL(k)`$. The polynomial representation ring of $`GL(k)`$ is isomorphic to the $`𝔽`$-algebra $`𝔽[z_1,z_2,\mathrm{},z_k]^{S_k}`$ of symmetric polynomials in the variables $`z_i`$. The rational representation ring $`R(GL(k))`$ is isomorphic to the $`𝔽`$-algebra of symmetric Laurent polynomials in the variables $`z_1,z_2,\mathrm{},z_k`$ or equivalently the localization $`𝔽[z_1,\mathrm{},z_k]^{S_k}[det[Z]^1]`$ of the algebra of symmetric polynomials obtained by inverting the character $`det[Z]=z_1z_2\mathrm{}z_k`$ of the determinant module. $`R(GL(k))`$ has a distinguished basis $`\{s_\lambda [Z]\lambda _{}^k\}`$ where $`s_\lambda [Z]`$ is the character of the irreducible finite-dimensional $`GL(n)`$-module $`V^\lambda `$ of highest weight $`\lambda `$. Explicitly, if $`m`$ is an integer such that $`m\lambda _k`$ then $`s_\lambda [Z]=det[Z]^ms_{\lambda +(m^k)}[Z]`$ where $`m^k=(m,m,\mathrm{},m)^k`$, $`\lambda +(m^k)`$ is the vector sum in $`^k`$ (which is a partition) and $`s_{\lambda +(m^k)}[Z]`$ is the Schur polynomial. For $`v^k`$, define its dual weight $`v^{}=(v_k,v_{k1},\mathrm{},v_1)`$. If $`\lambda _{}^k`$ then so is $`\lambda ^{}`$; it satisfies $`(V^\lambda )^{}V^\lambda ^{}`$ where $`V^{}`$ is the contragredient dual of $`V`$. Denote by $`\overline{c}_{\mu \nu }^\lambda `$ the tensor product multiplicities: $$s_\mu [Z]s_\nu [Z]=\underset{\lambda _{}^k}{}\overline{c}_{\mu \nu }^\lambda s_\lambda [Z]$$ for $`\lambda ,\mu ,\nu _{}^k`$. It is well-known that (8) $$\overline{c}_{\mu \nu }^\lambda =dim(V^\lambda ^{}V^\mu V^\nu )^{GL(n)}$$ where $`V^G`$ denotes the submodule of $`V`$ fixed by $`G`$. Moreover for any finite-dimensional $`G`$-module, (9) $$dim((V^{})^G)=dim(V^G).$$ Let (10) $$\overline{H}(Z)P[X]=P[X(1q)Z^{}]\mathrm{\Omega }[ZX].$$ Then by (5) and (6), (11) $$\overline{H}(Z)=\underset{\nu _{}^k}{}s_\nu [Z]H_\nu ^q.$$ ###### Proposition 5. For $`\nu _{}^k`$, (12) $$H_\nu ^q=\underset{\lambda ,\mu 𝒫^k}{}\overline{c}_{\mu \nu }^\lambda s_\lambda [X]s_\mu [X(q1)]^{}.$$ ###### Proof. Let $`Y`$ be another set of variables and let $`\overline{H}(Z)`$ act on the $`X`$ variables. We have $$\begin{array}{cc}\hfill \overline{H}(Z)\mathrm{\Omega }[XY]& =\mathrm{\Omega }[XY]\mathrm{\Omega }[(q1)Z^{}Y]\mathrm{\Omega }[XZ]\hfill \\ & =\mathrm{\Omega }[XY]\underset{\nu _{}^k}{}\underset{\lambda ,\mu 𝒫^k}{}s_\nu [Z]\overline{c}_{\mu ^{}\lambda }^\nu s_\mu [(q1)Y]s_\lambda [X].\hfill \end{array}$$ The tensor product multiplicity can be rewritten using (8) and (9) so that $`\overline{c}_{\mu ^{}\lambda }^\nu =\overline{c}_{\mu \nu }^\lambda `$. Clearly we have for fixed $`\gamma 𝒫^k`$ and $`\nu _{}^k`$ $`H_\nu ^qs_\gamma [X]`$ $`=\overline{H}(Z)\mathrm{\Omega }[XY]|_{s_\nu [Z]s_\gamma [Y]}`$ $`={\displaystyle \underset{\lambda ,\mu 𝒫^k}{}}\overline{c}_{\mu \nu }^\lambda s_\mu [(q1)Y]^{}s_\gamma [Y],\mathrm{\Omega }[XY]s_\lambda [X]_Y`$ $`={\displaystyle \underset{\lambda ,\mu 𝒫^k}{}}\overline{c}_{\mu \nu }^\lambda s_\lambda [X]s_\mu [(q1)X]^{}s_\gamma [X]`$ It is seen that (12) holds on the Schur basis and hence on $`\mathrm{\Lambda }`$. ∎ ## 5. Linear independence The full strength of the following result is required later, when we work with operators on $`\mathrm{\Lambda }`$ that are infinite linear combinations of the $`H_\nu ^q`$. ###### Proposition 6. Let $`\{c_\nu 𝔽\nu _{}^k,|\nu |=d\}`$ be an arbitrary collection of scalars. Then the map $`F=_{\nu _{}^k,|\nu |=d}c_\nu H_\nu ^q`$ is a well-defined linear operator on $`\mathrm{\Lambda }`$. Moreover this operator is $`0`$ if and only if $`c_\nu =0`$ for all $`\nu _{}^k`$ with $`|\nu |=d`$. ###### Proof. Any infinite linear combination of operators of the form $`s_\lambda [X]s_\mu [X(q1)]^{}`$ with $`|\lambda |=|\mu |+d`$ is well-defined because its action on any given element of the basis $`\{s_\lambda [X/(q1)]\}`$ has only finitely many nonzero summands. The operator $`F=_{\nu _{}^k,|\nu |=d}c_\nu H_\nu ^q`$ is well-defined because for fixed partitions $`\lambda `$ and $`\mu `$, $`s_\lambda [X]s_\mu [X(q1)]^{}`$ appears in the formula given in Proposition 5 for only finitely many $`H_\nu ^q`$, namely for those such that $`\overline{c}_{\mu \nu }^\lambda 0`$. For $`\gamma _{}^k`$, set $`\alpha (\gamma )𝒫^k`$ to be the partition with $`\alpha (\gamma )_i=\mathrm{max}(\gamma _i,0)`$ and $`\beta (\gamma )𝒫^k`$ be the partition defined by $`\beta (\gamma )_i=min(\gamma _{k+1i},0)`$. By Proposition 5, for $`\tau 𝒫^k`$, (13) $$H_\gamma ^q(s_\tau [X/(q1)])=\underset{\lambda ,\mu 𝒫^k}{}\overline{c}_{\mu \gamma }^\lambda s_\lambda [X]s_\mu [X(q1)]^{}(s_\tau [X/(q1)]).$$ The Littlewood Richardson rule implies that the coefficient $`\overline{c}_{\mu \gamma }^\lambda =0`$ unless $`\mu \beta (\gamma )`$. Since $`\{s_\mu [X(q1)]\}`$ and $`\{s_\tau [X/(q1)]\}`$ are dual bases, we calculate that $`H_\gamma ^q(s_\tau [X/(q1)])=0`$ if $`|\tau ||\beta (\gamma )|`$ and $`\tau \beta (\gamma )`$, and $`H_\gamma ^q(s_{\beta (\gamma )}[X/(q1)])=s_{\alpha (\gamma )}[X]`$. It follows now that $`F=0`$ if and only if $`c_\gamma =0`$ for all $`\gamma `$. For if all of the coefficients are not zero, then $`\gamma `$ is chosen such that $`c_\gamma `$ is non-zero and $`|\beta (\gamma )|`$ is a minimum. We see then that (14) $$F(s_{\beta (\gamma )}[X/(q1)])=\underset{\beta (\nu )=\beta (\gamma )}{}c_\nu s_{\alpha (\nu )}[X]0$$ ## 6. Connection with generalized Kostka polynomials Next it is shown that the operators $`H_\nu ^q`$ have the same relation to the generalized Kostka polynomials of that the components of Garsia’s modified Hall-Littlewood vertex operator has to the Kostka-Foulkes polynomials. Let us recall the definition of the generalized Kostka polynomials. Let $`\eta =(\eta _1,\eta _2,\mathrm{},\eta _t)`$ be a sequence of positive integers summing to $`n`$. Let $$\mathrm{Roots}_\eta =\{(i,j)1i\eta _1+\mathrm{}+\eta _k<jn\text{ for some }k\text{ }\}$$ be the set of strictly upper block triangular positions in an $`n\times n`$ matrix with diagonal block sizes given by $`\eta `$. Let $`Z`$ be the sequence of variables $`z_1`$ through $`z_n`$. Define the formal power series $`B_\eta [Z;q]`$ by $$B_\eta [Z;q]=\underset{(i,j)\mathrm{Roots}_\eta }{}\frac{1}{1qz_i/z_j}.$$ Let $`J`$ be the antisymmetrizer $`J=_{\sigma S_n}\mathrm{sign}(\sigma )\sigma `$. Define the linear operator $`\pi :𝔽[z_1^\pm ,\mathrm{},z_n^\pm ]R(GL(n))`$ given by $`\pi (f)=J(z^\rho )^1J(z^\rho f)`$. Let $`\gamma ^n`$. Define the generating series $$_{\eta ,\gamma }[Z;q]=\pi \left(z^\gamma B_\eta [Z;q]\right).$$ Since $`\{s_\lambda [Z]\lambda _{}^n\}`$ is a basis of $`R(GL(n))`$, define $`K_{\lambda \gamma \eta }(q)[[q]]`$ by $$_{\eta ,\gamma }[Z;q]=\underset{\lambda _{}^n}{}s_\lambda [Z]K_{\lambda \gamma \eta }(q).$$ It is shown in that $`K_{\lambda \gamma \eta }(q)[q]`$. By developing the product of geometric series in $`B_\eta [Z;q]`$ and using the shifted skew symmetry of the index for irreducible characters $$s_v[Z]=\mathrm{sign}(\sigma )s_{\sigma (v+\rho )\rho }[Z]$$ for $`v^n`$ and $`\sigma S_n`$, one obtains the following expression : (15) $$K_{\lambda \gamma \eta }(q)=\underset{\sigma S_n}{}\mathrm{sign}(\sigma )\underset{\begin{array}{c}m:\mathrm{Roots}_\eta \\ \mathrm{wt}(m)=\sigma (\lambda +\rho )(\gamma +\rho )\end{array}}{}q^{|m|}$$ where $`ϵ_1`$ through $`ϵ_n`$ is the standard basis of $`^n`$, $`|m|=_{(i,j)\mathrm{Roots}_\eta }m(i,j)`$, and $`\mathrm{wt}(m)=_{(i,j)\mathrm{Roots}_\eta }m(i,j)(ϵ_iϵ_j)`$. Given $`\eta `$ and $`\gamma `$, write $`\gamma =(\gamma ^{(1)},\gamma ^{(2)},\mathrm{},\gamma ^{(t)})`$ with $`\gamma ^{(i)}^{\eta _i}`$ and assume that $`\gamma ^{(i)}_{}^{\eta _i}`$. ###### Proposition 7. (16) $$H_{\gamma ^{(1)}}^qH_{\gamma ^{(2)}}^q\mathrm{}H_{\gamma ^{(t)}}^q=\underset{\lambda _{}^n}{}K_{\lambda ,\gamma ,\eta }(q)H_\lambda ^q.$$ ###### Proof. Let $`Z^n`$ be the entire set of $`n`$ variables $`z_1`$ through $`z_n`$. Break this set into $`t`$ collections of successive variables such that the $`i`$-th collection $`Z^{(i)}`$ has size $`\eta _i`$ for $`1it`$. We have the relation $`H(U^k)H(V^{\mathrm{}})=\mathrm{\Omega }[qU^{}V]H(U^k,V^{\mathrm{}}),`$ which is verified by showing it holds when evaluated on an arbitrary $`P\mathrm{\Lambda }`$: $`H(U^k)H(V^{\mathrm{}})P[X]`$ $`=P[X(1q)(U^{}+V^{})]\mathrm{\Omega }[UX]`$ $`\mathrm{\Omega }[V(X(1q)U^{})]R(U^k)R(V^{\mathrm{}})`$ $`=P[X(1q)(U^{}+V^{})]\mathrm{\Omega }[(U+V)X]`$ $`\mathrm{\Omega }[qU^{}V]R(U^k,V^{\mathrm{}})`$ It follows that (17) $$H(Z^{(1)})H(Z^{(2)})\mathrm{}H(Z^{(t)})=H(Z^n)\underset{1i<jt}{}\mathrm{\Omega }[q(Z^{(i)})^{}Z^{(j)}]$$ Observe that $`H_{\gamma ^{(1)}}^qH_{\gamma ^{(2)}}^q\mathrm{}H_{\gamma ^{(t)}}^q`$ is the coefficient of $`z^\gamma `$ on the left hand side of (17). The coefficient of $`z^\gamma `$ on the right hand side is $`_mq^{|m|}H_{\gamma +\mathrm{wt}(m)}^q`$ where $`m`$ runs over the functions $`m:\mathrm{Roots}_\eta `$. By Proposition 6 the operators $`H_\lambda ^q`$ are independent for $`\lambda _{}^n`$. Hence we may take the coefficient of $`H_\lambda ^q`$ on both sides. Using Corollary 4 the right hand side becomes precisely the the expression (15) of $`K_{\lambda \gamma \eta }(q)`$. ∎ If $`\gamma `$ is such that $`\gamma ^{(i)}𝒫^{\eta _i}`$, define the symmetric function (18) $$_{(\gamma ^{(1)},\mathrm{},\gamma ^{(t)})}[X;q]=H_{\gamma ^{(1)}}^qH_{\gamma ^{(2)}}^q\mathrm{}H_{\gamma ^{(t)}}^q\mathrm{\hspace{0.17em}1}.$$ Then (19) $$_{(\gamma ^{(1)},\mathrm{},\gamma ^{(t)})}[X;q]=\underset{\lambda 𝒫^n}{}K_{\lambda \gamma \eta }(q)s_\lambda [X]$$ which is a finite sum, unlike the expansion (16) which is an infinite sum. At $`q=1`$ this is the expansion of the product of Schur functions $`s_{\gamma ^{(1)}}\mathrm{}s_{\gamma ^{(t)}}`$ in the Schur basis. We have shown that the coefficients of the expansions of $`_{\eta ,\gamma }[Z;q]`$ in terms of irreducible characters, and $`H_{\gamma ^{(1)}}^q\mathrm{}H_{\gamma ^{(t)}}^q`$ in terms of the $`H_\lambda ^q`$ for $`\lambda _{}^n`$, are the same. Consequently, at least on the level of characters, the questions of regarding a certain Grothendieck ring of graded modules, can be translated into questions regarding the span of the above operators. ## 7. Spaces of vertex operators and commutation relations By manipulating the order of the variables we may derive several sorts of explicit commutation relations. These are applied to prove basis theorems for spaces of our operators. Let $`𝒢(k,n)`$ denote the $`[q]`$-span of operators of the form $`H_\mu ^qH_\nu ^q`$ where $`\mu ^k`$ and $`\nu ^n`$, and $`𝒢_{}(k,n)`$ the $`[q]`$-span of such operators with the concatenated weight dominant, that is, $`(\mu ,\nu )_{}^{k+n}`$. The following are vertex operator analogues of results proven in for Grothendieck groups. Our proofs have the advantage of being explicit and working over $`[q]`$ instead of $`[q,q^1]`$. Moreover our relations among vertex operators can be lifted to relations in the Grothendieck groups. ###### Theorem 8. $`𝒢(k,n)=𝒢_{}(k,n)`$. ###### Theorem 9. If $`k>n`$ then $`𝒢(k,n)𝒢(k1,n+1)`$ and $`𝒢(n,k)𝒢(n+1,k1)`$. ###### Theorem 10. $`𝒢(k,n)=𝒢(n,k)`$. Consider the following generating function equation that follows easily from manipulating the operators on an arbitrary symmetric function. (20) $$\begin{array}{cc}\hfill H(& U^k,V^{\mathrm{}})H(W^mZ^n)\mathrm{\Omega }[q(U^{}W+V^{}W+V^{}Z)]\hfill \\ & =H(U^k,W^m)H(V^{\mathrm{}},Z^n)\mathrm{\Omega }[q(U^{}V+W^{}V+W^{}Z)](1)^\mathrm{}m\underset{i=1}{\overset{m}{}}w_i^{\mathrm{}}\underset{i=1}{\overset{\mathrm{}}{}}v_i^m\hfill \end{array}$$ Setting $`\mathrm{}=m=1`$ in this formula gives enough relations to prove Theorem 8. For brevity of notation, define $`|_\beta ^\alpha =|\alpha ||\beta |`$. If $`\mu ,\nu _{}^n`$ then we will say that $`\nu /\mu `$ is a vertical strip, and denote this by $`\nu /\mu 𝒱`$, provided that $`\nu _i\mu _i\{0,1\}`$ for all $`1in`$. ###### Lemma 11. For all $`a,b`$ and $`\mu _{}^k`$ and $`\nu _{}^n`$, $`{\displaystyle \underset{\alpha /\mu 𝒱}{}}{\displaystyle \underset{\nu /\beta 𝒱}{}}(q)^{|_\mu ^\alpha +|_\beta ^\nu }(H_{(\alpha ,a+|_\beta ^\nu )}^qH_{(b|_\mu ^\alpha ,\beta )}^qqH_{(\alpha ,a+|_\beta ^\nu +1)}^qH_{(b|_\mu ^\alpha 1,\beta )}^q)=`$ (21) $`{\displaystyle \underset{\alpha /\mu 𝒱}{}}{\displaystyle \underset{\nu /\beta 𝒱}{}}(q)^{|_\mu ^\alpha +|_\beta ^\nu }(qH_{(\alpha ,b+|_\beta ^\nu )}^qH_{(a|_\mu ^\alpha ,\beta )}^qH_{(\alpha ,b+|_\beta ^\nu 1)}^qH_{(a|_\mu ^\alpha +1,\beta )}^q)`$ ###### Proof. From equation (20), $$\begin{array}{cc}& H(U^k,v)H(w,Z^n)\mathrm{\Omega }[q(U^{}w+v^{}w+v^{}Z)]\hfill \\ \hfill =& H(U^k,w)H(v,Z^n)\mathrm{\Omega }[q(U^{}v+w^{}v+w^{}Z)](w/v)\hfill \end{array}$$ The desired identity is obtained by taking the coefficient of $`u^\mu v^aw^bz^\nu `$. ∎ Note that if $`b=a+1`$, then (21) reduces significantly to the following identity. $`{\displaystyle \underset{\alpha /\mu 𝒱}{}}{\displaystyle \underset{\nu /\beta 𝒱}{}}(q)^{|_\mu ^\alpha +|_\beta ^\nu }H_{(\alpha ,a+|_\beta ^\nu )}^qH_{(a+1|_\mu ^\alpha ,\beta )}^q`$ (22) $`={\displaystyle \underset{\alpha /\mu 𝒱}{}}{\displaystyle \underset{\nu /\beta 𝒱}{}}q(q)^{|_\mu ^\alpha +|_\beta ^\nu }H_{(\alpha ,a+|_\beta ^\nu +1)}^qH_{(a|_\mu ^\alpha ,\beta )}^q`$ Relations (21) and (22) suffice to prove Theorem 8. Before giving the proof an example is helpful. ###### Example 12. We wish to show that $`H_{(22)}^qH_{(41)}^q`$ is a linear combination of $`H_\alpha ^qH_\beta ^q`$ with $`\alpha ,\beta ^2`$ such that $`(\alpha ,\beta )`$ is dominant. $`H_{(22)}^qH_{(41)}^qqH_{(23)}^qH_{(40)}^qqH_{(32)}^qH_{(31)}^q+q^2H_{(33)}^qH_{(30)}^q`$ $`qH_{(23)}^qH_{(31)}^q+q^2H_{(24)}^qH_{(30)}^q+q^2H_{(33)}^qH_{(21)}^qq^3H_{(34)}^qH_{(20)}^q=`$ $`qH_{(24)}^qH_{(21)}^qq^2H_{(25)}^qH_{(20)}^qq^2H_{(34)}^qH_{(11)}^q+q^3H_{(35)}^qH_{(10)}^q`$ $`H_{(23)}^qH_{(31)}^q+qH_{(24)}^qH_{(30)}^q+qH_{(33)}^qH_{(21)}^qq^2H_{(34)}^qH_{(20)}^q`$ By Corollary 4 many terms vanish and others cancel with each other. The terms $`H_{(23)}^qH_{(40)}^q`$, $`H_{(23)}^qH_{(31)}^q`$, $`H_{(34)}^qH_{(20)}^q`$, and $`H_{(34)}^qH_{(11)}^q`$ are all zero. The terms $`q^2H_{(24)}^qH_{(30)}^q`$ and $`q^2H_{(33)}^qH_{(30)}^q`$ cancel and so do $`qH_{(24)}^qH_{(21)}^q`$ and $`qH_{(33)}^qH_{(21)}^q`$. When this relation is reduced and $`H_{(22)}^qH_{(41)}^q`$ is expressed alone on the left hand side of the equation we have $`H_{(22)}^qH_{(41)}^q=`$ $`qH_{(32)}^qH_{(31)}^qq^2H_{(33)}^qH_{(21)}^q+q^2H_{(43)}^qH_{(20)}^q`$ $`q^3H_{(44)}^qH_{(10)}^qqH_{(33)}^qH_{(30)}^q`$ On the right hand side of this equation, only $`H_{(32)}^qH_{(31)}^q`$ does not have the property that the concatenated indexing weights are dominant. We apply relation (22) to this operator: $`H_{(32)}^qH_{(31)}^qqH_{(33)}^qH_{(30)}^qqH_{(42)}^qH_{(21)}^q+q^2H_{(43)}^qH_{(20)}^q=`$ $`qH_{(33)}^qH_{(21)}^qq^2H_{(34)}^qH_{(20)}^qq^2H_{(43)}^qH_{(11)}^q+q^3H_{(44)}^qH_{(10)}^q`$ Therefore $`H_{(22)}^qH_{(41)}^q=`$ $`(q^2q)H_{(33)}^qH_{(30)}^q+q^2H_{(42)}^qH_{(21)}^qq^3H_{(43)}^qH_{(11)}^q`$ $`+(q^2q^3)H_{(43)}^qH_{(20)}^q+(q^4q^3)H_{(44)}^qH_{(10)}^q`$ We now give the proof of Theorem 8. ###### Proof. It is enough to show that if $`(\mu ,a)_{}^k`$ and $`(b,\nu )_{}^n`$ then $$H_{(\mu ,a)}^qH_{(b,\nu )}^q𝒢_{}(k,n).$$ If $`b=a+1`$ then in the relation (22) the only term that is not indexed by weights such that $`(\alpha ,\beta )_{}^{k+n}`$, is $`H_{(\mu ,a)}^qH_{(a+1,\nu )}^q`$, and hence it is in $`𝒢_{}(k,n)`$. Otherwise assume that $`b>a+1`$. All of the terms of the equation (21) are of the form $`H_{(\alpha ,a+|_\beta ^\nu )}^qH_{(b|_\mu ^\alpha ,\beta )}^q`$, $`H_{(\alpha ,a+|_\beta ^\nu +1)}^qH_{(b|_\mu ^\alpha 1,\beta )}^q`$, $`H_{(\alpha ,b+|_\beta ^\nu )}^qH_{(a|_\mu ^\alpha ,\beta )}^q`$, or $`H_{(\alpha ,b+|_\beta ^\nu 1)}^qH_{(a|_\mu ^\alpha +1,\beta )}^q`$. Let $`H_{(\theta ,c)}^qH_{(d,\gamma )}^q`$ be one of these terms after it has been rewritten using Corollary 4 so that $`(\theta ,c)_{}^k`$ and $`(d,\gamma )_{}^n`$. Consider $`H_{(\theta ,c)}^qH_{(d,\gamma )}^q`$ of the first form since the others follow from essentially the same remark. We verify that $`dc<ba`$ (unless $`H_{(\theta ,c)}^qH_{(d,\gamma )}^q=H_{(\mu ,a)}^qH_{(a+1,\nu )}^q`$) and hence by induction the theorem is true since if $`dc0`$ then $`(\theta ,c,d,\gamma )_{}^{k+n}`$. By definition, $$\begin{array}{cc}\hfill c& =\underset{1ik1}{\mathrm{min}}\{a+|_\beta ^\nu ,\alpha _i+(ki)\}\hfill \\ \hfill d& =\underset{1in1}{\mathrm{max}}\{b|_\mu ^\alpha ,\beta _ii\}\hfill \end{array}$$ We note that $$c\underset{1ik1}{\mathrm{min}}\{a+|_\beta ^\nu ,\mu _i+(ki)\}\mathrm{min}\{a+|_\beta ^\nu ,a+1\}$$ and hence $`ca`$ with equality if and only if $`|_\beta ^\nu =0`$. Similarly $$d\underset{1in1}{\mathrm{max}}\{b|_\mu ^\alpha ,\nu _ii\}\mathrm{max}\{b|_\mu ^\alpha ,bi\}$$ and so $`db`$ with equality if and only if $`|_\mu ^\alpha =0`$. ∎ The generalization of this statement, that $`H_{\gamma ^{(1)}}^qH_{\gamma ^{(2)}}^q\mathrm{}H_{\gamma ^{(t)}}^q`$ with $`\gamma ^{(i)}^{\eta _i}`$ is in the $`[q]`$ span of the operators $`H_{\alpha ^{(1)}}^qH_{\alpha ^{(2)}}^q\mathrm{}H_{\alpha ^{(t)}}^q`$ with $`\alpha ^{(i)}^{\eta _i}`$ and $`(\alpha ^{(1)},\alpha ^{(2)},\mathrm{},\alpha ^{(t)})`$ a dominant weight, is conjectured to be true , but does not seem to follow easily from these relations. By setting $`\mathrm{}=1`$ and $`m=0`$ in (20), we immediately obtain another relation among the operators which gives us another basis theorem. ###### Lemma 13. For all $`a`$, $`\mu _{}^k`$, and $`\nu _{}^n`$, (23) $$\underset{\begin{array}{c}\beta _{}^n\\ \nu /\beta 𝒱\end{array}}{}(q)^{|_\beta ^\nu }H_{(\mu ,a+|_\beta ^\nu )}^qH_\beta ^q=\underset{\begin{array}{c}\alpha _{}^k\\ \alpha /\mu 𝒱\end{array}}{}(q)^{|_\mu ^\alpha }H_\alpha ^qH_{(a|_\mu ^\alpha ,\nu )}^q.$$ ###### Proof. From equation (20), (24) $$H(U^k,v)H(Z^n)\mathrm{\Omega }[qv^{}Z]=H(U^k)H(v,Z^n)\mathrm{\Omega }[qvU^{}]$$ The stated identity is obtained by taking the coefficient of $`u^\mu v^az^\nu `$. ∎ Before proving Theorem 9 we give an example. ###### Example 14. We wish to write the composition of operators $`H_{(53)}^qH_{(2)}^q`$ as a linear combination of operators $`H_\alpha ^qH_\beta ^q`$ with $`\alpha ^1`$ and $`\beta ^2`$. $$H_{(53)}^qH_{(2)}^qqH_{(54)}^qH_{(1)}^q=H_{(5)}^qH_{(32)}^qqH_{(6)}^qH_{(21)}^q$$ $$H_{(54)}^qH_{(1)}^qqH_{(55)}^qH_{(0)}^q=H_{(5)}^qH_{(41)}^qqH_{(6)}^qH_{(31)}^q$$ $$H_{(55)}^qH_{(0)}^q=H_{(5)}^qH_{(50)}^qqH_{(6)}^qH_{(40)}^q$$ Making repeated substitutions, we have the relation $`H_{(53)}^qH_{(2)}^q=`$ $`q^2H_{(5)}^qH_{(50)}^qq^3H_{(6)}^qH_{(40)}^q+qH_{(5)}^qH_{(41)}^q`$ $`q^2H_{(6)}^qH_{(31)}^q+H_{(5)}^qH_{(32)}^qqH_{(6)}^qH_{(21)}^q`$ We now give the proof of Theorem 9. ###### Proof. We prove that if $`k>n`$ then $`𝒢(k,n)𝒢(k1,n+1)`$ as the other part is proven in the same manner. It is enough to show that if $`k>n`$, $`(\mu ,a)_{}^k`$ and $`\nu _{}^n`$ then $$H_{(\mu ,a)}^qH_\nu ^q𝒢_{}(k1,n+1)$$ We use (23) with $`k1`$ and $`n`$. The proof proceeds by induction on $`\mu _1a`$. If $`\mu _1a=0`$, then there is exactly one term on the left hand side of equation (23), namely $`H_{(\mu ,a)}^qH_\nu ^q`$. The right hand side is a $`[q]`$-linear combination of $`H_\gamma ^qH_\rho ^q`$ with $`\gamma ^{k1}`$ and $`\rho ^{n+1}`$. If $`\mu _1a>0`$ then the left hand side is a linear combination of operators of the form $`H_{(\gamma ,c)}^qH_\beta ^q`$ with $`\gamma _1=\mu _1`$ and $`c`$ $`=\underset{1ik1}{\mathrm{min}}\{a+|_\beta ^\nu ,\mu _i+(ki)\}`$ $`\underset{1ik1}{\mathrm{min}}\{a+|_\beta ^\nu ,a+(ki)\}.`$ It follows that $`ca`$ with equality if and only if $`\nu =\beta `$. The right hand side of this equation is again an element of $`𝒢(k1,n+1)`$. ∎ Finally we prove Theorem 10. ###### Proof. We shall prove $`𝒢(k+\mathrm{},k)𝒢(k,k+\mathrm{})`$ for $`k,\mathrm{}>0`$, the other inclusion being similar. Consider (20) with $`m=0`$ and $`n=k`$ so $`U=U^k`$, $`V=V^{\mathrm{}}`$, $`Z=Z^k`$: (25) $$H(U,V)H(Z)\mathrm{\Omega }[qV^{}Z]=H(U)H(V,Z)\mathrm{\Omega }[qU^{}V].$$ Take the coefficient of $`u^\alpha v^\beta z^\gamma `$ in this equation, where $`(\alpha ,\beta )_{}^{k+\mathrm{}}`$ and $`\gamma _{}^k`$. The entire right hand side consists of terms in $`𝒢(k,k+\mathrm{})`$. Expanding the expression $`\mathrm{\Omega }[qV^{}Z]=(1qz_i/v_j)`$, the left hand side is in the set $$H_{\alpha ,\beta }^qH_\gamma ^q+[q]\underset{\beta ^{},\gamma ^{}}{}H_{\alpha ,\beta ^{}}^qH_\gamma ^{}^q$$ where $`\beta ^{}^{\mathrm{}}`$ such that $`\beta _j^{}\beta _j+k`$ for all $`1j\mathrm{}`$ and $`|\beta ^{}|>|\beta |`$. Rewriting a typical term $`H_{\alpha ,\beta ^{}}^q`$ by Corollary 4, one obtains either $`0`$ or up to sign $`H_{\alpha ^{},\beta ^{\prime \prime }}^q`$, say. This means $`\alpha _1^{}+(k+\mathrm{}1)`$ is the largest part of the weight $`(\alpha ,\beta ^{})+\rho ^{(k+\mathrm{})}`$. Now $`\alpha _1\beta _j`$ for all $`1j\mathrm{}`$ by the dominance of $`(\alpha ,\beta )`$, so $$\begin{array}{cc}\hfill \alpha _1+k+\mathrm{}1& \beta _j+k+\mathrm{}1\beta _j^{}+\mathrm{}1\hfill \\ & \beta _j^{}+\mathrm{}j.\hfill \end{array}$$ It follows that $`\alpha _1^{}=\alpha _1`$, $`(\alpha ^{},\beta ^{\prime \prime })_{}^{k+\mathrm{}}`$ and $`|\alpha ^{}|+|\beta ^{\prime \prime }|>|\alpha |+|\beta |`$. There are only finitely many elements of $`_{}^{k+\mathrm{}}`$ with these properties, so the terms $`H_{(\alpha ^{},\beta ^{\prime \prime })}^qH_\gamma ^{}^q`$ can again be rewritten in the same way, and the process terminates. ∎ ## 8. Recovering identities via commutation relations The commutation relations in the previous section may be used to recover some identities among the operators $`H_\nu ^q`$ that correspond to known identities among generalized Kostka polynomials. ###### Proposition 15. For $`a`$ and positive integers $`k`$ and $`n`$, (26) $$H_{(a^n)}^qH_{(a^k)}^q=H_{(a^k)}^qH_{(a^n)}^q$$ ###### Proof. In the proof of Theorem 10, take $`\alpha =(a^k)`$, $`\beta =(a^{\mathrm{}})`$, and $`\gamma =(a^k)`$. The proof shows there is only one term on either side of the identity coming from (25), which agrees with (26) when $`n=k+\mathrm{}`$. ∎ ###### Proposition 16. (27) $$H_{(a^k)}^qH_{((a+1)^k)}^q=q^kH_{((a+1)^k)}^qH_{(a^k)}^q.$$ ###### Proof. When Lemma 13 is applied with $`\mu =(a^{k1})`$, $`a`$, and $`\nu =((a+1)^k)`$, there are no surviving terms on the right and exactly two on the left, indexed by $`\beta =\nu =((a+1)^k)`$ and $`\beta =(a^k)`$. Applying Corollary 4 to the term with $`\beta =(a^k)`$ the desired identity is obtained. ∎ ###### Proposition 17. For all $`a`$ and $`k1`$, (28) $$H_{(a^k)}^qH_{(a^k)}^q=H_{(a^{k+1})}^qH_{(a^{k1})}^q+q^kH_{((a+1)^k)}^qH_{((a1)^k)}^q.$$ ###### Proof. Apply Lemma 13 with $`n=k1`$, $`\mu =(a^k)`$, and $`\nu =(a^{k1})`$. On the left side one has the single nonvanishing term $`H_{(a^{k+1})}^qH_{(a^{k1})}^q`$ corresponding to the summand $`\beta =\nu =(a^{k1})`$. The right side has two nonvanishing terms indexed by $`\alpha =\mu =(a^k)`$ and $`\alpha =\mu +(1^k)=((a+1)^k)`$. The first of these terms is $`H_{(a^k)}^qH_{(a^k)}^q`$. The second is $`(q)^kH_{((a+1)^k)}^qH_{(ak,a^{k1})}^q`$. But $`H_{(ak,a^{k1})}^q=(1)^{k1}H_{((a1)^k)}^q`$ by Corollary 4. ∎ This identity can be viewed as the trace of a short exact sequence that resolves the ideal of a nilpotent conjugacy class closure over the coordinate ring of a minimally larger one . The version of this identity for the fermionic form of generalized Kostka polynomials appears in . This $`q`$-character identity is put in a more general context in . ## 9. Generalization of Garsia-Procesi defining recurrence for Kostka-Foulkes polynomials Manipulations of the definition allow us to derive commutation relations between the $`H_\nu ^q`$ and the operators $`e_k^{}`$. As a consequence we obtain a generalization of a defining recurrence for the Kostka-Foulkes polynomials given in . Let $`E(u)`$ be the generating function of operators on $`\mathrm{\Lambda }`$ defined by (29) $$E(u)P[X]=P[Xu].$$ By (4) we have $$P[Xu]=\underset{\lambda 𝒫}{}s_\lambda ^{}(P)[X]s_\lambda [u]=\underset{k0}{}e_k^{}(P)[X](u)^k.$$ In other words (30) $$e_k^{}P[X]=(1)^kP[Xu]|_{u^k}.$$ The commutation relation of $`H(Z^n)`$ and $`E(u)`$ is: $$\begin{array}{cc}\hfill E(u)H(Z^n)P[X]& =E(u)P[X(1q)Z^{}]\mathrm{\Omega }[XZ]R(Z^n)\hfill \\ & =P[Xu(1q)Z^{}]\mathrm{\Omega }[(Xu)Z]R(Z^n)\hfill \\ & =\mathrm{\Omega }[uZ]H(Z^n)E(u)P[X]\hfill \end{array}$$ Taking the coefficient of $`(u)^kz^\lambda `$ on both sides of this equation we obtain the following relation. ###### Proposition 18. Let $`\lambda _{}^n`$. Then $$e_k^{}H_\lambda ^q=\underset{\begin{array}{c}\beta _{}^n\\ \lambda /\beta 𝒱\end{array}}{}H_\beta ^qe_{k|\lambda |+|\beta |}^{}$$ Let $`\eta =(\eta _1,\eta _2,\mathrm{},\eta _t)`$ be a fixed sequence of positive integers summing to $`n`$. For any weight $`\gamma ^n`$, write $`\gamma ^{(1)}^{\eta _1}`$ for the first $`\eta _1`$ parts of $`\gamma `$, $`\gamma ^{(2)}^{\eta _2}`$ for the next $`\eta _2`$ parts of $`\gamma `$, etc. ###### Proposition 19. Let $`k`$ be a fixed positive integer, $`\eta =(\eta _1,\eta _2,\mathrm{},\eta _t)`$ a sequence of positive integers summing to $`n`$, $`\alpha _{}^n`$, and $`\gamma ^n`$ such that $`\gamma ^{(i)}_{}^{\eta _i}`$ for all $`i`$ and $`|\gamma ||\alpha |=k`$. Then $$\underset{\begin{array}{c}\nu ^n\\ |\gamma ||\nu |=k\\ \nu ^{(i)}_{}^{\eta _i}\\ \gamma ^{(i)}/\nu ^{(i)}𝒱\end{array}}{}K_{\alpha ,\nu ,\eta }(q)=\underset{\begin{array}{c}\lambda _{}^n\\ |\lambda ||\alpha |=k\\ \lambda /\alpha 𝒱\end{array}}{}K_{\lambda ,\gamma ,\eta }(q)$$ ###### Proof. Let $`\eta `$ and $`\gamma `$ be as above. For the moment assume that the entries of $`\gamma `$ are positive. Apply $`e_k^{}`$ to (16) and apply Proposition 18 to both sides to commute $`e_k^{}`$ to the right of the $`H`$ operators: $$\underset{\begin{array}{c}\nu ^n\\ \nu ^{(i)}_{}^{\eta _i}\\ \gamma ^{(i)}/\nu ^{(i)}𝒱\end{array}}{}H_{\nu ^{(1)}}^qH_{\nu ^{(2)}}^q\mathrm{}H_{\nu ^{(t)}}^qe_{k|\gamma |+|\nu |}^{}=\underset{\lambda _{}^n}{}\underset{\begin{array}{c}\beta _{}^n\\ \lambda /\beta 𝒱\end{array}}{}K_{\lambda ,\nu ,\eta }(q)H_\beta ^qe_{k|\lambda |+|\beta |}^{}.$$ Since $`\gamma `$ has positive parts, all the $`\nu `$ have nonnegative parts. Expanding the left hand side using (16) and applying the resulting operators to $`1\mathrm{\Lambda }`$, we obtain $$\underset{\begin{array}{c}\nu ^n\\ |\gamma ||\nu |=k\\ \nu ^{(i)}_{}^{\eta _i}\\ \gamma ^{(i)}/\nu ^{(i)}𝒱\end{array}}{}\underset{\mu _{}^n}{}K_{\mu ,\nu ,\eta }(q)H_\mu ^q\mathrm{\hspace{0.17em}1}=\underset{\lambda _{}^n}{}\underset{\begin{array}{c}\beta _{}^n\\ |\lambda ||\beta |=k\\ \lambda /\beta 𝒱\end{array}}{}K_{\lambda ,\nu ,\eta }(q)H_\beta ^q\mathrm{\hspace{0.17em}1}.$$ Assume for the moment that $`\alpha `$ has nonnegative parts. For such $`\alpha `$, $`H_\alpha ^q\mathrm{\hspace{0.17em}1}=s_\alpha [X]`$. Taking the coefficient of $`s_\alpha [X]`$ on both sides, we obtain the desired relation. The statement is true in general since $`K_{\lambda (a^n),\gamma (a^n),\eta }(q)=K_{\lambda ,\gamma ,\eta }(q)`$ for all integers $`a`$. ∎ The recurrence for the Kostka-Foulkes polynomials given in is recovered by setting $`\eta =(1^n)`$ in Proposition 19, as $`H_{(m)}^q`$ in our notation is $`H_m`$ in theirs. In the case $`\eta =(1^n)`$ the situation is particularly nice; the nondominant operators can be made dominant using the relation $`H_{(m)}^qH_{(m+1)}^q=qH_{(m+1)}^qH_{(m)}^q`$. For general $`\eta `$ the identities needed to rewrite the nondominant terms as dominant ones, can get complicated and produce negative signs. ## 10. Jing’s operators We give corresponding constructions for Jing’s Hall-Littlewood vertex operators \[9, Ex. III.5.8\]. Define $$\begin{array}{cc}\hfill \overline{B}(Z^k)P[X]& =P[XZ^{}]\mathrm{\Omega }[XZ(1q)]\hfill \\ \hfill B(Z^k)P[X]& =P[XZ^{}]\mathrm{\Omega }[XZ(1q)]R(Z^k)\hfill \\ \hfill B_\nu ^q& =B(Z^k)|_{z^\nu }\hfill \end{array}$$ for $`\nu ^k`$. The generating series of operators $`B(z)`$ (for a single variable $`z`$) and $`\overline{B}(Z^k)`$ defined here, coincide with the operators $`B(z)`$ and $`B(z_1,z_2,\mathrm{},z_k)`$ in the notation of \[9, Ex. III.5.8\]. $`B(z)`$ in our notation is the operator $`H(z)`$ in Jing’s . However the operators $`B_\nu `$ themselves are not studied by Jing or Macdonald. Let $`F`$ be the plethystic operator $`FP[X]=P[X(1q)]`$, with inverse operator $`F^1P[X]=P[X/(1q)]`$. Then it is not hard to show that (31) $$B(Z^k)=FH(Z^k)F^1.$$ This gives the following analogue of (16) for $`B(Z^k)`$. (32) $$B_{\gamma ^{(1)}}^qB_{\gamma ^{(2)}}^q\mathrm{}B_{\gamma ^{(t)}}^q=\underset{\lambda _{}^n}{}K_{\lambda ,\gamma ,\eta }(q)B_\lambda ^q.$$
warning/0001/astro-ph0001315.html
ar5iv
text
# Disk-Halo Interaction - I. Three-Dimensional Evolution of the Galactic Disk ## 1 Introduction This is the first of three papers where the dynamics of the Galactic fountain and its effects in the Galaxy are presented. The present paper describes three-dimensional hydrodynamical simulations of the disk gas powered by isolated and clustered supernovae and its interaction with the halo. In paper II (Avillez, 1999b) a discussion is conducted on the dynamics of the Galactic fountain in the halo and the formation of HI clouds condensing from the fountain flow. In paper III a discussion of the present day modelling of the disk-halo interaction is carried out. This includes a detailed presentation and comparison between the existing models, their strengths and weaknesses and how their predictions compare with observations. The approach adopted here assumes the Galactic fountains are intimately related to the vertical structure of the thick gas disk and to the rate occurrence of supernovae per unit area in the disk. The hot gas released by the explosions breaks through the thick disk, escaping into the halo via the diffuse ionized medium, and forming large scale fountains. The term “galactic fountain” has been used to define a global circulation model where the disk gas rises up into the halo, cools and returns to the disk. The returning cold gas contitutes some of the observed HI clouds in the halo (see Wakker & van Woerden, 1997). The latest developments on the study of the physics of the HI gas in the Galactic halo are reviewed in Wakker et al. (1999). Several mechanisms have been proposed for the disk-halo-disk circulation, taking into account the vertical structure of the Galactic disk and the distribution of O and B stars in the disk. These comprise two major models: the classical fountain (CF) model (Shapiro & Field, 1976; Bregman, 1980; Kahn, 1981) and the chimney model (Tomisaka & Ikeuchi, 1986; Norman & Ikeuchi, 1989). The CF models assumed the fountain gas being originated in type II supernovae randomly distributed in the disk. As a consequence the gas gains enough energy to expand buoyantly into the halo where it cools and condenses into clouds that would rain over the disk. The maximum velocity these clouds would have is 70 km/s (Kahn, 1981), ruling out HVCs (clouds with descending velocities greater than 100 km/s) as a product of a Galactic fountain. These models did not take into account the presence of a thick disk composed of a multiphase medium with different scale heights commonly known as the Lockman (with a scale height $`h_z500`$ pc) (Lockman, 1984; Lockman et al., 1986) and the Reynolds (with a scale height $`h_z11.5`$ kpc) (Reynolds, 1987) layers. The presence of such a thick gas disk constrains the dynamics of the hot intercloud medium on its way out of the disk. The detection of holes with sizes varying between a few hundred and a kilo parsec in the emission maps of nearby spiral galaxies (M31 and M33) by Brinks & Bajaja (1986) and Deul & Hartog (1990) and in the Milky Way by Heiles (1984) and Koo et al (1992) suggest that associations of O and B stars exploding in a correlated fashion could be responsible for the formation of superbubbles that blow holes in the disk. As a consequence, they release their hot inner parts into the lower halo through tunnels physically connected to the underlaying OB associations resembling chimneys (Tomisaka & Ikeuchi (1986), Norman & Ikeuchi (1989), Mac Low et al., 1989). The energies involved in such a phenomenon are of the order of $`10^{53}`$ erg (Heiles, 1991). As the hot gas rises up to a height of 10 kpc, it cools and returns to the Galactic disk, forming a fountain. The returning clouds would then acquire high velocities. Furthermore, chimneys reduce the volume filling factor of the hot gas in the disk reducing it from 0.7 predicted by the three-phase model of McKee & Ostriker (1977) to some 0.2 as predicted by Norman & Ikeuchi (1989). Chimneys have been regarded as the most efficient way to expand the hot disk gas into the halo and as indirectly being reponsible for the formation of HVCs (Kuntz & Danly, 1996; see reviews by Wakker & van Woerden 1997). The major drawback is the number of OB associations required to produce chimneys in such a way as to guarantee a continuous injection of hot gas into the halo. Rosen et al. (1993) and Rosen & Bregman (1995) developed an integrated model where two co-spatial fluids representing the stars and gas in the interstellar medium were used. These models reproduced the presence of a multiphase media with cool, warm and hot intermixed phases having mean scale heights compatible with those observed in the Galaxy. However, the structure and properties of the ISM changed according to the overall energy injection rate. The best range of supernova rates that allow the reproduction of the ISM stratification are closed to the observed in the Galaxy. The resulting volume filling factors ($`50\%`$ for the hot medium, $`2530\%`$ for the warm gas and $`20\%`$ for the cold gas (Rosen et al., 1999)) are slightly different from those expected for the Galactic disk (see Spitzer, 1990; Ferrière, 1995). Substantial progress has been made in modelling the disk-halo interaction and also the Galactic fountain since it was first suggested by Shapiro & Field in 1976. The models described above give a two-dimensional description of the dynamics of the disk and halo gas, and simulate the vertical distribution of IVCs and HVCs in the halo. Simplified conditions for the stars and ISM have been taken into account. The Galactic fountain models assumed that supernovae were randomly distributed in the disk and occurred with a rate of 3 per century, corresponding to a massflux of hot ($`T10^6`$ K) material into the halo of the order of $`10^{19}`$ g cm s<sup>-1</sup>, whereas global models such as those of Rosen & Bregman (1995) showed that it is very difficult to reproduce, in the same simulation, both the structure of the ISM near the disk and, at the same time, the presence of HI gas with high and intermediate velocities in the halo without increasing the volume filling factor of the hot gas in the disk. Both the CF and the global models of the disk-halo interaction show that large scale outflows may be responsible for the formation of some of the observed IVCs and HVCs, providing the hot intercloud medium has a temperature of T$`{}_{}{}^{}=10^6`$ K and the density $`\rho =1.67\times 10^{24}`$ g cm<sup>-3</sup>. This corresponds to mass influx of HVCs onto the disk of $`24\text{ }M_{}`$ yr<sup>-1</sup>. A result in accordance with the estimates carried out by Wakker (1990). The chimney models neglected the presence of a dynamical thick disk which would provide constraints to the structure of the chimneys. The formation of the chimneys was carried out in a medium where no other phenomenon was present. Furthermore, the two-dimensional calculations impose natural limitations on the evolution of the gas and clouds. The absence of a third dimension constrains the motion to a vertical plane perpendicular to the Galactic disk. If three-dimensional evolution is considered, the overall structure of the flow may suffer modifications that lead to the generation of phenomena that may not be identified in two-dimensions. The principal objective of the research reported here has been to develop a three-dimensional model that can account for the collective effects of type Ib, Ic and II supernovae on the structure of the interstellar medium in the galactic disk and halo. This model should, accordingly, account for the formation of the major features already observed in the Galaxy; these include the Lockman and Reynolds layers, large scale outflows, chimneys, and HI clouds, features that all contribute to the overall disk-halo-disk cycle known as the Galactic fountain. Section 2 deals with numerical modelling model where the model of the Galactic disk, rate of supernovae , boundary and initial conditions and numerical scheme are presented. Section 3 presents the evolution of the simulations, comprising the global evolution, the description of the simulations as they reach the steady state and the distribution of cold gas in the simulated disk. In section 4, a discussion on the simulations and the major predictions is carried out, and a comparison with existing observations is made. Finally, section 5 presents the conclusions and future prospects. ## 2 Numerical Modelling ### 2.1 Model of the Galaxy The study of the evolution of the Galactic disk rests in the realization that the Milky Way has a thin and a thick disk of gas in addition to a stellar disk. The thin gas disk has a characteristic thickness comparable to that of the stellar disk of Population I stars. The thick gas disk is composed of warm neutral and ionized gases with different scale heights - 500 pc (Lockman et al., 1986) and 950 pc (Reynolds, 1987), respectively. The stellar disk has a half thickness of 100 pc and the vertical mass distribution, $`\rho _{}`$, inferred from the star kinematics, given by (Avillez et al., 1998), $$\rho _{}=\rho _,sech^2\left[\left(2\pi G\beta _{}\rho _{}\right)^{1/2}z\right]$$ (1) where $`z`$ varies between $`100`$ pc and $`100`$ pc, $`\rho _,=3.0\times 10^{24}`$ g cm<sup>-3</sup> is the mass density contributed by Population I stars near the Galactic plane (Allen, 1991) and the constant $`\beta _{}=1.9\times 10^{13}`$ cm<sup>-2</sup> s<sup>2</sup>. This mass distribution generates a local gravitational potential, $`\mathrm{\Phi }`$, of the form $$\mathrm{\Phi }=\frac{2}{\beta _{}}\mathrm{ln}\mathrm{cosh}\left[\left(2\pi G\rho _{}\beta _{}\right)^{1/2}z\right].$$ (2) The stellar disk is populated by supernovae types Ib, Ic and II, whose major fraction occurs along the spiral arms of the Galaxy close to or in HII regions (Porter & Filippenko, 1987). Supernovae types Ib and Ic have progenitors with masses $`M15\text{ }M_{}`$, whereas Type II SNe originate from early B-type precursors with $`9\text{ }M_{}M15\text{ }M_{}`$. The rates of occurrence of supernovae types Ib+Ic and II in the Galaxy are $`2\times 10^3`$ yr<sup>-1</sup> and $`1.2\times 10^2`$ yr<sup>-1</sup> respectively (Cappellaro et al., 1997). Similar rates have been found by Evans et al. (1989) in a survey of 748 Shapley-Ames galaxies. The total rate of these supernovae in the Galaxy is $`1.4\times 10^2`$ yr<sup>-1</sup>. Sixty percent of these supernovae occur in OB associations, whereas the remaining $`40\%`$ are isolated events (Cowie et al., 1979). ### 2.2 Basic Equations and Numerical Methods The evolution of the disk gas is described by the equations of conservation of mass, momentum and energy: $$\frac{\rho }{t}+\left(\rho 𝐯\right)=0;$$ (3) $$\frac{\left(\rho 𝐯\right)}{t}+\left(\rho \mathrm{𝐯𝐯}\right)=p\rho \mathrm{\Phi };$$ (4) $$\frac{\left(\rho e\right)}{t}+\left(\rho e𝐯\right)=p𝐯n^2\mathrm{\Lambda };$$ (5) where $`\rho `$, $`p`$, $`e`$, and $`𝐯`$ are the mass density, pressure and specific energy and velocity of the gas, respectively. The set of equations is complete with the equation of state of ideal gases. $`\mathrm{\Lambda }`$ is the functional approximation of the cooling functions of Dalgarno & McCray (1972), Raymond et al. (1976) and the isochoric curves of Shapiro & Moore (1976), except in the range of temperatures of $`10^55\times 10^6`$ K where the simple power law of the temperature (Kahn, 1976) $$\mathrm{\Lambda }=1.3\times 10^{19}T^{0.5}\text{erg cm}\text{3}\text{ s}\text{-1}$$ (6) is applied. Between $`5\times 10^6`$ and $`5\times 10^7`$ the cooling function has been approximated by $`T^{0.333}`$ (Dorfi, 1997). In order to avoid any cooling below zero temperature, $`\mathrm{\Lambda }`$ is zero below 200 K. The equations of evolution are solved by means of a three-dimensional hydrodynamical scheme using the piecewise parabolic method (Collela & Woodward, 1984) and the adapted mesh refinement algorithm of Berger & Collela (1989). Three levels of refinement are used. Each is created with a refinement factor of two. The maximum resolution obtained with this method is 1.25 pc. ### 2.3 Computational Domain and boundary Conditions The simulations were carried out using a cartesian grid centred on the Galactic plane with an area of 1 kpc<sup>2</sup> and extending from -4 kpc to 4 kpc. Grid resolutions of 5 pc and 10 pc were used for $`270z270`$ pc and for $`z250`$ pc and $`z+250`$ pc, respectively. The cells located at $`270z250`$ pc and $`250z270`$ pc of the coarser and finer grids overlap to ensure continuity between the two grids. The solution in the finer grid, at ranges $`270z250`$ pc and $`250z270`$ pc, results from a conservative interpolation from the data in the coarser cells to the fine grid cells as prescribed by Berger & Collela (1989). The boundary conditions along the vertical axis are periodic; outflow boundary conditions are used in the upper and lower parts of the grid parallel to the Galactic plane. ### 2.4 Initial Conditions The initial configuration of the system in the computational domain considers the vertical distribution of the disk and halo gases in accordance with the Dickey & Lockman (1990) profile, and reproduces the distribution of the thin and thick disk gas in the solar neighborhood. The gas is initially in hydrostatic equilibrium with the gravitational field, thus requiring an initial gas temperature as defined by the density distribution. ### 2.5 Supernovae in the Simulated Disk Throughout the simulations the presence of the stellar disk is required with a thickness of 200 pc and an area of 1 kpc<sup>2</sup>, corresponding to a volume of $`V=2\times 10^8`$ pc<sup>3</sup>. Isolated and clustered supernovae of types Ib, Ic and II occur in the volume $`V`$ at rates obtained from the values discussed in §2.1 and given by $$\sigma \frac{V}{V_G}=1.42\times 10^3\sigma \text{ yr}\text{-1}$$ (7) where $`V_G=1.4\times 10^{11}`$ pc<sup>3</sup> is the volume of the Galactic disk with a radius of 15 kpc and a thickness of 200 pc, and $`\sigma `$ is the rate of supernovae observed in the volume $`V_G`$. Table 1 presents the rates of supernovae in the Galaxy as well as in the simulated disk. In the simulated stellar disk, isolated and clustered supernovae form at time intervals of $`1.26\times 10^5`$ yr and $`8.4\times 10^4`$ yr respectively. The time interval between two sucessive superbubbles in the volume $`V`$ varies between $`1.9\times 10^6`$ yr and $`5.23\times 10^6`$ yr (Ferrière, 1995), but following Norman and Ikeuchi (1989), and adopting an average value of $`N_{SN}=30`$ supernovae per superbubble, the time interval of formation of superbubbles in the volume $`V`$ is $$\frac{N_{SN}}{\sigma _{OB}}=2.5\times 10^6\text{yr},$$ (8) the value used in this study. Each supernova is set up in the beginning of phase II, with a thermal energy content of (Kahn, 1975) $$E_{Therm}=0.36\rho _{}a$$ (9) and a kinetic energy content of $$E_{Kin}=0.138\rho _{}a,$$ (10) where $`a=2E/\rho _{}`$, $`\rho _{}`$ is the local density of the medium where the supernova occurs, and $`E`$ is the energy of the explosion. The generation of supernovae in the stellar disk is carried out in a semi-random way, where the coordinates of the supernova are determined from three new random numbers converted to the dimensions of the stellar disk. After the determination of the random numbers, and therefore, the location of the new supernova in the grid, several tests are carried out in order to determine the type of supernova whether it occurs isolated in space or within an OB association, or even to determine if the new location must be dismissed. The location of a new supernova in the grid is based on the following constraints: (a) vertical density distribution of population I stars given by equation (1), (b) type of new supernova, (c) rates of occurrence of supernovae types Ib+Ic and II and (d) rate of formation of superbubbles. When a supernova is due to appear in a specific location, a supernova generator checks for the occurrence of a previous supernova in that location. If this test is negative, then a new supernova is set up, providing its generation is compatible with the time interval between the previous supernova of the same type. If the test is positive, then the time delay between the old supernova and the new one is checked. Now the decision is based on the comparison of the this time delay and that of a new supernova within a superbubble. If the delay is smaller than the time delay within the superbubble, then no supernova is generated and the generator chooses a new position from a new set of random numbers, otherwise the supernova is set up. When the type of the new supernova is known, the algorithm determines the radius of the supernova in the beginning of phase II. This is acomplished by equating the amount of mass of the progenitor star released during its explosion and the mass of the ISM engulfed by the blast wave. The mass of the progenitor star is a ramdom value determined from the interval of masses characteristic to the progenitors of that type of supernova, e.g. if the next supernova to occur is a type Ib, then the amount of mass that is released during the explosion is found within the interval $`[15,30]`$ $`M_{}`$. ## 3 Evolution of the Simulations The simulations were carried out for a period of 1 Gyr with supernovae being generated at time 0. The simulations start from a state of hydrostatic equilibrium breaking up during the first stages of the simulation. The system evolves into a statistical steady equilibrium where the overall structure of the ISM seems similar on the global scale. ### 3.1 Global Evolution During the first 50 Myr of simulated time there is an imbalance between cooling and heating of the gas in the disk. Initially there is an excess of radiative cooling over heating because of the small number of supernovae during this period. The gas originally located in the lower halo starts cooling and moves ballistically towards the midplane, colliding there with gas falling from the opposite side of the plane (Figure 1). After 15 Myr of evolution, the major fraction of the initial mass is confined to a slab having a characteristic thickness of 100-150 pc (Figure 1 (d)). A thin wiggly disk of cold gas forms at the midplane and has a thickness of few tens of parsec. As the supernovae occur, they warm up the gas in the slab, gaining enough energy to overcome the gravitational pull of the disk and, therefore, expanding upwards (Figure 1 (e) and (f)) redistributing matter and energy in the computational domain. After the over-expansion, a dynamic equilibrium is set up between upward and downward flowing gas (Figure 2). During the first 10-20 Myr, the descending gas has a maximum value of $`3.4\times 10^{19}`$ g cm<sup>-2</sup> s<sup>-1</sup> (that is $`4.8\times 10^2`$ $`M_{}`$ kpc<sup>-2</sup> yr<sup>-1</sup>). The ascending gas massflux peaks at $`2.8\times 10^{19}`$ g cm<sup>-2</sup> s<sup>-1</sup> ($`3.9\times 10^2`$ $`M_{}`$ kpc<sup>-2</sup> yr<sup>-1</sup>). The expansion of the disk gas diminishes during the next 20 Myr, with the consequent decrease in the total massflux. Between 60 and 75 Myr, a balance between descending and ascending flows sets in. Both the ascending and descending gases have massfluxes of approximately $`3\times 10^{20}`$ g cm<sup>-2</sup> s<sup>-1</sup> ($`4.2\times 10^3`$ $`M_{}`$ kpc<sup>-2</sup> yr<sup>-1</sup>), which corresponds to a total inflow rate of 2.97 $`M_{}`$ yr<sup>-1</sup> in the Galaxy on one side of the Galactic plane. Such a dynamic equilibrium is observed during the rest of the simulations, indicating that a quasi-equilibrium between the descending and ascending gases dominates the evolution of the disk-halo gas. However, there are some periods when one of the flows dominates over the other. This is related to periodic imbalances between cooling and heating leading to compressions and expansions of the disk gas. As the balance between descending and ascending flows is reached, the disk evolves into a statistical steady equilibrium where, in general, the disk gas is composed of a multiphase medium extending into the halo, forming a frothy region laced by sheets and crossed by HI clouds and well collimated chimneys as can be seen in Figure 3. The figure presents the vertical distribution of the density up to 1 kpc on either side of the plane. The disk gas shows a stratified distribution where neutral and ionized phases co-exist having different scale-heights. The neutral phase is composed by cold ($`T10^3`$ K) and warm ($`10^3<T<10^4`$ K) gas whereas the ionized phase is composed by a warmer gas with $`10^4T<10^5`$ K (hereafter identified as warm ionized gas) and hot gas ( $`T10^5`$ K). The neutral component is mainly distributed in the layer $`\left|z\right|500`$ pc, forming a thick HI disk and having an average temperature of $`10^4`$ K. Cold gas is present, predominantly in a thin but irregular layer (thin HI disk) around the Galactic plane, having a characteristic thickness of tens of parsec (Figures 3 and 4), whereas the hot gas is mainly distributed above 1.5 kpc. Gas with temperatures varying between $`10^4`$ and $`10^5`$ K fill up to $`80\%`$ of the disk volume, whereas gas with temperatures smaller than $`10^4`$ K, fills on average (Figure 5), $`50\%`$ of the disk volume. The $`T10^4`$ gas is mainly located in a layer with half thickness of 500 pc ($`|z|500`$ pc) (see also Figure 3), although some of this gas extends to 1.5 kpc, but with a very small volume filling factor. The warm ionized gas dominates the region between 500 pc and 1.5 kpc, although its presence is detected up to 2.5 kpc. Gas with temperatures larger than $`10^5`$ K fill on average $`1523\%`$ of the disk volume below 1 kpc. With the increase of $`z`$, the volume occupied by the hot gas increases, becoming the dominant medium at $`z1.5`$ kpc (Figure 5). The density distribution of the disk gas decreases with $`z`$. At $`\left|z\right|200`$ pc it can be approximated by a gaussian followed by an exponential distribution (Figure 4). The gaussian is overshot by the exponential distribution with a smooth decrease followed by a steep increase in the density profile (Figure 4) on either side of the plane. This suggests the presence of an interface between the thin disk and the thick HI disk. Such an interface is fed by the hot gas flowing upwards from the thin disk in large scale outflows ($`l200`$ pc or more) interacting with the cooler thick disk gas overlaying it. Such a configuration becomes Rayleigh-Taylor unstable, and therefore the upward flow acquires a finger-like structure during its expansion through the thick disk. This density distribution gives the appearance of a stratified distribution which is compatible to the volume filling factors of the different phases of the ISM at different $`z`$ (see Figure 5). The ionized component of the disk gas extends upwards up to $`z\pm 1.5`$ kpc. Its density decreases smoothly up to 1.4-1.5 kpc, where it has a steep decrease. This region acts as an interface between the thick disk and the halo gas. For $`z>1.5`$ kpc the gas has lower densities and temperatures $`T10^6`$ K. ### 3.2 Statistical Steady Evolution The disk gas shows a steady equilibrium where its structure and dynamics are similar throughout the rest of the simulations. That is, the disk has the same structure at different times and locations, although it experiences changes due to local events such as supernovae, superbubbles, chimneys, collisions with HI clouds, shock waves, etc. In order to stress this point, images showing the time evolution of the disk gas at different locations are presented in Figures 6-11. The figures were taken in directions perpendicular and parallel to the midplane at 500 and 800 Myr of evolution time. The $`xz`$slices (Figures 6-9) were taken at 30, 220, 410, 600, 790 and 980 pc from the edge of the computational domain, whereas the $`xy`$slices (Figures 10-11) were taken at $`z=250`$, $`0`$ and $`250`$ pc. The density distributions of the disk gas shown in Figures 6, 8 and 10 are complemented with the corresponding temperature maps in Figures 7, 9 and 11. The figures show a large number of features comprising, among others, HI clouds and sheets, bubbles and superbubbles, interacting supernovae forming networks of hot gas, large regions of hot gas crossing the $`xy`$slices at $`z=250`$ and 250 pc, etc. These features can be summarised as follows: * A wiggly thin and a cold disk, having a characteristic thickness varying between a few parsec and tens of parsec (Figures 6-9). There are regions where the thin disk may have disappeared as result of a large number of supernovae in the neighborhood as shown in Figure 8 (a) at $`200x400`$ pc. The region is crossed by a stream of hot gas as can be seen in the corresponding temperature map. * Vertical structures connected to the underlying thin disk and appearing to wiggle around like worms. These structures are the debris of broken shells and supershells, that occurr displaced from the midplane. * Sheets of cold gas distributed in the disk that result from the breakup of larger structures (Figures 6 \- 11). The sheets have thicknesses of a few pc and widths of several tens of parsec, even hundreds of parsec. The sheets observed at $`z=\pm 250`$ pc seem to connect to each other, forming a “network” of cold gas. In the plane, the cold gas forms large regions of very high density and low temperature (Figure 10 (b) the darkest blue regions). * Small cloudlets in the vicinity of worms and other sheet-like structures. The smallest cloudlets detected in the images have 1.25 pc size - this corresponds to the maximum resolution used in the disk. The cloudlets result from the breakup of the worms and other clouds which have been swept up by blast waves, involved in cloud-cloud collisions, etc. * Supernovae and superbubbles at different phases of their evolution surrounded by shells with variable thicknesses. Supernovae and superbubbles located above the thin gas disk show an elongated appearance as result of the density gradient along the $`z`$direction (Figures 6-8). * Networks of hot gas connecting supernovae. These can be observed in both the $`xz`$slices and $`xy`$slices of the temperature distribution of the disk gas. The network of hot gas is not confined to the plane ($`z=0`$), but spreads in the volume of the disk between $`z=250`$ and $`250`$ pc. * Reservoirs of hot gas (large regions of hot gas) moving away from the plane (Figures 7, 8 and 11). In general the disk gas is pervaded by isolated supernovae and superbubbles (large supernova structures generated by sucessive supernovae). Supernovae change the local structure of the interstellar medium, but are unable to change the global structure. Individual and clustered supernovae dominate the local environment, depending on their spatial location. Isolated supernovae change the structure of the inner parts of the disk whereas supernovae in OB associations dominate the upper regions of the disk. As the blast waves from supernovae expand in the disk, they sweep up clouds and sheets, triggering their disruption into smaller structures. At $`z=\pm 250`$ pc the presence of sheets results from: (a) cold gas descending from above and acquiring sheet-like forms as a result of Rayleigh-Taylor instabilities that occurr in larger clouds and (b) the breaking up of shells and supershells that expand upwards and are displaced from the Galactic plane. Young supernovae expand into a highly unstructured medium interacting with other supernovae or cold gas in the form of sheets or clouds (Figures 6-11). During their expansion, the remnants are preceeded by shock waves engulfing the shells of other supernovae and triggering their disruption. If the average shock speed in the shell is $`v_s1\times 10^4`$ km s<sup>-1</sup>, then the transmitted shock takes a time $$t_{cross}\frac{\mathrm{\Delta }R}{v_s}$$ (11) to cross the old shell, $`\mathrm{\Delta }R`$ being the thickness of the shell normally 5 to 10 pc thick. Thus, the blast wave crosses the shell in $`t_{cross}10^4`$ yr, generating a large pressure gradient which is enough to disrupt the old shell. As a consequence, the hot gas flows freely between the remnants forming a network of channels of low density gas. On average, after $`5\times 10^5`$ years, a type II supernova releases $`295\text{ }M_{}`$ of hot gas into the surrounding medium. Therefore, when several isolated supernovae merge together, they contribute to the formation of “reservoirs” of hot gas with enough energy to overcome the gravitational pull of the stellar disk expanding buoyantly upwards (Figures 5, 6 \- 9). In its ascending motion, the hot gas interacts with the denser gas distributed in the thick disk. Such a configuration is Rayleigh-Taylor unstable and as a consequence, the hot gas acquires a finger-like structure with a mushroom cap. As it moves upwards it appears to carve the cooler layers that compose the thick disk (Figure 3). Bubbles and superbubbles occurring in the outer parts of the thin disk, at some $`100`$ pc above and below the Galactic plane, expand vertically due to the local stratification of the ISM becoming elliptical (Figures 6 and 8). The vertical elongation of the shell surrounding the bubbles/superbubbles depends not only on the local medium, but also on the amount of energy released by the supernova explosions involved in the production of the bubble/superbubble. During their vertical expansion the shells, eventually become Rayleigh-Taylor unstable and blow their caps, releasing their hot inner parts into the ISM through narrow channels. The vertical expansion of the hot inner parts of the bubbles is halted by pressure effects due to the thick disk gas, while superbubbles break through it, blowing holes in the thick disk gas. The presence of such structures is shown in Figures 3 and 13-13. These structures may be classified as mini-chimneys and chimneys depending on their vertical elongation and on the presence of actual tunnels breaking through the disk gas as can be seen in Figure 3 (c) and (d). The tunnel observed in the figure crosses the thick gas disk and has a characteristic width of approximately 120 pc. The tunnel maintains the same appearance providing enough supernovae occur in the underlying association. The expansion of hot gas in well collimated structures may not have enough energy to break through the gas disk (Figure 3 (a) and (b)). Such a structure may be classified as a mini-chimney. The walls of the tunnels are formed by the remains of the broken shells whose appearance resembles worms crawling out of the disk. The average width of mini-chimneys varies between 50 and 60 pc whereas that of chimneys ranges from 100 pc to a few hundred parsec, depending on the number of supernova explosions inside the superbubble. Figure 13 shows the sequential evolution of the disk gas in a region where a chimney is formed, while Figure 13 presents the HI positional maps of the same region. The chimney, formed from a superbubble with a vertical elongation of 200 pc and a width of 100 pc, is located in the region $`350x450`$ pc in the southern hemisphere (hereafter called region A). The base of the superbubble is located in the midplane and connects with two merging bubbles located in the northern hemisphere at $`300x480`$ pc and $`40z150`$ pc. After a million years, these two bubbles merge, creating a large cavity extending from $`350x480`$ pc and $`50z190`$ pc (Figures 13 and 13 (b)) and hereafter named region B. At time 203 Myr, a supernova with radius of some 50 pc (the supernova is well inside phase III of its evolution) is observed to occur at $`x=400`$ pc, $`z=100`$ pc, inside region A. similar explosions occurred in the superbubble leading to its expansion and break up of the top of its shell (Figure 13 (b). This is corroborated by the HI emission map in Figure 13 (b). A narrow tunnel of low density gas is surrounded by thick walls (thicknesses of a few parsec). The top of the chimney has a width of some 80-90 pc. The hot inner parts of the superbubble are released upwards through sucessive injections due to different supernovae occurrences inside the cavity. As the gas expands out of the superbubble, it pushes upwards from $`z=300`$ to $`z=450`$ pc, debris of old shells in a period of 200 Myr. Continuing explosions inside the superbubble provoke the escaping of hot gas not only in the southern hemisphere, but also inject gas into region B (Figure 13 (c)). As a consequence the upper parts of the shell surrounding region B also breaks up and an injection of hot gas into the thick disk through a mini-chimney is induced. The gas released by the mini-chimney is unable to break through the thick disk and acquires a finger-like configuration until it cools down to a recombination temperature and merges together with the surrounding medium (Figure 13). The figure shows the evolution of the same region in space as Figures 13 and 13, but at later times: 209-211 Myr. After a period of a few million years, the width of region A varies between 200 pc near the plane to 110 pc on the upper parts of the shell (Figure 13 (c)). ### 3.3 Cold Gas in the Disk HI emission maps showing the vertical distribution of the cold gas in the disk and the corresponding temperature maps, taken at along the plane crossing the midplane at $`y=220`$ pc, are presented Figures 15-15. The figures present a time evolution of the cold gas in the disk during a period of four million years and show 20 equally spaced contours representing values of column density between $`\mathrm{log}(N_{HI})=18`$ and $`\mathrm{log}(N_{HI})=20`$, where $`N_{HI}`$ is in cm<sup>-2</sup>. The concentration of contours in a specific direction indicates a large column density of the gas there. In general the figures show HI emission concentrated in a wiggly cold structure (thin HI disk) distributed along the Galactic plane, having a thickness of a few tens of parsec, and in some cases being as thin as 10-20 pc. The presence of a very thin HI disk (with $`\mathrm{\Delta }z1020`$ pc) is an occasional phenomenon and results from local variations in the structure of the thin disk as result of, for example, correlated supernovae located near the plane or collisions of high velocity clouds with the thin disk. The occurrence of the supernovae leads to the formation of tunnels of hot gas crossing the Galactic plane, and therefore, to the absence of the thin disk observed in some of the $`xz`$-maps (see Figure 9). Collisions of high velocity clouds, may compress and drag with the cloud parts of the thin disk. The largest column density observed in the images is caused by the low scale height of the cold gas. The figures show the presence of low column density regions (absence of contours) persisting for long periods. These regions result from supernovae that occurred some time in the past, and that have blown holes in the Galactic disk. As the expanding gas cools, a thin shell of colder gas is formed behind the blast-wave generated by the supernova. The vertical structure of the HI holes is observed in most of the figures. Some are partially surrounded by thin sheets of cold gas, which result from old shells. The cone-like structures of low column density surrounded by cold sheets are found located on each side of the Galactic plane. Because of the high column density in the plane, the cones slightly change its local structure, but are unable to break the thin disk. Shells displaced from the midplane have a non-spherical appearance, resulting from the local medium not being smooth (on the small scale - 5 pc) and partly from the steep variation of the density along the $`z`$direction - the shells acquire a ovoid form (Figures 13, 15 \- 15). As the bubbles expand upwards the shells break, releasing the hot inner parts of the remnant into the surrounding medium. Therefore, the remaining parts of the shells become the walls of the cone-like structures. These walls have a wiggly appearance resembling “worms” crawling into the halo. Face-on maps of HI emission from the midplane, taken in a six million years period (Figure 16), show high column density regions spread over the plane surrounding holes and channels of low column density gas. The holes and tunnels form a large scale network of low density gas spread over the Galactic disk. The tunnels have widths of some 100 pc and lengths in the range of 200-400 pc. This network forms as a result of young supernovae occurring near to old remnants. High velocity shock waves sweep up the old remnants, disrupting their shells and leading to a movement of the hot gas inside the young remnants into the low density medium surrounding it. The blast wave may suffer several reflections inside the tunnel, leading to the formation of small clouds as result of the disruption of the shells and condensation into clouds of the gas swept several times by the blast waves. This gas is compressed during each swept up to four times the local density (assuming that these shocks are strong). The increase in the density leads to an increase in the cooling of the material and therefore to its condensation into clouds. Most of the gas observed in the HI emission maps forms clouds with sheet like appearance and connected to each other. Clouds have sizes/thicknesses varying between few parsec to hundreds of parsec. The larger clouds suffer Rayleigh-Taylor instabilities and break up into sheet-like structures that fall towards the Galactic plane. However, isolated structures with large sizes, identified near the plane are moving towards it colliding with it. The collision leads to the compression of the thin disk where higher density gas accumulates forming possible giant molecular clouds where star formation would occur. Column density profiles of the HI gas along a column of 400 pc (with $`200z200`$ pc) versus velocity are presented in Figure 17. The figure presents profiles convolved with a Gaussian that has a velocity dispersion that corresponds to the local gas temperature, being each cell of the column normalized to the local gas density. The profiles are symmetric with respect to the central peak and have FWHM values varying between 6.1 and 10.5 km s<sup>-1</sup>, being these the extreme cases. The average values of FWHM are $`9.5`$ km/s. The major fraction of cold gas in the disk has positive and negative velocities varying between $`20`$ and $`20`$ km/s. However, there are clouds with larger velocities indicating a fast approach to the midplane, where they collide with cold gas falling from the opposite hemisphere, or an escape from the thin disk (positive velocities). ## 4 Discussion The results of large-scale modelling of a section of a Galactic arm, located at 8.5 kpc from the Galactic centre are presented in this paper. It is a three-dimensional model that assumes the ISM as an integrated system where the collective effects of supernovae (isolated and clustered) are taken into account. The supernovae occur in a medium strongly structured by previous supernovae, ionization fronts and radiative cooling. The model considers the interstellar gas distributed in a smooth thin disk with the vertical distribution of the cool and warm neutral HI gas given by Dickey & Lockman (1990) and warm ionized gas given by Reynolds (1987). It expli- citly includes both correlated and random supernovae in a manner compatible with observations. Sixty percent of the supernovae occur within associations, whereas the remaining occur isolated. The supernovae belong to types Ib, Ic and II, and their masses are taken into account, as mass loading is allowed. Once disrupted by the explosions, the disk never returns to its initial state. Instead, it approaches a state where a thin HI disk is formed in the Galactic plane, overlayed by a thick disk of warm gas. The thick disk is generated as long as enough supernovae occur in the disk, regardless of the initial density profile adopted for the disk gas (§3.1), and, therefore, its formation and stability are directly correlated to the supernova rate per unit area in the simulated disk. The density distribution of the disk gas in $`z`$ can be approximated by a linear combination of two gaussians and an exponential with the general form, based on the model of Dickey & Lockman (1990) $$n(z)=\frac{n_{}}{0.566}\left[Ae^{Z^2/2B^2}+Ce^{Z^2/2D^2}+Ee^{Z/F}\right]$$ (12) where $`n`$ is the number density of the ISM phase, $`n_{}`$ is the number density at the Galactic plane, $`A`$, $`C`$ and $`E`$ are paremeters written in units cm<sup>-3</sup> and $`B`$, $`D`$ and $`F`$ are written in pc. The density distribution yields an effective scale-height $$H_z=\frac{1}{n_{}}_0^+\mathrm{}n(z)𝑑z.$$ (13) Noting that $$_0^+\mathrm{}e^{Z^2}𝑑Z=\frac{\sqrt{\pi }}{2},$$ (14) and using (13) the effective scale-height of the density distribution in equation (12) is given by $$H_z=\frac{1}{0.566}\left[(AB+CD)\sqrt{\frac{\pi }{2}}+EF\right].$$ (15) For the density distribution of the disk gas in the Galaxy, Dickey & Lockman (1990, hereafter denoted by the DL distribution) inferred the parameters shown in the first row of Table 3. The table presents, in rows 2-5, the parameters corresponding to density distributions measured along four lines of sight crossing the disk at: (a) $`x=30`$, $`y=30`$ pc, (b) $`x=300`$, $`y=250`$ pc, (c) $`x=600`$, $`y=500`$ pc and (d) $`x=800`$, $`y=800`$ pc. These distributions are the best fits to the underlying curves of the density distribution as shown in Figure 18. As can be seen in Table 3, the coefficients $`C,D,E`$ and $`F`$ are similar between all the fits and the DL distribution, whereas the $`A`$ and $`B`$ parameters for the fits and for the DL distribution have variations as much as $`15\%`$. The difference in the extreme case indicates that, in the simulations, most of the cold gas in the disk is concentrated near the plane and thus has a smaller effective scale-height. Despite the differences between the DL distribution and those obtained in the simulations, the thick gas disk has a distribution compatible with the presence of two gas components having different scale heights: a neutral layer with a scale height of 500 pc (warm HI disk) and an ionized component extending to a height of 1 to 1.5 kpc above the thin HI disk. Similar distribution of the thick gas disk and the HI and HII thick disks has been observed in the Galaxy (Lockman, 1984; Lockman et al., 1986; Reynolds 1987) and in external galaxies such as NGC 891 (Dettmar, 1992). Studies of the z-distribution of HI in edge-on galaxies (NGC 4565 and NGC 891) carried out by Rupen (1991) have shown the presence of an HI layer around the galactic midplane. This layer has a thickness that varies from 120 pc to 300 pc in both galaxies. The average physical properties of the different phases in the simulated disk in the steady state evolution are summarized in Table 3. These values are in agreement with those derived by Karberla & Kerp (1999) for the best-fit parameters required for a statistical steady equilibrium of the halo and disk gas in the Galaxy. ### 4.1 Volume Filling Factors of ISM Phases The volume filling factors of the different ISM phases have been presented in §3.1 and the corresponding best-fit profiles are presented in Figure 5. The distribution of the hot gas varies from a occupation volume of some $`23\%`$ in the inner disk (below 500 pc) to $`100\%`$ above 2.5 kpc. The thin disk is populated by a cold medium with a volume-filling factor smaller than $`5\%`$. The warm neutral gas dominates below 500 pc occupying a volume of some $`50\%`$, whereas the warm ionized medium dominates between 500 and $`1500`$ pc. At some 1500 pc, both the warm ionized and hot medium fill the same volume. This region corresponds to the disk-halo interface described above. In what follows, we shall consider the disk-halo interface as the height at which the two media have the same volume-filling factor. As the warm neutral and ionized media dominate different regions, it gives the appearance that each of these phases constitute different layers. However, both media are mixed and have different scale-heights. The region dominated by the ionized medium is known as the Reynold’s layer (Reynolds, 1987). The filling factors of the different phases observed in this study are similar to those estimated by Spitzer (1990) and Ferrière (1995). These authors estimated a volume-filling factor of the hot gas in the Galactic disk of $`0.2`$. In addition, Ferrière (1995) estimated volume-filling factors for the cold and neutral gases in the disk similar to those shown in Figure 5 and discussed above. Ferrière (1998) estimated the variation of the volume-filling factor of the hot gas with $`z`$. She finds that, at the solar circle, the volume occupation of the hot gas increases from 0.15 at $`z=0`$ pc to 0.23 at $`z=200`$ pc and decreases gradually for $`z>200`$ pc (see Figure 12 in Ferrière, 1998). The latter result seems to be in conflict with the calculations reported in this paper. Such apparent conflict is resolved if one considers that the calculations carried out by Ferrière reflect the distribution of hot gas inside cavities of supernova remnants and superbubbles, whose distribution decreases with $`z`$, while the present paper also considers hot gas that bouyantly rose from interacting supershells. Ferrière’s model does not include that gas at all, and so does not model the high-$`z`$ region fully. ### 4.2 Large Scale Outflows and Local Outbursts The simulations show that the thick gas disk is fed with the hot gas coming from the thin disk through two major processes: large scale outflows and local outbursts (known as chimneys). Large scale outflows are the result of the buoyant expansion of hot disk gas concentrated in large reservoirs located on either side of the thin HI disk (as shown in the density profiles). These reservoirs are formed by the hot inner parts of isolated supernovae randomly distributed in the stellar disk. Such hot gas has enough energy to be held gravitationally and expands buoyantly upwards, generating large scale outflows (Figure 3). The ascending flow triggers the growth of Rayleigh-Taylor instabilities as it interacts with the cooler, denser medium in the thick disk. Therefore, the hot gas acquires a finger-like structure with a mushroom cap on the tip of the finger. During the expansion of the finger, further instabilities develop, until the major fraction of the ascending gas cools and merges with the surrounding medium. The gas in the finger cools from the outside towards the centre, thus giving it the appearance of being enveloped by a thin sheet of cooler gas (Figures 3 and 13 at $`x=300,y=100`$ pc). A structure called the “anchor”, with properties similar to those described above, has been observed in the southern Galactic hemisphere and reported by Normandeau & Basu (1998). Chimneys result from correlated supernovae, which generate superbubbles that acquire an elongated shape due to the stratification of the ISM in the $`z`$direction. Superbubbles are able to expand and blow holes in the disk, injecting high speed gas, which forces its way out through relatively narrow channels with widths of some 100 - 150 pc. They provide a connection between the thin disk and the upper parts of the Reynolds layer. One can identify mini-chimneys and chimneys from the break up of shells and supershells located at the upper parts of the thin disk. Mini-chimneys result from isolated supernovae that have had their shells broken and released to hot gas into the surrounding medium (the broken shell acts as the walls of a small tunnel produced by the blow out), whereas chimneys are large tunnels of hot gas crossing the thick HI disk. These structures are well collimated, with a maximum width varying between 50 pc for mini-chimneys and 150 pc for chimneys, and vertical extensions varying between few tens of parsec (mini-chimneys) and 500 pc or more for the chimneys. This is corroborated by the HI position maps, taken at different locations of the disk along lines of sight parallel to the plane, showing the presence of broken shells (“worms”) wiggling around in the $`z`$-direction. Face-on HI maps of the disk taken at $`\left|z\right|250`$ pc show the presence of low density regions with elliptical or circular forms and having a large range of sizes. The major fraction of these holes have been produced by large cavities of hot gas resulting from clustered supernovae, and therefore by chimneys. It is hard to identify the holes resulting from mini-chimneys, suggesting that their maximum elongation is some 150 pc in the $`z`$-direction. Surveys of the Galactic disk (Heiles, 1984, 1990) and of M31 (Brinks & Bajaja, 1986) and M33 (Deul & Hartog, 1990) have revealed the presence of roughly elliptical holes in the HI maps, with sizes varying between 40 and 1000 pc. Some holes show a clear shell structure, although for most holes the shells are not resolved. The majority of the shells have deviations in the local velocity field with expansion velocities of 10-30 km s<sup>-1</sup>. The estimated kinematic ages of these holes varies from 2.6 to 30 Myr and the energy requirements are $`10^{49}`$ to $`10^{53}`$ erg. A correlation between the holes in the HI maps with single HII regions and OB associations in these galaxies has been established by Deul & Hartog (1990). Heiles (1984) identified the presence of holes in the HI emission maps of the Galactic disk. In addition, sheets of HI gas with a vertical structure have been identified in these maps by Heiles and Koo et al. (1992). These vertical structures have been associated with broken shells and supershells in the Galaxy, and have been proposed as being the walls of chimneys by Tomisaka & Ikeuchi (1986) and Norman & Ikeuchi (1989). HI maps of NGC 891 show a series of low density channels, similar to the ones shown in this paper. Most recently, Normandeau et al. (1996) identified a superbubble in the Perseus arm underlying an association of nine O-type stars. The superbubble has been created by stellar winds from these stars (Basu et al., 1999) - no supernova has been identified within the W4 region - and has an elongated appearance resulting from the vertical stratification of the disk gas. Although in HI emission maps the region looks like a chimney, such a view is altered in the H$`\alpha `$ emission maps of the region, which clearly shows the presence of a superbubble. ### 4.3 Numerical Resolution Numerical resolution seems to be an issue in the simulations reported in this paper and in Paper II. The major constraint imposed by numerical resolution is the prevention of the formation of very small scale structures resulting from thermal instabilities and condensations due to radiative cooling. Furthermore, the amount of cooling may be substantially increased if the resolution is great enough to allow for the development of turbulent shear layers in regions of outflow and infall. The numerical approach adopted to minimize such problems combined the usage of a high-resolution scheme (the PPM) to advance the solution in time with the dynamical refinement of the computational grid whenever required. The latter was set up in the region located at $`150z250`$ pc (hereafter referred to disk-grid), whereas in the rest of the computational domain no dynamical refinement was used - the maximum resolution being 10 pc. The resolutions in the disk-grid were 5, 2.5 and 1.25 pc (this was the smallest grid size allowed during the calculations). The PPM is a third-order Godunov scheme, which does not rely on artificial viscosity to spread sharp variations of the flow variables. These discontinuities are represented by sharp profiles, free of spurious oscillations, of the flow variables and are treated as solutions to the Riemann problem. ## 5 Future Efforts and Final Remarks The major drawback of the simulations is the absence of magnetic fields that may constrain the upward motions of the disk gas. Understanding the disk-halo interaction requires simulations that couple together the major consituents of the ISM in the disk and halo, i.e, gas, magnetic fields, heating and cooling. The presence of magnetic fields may provide a confinement effect over the disk gas, providing the scale height of its distribution is much larger than that of the disk gas. For an average density distribution such as that observed in the simulations, the scale height is of the order of $`H_g170`$ pc, which is comparable to the scale height of the thin component of the Galactic magnetic field - $`H_B170`$ pc (Thompson & Nelson, 1980). Tomisaka (1998), using a magnetic field decreasing as the square root of the gas density, shows that no confinement occurs for the superbubbles in the disk. However, the presence of a thicker component of the magnetic field, with a scale height of 1.2 kpc (Han & Qiao, 1994), confines the expansion of superbubbles, located at $`\left|z\right|300`$ pc, to more than 20 Myr. Furthermore, as has been shown here, such study must be three-dimensional and independent of the grid size. That is, a 3D grid with a thickness of only a few hundred parsec generates an overpressured region in the disk, and thus the simulations evolve in a two-dimensional fashion rather than three-dimensional. The time delay between supernovae and superbubbles is so large that when the next supernova explodes, the previous one has already expanded to few hundred parsec in size, and therefore crossed the boundaries of the grid. The younger supernova has a large probability of occurring inside the old cavity and, as such, it expands in an evacuated medium with velocities larger than that of a supernova in phase II. Such simulations are difficult to carry out due to present-day computational facilities. However, efforts are under way, by the author, to provide a new model where magnetic fields come into play in a medium where the disk-halo structure is regulated by supernovae. Phenomena such as high and intermediate velocity clouds are regarded as products of the Galactic fountain and their dynamics, structure and evolution are described in paper II (Avillez, 1999b). In addition, some of the items described in this paper require full discussion which will be presented in future papers in the series. ## Acknowledgments This paper is dedicated to the memory of Prof. Franz D. Kahn. He will be sadly missed. I would like to thank the anonymous referee whose comments and suggestions led to the improvement of this manuscript.
warning/0001/hep-ph0001331.html
ar5iv
text
# References SEARCH FOR AMBIENT NEUTRALINO DARK MATTER AT ACCELERATOR Tai-Fu Feng<sup>1,2</sup>, Xue-Qian Li<sup>1,2</sup>, Wen-Gan Ma<sup>1,3</sup> and X. Zhang<sup>1,4</sup> 1 CCAST (World Laboratory) P.O. Box 8730, 100080, Beijing, China. 2 Department of Physics, Nankai University, Tianjin, 300070, China. 3 Department of Modern Physics, University of Science and Technology of China, Hefei, 230026, China 4 Institute of High Energy Physics, P.O. Box 918-4, 100039, Beijing, China. Abstract We investigate the possibility of using accelarator beam particles to collide with the ambient neutralino dark matter particles in cosmic rays as a way to search for the cold dark matter. We study in detail its inelastic and elastic scattering with the projectile particles at electron-positron colliders and discuss the possible experimental signals and the relevent background. There is strong evidence for the existence of a large amount of cold dark matter (CDM) in the universe. The leading candidate of the cold dark matter particles in the minimal supersymmetric standard model with R parity is neutralino . The possibility of sneutrino being a candidate of CDM is ruled out unless it as heavy as 17 TeV which is unfavorable in all resonable theories. In the traditional method of directly searching for such CDM SUSY particles, the velocity of the SUSY CDM particles is very small , their kinetic energy is not sufficient to cause any inelastic scattering, so that a clear observation would be difficult due to existence of a flood of background trajectories. In this paper we propose a new detection mechanism of neutralino cold dark matter by using the accelarator beam particles to collide with the ambient neutralino dark matter particles. Namely we use the projectile particles produced in accelerator ($`e^\pm ,p(\overline{p})`$ or even $`\gamma `$) to be incident on the dark matter particles. Because most of the cosmic ray particles interact with ordinary matter via electromagnetic processes, they cannot penetrate into the tunnel with high vacuum, but the dark matter particles only participate in ”weak” interaction and can come into the tunnel. Concretely, we let the beam (say, $`e^{}`$) go through a highly vacuumized tunnel while the colliding beam ($`e^+`$) is turned off. Once the projectile electrons bombard on neutralinos ($`\stackrel{~}{\chi }_1^0`$), because of the large available energy of $`e^{}`$, inelastic processes may occur as $$\stackrel{~}{\chi }_1^0+e^{}\stackrel{~}{\chi }_1^{}+\nu _e.$$ (1) The charged SUSY particles $`\stackrel{~}{\chi }_1^{}`$ would make trajectories in detector with magnetic field. Because they are much heavier than proton and pion etc. ordinary particles, they can be identified easily from background products. There are also elastic processes $$\stackrel{~}{\chi }_1^0+e^{}\stackrel{~}{\chi }_1^0+e^{},\nu _e+e^{},$$ (2) where the projectile electron declines from its beam direction. However, this process might be smeared with the background effect such as $$n+e^{}n+e^{},$$ where $`n`$ is the nucleon left in the tunnel, even though it is highly vacuumized. We will discuss the background problem again later in this letter. The density of cold dark matter in our ambient universe has been estimated . There have also been many theoretical works concerning the SUSY relic density after the Big Bang . Thus we can immediately evaluate the flux of the SUSY cold dark matter . The observed event number of the suggested reaction can be obtained as $$N=\rho _{DM}\rho _{beam}|\stackrel{}{v}_{rel}|\sigma Slt=\rho _{DM}\frac{1}{2}L\sigma Slt,$$ (3) where $`\rho _{DM},\rho _{beam}`$ are the densities of dark matter and beam particles respectively, $`|\stackrel{}{v}_{rel}|`$ is their relative velocity, $`\sigma `$ is the cross section of the reaction, $`S`$ is the cross section of the beam, $`l`$ is the length of the available detection region, $`t`$ is the time duration for measurement. Since the velocity of the coming SUSY particles is much lower than that of the beam particle, $`\rho _{beam}|\stackrel{}{v}_{rel}|L/2`$ where $`L`$ is the named luminosity. In 1972, a peculiar event of heavy cosmic ray particle was observed in the cloudy chamber of the Yunan Cosmic Ray Station (YCRS) . Recently, the event was re-analyzed and it is identified as that a heavy neutral particle $`C^0`$ came in and bombarded on a proton to produce a heavy charged particle $`C^+`$ as well as a proton and $`\pi ^{}`$. Their analysis confirmed that the mass of the heavy neutral cosmic ray particle $`C^0`$ is greater than 43 GeV and the mass difference $$\mathrm{\Delta }M=M_{C^+}M_{C^0}<0.270\mathrm{GeV}.$$ If taking this result seriously, one would be tempted to conclude that the coming neutral $`C^0`$ is a SUSY dark matter particle $`\stackrel{~}{\chi }_1^0`$ and the produced heavy charged particle is $`\stackrel{~}{\chi }_1^+`$ accordingly. In this case, the available energy of the Beijing Electron-Positron Collider (BEPC) is sufficient to cause an inelastic scattering where the projectile particle $`e^{}`$ of 2 GeV hits the coming $`\stackrel{~}{\chi }_1^0`$ to produce $`\stackrel{~}{\chi }_1^{}`$. If the mass difference of $`\stackrel{~}{\chi }^\pm `$ and $`\stackrel{~}{\chi }^0`$ is as large as a few tens GeV, the BEPC energy is not sufficient, and one needs to invoke machines with higher energy, such as, LEP or hadron colliders. In this paper we will take $`\mathrm{\Delta }M<1`$ GeV for a detail study, and then we will briefly give a general discussion on the case for larger $`\mathrm{\Delta }M`$. In calculation of the cross section $`\sigma `$ in eqs.(1), (2) and (3) without losing generality we assume that only one generation sfermion is light and one can neglect the sfermion mixing among different generations. The mass matrices of neutralinos and charginos can be found in Ref. For the mixing of the first generation slepton (right and left fields), we have $$M_{\stackrel{~}{e}}^2=\left(\begin{array}{cc}\frac{e^2(\upsilon _1^2\upsilon _2^2)(12c_w^2)}{8s_wc_w}+M_e^2+M_L^2& \frac{1}{\sqrt{2}}\left(\sqrt{2}M_e\mu +\upsilon _1h_{s_l}^1\right)\\ \frac{1}{\sqrt{2}}\left(\sqrt{2}M_e\mu +\upsilon _1h_{s_l}^1\right)& \frac{e^2(\upsilon _1^2\upsilon _2^2)}{4c_w^2}+M_e^2+M_R^2\end{array}\right).$$ (4) The mixing matrix $`Z_{\stackrel{~}{e}}`$ is defined as $$Z_{\stackrel{~}{e}}^{}M_{\stackrel{~}{e}}^2Z_{\stackrel{~}{e}}=diag(M_{\stackrel{~}{e}_1}^2,M_{\stackrel{~}{e}_2}^2).$$ (5) The mass of the electron-sneutrino is: $$M_{\stackrel{~}{\nu }_e}^2=M_L^2\frac{e^2(\upsilon _1^2\upsilon _2^2)}{8s_w^2c_w^2}.$$ (6) At tree level, the squared mass, $`M_{H_1^0}^2`$, of the light Higgs boson has an upper bound which is given by $`M_Z^2\mathrm{cos}^22\beta `$. This is already below the experimental lower bound of LEP2. However, radiative corrections can raise the upper bound on $`M_{H_1^0}^2`$ dramatically. The dominant contribution is $$M_{H_1^0}^2=\frac{3M_t^4}{2\pi ^2\upsilon ^2}\mathrm{ln}\frac{M_{\stackrel{~}{t}}^2}{M_t^2},$$ (7) where $`M_t`$ is the top quark mass and $`M_{\stackrel{~}{t}}`$ the top-squark mass. In our numerical calculation, we take the correction into account. When the kinematics is permissive, several inelastic reactions such as $`e^{}+\stackrel{~}{\chi }_1^0\stackrel{~}{\nu }_e+W^{}(H_1^{})`$, $`e^{}+\stackrel{~}{\chi }_1^0\stackrel{~}{e}_i^{}+Z^0(H^0,A^0)`$ $`(i=1`$, $`2)`$ etc. can occur. For the moment we consider only inelastic channels $`e^{}+\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^{}+\nu _e`$ and as well the elastic processes $`e^{}+\stackrel{~}{\chi }_1^0e^{}+\stackrel{~}{\chi }_1^0`$. For the channel $`e^{}+\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^{}+\nu _e`$, the amplitudes are given as $`_{\stackrel{~}{e}_i^{}}^s={\displaystyle \frac{i}{sM_{\stackrel{~}{e}_i^{}}^2}}{\displaystyle \underset{\sigma _1\sigma _2}{}}A_{\sigma _1}^{(1)}B_{\sigma _2}^{(1)}\overline{\upsilon }(p_2)\omega _{\sigma _1}u(p_1)\overline{u}(p_4)\omega _{\sigma _2}\upsilon (p_3),`$ $`_W^u={\displaystyle \frac{i}{uM_W^2}}{\displaystyle \frac{e^2}{2s_w^2}}{\displaystyle \underset{\sigma }{}}C_\sigma ^{(1)}\upsilon ^T(p_3)C^1\gamma _\mu \omega _{}u(p_1)\overline{u}(p_4)\gamma ^\mu \omega _\sigma C\overline{\upsilon }^T(p_2)`$ $`={\displaystyle \frac{i}{uM_W^2}}{\displaystyle \frac{e^2}{2s_w^2}}{\displaystyle \underset{\sigma }{}}C_\sigma ^{(1)}\overline{u}(p_3)\gamma _\mu \omega _{}u(p_1)\overline{u}(p_4)\gamma ^\mu \omega _\sigma u(p_2),`$ $`_{\stackrel{~}{\nu }_e}^t={\displaystyle \frac{i}{tM_{\stackrel{~}{\nu }_e}^2}}{\displaystyle \underset{\sigma _1\sigma _2}{}}D_{\sigma _1}^{(1)}E_{\sigma _2}^{(2)}\overline{\upsilon }(p_2)\omega _{\sigma _1}\upsilon (p_3)\overline{u}(p_4)\omega _{\sigma _2}u(p_1),`$ (8) where the couplings are $`A_{}^{(1)}={\displaystyle \frac{e}{\sqrt{2}s_wc_w}}Z_{\stackrel{~}{e}}^{1i}(Z_N^{11}s_w+Z_N^{21}c_w),`$ $`A_+^{(1)}={\displaystyle \frac{\sqrt{2}e}{c_w}}Z_{\stackrel{~}{e}}^{2i}Z_N^{11},`$ $`B_{}^{(1)}=0,`$ $`B_+^{(1)}={\displaystyle \frac{e}{s_w}}Z_{\stackrel{~}{e}}^{1i}Z_{}^{1j},`$ $`C_{}^{(1)}=Z_N^{21}Z_+^{1j}{\displaystyle \frac{Z_N^{41}Z_+^{2j}}{\sqrt{2}}},`$ $`C_+^{(1)}=Z_N^{21}Z_{}^{1j}+{\displaystyle \frac{Z_N^{3i}Z_{}^{2j}}{\sqrt{2}}},`$ $`D_{}^{(1)}={\displaystyle \frac{e}{\sqrt{2}s_wc_w}}(Z_N^{11}s_wZ_N^{21}c_w),`$ $`D_+^{(1)}=0,`$ $`E_{}^{(1)}={\displaystyle \frac{e}{s_w}}Z_+^{1j},`$ $`E_+^{(1)}=0.`$ (9) $`Z_N`$, $`Z_+`$, $`Z_{}`$ are defined in and $`Z_{\stackrel{~}{e}}`$ is given in (5). The amplitudes for $`e^{}+\stackrel{~}{\chi }_1^0e^{}+\stackrel{~}{\chi }_i^0`$ (when i=1, it is the elastic scattering case) are $`_{\stackrel{~}{e}_j^{}}^s={\displaystyle \frac{i}{sM_{\stackrel{~}{e}_j^{}}^2}}{\displaystyle \underset{\sigma _1\sigma _2}{}}A_{\sigma _1}^{(2)}B_{\sigma _2}^{(2)}\overline{\upsilon }(p_2)\omega _{\sigma _1}u(p_1)\overline{u}(p_3)\omega _{\sigma _2}\upsilon (p_4),`$ $`_{Z^0}^t={\displaystyle \frac{i}{tM_Z^2}}{\displaystyle \underset{\sigma _1\sigma _2}{}}C_{\sigma _1}^{(2)}D_{\sigma _2}^{(2)}\overline{u}(p_3)\gamma _\mu \omega _{\sigma _1}u(p_1)\overline{\upsilon }(p_2)\gamma ^\mu \omega _{\sigma _2}\upsilon (p_4),`$ $`_{\stackrel{~}{e}_j^{}}^u={\displaystyle \frac{i}{uM_{\stackrel{~}{e}_j^{}}^2}}{\displaystyle \underset{\sigma _1\sigma _2}{}}E_{\sigma _1}^{(2)}F_{\sigma _2}^{(2)}\overline{u}(p_3)\omega _{\sigma _1}u(p_2)\overline{u}(p_4)\omega _{\sigma _2}u(p_1).`$ (10) The couplings are written as $`A_{}^{(2)}={\displaystyle \frac{e}{\sqrt{2}s_wc_w}}Z_{\stackrel{~}{e}}^{1j}(Z_N^{11}s_w+Z_N^{21}c_w),`$ $`A_+^{(2)}={\displaystyle \frac{\sqrt{2}e}{c_w}}Z_{\stackrel{~}{e}}^{2j}Z_N^{11},`$ $`B_{}^{(2)}={\displaystyle \frac{\sqrt{2}e}{c_w}}Z_{\stackrel{~}{e}}^{2j}Z_N^{1i},`$ $`B_+^{(2)}={\displaystyle \frac{e}{\sqrt{2}s_wc_w}}Z_{\stackrel{~}{e}}^{1j}(Z_N^{1i}s_w+Z_N^{2i}c_w),`$ $`C_{}^{(2)}={\displaystyle \frac{e}{s_wc_w}}({\displaystyle \frac{1}{2}}s_w^2),`$ $`C_+^{(2)}={\displaystyle \frac{es_w}{c_w}},`$ $`D_{}^{(2)}={\displaystyle \frac{e}{2s_wc_w}}Z_N^{41}Z_N^{4i},`$ $`D_+^{(2)}={\displaystyle \frac{e}{2s_wc_w}}Z_N^{31}Z_N^{3i},`$ $`E_{}^{(2)}={\displaystyle \frac{\sqrt{2}e}{c_w}}Z_{\stackrel{~}{e}}^{2j}Z_N^{11},`$ $`E_+^{(2)}={\displaystyle \frac{e}{\sqrt{2}s_wc_w}}Z_{\stackrel{~}{e}}^{1j}(Z_N^{11}s_w+Z_N^{21}c_w),`$ $`F_{}^{(2)}={\displaystyle \frac{e}{\sqrt{2}s_wc_w}}Z_{\stackrel{~}{e}}^{1j}(Z_N^{1i}s_w+Z_N^{2i}c_w),`$ $`F_+^{(2)}={\displaystyle \frac{\sqrt{2}e}{c_w}}Z_{\stackrel{~}{e}}^{2j}Z_N^{1i}.`$ (11) With the amplitude, we can easily obtain the cross sections by integrating over the phase space of final states. It is noted that we carry out all the calculations in the laboratory frame because the velocity of the heavy dark matter particles is very small compared to that of the projectile beam particles. In our numerical calculations, we take $`\alpha =1/128.8,M_Z=91.12`$ GeV, $`M_W=80.22`$ GeV and first assume that $`M_{\stackrel{~}{\chi }^{}}M_{\stackrel{~}{\chi }_1^0}`$ is about 1 GeV. We will also present the results for larger mass difference later. For $`e^{}+\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^{}+\nu _e`$ we take the SUSY parameters as $`\mathrm{tan}\beta =5,M_{\stackrel{~}{\chi }_2^{}}=200`$ GeV. In the propagators we choose the typical values for the SUSY particle pole masses as $`M_{\stackrel{~}{e}_1}=110`$ GeV $`M_{\stackrel{~}{e}_2}=200`$ GeV, $`M_{\stackrel{~}{\nu }_e}=110`$ GeV. For the Higgs masses, we have $`M_{H_1^0}=110`$ GeV, $`M_{H_2^0}=300`$ GeV and $`M_{A^0}=110`$ GeV which are commonly adopted in literatures. In the computations, we consider three possible masses of $`\stackrel{~}{\chi }_1^0`$ as 40 GeV, 50 GeV, and 80 GeV. We find that the cross section of $`e^{}+\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^{}+\nu _e`$ is of order 100$`\mu `$b, which is the typical value for the weak interactions. We tabulate event numbers of inelastic process $`e^{}+\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^{}+\nu _e`$ for luminosities and energies of various $`e^+e^{}`$ colliders in Table 1. | | $`c\tau `$ factory | BEPC | VEPP-4M | CESR | KEKB | PEP-II | LEP | | --- | --- | --- | --- | --- | --- | --- | --- | | $`M_{\stackrel{~}{\chi }_1^0}=40`$ $`M_{\stackrel{~}{\chi }_1^{}}=41`$ | 800 | 20 | 250 | 392 | 375 | 850 | 4950 | | $`M_{\stackrel{~}{\chi }_1^0}=50`$ $`M_{\stackrel{~}{\chi }_1^{}}=51`$ | 1400 | 40 | 380 | 538 | 625 | 1410 | 1300 | | $`M_{\stackrel{~}{\chi }_1^0}=80`$ $`M_{\stackrel{~}{\chi }_1^{}}=81`$ | 16350 | 490 | 11830 | 18676 | 51875 | 32032 | 26850 | | $`M_{\stackrel{~}{\chi }_1^0}=40`$ $`M_{\stackrel{~}{\chi }_1^{}}=60`$ | $``$ | $``$ | $``$ | $``$ | $``$ | $``$ | 2106 | | $`M_{\stackrel{~}{\chi }_1^0}=40`$ $`M_{\stackrel{~}{\chi }_1^{}}=90`$ | $``$ | $``$ | $``$ | $``$ | $``$ | $``$ | 2711 | Table 1 The estimated event number for $`e^{}+\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^{}+\nu _e`$. Here the length of detection region $`l`$ is taken as 1 m and the masses are in GeV. In the third row, the number is unreasonably large and it is because of the pole effect of the propagator and related to the chosen parameters. The last two rows correspond to larger $`\mathrm{\Delta }M=20,\mathrm{\hspace{0.33em}50}`$ GeV, and only the LEP beam energy is capable of making such inelastic process. For the elastic processes $`e^{}+\stackrel{~}{\chi }_1^0\stackrel{~}{\chi }_1^0+e^{}`$, we evaluated the cross sections in terms of the parameters introduced above and obtain that the event number of such reactions at the BEPC luminosity and energy is still about 50 per year. Experimentally we expect to observe inelastic scattering between neutralino and the beam particles at $`e^+e^{}`$ machines. Because all SUSY particles are much heavier than ordinary SM particles, the trajectories of charged SUSY and SM particles can be easily distinguished in a strong magnetic field. At lower beam energies, such as BEPC, the kinetic energy of the produced chargino is relatively small, so that one can easily detect them. The third row of Table 1 show very large number of events, the reason is due to the pole of the propagator. But unless the matching happens to occur this way, it is not the case. So we do not take such large numbers seriously, of course, if the beam energy can be adjusted freely in a wide range, the threshold effects might manifest, but the possibility in practice is slim. Now let us turn to study the elastic processes where the background may contaminate the situation. The observation is based on measuring the electrons scattered from the SUSY dark matter particles in $`e^{}+\stackrel{~}{\chi }_1^0e^{}+\stackrel{~}{\chi }_1^0`$. The background is from the electrons scattered from nucleons of the remnant atmosphere in the vacuumized tunnel $`e^{}+ne^{}+n`$. At lower energies, the cross section of scattering can be easily computed and the amplitude is $$=\frac{G_F}{2\sqrt{2}}\overline{n}\gamma _\mu [(1+\frac{4}{3}\mathrm{sin}^2\theta _W)+\gamma _5]n\overline{e}[(1+4\mathrm{sin}^2\theta _W)+\gamma _5]e.$$ (12) Then we can obtain the cross section. At the same length, the background events are at least 1000 times larger than the expected events at $`1.031510^6pa`$. However, the situation can be remedied by careful analysis on kinematics. The maximal differential cross sections of either $`e^{}+\stackrel{~}{\chi }_1^0e^{}+\stackrel{~}{\chi }_1^0`$ or $`e^{}+ne^{}+n`$ occur near $`\theta =\pi /2`$. We have also the energy of the scattered electron $`E^{}`$ as $$E^{}=\frac{EM}{E(1\mathrm{cos}\theta )+M},$$ (13) where $`E`$ is the beam energy and in the expression the mass of electron is ignored, $`M`$ is the mass of either the SUSY particle or nucleon. Obviously, if $`ME`$ we have $`E^{}E`$, but as $`EM`$, at $`\theta =\pi /2`$, $`E^{}=ME/(M+E)`$. In the BEPC case, if the electron is scattered from a SUSY particle, $`E^{}2`$ GeV, but when it is from the background nucleon, $`E^{}0.6`$ GeV. Therefore, if we set a reasonable cut at, say, 1 GeV for the kinetic energy of the scattered electron, we can effectively eliminate the background (the case for $`\theta =0`$ vicinity should be excluded). However, if the beam energy is much greater than $`M`$, for example at LEP and future colliders, the energy difference of electron scattered from a heavy SUSY particle or a nucleon is negligible, so that one cannot distinguish the source of scattering and the expected events would be drowned in the ocean of background. Therefore, observation of SUSY dark matter via elastic scattering can only be feasible at low energy accelerator, and the energy of BEPC or charm-tau factory is ideal. Moreover, in that case, the mass difference of the neutral and charged SUSY particles has no effects at all. It is noted that the main background is due to scattering between the beam electron and atmospheric neutron, since only neutron is neutral in the remanat atmosphere. In fact, the collision between beam electron and other charged particles in the atmosphere such as proton and electron can be easily determined by existence of trajectories of the bounced charged particles and the scattered beam electron. Therefore our signal can be clearly singled out from the background. In summary, we have proposed a new method of detection of the cold dark matter and analyzed the possibility of observing inelastic scattering processes between beam particles and the ambient neutralino cold dark matter particles. We also investigate the situation for elastic scattering. Our results suggest that if the mass gap between the lightest SUSY particles $`\stackrel{~}{\chi }_1^0`$ and their corresponding charged SUSY particles $`\stackrel{~}{\chi }_1^{}`$ is less than 1 GeV, the BEPC energy is enough to cause inelastic processes where the products can be easily identified in present experimental facilities. For large mass gap, one needs to invoke larger machines such as LEP etc. Besides we also discuss the possibility of measuring elastic scattering and find that by carefully setting a kinematic cut for the scattered electron energy, it is possible to identify if they are scattered by the incident SUSY particles from large background at BEPC, but not at accelerators with higher energies. Our conclusion is that using beam particles to bombard on the coming SUSY particles may cause inelastic and elastic processes whose products are measurable and can be identified. We should point out that Errede and Luk in Ref. have outlined a scheme for searching for terrestrial dark matter at Tevatron. However they have not studied in detail the processes of the beam particles colliding with heavy cold dark matter particles, such as the neutralino. As pointed out in this paper, the heaviness of the dark matter particles help identify the signals from the background. Acknowledgment: This work is partly supported by the National Natural Science Foundation of China. We thank T. Han, S. Nussinov and R. Peccei for discussions. We thank R. Cousins for bringing Ref into our attention during the final stage of this work.
warning/0001/math-ph0001028.html
ar5iv
text
# A generalization of Jaynes’ principle: an information-theoretic interpretation of the minimum principles of quantum mechanics and gravitation ## 1 INTRODUCTION As Laplace famously observed, Newtonian mechanics is deterministic to an idealized intelligent being, but (as Laplace observed less-famously in the same passage) practical realities demand a probabilistic mechanics <sup>1</sup><sup>1</sup>1Laplace, P.-S., Laplace’s Théorie Analytique des Probabilités, 3rd ed., Paris: Courcier, 1820. Reprinted in volume 7 of his Oeuvres Complètes, Paris: Gauthier-Villars, 1886, see the Introduction, pp. vi-ix.. In the century after Laplace’s observation, two forms of probabilistic mechanics were discovered: statistical mechanics, and quantum mechanics. While the former maintained at a detailed level the determinism of Newtonian mechanics, acquiring its probabilistic nature only when describing macroscopic observations of large ensembles of particles, quantum mechanics introduced a probabilistic nature at the most fundamental level. In a sharpening of Laplace’s ”principle of insufficient reason”, Jaynes cast statistical mechanics into a novel form. Jaynes equated the entropy of statistical mechanics $$\underset{i}{}p_i\mathrm{log}(p_i)$$ (1) to the entropy of Shannon’s information theory, and equated the principle of maximum entropy with an information-theoretic inference law, Jaynes’ principle, which asserts that the maximum entropy probability distribution is that distribution which is the least-biased estimate for such a distribution <sup>2</sup><sup>2</sup>2Jaynes, E.T., Phys. Rev.,106,620 (1957). The goal here is to generalize Jaynes’ principle to the continuous probability distributions of quantum mechanics, and to demonstrate how the extremum principles of quantum mechanics, the Schrödinger equation and the Dirac equation, may be viewed as statements for formulating least-biased estimates of those continuous probability distributions. My thesis <sup>3</sup><sup>3</sup>3Van Drie, J.H., thesis,Calif. Inst. of Tech., 1978, unpublished explored the conceptual implications of this view. In this work, we focus on the surprising consequence of this perspective that the laws of gravitation as defined by General Relativity may also be viewed as a minimum-information principle, and show how this leads naturally to a set of field equations for non-Riemannian nonEuclidean geometries. ## 2 ENTROPY AS A MEASURE OF CORRELATION As pointed out in an earlier work <sup>4</sup><sup>4</sup>4Van Drie, J.H., http://xxx.lanl.gov/abs/math-ph/0001024, the Boltzmann / Shannon formula for the entropy of a discrete distribution may be viewed as a measure of the correlation between two distinct distributions. This was demonstrated, by showing that the entropy functional is non-decreasing as ”correlation-destroying transformations” are applied to the distribution. This perspective yields yet another view of Jaynes’ principle: the least-biased distribution is that which displays the least correlation between the variables consistent with the constraints. Also, this perspective leads to yet another view of the Second Law of thermodynamics: correlations tend to spontaneously decay, and are highly unlikely to arise spontaneously in an isolated system. It is useful to recall the fundamental importance that the concept of correlation played in Maxwell’s original derivation of the velocity distribution of the atoms of an ideal gas <sup>5</sup><sup>5</sup>5Maxwell,J.C., Phil. Soc., 1860. He placed two assumptions on the velocity distribution $`\mathrm{\Phi }(v_x,v_y,v_z)d^3v`$: (1) there should be no preferred orientation, $`\mathrm{\Phi }(v_x,v_y,v_z)d^3v=\mathrm{\Phi }(v)d^3v`$, where $`v^2=v_x^2+v_y^2+v_z^2`$, and (2) the velocities along one direction should not be correlated with those along another direction, $$\mathrm{\Phi }(v_x,v_y,v_z)dv_xdv_ydv_z=\varphi (v_x)dv_x\varphi (v_y)dv_y\varphi (v_z)dv_z$$ (2) These assumptions lead to Maxwell’s velocity distribution law, $$\mathrm{\Phi }(v_x,v_y,v_z)d^3v=exp(\alpha v^2)v^2dvd\mathrm{\Omega }$$ (3) where $`\alpha `$ is a positive constant (shown by Boltzmann to be 1/kT), and $`d\mathrm{\Omega }`$ is the element of surface integration in velocity space, $`\mathrm{sin}(v_\theta )dv_\theta dv_\varphi `$. The ability of a distribution over multiple variables to be expressed as the product of distributions over single variables is the hallmark of an uncorrelated distribution. ## 3 KINETIC ENERGY AS A CORRELATION MEASURE The term in the Hamiltonian associated with the Schrödinger equation, $`<\mathrm{\Psi }|\mathrm{\Delta }|\mathrm{\Psi }>`$ is commonly called the ”kinetic energy” (multiplied by a suitable units-dependent constant), by analogy to the corresponding term in the classical Hamiltonian, where $`\mathrm{\Delta }`$ is the Laplacian operator, $$\mathrm{\Delta }=\frac{^2}{x^2}\frac{^2}{y^2}\frac{^2}{z^2}$$ (4) It was originally suggested by this author in unpublished work <sup>6</sup><sup>6</sup>6Van Drie, J.H., Candidacy report, Calif. Inst. of Tech., 1975, and later by Sears, Dinur and Parr <sup>7</sup><sup>7</sup>7Sears, S., Parr, R., Dinur, U., Isr. J. Chem., 19, 165 (1980) that this expression represents an entropy expression. This assertion rested on intuitive arguments, leaving open the question ”what mathematical property is common to both the kinetic energy of quantum mechanics and the entropy of statistical mechanics?”. The answer which will be provided here is that both expressions are quantitative measures of correlation, or the lack thereof; the idea that the quantum mechanical kinetic energy in some instances measures correlation is an old concept from molecular quantum mechanics <sup>8</sup><sup>8</sup>8Löwdin,P.O., Adv. Chem. Phys., 2,207, (1959). Let us consider two spaces $`M_1`$ and $`M_2`$, of dimension $`n_1`$ and $`n_2`$ respectively, and the cross-product space $`M_1\times M_2`$ of dimension $`n_1+n_2`$. Furthermore, let us consider a representation of the group of rotations and translations on $`M_1,\gamma _1`$, and a representation on $`M_2,\gamma _2`$. As in quantum mechanics, let us assume that a metric exists which allows us to define a probability distribution over $`M_k`$ from $`\gamma _k`$, $`\rho _k=(\gamma _k,\gamma _k)`$ for each point in $`M_k`$. One can construct the product representation, $`\gamma _1\times \gamma _2`$, as a representation over $`M_1\times M_2`$. Such a representation on $`M_1\times M_2`$ is by definition uncorrelated relative to the variables $`M_1`$ vis-à-vis $`M_2`$, since it can be written as the product of a representation on $`M_1`$ and one on $`M_2`$. Consider an operator on representations of $`M_1,O_1`$, and an operator on representations of $`M_2`$, and the composition of these on $`M_1\times M_2`$, $`O_{1\times 2}`$. We assert that all such operators which obey the following relationship may be considered as measures of correlation between the variables of $`M_1`$ and those of $`M_2`$: $$O_{1\times 2}=O_1+O_2$$ (5) Denoting the expectation value of an operator $`O_k`$ over the space $`M_k`$ against the representation $`\gamma _k`$ by $`<\gamma _k,O_k\gamma _k>`$, which equals $$<\gamma _k,O_k\gamma _k>=_{M_k}(\gamma _k,O_k\gamma _k)$$ (6) we see that for operators which are considered measures of correlation against $`M_1`$ and $`M_2`$, i.e. those operators obeying the relation ( 5), $$<\gamma _1\times \gamma _2,O_{1\times 2}\gamma _1\times \gamma _2>=<\gamma _1O_1\gamma _1>+<\gamma _2O_2\gamma _2>$$ (7) Considering these expectation values as measures of correlation of their corresponding representations, denoted $`I[\gamma ]`$, this allows us to interpret the above equation to say that, for uncorrelated representations, the measure of correlation is additive: $$I[\gamma _1\times \gamma _2]=I[\gamma _1]+I[\gamma _2]$$ (8) This same relationship holds for the Boltzmann / Shannon entropy of two discrete distributions, $`P=\{p_i\}_{i=1}^n`$ and $`Q=\{q_j\}_{j=1}^m`$, where the product distribution $`PQ=\{p_iq_j\}_{i,j=1}^{n,m}`$, $$S(PQ)=S(P)+S(Q)$$ (9) Note that the Laplacian operator, the kinetic energy term of the minimum principle for the Schrödinger equation, explicitly obeys the property ( 5) for the n translational variables $`\{x_k\}_{k=1}^n`$ of n-dimensonal space M, since for any m-dimensional subspace $`\{x_k\}_{k=1}^mM_1`$, and its complement $`\{x_k\}_{k=m+1}^nM_2`$, $`\mathrm{\Delta }_M=\mathrm{\Delta }_{M_1}+\mathrm{\Delta }_{M_2}`$. Intriguingly, R. A. Fisher applied the term information to the expectation value of the Laplacian, in an obscure and unelaborated reference <sup>9</sup><sup>9</sup>9Fisher, R.A., Stat. Methods and Sci. Inference, NY: Hafner, 1956, eqn. 155, but in the Introduction, he maintains this has no relation to the concepts in the Mathematical Theory of Communication, i.e. information theory. Even for 4-component spinor representations in relativistic spacetime, where the relevant quantum mechanical equation is the Dirac equation and the corresponding term in the minimum principle is the expectation value of the operator $`i/`$ <sup>10</sup><sup>10</sup>10Feynman R.P., Quantum Electrodynamics, London: Benjamin/Cummings, 1962, the property ( 5) holds. For the flat spacetime metric $`g_{\mu \nu }`$, a set of 4x4 matrices defined over the components of the spinors exist which obey the property $$\gamma _{(\mu }\gamma _{\nu )}=\frac{1}{2}(\gamma _\mu \gamma _\nu +\gamma _\nu \gamma _\mu )=g_{\mu \nu }$$ (10) and the operator $`i/`$ may be written $$i/=i\underset{\mu }{}\gamma _\mu \frac{}{x^\mu }$$ (11) As with the Laplacian, the property ( 5) is evident from the definition of this operator; hence even in relativistic spacetime, the minimum principle contains a term which we may call a measure of correlation. ## 4 THE GENERALIZED JAYNES’ PRINCIPLE Jaynes asserted that maximizing the entropy $`_kp_k\mathrm{log}(p_k)`$ over all distributions $`p_k`$ subject to the constraints of a given energy $`E=E_kp_k`$ and normalization $`_kp_k=1`$ may be viewed as a statement that the distribution $`p_k`$ is the least-biased distribution for $`p_k`$ subject to these constraints. $`\delta _{p_k}\{{\displaystyle \underset{k}{}}p_klog(p_k)`$ $``$ $`\lambda _1({\displaystyle \underset{k}{}}E_kp_k)`$ (12) $``$ $`\lambda _2({\displaystyle \underset{k}{}}p_k)\}=0`$ (13) $`p_k`$ $`=`$ $`\alpha exp(\beta E_k)`$ (14) where $`\lambda _1`$ and $`\lambda _2`$ are Lagrangian multipliers, and $`\delta _{p_k}`$ denotes varying over all possible $`\{p_k\}`$. Viewing the Laplacian operator as a measure of correlation in a representation $`\gamma `$, we assert that the minimum principle for the Schrödinger equation may be viewed as the statement that $`\gamma `$ represents the least-biased representation subject to the constraints of normalization and in the presence of a potential V(x): $$\delta _\gamma \{<\gamma ,\mathrm{\Delta }\gamma >\lambda _1V(x)(\gamma ,\gamma )\lambda _2(\gamma ,\gamma )\}=0$$ (15) One of the many conceptual implications of this view is that it allows us to understand the physical basis for why the electron does not collapse onto the nucleus of an atom: the tendency to minimize its potential energy by withdrawing into the nucleus is counterbalanced by the tendency of its distribution to resist achieving such a highly-correlated state. An amusing application of this principle is the case where $`\gamma `$ is a vector representation, the normal to the surface of a soap-film. Minimizing $`<\gamma ,\mathrm{\Delta }\gamma >`$ subject to the constraint that the film adhere to a specified 1-dimensional wire frame gives the equation for the equilibrium configuration of such surfaces. This tendency of minimizing $`<\gamma ,\mathrm{\Delta }\gamma >`$ to function like a surface tension can also be understood by recalling Maxwell’s observation about the Laplacian: the Laplacian of a function is proportional to the difference between that function and that function’s average value over a ball of radius $`ϵ`$ <sup>11</sup><sup>11</sup>11Maxwell, J.C., Treatise on Electricity and Magnetism, vol. 1, New York: Dover, 1991, p. 31, a property well-known in the numerical analysis of Laplace’s equation. For nonEuclidean geometries, a more refined definition of the Laplacian must be used to ensure that the property ( 5) is satisfied. The generalized Laplacian of deRham, <sup>12</sup><sup>12</sup>12deRham, G., Variétés Différentiables, Paris: Hermann, 1960, p. 125, $`\mathrm{\Delta }=d\delta +\delta d`$, where d is the exterior derivative and $`\delta =d`$, $``$ the Hodge dual operator, may be tediously shown to possess the property ( 5), through the use of deRham’s forms double. This allows us to write the most general form of this generalized Jaynes’ principle, namely $$\delta _\gamma \{<\gamma ,\mathrm{\Delta }\gamma >\}=0$$ (16) subject to constraints, among them $`<\gamma ,\gamma >=1`$, where $`\gamma `$ is understood to be any representation over a nonEuclidean space, and $`\mathrm{\Delta }`$ is understood to be deRham’s Laplacian. ## 5 APPLICATION TO NONEUCLIDEAN GEOMETRIES For a Riemannian geometry, deRham gave the explicit formula for his Laplacian applied to a tensor of arbitrary rank p<sup>13</sup><sup>13</sup>13deRham,ibid., p. 131: $`\mathrm{\Delta }\alpha _{k_1k_2\mathrm{}k_p}`$ $`=`$ $`\alpha _{k_1k_2\mathrm{}k_p;i}^{;i}+{\displaystyle \underset{\nu =1}{\overset{p}{}}}(1)^\nu R_{.i.k_\nu }^{h.i.}\alpha _{hk_1\mathrm{}\widehat{k_\nu }\mathrm{}k_p}`$ (18) $`+2{\displaystyle \underset{\mu <\nu }{\overset{1\mathrm{}p}{}}}(1)^{\mu +\nu }R_{.k_\nu .k_\mu }^{h.i.}\alpha _{ihk_1\mathrm{}\widehat{k_\mu }\mathrm{}\widehat{k_\nu }\mathrm{}k_p}`$ where $`\widehat{k_\mu }`$ denotes that subscript is dropped from the enumerated list of indices, and where deRham’s notation of the covariant derivative $`_i`$ is replaced by the notation of Misner et al.<sup>14</sup><sup>14</sup>14Misner, C.W., Thorne, K.S., and Wheeler, J.A., Gravitation, San Francisco: W.H.Freeman, 1973where the covariant derivative is denoted by $`;\alpha `$, and where $`R_{\beta \mu \nu }^\alpha `$ is the Riemann curvature tensor. In spacetime, applying deRham’s Laplacian to the metric, $`g_{\mu \nu }`$, we see that the above formula reduces to $$\mathrm{\Delta }g_{\mu \nu }=g_{\mu \nu ;\alpha }^{;\alpha }+R_{\mu \nu }$$ (19) where $`R_{mu\nu }`$ is the Ricci tensor, the contraction of the Riemann tensor. The term $`g_{\mu \nu ;\alpha }^{;\alpha }`$ is zero, by the covariant constancy of the metric, and hence the measure of correlation, the expectation value of the Laplacian, is $$<g^{\mu \nu }\mathrm{\Delta }g_{\mu \nu }>=g^{\mu \nu }R_{\mu \nu }𝑑\tau =R𝑑\tau $$ (20) where $`d\tau `$ is the volume element of integration over spacetime, and R is the scalar curvature, the contraction of the Ricci tensor. Minimizing $`R𝑑\tau `$ over all metrics is the Hilbert variational principle, and leads to Einstein’s equations of General Relativity in free space<sup>15</sup><sup>15</sup>15Misner et al., ibid., Ch. 21. Hence, the generalized Jaynes principle states that Einstein’s equations of General Relativity in free space may be interpreted as asserting that the metric is the least-biased metric, or the minimally-correlated metric. In the nonEuclidean spaces of Cartan, the fundamental quantities are not the metric, but rather the 1-forms of the repere mobile $`\omega _\mu `$ and the connection 1-forms $`\omega _\nu ^\mu `$ <sup>16</sup><sup>16</sup>16Slebodzinski, W., Exterior Forms and their Applications, Warsaw: Polish Scientific Publishers, 1970, section 124. This originated from Cartan, E., Ann. Ec. Norm.,40, 325 (1923), reprinted in his Oeuvres Complètes, partie III, vol. 1, Paris: Gauthiers-Villars, p. 659. He showed that, for these more general nonEuclidean spaces, an additional invariant arises, the torsion; two such spaces are equivalent if both the torsion and curvature are equal. Cartan’s structure equations define the torsion $`\mathrm{\Omega }^\mu `$ and curvature $`\mathrm{\Omega }_\nu ^\mu `$: $`d\omega ^\mu +\omega _\nu ^\mu \omega ^\nu `$ $`=`$ $`\mathrm{\Omega }^\mu `$ (21) $`d\omega _\nu ^\mu +\omega _\alpha ^\mu \omega _\nu ^\alpha `$ $`=`$ $`\mathrm{\Omega }_\nu ^\mu `$ (22) For Riemannian spaces, the torsion is zero, and the problem of equivalence reduces to the study of the curvature form. While Einstein’s equations of General Relativity allows one to write field equations for the curvature, the issue of field equations for torsion and curvature has received less attention. The natural extension of the above ideas to such spaces with torsion is to consider the following equations: $$\delta _{\omega _\mu ,\omega _\nu ^\mu }\{<\omega ^\mu \mathrm{\Delta }\omega _\mu >\}=0$$ (23) where the minimization is taken over all orthonormal bases $`\omega _\mu `$ and connection forms $`\omega _\nu ^\mu `$. Of the numerous possible forms to choose from, the basis 1-forms seems the most natural, in that, like the metric, they represent a generalized potential, derivatives of which lead to a generalized force; derivatives of generalized forces can then be linked to source terms, like mass/energy density. The details of the above equation will be explored in a future work. ## 6 IMPLICATIONS OF THIS EQUATION An interesting question is “Are the principles of structure independent of scale?”. One of the reasons the study of fractals in biological settings has generated such enthusiasm is that it implicitly answers that question Yes over the scales ranging from the size of macromolecules to the size of plants. This equation appears to suggest that the principles of structure may be independent of scale over an even wider range, from the atomic scale to the astrophysical scale. Of course, at each scale the forces that are relevant are different, and hence the resulting structures are different (the constraints that must be imposed in the generalized Jaynes principle).
warning/0001/cond-mat0001027.html
ar5iv
text
# The Hypergraphite: A possible extension of graphitic network ## I Introduction Graphite has many intriguing aspects from scientific and industrial view points. It has high electronic conductivity and large anisotropic diamagnetic susceptibility. The electronic structure is semi-metallic. An important point is that these characteristic features of an electronic structure originate from a particular lattice structure of graphite. The lattice structure is a stack of honeycomb networks of carbon atoms. A single layer of graphite is called graphene. Our interest in graphite has deepened much by findings of rich physics not only in pure graphite itself but also in various graphitic materials. For example, fullerenes and carbon nanotubes are regarded as a deformed graphitic structure. The discoveries have revealed that the topology of $`sp^2`$ carbon networks crucially influences their $`\pi `$-electronic structures. An example showing effects of topology is that electronic structures of single wall carbon nanotubes depend on the chiral vectors. The $`\pi `$-electronic structures show metallic or semi-conducting behavior. Interestingly, recent experiments confirmed this interplay between the topology of $`sp^2`$ carbons and the $`\pi `$-electronic states. Let us start from the graphene to understand the electronic structures of graphite and related materials. The electronic structure of the graphene around the Fermi energy is composed of a $`\pi `$ (or $`\pi ^{}`$)-band. The $`\pi `$-band and the $`\pi ^{}`$-band are the highest valence band and the lowest conduction band, respectively. The two bands degenerate at the edge of the first Brillouin zone (1st BZ), called K (or K’)-point, and show $`k`$-linear dispersion at these points. Thus, the graphene is a zero-gap semiconductor. Namely, the density of states (DOS) becomes zero at the Fermi energy and linearly increases with leaving from the Fermi energy. Another interesting $`\pi `$-electron system is nanometer-sized graphite called “nanographite”. Nanographites are a class of mesoscopic systems which are situated between aromatic molecules and graphite. In nanographite, the presence of edges, their shapes and size crucially affect their $`\pi `$-electronic structures. We introduced ribbon models of a graphene to study effects of the edge on their $`\pi `$-electronic structures theoretically. The typical edge shapes are a zigzag edge or an armchair edge. Hence, we investigated zigzag ribbons, which are graphene ribbons with only zigzag edge, and armchair ribbons, which are graphene ribbons with only armchair edge. There is a clear difference between $`\pi `$-electronic structures of these two ribbon models. In case of zigzag ribbons, a localized states appear around the edge at the Fermi energy. While such localized states do not exist on the edge of armchair ribbons. Thus appearance of the localized states is a topological effect of zigzag ribbons. The localized states on zigzag ribbons were named “edges states”, because they exist around edges of zigzag ribbons. A solution of the edge states was constructed on a semi-infinite graphene with a zigzag edge. This solution shows that the edge states is a non-bonding orbital (NBO). Band structures of zigzag ribbons possess a pair of almost flat bands near the Fermi energy. These flat bands, i.e. edge states, induce a sharp peak in DOS at the Fermi energy. Therefore, zigzag ribbons are expected to show unusual properties, e.g. spin polarization, lattice deformation, magnetic field effects, transport properties, etc. The appearance of edge states on zigzag ribbons is also recognized in terms of the first-principles calculations within the framework of the local density approximation. Now it is natural to ask a question whether other structures which have edge states due to a topological reason exist or not. This is because interesting phenomena have been found only for zigzag ribbons of a graphene. In this paper, we report on a method to construct an extension of the graphitic network. The extended networks possess the same characteristics that the graphene shows both in the topological network and in the electronic structures. The networks are an AB bipartite network. When a proper boundary condition is subjected to them, edge states appear in these networks. Besides, the bulk states of these networks become a zero-gap semiconductor. And the highest valence band and the lowest conduction band degenerate at the Fermi energy in 1st BZ and the two bands show $`k`$-linear dispersion at the Fermi energy. We name the networks “hypergraphite”, because they are regarded as an extension of the graphitic network. The organization of this paper is the following. In section II, we argue about edge states. In particular, we introduce networks in which edge states appear under a proper boundary condition. In section III, we argue about bulk states of these networks by using a numerical calculation and an effective mass approximation. In section IV, we give a definition of the hypergraphite. ## II EDGE STATES The purpose of this section is to show several networks in which edge states appear. First, we recall a solution of edge states on a semi-infinite graphene with a zigzag edge. Then we consider the relationship between the topology of zigzag ribbons and the solution of edge states. This consideration leads us to find a method to construct these networks. Next, we give solutions of edge states which appear in these networks. We show band structures of slab model of them. In this section and the next section, we use a single-band tight-binding model (s-TBM). This is because we study the relation between topology of networks and electronic structures. The Hamiltonian is written as $`H_{TBM}=t{\displaystyle \underset{i,j,\sigma }{}}(c_{i,\sigma }^{}c_{j,\sigma }+\mathrm{h}.\mathrm{c}.).`$ where $`c_{i,\sigma }^{}(c_{i,\sigma })`$ creates (annihilates) an electron with spin, $`\sigma `$, on the $`i`$-th site (atom). We let the on-site energy zero, which is always possible for s-TBM. A bond connection is represented by $`i,j`$ and taken only between connected sites with hopping integral, $`t`$. Hereafter we assume that the hopping integral, $`t`$, is unity for simplicity. And we consider only a paramagnetic state. In addition, each site is occupied by one electron on the average. ### A EDGE STATES ON A ZIGZAG RIBBON We recall how to construct the solution of the edge states on a semi-infinite graphene with a zigzag edge. The solution, $`\varphi _{k,D}(x,z)`$, is for the $`\pi `$-electron systems (Fig.1(a)). Here, $`k`$ and $`D`$ denote a wave vector along the edge ($`x`$-direction) and a dumping factor inward the ribbon ($`z`$-direction), respectively. Each atom at the zigzag edge is assumed to be terminated by a hydrogen atom. The honeycomb network is AB bipartite and each atomic site can be classified into A- or B-site. In terms of this classification, all of the zigzag-edge sites of a semi-infinite graphene is in one of the two sublattices. In this paper, we define sites in the sublattice containing the edge sites as A-sites and sites in the other sublattice as B-sites. We list some characteristics of $`\varphi _{k,D}(x,z)`$. 1) The amplitude is non-zero only on A-sites. This means that edge states are NBOs. 2) Edge states are a dumping wave with a dumping factor, $`D=2\mathrm{cos}\frac{k}{2}`$, when $`\frac{2\pi }{3}<k<\pi `$. 3) The edge state completely localizes at the edge, ($`D=0`$), when $`k=\pi `$. The amplitude are +1 and -1 alternately along the edge except for normalization. 4) The edge state approaches a propagating wave ($`D1`$), when $`k\frac{2\pi }{3}`$. This limit, $`\varphi _{k2\pi /3,D1}`$, coincides with bulk states at the K point in BZ of the graphene. A construction method of the solution of the edge states in a semi-infinite graphene with a zigzag edge is briefly summarized as follows. We use a notation ($`x,nj`$), ($`j=A`$ or $`B`$), to specify a site. Here, $`x`$ denotes a position of a cell represented by dashed lines in $`x`$-direction (See Fig. 1(b)). The cell contains two types of sites of both A- and B-sites. $`nA`$ (or $`nB`$) represents A sites (or B sites) in the $`n`$-th zigzag-chain in the cell, respectively. We start from setting a trial function, $`\psi _k(x,1A)=exp(ikx)`$, to construct the solution of edge states. Next, we determine amplitudes of $`\psi _k(x,iA),(i=2,\mathrm{},\mathrm{})`$, so that $`\psi _k(x,jB),(j=1,\mathrm{},\mathrm{})`$, become zero. The final expression becomes an eigen function with zero amplitude on B-sites, i.e. edge states. ### B NETWORKS WITH EDGE STATES We consider the relationship between topology of zigzag ribbons and the above method to construct the wave functions. Usually, a network of zigzag ribbons is regarded as assembly of hexagons. However, we give another topological view point to the network of zigzag ribbons. It is that the network of zigzag ribbons are regarded as a network made by linking A-site of the $`i`$-th AB bipartite zigzag-chain and B-site of the ($`i+1`$)-st one by extra bond ($`i=1,\mathrm{},\mathrm{}`$) (Fig. 1(c)). A characteristic of the view point is to connect zigzag-chains which are an AB bipartite lattice and have a NBO as an eigen-function. We conjecture conditions for networks having edge states from the topological view point as follows. 1) A network is regarded as one made by successively connecting units. The unit is an AB bipartite lattice which has a NBO as an eigen-function. 2) The connection is made by linking A-sites of the $`i`$-th bipartite lattice with B-sites of the ($`i+1`$)-st bipartite lattice by extra bonds ($`i=1,2,\mathrm{},\mathrm{}`$). Fig.1. (a) A solution of an edge state for a semi-infinite graphene with a zigzag edge. Closed circles and opened circles represent A-sites and B-sites, respectively. The wave function has finite amplitude at A-sites. $`D`$ is a dumping factor, $`D=2\mathrm{cos}\frac{k_x}{2}`$. (b) This figure shows the notation which we use in text. The arrow represents a unit vector. Dash line shows a unit cell of a semi-infinite graphene with a zigzag edge. $`N`$ represents the numbers of zigzag chains. (c) A way to construct a zigzag ribbon from zigzag chains. Black bonds are extra bonds to connect zigzag chains Fig.2. (a) The three-dimensional three-fold coordinated network (the 3-3 network). $`N`$ represents the number of blocks shown in Fig. 2(c). Closed circles and opened circles represent A-sites and B-sites, respectively. (b) 1D zigzag chains aligned in the x-direction. (c) A structure made by connecting zigzag chains aligned in x-direction and y-direction. Here, we introduce three examples which satisfy the above conditions. They are the three-dimensional three-fold coordinated network (the 3-3 network), the diamond structure (DS) and the three-dimensional five-fold coordinated network (the 3-5 network). We show the 3-3 network in Fig. 2(a), the DS in Fig. 3(a), and the 3-5 network in Fig. 4. Of course, edge states appear in all of them under a proper boundary condition as shown in the next subsection. We show that they satisfy the above conditions. The 3-3 network is regarded as a network connecting AB bipartite lattices as shown in Fig. 2(b). The AB bipartite lattice is made by combining a set of zigzag chains aligned in $`x`$-direction (Fig. 2(c)) and in $`y`$-direction, alternately. A NBO exists on the AB bipartite lattice as an eigen function. The DS and the 3-5 network are regarded as networks of connected honeycomb lattices and of connected square lattices, respectively. Naturally, a honeycomb lattice and a square lattice are an AB bipartite lattice and have a NBO as an eigen-function. Fig.2. (d) A solution of the edge state for the semi-infinite 3-3 network with zigzag surface. The wave function of the edge states has finite amplitude at the sites indicated by closed circles. Dashed lines represent the next layer to the layer shown by solid lines. And two arrows represent two-dimensional unit vectors. The wave function is given by $`\mathrm{\Psi }_{n,m}^{lx}=D^{l1}\mathrm{exp}(i(k_xn+k_ym))`$, $`\mathrm{\Psi }_{n,m}^{ly}=D^{l1}(2\mathrm{cos}k_x)\mathrm{exp}(i(k_xn+k_ym))`$ and $`D=4\mathrm{cos}k_x\mathrm{cos}k_y`$. $`D`$ is a dumping factor. (e) The 1st BZ of a slab model of the 3-3 network. Edge states emerge in the shadowed region. (f) A band structure and DOS of the 3-3 network. Here, we adopt a slab model with $`N=10`$. ### C SOLUTIONS OF EDGE STATES We show solutions of edge states which appear in the 3-3 network and the DS under a boundary condition. The boundary condition is an open boundary condition with a surface. The surface is composed of A-sites being only in a connected AB bipartite lattice used as a unit to make the full network. We call the surface ”zigzag surface”. A dumping factor, $`D`$, and an area in which edge states appear in 1st BZ are determined from the solution. We show solutions of edge states in Fig. 2(d) for the 3-3 network and in Fig. 3(b) for the DS. In this paper, we do not show a solution of the 3-5 network, but a similar solution is constructed for it. Here, we note that s-TBM is used when solutions of edge states are constructed. Namely, we assume that a hopping integral, $`t`$, is the same and unity between any connected sites in any direction. Fig.3. (a) A diamond structure. $`N`$ represents the number of honeycomb lattices. (b) A solution of the edge states for semi-infinite DS i,e, with (111)-surface. In each panel, dashed lines represent a next layer to a layer represented by solid lines. The wave function of the edge states has the amplitude at the sites indicated by closed circles. And two arrows represent two-dimensional unit vectors. The wave function is given by $`\mathrm{\Psi }_{n,m}^l=(D)^{l1}\mathrm{exp}(i(k_xn+k_ym))`$ and $`D=e^{i(2k_x+k_y)}+e^{i(k_x2k_y)}+e^{i(k_x+k_y)}`$. $`D`$ is a dumping factor. (c) The 1st BZ of a slab model of DS. Edge states emerge in the shadowed region. (d) A band structure and DOS of a slab model of DS. Here, we adopt a slab model with $`N=20`$. ### D BAND STRUCTURES OF SLAB MODELS The electronic band structures of a slab model are shown in Fig. 2(f) for the 3-3 network and in Fig. 3(d) for the DS. They are obtained by solving the eigenvalue equations of s-TBM for each structure numerically. Since, we consider the case that the electron number is the same as the number of atomic sites, the Fermi energy becomes just zero for these models. A pair of almost flat bands exists at the Fermi energy in both band structures. There is a sharp peak at the Fermi energy in DOS for the 3-3 network and the DS. We recognize appearance of edge states from these results. Edge states appear in the 3-3 network and the DS which satisfy the conditions described in subsection B. In this paper, we show only two examples, but we have found many networks which satisfy the conditions and confirmed that edge states appear in these networks. In addition, the dimension of the structure does not have to be three or two. It is possible to design a higher dimensional structure with edge states. A dumping factor for these networks is obtained by using a transfer matrix method (See Appendix. A). Fig.4. A three-dimensional five-fold coordinated network. $`N`$ represents the number of square lattices. ## III BULK STATES In this section, we argue about bulk states of the networks which satisfy the conditions described in the last section. Electronic structures of the 3-3 network and the DS with a zigzag surface are similar to that of the graphene. Here, we have a question of how the electronic structure of bulk states is. Bulk states of the 3-3 network and the DS are made use of in this section as examples. ### A BAND STRUCTURES We show electronic band structures and DOS in Fig. 5(a) for the 3-3 network and in Fig. 6(a) for the DS, respectively. The Fermi energy is just zero in the model. At first, we explain the band structure of the 3-3 network. As shown in Fig. 5(a), the lowest conduction band and the highest valence band degenerate at a point on the $`\mathrm{\Gamma }X`$ line. At this point, two bands have $`k`$-linear dispersion. This is the same characteristic that the $`\pi `$-states of the graphene have. But, in the case of the 3-3 network, the lowest conduction band and the highest valence band degenerate on a closed line in 1st BZ. This line is shown in Fig 5(b). Recall that the lowest conduction band and the highest valence band degenerate at only $`K`$ (or $`K^{}`$) point in the case of the graphene. This line for the 3-3 network is understood to be an extension of the $`K`$ point of the graphene. Hence, we call this line “$`k`$-line”. This point is argued by using the $`kp`$ approximation in next subsection again. In addition, the system is a zero-gap semiconductor as seen in DOS. Next, we explain the band structure of the DS. There is a $`k`$-line, where the lowest conduction band and the highest valence band degenerate and two bands possess $`k`$-linear dispersion at W point. These $`k`$-lines are shown in Fig. 6(b). The system is also a zero-gap semiconductor as seen in DOS. Fig.5. (a) A band structure and DOS of a bulk of the 3-3 network. (b) The 1st BZ of the 3-3 network. In this figure, a region which contain $`k_x0`$ and $`k_y0`$ is showed. The $`k`$-line is represented by the thick line. Fig.6. (a) A band structure and DOS of a bulk of the DS. (b) The 1st BZ of the DS. In this figure, a region which contain $`k_x0`$ and $`k_y0`$ is showed. The $`k`$-lines are represented by the thick lines. The electronic structures of both the 3-3 network and the DS show the same characteristic. However, there is a difference in the shape of $`k`$-line of these two structures. The $`k`$-line is closed in 1st BZ in a case of the 3-3 network, while the $`k`$-line is open in 1st BZ in a case of the DS. We have studied many networks which satisfy the conditions described in the last section except the 3-3 network and the DS. The most of the networks show the same characteristics that the 3-3 network or the DS shows. One point which we have to remark is that in some structures made by our construction method, the bulk band structure is not a zero-gap semiconductor but a metal. In these exceptional cases, edge states appear in these networks whose bulk states become metallic. ### B K$``$P APPROXIMATION In this subsection, we study bulk states of the 3-3 network and the DS using the $`kp`$ approximation . The results show again that there are bands with $`k`$-linear dispersion in these networks. First, we study the 3-3 network by the $`kp`$ approximation. We start from obtaining wave functions on a $`k`$-line by diagonalizing the single-band tight-binding Hamiltonian. Wave functions are represented in a Bloch form as $`\psi _𝐤(𝐫)={\displaystyle \frac{1}{\sqrt{N}}}{\displaystyle \underset{j=1}{\overset{4}{}}}{\displaystyle \underset{𝐑_i}{}}\mathrm{exp}(i𝐤𝐑_i)b_j(𝐤)\varphi _j(𝐫𝐑_i),`$ (1) where $`\varphi _j(𝐫𝐑_i)`$ is a wave function of the $`j`$-th site in a unit cell. The vector, $`𝐑_i`$, represents the position of the $`i`$-th cell. $`b_j(𝐤)`$ is an amplitude which should be determined. $`N`$ is a number of unit cells. Here, we assume that $`\varphi _j(𝐫𝐑_i)`$ are orthonormal: $`(\varphi _j(𝐫𝐑_i),\varphi _j^{}(𝐫𝐑_i^{}))=\delta _{𝐑_i,𝐑_i^{}}\delta _{j,j^{}}.`$ The eigenvalue equation of the 3-3 network becomes following, $`H(𝐤)𝐛(𝐤)=\epsilon 𝐛(𝐤).`$ (2) The matrix, $`H(𝐤)`$, is defined by, $`H(𝐤)=\left(\begin{array}{cc}\mathrm{𝟎}& 𝐀\\ 𝐀^{}& \mathrm{𝟎}\end{array}\right),`$ (5) where $`𝐀=\left(\begin{array}{cc}(1+\mathrm{exp}(i\sqrt{3}k_xa)& \mathrm{exp}(i(\frac{\sqrt{3}}{2}k_xa+\frac{\sqrt{3}}{2}k_ya+3k_za)\\ 1& (1+\mathrm{exp}(i\sqrt{3}k_ya))\end{array}\right).`$ (8) a is the bond length. A vector $`𝐛(𝐤)`$ is given by $`𝐛(𝐤)=\left(\begin{array}{c}b_1(𝐤)\\ b_2(𝐤)\\ b_3(𝐤)\\ b_4(𝐤)\end{array}\right).`$ We can deduce equations determining a $`k`$-line on which energy is equal to zero from the above eigenvalue equation (Eq. (2)). The condition, $`\mathrm{det}𝐀=0`$, gives $`k`$-line equations, $`\{\begin{array}{c}\mathrm{cos}\frac{\sqrt{3}}{2}K_xa\mathrm{cos}\frac{\sqrt{3}}{2}K_ya=\frac{1}{4}\hfill \\ k_z=0\hfill \end{array},`$ (11) where $`𝐊=(K_x,K_y,0)`$ is a point on $`k`$-line. We need wave functions on $`k`$-line to apply the $`kp`$ approximation. These are given in a Bloch form as, $`\psi _A(𝐫)`$ $`=`$ $`{\displaystyle \frac{1}{L\sqrt{N}}}{\displaystyle \underset{𝐑_i}{}}\{\mathrm{exp}(i𝐊𝐑_i)(\mathrm{exp}(i\beta ))Y\varphi _1(𝐫𝐑_i)+X\varphi _2(𝐫𝐑_i))\},`$ (12) $`\psi _B(𝐫)`$ $`=`$ $`{\displaystyle \frac{1}{L\sqrt{N}}}{\displaystyle \underset{𝐑_i}{}}\{\mathrm{exp}(i𝐊𝐑_i)(Y\varphi _3(𝐫𝐑_i)\mathrm{exp}(i\alpha )X\varphi _4(𝐫𝐑_i))\},`$ (13) where $`X=\sqrt{2\mathrm{cos}\alpha },Y=\sqrt{2\mathrm{cos}\beta },L=\sqrt{2\mathrm{cos}\alpha +2\mathrm{cos}\beta },`$ and $`\alpha ={\displaystyle \frac{\sqrt{3}}{2}}K_xa,\beta ={\displaystyle \frac{\sqrt{3}}{2}}K_ya.`$ We introduce a wave vector k measured from a point on $`k`$-line and define two functions, $`\mathrm{\Phi }_{j,𝐤}(𝐫)=\mathrm{exp}(i𝐤𝐫)\psi _j(𝐫),`$ (14) with $`j=A,B`$. These functions are orthonormal: $`(\mathrm{\Phi }_{j,𝐤},\mathrm{\Phi }_{j^{},𝐤^{}})=\delta _{jj^{}}\delta _{\mathrm{𝐤𝐤}^{}}.`$ The wave function near the point can be expanded as $`\mathrm{\Psi }(𝐫)={\displaystyle \underset{j=A,B}{}}{\displaystyle \frac{d𝐤}{(2\pi )^3}C_j(𝐤)\mathrm{\Phi }_{j,𝐤}(𝐫)}={\displaystyle \underset{j=A,B}{}}F_j(𝐫)\psi _j(𝐫),`$ (15) where $`F_j(𝐫)`$ is the envelop function defined by $`F_j(𝐫)={\displaystyle \frac{d𝐤}{(2\pi )^3}\mathrm{exp}(i𝐤𝐫)C_j(𝐤)}.`$ (16) Substituting Eq.(15) into the Schrödinger equation, we obtain the following $`kp`$ equation, $`\{\begin{array}{c}\frac{1}{L^2}\mathrm{exp}(i(\alpha +\beta ))(\frac{\sqrt{3}}{2}a\mathrm{tan}\alpha \widehat{k_x}\frac{\sqrt{3}}{2}a\mathrm{tan}\beta \widehat{k_y}+i3a\widehat{k_z})F_A(𝐫)=EF_B(𝐫)\hfill \\ \frac{1}{L^2}\mathrm{exp}(i(\alpha +\beta ))(\frac{\sqrt{3}}{2}a\mathrm{tan}\alpha \widehat{k_x}\frac{\sqrt{3}}{2}a\mathrm{tan}\beta \widehat{k_y}i3a\widehat{k_z})F_B(𝐫)=EF_A(𝐫)\hfill \end{array},`$ (19) where $`\widehat{k}_{x,y,z}`$ are defined as $`\widehat{k}_{x,y,z}=i_{x,y,z}`$. Here, we rewrite the above expression with $`\widehat{k_{}}`$ and $`\widehat{k_{}}`$, which are components of the operator, $`\widehat{𝐤}`$, perpendicular to $`k`$-line and parallel to $`k`$-line in $`k_xk_y`$-plane, respectively. They are given by $`\left(\begin{array}{cc}\hfill \widehat{k}_{}& \\ \hfill \widehat{k}_{}& \end{array}\right)=\left(\begin{array}{cc}\hfill \mathrm{cos}\theta _𝐊& \hfill \mathrm{sin}\theta _𝐊\\ \hfill \mathrm{sin}\theta _𝐊& \hfill \mathrm{cos}\theta _𝐊\end{array}\right)\left(\begin{array}{cc}\hfill \widehat{k}_x& \\ \hfill \widehat{k}_y& \end{array}\right),\{\begin{array}{c}\mathrm{cos}\theta _𝐊=\frac{1}{\mathrm{\Delta }_𝐊}\mathrm{tan}\alpha \\ \mathrm{sin}\theta _𝐊=\frac{1}{\mathrm{\Delta }_𝐊}\mathrm{tan}\beta \end{array},`$ $`\mathrm{\Delta }_𝐊=\sqrt{\mathrm{tan}^2\alpha +\mathrm{tan}^2\beta }.`$ By using the components, $`\widehat{k}_{}`$ and $`\widehat{k}_{}`$, Eq. (19) becomes $`\left(\begin{array}{cc}0& \mathrm{exp}(i(\alpha +\beta ))(\xi _𝐊\widehat{k_{}}i\eta \widehat{k_z})\\ \mathrm{exp}(i(\alpha +\beta ))(\xi _𝐊\widehat{k_{}}+i\eta \widehat{k_z})& 0\end{array}\right)\left(\begin{array}{c}F_A(𝐫)\\ F_B(𝐫)\end{array}\right)=E\left(\begin{array}{c}F_A(𝐫)\\ F_B(𝐫)\end{array}\right),`$ (26) where $`\{\begin{array}{c}\xi _𝐊=\frac{\sqrt{3}}{2}a\mathrm{\Delta }_𝐊/L^2\hfill \\ \eta =3a/L^2\hfill \end{array}.`$ The final expression is the same as that of the $`kp`$ equation for the graphene around K (or K’) point. We can obtain the energy eigenvalue around $`k`$-line from (26), which is $`E=\sqrt{\xi _𝐊^2k_{}^2+\eta ^2k_z^2}.`$ (27) Here, $`k_{}`$ and $`k_z`$ are a component of the wave vector, $`𝐤`$, defined above. Thus, the energy dispersion around a $`k`$-line is in proportion to $`k`$, which is a distance from the $`k`$-point to the $`k`$-line in plane perpendicular to k-line. Next, we study the DS by using the $`kp`$ approximation. We have several $`k`$-lines in 1st BZ in this case. We choose a $`k`$-line, $`\{\begin{array}{c}K_y=0\hfill \\ K_z=\frac{2\pi }{a}\hfill \end{array},`$ (30) to demonstrate the $`kp`$ approximation. From the same procedure, we can obtain the following $`kp`$ equation, $`\left(\begin{array}{cc}0& \mathrm{exp}(i\frac{K_xa}{4})(\xi _{K_x}\widehat{k}_y+i\eta _{K_x}\widehat{k}_z)\\ \mathrm{exp}(i\frac{K_xa}{4})(\xi _{K_x}\widehat{k}_yi\eta _{K_x}\widehat{k}_z)& 0\end{array}\right)\left(\begin{array}{c}F_A(𝐫)\\ F_B(𝐫)\end{array}\right)=E\left(\begin{array}{c}F_A(𝐫)\\ F_B(𝐫)\end{array}\right),`$ (37) where $`\xi _{K_x}=a\mathrm{sin}({\displaystyle \frac{K_xa}{4}}),\eta _{K_x}=a\mathrm{cos}({\displaystyle \frac{K_xa}{4}}).`$ This equation is again the same form as that of the graphene. We can obtain the energy eigenvalue around the $`k`$-line from Eq.(37), which becomes $`E=\sqrt{\xi _{K_x}^2k_y^2+\eta _{K_x}^2k_z^2}.`$ (38) Since we consider k-line which is parallel to $`k_x`$-axis, this energy eigenvalue is proportional to a component $`k`$ of the wave number, $`𝐤`$, in a plane perpendicular to k-line, i.e the $`k_yk_z`$-plane. The point ($`0,0,\frac{2\pi }{a}`$), where Eq.(37) possesses a singularity, is an exception. The energy eigenvalue is proportional to only $`k_z`$ on this point, because $`\alpha _{K_x}`$ becomes zero. Dispersion on a $`k`$-line is proportional to a wave number for all direction in plane being perpendicular to the $`k`$-line except for some singular points. From this results, $`k`$-line is regarded as locus of a point where conduction and valence bands degenerate with $`k`$-linear dispersion, i.e. the $`K`$ (or $`K^{}`$) point of the BZ in the graphene. This is a reason why we use a word “ $`k`$-line”. ## IV DISCUSSION In this paper, we use a single-band tight binding model for explaining nature of networks which satisfy the conditions described in the section IIB. In the model, we assume that the hopping integral, $`t`$, is the same in any direction, in order to make a simple argument. But, this assumption is not necessarily required for networks having edge states. In a selected class of networks, we can prove the existence of edge states on them (See appendix A). Here, hypergraphite is defined to be a network which satisfies following two conditions. One is on the topology of a network given by hopping integrals and the other is on the electronic structure. The topology of network given by the hopping integrals is regarded as being constructed by the following method. 1. Prepare an ($`N1`$)-dimensional AB bipartite network having a NBO whose amplitude is finite only on A-sites. The number of A-sites and that of B-sites in a unit cell are equal. 2. Line up the copies of the network along z-direction. We define A-sites and B-sites on the $`i`$-th AB bipartite electronic network as the same way on the first network. 3. Connect B-sites on the first network with A-sites on the second network by extra bonds. 4. Connect B-sites on the $`i`$-th network with A-sites on the ($`i+1`$)-st network by the same way which is used in 3) ($`i=1,2,\mathrm{},\mathrm{}`$). Next, consider the s-TBM in the network, the electronic structure constructed above fashion has to show following characteristics. 1. Edge states appear in this network with a zigzag surface. 2. The highest valence band and the lowest conduction band degenerate at the Fermi energy and the Fermi surface becomes an ($`N2`$)-dimensional surface. 3. The highest valence band and the lowest conduction band possess $`k`$-liner dispersion at points on the ($`N2`$)-dimensional Fermi surface. 4. The bulk system is a zero-gap semiconductor. Here, note that the Fermi level is set on the center of the band structure, because we consider the case of a half-filled AB bipartite network. From the above definition on a network, the appearance of edge states on hypergraphite itself is natural. This is because a NBO which exists on a basic AB bipartite network coincides with a completely localized edge states in a hypergraphite. But it is not clear whether hypergraphite become a zero-gap semiconductor and whether it has bands with $`k`$-linear dispersion at the Fermi energy. This important point is left to be solved in general. Here, we list possible scenarios how hypergraphite networks and their characters are realized in real materials. We have shown that the three-dimensional three-fold coordinated network and the diamond structure are classified as the hypergraphite. From the point of geometric topology, the cubic diamond and the cubic silicon possess the topology of the diamond structure. We also find that the 3-3 network is created by silicon in $`\alpha `$-ThSi<sub>2</sub>. In these real materials, the electronic and geometric structures are determined by not only $`s`$-electrons but also $`p`$-electrons of carbon or silicon atoms. Actually, $`sp^3`$ or $`sp^2`$ hybridized orbitals form these interesting lattice structure and electronic properties. Hence, no electronic characteristic of hypergraphite is found around the Fermi energy. But, we notice the characteristics of the hypergraphite in the lowest two bands. Examples are 1) gap closing at X-point in the lowest two $`sp^3`$-hybridized bands of diamond or silicon and 2) gap closing at a point on the $`\mathrm{\Gamma }`$-X line in the lowest two $`sp^2`$-hybridized bands of CaSi<sub>2</sub> in the $`\alpha `$-ThSi<sub>2</sub> structure. This is because the two bands are mainly composed of 2$`s`$ or 3$`s`$ electrons. ## V SUMMARY We propose a class of $`N`$-dimensional networks. These networks have the same characteristics of topology of hopping integrals and an electronic structure that the graphene has. A characteristic of topology of hopping integrals is AB bipartite network. Particular properties of the electronic structure are 1) appearance of edge states under a zigzag surface and 2) existence of a $`k`$-line with $`k`$-linear dispersion, and 3) being a zero-gap semiconductor. From the above reason, we regard the class of networks as an extended graphitic network and name them “hypergraphite”. ## VI ACKNOWLEDGMENTS The author would like to thank M. Fujita, M. Igami, K. Wakabayashi, S. Okada, K. Nakada and K. Kusakabe. Numerical calculations were performed on the Fujitsu VPP500 of computer center at Institute for Solid States Physics, University of Tokyo and the NEC SX3 of computer center at Institute for Molecular Science, Okazaki National Institute. This work is supported by Grant-in-Aid for Scientific Research nos.10309003 and 11740392 from the Ministry of Education, Science and Culture, Japan. ## A Transfer Matrix Method In this appendix, we show that the edge states appear in a selected class of $`N`$-dimensional networks which are composed of ($`N1`$)-dimensional AB bipartite network. Naturally, we consider the $`N`$-dimensional networks which satisfy the conditions of hypergraphite. We consider a semi-infinite slab model, on which the single-band tight- binding model is constructed. In this model, hopping integrals do not have to be unity. We call an $`i`$-th ($`N1`$)-dimensional network $`\mathrm{\Lambda }^i`$. ($`i=1,2,\mathrm{},n,\mathrm{},\mathrm{}`$) It is assumed that there exists a NBO on $`\mathrm{\Lambda }^1`$. We consider only AB bipartite networks, where 1) the number of A-sites and that of B-sites are the same and 2) the NBO has a finite amplitude on every A-site of $`\mathrm{\Lambda }^1`$ and all of B-sites are node for the NBO. Since all of $`\mathrm{\Lambda }^i`$ has the same network as that of $`\mathrm{\Lambda }^1`$, all of $`\mathrm{\Lambda }^i`$ has a NBO which is the copy of the first NBO. Then, we naturally define A-sites and B-sites on $`\mathrm{\Lambda }^i`$ so that B-sites are always node for the NBO. A surface of the network is composed by only A-sites on $`\mathrm{\Lambda }^1`$. We also assume that the ($`n1`$)-dimensional AB bipartite network is a lattice having a unit cell and is periodic in all ($`n1`$)-directions. The numbers of A-sites and B-sites in the unit cell of $`\mathrm{\Lambda }^i`$ are equal and represented by $`N_0`$. Then, we can obtain wave functions on $`\mathrm{\Lambda }^i`$ in a Bloch form as, $`|\psi _i(𝐤)={\displaystyle \underset{j,l}{}}{\displaystyle \underset{𝐫_i}{}}u_{i,j,l}(𝐤)\mathrm{exp}(i𝐤𝐫_i)c_{i,𝐫_i,j,l}^{}|0.`$ (A1) Here $`𝐤`$ and $`𝐫_i`$ represent an ($`N1`$)-dimensional wave vector and a position vector for a unit cell on $`\mathrm{\Lambda }^1`$, respectively. Two labels, $`j`$ and $`l`$, indicate sublattices ($`j=`$ A or B) and a site of the $`j`$-sublattice in the cell ($`l=1,\mathrm{},N_0`$), respectively. An operator, $`c_{i,𝐫_i,j,l}^{}`$, creates an electron on a site indexed by $`j`$ and $`l`$ in a cell at $`𝐫_i`$ on $`\mathrm{\Lambda }^i`$. We introduce an $`N_0`$-dimensional vector $`𝐮_{i,j}=(u_{i,j,l}(𝐤))`$. Then we can rewrite the Schrödinger equation on the total $`N`$-dimensional network in a form as, $`\begin{array}{ccc}ϵ𝐮_{1,A}& =& T^{in}𝐮_{1,B},\hfill \\ ϵ𝐮_{i,A}& =& T^{in}𝐮_{i,B}+T^{ex}𝐮_{i1,B},\hfill \\ ϵ𝐮_{i,B}& =& T^{in}𝐮_{i,A}+T^{ex}𝐮_{i+1,A}.\hfill \end{array}`$ (A5) Here, matrices, $`T^{in}`$ and $`T^{ex}`$, are functions of k. An ($`N_0\times N_0`$) matrix $`T^{in}`$ represents the transfer from B-sites to A-sites in $`\mathrm{\Lambda }^i`$. $`T^{ex}`$ is another ($`N_0\times N_0`$) matrix representing transfers from B-sites in $`\mathrm{\Lambda }^{i1}`$ to A-sites in $`\mathrm{\Lambda }^i`$. $`T^{in}`$ is a matrix representing transfers from A-sites to B-sites in $`\mathrm{\Lambda }^i`$. In the first place, we discuss the simplest case, i.e. the unit cell in $`\mathrm{\Lambda }^i`$ has only one A site and one B site. Then $`T^{in},T^{ex}`$ and $`𝐮_{i,j}`$ become a scalar. ($`i=1,2,\mathrm{},n,\mathrm{}\mathrm{},j=AorB`$). We seek for a dumping wave with $`ϵ=0`$. Then we have the following equations with a transfer matrix, $`T_j`$, $`\left(\begin{array}{c}u_{(i+1),A}\\ u_{i,B}\end{array}\right)=T_A\left(\begin{array}{c}u_{i,B}\\ u_{i,A}\end{array}\right)=\left(\begin{array}{cc}0& \frac{T^{in}}{T^{ex}}\\ 1& 0\end{array}\right)\left(\begin{array}{c}u_{i,B}\\ u_{i,A}\end{array}\right),`$ (A14) $`\left(\begin{array}{c}u_{i,B}\\ u_{i,A}\end{array}\right)=T_B\left(\begin{array}{c}u_{i,A}\\ u_{i1,B}\end{array}\right)=\left(\begin{array}{cc}0& \frac{T^{ex}}{T^{in}}\\ 1& 0\end{array}\right)\left(\begin{array}{c}u_{i,A}\\ u_{i1,B}\end{array}\right),`$ (A23) with another condition, $`u_{1,B}=0`$. Here, we assume that $`T^{ex}`$ and $`T^{in}`$ are not equal to zero. This is natural because, a) if $`T^{in}`$ is zero, there exists a completely localized edge state for a given $`𝐤`$, and b) if $`T^{ex}=0`$, there is no solution with $`u_{1,A}0`$ and $`u_{1,B}=0`$. The problem is that on what conditions this set of equations has a dumping wave toward the $`z`$-direction. If we multiply two transfer matrices, we have, $`\left(\begin{array}{c}u_{i+1,B}\\ u_{i+1,A}\end{array}\right)=\left(\begin{array}{cc}\frac{T^{ex}}{T^{in}}& 0\\ 0& \frac{T^{in}}{T^{ex}}\end{array}\right)\left(\begin{array}{c}u_{i,B}\\ u_{i,A}\end{array}\right).`$ (A30) Then, we can easily see that a solution can be chosen to satisfy either $`u_{i,A}0`$ or $`u_{i,B}0`$. Because $`u_{1,B}=0`$ to satisfy the boundary condition of edge states, we have a solution, $`u_{i,A}=(1)^{i1}({\displaystyle \frac{T^{in}}{T^{ex}}})^{i1}u_{1,A}.`$ (A31) Thus, in region of reciprocal lattice space where $`𝐤`$ satisfies $`|T^{in}/T^{ex}|<1`$, we have a dumping wave. The dumping factor $`D(𝐤)`$ is given by $`T^{in}/T^{ex}`$. Next, we have to show that a wave vector satisfying $`|D(𝐤)|<1`$ exists in a given network whose $`T^{in},T^{ex}`$ and $`u_{i,j}`$ are a scalar. But, we have assumed that a zero energy states exist on $`\mathrm{\Lambda }^1`$, which is given by k<sub>0</sub> satisfying $`T^{in}(𝐤_0)=0`$. Namely, this state is a localized eigen-state. Thus existence of a solution with $`D=0`$ is assumed. In usual networks with finite number of bonds for each site, $`T^{in}`$ and $`T^{ex}`$ are analytic functions of $`k_j`$ ($`j=1,\mathrm{},N_0`$). Hence, one can assume the continuity of $`T^{in}`$ (and $`T^{ex}`$) as a function of $`k_j`$. Then we have solutions with $`0<|D|<1`$, which gives the degenerate edge states. The graphene and the 3-5 network shown in this paper correspond to this simplest case. In case of graphene, $`T^{ex}=1`$ and $`T^{in}=2\mathrm{cos}\frac{k}{2}`$. Hence, the dumping factor $`D`$ becomes $`2\mathrm{cos}\frac{k}{2}`$. While, in case of the 3-5 network, $`T^{ex}=1`$ and $`T^{in}=2(\mathrm{cos}\frac{k_x}{2}+\mathrm{cos}\frac{k_y}{2}`$). Hence, the dumping factor $`D`$ becomes $`D=2(\mathrm{cos}\frac{k_x}{2}+\mathrm{cos}\frac{k_y}{2}`$). So far, we have assumed that $`T^{ex},T^{in}`$ and $`u_{i,j}`$ are scalars. In general, $`T`$s and $`u_{i,j}`$ could be a matrix and a vector, respectively. Then there is a case which $`T^{ex}`$ become a singular matrix. Hence, it is impossible to derive analogue of Eqs.(A14) and (A23) from Eq.(A5). But, we may be able to obtain analogous determination equations for an $`N`$-dimensional network. In case when $`(T^{ex})^1`$ exist, we see that an eigenvector of each A-sites on $`\mathrm{\Lambda }_i`$ is given by $`D^{i1}u_{1,A}`$ and $`u_{i,B}=0`$. Thus, the eigenvector of each A-sites on $`\mathrm{\Lambda }_i`$ is represented by eigenvector of each A-sites on $`\mathrm{\Lambda }_{i1}`$ and a dumping factor $`D`$ as, $`𝐮_{i+1,A}=D𝐮_{i,A}.`$ (A32) Substituting Eq.(A32) to Eq.(A5), we have, $`T^{in}𝐮_{i,A}=DT^{ex}𝐮_{i,A}.`$ (A33) The dumping factor, $`D`$, is determined by this equation. We can judge whether edge states emerge at an edge of the $`N`$-dimensional network from it. When $`D=0`$, the edge state is completely localized at the surface. When $`0<|D|<1`$, the edge state becomes a dumping wave. When $`|D|=1`$, the state coincides with a bulk state. When $`1<|D|`$, the solution becomes unphysical. We apply this formulation to the 3-3 network. The equation becomes $`\left(\begin{array}{cc}(1+\mathrm{exp}(2ik_x)& \mathrm{exp}(i(k_x+k_y))\\ 1& (1+\mathrm{exp}(2ik_y))\end{array}\right)\left(\begin{array}{c}u_{1,A}\\ u_{2,A}\end{array}\right)=D\left(\begin{array}{cc}0& 0\\ 1& 0\end{array}\right)\left(\begin{array}{c}u_{1,A}\\ u_{2,A}\end{array}\right).`$ (A42) With this equation, we obtain a dumping factor for the 3-3 network. It is given by $`D=4\mathrm{cos}k_x\mathrm{cos}k_y`$.
warning/0001/hep-th0001013.html
ar5iv
text
# 1 Introduction ## 1 Introduction Randall and Sundrum recently explained the large hierarchy between the Planck scale and the weak scale in terms of a model where our observable universe corresponds to a $`(3+1)`$-dimensional boundary of a $`(4+1)`$-dimensional manifold. The extra spacelike dimension in this model is non-periodic and finite or eventually even of infinite length , and ordinary matter is restricted to the boundary while gravitational modes may propagate in the bulk. This model attracted a lot of attention, see and references there. Usually the Randall–Sundrum model is investigated under the assumption that the coupling of gravitational modes to the matter on the boundary is governed by the restriction of the Einstein tensor to the boundary. This led in particular to the claim that the Randall–Sundrum model with a small non-periodic dimension would yield antigravity . In the present paper we reconsider the gravitational evolution equations and the Newtonian limit of the Randall–Sundrum model. Contrary to the common assumption, the coupling of matter to gravity in this model does not appear through an Einstein equation on the boundary but through Neumann type boundary conditions for the Einstein equation in the bulk. This deviation from Einstein gravity on the boundary cures the antigravity problem: Gravity on the boundary is attractive and has a correct Newtonian limit. The observation that the Randall–Sundrum model implies first order equations for the metric on the boundary instead of second order equations does not depend on whether one uses the Einstein–Hilbert term or an Einstein term for the gravitational action in the bulk. However, with the Einstein–Hilbert term the system of gravitational evolution equations is overdetermined and therefore we use an Einstein term. To explain this, we point out several differences between the Randall–Sundrum model and Kaluza–Klein theory in the next section before we address the equations of motion for the metric and the Newtonian limit in the Randall–Sundrum model. To avoid confusion, we count the dimension of spaces with Minkowski signature explicitly in the form $`d+1`$, with $`d=3,4`$. ## 2 Differences to Kaluza–Klein theory One might presume that gravity should contribute an Einstein–Hilbert term to the action of the Randall–Sundrum model, and that gravity in $`3+1`$ dimensions should arise in a similar way as in a Kaluza–Klein theory with periodic dimensions. In Kaluza–Klein theories with small periodic internal dimensions the low-energy degrees of freedom are restricted to zero modes which are separated by a large mass gap from the massive modes. The zero modes are indepedent from the internal coordinates and the resulting low-energy theory is genuine $`(3+1)`$-dimensional. However, the periodicity constraints are instrumental for the emergence of the mass gap. Contrary to Kaluza–Klein theory, boundary conditions in a bounded non-periodic dimension have to be fixed dynamically by the equations of motion, and a priori this does not imply a restriction to zero modes separated by a mass gap. By the same token, we have to subtract a complete divergence from the Einstein–Hilbert term in the action of the Randall–Sundrum model. For an explanation of this point, consider the Einstein–Hilbert action with a cosmological term in the $`(4+1)`$-dimensional universe of the Randall–Sundrum model ($`d^5x=d^4xdx^5`$, $`0x^5L`$): $$S_{EH}=d^5x\sqrt{g}\left(\frac{\mu ^3}{2}g^{MN}R_{MN}\mathrm{\Lambda }\right),$$ $$\delta S_{EH}=d^5x\sqrt{g}\delta g^{MN}\left(\frac{\mu ^3}{2}\left(R_{MN}\frac{1}{2}g_{MN}g^{KL}R_{KL}\right)+\frac{1}{2}g_{MN}\mathrm{\Lambda }\right)$$ (1) $$+\frac{\mu ^3}{2}d^4x\sqrt{g}(g^{MN}\delta \mathrm{\Gamma }^5{}_{MN}{}^{}g^{5N}\delta \mathrm{\Gamma }^M{}_{MN}{}^{})|_{x^5=0}^L.$$ If matter degrees of freedom could propagate on the whole manifold and if the fields would be periodic, then $``$ the boundary terms would cancel and $``$ we could perform a Fourier decomposition of the degrees of freedom and throw away the massive Kaluza–Klein modes. This would then correspond to original Kaluza–Klein theory and yield low-dimensional Einstein gravity in the usual way. However, the space-time points $`x^0,\mathrm{}x^3,x^5=0`$ and $`x^0,\mathrm{}x^3,x^5=L`$ are different physical points in the Randall–Sundrum model and periodicity is not required (and cannot be required by causality). Furthermore, matter degrees of freedom are supposed to be fixed to the $`(3+1)`$-dimensional boundaries, and therefore variation of corresponding action principles yields homogeneous equations for the gravitational degrees of freedom in the bulk, while the coupling to the matter degrees of freedom arises from the variation on the boundaries. Below we will point out that the gravitational potential in this theory does not correspond to a three-dimensional Greens function for Dirichlet boundary conditions at infinity, but to a four-dimensional Greens function for Neumann boundary conditions on three-dimensional boundaries. A priori this implies deviations from the ordinary Newton potential in three dimensions. However, for a small non-periodic extra dimension of length $`L`$ the leading terms in the Newtonian limits of the Randall–Sundrum model and Einstein gravity agree if the naive estimate on the relation between the $`(3+1)`$\- and $`(4+1)`$-dimensional Planck masses (inferred from the corresponding relation in Kaluza–Klein theory) is augmented by a factor 3. Another difference to Kaluza–Klein theory concerns the fact, that for a small non-periodic extra dimension the deviations from the Newtonian limit of Einstein gravity are not suppressed by a term $`\mathrm{exp}(r/L)`$ but correspond to an expansion in $`r/V_3^{1/3}1`$, where $`V_3`$ is the 3-volume of a time slice of the $`(3+1)`$-dimensional boundary. As a consequence, in the large $`V_3`$ limit the gravitational potential has the usual form, with the correction term corresponding to a renormalization of the Planck mass. ## 3 The gravitational potential in the Randall–Sundrum model We have seen that the excitation of space-time curvature in the Randall–Sundrum model does not arise due to matter sources in the bulk equations, but through boundary conditions on the gravitational field arising from boundary equations of motion. This raises the issue of the Newtonian limit for the gravitational field on the boundary, which we examine through the $`(4+1)`$-dimensional action $$S=_𝒱d^5x\sqrt{g}(\frac{\mu ^3}{2}g^{KL}(\mathrm{\Gamma }^M{}_{NK}{}^{}\mathrm{\Gamma }_{}^{N}{}_{ML}{}^{}\mathrm{\Gamma }^M{}_{NM}{}^{}\mathrm{\Gamma }_{}^{N}{}_{KL}{}^{})\mathrm{\Lambda })+\underset{i=1}{\overset{2}{}}_{𝒱_i}d^4x_i.$$ (2) Here $`𝒱_i`$ are the two connected components of the $`(3+1)`$-dimensional boundary and $`_i`$ denotes the Lagrangians for the matter degrees of freedom on the boundary components. The Lagrangians $`_i`$ may also contain cosmological terms on the boundary. Coordinates $`x^0,x^1,x^2,x^3,x^5`$ are chosen such that the two boundary components correspond to $`x^5=0`$ and $`x^5=L`$, respectively. The gravitational part in (2) is fixed from two requirements: $``$ The Einstein tensor is the leading derivative term in any evolution equation for the metric on a Riemannian manifold, and this should also hold true in the present model, since there is no symmetry prohibiting this leading curvature term. $``$ At the same time, the full Einstein–Hilbert Lagrangian (as well as a leading higher curvature term in the action) would not give consistent boundary equations of motion, due to the second derivatives on the metric tensor: The divergence included in the Einstein–Hilbert Lagrangian yields boundary terms $`_5\delta g_{\mu \nu }`$ which have no counterpart in the $`\delta _i`$ terms and overdetermine the boundary value problem for the metric. Therefore, we used Einstein’s well-known Lagrangian (adapted to $`4+1`$ dimensions) $$_E=\frac{\mu ^3}{2}\sqrt{g}g^{KL}(\mathrm{\Gamma }^M{}_{NK}{}^{}\mathrm{\Gamma }_{}^{N}{}_{ML}{}^{}\mathrm{\Gamma }^M{}_{NM}{}^{}\mathrm{\Gamma }_{}^{N}{}_{KL}{}^{}).$$ This subtracts the divergence term from $`\sqrt{g}R`$ and yields the full Einstein tensor in the bulk. In analyzing (2) it is convenient to choose the bulk coordinate orthogonal to the boundaries: $`g_{\mu 5}|_{x^5=0,L}=0`$, $`0\mu 3`$. Variation of (2) yields again a sum of a $`(4+1)`$-dimensional integral and an integral over the boundary, implying gravitational equations of motion in the bulk $$R_{MN}=\frac{2\mathrm{\Lambda }}{3\mu ^3}g_{MN}$$ (3) and on the boundary: $$g^{\lambda \nu }_\mu g_{\lambda \nu }|_{x^5=0,L}=g^{55}_\mu g_{55}|_{x^5=0,L},$$ (4) $$_5g_{\mu \nu }|_{x^5=0}=\frac{2}{\mu ^3}g_{55}\left(T_{\mu \nu }^{(1)}\frac{1}{3}g_{\mu \nu }g^{\kappa \lambda }T_{\kappa \lambda }^{(1)}\right),$$ (5) $$_5g_{\mu \nu }|_{x^5=L}=\frac{2}{\mu ^3}g_{55}\left(T_{\mu \nu }^{(2)}\frac{1}{3}g_{\mu \nu }g^{\kappa \lambda }T_{\kappa \lambda }^{(2)}\right).$$ (6) Here $`g_{\mu \nu }`$ denotes the tangent components of the metric tensor on the boundary, and (4) arises from boundary terms $`\delta g^{5\mu }`$, while (5) and (6) arise from boundary terms $`\delta g^{\mu \nu }`$. No boundary terms $`\delta g^{55}`$ appear. The energy momentum tensors on the boundary components are $$T_{\mu \nu }^{(i)}=\frac{2}{\sqrt{g}}\frac{\delta _i}{\delta g^{\mu \nu }},$$ and as usual in this kind of variational problems, the boundary equations amount to boundary conditions for the bulk equations of motion. Eq. (4) has two implications: On the one hand it tells us that the determinant $`g_{(4)}`$ of the metric induced on the boundary determines the boundary value of $`g_{55}`$ up to a constant factor, and on the other hand it ensures invariance of (2) under diffeomorphisms $`x^Mx^Mϵ^M(x)`$ which leave the boundary invariant: $`ϵ^5(x)|_{x^5=0,L}=0`$. To examine the gravitational potential emerging in the Randall–Sundrum model, it is useful to reformulate the evolution equations for spatially closed $`(3+1)`$-dimensional boundary universes, i.e. we consider $`x^5`$ as a radial coordinate between two spherical shells at radii $`ax^5=rb`$. Eqs. (5,6) then read $$\frac{}{r}g_{\mu \nu }|_{r=a}=\frac{2}{\mu ^3}g_{55}\left(T_{\mu \nu }^{(1)}\frac{1}{3}g_{\mu \nu }g^{\kappa \lambda }T_{\kappa \lambda }^{(1)}\right).$$ (7) $$\frac{}{r}g_{\mu \nu }|_{r=b}=\frac{2}{\mu ^3}g_{55}\left(T_{\mu \nu }^{(2)}\frac{1}{3}g_{\mu \nu }g^{\kappa \lambda }T_{\kappa \lambda }^{(2)}\right).$$ (8) In the Newtonian approximation we consider weakly coupled gravitational systems on time scales much shorter than the age of the universe and length scales far below the Hubble radius. In this approximation cosmological background metrics can very well be approximated by a local Minkowski background. The weak field approximation $`g_{MN}=\eta _{MN}+h_{MN}`$ for static sources $`T_{00}^{(i)}=\varrho _i`$ on the boundary yields a Neumann type boundary problem for the gravitational potential $`U=h_{00}/2`$: $$\mathrm{\Delta }U=0,$$ (9) $$\frac{}{r}U|_{r=a}=\frac{2}{3\mu ^3}\varrho _1,$$ (10) $$\frac{}{r}U|_{r=b}=\frac{2}{3\mu ^3}\varrho _2.$$ (11) As a consequence, the gravitational interaction between matter components on the boundary arises through a four-dimensional Greens function adapted to Neumann boundary conditions: $$U(\text{r})=_Vd^3\text{r}^{}\left(G(\text{r},\text{r}^{})\frac{}{r^{}}U(\text{r}^{})U(\text{r}^{})\frac{}{r^{}}G(\text{r},\text{r}^{})\right)|_{r^{}=a}^{r^{}=b}$$ (12) $$=U\frac{2}{3\mu ^3}_{r^{}=a}d^3\text{r}^{}G(\text{r},\text{r}^{})\varrho _1(\text{r}^{})\frac{2}{3\mu ^3}_{r^{}=b}d^3\text{r}^{}G(\text{r},\text{r}^{})\varrho _2(\text{r}^{}).$$ Here $`U`$ is the average value of $`U`$ on the boundary, and $`d^3\text{r}^{}`$ is the spatial volume element on the boundary $`V`$ of a time slice $`V`$ of $`𝒱`$. The Greens function for the Neumann boundary problem is defined by the requirements $$\mathrm{\Delta }^{}G(\text{r},\text{r}^{})=\delta (\text{r}\text{r}^{}),$$ $$\frac{}{r^{}}G(\text{r},\text{r}^{})|_{r^{}=a}=\frac{}{r^{}}G(\text{r},\text{r}^{})|_{r^{}=b}=\frac{1}{2\pi ^2(a^3+b^3)}$$ and we have calculated it for a spatial four-manifold bounded by two concentric three-spheres: $$4\pi ^2G(\text{r},\text{r}^{})=\frac{1}{a^3+b^3}(\frac{b^3}{r_>^2}\frac{a^3}{r_<^2})+\underset{l=1}{\overset{\mathrm{}}{}}(\frac{r_<^l}{r_>^{l+2}}+\frac{l+2}{l}\frac{r^lr^l}{b^{2l+2}a^{2l+2}}$$ (13) $$+\frac{a^{2l+2}}{b^{2l+2}a^{2l+2}}(\frac{r^l}{r^{l+2}}+\frac{r^l}{r^{l+2}}+\frac{l}{l+2}\frac{b^{2l+2}}{r^{l+2}r^{l+2}}))\frac{\mathrm{sin}((l+1)\theta )}{\mathrm{sin}\theta }.$$ Here $`\theta `$ denotes the angle between the four-dimensional vectors r and $`\text{r}^{}`$, and like in three-dimensional multipole expansions $`r_<`$ is the smaller of the two radii $`r`$ and $`r^{}`$, while $`r_>`$ is the larger radius. If we choose the three-sphere at $`r=a`$ as the time slice of our $`(3+1)`$-dimensional universe in this scenario, the gravitational potential between ordinary matter sources and probes arises in the limit $`r,r^{}a`$, and the distance between source and probe within this three-sphere is $`d=a\mathrm{sin}\theta `$. Up to an irrelevant constant term the gravitational potential of a mass $`m`$ on the 3-sphere $`S_a^3`$ of radius $`a`$ follows from (13,12) $$U(\theta )=\frac{m}{6\pi ^2\mu ^3a^2}\underset{l=1}{\overset{\mathrm{}}{}}\left(1+\frac{a^{2l+2}}{b^{2l+2}a^{2l+2}}\left(3+\frac{2}{l}+\frac{l}{l+2}\frac{b^{2l+2}}{a^{2l+2}}\right)\right)\frac{\mathrm{sin}((l+1)\theta )}{\mathrm{sin}\theta }.$$ (14) Here $`\theta `$ is the angle between the source $`m`$ of the gravitational field and the point where it is probed. A large internal dimension corresponds to $`ba`$ and yields $$U(\theta )|_{ba}=\frac{m}{3\pi ^2\mu ^3a^2}\underset{l=1}{\overset{\mathrm{}}{}}\frac{l+1}{l+2}\frac{\mathrm{sin}((l+1)\theta )}{\mathrm{sin}\theta }.$$ (15) In the other case of small internal length $`La,b=a+L`$ we find $$U(\theta )|_{a,b=a+LL}=\frac{m}{3\pi ^2\mu ^3aL}\underset{l=1}{\overset{\mathrm{}}{}}\frac{l+1}{l(l+2)}\frac{\mathrm{sin}((l+1)\theta )}{\mathrm{sin}\theta }.$$ (16) For comparison, the genuine three-dimensional gravitational potential on a 3-sphere of radius $`a`$ is: $$𝒰(\theta )=\frac{m}{4\pi m_{Pl}^2a}\mathrm{cot}\theta =\frac{2m}{\pi ^2m_{Pl}^2a}\underset{l=1}{\overset{\mathrm{}}{}}\frac{l}{(2l1)(2l+1)}\frac{\mathrm{sin}(2l\theta )}{\mathrm{sin}\theta },$$ (17) where $`m_{Pl}=(8\pi G_N)^{1/2}`$ is the reduced Planck mass on $`S_a^3`$. As expected from a higher-dimensional potential, $`U(\theta )|_{ba}`$ has a stronger singularity for $`\theta 0`$ than the ordinary Newton potential on $`S_a^3`$. The case of small non-periodic extra dimension is more subtle: The odd-$`l`$ modes of $`U(\theta )|_{a,b=a+LL}`$ are absent in the classical inherently three-dimensional potential $`𝒰(\theta )`$, but the even modes agree if the naive Kaluza–Klein type relation between $`\mu `$ and $`m_{Pl}`$ is augmented by a factor 3: $$3\mu ^3L=m_{Pl}^2.$$ (18) Contrary to Kaluza–Klein theory, the relation (18) in the present theory eliminates the parameter $`L`$ completely from the correction term to the ordinary Newton potential: $$U(\theta )|_{a,b=a+LL}𝒰(\theta )=\frac{m}{4\pi ^2m_{Pl}^2a}\underset{l=1}{\overset{\mathrm{}}{}}\frac{2l+1}{l(l+1)}\frac{\mathrm{sin}((2l+1)\theta )}{\mathrm{sin}\theta },$$ (19) and therefore the corrections to the Newton potential are not suppressed by a factor $`\mathrm{exp}(d/L)`$ but correspond to an expansion in $`d/a`$. In the limit $`a\mathrm{}`$, the correction term corresponds to a renormalization of the three-dimensional Planck mass, but the functional dependence on the distance between source and probe is just that of the three-dimensional Newton potential. We finally would like to point out that the 4D Poincaré invariant metric of Randall and Sundrum complies with the boundary equations (46): In the present conventions the metric arises from the Ansatz $$g_{\mu \nu }=\mathrm{exp}(2\sigma (x^5))\eta _{\mu \nu },g_{55}=1$$ under the assumption of boundary cosmological terms: $$_1=\lambda _1\mathrm{exp}(4\sigma (0)),$$ $$_2=\lambda _2\mathrm{exp}(4\sigma (L))$$ corresponding to boundary energy momentum tensors $$T_{\mu \nu }|_{x^5=0}=\lambda _1g_{\mu \nu }|_{x^5=0},$$ $$T_{\mu \nu }|_{x^5=L}=\lambda _2g_{\mu \nu }|_{x^5=L}.$$ The Einstein equation in the volume yields again (cf. eq. (7) in Ref. ) $$\sigma _{}^{}{}_{}{}^{2}=\frac{\mathrm{\Lambda }}{6\mu ^3},$$ and the boundary equations imply $$\sigma ^{}=\frac{1}{3\mu ^3}\lambda _1=\frac{1}{3\mu ^3}\lambda _2,$$ i.e. eq. (11) from Ref. is only rescaled by a factor 4 $$\mathrm{\Lambda }=\frac{2}{3\mu ^3}\lambda _i^2.$$ We conclude that the criticism of the small-$`L`$ Randall–Sundrum model was based on an incorrect set of gravitational evolution equations and not justified. Gravity in the Randall–Sundrum model is not repulsive, and it has a correct Newtonian limit. Acknowledgement: RD thanks Richard Altendorfer for an interesting discussion.
warning/0001/cond-mat0001387.html
ar5iv
text
# Reaction-Controlled Diffusion ## I Introduction There has been considerable effort to elucidate the properties and conditions of anomalous diffusive behavior. A simple physical realization is given by diffusion on a fractal lattice , where due to the increasing number of paths within the lattice, the time for a diffusion process will be prolongated. Also diffusion in random media with quenched disorder may be anomalous. Depending on the distribution of barrier heights (or depths of traps) one may observe normal diffusive or subdiffusive behavior, respectively, if the available number of diffusive paths is reduced by the presence of obstacles . Here we discuss a quite different situation in which diffusion is activated by the presence of particles or excitations which also propagate diffusively, but in the course of time decay. As a result the activated diffusion is rendered anomalous because the number of available paths decreases with time. However, the resulting structure of diffusive paths is not static, but evolves temporally. One may call this phenomenon dynamical fractality or dynamical disorder, depending on how the spatial distribution of excitations evolves in time. We model this scenario by starting from a two-component system consisting of distinct particle species $`A`$ and $`B`$, with local time-dependent densities $`\rho _A(\stackrel{}{x},t)`$ and $`\rho _B(\stackrel{}{x},t)`$. An $`A`$ particle is allowed to perform hopping processes between adjacent neighboring sites on a lattice, provided there are one or more $`B`$ particles present in its vicinity. To be more specific, an $`A`$ particle hops from a site $`j`$ to a neighboring point $`i`$ subject to the condition that this site $`i`$ is already occupied with a particle of species $`B`$, and with a rate proportional to the local $`B`$ particle density. Obviously, such an effective attractive interaction strongly influences the diffusive mobility of the $`A`$ species: Their mean-square displacement $`\stackrel{}{x}_A(t)^2`$ will depend on the time evolution of the local $`B`$ density $`\rho _B(\stackrel{}{x},t)`$. A non-trivial temporal behavior for $`\rho _B(\stackrel{}{x},t)`$ will result if we submit the $`B`$ species locally to diffusion-limited reactions such as $`n`$-th order annihilation $`nB\mathrm{}`$ (at the same or adjacent lattice points) or combined annihilation ($`n2`$) and spontaneous offspring production $`B(m+1)B`$ (the $`B`$ particles then perform branching and annihilating random walks, BARW). Once the time-dependence of $`\rho _B(t)`$ has been determined, we shall see that the $`A`$ kinetics is in the long-time limit to good approximation described on the basis of the associated mean-field rate equation. When the $`B`$ species is in an active state, i.e., $`\rho _B(\stackrel{}{x},t\mathrm{})=\rho _B^{\mathrm{}}>0`$, with a basically homogeneous distribution in space, the $`A`$ particles will display normal diffusive behavior, with a diffusion constant $`D_A\rho _B^{\mathrm{}}`$. In such a situation one has dynamical disorder, but there is always a finite fraction of sites available for hopping. However, an inactive phase, or the BARW critical point, are described either by an exponential decay $`\rho _B(\stackrel{}{x},t\mathrm{})e^{\lambda t}`$, in which case the $`A`$ particles remain localized, or by a power-law decrease $`\rho _B(\stackrel{}{x},t\mathrm{})t^\alpha `$ with a characteristic exponent $`\alpha >0`$. The diminishing density of $`B`$ particles reduces the induced mobility of the $`A`$ species, and these competing effects lead to subdiffusive behavior $`\stackrel{}{x}_A(t)^2t^{1\alpha }`$ for $`\alpha <1`$. In the borderline case $`\alpha =1`$ one has merely logarithmic growth $`\stackrel{}{x}_A(t)^2\mathrm{ln}t`$. Intuitively some of this behavior can be easily understood in the limit of vanishing diffusion of the B-particles. We allow for multiple occupation, i.e., the site occupation number can be any integer between zero and infinity. When the occupation number of a $`B`$ particles at a certain lattice point is nonzero, that site is available for the $`A`$ species. In an inactive state the number of $`B`$ particles will be permanently reduced by reactions leading to a decreasing density of available sites for the $`A`$ species. This procedure can be viewed as an effective “thinning out” of lattice sites that leads to subdiffusive behavior for the $`A`$ particles reminiscent of (but distinct to) the mathematically considerably more complex phenomenon of diffusion on a fractal lattice. ## II The model Here we present a more precise definition of our model in terms of a master equation which we formulate in the standard Fock-space formulation , sometimes called “quantum Hamiltonian formalism” , particularly for particles with hard-core repulsion. Physically, this interaction is usually insignificant unless exclusion between different particle species or external driving forces need to be taken into account. This is intuitively clear for annihilating processes where the particle density tends to zero at late times , but remains true also in the absence of particle reactions even in one dimension. In some models, however, e.g., the annihilation-fission model or pair contact process with diffusion , site occupation number constraints do play a crucial role. Hence, usually the specific choice of model is prescribed by the mathematical treatment used to analyze it. In the present context where we shall employ mean-field techniques and renormalization group arguments it is more advantageous to consider particles without site exclusion (for recent reviews, see e.g. Refs. ). We consider a system consisting of two different types of particles denoted as $`A`$ and $`B`$. The time evolution can be represented through an evolution operator $`L`$ . The corresponding annihilation and creation operators are written as $`a_i`$ ($`b_i`$) and $`a_i^{}`$ ($`b_i^{}`$), where the index indicates a lattice point in $`d`$ space dimensions. For example, the normal hopping process of species $`B`$ from a site $`j`$ to its neighbor $`i`$ is described by the evolution operator $`D(b_i^{}b_jb_j^{}b_j)`$, and for the entire lattice therefore $$L_B=D\underset{(ij)}{}\left(b_i^{}b_j^{}\right)\left(b_jb_i\right),$$ (1) where $`D`$ is the hopping rate or diffusion constant. An analogous expression would describe free diffusion of the $`A`$ particles. Here, however, we examine the situation that such a process is only allowed if there is at least one $`B`$ particle present at site $`i`$. If no representative of species $`B`$ is available at that site, an $`A`$ particle cannot move there. The time evolution operator for that process is proportional to $`(a_i^{}a_ja_j^{}a_j)b_i^{}b_i`$. The corresponding hopping process will occur provided an $`A`$ particle is in fact present at site $`j`$ and at least one $`B`$ particle occupies site $`i`$. Moreover, its rate is actually proportional to the number of $`B`$ particles present at site $`i`$. For the full system we obtain $$L_A=\stackrel{~}{D}\underset{(i,j)}{}\left(a_i^{}a_j^{}\right)\left(a_jb_i^{}b_ia_ib_j^{}b_j\right).$$ (2) Here, $`\stackrel{~}{D}`$ denotes the induced transition rate for the dynamical process of species $`A`$. In contrast to species $`A`$, the $`B`$ particles are subject to local reactions. A decreasing number of $`B`$ particles will lead to a slowing down for the motion of $`A`$’s through the lattice. For $`n`$-th order annihilation reactions $`nB\mathrm{}`$, the nonequilibrium evolution operator reads $$L_R=\lambda _n\underset{i}{}\left(1b_{i}^{}{}_{}{}^{n}\right)b_i^n.$$ (3) Obviously, this operator describes the annihiliation of $`n`$ particles of type $`B`$ at a lattice site $`i`$ provided such particles are available; $`\lambda _n`$ denotes the corresponding rate. Similarly, spontaneous branching processes $`B(m+1)B`$ with rate $`\sigma _m`$ are described by $$L_P=\sigma _m\underset{i}{}\left(b_{i}^{}{}_{}{}^{m}1\right)b_i^{}b_i.$$ (4) Together with Eq. (1), $`L_R`$ and $`L_P`$ represent the time evolution operator for branching annihilating random walks (BARW). The complete dynamics is determined by $`L=L_A+L_B+L_R(+L_P)`$, and may be encoded into a time-dependent “state vector” $$|F(t)=\underset{n_i}{}P(\stackrel{}{n},t)|\stackrel{}{n}.$$ (5) Here $`P(\stackrel{}{n},t)`$ is the evolving probability distribution for the unrestricted site occupation numbers $`\stackrel{}{n}=\{n_i\}`$ for both $`A`$ and $`B`$ particles, and $`|\stackrel{}{n}`$ a basic vector containing all possible entries $`n_i=0,1,2,\mathrm{}\mathrm{}`$, i.e., the eigenvalues of the second-quantized bosonic particle number operators $`a_i^{}a_i`$ and $`b_i^{}b_i`$, respectively. The state $`|0`$ represents the vacuum with no particles present, $`a_i|0=0=b_i|0`$. The state vector obeys the equation of motion $$_t|F(t)=L|F(t),$$ (6) or formally $`|F(t)=e^{Lt}|F(0)`$. The nonequilibrium operator $`L`$ corresponds to, and is obtained from the evolution operator $`L^{}`$ of the classical master equation that can generally be written as $$_tP(\stackrel{}{n},t)=L^{}P(\stackrel{}{n},t),$$ (7) and the matrix elements of $`L`$ and $`L^{}`$ are uniquely related to each other. The time-dependent average of an arbitrary physical quantity $`G(\stackrel{}{n})`$ with the probability distribution $`P(\stackrel{}{n},t)`$ can be cast into a “matrix element” form for the corresponding second-quantized operator $`G(t)`$ $$G(t)=\underset{n_i}{}P(\stackrel{}{n},t)G(\stackrel{}{n})=\mathrm{\Psi }|G|F(t),$$ (8) with the projection state $`\mathrm{\Psi }|=0|\mathrm{exp}_i(a_i+b_i)`$. Using the relation $`\mathrm{\Psi }|L=0`$, the evolution equation for an arbitrary operator $`G`$ becomes $$_tG=\mathrm{\Psi }|[G,L]|F(t).$$ (9) All the dynamical equations governing the classical problem are thus determined by the commutation rules of the underlying operators and the structure of the evolution operator $`L`$. In our case the dynamics of the model is given by induced hopping processes for the $`A`$ particles and diffusion-limited reactions for the $`B`$ species, which we shall assume to be distributed randomly at the initial time $`t=0`$. As a final step, we employ coherent basis states to represent the matrix element (8) by means of a path integral , and take the continuum limit. Absorbing factors containing the lattice constant into the diffusion and reaction rates, we may compute averages with a dynamical weight $`\mathrm{exp}\left(𝒜[\widehat{a},a,\widehat{b},b]\right)`$ consisting of contributions to the bosonic field action $`𝒜`$ describing the ordinary $`B`$ diffusion $$𝒜_B[\widehat{b},b]=d^dx𝑑t\widehat{b}\left(_tbD^2b\right),$$ (10) the pure $`n`$-th order annihilation reactions $$𝒜_R[\widehat{b},b]=\lambda _nd^dx𝑑t\left(1\widehat{b}^n\right)b^n,$$ (11) or offspring production processes, $$𝒜_P[\widehat{b},b]=\sigma _md^dx𝑑t\left(1\widehat{b}^m\right)\widehat{b}b,$$ (12) respectively. Finally the $`A`$ diffusion as induced by the coupling to the $`B`$ species, $`𝒜_A[\widehat{a},a,\widehat{b},b]={\displaystyle }d^dx{\displaystyle }dt\widehat{a}[_ta\stackrel{~}{D}(^2a)\widehat{b}b`$ (13) $`+\stackrel{~}{D}a^2(\widehat{b}b)].`$ (14) Notice that $`\widehat{b}(\stackrel{}{x},t)b(\stackrel{}{x},t)`$ represents the local density $`\rho _B(\stackrel{}{x},t)`$ (when appropriate ensemble averages are taken); the $`A`$ diffusion is thus mediated by the presence of $`B`$ particles. We remark that apart from the continuum limit, the mapping of the master equation onto the above field theory is exact and involves no further approximations. (We have omitted the boundary contributions stemming from the initial conditions and the projection state here.) ## III General considerations Clearly, the dynamic process for the $`A`$ particles as defined above is induced by the coupling to the reactive $`B`$ species only. When there are no $`B`$ particles present, $`\rho _B(x,t)=0`$, the $`A`$ dynamics obviously ceases. Indeed, it turns out that there appears no noise in the dynamic equation governing the $`A`$ kinetics, which would formally appear as a contribution $`\widehat{a}\widehat{a}`$ (or higher powers of $`\widehat{a}`$) in the dynamic functional. In fact, any stochasticity emerges as a result of spatio-temporal fluctuations for the $`B`$ species (essentially reaction noise here). In order to further elucidate this point, we may derive effective Langevin-type equations for the local densities $`\rho _A`$ and $`\rho _B`$. To this end, we need to perform the shifts $`\widehat{a}=1+\stackrel{~}{a}`$, $`\widehat{b}=1+\stackrel{~}{b}`$, which take care of the annihilation operators appearing in the projection state $`\mathrm{\Psi }|`$, see Ref. . To be specific, let us consider the case of $`B`$ pair annihilation reactions. Omitting temporal boundary terms describing the initial configuration, the new action becomes $`𝒜[\widehat{a},a,\widehat{b},b]={\displaystyle }d^dx{\displaystyle }dt[\stackrel{~}{a}(_taD^{}^2a\stackrel{~}{D}(^2a)b`$ (15) $`+\stackrel{~}{D}a(^2b))\stackrel{~}{D}\stackrel{~}{a}(^2a)\stackrel{~}{b}b+\stackrel{~}{D}\stackrel{~}{a}a^2(\stackrel{~}{b}b)`$ (16) $`+\stackrel{~}{b}(_tbD^2b+2\lambda b^2)+\lambda \stackrel{~}{b}^2b^2].`$ (17) Here, we have allowed for additional ordinary $`A`$ diffusion processes with rate $`D^{}`$. This dynamic action is equivalent to the following set of coupled Langevin equations, $`_ta=D^{}^2a+\stackrel{~}{D}(^2a)b\stackrel{~}{D}a(^2b)+\zeta ,`$ (18) $`_tb=D^2b2\lambda b^2+\eta ,`$ (19) where the fluctuating forces with zero mean are characterized by the noise correlations $`\zeta (\stackrel{}{x},t)\zeta (\stackrel{}{x}^{},t^{})=0,`$ (20) $`\zeta (\stackrel{}{x},t)\eta (\stackrel{}{x}^{},t^{})=\stackrel{~}{D}[^2a(\stackrel{}{x},t)]b(\stackrel{}{x},t)\delta (\stackrel{}{x}\stackrel{}{x}^{})\delta (tt^{})`$ (21) $`\stackrel{~}{D}a(\stackrel{}{x},t)^2[b(\stackrel{}{x},t)\delta (\stackrel{}{x}\stackrel{}{x}^{})\delta (tt^{})],`$ (22) $`\eta (\stackrel{}{x},t)\eta (\stackrel{}{x}^{},t^{})=2\lambda b(\stackrel{}{x},t)^2\delta (\stackrel{}{x}\stackrel{}{x}^{})\delta (tt^{}).`$ (23) Taking averages, we may then identify $`\rho _A(t)=a(\stackrel{}{x},t)`$ and $`\rho _B(t)=b(\stackrel{}{x},t)`$, as Eqs. (18) and (19) obviously generalize the mean-field rate equations for the local particle densities. The reaction noise for the $`B`$ species displays the characteristic negative correlations (“imaginary noise”), which reflect the particle anticorrelations induced by the annihilation reaction . When there are no $`B`$ particles left \[$`b(\stackrel{}{x},t)=0`$\], the fluctuations cease, characteristic of an absorbing inactive state. As anticipated, no noise contributions exist for the pure $`A`$ dynamics, but there appear $`A`$-$`B`$ noise cross-correlations. (Notice that pure diffusion noise does not appear explicitly here.) Next, let us study what happens when the $`A`$ particles are subject to an additional random force that leads to ordinary diffusion, i.e., the term $`D^{}`$ in the action (17). Obviously, one should expect that the induced diffusion $`\stackrel{~}{D}`$ is suppressed in this situation, and in the long-time limit standard diffusion prevails. This becomes indeed clear through simple power counting, introducing a momentum scale $`\kappa `$, i.e., $`[x]=\kappa ^1`$, and measuring time scales as $`[t]=\kappa ^2`$, as appropriate for diffusive dynamics. Then $`[D]=[D^{}]=\kappa ^0`$ become dimensionless, and we infer the field scaling dimensions $`[\widehat{a}]=[\widehat{b}]=[\stackrel{~}{a}]=[\stackrel{~}{b}]=\kappa ^0`$ and $`[a]=[b]=\kappa ^d`$, as to be expected for $`d`$-dimensional particle densities. The remaining couplings (reaction rates) acquire the scaling dimensions $$[\sigma _m]=\kappa ^2,[\lambda _n]=\kappa ^{2(n1)d},[\stackrel{~}{D}]=\kappa ^d.$$ (24) A positive scaling dimension means that the corresponding parameter is relevant in the renormalization-group (RG) sense. E.g., the branching rate $`\sigma _m`$ carries the dimensions of a “mass” term, and indeed represents the decisive control parameter for BARW: In mean-field theory, the critical point must be at $`\sigma _m=0`$, and is therefore described by the pure annihilation model, while for any positive $`\sigma _m`$ there will be only an active phase characterized by exponential correlations. The annihilation rate is relevant for $`d<2/(n1)`$ dimensions, and irrelevant for $`d>2/(n1)`$ . Hence we identify the upper critical dimension, below which fluctuations in fact dominate the asymptotic behavior, as $`d_c(n)=2/(n1)`$ for $`n`$-th order annihilation processes . Thus, for $`n>3`$ fluctuations are not too important in any physical dimension $`d1`$. Furthermore we notice that the coupling $`\stackrel{~}{D}`$ is irrelevant, i.e., compared to the other parameters in the theory its influence should become negligible in the asymptotic long-time, long-wavelength limit. Evidently, $`\rho _B(\stackrel{}{x},t)`$ either vanishes (inactive phase) or approaches a constant $`\rho _B^{\mathrm{}}`$ (active phase) as $`t\mathrm{}`$. In the former case, normal $`B`$ diffusion, if present ($`D^{}>0`$), will dominate; in the latter situation, the combined quantity $`\stackrel{~}{D}\rho _B^{\mathrm{}}`$ will effectively act as an ordinary diffusion constant, numerically renormalizing $`D^{}`$. In any case, we see that the ordinary $`A`$ diffusion process is not qualitatively affected by the induced hopping through attractive coupling to the $`B`$ density, and the associated noise cross-correlations. Also when $`D^{}=0`$, as in our original model, and in a system with an initially finite number of $`B`$ particles, asymptotically the $`A`$ particles either remain localized or display standard diffusion. In this respect, in numerical simulations the induced anomalous diffusion in which we are interested here would appear as a crossover feature in the long-time kinetics and correspond to corrections to scaling to the leading asymptotic time dependence. In an infinite system, however, with initially finite $`B`$ density, the anomalous diffusion regime will persist indefinitely. A corollary of these observations is that the rate $`\stackrel{~}{D}`$ does not acquire any non-trivial frequency or time dependence in the infrared. In the field theory language, we note that neither diffusive propagator for the $`A`$ or $`B`$ species can be renormalized by the $`(\widehat{a}\widehat{b}ba)`$ four-point vertex in the unshifted action (14), or equivalently, the three- and four-point vertices in the shifted action (17). Consequently, the renormalization for the vertex functions $`\mathrm{\Gamma }_{\widehat{a}\widehat{b}ba}`$ or $`\mathrm{\Gamma }_{\stackrel{~}{a}\stackrel{~}{b}ba}`$ and $`\mathrm{\Gamma }_{\stackrel{~}{a}ba}`$, respectively, can be determined to all orders in the perturbation expansion (with respect to $`\stackrel{~}{D}`$) by means of a Bethe–Salpeter equation, or equivalently, a geometric series of loops containing just the $`A`$ and $`B`$ propagator. This leads to the renormalized wavevector- and frequency-dependent coupling $$\stackrel{~}{D}_R(\stackrel{}{q},\omega )=\stackrel{~}{D}\left[q^2+\stackrel{~}{D}\frac{d^dp}{(2\pi )^d}\frac{p^2[(\stackrel{}{q}\stackrel{}{p})^2p^2]}{i\omega +D^{}p^2+Dp^2}\right]^1,$$ (25) where $`\stackrel{}{q}`$ and $`\omega `$ denote the momentum and frequency transfer between the $`A`$ and $`B`$ particles. We may now set $`D^{}=0`$ again, and investigate the long-wavelength limit $`\stackrel{}{q}0`$, $$\frac{}{q^2}\stackrel{~}{D}_R(\stackrel{}{q},\omega )|_{q=0}=\stackrel{~}{D}\left[1+\frac{\stackrel{~}{D}}{D}\left(\frac{\omega }{D}\right)^{\frac{d}{2}}\frac{d^dk}{(2\pi )^d}\frac{k^2}{i+k^2}\right]^1,$$ (26) where $`k^2=Dp^2/\omega `$. Thus, as $`\omega 0`$, the fluctuation corrections vanish (provided the integral is regularized in the ultraviolet with an appropriate cutoff), and the renormalized coefficient $`\stackrel{~}{D}_R`$ in Eq. (26) approaches the original “bare” constant $`\stackrel{~}{D}`$. This is to be contrasted with the infrared-singular behavior of e.g. the $`B`$ pair annihilation rate, for which an analogous procedure yields $`\lambda _R(\stackrel{}{q},\omega )=\lambda \left[1+{\displaystyle \frac{\lambda }{D}}{\displaystyle \frac{d^dp}{(2\pi )^d}\frac{1}{i\omega /D+q^2/4+p^2}}\right]^1,`$ (27) $`\lambda _R(0,\omega )=\lambda \left[1+{\displaystyle \frac{\lambda }{D}}\left({\displaystyle \frac{\omega }{D}}\right)^{\frac{d2}{2}}{\displaystyle \frac{d^dk}{(2\pi )^d}\frac{1}{i+k^2}}\right]^1.`$ (28) For $`d>d_c(2)=2`$, again $`\lambda _R(0,0)=\lambda `$ is just the original rate constant, resulting in the mean-field power law $`\rho _B(t)t^1`$. However, for $`d<d_c(2)=2`$, $`\lambda _R(0,\omega )\omega ^{1d/2}`$ vanishes for low frequencies. Inserting the corresponding effective time-dependent rate $`\lambda _R(t)t^{1+d/2}`$ into Eq. (19) leads to the correct slower algebraic decay $`\rho _B(t)t^{d/2}`$. In summary, the $`B`$ process itself is, per definition of our model, not influenced by the $`A`$ dynamics. In the renormalization group treatment, this is reflected by the fact that the coupling $`\stackrel{~}{D}`$ is irrelevant, and thus does not affect the long-time behavior. Yet the induced hopping rate $`\stackrel{~}{D}`$ is of course crucial for the $`A`$ species kinetics, and must be kept even in the mean-field approximation. We may thus solve for the $`B`$ kinetics first, and then explore its influence on the induced $`A`$ diffusion. Henceforth, we shall again set $`D^{}=0`$, as otherwise simple ordinary $`A`$ diffusion would ensue, with $`\stackrel{~}{D}`$ then irrelevant also for the $`A`$ kinetics, and the entire coupling of the $`A`$ and $`B`$ processes would disappear asymptotically. In the following, we shall study the $`A`$ kinetics assuming a spatially homogeneous, but time-dependent distribution of $`B`$ particles, which leads us to a mean-field description. ## IV Mean-field evolution equations ### A Annihilation kinetics Let us assume we can ignore spatial fluctuations for the $`B`$ species entirely, and ignore the reaction noise. For the $`n`$-th order annihilation processes, we saw that this as at least a qualitatively correct description for $`d>d_c(n)=2/(n1)`$, i.e., for $`d>2`$ in the case of pair annihilations, $`d>1`$ for the third-order process $`3B\mathrm{}`$, and in any physical dimension for $`n>3`$. The evolution equation can either be obtained directly from the non-equilibrium operator $`L_R`$ in Eq. (3) and the equation of motion (9), or from solving for the stationarity condition $`\delta 𝒜/\delta \widehat{b}=0`$ for the action $`𝒜=𝒜_B+𝒜_R`$, Eqs. (10) and (11), setting $`D=0`$. (Notice that $`\delta 𝒜/\delta b=0`$ is always solved by $`\widehat{b}=1`$.) Either procedure results in the obvious mean-field rate equation $$_t\rho _B(t)=n\lambda _n\rho _B(t)^n,$$ (29) which is readily integrated for $`n>1`$, $$\rho _B(t)=\frac{\rho _B(0)}{\left(1+t/\tau \right)^{1/(n1)}},\tau =\frac{\rho _B(0)^{1n}}{n(n1)\lambda _n},$$ (30) i.e., for $`t\tau `$ the $`B`$ density decays algebraically $`t^{1/(n1)}`$ in this approximation, while of course for $`n=1`$ $$\rho _B(t)=\rho _B(0)e^{\lambda _1t}.$$ (31) In the same manner, we may obtain the evolution equation for the $`A`$ species, or just consider Eq. (18) for $`D^{}=0`$ and vanishing noise. In the spirit of mean-field theory, we assume a homogeneous $`B`$ density, and obtain $$_t\rho _A(\stackrel{}{x},t)=\stackrel{~}{D}\rho _B(t)^2\rho _A(\stackrel{}{x},t).$$ (32) Again, this equation can be solved exactly, considering a delta-like density distribution for the $`A`$ species at the initial time $`t=0`$. As in this mean-field approach the $`\rho _B(t)`$ is assumed to be spatially uniform, the $`A`$ species will be Gaussian distributed in space, just like in ordinary diffusion, $$\rho _A(\stackrel{}{x},t)=\left(\frac{1}{2\pi \stackrel{}{x}_A^2(t)}\right)^{d/2}\mathrm{exp}\left(\frac{\stackrel{}{x}^2}{2\stackrel{}{x}_A^2(t)}\right).$$ (33) However, the $`B`$ decay (or lattice depletion) will be reflected in the anomalous time-dependence of the width (mean-square displacement). A straightforward brief calculation yields $$\stackrel{}{x}_A^2(t)=2\stackrel{~}{D}_0^t\rho _B(t^{})𝑑t^{}.$$ (34) For $`n=1`$, i.e., the simple exponential decay (31), the result is $$\stackrel{}{x}_A^2(t)=\frac{2\stackrel{~}{D}\rho _B(0)}{\lambda _1}\left(1e^{\lambda _1t}\right).$$ (35) Initially ($`\lambda _1t1`$) one finds normal diffusion with effective diffusion constant $`\overline{D}=\stackrel{~}{D}\rho _B(0)`$, but at long times the mean-square displacement approaches a constant, and the $`A`$ particles remain localized in a region of volume $`\stackrel{}{x}_A^2(t\mathrm{})^{d/2}=(2\stackrel{~}{D}\rho _B(0)/\lambda _1)^{d/2}`$. Given that this simple process is characterized by short-range correlations in space and time only, we do not expect any considerable modification through fluctuation effects. In the pair annihilation case, $`n=2`$, one finds $$\stackrel{}{x}_A^2(t)=2\stackrel{~}{D}\rho _B(0)\mathrm{ln}\left(1+\frac{t}{\tau }\right),$$ (36) while the mean-field result for $`n>2`$ reads $$\stackrel{}{x}_A^2(t)=2\stackrel{~}{D}\rho _B(0)\frac{n1}{n2}\tau \left[\left(1+\frac{t}{\tau }\right)^{\frac{n2}{n1}}1\right].$$ (37) In the asymptotic regime $`t\tau `$, this implies anomalous diffusion according to $$\stackrel{}{x}_A^2(t)t^{2/(2+\mathrm{\Theta })}$$ (38) with a positive exponent $`\mathrm{\Theta }=2/(n2)`$ indicating subdiffusive behavior. In the limit $`n\mathrm{}`$ we have $`\mathrm{\Theta }0`$, and conventional diffusion is recovered. The reason is of course that for large $`n`$ the depleting reactions become very unlikely, as $`n`$ particles are required to meet at the same lattice site. Thus, low-order $`B`$ species reactions are much more effective in slowing down the $`A`$ diffusion. The time scale for the crossover to the pure algebraic decay of the $`B`$ particle density and subsequently for the anomalous $`A`$ diffusion is given by $`\tau \rho _B(0)^{1n}/\lambda _n`$. The crossover to the asymptotic slow dynamics is fast for large initial densities and reaction rates. The above analysis should be qualitatively correct for $`n>3`$, as the corresponding critical dimension $`d_c(n)<1`$. For $`n=2`$, i.e., $`B`$ pair annihilation processes in $`d2`$ dimensions, we know that at long times anticorrelations develop : Initially close-by particles disappear quickly, and only widely separated ones survive. This effective “repulsion” should result in a roughly uniform spatial $`B`$ distribution even for a clustered initial configuration. Given that the coupling coefficient $`\stackrel{~}{D}`$ itself does not renormalize, we therefore expect that our decoupling assumption leading to Eq. (32) should represent a fair approximation, provided the correct time dependence of the $`B`$ density is inserted. For $`d<2`$, the asymptotic result is $$\rho _B(t)t^{d/2},$$ (39) see Ref. and also Sec. III following Eq. (28), whence $$\stackrel{}{x}_A^2(t)=2\overline{D}t^{1d/2}$$ (40) with an appropriate effective rate $`\overline{D}\stackrel{~}{D}/(1d/2)`$. In low dimensions, this algebraic subdiffusive behavior with $`\mathrm{\Theta }=2d/(2d)`$ replaces the logarithmic law (36). At the critical dimensions $`d_c(2)=2`$, one finds the typical logarithmic corrections $$\rho _B(t)t^1\mathrm{ln}t,$$ (41) implying $$\stackrel{}{x}_A^2(t)\stackrel{~}{D}(\mathrm{ln}t)^2,$$ (42) which also describes slower kinetics than given by the mean-field result (36). For the case of $`n=3`$ at its critical dimension $`d_c(3)=1`$, $$\rho _B(t)(t^1\mathrm{ln}t)^{1/2},$$ (43) and one would therefore expect the leading time dependence $$\stackrel{}{x}_A^2(t)\stackrel{~}{D}(t\mathrm{ln}t)^{1/2},$$ (44) i.e., essentially a square-root power law with logarithmic corrections. ### B BARW kinetics We now extend the $`B`$ dynamics and include branching processes of the form $`B(m+1)B`$ with rate $`\sigma _m`$, described by Eqs. (4) or (12). The mean-field rate equation (29), with $`n2`$, is then replaced by $$_t\rho _B(t)=n\lambda _n\rho _B(t)^n+m\sigma _m\rho _B(t),$$ (45) which has two stationary solutions $`\rho _B=0`$ (inactive phase) and $$\rho _B^{\mathrm{}}=\left(\frac{m\sigma _m}{n\lambda _n}\right)^{1/(n1)}$$ (46) (active phase). For any $`\sigma _m>0`$, the latter turns out to be stable, i.e., BARW are always in the active phase in the mean-field approximation. The explicit solution of Eq. (46) furthermore shows that the asymptotic density $`\rho _B^{\mathrm{}}`$ is exponentially approached, $$\rho _B(t)=\frac{\rho _B^{\mathrm{}}}{\left[1+Ce^{(n1)m\sigma _mt}\right]^{1/(n1)}},$$ (47) where $`C=[\rho _B^{\mathrm{}}/\rho _B(0)]^{n1}1`$. Again, Eq. (32) is solved by the Gaussian distribution (33) with mean-square displacement (34). The ensuing integral is readily calculated for some special cases, e.g., for $`n=2`$ $$\stackrel{}{x}_A^2(t)=2\stackrel{~}{D}\rho _B^{\mathrm{}}\left[t+\frac{1}{m\sigma _m}\mathrm{ln}\left(1+Ce^{m\sigma _mt}\right)\right],$$ (48) whereas for $`n=3`$ $$\stackrel{}{x}_A^2(t)=\frac{\stackrel{~}{D}\rho _B^{\mathrm{}}}{m\sigma _m}\mathrm{ln}\left(\frac{\sqrt{1+Ce^{2m\sigma _mt}}+1}{\sqrt{1+Ce^{2m\sigma _mt}}1}\right).$$ (49) In general, asymptotically normal diffusion with effective diffusion coefficient $`\stackrel{~}{D}\rho _B^{\mathrm{}}`$ is recovered in the active state, $$\stackrel{}{x}_A^2(t)=2\stackrel{~}{D}\rho _B^{\mathrm{}}t.$$ (50) The properties of the active phase with an asymptotically homogeneous $`B`$ density are not much influenced by fluctuations, and hence Eq. (50) should aptly describe the ensuing $`A`$ kinetics even beyond mean-field theory. For the possible existence of an inactive phase, and the characterization of the ensuing critical behavior, fluctuation effects are however of utmost importance for $`n=2`$, and it turns out that the cases of odd and even offspring number $`m`$ need to be distinguished. For odd $`m`$, aside from all lower-order branchings, first-order decay processes $`B\mathrm{}`$ are generated, and become sufficiently efficient to shift the critical point to $`\sigma _c>0`$ for $`d2`$ dimensions. The emerging transition at $`\sigma _c>0`$ can be shown to be in the generic directed-percolation (DP) universality class . The inactive phase is then governed by exponential $`B`$ density decay, whereupon the $`A`$ species will become localized according to Eq. (35). At the critical point itself, the $`B`$ species density decays according to a power law $`\rho _B(\stackrel{}{x},t)t^\alpha `$, with $`\alpha =\beta /z\nu _{}`$ given by DP critical exponents in $`d=1`$ and $`d=2`$, respectively. This would suggest $`\stackrel{}{x}_A^2(t)t^{1\alpha }`$; yet the $`B`$ density is far from uniform at the critical point, and is instead characterized by the appearance of fractal density clusters. While we would still expect subdiffusive behavior for the $`A`$ species with $`\mathrm{\Theta }>0`$, this exponent will likely be influenced by the power-law correlations in the critical $`B`$ density. In the case of even $`m`$, for which the $`B`$ particle number parity is locally conserved under the reactions, a non-trivial transition with $`\sigma _c>0`$ is possible only for $`dd_c^{}4/3`$ dimensions. The inactive phase is then given by the pure pair annihilation theory, and consequently Eq. (40) should provide a fair description for the ensuing anomalous $`A`$ diffusion. The critical behavior is governed by a different parity-conserving universality class, with $`\alpha <1/2`$. In this instance, we again expect the above mean-field description to be rather inaccurate. ## V Conclusions We have studied a novel mechanism to induce anomalous diffusion. Whenever an active particle of the $`A`$ species performs a random walk on a lattice, it may visit a certain lattice site only provided this site is already occupied by at least one $`B`$ particle. The random walk is prolongated when the $`B`$ particles react with each other in such a manner that the $`B`$ species density is decreasing. If that decay is exponential (first-order reaction), then after a short time interval (given by the inverse decay rate) the $`B`$ species has disappeared and a further visit of an $`A`$ particle at that site is impossible. As a consequence the $`A`$ species, after some initial mixing, remains localized. When the $`B`$ species undergo reactions of higher order, requiring at least two $`B`$ particles to meet at a lattice site, an algebraic decay ensues that allows hopping processes for the $`A`$ species to occur for a much longer period. However, the random walk process is slowed down considerably as the $`B`$ density diminishes, resulting in a much shorter mean-square displacement of $`A`$ particles as compared with conventional diffusion. The emerging anomalous diffusion is governed by power laws or logarithmic behavior that can (approximately) be related to the asymptotic time behavior of the reacting $`B`$ particle density. In this instance one may view this process as resembling diffusion on a dynamical fractal. Only when at long times the $`B`$ density remains finite and nearly homogeneous, conventional $`A`$ diffusion is recovered. This situation corresponds to diffusion with dynamical disorder, where in the long-time limit the $`B`$ particles, with largely decayed fluctuations, merely resemble a quasi-static inhomogeneous background for the $`A`$ kinetics. The consistent mathematical treatment of diffusion on a static fractal, as well as induced diffusion processes on critical (isotropic or directed) percolation clusters or near BARW critical points remains an open problem that requires more sophisticated analysis beyond the largely mean-field approach presented here. ###### Acknowledgements. We acknowledge fruitful discussions with Hans-Karl Janssen, Beate Schmittmann, and Olaf Stenull.
warning/0001/astro-ph0001519.html
ar5iv
text
# Early spectroscopic observations of Nova (V1494) Aquilae 1999 No.2 Based on the data obtained at the David Dunlap Observatory, University of Toronto ## 1 Introduction Nova Aql 1999 No. 2 (=V1494 Aql) was discovered visually by A. Pereira on Dec. 1.875 UT, 1999 (Pereira et al. 1999) at magnitude m<sub>vis</sub>=6.0. The spectroscopic confirmation was given by subsequent low-resolution observations revealing hydrogen Balmer series with P-Cyg profiles, Mg II 448.1 nm (or He I 447.1 nm) and O I 777.4 nm lines, all with P-Cyg profiles. The very early spectra showed the H$`\alpha `$ absorption component to be blueshifted by about 1020 km s<sup>-1</sup> in respect to the emission peak (Fujii et al. 1999). In contrast to this latter measurement, Moro et al. (1999) reported the absorption component of the H$`\alpha `$ P-Cyg profile at a blueshift of $``$1850 km s<sup>-1</sup>. Additionally, Fe II multiplets at 492, 502 and 517 nm were observed in emission with P-Cyg profile. Further low-resolution spectroscopy was reported by W. Liller (Liller et al. 1999), who took a CCD spectrogram with an objective prism, that showed H$`\alpha `$ in emission at a level of 31 percent above the local continuum. Early photometric observations consist of mainly visual estimates carried out by amateur astronomers published partly in a number of IAU Circulars and partly in observing reports collected by the VSNET group (http://www.kusastro.kyoto-u.ac.jp/vsnet). Thanks to the fast access to the newly obtained magnitudes, the realtime brightness evolution could be followed. The star reached a maximum brightness of 4.0v shortly after the discovery, which was followed by a rapid decline (see below). Few photoelectric data were published which gave a similar picture to that based on visual estimates. V1494 Aql was also detected at 0.85 and 0.45 mm using SCUBA on the JCMT (Pontefract et al. 1999) V1494 Aql is the brightest nova in the northern hemisphere after Nova (V1500) Cygni 1975 – Nova (V1974) Cygni 1992 was fainter in maximum by about 0.2 mag, Warner 1995 –, therefore, it provides a good opportunity to carry out thorough studies of a nova explosion by various instruments. Unfortunately, its celestial position disables continuous follow-up observations in early 2000. Therefore, the spectroscopic evolution, especially the formation of the nebular spectrum is difficult to monitor. The main aim of this paper is to present spectra of the nova taken about a week, 19 and 28 days after maximum and to discuss the meaning of the observed spectra. In addition, we also estimate the absolute magnitude and the distance of the nova given by the characteristic light curve parameters ($`t_2`$ and $`t_3`$). ## 2 The spectroscopic observations The spectroscopic observations were carried out with the Cassegrain-spectrograph attached to the 1.88-m telescope of the David Dunlap Observatory (Richmond Hill, Canada). The spectra were obtained on four nights in December, 1999. The detector was a Thomson 1024 x 1024 CCD chip (with a 6 e<sup>-</sup> readout noise). A low resolution broad band spectrum was taken on the first night of observations, while we took only medium resolution spectra to study the line profiles in detail on the following three nights. The gratings, observed spectral regions and the resolution at the central wavelength are summarized in Table 1. The spectra were reduced with standard IRAF tasks, including bias removal, flat-fielding, aperture extraction (with the task doslit) and wavelength calibration. For the latter, two FeAr spectral lamp exposures were used, which were obtained immediately before and after every stellar exposures. The integration times varied between 10 seconds and 10 minutes, depending on the actual resolving power and wavelength range. The obtained low-resolution spectrum is plotted in Fig. 1, where the most characteristic features are also marked. Fig. 1 clearly illustrates the typical principal plus diffuse-enhanced spectrum. The prominent Balmer series and Fe II multiplets have P-Cyg profiles that were already reported by Moro et al. (1999). Note the Fe II and He I blend at 4500 Å. The presence of strong Fe II multiplets suggest that V1494 Aql belongs to the “Fe II” class of classical novae defined by Williams (1992). Adopting the interpretation of Williams (1992), this means that the line formation happens in a post-outburst wind. The higher-resolution view of the blue region is presented in Fig. 2, where the hydrogen profiles show interesting substructures in absorption. This system of narrow absorption components of H$`\gamma `$ and H$`\delta `$ lines can be seen in Fig. 3a. The symmetric sharp absorption lines are shifted approximately by $`\pm `$400 and $`\pm `$1000 km s<sup>-1</sup> relative to the central emission near the rest wavelength. H$`ϵ`$ is shown for comparison, because of the similar complexity. The evolution of the H$`\alpha `$ profile was followed through 4 weeks after the maximum. The earlier non-Gaussian emission peak changed to a saddle-shaped profile with two maxima at $``$1050 km s<sup>-1</sup> and +1200 km s<sup>-1</sup>. (Fig. 3b). Unfortunately, the possible similar narrow absorption lines are strongly affected by the atmospheric telluric lines. ## 3 The light curve To estimate the light curve parameters, we used visual magnitude estimates by amateur astronomers that were made available by the VSNET group (this data also includes the observations published in the IAU Circulars). Between the discovery on Dec. 1, 1999 and Jan. 5, 2000, almost 420 individual points were collected. As the typical uncertainty of a visual estimate is about $`\pm `$0.3 mag, we have determined a mean light curve by taking 0.1-day bins and calculating the mean value from the individual points. This led to an averaged light curve containing 117 points, which was further noise-filtered by a simple Gaussian smoothing with 0.08 days FWHM. This approach can only be used when the original light curve is dense enough and therefore, the observational scatter can be properly averaged (see Kiss et al. 1999 illustrating this method). The resulting visual light curve is plotted in Fig. 4, where the maximum, $`t_2`$ and $`t_3`$ are also indicated. The derived parameters are: $`t_0`$=2451515.9$`\pm `$0.1 days (1999 Dec 3.4 UT), $`t_2`$=6.6$`\pm `$0.5 days, $`t_3`$=16$`\pm `$0.5 days. Consequently, V1494 Aql is a fast nova and the spectroscopic observations were done 6, 7, 19 and 28 days after the maximum. We note, that the statistical relationship between $`t_2`$ and $`t_3`$ ($`t_32.75t_2^{0.88}`$, Warner 1995) agrees well with the observed values, as the determined $`t_2`$ would imply a $`t_3`$ of 14.5 days. The most striking post-maximum feature is the additional cyclic brightness change following the fast decline. The characteristic period of this secondary variation is about 7 days with an approximate semi-amplitude of 0.3–0.4 mag. It occured 10 days after the maximum light. Fortunately, the nova was at 7.0 mag (maximum plus 3 mag) around the mean level of the cyclic change, therefore, $`t_3`$ is only weakly affected. The occurence of the quasi-periodic brightness oscillations indicated the start of the transition phase, when the fast novae often show variations with quasi-periods ranging from 5 days (GK Per) to 25 days (DK Lac, Warner 1995), generally with a time-scale being close to $`t_2`$. The underlying physical mechanism could be related either to accretion disk phenomena (Leibowitz 1993) or to the contracting photosphere of the white dwarf (Bianchini et al. 1992). Three maximum magnitude versus rate of decline (MMRD) relations were used to calculate visual absolute magnitude (Della Valle & Livio 1995, Capaccioli et al. 1989 and Schmidt 1957). They gave $``$8.84 mag, $``$8.87 mag and $``$8.72 mag for M<sub>V</sub> and M<sub>vis</sub>, respectively. Another method, involving the absolute magnitude 15 days after maximum (Capaccioli et al. 1989), resulted in M<sub>V</sub>(max)=$`8.5\pm 0.15`$ mag (formal error), however, this value has higher uncertainty due to the brightness oscillation in the transition phase. All of these values suggest an approximate visual absolute magnitude of $`8.8\pm 0.2`$ mag being typical among the fast novae. The distance can be determined from the apparent maximum brightness (m<sub>vis</sub>=4.0 mag). Neglecting the interstellar reddening, the distance is d=3.6$`\pm `$0.3 kpc. The absence of strong Na I D absorption may suggest that the reddening is low indeed. The mass of the white dwarf in V1494 Aql was estimated using Eq. (6) of Livio (1992) assuming $`(BV)_0^{max}0`$ (i.e. M$`{}_{\mathrm{B}}{}^{max}M_\mathrm{V}^{max}`$) (Warner 1995). It resulted in M<sub>WD</sub>=1.1 M, however, this value should be only considered as a rough estimate. Finally, the calculated distance and the apparent brightness of the likely progenitor (Pereira et al. 1999) give the pre-outburst visual absolute magnitude of the system. There is a star in the USNO-A2.0 catalogue very close to the position of the nova with red mag 15.6 and blue mag 17.4. These values correspond to an approximate visual apparent magnitude 16.5, that results in M<sub>vis</sub>=3.7 mag for the progenitor, which is normal for quiescent novae (Ringwald et al. 1996). This result excludes the possibility of a giant secondary component, like in GK Per or RS Oph. ## 4 Conclusions The overall appearence of the low-resolution spectrum covering the whole optical range strongly resembles Nova LMC 1988 No.1 (see Fig. 1 in Williams 1992) being a nova with characteristic “Fe II” type spectrum. Also, it is very similar to the spectrum of the fast ($`t_2`$=5 days) nova, Nova LMC 1988 No. 2 (Sekiguchi et al. 1989, Williams et al. 1991). Sekiguchi et al. (1989) found two distinct sharp absorption systems in the Balmer lines resembling our medium-resolution observations of the H$`\gamma `$ and H$`\delta `$ profiles. The presence of such complex absorption systems is a common phenomenon in novae (Sekiguchi et al. 1989), which may usually be traced even to H$`ϵ`$–H8 in the Balmer series. Unfortunately, we have obtained only one spectrum in that spectral region, therefore, we have no information on the time evolution and no firm conclusion can be drawn about the underlying physical reasons. One would speculatively expect such symmetric ($`\pm `$400 and $`\pm `$1000 km s<sup>-1</sup>) and narrow absorption lines in an axisymmetric wind during the common envelope (CE) phase in the outburst, when the CE phase produces a density contrast between the equatorial and polar directions (Livio et al. 1990). This is supported by the observed H$`\alpha `$ profiles, which showed first an emission profile with one main component (simultaneously with the H$`\gamma `$ and H$`\delta `$ observations) and then the split into two main peaks after 19 days. The saddle-shaped H$`\alpha `$ profile suggest the non-spherical symmetry of the nova shell. Most recently, Gill & O’Brien (1999) presented an ensemble of calculated emission-line profiles from model nova shells with various symmetries. The observed shape of the H$`\alpha `$ line corresponds to an equatorial ring seen most probably nearly edge-on. The difference of the blue and red peaks in the H$`\alpha `$ can be associated with possible small-scale clumpiness in the shell, which produces an increase in brightness in the receding half of the shell (see Gill & O’Brien 1999 concerning the observed asymmetries in the emission-line profiles of V705 Cas). Adopting the calculated distance of 3.6 kpc and and expansion velocity of the nova shell of 2000 km s<sup>-1</sup> given by the radial velocity of the absorption component of the P-Cyg profiles, Eq. (5.10) in Warner (1995) gives 0$`\stackrel{}{.}`$11 for the angular radius of the shell after 1 year. This is within range of the HST. Therefore, subsequent spectroscopy and high-resolution imaging may confirm the existence of the hypothetic non-spherical structure. Evidence suggests this to be an equatorial ring. ###### Acknowledgements. This research was supported by Hungarian OTKA Grants #F022249, #T022259 and Szeged Observatory Foundation. Fruitful discussions with J. Vinkó are gratefully acknowledged. The referee (Dr. M. Orio) has greatly improved the paper with her helpful comments and suggestions. The NASA ADS Abstract Service was used to access data and references.
warning/0001/hep-ph0001060.html
ar5iv
text
# 1 Ref. G uses 1P-1S splitting of the charmonium system to calibrate (1/a). Reference A quotes two different results for two sets of parameters. All results are at the scale 𝜇=2 GeV IUHET-418 Lattice simulations of the strange quark mass and Fritzsch texture Biswajoy Brahmachari Physics Department, Indiana University, Bloomington IN-47405, USA Abstract A number of numerical simulations of lattice gauge theory have indicated a low mass of strange quark in 100 MeV range at the scale of $`\mu =2`$ GeV. In the unquenched case, which is improved over the simulation in the quenched approximation by the inclusion of $`u`$ and $`d`$ sea quark effects, one sees a further downward trend. Here the fermion mass spectrum of the Fritzsch texture is recalculated. In a single step supersymmetric GUT with $`M_X10^{16}`$ GeV such values of the strange quark mass can be obtained for low values of $`\mathrm{tan}\beta `$. Experimental numbers $`m_t^{pole}=173\pm 6`$ GeV and $`4.1<m_b(m_b)<4.4`$ GeV are used in this study. Since the scenario is supersymmetric, gaugino loop diagrams contribute to the masses in addition to usual tree level Yukawa contributions. Upper bound of the mixing parameter $`V_{cb}`$ is taken at 0.045 The strange quark mass has been traditionally calculated using the current algebra mass ratio $$\frac{m_s}{m_u+m_d}=12.6\pm 0.5$$ (1) Equation (1) is evaluated using values of ($`m_u+m_d`$) which are reported in calculations of QCD finite energy sum-rules (FESR). At the two-loop level of perturbative QCD calculations which include non-perturbative corrections up to dimension six one has the result $$(m_u+m_d)(1\mathrm{GeV})=15.5\pm 2.0\mathrm{MeV}$$ (2) For $`\alpha _s(m_Z)=0.118`$ results of (1) and (2) together lead to $$m_s(1\mathrm{GeV})=195\pm 28\mathrm{MeV}\mathrm{or}m_s(2\mathrm{GeV})=150\pm 21\mathrm{MeV}.$$ (3) The ratio $$m_s(2\mathrm{GeV})/m_s(1\mathrm{GeV})=0.769\mathrm{for}\alpha _s(m_Z)=0.118$$ (4) can be obtained by solving renormalization group equations. A systematic uncertainty in this result remains in reconstruction of so called ‘spectral function’ from experimental data of resonances. When a different functional form of the resonance is adopted, and three loop order perturbative QCD theory is used one obtains $$(m_u+m_d)(1\mathrm{GeV})=12.0\pm 2.5\mathrm{MeV}$$ (5) With (5) and (1) one gets $$m_s(1\mathrm{G}\mathrm{e}\mathrm{V})=151\pm 32\mathrm{MeV}\mathrm{or}m_s(2\mathrm{G}\mathrm{e}\mathrm{V})=116\pm 24\mathrm{MeV}.$$ (6) Again this translation from the scale of 1 GeV to the scale of 2 GeV is obtained for the case $`\alpha _s=0.118`$. It has been remarked in Ref. that it is indeed difficult to account for vacuum fluctuations, or sea quark effects generated by quarks of small masses in perturbative QCD calculations. Thus, numerical simulations of strange quark mass on a lattice becomes rather attractive, especially if the simulation includes virtual light quark loop effects. Up and down type quarks differ only in the $`U(1)_{em}`$ quantum numbers in an effective theory where the gauge symmetry is $`SU(3)_c\times U(1)_{em}`$. Lattice calculations in current literature have neglected effects of $`U(1)_{em}`$ which distinguishes up quarks from down quarks. Let us note that we are describing the lattice in terms of a theory at the scale of a few GeVs where light quark masses are to be described in terms of observables relevant to their own scales, which are meson masses and decay constants. Thus, lattice simulation determines $`m_s`$, $`\frac{m_u+m_d}{2}`$ and the lattice spacing $`a`$ using three hadronic observables. They can be chosen, for example, $`M_\pi ,M_K^{}`$, and $`f_\pi `$. Due to the structure of equations which are to be fitted, the scale $`a`$ can also be taken as a function of some other observable, for example, it may be chosen as $`a(M_n)`$ or $`a(M_\mathrm{\Delta })`$ etc. The result depend on the choice of the observable that fits the lattice spacing. The best choice would be the one which has minimum experimental uncertainty and the best result would be a clever weighted average of results from various choices. A test of the simulation is obviously to see whether results from various observables are statistically consistent with each other. Next question is how do we describe quark masses when the theory is living on a discritized lattice. Various definitions or formalisms of quark masses on a lattice have been suggested. Ref. uses the definition in terms of hopping parameter $`\kappa `$ of the lattice $$am_{bare}=log(1+(1/2\kappa 1/2\kappa _c)),$$ (7) for a Wilson-like fermion. In the continuum limit we have $`a0`$, and there one gets the hopping parameter $`\kappa =\kappa _c=1/8`$. A smaller hopping parameter makes the lattice more sticky, and fermions remain on lattice points for a longer time. This make them look as if they were more massive. There are various other formalisms of defining the mass of fermions on the lattice such as stagerred fermions or domain wall fermions. Most calculations, however, use the Wilson action for various definitions of the fermion mass. Next, to compare the result with experiment, one has to calculate the $`\overline{MS}`$ mass at a scale $`\mu `$ starting from the lattice estimate of the bare mass (7) using, for example, the mass renormalization constant $`Z_m(\mu )`$ relating the lattice regularization scheme to the continuum regularization scheme. The lattice regularization prescription is given in Ref. The $`Z_m`$ constant for various formalisms such as Wilson-like or Staggered are given in table1 of . Final results of the physical quark mass for various definitions of the fermion on a lattice differ $`O(a)`$ among each other and one expects to get the same result of the physical quark mass in the continuum limit when $`a0`$. Beyond the minimal lattice simulation of light quark masses using the heavy quark effective theory, the next step would be to incorporate sea quark effects. From the conservation of energy it can be understood that it is easiest to produce lightest of the quarks virtually. Indeed such simulations have been performed. They are termed $`n_f=2`$ unquenched lattice simulations. The detailed processes of numerical simulations are described by respective authors. However, we have summarized the results of recent studies are in table 1 in the order: (A), (B), (C), (D), (E), (F), and (G). On the experimental front the bottom quark mass is in the range $$4.1<m_b(m_b)<4.4\mathrm{GeV}$$ (8) according to the review of particle physics (PDG) tables. Theoretically, one re-expresses the bottom quark mass in terms of parameters of the Minimal Supersymmetric Standard Model(MSSM). The tree level contribution which is related straight to the Yukawa texture, and the one loop contribution due to the dominant gaugino loop can be accounted individually. Then one can write down the relation $`m_b`$ $`=`$ $`m_b^{texture}+m_b^{SUSY}`$ (9) $`=`$ $`h_b{\displaystyle \frac{V_F}{\sqrt{2}}}\mathrm{cos}\beta +m_b{\displaystyle \frac{8}{3}}g_3^2{\displaystyle \frac{\mathrm{tan}\beta }{16\pi ^2}}{\displaystyle \frac{m_{\stackrel{~}{g}}\mu }{m_{eff}^2}}.`$ Here $`m_{\stackrel{~}{g}}`$ is the gluino mass $`\mu `$ is the $`\mu `$ parameter and $`m_{eff}`$ is averaged supersymmetry breaking mass scale. This paper discusses a scenario where the first term of the RHS of (9) comes from diaginalizing a Fritzsch Yukawa texture. The second term can be estimated to be around $`\pm 2`$ GeV. In the supersymmetric case it will be satisfactory if the Fritzsch Yukawa contribution is in the range $$2.1<m_b^{texture}<6.4\mathrm{GeV}$$ (10) Next question is concerning $`\mathrm{tan}\beta `$. Supersymmetry, together with the gauge quantum number structure of fermions demands that at least two Higgs doublets are necessary. Hence the ratio of the VEVs of two Higgs doublets is an unavoidable parameter given the value of the effective four-Fermi coupling $`V_F`$. There are perturbative bounds on $`\mathrm{tan}\beta `$ in the context of grand unified theories (They can be extended to supersymmetric theories without grand unification if MSSM is valid up to a certain high scale, say $`10^{19}`$ GeV). In practice the very low-valued regions of the parameter space for $`\mathrm{tan}\beta `$ are forbidden from perturbative considerations. See for example. Furthermore high values of $`\mathrm{tan}\beta `$ have constraints from charge and color breaking. This is especially true if Yukawa couplings are at the fixed point region. The intermediate regions of $`\mathrm{tan}\beta `$ are definitely allowed. To make a safe case let us choose the range for the purpose of this paper $$\mathrm{tan}\beta =230.$$ (11) Now let us focus on the texture. It has been noted that the quark mixing angle $`V_{us}`$, which is a dimension-less quantity, can be thought of as a ratio of the mass scales of flavor symmetry breaking. These symmetries lead to the mass hierarchy between families. Phenomenologically of course, the ratio of the masses of the first and the second generation satisfies well the relation $$\mathrm{tan}\theta _c=\sqrt{\frac{m_d}{m_s}}.$$ (12) If there are two Higgs doublets instead, (12) remains untouched as the ratio of the VEVs of the doublets cancel in the ratio on the RHS. Thus it cannot feel $`\mathrm{tan}\beta `$. Suppose in a two generation case rotation angles of the up and the down sectors are $`\theta _u`$ and $`\theta _d`$. Then the combined quark mixing matrix $`V=O_uO_{d}^{}{}_{}{}^{}`$ will give $`\theta _c=\theta _u\pm \theta _d`$. This observation plays a role in the Fritzsch mass matrices. Fritzsch mass matrices can be thought of as a set of mass matrices which generalizes (12) to the following form $$\theta _c=\theta _d\pm \theta _u=\mathrm{tan}^1\sqrt{\frac{m_d}{m_s}}\pm \mathrm{tan}^1\sqrt{\frac{m_u}{m_c}}.$$ (13) Where $`m_i`$ are eigenvalues of Fritzsch mass matrices. In the three generation case Fritzsch textures for up and down sectors are given by $$M_U=\left(\begin{array}{ccc}0& ae^{ir}& 0\\ ae^{ir^{}}& 0& be^{ih}\\ 0& be^{ih^{}}& ce^{iq}\end{array}\right)M_D=\left(\begin{array}{ccc}0& Ae^{iR}& 0\\ Ae^{iR^{}}& 0& Be^{iH}\\ 0& Be^{iH^{}}& Ce^{iQ}\end{array}\right).$$ (14) The phases $`r,R,r^{},R^{},h,H,h^{},H^{},q,Q`$ can be absorbed in the redefinition of quark fields and individual mass matrices can be made real. However the weak charge changing current consists of (couples to) gauge eigenstates. Furthermore the structure of the weak charge changing current is $`\psi _L^U\gamma ^\mu \psi _L^D`$. Consequently the weak mixing matrix must contain some combination of phases. It can be shown that residual phases of the weak mixing matrix are $`\sigma `$ $`=`$ $`(rR)(hH)(h^{}H^{})+(qQ)`$ $`\tau `$ $`=`$ $`(rR)(h^{}H^{}).`$ (15) when the weak mixing matrix is expressed as $$O_U\left(\begin{array}{ccc}1& 0& 0\\ 0& e^{i\sigma }& 0\\ 0& 0& e^{i\tau }\end{array}\right)O_{D}^{}{}_{}{}^{1}.$$ (16) In (16) $`O_U`$ and $`O_D`$ diagonalizes $`_U`$ and $`_D`$ which are precisely those in (14) but in the limit when all the phases vanish. Which are very simply $$_U=\left(\begin{array}{ccc}0& a& 0\\ a& 0& b\\ 0& b& c\end{array}\right);_D=\left(\begin{array}{ccc}0& A& 0\\ A& 0& B\\ 0& B& C\end{array}\right).$$ (17) Let us consider the limit $`m_u<<m_c<<m_t`$ and $`m_d<<m_s<<m_b`$. We can approximately re-write (17) as following $$_U=\left(\begin{array}{ccc}0& a& 0\\ a& m_c& 0\\ 0& 0& m_t\end{array}\right);_D=\left(\begin{array}{ccc}0& A& 0\\ A& m_d& 0\\ 0& 0& m_t\end{array}\right).$$ (18) Now we easily see that $`_U`$ can be diagonalized by the rotation $$O_U=\left(\begin{array}{ccc}\mathrm{cos}\theta _1^u& \mathrm{sin}\theta _1^u& 0\\ \mathrm{sin}\theta _1^u& \mathrm{cos}\theta _1^u& 0\\ 0& 0& 1\end{array}\right)\left(\begin{array}{ccc}1& 0& 0\\ 0& \mathrm{cos}\theta _2^u& \mathrm{sin}\theta _2^u\\ 0& \mathrm{sin}\theta _2^u& \mathrm{cos}\theta _2^u\end{array}\right)$$ (19) where $`\mathrm{tan}\theta _1^u=\sqrt{\frac{m_u}{m_c}}`$ and $`\mathrm{tan}\theta _2^u=\sqrt{\frac{m_c}{m_t}}`$. $`_D`$ can be diagonalized similarly. In the limit of strongly hierarchical eigenvalues one may use the approximation $$\mathrm{cos}\theta _i^u1\mathrm{cos}\theta _i^d1.$$ (20) Furthermore let us define the quantities $`\mu _1=\sqrt{{\displaystyle \frac{m_u}{m_c}}}\mu _2=\sqrt{{\displaystyle \frac{m_c}{m_t}}}\nu _1=\sqrt{{\displaystyle \frac{m_d}{m_s}}}\nu _2=\sqrt{{\displaystyle \frac{m_s}{m_b}}}`$ (21) Then the mixing matrix (16) takes the following form $$\left(\begin{array}{ccc}1& \nu _1+\mu _1e^{i\sigma }& \mu _1(\nu _2e^{i\sigma }\mu _2e^{i\tau })\\ \mu _1+\nu _1e^{i\sigma }& \mu _1\nu _1+\mu _2\nu _2e^{i\sigma }+e^{i\sigma }& \nu _2e^{i\sigma }\mu _2e^{i\tau }\\ \nu _1(\mu _2e^{i\sigma }\nu _2e^{i\tau })& \mu _2e^{i\sigma }\nu _2e^{i\tau }& \mu _2\nu _2e^{i\sigma }+e^{i\tau }\end{array}\right).$$ (22) A detailed derivation of these relations are given in Ref. A cancelation among two terms in the expression for $`V_{cb}`$ in (22) is needed. To achieve this one can choose $`\sigma \tau {\displaystyle \frac{\pi }{2}}\mathrm{this}\mathrm{gives}V_{cb}\sqrt{{\displaystyle \frac{m_s}{m_b}}}\sqrt{{\displaystyle \frac{m_s}{m_t}}}.`$ (23) It is easy to check that (23) makes the top quark mass too light to be experimentally true. Thus, it is worth asking the question whether if the Fritzsch texture were valid at the GUT scale instead, in other words, if the flavor symmetries were exact only above the GUT scale, could a miracle of renormalization group evolution of the masses and mixing angles make the Fritzsch relations valid at low energy. Here we study their idea. The renormalization of the full $`3\times 3`$ complex Yukawa matrices and their renormalization up to the GUT scale $`M_X=10^{16}`$ GeV is what follows. Let us set our notations of mixing angles, the non-removable phase and eigenvalues of the Yukawa matrices. We adopt the parameterization $$\left(\begin{array}{ccc}s_1s_2c_3+c_1c_2e^{i\varphi }& c_1s_2c_3s_1c_2e^{i\varphi }& s_2s_3\\ s_1c_2c_3c_1s_2e^{i\varphi }& c_1c_2c_3+s_1s_2e^{i\varphi }& c_2s_3\\ s_1s_3& c_1s_3& c_3\end{array}\right).$$ (24) There is a detailed proof in Ref. that in this parameterization eigenvalues $`y_i`$ of the Yukawa textures, three CKM mixing angles and the CP violating phase $`\varphi `$ satisfy the renormalization group equations $`16\pi ^2{\displaystyle \frac{d}{dt}}\varphi =0,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}\mathrm{tan}\theta _1=y_t^2\mathrm{sin}^2\theta _3,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}\mathrm{tan}\theta _2=y_b^2\mathrm{sin}^2\theta _3,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}\mathrm{tan}\theta _3=y_t^2y_b^2,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}y_u=c_i^ug_i^2+3y_t^2+y_b^2\mathrm{cos}^2\theta _2\mathrm{sin}^2\theta _3,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}y_c=c_i^ug_i^2+3y_t^2+y_b^2\mathrm{sin}^2\theta _2\mathrm{sin}^2\theta _3,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}y_t=c_i^ug_i^2+6y_t^2+y_b^2cos^2\theta _3,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}y_d=c_i^dg_i^2+y_t^2\mathrm{sin}^2\theta _1\mathrm{sin}^2\theta _3+3y_b^2+y_\tau ^2,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}y_s=c_i^dg_i^2+y_t^2\mathrm{cos}^2\theta _1\mathrm{sin}^2\theta _3+3y_b^2+y_\tau ^2,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}y_b=c_i^dg_i^2+y_t^2\mathrm{cos}^2\theta _3+6y_b^2+y_\tau ^2,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}y_e=c_i^eg_i^2+3y_b^2+y_\tau ^2,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}y_\mu =c_i^eg_i^2+3y_b^2+y_\tau ^2,`$ $`16\pi ^2{\displaystyle \frac{d}{dt}}\mathrm{ln}y_\tau =c_i^eg_i^2+3y_b^2+4y_\tau ^2.`$ (25) Multipliers $`c_j^i`$ of gauge couplings in (18) are well known. They can be found in . We have solved these one-loop equations numerically using Mathematica NDSolve subroutine. The flow chart follows this line. Taking all experimentally possible values of the eigenvalues but only central values of the angles at low energy, we have evolved the set to the GUT scale using (18). At the GUT scale predictions for $`V_{cb}`$ $`V_{us}`$ and $`V_{ub}`$ are calculated assuming that Fritzsch relations are valid only at the GUT scale and beyond. While translating predictions of CKM entries back to low energy using (18), we have used exact values of the angles not central values. Thereafter we have checked whether each individual value of masses and mixings remain within experimentally allowed ranges. For the strange quark mass values quoted in table1 are used. For all other masses and mixings the experimental values are taken from the review of particle physics. Our results are given in table2 In conclusion, implications of results of $`n_f=2`$ unquenched lattice simulations of the strange quark mass in the context of the Fritzsch texture are studied. Previous calculations in this line exist in the literature. We have two new aspects. Because the combined effect of charge and color breaking and perturbative unitarity of the Yukawa couplings may rule out the large $`\mathrm{tan}\beta `$ scenario we have studied the low $`\mathrm{tan}\beta `$ scenario. Moreover, we have included corrections of supersymmetric origin in the study. Supersymmetric corrections to the bottom quark mass and the low $`\mathrm{tan}\beta `$ scenario goes hand in hand. This is in the sense that in the low $`\mathrm{tan}\beta `$ regime Fritzsch texture demands a large Yukawa contribution to the bottom quark mass. This is partially canceled by the supersymmetric loop corrections. Thus, the original Fritzsch texture is consistent with experimental data if it holds at the GUT scale. The strange quark mass emerges in the range 60-70 MeV at $`\mu =2`$ GeV for the central values of $`\alpha _s=0.118`$ and $`m_t^{pole}=173`$ GeV. This range is consistent with $`n_f=2`$ sea quark effect improved (unquenched) lattice simulations of the strange quark mass. We thank S. Gottlieb for discussions and M. S. Berger for lending his Fortran routine to calculate the $`\eta _i`$ scaling factors of Ref.. This research was supported by U.S Department of Energy under the grant number DE-FG02-91ER40661.
warning/0001/hep-ex0001004.html
ar5iv
text
# Phys. Lett. B 403 (1997) 185UMS/HEP/96-001FERMILAB-Pub-96-206-E Observation of 𝐷-𝜋 Production Correlations in 500 GeV 𝜋⁻-𝑁 Interactions ## Abstract We study the charge correlations between charm mesons produced in 500 GeV $`\pi ^{}N`$ interactions and the charged pions produced closest to them in phase space. With 110,000 fully reconstructed D mesons from experiment E791 at Fermilab, the correlations are studied as functions of the $`D\pi D`$ mass difference and of Feynman $`x`$. We observe significant correlations which appear to originate from a combination of sources including fragmentation dynamics, resonant decays, and charge of the beam. While the production of heavy quarks can be calculated in perturbative Quantum Chromodynamics (QCD), the evolution of these quarks into hadrons remains one of the most challenging aspects of nonperturbative QCD. Correlations between charm mesons and the charged pions produced closest to them in phase space provide information on how quarks evolve into hadrons. Fragmentation dynamics, resonances, and beam effects can each produce such correlations. The relative importance of these mechanisms must be determined from data. During fragmentation, correlations could be produced because $`\overline{q}q`$ pairs from the vacuum are neutral. For example, if a $`c`$ quark combines with a $`\overline{d}`$ from such a pair to form a $`D^+`$, the remaining $`d`$ is close by in phase space and is likely to become part of the closest pion, which we call the “associated pion”. Thus, $`D^+\pi ^{}`$ ($`D^{}\pi ^+`$) would be favored and $`D^+\pi ^+`$ ($`D^{}\pi ^{}`$) disfavored. Similarly, $`D^0\pi ^+`$ ($`\overline{D^0}\pi ^{}`$) would be favored and $`D^0\pi ^{}`$ ($`\overline{D^0}\pi ^+`$) disfavored. Resonances produce the same favored associations. $`D^+`$ decay associates a $`\pi ^+`$ with a $`D^0`$ while $`D^{}`$ decay associates a $`\pi ^{}`$ with a $`\overline{D^0}`$. Qualitatively, $`D^{}`$ decays produce the same correlations. The charge of the beam particle can also lead to charge correlations. Using a $`\pi ^{}`$ beam can lead to the association of both charm mesons and anticharm mesons with negative pions, especially in the forward (beam) direction. Two distinct but related mechanisms can lead to this result. If the charm quark (antiquark) produced in a hard interaction coalesces with the antiquark (quark) from the beam particle to form the charm (anticharm) meson, the remaining quark (antiquark) from the beam can become part of a negative pion, but not part of a positive pion. If neither the quark nor the antiquark from the beam pion is used in making the charm meson, both are available to form negative pions but not positive pions. By comparing the charge correlations of different species of charm mesons and antimesons with associated pions, and by studying them as functions of Feynman $`x`$ ($`x_F`$) of the charm meson, one can hope to disentangle some of these processes. Evidence of such correlations between $`B`$ mesons and associated light mesons, ascribed to resonances, has been observed in $`Z^0`$ decays at LEP by the OPAL collaboration. In this letter, we report the first observation of fragmentation and beam-related production correlations for charm mesons. We use $`D^0K^{}\pi ^+`$, $`D^+K^{}\pi ^+\pi ^+`$, and $`D^+D^0\pi ^+`$ signals (and their charge conjugate decays) from experiment E791 at Fermilab for this study. The data were recorded using a 500 GeV/c $`\pi ^{}`$ beam interacting in five thin target foils (one platinum, four diamond) separated by gaps of about 1.4 cm. The detector, described elsewhere in more detail, is a large-acceptance, forward, two-magnet spectrometer. Its key components for this study include eight planes of multiwire proportional chambers, six planes of silicon microstrip detectors (SMD) before the target for beam tracking, a 17-plane SMD system and 35 drift chamber planes downstream of the target for track and vertex reconstruction, and two multicell threshold Čerenkov counters for charged particle identification. During event reconstruction, all events with evidence of well-separated production and decay vertices were retained as charm decay candidates. For this study, we require the candidate charm decay vertex to be located well outside the target foils and to be at least 8$`\sigma _\mathrm{\Delta }`$ downstream of the primary vertex (where $`\sigma _\mathrm{\Delta }`$ is the error in the measured longitudinal separation between the vertices $`350\mu `$m). The momentum vector of the candidate D must point back to the primary vertex with impact parameter less than 80$`\mu `$m. The momentum of the D transverse to the line joining the primary and secondary vertices must be less than 0.35 GeV/$`c`$. Each decay track must pass closer to the secondary vertex than to the primary vertex. Finally, the track assigned to be the kaon in the charm decay must have a signature in the Čerenkov counters consistent with the kaon hypothesis. The $`D^\pm `$ candidates are found from the $`D^0/\overline{D^0}`$ samples by adding $`\pi ^\pm `$ tracks and requiring that the mass difference $`\mathrm{\Delta }m`$ = $`M(D\pi )M(D)`$ be consistent with the $`D^{}D\pi `$ hypothesis. The final signal sizes are obtained by fitting the invariant mass spectra as Gaussian signals and linear backgrounds. For $`D^0`$, $`\overline{D^0}`$, $`D^+`$, $`D^{}`$, $`D^+`$, and $`D^{}`$, the fits yield $`22587\pm 210`$, $`24237\pm 216`$, $`24569\pm 204`$, $`29649\pm 238`$, $`4997\pm 84`$ and $`6048\pm 93`$ events, respectively. The r.m.s. mass resolutions, $`\sigma _D`$, used later in defining signal and background bands, are $`13\mathrm{MeV}/c^2`$, $`13\mathrm{MeV}/c^2`$, and $`14\mathrm{MeV}/c^2`$ for $`D^0`$, $`D^+`$, and $`D^+`$, respectively. For each D found in an event, all tracks originating from the primary vertex and producing a pion signature in the Čerenkov counters are combined with the D. Among these combinations, the pion that forms the smallest invariant mass difference ($`\mathrm{\Delta }m_{min}`$) with the D decay products is selected as the associated pion. We define the correlation parameter $`\alpha `$ as $`(1)`$ $$\alpha (D)\frac{N_i(D\pi ^q)N_i(D\pi ^q)}{N_i(D\pi ^q)+N_i(D\pi ^q)},$$ where $`q`$ = +, $``$, $``$, +, $``$, + for $`D=D^0`$, $`\overline{D^0}`$ , $`D^+`$, $`D^{}`$, $`D^+`$, and $`D^{}`$, respectively, and $`N_i(D\pi ^q)`$ denotes the number of charm mesons for which the selected pion has the charge $`q`$. In the absence of correlations $`\alpha `$ is zero, and in maximally correlated cases it is unity. We first study the $`D\pi `$ correlations as functions of $`\mathrm{\Delta }m_{min}`$ for $`\mathrm{\Delta }m_{min}<0.74`$ GeV/$`c^2`$. The number of $`D\pi `$ signal combinations in each $`\mathrm{\Delta }m_{min}`$ bin is determined by subtracting from the $`\mathrm{\Delta }m_{min}`$ distribution for $`D`$ candidates (mass within $`\pm 2.5\sigma _D`$ of the nominal $`D`$ mass) the appropriately normalized $`\mathrm{\Delta }m_{min}`$ distribution for background events (mass between 3.0 $`\sigma _D`$ and 5.5 $`\sigma _D`$ from the nominal $`D`$ mass). The correlation parameters for background-subtracted signals (before additional corrections) and background regions are listed in Table 1. The signal correlations differ significantly from the background correlations. We note that replacing the $`D`$ candidate in an event with a $`D`$ of the same species from another event, while keeping the rest of the event the same, produces correlations consistent with those of the background. We use a Monte Carlo simulation of the experiment and the LUND event generator (PYTHIA 5.7/JETSET 7.3) to model the effects of our apparatus and reconstruction. This simulation describes the geometry, resolution, noise, and efficiency of all detectors, as well as interactions and decays in the spectrometer. The detected $`D^{}/D`$ production ratio in the Monte Carlo matches our data well. As with real events, the associated pion for each reconstructed D meson is selected. By matching the selected pion’s momentum vector with the momenta of all generated particles, we determine whether the selected pion track is a real track or a ghost (false) track. Selecting a ghost pion (not matched to any generated track) or a real pion not matched to the true associated pion can cause smearing in $`\mathrm{\Delta }m_{min}`$ and dilution of the correlation. Selecting a pion with the same charge as the associated pion but with different momentum smears events in $`\mathrm{\Delta }m_{min}`$. Selecting a pion with the opposite charge smears events in $`\mathrm{\Delta }m_{min}`$ and also dilutes the correlation. To account for the effects of ghost tracks, smearing, dilution, and acceptance on the correlations as functions of $`\mathrm{\Delta }m_{min}`$, we employ a matrix formalism. For the $`D^+`$, the observed number of $`D^+\pi ^{}`$ combinations $`O_j^+`$ in the $`j^{th}`$ bin of $`\mathrm{\Delta }m_{min}`$ can be written as $`(2)`$ $$O_j^+=\underset{i}{}S_{ji}^1A_i^+N_i^++\underset{i}{}S_{ji}^2A_i^{+\pm }N_i^{+\pm }+G_j^+O_j^+$$ where $`N_i^+`$ denotes the true number of $`D^+\pi ^{}`$ events in the $`i`$th bin of $`\mathrm{\Delta }m_{min}`$, $`A_i^+`$ the acceptance probability, and $`G_j^+`$ the ghost track rate for $`D^+\pi ^{}`$ combinations. The matrix $`S^1`$ describes smearing in the absence of dilution while the matrix $`S^2`$ describes smearing and dilution when the wrong sign pion is selected. The smearing matrices $`S^1`$ and $`S^2`$, the acceptance coefficients $`A^+`$ and $`A^{++}`$, and the ghost track rates $`G^+`$ and $`G^{++}`$ are determined from the Monte Carlo. The coupled matrix equations in (2) are solved to obtain the true distributions $`N_i^+`$ and $`N_i^{++}`$. Corrected $`\mathrm{\Delta }m_{min}`$ distributions are shown in Figure 1. The corrected correlation parameters for $`D`$, $`\alpha (D)`$, for $`\overline{D}`$, $`\alpha (\overline{D})`$, and for the $`D`$ and $`\overline{D}`$ combined, $`\alpha (D,\overline{D})`$ are presented in column 4 of Table 1. The statistical and systematic errors assigned to the final measurements, shown first and second respectively, are also given in Table 1. These errors are propagated through the matrix formalism. The systematic errors account for uncertainties in the Monte Carlo simulation of the detector (their effects on dilution, smearing, ghost tracks, and acceptance), analysis cuts, background subtraction, kaon misidentification, and binning (in decreasing order of importance as listed). For each data point, the systematic uncertainties due to these sources are added in quadrature. The systematic uncertainties due to statistical fluctuations in the Monte Carlo are negligible. To verify the results produced by the matrix formalism, we also estimate the correlations using simple dilution factors (summed over all bins of $`\mathrm{\Delta }m_{min}`$). For $`D^+`$, the true number of combinations, $`N_t^+`$ and $`N_t^{++}`$, can be expressed in terms of the reconstructed combinations $`N_r^+`$ and $`N_r^{++}`$ as $`(3)`$ $$N_r^+=(1d_+)N_t^++d_{+\pm }N_t^{+\pm },$$ where the dilution factor $`d_+`$ denotes the probability that a true $`D^+\pi ^{}`$ combination is reconstructed as a $`D^+\pi ^\pm `$. The results from this technique are consistent with those reported in Table 1. All studies and corrections have been done within the framework of the LUND PYTHIA/JETSET model. The dilution factors $`d_{ab}`$ in Eq. (3) are typically of order $`0.20.3`$. In our Monte Carlo, $`d_+d_{++}`$ but $`d_+`$ is less than $`d_{}`$. The difference between $`d_{}`$ and $`d_+`$ is almost independent of $`x_F`$ with a typical value near 0.06. Varying some of the JETSET fragmentation parameters to reproduce our inclusive $`D^+/D^{}`$ production asymmetries as a function of $`x_F`$, as described in ref., leads to results consistent with those in Table 1. A fundamentally different model of hadron production might change the differences between the $`d`$’s discussed above by a few times 0.01, which would in turn change the measured correlation parameters. For example, reducing ($`d_{}d_+`$) from 0.06 to 0.05 would increase $`\alpha (D^{})`$ by 0.02 – 0.03. In Figs. 1(a) and (b) we present the numbers of $`D^0\pi ^\pm `$ and $`\overline{D^0}\pi ^{}`$ combinations as functions of $`\mathrm{\Delta }m_{min}`$. In both of these plots the combinations differ mainly in the $`D^\pm `$ resonance region ,the first 75$`\mathrm{MeV}/\mathrm{c}^2`$ bin. Using a $`\pm 2.5\sigma `$ cut on the $`D^+D^0`$ and $`D^{}\overline{D^0}`$ mass difference, we separate the final $`D^0\pi ^+`$ and $`\overline{D^0}\pi ^{}`$ samples into resonance ($`res`$) and continuum ($`cont`$) contributions to obtain $`\alpha (D_{res}^0)=0.98\pm 0.04`$ and $`\alpha (\overline{D^0}_{res})=0.98\pm 0.02`$. For pure resonance, $`\alpha `$ would be near 1. The measured values serve as a check of our method. The continuum measurements are $`\alpha (D_{cont}^0)=0.07\pm 0.03`$ and $`\alpha (\overline{D^0}_{cont})=0.17\pm 0.03`$. In Figs. 1(c) and (d) we present the $`D^+\pi ^{}`$ and $`D^{}\pi ^\pm `$ combinations. In both these plots the combinations differ over a broad range in $`\mathrm{\Delta }m_{min}`$. In Figs. 1(e) and (f) we present the $`D^+\pi ^{}`$ and $`D^{}\pi ^\pm `$ combinations. A pattern similar to that for $`D^\pm `$ is manifest. The plots for charm mesons and anticharm mesons clearly differ. These differences also appear in column (4) of Table 1, and indicate the presence of significant beam-related effects. To investigate beam-related effects in more detail, we study the $`x_F`$ dependence of the $`D^+`$ and $`D^+`$ correlations. We do not show the correlations for $`D^0`$’s since many $`D^0`$’s are decay products of either $`D^0`$ and $`D^+`$, making interpretation difficult. In Fig. 2, we plot $`\alpha `$ as a function of $`x_F`$, for both particle and antiparticle for $`D^+`$ and $`D^+`$. The distributions are corrected using the simple dilution factor technique. We observe that $`\alpha (D^+)`$ rises slightly with $`x_F`$ but $`\alpha (D^{})`$ falls sharply to negative values for $`x_F>0.2`$. In both cases, the $`D`$’s are more likely to be associated with $`\pi ^{}`$’s at high $`x_F`$ where beam effects seem to be important. There appears to be less dependence of $`\alpha `$ on $`x_F`$ for the $`D^\pm `$. Further beam-related studies with Monte Carlo data suggest the correlation asymmetries cancel under neutral beam conditions and are in fact symmetric. This is effectively accomplished when the combined particle and antiparticle correlations are computed. In Table 1, we show the combined and symmetrized correlation parameters to be $`\alpha (D^0,\overline{D^0})=0.29\pm 0.02\pm 0.03`$, $`\alpha (D^+,D^{})=0.21\pm 0.02\pm 0.03`$, and $`\alpha (D^+,D^{})=0.23\pm 0.04\pm 0.03`$. These results indicate that fragmentation dynamics and resonant decays produce substantial correlations between $`D`$ mesons and their associated pions. All three combined correlation levels are approximately equal, although the correlations for neutral and charged $`D`$ mesons are dominated by resonant and continuum regions of $`\mathrm{\Delta }m_{min}`$, respectively. In summary, we observe significant production correlations between $`D`$ mesons and their associated pions. Some of these correlations are associated with fragmentation dynamics, some with resonances, and some with the charge of the beam. In addition to providing information on how heavy quarks evolve into hadrons, such correlations may provide tools for tagging flavor in CP violation studies in heavy flavor systems. We gratefully acknowledge the staffs of Fermilab and of all the participating institutions. This research was supported by the Brazilian Conselho Nacional de Desenvolvimento Científico e Technológio, the Mexican Consejo Nacional de Ciencia y Tecnologica, the Israeli Academy of Sciences and Humanities, the U.S. Department of Energy, the U.S.-Israeli Binational Science Foundation and the U.S. National Science Foundation. Fermilab is operated by the Universities Research Association, Inc., under contract with the United States Department of Energy.
warning/0001/hep-ph0001327.html
ar5iv
text
# Loop-after-loop contribution to the second-order Lamb shift in hydrogenlike low-𝑍 atoms ## Introduction In the low-$`Z`$ region, calculations of radiative corrections in bound-state QED have historically relied on a (semi-) analytic expansion in powers of the external binding field $`Z\alpha `$. Calculations based on this perturbative approach have made an enormous advance during the last 50 years and achieved an excellent agreement with experiments (see, for example, a recent review ). However, calculations in higher orders in $`Z\alpha `$ become increasingly complex, as the number of terms in each higher order increases rapidly. Beside this, it is difficult to estimate the contribution of unevaluated higher-order terms. These are the reasons why the exact numerical treatment of radiative corrections is highly appreciated even in the low-$`Z`$ region. It allows to test the reliability of methods based on an expansion in $`Z\alpha `$ and can provide even more accurate results than analytic perturbative calculations. Some examples of this are the calculation of the self-energy correction to the hyperfine splitting in muonium performed by Blundell and coworkers , the calculation of the relativistic recoil correction for hydrogen by Shabaev et al. , and the evaluation of the first-order self-energy correction for $`Z=15`$ by Jentschura et al. . The aim of the present work is a numerical evaluation of the loop-after-loop contribution to the second-order Lamb shift of the ground state in hydrogen-like atoms to all orders in $`Z\alpha `$ in the low-$`Z`$ region. Analytic calculations of the $`Z\alpha `$-expansion coefficients for this contribution were previously performed by Eides and coworkers and Pachucki in order $`\alpha ^2(Z\alpha )^5`$ and by Karshenboim in order $`\alpha ^2(Z\alpha )^6\mathrm{ln}^3(Z\alpha )^2`$. The first calculation of the loop-after-loop correction without an expansion in $`Z\alpha `$ was carried out by Mitrushenkov et al. for high-$`Z`$ atoms. Recently, this correction was calculated to all orders in $`Z\alpha `$ for the entire range of nuclear charge numbers by Mallampalli and Sapirstein . A fit to the data from Ref. confirms the analytic result of order $`\alpha ^2(Z\alpha )^5`$ but it is in a significant disagreement with Karshenboim’s result of order $`\alpha ^2(Z\alpha )^6\mathrm{ln}^3(Z\alpha )^2`$. The subsequent calculation by Goidenko et al. , also non-perturbative in $`Z\alpha `$, shows to be compatible with the analytic calculations. In this work, we perform an independent calculation of the loop-after-loop correction and investigate possible reasons for the discrepancy between different calculations. Relativistic units are used in this article ($`\mathrm{}=c=m=1`$). ## I Basic formalism and numerical procedure The expression for the irreducible contribution of Fig 1a (we refer to it as the loop-after-loop correction) reads $$\mathrm{\Delta }E_{\mathrm{lal}}=\underset{\epsilon _n\epsilon _a}{}\frac{a|\mathrm{\Sigma }_R(\epsilon _a)|nn|\mathrm{\Sigma }_R(\epsilon _a)|a}{\epsilon _a\epsilon _n},$$ (1) where $`\mathrm{\Sigma }_R`$ denotes the renormalized self-energy operator, $`|a`$ indicates the initial state and the summation is performed over the spectrum of the Dirac equation. The term with $`\epsilon _n=\epsilon _a`$ corresponds to the reducible contribution and should be calculated together with the remaining diagrams in Fig. 1. The self-energy operator is defined by its matrix elements $`a|\mathrm{\Sigma }_R(\epsilon )|b`$ $`=`$ $`i\alpha {\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\omega {\displaystyle d^3𝐱_1d^3𝐱_2\psi _a^{}(𝐱_1)}`$ (4) $`\times \alpha _\mu G(\epsilon \omega ,𝐱_1,𝐱_2)\alpha _\nu \psi _b(𝐱_2)D^{\mu \nu }(\omega ,𝐱_{12})`$ $`\delta m{\displaystyle d^3𝐱\psi _a^{}(𝐱)\beta \psi _b(𝐱)},`$ where $`\alpha _\mu =(1,𝜶)`$; $`\beta `$, $`𝜶`$ are the Dirac matrices, $`G(\omega )=1/(\omega (1i0))`$ is the Dirac-Coulomb Green function, $`=(𝜶𝐩)+\beta m+V(𝐱)`$ is the Dirac-Coulomb Hamiltonian, $`\delta m`$ is the mass counterterm, and $`D^{\mu \nu }(\omega ,𝐱_{12})`$ is the photon propagator in a general covariant gauge $$D^{\mu \nu }(\omega ,𝐱_{12})=\frac{d𝐤}{(2\pi )^3}e^{i\mathrm{𝐤𝐱}_{12}}\left(\frac{g^{\mu \nu }}{\mathrm{k}^2+i0}+(1\lambda )\frac{\mathrm{k}^\mu \mathrm{k}^\nu }{(\mathrm{k}^2+i0)^2}\right)|_{\mathrm{k}^0=\omega }.$$ (5) To our knowledge, up to now all the practical self-energy calculations without an expansion in $`Z\alpha `$ were carried out in the Feynman gauge ($`\lambda =1`$) which is technically the easiest choice of the gauge. While the usage of the Feynman gauge in calculations of the self-energy matrix elements is natural in the high-$`Z`$ region, for low $`Z`$ it is known to provide a spurious contribution of order $`Z\alpha `$ which should be cancelled numerically to give a residual of order $`(Z\alpha )^4`$. This spurious term is known to vanish in the Fried-Yennie gauge ($`\lambda =3`$), which possesses remarkable infrared properties. Since the present work is aimed to a calculation of the loop-after-loop correction in the low-$`Z`$ region, we use the fact that this contribution is invariant in any covariant gauge and perform our calculations in the Fried-Yennie gauge. A general method which was used here for the calculation of the self-energy matrix elements can be found in Ref. , with some modifications due to a non-diagonal nature of the matrix elements and the different gauge. The self-energy matrix element is considered as a sum of two contributions originating from an expansion of the bound electron propagator in terms of interactions with the external field of the nucleus $$a|\mathrm{\Sigma }_R(\epsilon )|b=a|\mathrm{\Sigma }_R^{(0+1)}(\epsilon )|b+a|\mathrm{\Sigma }^{(2+)}(\epsilon )|b.$$ (6) Here, the first term contains zero and one Coulomb interaction with the nucleus, and the second term contains two and more interactions. They are calculated in momentum and coordinate space, respectively. The expression (1) for the loop-after-loop contribution contains a summation of non-diagonal self-energy matrix elements over the whole spectrum of the Dirac equation. To perform the summation, we use the B-splines method for the Dirac equation developed by Johnson et al. . In this method, the infinite summation in the spectral representation of the Green function with a fixed angular momentum quantum number is replaced by a finite sum over basis-set functions. A straightforward evaluation of the sum in Eq. (1) implies a computation of many self-energy matrix elements with highly-oscillating wave functions and is computationally intensive. To reduce the computational time significantly, we define a self-energy correction to the wave function, as proposed in Ref. $$|\phi _{SE}\mathrm{\Sigma }_R(\epsilon _a)|a.$$ (7) According to Eqs. (6) and (7), we write Eq. (1) as $`\mathrm{\Delta }E_{\mathrm{lal}}`$ $`=`$ $`{\displaystyle \frac{d^3𝐩_1}{(2\pi )^3}\frac{d^3𝐩_2}{(2\pi )^3}\phi _{SE}^{(0+1)}{}_{}{}^{}(𝐩_1)G^{\mathrm{red}}(\epsilon _a,𝐩_1,𝐩_2)\phi _{SE}^{(0+1)}(𝐩_2)}`$ (10) $`+2{\displaystyle \frac{d^3𝐩_1}{(2\pi )^3}d^3𝐱_2\phi _{SE}^{(0+1)}{}_{}{}^{}(𝐩_1)G^{\mathrm{red}}(\epsilon _a,𝐩_1,𝐱_2)\phi _{SE}^{(2+)}(𝐱_2)}`$ $`+{\displaystyle d^3𝐱_1d^3𝐱_2\phi _{SE}^{(2+)}{}_{}{}^{}(𝐱_1)G^{\mathrm{red}}(\epsilon _a,𝐱_1,𝐱_2)\phi _{SE}^{(2+)}(𝐱_2)},`$ where $`G^{\mathrm{red}}(\epsilon _a,𝐱_1,𝐱_2)`$, $`G^{\mathrm{red}}(\epsilon _a,𝐩_1,𝐩_2)`$, and $`G^{\mathrm{red}}(\epsilon _a,𝐩_1,𝐱_2)`$ are the reducible Dirac-Coulomb Green functions in coordinate, momentum, and mixed representations, respectively (by the mixed representation we mean the Fourier transform over one coordinate variable). As the first step of the numerical evaluation of Eq. (10), the effective wave functions $`\phi _{SE}^{(0+1)}(𝐩)`$ and $`\phi _{SE}^{(2+)}(𝐱)`$ are calculated on a grid and stored in an external file. Their computation is not much more intensive than an evaluation of a single self-energy matrix element. The most difficult part of the calculation is the evaluation of $`\phi _{SE}^{(2+)}(𝐱)`$. Working in the Fried-Yennie gauge, we do not encounter severe cancellations between zero-, one-, and many-potential terms, as occur in the case of the Feynman gauge. Still, significant cancellations arise in the computation of the Green function $`G^{(2+)}`$ which contains two and more interactions with the external field. In our implementation it is evaluated by a point-by-point subtraction of the two first terms of the Taylor expansion from the Dirac-Coulomb Green function (see Ref. for details) $$G^{(2+)}(\epsilon ,x_1,x_2)=G(\epsilon ,x_1,x_2)G(\epsilon ,x_1,x_2)|_{Z=0}Z\left(\frac{d}{dZ}G(\epsilon ,x_1,x_2)|_{Z=0}\right).$$ (11) To control the cancellations which arise in the low-$`Z`$ region, we monitor the corresponding Wronskian difference $`\mathrm{\Delta }_\kappa ^{(2+)}(\epsilon )`$ which can be calculated analytically ($`\mathrm{\Delta }_\kappa (\epsilon )`$ is the Wronskian of the solutions of the radial Dirac equation). Another numerical problem is the partial wave expansion. Its convergence is somewhat slower in the case of the Fried-Yennie gauge than in the Feynman gauge. In actual calculations we extended the summation up to sixty partial waves. It was performed before all numerical integrations were carried out. The remainder after the truncation of the sum was estimated taking into account the asymptotic behaviour of the expansion terms. Several checks were made of calculations of $`\phi _{SE}^{(0+1)}(𝐩)`$ and $`\phi _{SE}^{(2+)}(𝐱)`$. In one, we compared the diagonal self-energy matrix elements to the known results for the first-order self-energy contribution . We also calculated the irreducible contribution to the self-energy correction to the hyperfine splitting in H-like atoms and found a good agreement with Ref. . In the next step, we perform the radial integrations in Eq. (10). The Dirac-Coulomb Green function in coordinate space is evaluated using a finite basis set constructed from B-splines, after a transformation to a piecewise-polynomial representation as described in the Appendix. The momentum and the mixed representations of the Green function are obtained by the direct numerical Fourier transformation of the polynomial basis. After that, two-dimensional radial integrals in Eq. (10) are expressed as a linear combination of one-dimensional integrals and can be easily evaluated up to a desirable precision. In actual calculations we used a basis set consisting of 70 positive and 70 negative energy states. The stability of the final results with respect to the size of the cavity and the number of energy states was checked. ## II Numerical results and discussion In Table I and Fig. 2 we present the results of our calculation of the loop-after-loop contribution to the second-order Lamb shift of the ground state of hydrogenlike atoms, expressed in the standard form $$\mathrm{\Delta }E_{\mathrm{lal}}=\left(\frac{\alpha }{\pi }\right)^2\frac{(Z\alpha )^5}{n^3}G_{\mathrm{lal}}(Z\alpha ).$$ (12) The results of two previous non-perturbative calculations of this correction are presented in Table I and Fig. 2 as well. A comparison exhibits a good agreement of the present calculation with the results of Mallampalli and Sapirstein and a strong deviation from the results of Goidenko et al. . Let us consider possible reasons for this discrepancy. The method used in Ref. is based on the multiple commutator approach combined with the partial-wave renormalization (PWR) procedure. In the PWR method, the truncation of the partial-wave expansion fulfils the role of the regularization parameter. This shows that this method is non-covariant. Still, it can be used for the calculation of the diagonal first-order self-energy matrix elements, for which the PWR procedure is known to provide the correct result . In Ref. this renormalization procedure was investigated for the self-energy correction to an additional Coulomb screening potential. It was shown analytically that some spurious terms arise in different parts of the total self-energy correction due to the non-covariant nature of the renormalization procedure. According to Ref. , the spurious terms cancel each other if the perturbation is the Coulomb potential. The cancellation of the spurious contributions in the total self-energy correction holds no longer if the perturbation contains a magnetic photon (see Ref. and a conclusion remark in Ref. ). To consider this topic in more detail, we calculate the self-energy correction in the presence of the perturbing potential $`\alpha /r`$ both in the PWR scheme and using a covariant renormalization. For this choice of a perturbing potential, the total self-energy correction for a state $`|a`$ is $`d/(dZ)a|\mathrm{\Sigma }_R(\epsilon _a)|a`$. The results of calculations are listed in Table II. Our calculation confirms the conclusions from Ref. about $`a)`$ the presence of spurious terms in different parts of the correction and $`b)`$ their cancellation in the sum for this particular choice of a perturbing potential. Summarizing, we conclude that it is possible that the PWR method applied to the irreducible part of the second-order self-energy correction, can provide a nonzero spurious contribution. In order to compare our results with calculations based on an expansion in $`Z\alpha `$, we approximate our data for the function $`G_{\mathrm{lal}}`$ by a least-squares fit with five parameters $`a_{50}`$, $`a_{63}`$, $`a_{62}`$, $`a_{61}`$, and $`a_{60}`$ (the first index of the $`a`$ coefficients indicates the power of $`Z\alpha `$, the second corresponds to the power of $`\mathrm{ln}(Z\alpha )^2`$). A fit to our numerical results in Table I yields $$a_{50}=2.33\text{ }a_{63}=1.1.$$ (13) This is in a good agreement with the fitting coefficients from Ref. ($`a_{50}=2.3`$ or $`2.8`$ for different sets of data, $`a_{63}=0.9`$) but disagrees significantly with Karshenboim’s analytic result $`a_{63}=8/27`$ . In order to investigate this discrepancy in more detail, we note that the $`Z\alpha `$-expansion calculations of the loop-after-loop correction in Refs. were performed in the Fried-Yennie gauge like in the present work and, therefore, it is possible to compare the calculations on intermediate stages. So, we expand the inner electron propagators in diagram Fig. 1a in terms of interactions with the nuclear binding potential and calculate the first six terms of the expansion separately. The corresponding Feynman diagrams are presented in Fig. 3. These diagrams do not contain $`\phi _{SE}^{(2+)}(𝐱)`$ which is the most difficult part of the calculation. Therefore, we were able to calculate them for very low fractional $`Z`$. This is important for a reliable fitting of our data which vary very fast in the vicinity of $`Z=0`$. The remainder behaves more smoothly in the low-$`Z`$ region and its fitting is easier. In the calculation of the diagrams shown in Fig. 3, we use closed analytical expressions for the Dirac Green function with zero and one Coulomb interaction. In this way we eliminate the numerical uncertainty due to the finite basis set representation of the Green function. The numerical results for each diagram in Fig. 3 were approximated by least-squares fits with eight or seven parameters $`a_{50}`$, $`a_{6i}`$ ($`i=3,\mathrm{},0`$), $`a_{7i}`$ ($`i=3,2,1`$) (in the last case $`a_{71}`$ was omitted). In order to reduce the statistical uncertainty of the fitting procedure, a large number of points (twenty or more) was used. The stability of the fitting coefficients was checked with respect to the number of points, minimal and maximal nuclear charge numbers, and different fits. The numerical results and the fitting coefficients for diagrams in Fig. 3 are listed in Table III. We found a good agreement with results from Refs. for the coefficient $`a_{50}`$ and with Ref. for the coefficient $`a_{63}`$ originating from diagram Fig. 3f. The only discrepancy with the analytical calculations originates from diagram Fig. 3c. While this diagram should not contribute to order $`\alpha ^2(Z\alpha )^6\mathrm{ln}^3(Z\alpha )^2`$ according to Karshenboim, our calculation shows the presence of a cubed logarithm with coefficient $`a_{63}=0.652(30)`$. Summarizing, we conclude that our calculation of the loop-after-loop correction confirms the analytic result of Refs. for the coefficient $`a_{50}`$ ($`a_{50}=2.3`$). A fit to the numerical results yields $$a_{63}=0.958(30)\text{ }a_{62}=3.3(5)$$ (14) for the diagrams shown in Fig. 3, and $$a_{63}=0.05(7)\text{ }a_{62}=1.2(8)$$ (15) for the non-perturbative remainder. We note a remarkably slow convergence of the $`Z\alpha `$-expansion for the loop-after-loop contribution to the second-order Lamb shift. As an illustration, in Fig. 4 we plot the contributions of the first one, two, and three expansion terms together with the non-perturbative results. The expansion coefficients are taken from Eqs. (14) and (15). One can see that even for hydrogen the contribution of the first three expansion terms covers only about 50% of the total result. To obtain a reasonable fit to the numerical data even for very low $`Z`$, it is necessary to take into account at least four first expansion terms. This fact shows the necessity for non-perturbative (in $`Z\alpha `$) calculations of the total second-order Lamb shift in the low-$`Z`$ region. ## Acknowledgments I would like to thank Prof. Shabaev for his guidance and continued interest during the course of this work. Valuable conversations with S. G. Karshenboim and T. Beier are acknowledged. I am grateful for the kind hospitality to Prof. Eichler during my visit in autumn 1999. This work was supported by the Russian Foundation for Basic Research (Grant No. 98-02-18350) and by the program ”Russian Universities. Basic Research” (project No. 3930). ## Piecewise-polynomial representation of the Green function The B-splines method for the Dirac equation provides a finite set of radial wave functions with a fixed angular momentum quantum number which can be written in the form $$\phi _{\kappa ,n}^i(x)=\frac{1}{x}\underset{k}{}c_k^i(\kappa ,n,l)(xx_l)^k,$$ (16) assuming that $`x[x_l,x_{l+1}]`$. Here $`x_l`$ is the radial grid, the index $`i=1,2`$ indicates the upper and the lower component of the radial wave function, $`n`$ numbers the wave functions in the set, and $`\kappa `$ is the angular momentum quantum number. The radial Green function, defined by $$G_\kappa ^{ij}(\epsilon ,x_1,x_2)=\underset{n}{}\frac{\phi _{\kappa ,n}^i(x_1)\phi _{\kappa ,n}^j(x_2)}{\epsilon \epsilon _n},$$ (17) can be written in the piecewise-polynomial representation as follows: $$G_\kappa ^{ij}(\epsilon ,x_1,x_2)=\frac{1}{x_1x_2}\underset{k_1k_2}{}A_{k_1k_2}^{ij}(\epsilon ,\kappa ,l_1,l_2)(xx_{l_1})^{k_1}(xx_{l_2})^{k_2},$$ (18) where $`x_1[x_{l_1},x_{l_1+1}]`$, $`x_2[x_{l_2},x_{l_2+1}]`$. The coefficients $`A_{k_1k_2}^{ij}`$ are $$A_{k_1k_2}^{ij}(\epsilon ,\kappa ,l_1,l_2)=\underset{n}{}\frac{c_{k_1}^i(\kappa ,n,l_1)c_{k_2}^j(\kappa ,n,l_2)}{\epsilon \epsilon _n}.$$ (19) The radial Green function in momentum space can be written in the same way using Fourier transformed basic polynomials $`G_\kappa ^{ij}(\epsilon ,p_1,p_2)`$ $`=`$ $`{\displaystyle \underset{l_1l_2}{}}{\displaystyle \underset{k_1k_2}{}}A_{k_1k_2}^{ij}(\epsilon ,\kappa ,l_1,l_2)\mathrm{\Pi }_{l_1}^{ik_1}(p_1)\mathrm{\Pi }_{l_2}^{jk_2}(p_2),`$ (20) $`\mathrm{\Pi }_l^{ik}(p)`$ $`=`$ $`4\pi s(L_i){\displaystyle _{x_l}^{x_{l+1}}}𝑑xx(xx_l)^kj_{L_i}(px),`$ (21) where $`L_{1,2}=|\kappa \pm 1/2|1/2`$; $`s(L_1)=1`$, $`s(L_2)=\kappa /|\kappa |`$; and $`j_L(z)`$ denotes the spherical Bessel function.
warning/0001/quant-ph0001074.html
ar5iv
text
# An Epistemological Derivation of Quantum Logic ## 1 Introduction In physics, the theory of statics comprises propositions about two basic observables, position and momentum. These propositions define ranges of values for each observable. We may then add to this a notion of phase space and a law of propagation associated with the physical system. In a classical physical theory the subsets of phase space are correlated with propositions about the ranges of values so that there is an obvious correspondence between the set theoretic operators of union and intersection and the logical connectives ‘and’ and ‘or’ - with set theoretic inclusion corresponding to logical implication. In a quantum mechanical system the propositions concerning ranges of the observables are correlated with the subspaces of a Hilbert space and the logical connectives between propositions correspond to set products, sums and complements with, once again, inclusion corresponding to implication. Obviously therefore the quantum mechanical and classical propositional calculi differ from an algebraic standpoint. This difference is precisely the following. In the classical propositional calculus the logical connectives between propositions obey the distributive law, $$a(bc)=(ab)(ac)$$ while in the propositional calculus of quantum mechanics $``$ and $``$ do not obey this law although they obey a modified version of it - the weaker modular law, $$\text{ if }ac\text{ then }a(bc)=(ab)(ac)$$ In both cases $``$ and $``$ obey the other usual assumptions sufficient to make $`\mathrm{`}\mathrm{`}ab\text{ iff }ab=a\text{ (or equivalently }ab=b\text{) ”}`$ a partial order. If we assume in addition to the modular law that every proposition can be written as the union of basic elements – atomicity – and any pair of these basic elements have a common complement – perspectivity – then we get the characterisation due to Birkhoff and Von Neumann of the lattice of quantum mechanical propositions as an infinite, modular, atomic, perspective lattice. Since all the nonclassical results of quantum mechanics arise from this distinction in algebraic structure many attempts have been made to explain the need for this modular law. Here it is derived from what we take to be the basic requirements demanded of any physical theory — that the theory contains an operator representing the analysis of quantities into distinct named parts — and that we can treat this operator in boolean languages i.e. we can identify substructures produced by the analysis operator with sublattices of boolean lattices subject to certain consistency requirements. To produce a model points from some model of set theory, $`Set`$, must be assigned to the variables of the theory. We have not specified which points in particular must be chosen from $`Set`$ and certain indistinguishable possibilities arise for models generated in this way. Choosing a single set of possibilities gives the usual boolean lattice as we would expect but we show that the quantum mechanical lattice is the lattice generated by taking all consistent possibilities. Quantum mechanics is therefore in a sense the fullest description of the physical world if the world is restricted to what can be modelled by structures of the above kind. ## 2 Definitions We have stated that the propositions of classical physics are points in boolean lattices. How is such a lattice defined ? A structure will be called a lattice iff to any pair of its elements $`x`$ and $`y`$ there correspond elements $`xy`$ , $`xy`$ with the operators $`,`$ (‘join’ and ‘meet’) satisfying $`xx=x,xx=x,`$ (idempotency) (1) $`xy=yx,xy=yx,`$ (commutativity) (2) $`x(yz)=(xy)z,x(yz)=(xy)z,`$ (associativity) (3) $`x(xy)=x(xy)=x,`$ (absorbtion) (4) $`,`$ then also define a partial order given by $`xy\text{ if }xy=x`$ (or, equivalently, $`xy=y)`$. If in addition $`,`$ satisfy $$x(yz)=(xy)(xy)$$ (5) the lattice is said to be distributive. If instead they satisfy the weaker $$xzx(yz)=(xy)(xy)$$ (6) the lattice is said to be modular. The following is an equivalent formulation of the modular law for finite lattices. Let $`B`$ be a lattice and $`h`$ a positive function, the height function, defined on $`B`$ where for all $`a,b,c,`$ in $`B`$ $$h(ab)=h(a)+h(b)+h(ab)$$ (7) Then $`B`$ satisfies the modular law. Clearly a distributive lattice supports a function satisfying (7). If there exist elements $`0,1`$ in the lattice satisfying $$0x1x$$ and if for every element $`x`$ in the lattice there is a $`y`$ with $$xy=0,xy=1$$ the lattice is said to be complemented. A complemented distributive lattice is called a boolean lattice. For $`x,y`$ elements of a lattice, if $`z`$ with $`xz=0,xz=1`$ $`yz=0,yz=1`$ then $`x`$ and $`y`$ are said to have a common complement $`z`$. Such $`x`$ and $`y`$ are said to be perspective. If every pair of points in a lattice are perspective the lattice is said to be perspective. Finally, an atom is a point $`x`$ such that $`y`$ in the lattice $$y<xy=0$$ ## 3 Outline of the Argument If we want to produce an analytic description of the world then, formally speaking, it must at least satisfy the conditions sketched at the end of section 1 above, which we restate here. Our central premise is that anything we think of as a physical theory must incorporate the idea of analysis of individuals into parts and the treatment as far as is consistently possible of these parts in languages containing (boolean) joins and meets and that any attempt to construct a model of such a theory will naturally reflect this. We can think of the physical theory as containing the collection of statements arising from analysing the universe into parts through measurement and the further statements that can be made about these parts in the boolean languages that underly any formal discussion. We first give a lengthier informal description of the properties that should be satisfied by a structure representing a physical theory and later define them precisely. Firstly, the structure should be nonempty – it should contain two points chosen from set theory call them 0 and 1, and, further, in the ordering relation we will define on the structure, 0 and 1 satisfy $`0x1x`$ in the structure. Secondly, a structure representing a physical theory should support an operator representing the analysis or division of an element into parts. Given any element $`a`$ in the structure the structure should also contain an element $`b_a`$ distinct from $`a`$, with $`a=b_aa`$ contained in the structure and given such a $`b_a`$ there should also exist a $`c_a`$ in the structure with $`a=b_ac_a`$ and such that $`ec_ae=c_a`$ and $`b_ae=b_a`$ iff $`e=0`$. We claim that to represent the natural notion of analysis of a whole into parts $``$ should be a binary operator generating a partial order. It is important to note here that in terms of using this property to generate such a structure the elements from $`Set`$ represented by $`b_a`$ and $`c_a,`$ above are not uniquely specified — there are many possible candidates in $`Set`$ for such elements. A chain between $`a_1`$ and $`a_h`$ in the structure is a set of distinct points $`\{a_1,\mathrm{},a_h\}`$ with $`ia_ia_{i+1}=a_i`$. We claim that to represent the idea of analysis $``$ should satisfy a further condition. Let a refinement of a chain be a larger chain containing it. Then refinements exist subject to the following. For a given $`a,b`$ in the structure either any path between $`a`$ and $`b`$ may be refined or there exists a bound $`d(a,b)`$ such that only those paths of length $`<d(a,b)`$ may be refined. We claim that if in addition $``$ is consistent with the existence of a lattice extending it and satisfying the same properties then this operator represents our natural notion of analysis i.e. the operator represents the ‘part of’ relation arising from measurement in a physical theory. Thirdly, these two principles give rise to many distinct elements and statements of the form $`a=bc`$ where $`a`$, $`b`$ and $`c`$ represent points from $`Set`$. Let $`M`$ be a set of statements generated in this way. While $``$ is consistent with the existence of a lattice with the given properties it does not generate it. In some cases it may be possible to embed $`M`$ in a natural language in such a way that the boolean operators of the language generate the lattice. However it may also be that the lattice we require may not be embeddable in such a language. This is the source of the difficulties in the interpretation of quantum mechanics. All such difficulties can be reduced to the problem of trying to construct the above lattice in terms of an analysis based on what is constructible in boolean lattices. The purpose of this paper is to produce such an analysis. As we have just said it may be possible to make statements about $`M`$ in a natural language, that is to say it may be possible to embed $`M`$ in a boolean lattice in such a way that the combined structure still satisfies the above requirements on an analytic operator. A physical theory should include the expansion of the analytic operator by statements which can be built up by a treatment of the operator in a natural language subject to a general demand of consistency. So a physical theory can make statements about the analytic operator in a language, L, at least strong enough to contain the statements that could arise in any natural discussion, i.e. L contains the boolean operators $`_L`$, $`_L`$ and $`{}_{L}{}^{}{}_{}{}^{}`$ or complement. Consider the statements which we may be able to make about $`M`$ in such a language. It may be that there is a structure generated from $`M`$ using $`_L`$, $`_L`$ and $`{}_{L}{}^{}{}_{}{}^{}`$ extending $`M`$ to a boolean lattice $`B_M`$ in such a way that $`B_M`$ agrees with the lattice operators, height function and possibly the $`0,1`$ of $`M`$ where they overlap and – restricting ourselves to extending the structure as a partition operator – such that the combined structure is still capable of extension to a lattice satisfying the properties described above. One obvious requirement that this extension principle be a consistent one is that we allow an extension only when it is consistent with all such extensions of the substructures of $`M`$. For a given $`M`$ there may be a number of ways of forming boolean extensions of the substructures of $`M`$, call them $`M_i`$, in such a way that they are mutually consistent and consistent with the other conditions outlined above. A collection $`\{B_{M_i}\}`$ of such extensions may be maximal i.e. there is no additional substructure $`M_j`$ such that any $`B_{M_j}`$ is consistent with the collection $`\{B_{M_i}\}`$. Call such a collection a cover of $`M`$. Then the existence of any such cover will imply the existence of a boolean extension of $`M`$, $`B_M`$, iff every cover of $`M`$ contains such a $`B_M`$. The third property that should be enjoyed by a physical theory is that if this condition holds the structure may be extended by some $`B_M`$. We are allowing $`M`$ and its subsets to be extended by the boolean operators of the language only if this happens in a consistent way. We will show that these properties are realised in a boolean lattice and any structure given by these properties contains such a lattice. In describing these properties we alluded to the fact that points chosen to name variables in the theory are not uniquely specified. It turns out we can generate a structure realising all of these possible choices simultaneously by first extending the second property to state that in addition to $`b_a,c_a`$ with $`b_ac_a=a`$ there is also a point $`b_a^{}b_a,c_a`$ with $`b_a^{}d_a=a`$. The structure which is generated by extending Principle II in this way is $`P_{n1}`$, Birkhoff and Von Neumann’s characterisation of the quantum mechanical lattice. ## 4 Construction Principles Let $`Const=\{c_1,\mathrm{}\}`$ where $`c_i`$ are distinct points chosen from some model of set theory, $`Set`$. Equivalently we may define the $`c_i`$ to be distinct elements in the language and $`c_i,c_j,c_ic_j`$ is true. In the last section we sketched a set of properties which a structure representing a set of physical propositions should have. We shall see that these properties are sufficient to represent such a set i.e. if we recast them as construction principles the structure they describe is that of the lattice of physical propositions. We describe these principles below. I The structure should be nonempty. Principle I *The structure should contain a substructure consisting of two points chosen from $`Const`$. Call the points $`0`$ and $`1`$. We define $`01=1`$ and in all that follows it is consistent that $`0x=x,x1=1`$ for any $`x`$ we construct.* II From the principle of analysis of a whole into parts we get the following Principle II *Given any element $`a`$, in the structure the structure should also contain an element $`b_a`$, distinct from $`a`$ chosen from $`Const`$, with the statement $`a=b_aa`$ contained in the structure and given such an $`a`$ and $`b_a`$ the structure contains a $`c_a`$ from $`Const`$ with $`a=b_ac_a`$ and such that $`ec_ae=c_a`$ and $`b_ae=b_a`$ iff $`e=0`$.* Whether $`b_a,c_a`$ are points already chosen in the structure is not defined. It is important to note here that in extending the structure according to Principle II we have not specified which elements of $`Const`$ are represented by $`b_a`$ and $`c_a`$. All that is demanded of them is that they should be distinct and different from $`a`$. Principle II states the existence of a condition to be satisfied by points from $`Const`$ without actually specifying those points. We restate here that a chain between $`a_1`$ and $`a_h`$ in the theory is a set of distinct points $`\{a_1,\mathrm{},a_h\}`$ with $`a_ia_{i+1}=a_ii`$. Let a refinement of a chain be a larger chain containing it. Then Principle II extends the structure subject to refinements satisfying the following condition; for a given $`a,b`$ in the structure either any path between $`a`$ and $`b`$ may be refined or there exists a bound $`d(a,b)`$ such that any path of length $`<d(a,b)`$ may be refined. Further $``$ should satisfy the requirements on a partial order and be consistent with the existence of a lattice extending it and satisfying the requirements outlined above, i.e. .15in (i) Principle II holds in the lattice subject to Principle IV. .15in (ii) Refinements exist in the lattice subject to the condition that for any $`a`$ and $`b`$ in the lattice there exists a bound $`d(a,b)`$ such that just those paths of length $`d(a,b)`$ may be refined. III Any natural language in which we might treat the analytic operator in Principle II contains the boolean operators. We extend the notion of a physical proposition to include statements we can make about this analytic principle in a natural language. Arising from this assumption the final principle is that the structure should include statements which are given by the possibility of extending substructures using the boolean operators of such languages. Let $`M`$ be such a substructure and let $`_M`$ and $`_M`$ as defined in $`M`$ be consistent with the boolean property (5). Our natural notion of being able to treat the propositions of $`M`$ in a rational language amounts to saying that they may be embedded in a structure equipped with boolean operators, $`_L`$, $`_L`$ and $`{}_{L}{}^{}{}_{}{}^{}`$, which are consistent with and extend the $``$ and $``$ of the original structure in such a way that $``$ and $``$ are consistent with the requirements described in the definition of Principle II. Thus the third principle generating the structure is defined as follows; A substructure $`M`$ comprises a set of statements of the form $`ab=c,a^{}b^{}=c^{}\mathrm{}`$ and statements regarding the function $`d`$, defined above, applied to points in $`M,d(a,b)=d_{ab}\mathrm{}`$. Let $`\{M_i,iI\}`$ be the set of substructures of $`M`$. For a given $`M_i`$ it may be possible to define a structure $`B_{M_i}`$, a boolean lattice, on points chosen from $`Const`$, containing $`M_i`$ and consistent with the structure as defined so far (where by consistency we mean that if we define a function $`h`$ on the structure by $`d(0,a)=h(a)`$, and $`h_B`$ is a height function on the distributive $`B_{M_i}`$, then $`_B`$, $`_B`$, $`h_B`$, $`0_B`$ and $`1_B`$ agree with $`,`$, $`h`$, 0 and 1, where defined in the structure so far, and the structure, extended by $`B_{M_i}`$, can still be extended to a lattice satisfying the requirements given above ). Next we define a cover of $`M`$. Given a collection of substructures of $`M`$, $`\{M_i:iI\}`$, suppose that for each $`M_i`$ in the collection there exists such a boolean lattice and that these lattices are consistent in the sense described above with each other and the rest of the structure so far defined. If $`\{B_{M_i}:iI\}`$ is not contained in a larger collection of boolean lattices with this property $`\{B_{M_i}:iI^{},I^{}I\}`$ we call $`\{B_{M_i}:iI\}`$ a cover of $`M`$. Such a cover represents a fullest possible mutually consistent treatment of the parts of $`M`$ in boolean languages in the manner described. If the existence of any such way of consistently extending parts of $`M`$ to such lattices necessarily implies the existence of some $`B_M`$, i.e. if every cover of $`M`$ contains a $`B_M`$, then we demand that the structure should be extended by some such $`B_M`$. In other words if granting that we can treat as many parts as possible of $`M`$ in a rational language implies that we have a boolean lattice containing $`M`$ then we may add some such lattice to the structure. Recasting this as a construction principle we can say that the structure should contain that substructure common to all boolean lattices containing $`M`$ and satisfying the consistency requirements described above. Principle III *For $`M`$ a given substructure, if all maximal coverings, $`\{B_{M_i}:iI,M_iM\}`$, contain a $`B_M`$, then the structure should contain a substructure common to all such $`B_M`$.* We claim that this completely describes our natural notion of what can be said about the products of analysis in every boolean language. In addition to these three principles we introduce an ad hoc assumption limiting the depth of the structure . In their characterisation of the lattice of quantum mechanical propositions Birkhoff and Von Neumann, for the sake of simplifying the proof, restrict their attention to lattices of depth bounded by some $`n\text{N}`$ \- that is to say the length of every chain in the lattice is bounded by $`n`$. We do not need a restriction on the length of chains in the structures to prove the general result of this paper characterising the propositional calculi of the theories of classical and quantum mechanics. However, in the interest of simplifying our proof we too will assume this ad hoc bound on the length of chains in the structure and under this assumption derive from our generating principles Birkhoff and Von Neumann’s restricted models of the propositional calculi. Again the structures we get in the absence of this ad hoc assumption are equivalent to the infinite models of Birkhoff and Von Neumann but the proof is more elaborate. We introduce the following Ad Hoc Principle IV*For some $`n`$ N Principles II and III hold subject to the requirement that for any $`a`$ in the structure $`d(0,a)n`$* Define $`d(0,a)=h(a)`$, called the height of $`a`$. ## 5 These four principles generate and are realised in a boolean lattice. We now show that these four principles are realised in a boolean lattice of depth $`n`$ and any structure generated by these four principles contains such a boolean lattice. Which is what we would expect to get from treating the products of an analytic principle like II in a language with an underlying boolean operator. Let $`B`$ be a model of a boolean lattice of depth $`n`$ i.e. $`h(1)=n`$. We first show that $`B`$ realises any structure constructed by I-IV, and hence these principles are consistent. Principle I stating the existence of a 0 and 1 in $`B`$ with $`0<x<1xB`$ is obviously satisfied in $`B`$. Let $`C`$ be a structure generated by Principles I-IV and let $`C`$ be realised in $`B`$ i.e. there exists a homomorphism $`f`$ mapping $`\{C\}`$ into $`B`$ and $`a,b,cCab=cf(a)f(b)=f(c),ab=cf(a)f(b)=f(c),h_C(a)=h_B(f(a))`$. Then we will show that the extension of $`C`$ by an application of Principle II is realised in $`B`$. Let $`a`$ be a point in $`C`$ realised in $`B`$. Then Principle II states that there should exist $`b_a`$ and $`c_a`$ with $`b_ac_a=a`$ and such that $`ec_ae=c_a`$ and $`b_ae=b_a`$ iff $`e=0`$ subject to this being consistent with the ad hoc Principle IV, i.e. subject to $`h(a)2`$. But if $`h(a)2`$ in $`B`$ then obviously $`B`$ contains such a $`b_a`$ and $`c_a`$ with $`b_ac_a=0`$. Hence Principle II is satisfied in $`B`$. Next we show that an extension of $`C`$ by an application of Principle III is realised in $`B`$. Let $`M`$ be a substructure of $`C`$. $`M`$ is realised in $`B`$. Then Principle III asserts the existence of a substructure containing $`M`$ common to all $`B_M`$. But $`B`$ itself obviously realises such a substructure. Since the extensions of $`C`$ above were subject to the restrictions of Principle IV we are done. We now show that Principles I-IV generate a model of $`B`$, the boolean lattice of depth $`n`$, and hence that any model generated by I-IV contains a submodel equivalent to $`B`$. Let 1 be given by Principle I. By repeated application of Principle II we can construct a tree $`T`$ of depth $`n`$ with $`2^n`$ points of height 1 (or atoms). Call them $`p_1,\mathrm{},p_{2^n}`$. A set of atoms $`q_1,\mathrm{},q_n,`$ are said to be independent if for any $`q_i`$ and any other $`q_{j_1},\mathrm{},q_{j_k}`$ in the set $`q_iq_{j_1}\mathrm{},q_{j_k}`$. $`\{p_1,\mathrm{},p_{2^n}\}`$ contains $`n`$ points $`q_1,\mathrm{},q_n`$ generating a boolean lattice of height $`n`$. We first demonstrate that $`\{p_1,\mathrm{},p_{2^n}\}`$ contains $`n`$ points $`q_1,\mathrm{},q_n`$ such that $`h(q_1\mathrm{}q_n)=n`$. We show this by induction. Set $`q_1=`$ any $`p_i\{p_1,\mathrm{},p_{2^n}\}`$. Then $`h(q_1)=1`$. We prove the induction step as follows. We assume there exists a $`P_i`$ in $`T`$ with $`h(P_i)=i`$ and $`q_1,\mathrm{},q_i`$ atoms in $`T`$ with $`q_j<P_iji`$. By construction of $`T`$ there is a $`P_{i+1}`$ in $`T`$ with $`P_{i+1}>P_i`$ and a $`q_{i+1}`$, equal to some $`p_j`$ , with $`q_{i+1}<P_{i+1}`$ , $`q_{i+1}q_jji`$, and $`q_{i+1}P_i`$ (if we can’t find such a $`q_{i+1}`$ then any $`p_j<P_{i+1}`$ would also be $`<P_i`$ and, since it follows from the construction of $`T`$ that $`P_{i+1}`$ is the join of some set of $`p_j<P_{i+1}`$ we would have $`P_{i+1}P_i`$ contradicting the fact that the $`P_i`$ arose from an application of Principle II to $`P_{i+1}`$). Now $`q_{i+1}_{ji}q_j`$ and $`h(_{ji+1}q_j)>h(_{ji}q_j)`$. But $`h(_{ji+1}q_j)h(_{ji}q_j)+h(q_j)`$. Therefore $`h(_{ji+1}q_j)=h(_{ji}q_j)+1`$. A set of atoms $`q_1,\mathrm{},q_n,`$ are said to be independent if for any $`q_i`$ and any other $`q_{j_1},\mathrm{},q_{j_k}`$ in the set $`q_iq_{j_1}\mathrm{},q_{j_k}`$. The lattices in any cover of $`M=\{q_1,\mathrm{},q_n\}`$ are consistent with the extension of $`M`$ as a lattice satisfying a height function i.e. a modular lattice. But any set of points $`\{q_1,\mathrm{},q_n\}`$ in a modular lattice with $`q_iq_j=0i,jn`$ and $`h(_{in}q_i)=n`$ are independent and their join, having height $`n`$, if it exists is equal to 1. Hence $`B_{\{q_1,\mathrm{},q_n\}}`$ is unique and Principle III ensures that the structure contains this lattice. ## 6 Construction Principle IIa Principles I-IV generate and are realised by the lattice of propositions of classical physics. As we use each principle to enlarge the structure we choose constants from $`Const`$, to satisfy the new relation generated by the principle. These new constants need only be related to points already chosen in a manner implied by the relations generated so far and these are the only relations they must satisfy. Therefore there may be many possible choices of constant to substitute in a relation generated by a given principle and hence many distinct possibilities for a structure realising Principles I-IV. Let us consider a larger structure than one given by Principles I-IV; the structure, call it Q, which comprises all such simultaneously possible structures. What meaning may we attach to Q? If a structure generated by Principles I-IV represents a physical description of the world then Q would contain all the physical descriptions that are simultaneously possible based on our treatment of analysis in a boolean language. Q then represents the fullest description of the world if the world is restricted to statements which belong in some model of this conception of a physical theory. Q is generated by strengthening Principle II in the following way. In the definition of Principle II, for a given $`a`$ in the structure we define elements $`b_a`$ and $`c_a`$, chosen from $`Const`$ , with $`b_ac_a=a`$. To generate $`Q`$ we define the stronger Principle IIa which says there also exists $`b_a^{}Const`$ distinct from $`b_a`$ and $`c_a`$ with $`b_a^{}c_a=a`$ and such that $`ec_ae=c_a`$ and $`b_a^{}e=b_a^{}`$ iff $`e=0`$. It is only this principle which must be strengthened to generate $`Q`$. The 0 and 1 of Principle I are, by definition, unique and no new points are introduced by Principle IV. The points introduced by Principle III are either uniquely defined by $``$ and $``$ from points already in the structure or else generated by taking complements. However we will show that the Principles I, III, IV and the augmented Principle IIa generate a structure in which each point has a non unique complement and so it is not necessary to augment Principle III. Formally we extend Principle II as follows to generate the structure comprising all possibilities that arise in creating a model in the manner described above. Principle $`II_A`$ *Let $`a`$ be a point in the structure to which we may apply the analytic Principle II (subject to Principle IV ) . Then in addition to $`b_a,c_a`$ distinct, with $`b_ac_a=a`$, there also exists $`b_a^{}`$ with $`b_a,c_a,b_a^{}`$ all distinct and $`b_a^{}c_a=a`$* Let $`Q`$ be the structure generated by Principles I, IIa, and III, together with the ad hoc Principle IV restricting the range of $`h`$. ## 7 Q generates and is realised in the lattice of Quantum Mechanical propositions We now show $`Q`$ realises and contains $`P_{n1}`$, the projective lattice of dimension $`n1`$, Birkhoff and Von Neumann’s characterisation of the lattice of quantum mechanical propositions. They define $`P_{n1}`$ as the modular atomic perspective lattice of height bounded by $`n`$ consisting of the subspaces of the projective lattice $`P_{n1}`$ under set intersection and linear sum. However the following equivalent characterisation will also be useful. $`P_{n1}`$ is a lattice defined as follows; Call the atoms of the lattice ’points’, the elements $`\lambda `$ with $`h(\lambda )=2`$ ’lines’, and the elements $`\pi `$ with $`h(\pi )=3`$ ’planes’. We say a point, $`p`$, is on a line, $`\lambda `$, when $`p\lambda `$ in the lattice. For a line, $`\lambda `$, and a plane, $`\pi `$, if $`\lambda \pi `$ the line is said to lie in the plane. $`P_{n1}`$ is then defined to have the following properties. P1 Two distinct points are on one and only one line. P2 If two lines lie in the same plane they have a nonempty intersection. P3 Every line contains at least three points. P3 The set of all points is spanned by $`n`$ points but not by fewer than $`n`$ points. i.e. there is a set of points $`p_1,\mathrm{},p_n`$ such that for $`p`$, a point in $`P_{n1},pp_1p_2,\mathrm{},p_n`$. ### 7.1 $`P_{n1}`$ contains $`Q`$ Let $`P_{n1}`$ be a modular atomic perspective lattice of height bounded by $`n`$. We will show $`P_{n1}`$ realises any statements constructed according to Principles I, IIa, III and IV (and hence these Principles are consistent). $`P_{n1}`$ obviously realises points 0 and 1 satisfying Principle I. Let $`Q`$ , a structure generated by I, IIa, III and IV, be realised in $`P_{n1}`$. Then as in the earlier case, given any $`a`$ in $`Q`$ to which Principle IIa can be applied subject to Principle IV, $`a`$ is realised as an element of height $`>1`$ in $`P_{n1}`$ and hence $`b_a,c_aP_{n1}`$ with $`b_ac_a=a`$ and $`b_ac_a=0`$ as required. Since $`P_{n1}`$ is perspective $`b_aP_{n1},b_a^{}b_a,c_a`$ with $`b_a^{}c_a=b_a^{}b_a=a`$ and $`b_a^{}c_a=0`$. Hence $`P_{n1}`$ realises this additional extension of $`Q`$ arising from the application of Principle IIa (subject to Principle IV). Let $`M`$ be a substructure of $`Q`$ realised in $`P_{n1}`$ and such that Principle III generates a substructure of a boolean lattice, $`B_M`$, containing $`M`$ and consistent with Principle IV as described above. Since $`M`$ is realised in $`P_{n1}`$ each point in $`M`$ may be expressed as a join of atoms in $`P_{n1}`$. Then there is a boolean lattice in $`P_{n1}`$ realising a $`B_M`$, i.e. there is a set of independent atoms in $`P_{n1}`$, $`p_1,\mathrm{},p_n`$ generating a boolean lattice in the same relation to $`M`$ as some $`B_M`$ . Suppose such a set does not exist. Let $`p_1,\mathrm{},p_n`$ be a set of atoms in $`P_{n1}`$ generating the points representing $`M`$. Further there is no smaller set of atoms with the same property. Then by hypothesis $`I,J\{1,\mathrm{},n\}IJ=\mathrm{}`$ and $`_{iI}q_i_{jJ}q_j\mathrm{}`$. Let $`B_{M_i}`$ be some boolean sublattice of $`P_{n1}`$ containing the $`q_i,iI`$. Then the existence of a boolean lattice in $`P_{n1}`$ containing $`M`$ is contradicted by the existence of the $`q_i,iI`$ with the above properties and therefore we have a cover of $`M`$ that contradicts the condition for the application of Principle III. Hence $`P_{n1}`$ realises an extension of $`Q`$ by application of Principle III. In the above $`P_{n1}`$ was shown to realise Principles IIa and III subject to Principle IV. ### 7.2 $`Q`$ contains $`P_{n1}`$ Here we show that a structure generated by I, IIa, III and IV, satisfies conditions (i)-(iv) defining $`P_{n1}`$. Let $`Q`$ be such a structure. (i) For any two distinct atoms $`a,bQabQ`$. Let $`a,b`$ be atoms in $`Q`$. In any given cover of $`M=[a,b]`$ the lattices in the cover are consistent with the existence of $`ab`$ and $`ab`$ satisfying a modular height function i.e. $`h(ab)=2,h(ab)=0`$. They are also consistent with Principle II by which $`c`$ with $`c(ab)=1`$ and $`ece=c,e(ab)=(ab)e=0`$. And , as in the earlier case, repeated application of Principle II to $`c`$ shows the existence of a set of distinct atoms $`q_1,\mathrm{},q_{n2}`$ s.t. in a modular lattice $`\{q_1,\mathrm{}q_{n2},a,b\}`$ generate a $`B_{\{q_1,\mathrm{},q_{n2},a,b\}}`$ and so any cover contains such a boolean lattice and $`Q`$ contains $`ab`$. (ii) Every line contains a third point. Let $`ab`$ define a line $`L,a,b,`$ atoms in $`Q`$. Then by Principle IIa there exists $`b^{}`$ in $`Q`$ with $`ab^{}=L`$. (iii)Two lines $`L,L^{}`$ lying in the same plane $`P`$ have a non empty intersection. In any given cover of $`M=[L_1,L_2,P]`$ the lattices in the cover are consistent with the existence of $`L_1L_2`$ and $`L_1L_2`$ with height determined as follows. $`h(L_1L_2)>2`$ since $`L_1L_2`$, $`h(L_1L_2)3`$ since $`L_1<P`$, $`L_2<P`$ and so $`h(L_1L_2)=3`$, $`L_1L_2=P`$ and $`h(L_1L_2)=h(L_1)+h(L_2)h(L_1L_2)=1.`$ Set $`L_1L_2=q^{}`$ then any set of lattices in the cover are consistent with the existence of a pair of points $`q_1,q_2`$ not contained in any of the other lattices with $`h(q_1)=h(q_2)=1`$, $`q_1L_1=L_1`$, $`q_2L_2=L_2`$ and hence since these obey a height function $`q_1q^{}=L_1`$, $`q_2q^{}=L_2`$, $`q_1q_2q_i=P`$. Arguing as in (i) we can show the existence of a set of atoms $`\{q_4,\mathrm{},q_n\}`$ such that any modular lattice containing $`\{q_1,q_2,q^{},q_4,\mathrm{},q_n\}`$ contains a $`B_{\{L_1,L_2,P\}}`$. Therefore there exists a $`B_{\{L_1,L_2,P\}}`$ in every cover containing $`L_1L_2`$ with $`h(L_1L_2)=1`$ and so $`h(L_1L_2)=1`$ in $`Q`$. (iv) Since Principle IIa implies Principle II and we have already shown Principles I, II, and III subject to Principle IV generate a boolean lattice of height $`n`$ condition P3 is satisfied in $`Q`$. ## 8 Conclusion In this paper we have shown how the exotic modular perspective lattice of quantum mechanical propositions $`P_{n1}`$ can be constructed from a more intuitive set of assumptions about a physical theory - in particular our assumptions about what constitutes an analytic procedure and those about the logical structure of the language in which we deal with this analytic procedure. Forming a model of everything we can construct in this way we generate the lattice of Quantum Mechanical propositions. Any physical theory contains a notion of measurement founded on analysis or the division of the whole into parts. This idea of ‘part of’ is represented by a partial ordering relation . If this partition arises from some finite process ‘inf’ and ‘sup’ may be taken giving a binary relation $``$ which can, potentially, be extended to a modular complemented lattice. If the lattice does not contain perspective elements we can construct this extension in actuality as follows. Any natural language in which we treat the partition operator contains the boolean operators $`_L`$ and $`_L`$ and $`{}_{L}{}^{}{}_{}{}^{}`$ which may be used subject to requirements of consistency to extend the structure generated by the partition operator to a lattice. However in attempting to construct a representation of partition in this way many points from Set may be chosen to represent the elements of the structure as they are generated - i.e. the elements may be named in a variety of ways. For a particular set of such choices the partition structure can be extended to a boolean lattice by the boolean operators of the language in the manner described above. If we consider the structure generated by each such set of choices to be a legitimate set of physical propositions the collection of all simultaneously possible such sets will be the fullest physical description of the world. However the complemented modular lattice to which the partial order can potentially be expanded now contains perspective elements and hence can not be expanded to a boolean lattice by the boolean operators of the language. Instead we can only embed fragments of this structure representing partition in our boolean languages. The model that is implied by the possibility of expanding these fragments in this way is $`P_{n1}`$. The central point to the development of $`P_{n1}`$ described here is that the expansion in terms of the usual boolean operators of the natural notion of partition, when we take in to account the freedom that arises in naming the elements of the structures, is sufficient to generate the problematic modular operator of $`P_{n1}`$. What are the consequences of this development of the quantum mechanical lattice of propositions for the interpretation of quantum mechanics ? Wave particle duality as observed in the two slit experiment is a consequence of $`P_{n1}`$ satisfying the perspectivity property derived above. Another consequence of the independence of the perspectivity relation, given by Principle IIa, from the principles generating a boolean lattice is that there exist propositions in $`P_{n1}`$ outside a given boolean lattice in $`P_{n1}`$. This notion of independence may be reformulated to give many of the interpretations of Quantum Mechanics. e.g. technically randomness can be defined in terms of such independence and this leads in a natural way to the probablistic interpretation of Quantum Mechanics . What has been done here is to provide from an examination of the nature of physical theories a set of generating principles for physical propositions that are formally stronger than, or independent of, principles generating a given boolean lattice of physical propositions – such formal independence being already the basis for most interpretations of Quantum Mechanics . ## 9 Bibliography ‘The logic of Quantum Mechanics’ G.Birkhoff, J. von Neumann Annals of Mathematics 37 1936, 823-43.
warning/0001/hep-ex0001029.html
ar5iv
text
# High energy photon colliders.Invited talk at the International Symposium on New Visions in Laser-Beam Interactions, October 11-15, 1999, Tokyo, Metropolitan University Tokyo, Japan. To be published in Nucl. Instr. and Meth. B. ## 1 Introduction. Fantastic progress in laser technique makes it possible now to consider seriously many different applications of lasers in particle beam physics. I hope that our Symposium (perhaps the first of a series) on New Visions in Laser-Beam Interactions will be very useful for progress in this new branch of science. In this talk I will report on developments in $`\gamma \gamma `$,$`\gamma `$e colliders (shortly Photon Colliders) with energies of about $`10^{12}`$ eV. The key element of the Photon Collider is a powerful laser which is used for production of high energy photons using backward Compton scattering. Such colliders provide a new unique way for the study of matter, similar to e<sup>+</sup>e<sup>-</sup> or pp colliders but even better for the study of some phenomena. The history of the $`\gamma \gamma `$ physics and photon colliders is closely connected with the history of the e<sup>+</sup>e<sup>-</sup> colliders. Since 1970 two-photon physics has been actively studied at e<sup>+</sup>e<sup>-</sup> storage rings in collisions of virtual photons. The spectrum of these photons is $`dn0.035d\omega /\omega `$, so that $`\gamma \gamma `$ luminosity was much lower than that in e<sup>+</sup>e<sup>-</sup> collisions. Nevertheless, these experiments have provided a lot of new information on the nature of elementary particles. The maximum energy of e<sup>+</sup>e<sup>-</sup> storage rings is limited by severe synchrotron radiation. To explore the energy region beyond LEP-II, linear e<sup>+</sup>e<sup>-</sup> colliders (LC) in the range from a few hundred GeV to about 1.5 TeV are under intense study around the world. Three specific projects NLC (North American) , TESLA (European) and JLC (Asian) have published their pre-conceptual design reports and intend to submit full conceptual design reports in 2001-2002. One team at CERN is working on the concept of a multi-TeV linear collider (CLIC) which will be able extend the energy range of LC in future. Unlike the situation in storage rings, in linear colliders each beam is used only once. This make it possible to ”convert” electrons to high energy photons to obtain colliding $`\gamma \gamma `$, $`\gamma `$e beams. Among various methods of $`e\gamma `$ conversion the best one is Compton scattering of the laser light on the high energy electrons. The basic scheme of a photon collider is shown in Fig. 1. Two electron beams after the final focus system are traveling toward the interaction point (IP) and at a distance of about 0.1-1 cm from the IP collide with the focused laser beams. After scattering, the photons have an energy close to that of the initial electrons and follow their direction to the interaction point (IP) (with some small additional angular spread of the order of $`1/\gamma `$), where they collide with a similar counter moving high energy beam or with an electron beam. With reasonable laser parameters (several Joules flash energy) one can “convert” most of the electrons into high energy photons. The luminosity of $`\gamma \gamma `$, $`\gamma `$e collisions will be of the same order of magnitude as the “geometric” luminosity of the basic $`ee`$ beams. Luminosity distributions in $`\gamma \gamma `$ collisions have the characteristic peaks near the maximum invariant masses with a typical width about 10 % (and a few times smaller in $`\gamma `$e collisions). High energy photons can have various polarizations, which is very advantageous for experiments. This idea was proposed by the author and colleagues many years ago and has been further developed and discussed in Refs - and many others papers. The physics at high energy $`\gamma \gamma `$,$`\gamma `$e colliders is very rich and no less interesting than with pp or e<sup>+</sup>e<sup>-</sup> collisions (some examples will be given below). This option has been included in the pre-conceptual design reports of all LC projects -, and work on the full conceptual design is under way. In the present climate of tight HEP budgets we should give very clear answers to the following questions: a) can $`\gamma \gamma `$,$`\gamma `$e collisions give new physics information in addition to e<sup>+</sup>e<sup>-</sup> collisions that could justify an additional collider cost ($``$15%, second interaction region, including detector)? b) is it technically feasible? c) are there enough people for the design and construction of a photon collider and then exploiting its unique science? Items a) and b) are discussed in the main part of this paper. As for the last question, the situation is the following. In the last two decades, the conception of photon colliders has been developed and discussed at many workshops. The bibliography on $`\gamma \gamma `$, $`\gamma `$e physics now numbers over 1000 papers, mostly theoretical. The next phase will require much wider participation of the experimental community. Now the work on photon colliders is being continued within the framework of the Worldwide Study on Physics and Detectors at LC, also the International Collaboration on Photon Colliders has recently been initiated. ## 2 Physics In general, the physics at e<sup>+</sup>e<sup>-</sup> and $`\gamma \gamma `$,$`\gamma `$e colliders is quite similar because the same new particles can be produced. However, the events are complimentary, because the cross sections depend differently on the parameters of the theories. If something new is discovered (Higgs, supersymmetry or … quantum gravity with extra dimensions), the nature of these new phenomena will be better undersood if they are be studied in different reactions. Some phenomena can best be studied at photon colliders. Below I will give several examples. The second aspect, important for physics study, is the luminosity attained by the collider. In the next section it will be shown that in the current LC designs the $`\gamma \gamma `$ luminosity in the high energy peak of the luminosity spectrum is about 20 % of the e<sup>+</sup>e<sup>-</sup> luminosity. However, if beams with smaller emittances are used, the $`\gamma \gamma `$ luminosity can be higher than that of e<sup>+</sup>e<sup>-</sup> collisions. That is because in e<sup>+</sup>e<sup>-</sup> collisions the luminosity is restricted by the collision effects (beamstrahlung, instabilities) which are absent in $`\gamma \gamma `$ collisions. Higgs boson The present ”Standard” model, which describes precisely almost everything at present energies, assumes existence of a very unique particle, the Higgs boson, which is thought to be responsible for the origin of particle masses. It is not found yet, but from existing experimental information it follows that, if it exists, its mass is about 100–200 GeV, i.e. lays in the region of the next linear colliders. In $`\gamma \gamma `$ collisions the Higgs boson will be produced as a single resonance. This process goes via the loop and its cross section is very sensitive to all heavy (even super-heavy) charged particles. The effective cross section is presented in Fig. 2 . Note that here $`L_{\gamma \gamma }`$ is defined as the $`\gamma \gamma `$ luminosity at the high energy luminosity peak ($`z=W_{\gamma \gamma }/2E_e>0.65`$ for $`x=4.8`$) with FWHM about 15%. For comparison, the cross sections of the Higgs production in e<sup>+</sup>e<sup>-</sup> collisions are shown. We see that for $`M_H=`$ 120–250 GeV the effective cross section in $`\gamma \gamma `$ collisions is larger than that in e<sup>+</sup>e<sup>-</sup> collisions by a factor of about 6–30. If the Higgs is light enough, its width is much less than the energy spread in $`\gamma \gamma `$ collisions. It can be detected as a peak in the invariant mass distribution or can be searched for by energy scanning using the very sharp ($`1\%`$) high energy edge of luminosity distribution . The total number of events in the main decay channels $`Hb\overline{b},WW(W^{}),ZZ(Z^{})`$ will be several thousands for a typical integrated luminosity of 10 fb<sup>-1</sup>. The scanning method also enables the measurement of the Higgs mass with a high precision. What is most remarkable in this process? The cross section of the process $`\gamma \gamma Hb\overline{b}`$ is proportional to $`\mathrm{\Gamma }_{\gamma \gamma }(H)\times Br(Hb\overline{b}`$). The branching ratio $`Br(Hb\overline{b})`$ can be measured with high precision in e<sup>+</sup>e<sup>-</sup> collisions in the process with the ”tagged” Higgs production: $`\text{e}\text{+}\text{e}\text{-}ZH`$ . As a result, one can measure the $`\mathrm{\Gamma }_{\gamma \gamma }(H)`$ width at photon colliders with an accuracy better than 2-3% ,. On the other hand, the value of this two-photon decay width is determined by the sum of contributions to the loop of all heavy charge particles with masses up to infinity. So, it is a unique way to ”see” particles which cannot be produced at the accelerators directly (maybe never). The measurement of the Higgs two-photon width reminds me of the experiment on the measurement of the number of neutrino generations at LEP. This experiment showned that there are only three light neutrinos, all of them were known already. But there could be more. That would be a great discovery! Measurement of the Higgs two-photon width is also some kind of counting of unknown particles. The Higgs two-gluon decay width is also sensitive to heavy particles in the loop, but only to those which have strong interactions (like quarks). These two measurements together with the $`\mathrm{\Gamma }_{Z\gamma }(H)`$ width, which could be measured in $`\gamma `$e collisions, will allow us to ”observe” and perhaps understand the nature of invisible heavy charged particles. This would be a great step forward. Charge pair production The second example is the charged pair production. It could be $`W^+W^{}`$ or $`t\overline{t}`$ pairs or some new, for instance, supersymmetric particles. Cross sections for the production of charged scalar, lepton, and top pairs in $`\gamma \gamma `$ collisions are larger than those in e<sup>+</sup>e<sup>-</sup> collisions by a factor of approximately 5–10; for WW production this factor is even larger, about 10–20. The corresponding graphs can be found elsewhere ,. The cross section of the scalar pair production, predicted in some theories, in collision of polarized photons near the threshould, is higher than that in e<sup>+</sup>e<sup>-</sup> collisions by a factor of 10–20(see figs in Refs ,). The cross section in the $`\gamma \gamma `$ collisions near the threshold is very sharp (while in e<sup>+</sup>e<sup>-</sup> it contains a factor $`\beta ^3`$) and can be used for measurement of particle masses. Note, that in e<sup>+</sup>e<sup>-</sup> collision two charged pairs are produced both via annihilation diagram with virtual $`\gamma `$ and $`Z`$ and also via exchange diagrams where new particles can contribute, while in $`\gamma \gamma `$ collisions it is pure QED process which allows the charge of produced particles to be measured unambiguously. This is a good example of complementarity in the study of the same particles in different types of collisions. Accessible masses In $`\gamma `$e collisions, charged particle with a mass higher than that in e<sup>+</sup>e<sup>-</sup> collisions can be produced (a heavy charged particle plus a light neutral), for example, supersymmetric charged particle plus neutralino or new W boson and neutrino. $`\gamma \gamma `$ collisions also provide higher accessible masses for particles which are produced as a single resonance in $`\gamma \gamma `$ collisions (such as the Higgs boson). Quantum gravity effects in Extra Dimensions This new theory suggests a possible explanation of why gravitation forces are so weak in comparison with electroweak forces. According to this theory the gravitational forces are as strong as electroweak forces at small distances in space with extra dimensions and became weak at large distances due to “compactification” of these extra dimensions. It turns out that this extravagant theory can be tested at linear colliders and according to T.Rizzo ($`\gamma \gamma WW`$) and K.Cheung ($`\gamma \gamma \gamma \gamma `$) photon colliders are sensitive up to a factor of 2 higher quantum gravity mass scale than e<sup>+</sup>e<sup>-</sup> collisions. Concluding remark. We have seen that the Higgs and charged pair cross sections in $`\gamma \gamma `$ collisions are higher that those in e<sup>+</sup>e<sup>-</sup>collisions at least by a factor of 5, so, even with 5 times lower $`\gamma \gamma `$ luminosity (as it is approximately in current designs) the number of events in e<sup>+</sup>e<sup>-</sup> and $`\gamma \gamma `$ collisions will be comparable (but physics complementary). However, the possibility of much larger $`\gamma \gamma `$ luminosity is not excluded, see below. ## 3 Lasers, optics The new key element at photon colliders is a powerful laser system which is used for e$`\gamma `$ conversion. Lasers with the required flash energies (several Joules) and pulse duration $``$ 1 ps already exist and are used at several laboratories, the main problem here is the high repetition rate, about 10–15 kHz. One very promising way to overcome this problem is discussed in this paper. It is an optical cavity approach, which allows a considerable reduction of the required peak and average laser power. ### 3.1 Requirements for the laser, wave length, flash energy The processes in the conversions region: Compton scattering and several other important phenomena have been considered in detail in papers ,,,, and references therein. There you can find formulae, figures and explanation of various phenomena in the conversion region as well as requirements for lasers for photon colliders. Laser parameters important for this task are: laser flash energy, duration of laser pulse, wave length and repetition rate. The required wave length follows from the kinematics of Compton scattering . In the conversion region a laser photon with the energy $`\omega _0`$ scatters at a small collision angle $`\alpha _0`$ on a high energy electron with the energy $`E_0`$. The maximum energy of scattered photons (in direction of electrons) $$\omega _m=\frac{x}{x+1}E_0;x=\frac{4E_0\omega _0\mathrm{cos}^2\alpha /2}{m^2c^4}15.3\left[\frac{E_0}{\text{TeV}}\right]\left[\frac{\omega _0}{eV}\right].$$ (1) For example: $`E_0`$ =250 GeV, $`\omega _0=1.17`$ eV ($`\lambda =1.06`$ $`\mu `$m) (Nd:Glass laser) $``$ $`x=4.5`$ and $`\omega /E_0=0.82`$. The energy of the backscattered photons grows with increasing $`x`$. However, at $`x>4.8`$ the high energy photons are lost due to e<sup>+</sup>e<sup>-</sup> creation in the collisions with laser photons ,. The maximum conversion coefficient (effective) at $`x10`$ is about 0.33 while at $`x<4.8`$ it is about 0.65 (one conversion length). The luminosity in the first case will be smaller by a factor of 4. Detailed study of dependence of the maximum $`\gamma \gamma `$ luminosity and monochromaticity on $`x`$ can be found elsewhere . In the laser focus at photon colliders the field is so strong that multi-photon processes can take place, for example, the electron can scatter simultaneously on several laser photons. It is preferable to work in a regime where these effects are small enough, because the shape of the photon spectrum in this case is sharper. Sometimes strong fields can be useful. Due to transverse motion of electrons in the laser wave the effective electron mass is increased and the threshold of e<sup>+</sup>e<sup>-</sup> production is shifted to the higher beam energies, a factor of 1.5–2 is possible without special problems “simply” by adding a laser power. For some tasks, such as the energy scanning of the low mass Higgs, the luminosity spectrum should be very sharp, that is only possible when multi-photon effects are small. From all this it follows that an existing powerful Terawatt solid state laser with the wave length about 1 $`\mu `$m can be used for photon colliders up to c.m.s. energies about 1 TeV. For low energy colliders (for study of the low mass Higgs, for instance), the doubling of the laser frequency may be useful, this can be done with high efficiency. In the calculation of the required flash energy one has to take into account the natural “diffraction” emittance of the laser beam , the maximum allowed value of the field strength (characterized by the parameter $`\xi ^2=e^2\overline{B^2}\lambda ^2/m^2c^4`$) , and the laser spot size at the conversion point which should be larger than that of the electron beam. In the collision scheme with the ”crab-crossing” <sup>1</sup><sup>1</sup>1The crab crossing scheme for beam collisions is obligatory in photon colliders for the removal of disrupted beams . In this scheme the electron bunches are collided with crossing angle $`\alpha _c`$. To preserve the luminosity the electron bunches are tilted (using an RF cavity) with respect to the direction of the beam motion on the angle $`\alpha _c/2`$. The required $`\alpha _c`$ for the projects considered is about 30 mrad . the electron beam is tilted in respect to the direction of motion that creates an additional effective transverse beam size $`\sigma _x=\sigma _z\alpha _c/2`$. The result of MC simulation of $`k^2`$ ($`k`$ is the conversion coefficient, $`k^2`$ is proportional to the $`\gamma \gamma `$ luminosity) for the electron bunch length $`\sigma _z=0.3`$ mm (TESLA project) as a function of the flash energy and parameter $`\xi ^2`$ (in the center of the laser bunch) are shown in figs. 3 and 4. In summary: the required laser flash energy is about 3–5 Joules, which is quite reasonable. However, the LC have a repetition rate of about 10–15 kHz, so the average power of the laser system should be about 50 kW. One possible solution is the multi-laser system which combines pulses into one train using Pockels cells . However, such a system will be very expensive . ### 3.2 Multi-pass laser systems To overcome the “repetition rate” problem it is quite natural to consider a laser system where one laser bunch is used for e$`\gamma `$ conversion many times. Indeed, one Joule laser flash contains about $`10^{19}`$ laser photons and only $`10^{10}10^{11}`$ photons are knocked out in the collision with one electron bunch. The simplest solution is to trap the laser pulse to some optical loop and use it many times. In such a system the laser pulse enters via the film polarizer and then is trapped using Pockels cells and polarization rotating plates. Unfortunately, such a system will not work with Terawatt laser pulses due to a self-focusing effect. Fortunately, there is one way to “create” a powerful laser pulse in the optical “trap” without any material inside. This very promising technique is discussed below. ### 3.3 Laser pulse stacking in an “external” optical cavity. Shortly, the method is the following. Using the train of low energy laser pulses one can create in the external passive cavity (with one mirror having some small transparency) an optical pulse of the same duration but with much higher energy (pulse stacking). This pulse circulates many times in the cavity each time colliding with electron bunches passing the center of the cavity. The idea of pulse stacking is simple but not trivial and not well known in the HEP community (and even to laser experts, though it is as old as the Fabry-Perot interferometer). This method is used now in several experiments on detection of gravitation waves. It was mentioned also in NLC ZDR though without analysis and further development. In my opinion, pulse stacking is very natural for photon colliders and allows not only to build a relatively cheap laser system for $`e\gamma `$ conversion but gives us the practical way for realization of the laser cooling, i.e. opens up the way to ultimate luminosities of photon colliders. As this is very important for photon colliders, let me consider this method in more detail . The principle of pulse stacking is shown in Fig.5. The secret consists in the following. There is a well known optical theorem: at any surface, the reflection coefficients for light coming from one and the other sides have opposite signs. In our case, this means that light from the laser entering through semi-transparent mirror into the cavity interferes with reflected light inside the cavity constructively, while the light leaking from the cavity interferes with the reflected laser light destructively. Namely, this fact produces asymmetry between cavity and space outside the cavity! Let R be the reflection coefficient, T the transparency coefficient and $`\delta `$ the passive losses in the right mirror. From the energy conservation $`R+T+\delta =1`$. Let $`E_1`$ and $`E_0`$ be the amplitudes of the laser field and the field inside the cavity. In equilibrium, $`E_0=E_{0,R}+E_{1,T}`$. Taking into account that $`E_{0,R}=E_0\sqrt{R}`$, $`E_{1,T}=E_1\sqrt{T}`$ and $`\sqrt{R}1T/2\delta /2`$ for $`R1`$ we obtain $`E_0^2/E_1^2=4T/(T+\delta )^2.`$ The maximum ratio of intensities is obtained at $`T=\delta `$, then $`I_0/I_1=1/\delta Q`$, where $`Q`$ is the quality factor of the optical cavity. Even with two metal mirrors inside the cavity, one can hope to get a gain factor of about 50–100; with multi-layer mirrors it can reach $`10^5`$. ILC(TESLA) colliders have 120(2800) electron bunches in the train, so the factor 100(1000) would be perfect for our goal, but even the factor of ten means a drastic reduction of the cost. Obtaining of high gains requires a very good stabilization of cavity size: $`\delta L\lambda /4\pi Q`$, laser wave length: $`\delta \lambda /\lambda \lambda /4\pi QL`$ and distance between the laser and the cavity: $`\delta s\lambda /4\pi `$. Otherwise, the condition of constructive interference will not be fulfilled. Besides, the frequency spectrum of the laser should coincide with the cavity modes, that is automatically fulfilled when the ratio of the cavity length and that of the laser oscillator is equal to an integer number 1, 2, 3… . For $`\lambda =1\mu m`$ and $`Q=100`$, the stability of the cavity length should be about $`10^7`$ cm. In the LIGO experiment on detection of gravitational waves which uses similar techniques with $`L4`$ km and $`Q10^5`$ the expected sensitivity is about $`10^{16}`$ cm. In comparison with this project our goal seems to be very realistic. In HEP literature I have found only one reference on pulse stacking of short pulses ($`1`$ ps) generated by FEL with the wave length of 5 $`\mu `$m. They observed pulses in the cavity with 70 times the energy of the incident FEL pulses, though no long term stabilization was done. Possible layout of the optics at the interaction region scheme is shown in Fig.6. In this variant, there are two optical cavities (one for each colliding electron beam) placed outside the electron beams. Another possible variant has only one cavity common for both electron beams. In this case, it is also possible to arrange two conversion points separated by the distance of several millimeters (as it is required for photon colliders), though the distribution of the field in the cavity is not completely stable in this case (though it may be sufficient for not too large a Q and , it can be made stable in more complicated optical system). Also, mirrors should have holes for electron beams (which does not change the Q factor of the cavity too much). The variant presented in fig.6 is simpler though it requires a factor of 2 higher flash energy. ## 4 Luminosity of photon colliders in current designs. Some results of simulation of $`\gamma \gamma `$ collisions at TESLA, ILC (converged NLC and JLC) and CLIC are presented below in Table 1. Beam parameters were taken the same as those in e<sup>+</sup>e<sup>-</sup> collisions with the exception of the horizontal beta function at the IP which is taken (quite conservatively) equal to 2 mm for all cases, that is several times smaller than that in e<sup>+</sup>e<sup>-</sup> collisions due to the absence of beamstrahlung. The conversion point(CP) is situated at distance $`b=\gamma \sigma _y`$. It is assumed that electron beams have 85% longitudinal polarization and laser photons have 100% circular polarization. We see that the $`\gamma \gamma `$ luminosity in the hard part of the spectrum $`L_{\gamma \gamma }(z>0.65)0.1L(geom)`$, numerically it is about $`(1/6)L_{\text{e}\text{+}\text{e}\text{-}}`$. <sup>2</sup><sup>2</sup>2this is because a) $`L_{\text{e}\text{+}\text{e}\text{-}}1.5L_{geom}`$, factor 1.5 (roughly) is due to the pinch effect: b) $`L_{geom}`$ in the case of photon colliders is larger than that in e<sup>+</sup>e<sup>-</sup> collisions by a factor about 2.5 (in the current projects) due to the smaller $`\beta `$-function Note, that the coefficient $`1/6`$ is not a fundamental constant. The $`\gamma \gamma `$ luminosity in these projects is determined only by “geometric” ee-luminosity. With some new low emittance electron sources or with laser cooling of electron beams after the damping ring (or photo-guns) one can get, in principle, $`L_{\gamma \gamma }(z>0.65)>L_{\text{e}\text{+}\text{e}\text{-}}`$. The limitations and technical feasibility are discussed in the next section. In addition to the $`\gamma \gamma `$ collisions, there is considerable $`\gamma `$e luminosity (see table) and it is possible to study $`\gamma `$e interactions simultaneously with $`\gamma \gamma `$ collisions. The normalized $`\gamma \gamma `$ luminosity spectra for a 0.5 TeV TESLA are shown in Fig.7(left). The luminosity spectrum is decomposed into two parts, with the total helicity of two photons 0 and 2. We see that in the high energy part of the luminosity spectra photons have a high degree of polarization, which is very important for many experiments. In addition to the high energy peak, there is a factor 5–8 larger low energy luminosity. It is produced by photons after multiple Compton scattering and beamstrahlung photons. Fortunately, these events have a large boost and can be easily distinguished from the central high energy events. In the same Fig.7(left) you can see the same spectrum with an additional “soft” cut on the longitudinal momentum of the produced system which suppresses low energy luminosity to a negligible level. Fig.7 (right) shows the same spectrum with a stronger cut on the longitudinal momentum. In this case, the spectrum has a nice peak with FWHM about 7.5%. On first sight such cut is somewhat artificial because one can directly select events with high invariant masses and the minimum width of the invariant mass distribution depends only on the detector resolution. However, there is a very important example when one can obtain a “collider resolution” somewhat better than the “detector resolution”; this is the case of only two jets in the event when one can restrict the longitudinal momentum of the produced system using the acollinearity angle between jets ($`Hb\overline{b},\tau \tau `$, for example). A similar table and distributions for the photon collider on the c.m.s. energy 130 GeV (Higgs collider) can be found in ref.. ## 5 Ultimate $`\gamma \gamma `$, $`\gamma `$e luminosities There is only one collision effect restricting the $`\gamma \gamma `$ luminosity, that is a process of coherent pair creation when the high energy photon is converted into an e<sup>+</sup>e<sup>-</sup> pair in the strong field of the opposing electron beam ,,. It becomes more important at larger collider energies or(and) very short bunches. Detailed analysis of ultimate luminosities at photon colliders was done in the ref. . In the current projects the $`\gamma \gamma `$ luminosities are determined by the “geometric” luminosity of the electron beams. Having electron beams with smaller emittances one can obtain a much higher $`\gamma \gamma `$ luminosity . Below are results of the simulation with the code which takes into account all main processes in beam-beam interactions . Fig.8 shows dependence of the $`\gamma \gamma `$ (solid curves) and $`\gamma `$e (dashed curves) luminosities on the horizontal beam size. The vertical emittance is taken as in TESLA(500), ILC(500) projects (see Table 1). The horizontal beam size was varied by change of horizontal beam emittance keeping the horizontal beta function at the IP constant and equal to 2 mm. One can see that all curves for $`\gamma \gamma `$ luminosity follow their natural behavior: $`Ł1/\sigma _x`$, with the exception of ILC at $`2E_0=1`$ GeV where at small $`\sigma _x`$ the effect of coherent pair creation is seen.<sup>3</sup><sup>3</sup>3This curve has also some ”bend” at large $`\sigma _x`$ that is connected with synchrotron radiation in quads (Oide effect) due to a large horizontal emittance. One can avoid this effect by taking larger $`\beta _x`$ and smaller $`ϵ_{nx}`$. This means that at the same collider the $`\gamma \gamma `$ luminosity can be increased by decreasing the horizontal beam size (see table 1) at least by one order ($`\sigma _x<10`$ nm is difficult due to some effects connected with the crab crossing). Additional increase of $`\gamma \gamma `$ luminosity by a factor about 3 (TESLA), 7(ILC) can be obtained by a further decrease of the vertical emittance . So, using beams with smaller emittances, the $`\gamma \gamma `$ luminosity at TESLA, ILC can be increase by almost 2 orders of magnitude. However, even with one order improvement, the number of “interesting” events (the Higgs, charged pairs) at photon colliders will be larger than that in e<sup>+</sup>e<sup>-</sup> collisions by about one order. This is a nice goal and motivation for photon colliders. In $`\gamma `$e collision (Fig.8, dashed curves), the behavior of the luminosity on $`\sigma _x`$ is different due to additional collision effects: beams repulsion and beamstrahlung. As a result, the luminosity in the high energy peak is not proportional to the “geometric” luminosity. There are several ways of decreasing the transverse beam emittances (their product): optimization of storage rings with long wigglers, development of low-emittance RF (or pulsed photo-guns) with merging many beams with low charge and emittance. Here some progress is certainly possible. Moreover, there is one method which allows further decrease of beam cross sections by two orders of magnitude in comparison with current designs, it is a laser cooling ,. This method is discused in my second talk at this Symposium. Other important aspects for photon colliders are removal of disrupted beams and backgrounds. Discussion of these problems can be found elsewhere , , . ## 6 Conclusion The physics program for photon $`\gamma \gamma `$,$`\gamma `$e colliders is very interesting and the additional cost of the second interaction region is certainly justified. There are no show-stoppers. All processes in the conversion and interaction regions and the limitations of attainable luminosity are well understood. There are ideas on laser and optical scheme designs. However, much remains to be done in terms of detailed studies and experimental tests. Special effort is required for the development of the laser and optics which are the key elements of photon colliders. The present laser technology has, in principle, all elements needed for photon colliders, the development of a practical scheme is the most pressing task now. One of the most promising methods is the optical cavity approach which allows a considerable reduction of the required peak and average laser power. A reduction of one order of magnitude is already sufficient, but for the TESLA collider with a large number of bunches in a train and large spacing between the bunches one can think about 2–3 orders, though this may be difficult due to other reasons. The $`\gamma \gamma `$ luminosity at photon colliders with energy below one TeV can be higher than that in e<sup>+</sup>e<sup>-</sup> collisions, typical cross sections are also several times higher, so one could consider an X-factory (X = Higgs, W, etc.). The main problem here is the generation of polarized electron beams with very small emittances (products of transverse emittances). Optimization of damping rings and development of low emittance multi-gun RF sources is the first step in this direction. The second step requires new technologies. The laser cooling of electron beams is one possible way of achieving ultimate $`\gamma \gamma `$ luminosity. Realization of this method depends on the progress of Laser Technology, especially promising is the method of the storage (stacking) of laser pulses in an optical cavity. Dear participants of the Symposium on New Vision in Laser Beam Interactions, and all laser experts, there is a possibility to build a unique instruments for study of the matter in the next decade: The High Energy Photon Collider. The development of the required laser systems is a very challenging task, we need your knowledge, experience and talent, join us in this exciting undertaking!
warning/0001/nlin0001010.html
ar5iv
text
# A new method to introduce additional separated variables for high-order binary constrained flows ## 1. Introduction For a finite-dimensional integrable Hamiltonian systems (FDIHS), let $`m`$ denote the number of degrees of freedom, and $`P_i,i=1,\mathrm{},m,`$ be functionally independent integrals of motion in involution, the separation of variables means to construct $`m`$ pairs of canonical separated variables $`v_k,u_k,k=1,\mathrm{},m`$, $$\{u_k,u_l\}=\{v_k,v_l\}=0,\{v_k,u_l\}=\delta _{kl},k,l=1,\mathrm{},m,$$ $`1.1`$ and $`m`$ functions $`f_k`$ such that $$f_k(u_k,v_k,P_1,\mathrm{},P_m)=0,k=1,\mathrm{},m,$$ $`1.2`$ which are called separated equations. The equations (1.2) give rise to an explicit factorization of the Liouville tori. For the FDIHSs with the Lax matrices admitting the $`r`$-matrices of the $`XXX,XXZ`$ and $`XYZ`$ type, there is a general approach to their separation of variables \[1-6\]. The corresponding separated equations enable us to express the generating function of canonical transformation in completely separated form as an abelian integral on the associated invariant spectral curve. The resulted linearizing map is essentially the Abel map to the Jacobi variety of the spectral curve, thus providing a link with the algebro-geometric linearization methods given by \[7-9\]. The separation of variables for a FDIHS requires that the number of canonical separated variables $`u_k`$ should be equal to the number $`m`$ of degrees of freedom. In some cases, the number of $`u_k`$ resulted by the normal method may be less than $`m`$ and one needs to introduce some additional canonical separated variables. So far very few models in these cases have been studied. These cases remain to be a challenging problem . The separation of variables for constrained flows of soliton equations has been studied (see, for example, \[4,10-14\]). In recent years binary constrained flows of soliton hierarchies have attracted attention (see, for example, \[15-22\]). However the separation of variables for binary constrained flows has not been studied. The degree of freedom for high-order binary constrained flows admitting $`2\times 2`$ Lax matrices $`M=\left(\begin{array}{cc}A(\lambda )& B(\lambda )\\ C(\lambda )& A(\lambda )\end{array}\right)`$ is a natural number $`2N+k_0`$. Via the Lax matrix $`M`$, $`N+k_0`$ pairs of canonical separated variables $`u_1,\mathrm{},u_{N+k_0}`$ can be introduced by the set of zeros of $`B(\lambda )`$ and $`v_k=A(u_k)`$, and $`N+k_0`$ separated equations can be found from the generation function of intrgrals of motions. In previous papers we pesented a method with two different ways for determining additional $`N`$ pairs of canonical separated variables and additional $`N`$ separated equations for first binary constrained flows with $`2N`$ degree of freedom. The main idea in is to construct two functions $`\stackrel{~}{B}(\lambda )`$ and $`\stackrel{~}{A}(\lambda )`$ and define $`u_{N+1},\mathrm{},u_{2N}`$ by the set of zeros of $`\stackrel{~}{B}(\lambda )`$ and $`v_{N+k}=\stackrel{~}{A}(u_{N+k})`$. The ways for constructing $`\stackrel{~}{B}(\lambda )`$ and $`\stackrel{~}{A}(\lambda )`$ in and are some different. In present paper we propose a completely different method from that in to introduce the additional $`N`$ separated variables and $`N`$ separated equations for high-order binary constrained flows with $`2N+k_0`$ degree of freedom. It is observed that the introduction of $`v_k`$ has some link with integrals of motion and should lead to the separated equations. We find that there are $`N`$ integrals of motion $`P_{N+k_0+1},\mathrm{},P_{2N+k_0}`$ among the $`2N+k_0`$ integrals of motion for the high-order binary constrained flows which commute with $`A(\lambda )`$ and $`B(\lambda )`$. This observation and property stimulate us to directly use the additional integrals of motion to define both the $`N`$ pairs of additional separated variables and $`N`$ separated equations by $`v_{N+k_0+j}=P_{N+k_0+j},j=1,\mathrm{},N`$. Then we can find the conjugated variables $`u_{N+k_0+j},1,\mathrm{},N,`$ commuting with $`A(\lambda )`$ and $`B(\lambda )`$. In contrast to the method in , this method is easer to be applied to the high-order binary constrained flows. We will also present the separation of variables of soliton equations. The first step is to factorize $`(1+1)`$dimensional soliton equations into two commuting $`x`$ and $`t`$FDIHSs via high-order binary constrained flows, namely the $`x`$ and $`t`$dependences of the soliton equations are separated by the $`x`$ and $`t`$FDIHSs obtained from the $`x`$\- and $`t`$-binary constrained flows. The second step is to produce separation of variables for the $`x`$ and $`t`$FDIHDs. Finally, combining the factorization of soliton equations with the Jacobi inversion problems for $`x`$ and $`t`$FDIHSs enables us to establish the Jacobi inversion problems for soliton equations. If the Jacobi inversion problem can be solved by the Jacobi inversion technique , one can obtain the solution in terms of the Riemann-theta function for soliton equations. We illustrate the method by KdV, AKNS and Kaup-Newell (KN) hierarchies. The paper is organized as follows. In section 2, we first recall the high-order binary constrained flows and factorization of the KdV hierarchy. Then propose the mehtod for introducing the $`N`$ pairs of additional separated variables. We illustrate the method by both first binary constrained flow and second binary constrained flow. Finally we present the separation of variables for KdV hierarchy. In section 3 and 4, the method is applied to the AKNS hierarchy and KN hierarchy, respectively. In fact this method can be applied to all high-order binary constrained flows and other soliton hierarchies admitting $`2\times 2`$ Lax pairs. ## 2. Separation of variables for the KdV equations We first recall the high-order binary constrained flows of the KdV hierarchy. ## 2.1 High-order binary constrained flows of the KdV hierarchy Consider the Schr$`\ddot{\text{o}}`$dinger equation $$\varphi _{xx}+(\lambda +u)\varphi =0$$ which is equivalent to the following spectral problem $$\varphi _x=U(u,\lambda )\varphi ,U(u,\lambda )=\left(\begin{array}{cc}0& 1\\ \lambda u& 0\end{array}\right),\varphi =\left(\genfrac{}{}{0pt}{}{\varphi _1}{\varphi _2}\right).$$ $`2.1`$ Take the time evolution law of $`\varphi `$ as $$\varphi _{t_n}=V^{(n)}(u,\lambda )\varphi ,$$ $`2.2`$ where $$V^{(n)}(u,\lambda )=\underset{i=0}{\overset{n+1}{}}\left(\begin{array}{cc}a_i& b_i\\ c_i& a_i\end{array}\right)\lambda ^{n+1i}+\left(\begin{array}{cc}0& 0\\ b_{n+2}& 0\end{array}\right),$$ $$a_0=b_0=0,c_0=1,a_1=0,b_1=1,$$ $$b_{k+1}=Lb_k=\frac{1}{2}L^{k1}u,,a_k=\frac{1}{2}b_{k,x},c_k=\frac{1}{2}b_{k,xx}b_{k+1}b_ku,k=1,2,\mathrm{},$$ $$L=\frac{1}{4}^2u+\frac{1}{2}^1u_x,=_x,^1=^1=1.$$ $`2.3`$ The compatibility condition of (2.1) and (2.2) gives rise to the n-th KdV equation which can be written as an infinite-dimensional Hamiltonian system $$u_{t_n}=2b_{n+2,x}=L^nu=\frac{\delta H_n}{\delta u},$$ $`2.4`$ with the Hamiltonian $`H_n=\frac{4b_{n+3}}{2n+3}`$ and $`\frac{\delta H_n}{\delta u}=2b_{n+2}.`$ For $`n=1`$ we have $$\varphi _{t_1}=V^{(1)}(u,\lambda )\varphi ,V^{(1)}=\left(\begin{array}{cc}\frac{1}{4}u_x& \lambda \frac{1}{2}u\\ \lambda ^2\frac{1}{2}u\lambda +\frac{1}{4}u_{xx}+\frac{1}{2}u^2& \frac{1}{4}u_x\end{array}\right),$$ $`2.5`$ and the equation (2.4) for $`n=1`$ is the well-known KdV equation $$u_{t_1}=\frac{1}{4}(u_{xxx}+6uu_x).$$ $`2.6`$ The adjoint spectral problem reads $$\psi _x=U^T(u,\lambda )\psi ,\psi =\left(\genfrac{}{}{0pt}{}{\psi _1}{\psi _2}\right).$$ $`2.7`$ By means of the formula in , we have $$\frac{\delta \lambda }{\delta u}=Tr[\left(\begin{array}{cc}\varphi _1\psi _1& \varphi _1\psi _2\\ \varphi _2\psi _1& \varphi _2\psi _2\end{array}\right)\frac{U(u,\lambda )}{u}]=\psi _2\varphi _1.$$ According to \[15-22\], the high-order binary $`x`$-constrained flows of the KdV hierarchy (2.4) consist of the equations obtained from the spectral problem (2.1) and the adjoint spectral problem (2.7) for $`N`$ distinct real numbers $`\lambda _j`$ and the restriction of the variational derivatives for the conserved quantities $`H_{k_0}`$ (for any fixed $`k_0`$) and $`\lambda _j`$: $$\mathrm{\Phi }_{1,x}=\mathrm{\Phi }_2,\mathrm{\Phi }_{2,x}=\mathrm{\Lambda }\mathrm{\Phi }_1u\mathrm{\Phi }_1,$$ $`2.8a`$ $$\mathrm{\Psi }_{1,x}=\mathrm{\Lambda }\mathrm{\Psi }_2+u\mathrm{\Psi }_2,\mathrm{\Psi }_{2,x}=\mathrm{\Psi }_1,$$ $`2.8b`$ $$\frac{\delta H_{k_0}}{\delta u}\underset{j=1}{\overset{N}{}}\frac{\delta \lambda _j}{\delta u}=2b_{k_0+2}+<\mathrm{\Psi }_2,\mathrm{\Phi }_1>=0.$$ $`2.8c`$ Hereafter we denote the inner product in $`\text{R}^N`$ by $`<.,.>`$ and $$\mathrm{\Phi }_i=(\varphi _{i1},\mathrm{},\varphi _{iN})^T,\mathrm{\Psi }_i=(\psi _{i1},\mathrm{},\psi _{iN})^T,i=1,2,\mathrm{\Lambda }=diag(\lambda _1,\mathrm{},\lambda _N).$$ The binary $`t_n`$-constrained flows of the KdV hierarchy (2.4) are defined by the replicas of (2.2) and its adjoint system for $`N`$ distinct real number $`\lambda _j`$ $$\left(\genfrac{}{}{0pt}{}{\varphi _{1j}}{\varphi _{2j}}\right)_{t_n}=V^{(n)}(u,\lambda _j)\left(\genfrac{}{}{0pt}{}{\varphi _{1j}}{\varphi _{2j}}\right),\left(\genfrac{}{}{0pt}{}{\psi _{1j}}{\psi _{2j}}\right)_{t_n}=(V^{(n)}(u,\lambda _j))^T\left(\genfrac{}{}{0pt}{}{\psi _{1j}}{\psi _{2j}}\right),j=1,\mathrm{},N,$$ $`2.9a`$ as well as the n-th KdV equation itself (2.4) in the case of the higher-order constraint for $`k_01`$ $$u_{t_n}=2b_{n+2,x}.$$ $`2.9b`$ (1) For $`k_0=0`$, we have $$b_2=\frac{1}{2}u=\frac{1}{2}<\mathrm{\Psi }_2,\mathrm{\Phi }_1>,i.e.,u=<\mathrm{\Psi }_2,\mathrm{\Phi }_1>.$$ $`2.10`$ By substituting (2.10), (2.8a) and (2.8b) becomes a finite-dimensional Hamiltonian system (FDHS) $$\mathrm{\Phi }_{1x}=\frac{F_1}{\mathrm{\Psi }_1},\mathrm{\Phi }_{2x}=\frac{F_1}{\mathrm{\Psi }_2},\mathrm{\Psi }_{1x}=\frac{F_1}{\mathrm{\Phi }_1},\mathrm{\Psi }_{2x}=\frac{F_1}{\mathrm{\Phi }_2},$$ $`2.11`$ $$F_1=<\mathrm{\Psi }_1,\mathrm{\Phi }_2><\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1>+\frac{1}{2}<\mathrm{\Psi }_2,\mathrm{\Phi }_1>^2.$$ Under the constraint (2.10) and the $`x`$-FDHS (2.11), the binary $`t_1`$-constrained flow obtained from (2.9a) with $`V^{(1)}`$ given by (2.5) can also be written as a $`t_1`$-FDHS $$\mathrm{\Phi }_{1,t_1}=\frac{F_2}{\mathrm{\Psi }_1},\mathrm{\Phi }_{2,t_1}=\frac{F_2}{\mathrm{\Psi }_2},\mathrm{\Psi }_{1,t_1}=\frac{F_2}{\mathrm{\Phi }_1},\mathrm{\Psi }_{2,t_1}=\frac{F_2}{\mathrm{\Phi }_2},$$ $`2.12`$ $$F_2=<\mathrm{\Lambda }^2\mathrm{\Psi }_2,\mathrm{\Phi }_1>+<\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>+\frac{1}{2}<\mathrm{\Psi }_2,\mathrm{\Phi }_1><\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1>$$ $$+\frac{1}{2}<\mathrm{\Psi }_2,\mathrm{\Phi }_1><\mathrm{\Psi }_1,\mathrm{\Phi }_2>+\frac{1}{8}(<\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Psi }_1,\mathrm{\Phi }_1>)^2.$$ The Lax representation for the $`x`$-constrained flow (2.8) and the $`t_n`$-constrained flow (2.9) can be deduced from the adjoint representation of (2.1) and (2.2) by using the method in $$M_x=[\stackrel{~}{U},M],M_{t_n}=[\stackrel{~}{V}^{(n)},M],$$ $`2.13`$ where $`\stackrel{~}{U}`$ and $`\stackrel{~}{V}^{(n)}`$ are obtained from $`U`$ and $`V^{(n)}`$ under the system (2.8), and the Lax matrix $`M`$ is of the form $$M=\left(\begin{array}{cc}A(\lambda )& B(\lambda )\\ C(\lambda )& A(\lambda )\end{array}\right).$$ The Lax matrix $`M`$ for $`x`$-FDHS (2.11) and $`t_1`$-FDHS (2.12) is given by $$A(\lambda )=\frac{1}{4}\underset{j=1}{\overset{N}{}}\frac{\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j}}{\lambda \lambda _j},B(\lambda )=1+\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{\psi _{2j}\varphi _{1j}}{\lambda \lambda _j},$$ $$C(\lambda )=\lambda +\frac{1}{2}<\mathrm{\Psi }_2,\mathrm{\Phi }_1>+\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{\psi _{1j}\varphi _{2j}}{\lambda \lambda _j}.$$ $`2.14`$ The generating function of integrals of motion for (2.11) and (2.12) yields $$A^2(\lambda )+B(\lambda )C(\lambda )P(\lambda )=\lambda +\underset{j=1}{\overset{N}{}}[\frac{P_j}{\lambda \lambda _j}+\frac{P_{N+j}^2}{(\lambda \lambda _j)^2}],$$ $`2.15`$ where $`P_1,\mathrm{},P_{2N}`$ are independent integrals of motion for the FDHSs (2.11) and (2.12) $$P_j=\frac{1}{2}\psi _{1j}\varphi _{2j}+(\frac{1}{2}\lambda _j+\frac{1}{4}<\mathrm{\Psi }_2,\mathrm{\Phi }_1>)\psi _{2j}\varphi _{1j}$$ $$+\frac{1}{8}\underset{kj}{}\frac{1}{\lambda _j\lambda _k}[(\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j})(\psi _{1k}\varphi _{1k}\psi _{2k}\varphi _{2k})+4\psi _{1j}\varphi _{2j}\psi _{2k}\varphi _{1k}],$$ $`2.16a`$ $$P_{N+j}=\frac{1}{4}(\psi _{1j}\varphi _{1j}+\psi _{2j}\varphi _{2j}),j=1,\mathrm{},N.$$ $`2.16b`$ It is easy to verify that $$F_1=2\underset{j=1}{\overset{N}{}}P_j,F_2=2\underset{j=1}{\overset{N}{}}(\lambda _jP_j+P_{N+j}^2).$$ $`2.17`$ With respect to the standard Poisson bracket $$\{f,g\}=\underset{j=1}{\overset{N}{}}(\frac{f}{\psi _{1j}}\frac{g}{\varphi _{1j}}+\frac{f}{\psi _{2j}}\frac{g}{\varphi _{2j}}\frac{f}{\varphi _{1j}}\frac{g}{\psi _{1j}}\frac{f}{\varphi _{2j}}\frac{g}{\psi _{2j}}),$$ $`2.18`$ by calculating the formulas like (2.31), it is easy to verify that $$\{A^2(\lambda )+B(\lambda )C(\lambda ),A^2(\mu )+B(\mu )C(\mu )\}=0,$$ $`2.19`$ which implies that $`P_1,..,P_{2N}`$ are in involution, (2.11) and (2.12) are FDIHSs and commute with each other. The construction of (2.11) and (2.12) ensures that if $`(\mathrm{\Psi }_1,\mathrm{\Psi }_2,\mathrm{\Phi }_1,\mathrm{\Phi }_2)`$ satisfies the FDIHSs (2.11) and (2.12) simultaneously, then $`u`$ defined by (2.10) solves the KdV equation (2.6). Set $$A^2(\lambda )+B(\lambda )C(\lambda )=\lambda \underset{k=0}{\overset{\mathrm{}}{}}\stackrel{~}{F}_k\lambda ^k,$$ $`2.20`$ where $`\stackrel{~}{F}_k,k=1,2,\mathrm{},`$ are also integrals of motion for both the $`x`$-FDHSs (2.11) and the $`t_n`$-binary constrained flows (2.9). Comparing (2.20) with (2.15), one gets $$\stackrel{~}{F}_0=1,\stackrel{~}{F}_1=0,\stackrel{~}{F}_k=\underset{j=1}{\overset{N}{}}[\lambda _j^{k2}P_j+(k2)\lambda _j^{k3}P_{N+j}^2],k=2,3,\mathrm{}.$$ $`2.21`$ By employing the method in , the $`t_n`$-FDIHS obtained from the $`t_n`$-binary constrained flow (2.9) is found to be of the form $$\mathrm{\Phi }_{1,t_n}=\frac{F_{n+1}}{\mathrm{\Psi }_1},\mathrm{\Phi }_{2,t_n}=\frac{F_{n+1}}{\mathrm{\Psi }_2},\mathrm{\Psi }_{1,t_n}=\frac{F_{n+1}}{\mathrm{\Phi }_1},\mathrm{\Psi }_{2,t_n}=\frac{F_{n+1}}{\mathrm{\Phi }_2},$$ $`2.22a`$ $$F_{n+1}=\underset{m=0}{\overset{n}{}}(\frac{1}{2})^{m1}\frac{\alpha _m}{m+1}\underset{l_1+\mathrm{}+l_{m+1}=n+2}{}\stackrel{~}{F}_{l_1}\mathrm{}\stackrel{~}{F}_{l_{m+1}},$$ $`2.22b`$ where $`l_11,\mathrm{},l_{m+1}1,\alpha _0=1,\alpha _1=\frac{1}{2},\alpha _2=\frac{3}{2},`$ and $$\alpha _m=2\alpha _{m1}+\underset{l=1}{\overset{m2}{}}\alpha _l\alpha _{ml1}\frac{1}{2}\underset{l=1}{\overset{m1}{}}\alpha _l\alpha _{ml},m3.$$ $`2.22c`$ The n-th KdV equation (2.4) is factorized by the $`x`$-FDIHS (2.11) and the $`t_n`$-FDIHS (2.22). (2) For $`k_0=1`$, one gets $$b_3=\frac{1}{8}(u_{xx}+3u^2)=\frac{1}{2}<\mathrm{\Psi }_2,\mathrm{\Phi }_1>.$$ $`2.23`$ By introducing $`q=u,p=\frac{1}{4}u_x,`$ (2.8a), (2.8b) and (2.23) can be written as a $`x`$-FDHS $$\mathrm{\Phi }_{ix}=\frac{F_1}{\mathrm{\Psi }_i},\mathrm{\Psi }_{ix}=\frac{F_1}{\mathrm{\Phi }_i},i=1,2,q_x=\frac{F_1}{p},p_x=\frac{F_1}{q},$$ $`2.24`$ $$F_1=<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1>+<\mathrm{\Psi }_1,\mathrm{\Phi }_2>q<\mathrm{\Psi }_2,\mathrm{\Phi }_1>+2p^2+\frac{1}{4}q^3.$$ Under the constraint (2.23), $`V^{(1)}`$ becomes $$\stackrel{~}{V}^{(1)}=\left(\begin{array}{cc}p& \lambda \frac{1}{2}q\\ \lambda ^2\frac{1}{2}q\lambda +<\mathrm{\Psi }_2,\mathrm{\Phi }_1>\frac{1}{4}q^2& p\end{array}\right).$$ $`2.25`$ Under the constraint (2.23) and the $`x`$-FDHS (2.24), the binary $`t_1`$-constrained flow consists of (2.9a) with $`V^{(1)}`$ replaced by $`\stackrel{~}{V}^{(1)}`$ and (2.9b) given by (2.6) can also be written as a $`t_1`$-FDHS $$\mathrm{\Phi }_{it_1}=\frac{F_2}{\mathrm{\Psi }_i},\mathrm{\Psi }_{it_1}=\frac{F_2}{\mathrm{\Phi }_i},i=1,2,q_{t_1}=\frac{F_2}{p},p_{t_1}=\frac{F_2}{q},$$ $`2.26`$ $$F_2=<\mathrm{\Lambda }^2\mathrm{\Psi }_2,\mathrm{\Phi }_1>+<\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>\frac{1}{2}q<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1>\frac{1}{2}q<\mathrm{\Psi }_1,\mathrm{\Phi }_2>$$ $$+p<\mathrm{\Psi }_1,\mathrm{\Phi }_1>p<\mathrm{\Psi }_2,\mathrm{\Phi }_2>+\frac{1}{2}<\mathrm{\Psi }_2,\mathrm{\Phi }_1>^2\frac{1}{4}q^2<\mathrm{\Psi }_2,\mathrm{\Phi }_1>.$$ The Lax representations for the $`x`$-FDHS (2.24) and the $`t_1`$-FDHS (2.26), which can can be deduced from the adjoint representation of (2.1) and (2.2), are given by (2.13) with $`\stackrel{~}{V}^{(1)}`$ defined by (2.25) and $`\stackrel{~}{U}`$ obtained from $`U`$ by using $`q`$ instead of $`u`$ as well as $`M`$ given by $$A(\lambda )=p+\frac{1}{4}\underset{j=1}{\overset{N}{}}\frac{\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j}}{\lambda \lambda _j},B(\lambda )=\lambda \frac{1}{2}q+\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{\psi _{2j}\varphi _{1j}}{\lambda \lambda _j},$$ $$C(\lambda )=\lambda ^2\frac{1}{2}q\lambda +\frac{1}{2}<\mathrm{\Psi }_2,\mathrm{\Phi }_1>\frac{1}{4}q^2+\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{\psi _{1j}\varphi _{2j}}{\lambda \lambda _j}.$$ $`2.27`$ The generating function of integrals of motion for (2.24) and (2.26) yields $$A^2(\lambda )+B(\lambda )C(\lambda )P(\lambda )=\lambda ^3+P_0+\underset{j=1}{\overset{N}{}}[\frac{P_j}{\lambda \lambda _j}+\frac{P_{N+j}^2}{(\lambda \lambda _j)^2}],$$ $`2.28`$ where $`P_0,\mathrm{},P_{2N}`$ are independent integrals of motion for the FDHSs (2.24) and (2.26) and $`P_0=\frac{1}{2}F_1`$, $$P_j=\frac{1}{2}\lambda _j^2\psi _{2j}\varphi _{1j}+\frac{1}{2}\lambda _j\psi _{1j}\varphi _{2j}\frac{1}{4}\lambda _jq\psi _{2j}\varphi _{1j}\frac{1}{4}q\psi _{1j}\varphi _{2j}$$ $$+\frac{1}{2}p(\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j})+\frac{1}{4}(<\mathrm{\Psi }_2,\mathrm{\Phi }_1>\frac{1}{2}q^2)\psi _{2j}\varphi _{1j}$$ $$+\frac{1}{8}\underset{kj}{}\frac{1}{\lambda _j\lambda _k}[(\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j})(\psi _{1k}\varphi _{1k}\psi _{2k}\varphi _{2k})+4\psi _{1j}\varphi _{2j}\psi _{2k}\varphi _{1k}],$$ $`2.29a`$ $$P_{N+j}=\frac{1}{4}(\psi _{1j}\varphi _{1j}+\psi _{2j}\varphi _{2j}),j=1,\mathrm{},N.$$ $`2.29b`$ We have $$F_1=2P_0,F_2=2\underset{j=1}{\overset{N}{}}P_j.$$ $`2.30`$ Similarly, it can be shown that (2.24) and (2.26) are FDIHSs and commute with each other. The KdV equation (2.6) is factorized by $`x`$-FDIHS (2.24) and $`t_1`$-FDIHS (2.26). If $`(\mathrm{\Psi }_1,\mathrm{\Psi }_2,p,\mathrm{\Phi }_1,\mathrm{\Phi }_2,q)`$ satisfies the FDIHSs (2.24) and (2.26) simultaneously, then $`u=q`$ solves the KdV equation (2.6). ## 2.2 The separation of variables for the KdV equations (1) For the case $`k_0=0`$, we first consider the separation of variables for FDIHSs (2.11) and (2.12). With respect to the standard Poisson bracket (2.18), it is found that for the $`A(\lambda )`$ and $`B(\lambda )`$ given by (2.14) we have $$\{A(\lambda ),A(\mu )\}=\{B(\lambda ),B(\mu )\}=0,\{A(\lambda ),B(\mu )\}=\frac{1}{2(\lambda \mu )}[B(\mu )B(\lambda )].$$ $`2.31`$ An effective way to introduce the separated variables $`v_k,u_k`$ and to obtain the separated equations is to use the Lax matrix $`M`$ and the generating function of integrals of motion. The commutator relations (2.31) and the equation (2.15) enable us to define the first $`N`$ pairs of the canonical variables $`u_1,\mathrm{},u_N`$ by the set of zeros of $`B(\lambda )`$ \[1-3\] $$B(\lambda )=1+\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{\psi _{2j}\varphi _{1j}}{\lambda \lambda _j}=\frac{R(\lambda )}{K(\lambda )},$$ $`2.32a`$ where $$R(\lambda )=\underset{k=1}{\overset{N}{}}(\lambda u_k),K(\lambda )=\underset{k=1}{\overset{N}{}}(\lambda \lambda _k),$$ and $`v_1,\mathrm{},v_N`$ by $$v_k=2A(u_k),k=1,\mathrm{},N.$$ $`2.32b`$ The commutator relations (2.31) guarantee that $`u_1,\mathrm{},u_N`$ and $`v_1,\mathrm{},v_N`$ satisfy the canonical conditions (1.1) \[1-3\]. Then substituting $`u_k`$ into (2.15) gives rise to the first $`N`$ separated equations $$v_k=2A(u_k)=2\sqrt{P(u_k)}=2\sqrt{u_k+\underset{j=1}{\overset{N}{}}[\frac{P_j}{u_k\lambda _j}+\frac{P_{N+j}^2}{(u_k\lambda _j)^2}]},k=1,\mathrm{},N.$$ $`2.33`$ The FDIHSs (2.11) and (2.12) have $`2N`$ degrees of freedom, we need to introduce the other $`N`$ pairs of canonical variables $`v_k,u_k,k=N+1,\mathrm{},2N`$. Notice that $`P_{N+j}`$ given by (2.16b) are integrals of motion for the FDIHSs (2.11) and (2.12), and satisfy $$\{B(\lambda ),P_{N+j}\}=\{A(\lambda ),P_{N+j}\}=0.$$ $`2.34`$ Thus we may define $$v_{N+j}=2P_{N+j}=\frac{1}{2}(\psi _{1j}\varphi _{1j}+\psi _{2j}\varphi _{2j}),j=1,\mathrm{},N,$$ $`2.35a`$ which also give rise to the separated equations. It is easy to see that if we take $$u_{N+j}=\text{ln}\frac{\varphi _{1j}}{\psi _{2j}},j=1,\mathrm{},N,$$ $`2.35b`$ then $$\{v_{N+j},u_{N+k}\}=\delta _{jk},\{v_{N+j},v_{N+k}\}=\{u_{N+j},u_{N+k}\}=0,j,k=1,\mathrm{},N,$$ $`2.36`$ $$\{B(\lambda ),u_{N+j}\}=\{A(\lambda ),u_{N+j}\}=\{B(\lambda ),v_{N+j}\}=\{A(\lambda ),v_{N+j}\}=0.$$ $`2.37`$ We have the following proposition. ###### Proposition 1 Assume that $`\lambda _j,\varphi _{ij},\psi _{ij}\text{R},i=1,2,j=1,\mathrm{},N`$. Introduce the separated variables $`u_1,\mathrm{},u_{2N}`$ and $`v_1,\mathrm{},v_{2N}`$ by (2.32) and (2.35). If $`u_1,\mathrm{},u_N,`$ are single zeros of $`B(\lambda )`$, then $`v_1,\mathrm{},v_{2N}`$ and $`u_1,\mathrm{},u_{2N}`$ are canonically conjugated, i.e., they satisfy (1.1). Proof. By following the similar method in \[1-6,23,24\], it is easy to show that $`v_1,\mathrm{},v_N`$ and $`u_1,\mathrm{},u_N`$ satisfy (1.1). Notice $`B^{}(u_k)0`$. Hereafter the prime denotes the differentiation with respect to $`\lambda `$. It follows from (2.36) and (2.37) that $$0=\{u_{N+k},B(u_l)\}=B^{}(u_l)\{u_{N+k},u_l\}+\{u_{N+k},B(\mu )\}|_{\mu =u_l}=B^{}(u_l)\{u_{N+k},u_l\},$$ $$\{v_k,u_{N+l}\}=2\{A(u_k),u_{N+l}\}$$ $$=2A^{}(u_k)\{u_k,u_{N+l}\}+\{A(\lambda ),u_{N+l}\}|_{\lambda =u_k}=2A^{}(u_k)\{u_k,u_{N+l}\},$$ $`2.38`$ which leads to $`\{u_{N+k},u_l\}=\{u_{N+k},v_l\}=0`$. Similarly we can show that $`\{v_{N+k},u_l\}=\{v_{N+k},v_l\}=0.`$ These together with (2.36) complete the proof. It follows from (2.32a) and (2.35b) that $$u=<\mathrm{\Psi }_2,\mathrm{\Phi }_1>=2\underset{j=1}{\overset{N}{}}(u_j\lambda _j),$$ $`2.39`$ $$\psi _{2j}\varphi _{1j}=2\frac{R(\lambda _j)}{K^{}(\lambda _j)},\frac{\varphi _{1j}}{\psi _{2j}}=e^{u_{N+j}},j=1,\mathrm{},N,$$ or $$\varphi _{1j}=\sqrt{\frac{2R(\lambda _j)e^{u_{N+j}}}{K^{}(\lambda _j)}},\psi _{2j}=\sqrt{\frac{2R(\lambda _j)e^{u_{N+j}}}{K^{}(\lambda _j)}},j=1,\mathrm{},N.$$ $`2.40`$ The separated equations are given by (2.33) and (2.35a). Replacing $`v_k`$ by the partial derivative $`\frac{S}{u_k}`$ of the generating function $`S`$ of the canonical transformation and interpreting the $`P_i`$ as integration constants, the equations (2.33) and (2.35a) give rise to the Hamilton-Jacobi equations which are completely separable and may be integrated to give the completely separated solution $$S(u_1,\mathrm{},u_{2N})=\underset{k=1}{\overset{N}{}}[^{u_k}2\sqrt{P(\lambda )}𝑑\lambda +2P_{N+k}u_{N+k}].$$ $`2.41`$ The linearizing coordinates are then $$Q_i=\frac{S}{P_i}=\underset{k=1}{\overset{N}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda ,i=1,\mathrm{},N,$$ $`2.42a`$ $$Q_{N+i}=\frac{S}{P_{N+i}}=2\underset{k=1}{\overset{N}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +2u_{N+i},i=1,\mathrm{},N.$$ $`2.42b`$ By using (2.17), the linear flow induced by (2.11) is then given by $$Q_i=\gamma _i+x\frac{F_1}{P_i}=\gamma _i+2x,Q_{N+i}=2\gamma _{N+i}+x\frac{F_1}{P_{N+i}}=2\gamma _{N+i},i=1,\mathrm{},N.$$ $`2.43`$ Hereafter $`\gamma _i,i=1,\mathrm{},2N,`$ are arbitrary constants. Combining the equation (2.42) with the equation (2.43) leads to the Jacobi inversion problem for the FDIHS (2.11) $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda =\gamma _i+2x,i=1,\mathrm{},N,$$ $`2.44a`$ $$\underset{k=1}{\overset{N}{}}[^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}d\lambda +u_{N+i}=\gamma _{N+i},i=1,\mathrm{},N.$$ $`2.44b`$ If, by using the Jacobi inversion technique , $`\varphi _{1j},\psi _{2j},<\mathrm{\Psi }_2,\mathrm{\Phi }_1>`$ given by (2.39) and (2.40) can be obtained from (2.44), then $`\varphi _{2j},\psi _{1j}`$ can be found from the first and the last equation in (2.11) by an algebraic calculation, respectively. The $`(\varphi _{1j},\varphi _{2j},\psi _{1j},\psi _{2j})`$ provides the solution to the FDIHS (2.11). By using (2.17), the linear flow induced by (2.12) is then given by $$Q_i=\overline{\gamma }_i+\frac{F_2}{P_i}t_1=\overline{\gamma }_i+2\lambda _it_1,$$ $$Q_{N+i}=2\overline{\gamma }_{N+i}+\frac{F_2}{P_{N+i}}t_1=2\overline{\gamma }_{N+i}+4P_{N+i}t_1,i=1,\mathrm{},N,$$ $`2.45`$ where $`\overline{\gamma }_i`$ are arbitrary constants. Combining the equation (2.42) with the equation (2.45) leads to the Jacobi inversion problem for the FDIHS (2.12) $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda =\overline{\gamma }_i+2\lambda _it_1,i=1,\mathrm{},N,$$ $`2.46a`$ $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+i}=\overline{\gamma }_{N+i}+2P_{N+i}t_1,i=1,\mathrm{},N.$$ $`2.46b`$ Finally, since the KdV equation (2.6) is factorized by the FDIHSs (2.11) and (2.12), combining the equation (2.44) with the equation (2.46) give rise to the Jacobi inversion problem for the KdV equation (2.6) $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda =\gamma _i+2x+2\lambda _it_1,i=1,\mathrm{},N,$$ $`2.47a`$ $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+i}=\gamma _{N+i}+2P_{N+i}t_1,i=1,\mathrm{},N.$$ $`2.47b`$ If, by using the Jacobi inversion technique , $`u`$ given by (2.39) can be found in terms of Riemann theta functions by solving (2.47), then $`u`$ provides the solution of the KdV equation (2.6). In general, since the n-th KdV equation (2.4) is factorized by the $`x`$-FDIHS (2.11) and the $`t_n`$-FDIHS (2.22), the above procedure can be applied to find the Jacobi inversion problem for the n-th KdV equation (2.4). We have the following proposition. ###### Proposition 2 The Jacobi inversion problem for the n-th KdV equation (2.4) is given by $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda =\gamma _i+2x$$ $$+t_n\underset{m=0}{\overset{n}{}}(\frac{1}{2})^{m1}\alpha _m\underset{l_1+\mathrm{}+l_{m+1}=n+2}{}\lambda _i^{l_{m+1}2}\stackrel{~}{F}_{l_1}\mathrm{}\stackrel{~}{F}_{l_m},i=1,\mathrm{},N,$$ $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+i}=\gamma _{N+i}$$ $$+t_n\underset{m=0}{\overset{n}{}}(\frac{1}{2})^{m2}\alpha _m\underset{l_1+\mathrm{}+l_{m+1}=n+2}{}(l_{m+1}2)\lambda _i^{l_{m+1}3}P_{N+i}\stackrel{~}{F}_{l_1}\mathrm{}\stackrel{~}{F}_{l_m},i=1,\mathrm{},N,$$ where $`l_11,\mathrm{},l_{m+1}1`$ and $`\stackrel{~}{F}_{l_1},\mathrm{}\stackrel{~}{F}_{l_m},`$ are given by (2.21). (2) For the case $`k_0=1`$, we now consider the separation of variables for FDIHSs (2.24) and (2.26). With respect to the standard Poisson bracket, it is found that the $`A(\lambda )`$ and $`B(\lambda )`$ given by (2.27) also satisfy commutator relation (2.31). In the same way, the first $`N+1`$ pairs of the canonical variables $`u_1,\mathrm{},u_{N+1}`$ can be introduced by the set of zeros of $`B(\lambda )`$ $$B(\lambda )=\lambda \frac{1}{2}q+\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{\psi _{2j}\varphi _{1j}}{\lambda \lambda _j}=\frac{R(\lambda )}{K(\lambda )},$$ $`2.48a`$ where $$R(\lambda )=\underset{k=1}{\overset{N+1}{}}(\lambda u_k),K(\lambda )=\underset{k=1}{\overset{N}{}}(\lambda \lambda _k),$$ and $`v_1,\mathrm{},v_{N+1}`$ by $$v_k=2A(u_k),k=1,\mathrm{},N+1.$$ $`2.48b`$ Then substituting $`u_k`$ into (2.28) gives rise to the first $`N+1`$ separated equations $$v_k=2A(u_k)=2\sqrt{P(u_k)}=2\sqrt{u_k^3+P_0+\underset{j=1}{\overset{N}{}}[\frac{P_j}{u_k\lambda _j}+\frac{P_{N+j}^2}{(u_k\lambda _j)^2}]},$$ $$k=1,\mathrm{},N+1.$$ $`2.49`$ The additional $`N`$ pairs of canonical variables can also be defined by the same way $$v_{N+1+j}=2P_{N+j}=\frac{1}{2}(\psi _{1j}\varphi _{1j}+\psi _{2j}\varphi _{2j}),j=1,\mathrm{},N,$$ $`2.50a`$ $$u_{N+1+j}=\text{ln}\frac{\varphi _{1j}}{\psi _{2j}},j=1,\mathrm{},N.$$ $`2.50b`$ In the same way we can show the following proposition. ###### Proposition 3 Assume that $`\lambda _j,\varphi _{ij},\psi _{ij}\text{R},i=1,2,j=1,\mathrm{},N`$. Introduce the separated variables $`u_1,\mathrm{},u_{2N+1}`$ and $`v_1,\mathrm{},v_{2N+1}`$ by (2.48) and (2.50). If $`u_1,\mathrm{},u_{N+1},`$ are single zeros of $`B(\lambda )`$, then $`v_1,\mathrm{},v_{2N+1}`$ and $`u_1,\mathrm{},u_{2N+1}`$ are canonically conjugated, i.e., they satisfy (1.1). It follows from (2.48) and (2.50) that $$u=q=2\underset{j=1}{\overset{N+1}{}}u_j2\underset{j=1}{\overset{N}{}}\lambda _j,$$ $`2.51a`$ $$\varphi _{1j}=\sqrt{\frac{2R(\lambda _j)e^{u_{N+1+j}}}{K^{}(\lambda _j)}},\psi _{2j}=\sqrt{\frac{2R(\lambda _j)e^{u_{N+1+j}}}{K^{}(\lambda _j)}},j=1,\mathrm{},N.$$ $`2.51b`$ The separated equations (2.49) and (2.50a) may be integrated to give the completely separated solution for the generating function $`S`$ of the canonical transformation $$S(u_1,\mathrm{},u_{2N+1})=\underset{k=1}{\overset{N+1}{}}^{u_k}2\sqrt{P(\lambda )}𝑑\lambda +2\underset{k=1}{\overset{N}{}}P_{N+k}u_{N+1+k},$$ $`2.52`$ where $`P(\lambda )`$ is given by (2.28). In the exactly same way, one gets the Jacobi inversion problem for the FDIHS (2.24) $$\underset{k=1}{\overset{N+1}{}}^{u_k}\frac{1}{\sqrt{P(\lambda )}}𝑑\lambda =\gamma _0+2x,$$ $`2.53a`$ $$\underset{k=1}{\overset{N+1}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda =\gamma _i,i=1,\mathrm{},N,$$ $`2.53b`$ $$\underset{k=1}{\overset{N+1}{}}[^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}d\lambda +u_{N+1+i}=\gamma _{N+i},i=1,\mathrm{},N,$$ $`2.53c`$ the Jacobi inversion problem for the FDIHS (2.26) $$\underset{k=1}{\overset{N+1}{}}^{u_k}\frac{1}{\sqrt{P(\lambda )}}𝑑\lambda =\gamma _0,$$ $`2.54a`$ $$\underset{k=1}{\overset{N+1}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda =\gamma _i+2t_1,i=1,\mathrm{},N,$$ $`2.54b`$ $$\underset{k=1}{\overset{N+1}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+1+i}=\gamma _{N+i},i=1,\mathrm{},N.$$ $`2.54c`$ Finally we have the following proposition. ###### Proposition 4 The Jacobi inversion problem for the KdV equation (2.6) $$\underset{k=1}{\overset{N+1}{}}^{u_k}\frac{1}{\sqrt{P(\lambda )}}𝑑\lambda =\gamma _0+2x,$$ $`2.55a`$ $$\underset{k=1}{\overset{N+1}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda =\gamma _i+2t_1,i=1,\mathrm{},N,$$ $`2.55b`$ $$\underset{k=1}{\overset{N+1}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+1+i}=\gamma _{N+i},i=1,\mathrm{},N.$$ $`2.55c`$ If, by using the Jacobi inversion technique , $`u`$ given by (2.51a) can be found in terms of Riemann theta functions by solving (2.55), then $`u`$ provides the solution of the KdV equation (2.6). In general, since the n-th KdV equation (2.4) is factorized by the $`x`$-FDIHS (2.24) and the $`t_n`$-FDIHS obtained from (2.9) under (2.24), the above procedure can be applied to find the Jacobi inversion problem for the n-th KdV equation (2.4). (3) The method can be applied to all high-order binary constrained flows (2.8) and (2.9) as well as the whole KdV hierarchy. For any fixed $`k_0`$, by introducing the so-called Jacobi-Ostrogradsky coordinates, the high-order binary $`x`$-constrained flow (2.8) can be transformed into a $`x`$-FDIHS with degree of freedom $`2N+k_0`$. Under the $`x`$-FDIHS, the binary $`t_n`$-constrained flow (2.9) can also be transformed into a $`t_n`$-FDIHS. The Lax representation for the $`x`$\- and $`t_n`$-FDIHS can be deduced from the adjoint representation of (2.1) and (2.2) by using the method in . By means of the Lax matrix we can introduce the first $`N+k_0`$ canonical variables $`u_1,\mathrm{},u_{N+k_0}`$ by the set of zeros of $`B(\lambda )`$ and $`v_k=2A(u_k),k=1,\mathrm{},N+k_0`$. Then the additional $`N`$ canonical separated variables can be defined by $$v_{N+k_0+j}=2P_{N+j}=\frac{1}{2}(\psi _{1j}\varphi _{1j}+\psi _{2j}\varphi _{2j}),u_{N+k_0+j}=\text{ln}\frac{\varphi _{1j}}{\psi _{2j}},j=1,\mathrm{},N.$$ Finally, since the n-th KdV equation (2.4) is factorized by the $`x`$-FDIHS and the $`t_n`$-FDIHS, in the exactly same way we can obtain the Jacobi inversion problem for (2.4). The above scheme can be applied to all soliton equations admitting $`2\times 2`$ Lax pairs. ## 3. The separation of variables for the AKNS equations ## 3.1 Binary constrained flows of the AKNS hierarchy For the AKNS spectral problem $$\varphi _x=U(u,\lambda )\varphi ,U(u,\lambda )=\left(\begin{array}{cc}\lambda & q\\ r& \lambda \end{array}\right),\varphi =\left(\genfrac{}{}{0pt}{}{\varphi _1}{\varphi _2}\right),u=\left(\genfrac{}{}{0pt}{}{q}{r}\right),$$ $`3.1`$ Take $$\varphi _{t_n}=V^{(n)}(u,\lambda )\varphi ,V^{(n)}(u,\lambda )=\underset{i=0}{\overset{n}{}}\left(\begin{array}{cc}a_i& b_i\\ c_i& a_i\end{array}\right)\lambda ^{ni},$$ $`3.2`$ where $$a_0=1,b_0=c_0=0,a_1=0,b_1=q,c_1=r,a_2=\frac{1}{2}qr,\mathrm{},$$ $$\left(\genfrac{}{}{0pt}{}{c_{k+1}}{b_{k+1}}\right)=L\left(\genfrac{}{}{0pt}{}{c_k}{b_k}\right),a_k=^1(qc_krb_k),k=1,2,\mathrm{},$$ $`3.3`$ $$L=\frac{1}{2}\left(\begin{array}{cc}2r^1q& 2r^1r\\ 2q^1q& +2q^1r\end{array}\right).$$ The AKNS hierarchy associated with (3.1) and (3.2) reads $$u_{t_n}=\left(\genfrac{}{}{0pt}{}{q}{r}\right)_{t_n}=JL^n\left(\genfrac{}{}{0pt}{}{r}{q}\right)=J\frac{\delta H_{n+1}}{\delta u},n=1,2,\mathrm{},$$ $`3.4`$ $$J=\left(\begin{array}{cc}0& 2\\ 2& 0\end{array}\right),H_n=\frac{2a_{n+1}}{n+1},\left(\genfrac{}{}{0pt}{}{c_n}{b_n}\right)=\frac{\delta H_n}{\delta u},n=1,2,\mathrm{}.$$ For $`n=2`$ we have $$\varphi _{t_2}=V^{(2)}(u,\lambda )\varphi ,V^{(2)}=\left(\begin{array}{cc}\lambda ^2+\frac{1}{2}qr& q\lambda \frac{1}{2}q_x\\ r\lambda +\frac{1}{2}r_x& \lambda ^2\frac{1}{2}qr\end{array}\right),$$ $`3.5`$ and the AKNS equation (3.4) for $`n=2`$ reads $$q_{t_2}=\frac{1}{2}q_{xx}+q^2r,r_{t_2}=\frac{1}{2}r_{xx}r^2q.$$ $`3.6`$ The adjoint AKNS spectral problem is of the same form as equation (2.7). We have $$\frac{\delta \lambda }{\delta u}=\left(\genfrac{}{}{0pt}{}{\frac{\delta \lambda }{\delta q}}{\frac{\delta \lambda }{\delta r}}\right)=Tr[\left(\begin{array}{cc}\varphi _1\psi _1& \varphi _1\psi _2\\ \varphi _2\psi _1& \varphi _2\psi _2\end{array}\right)\frac{U(u,\lambda )}{u}]=\left(\genfrac{}{}{0pt}{}{\psi _1\varphi _2}{\psi _2\varphi _1}\right),$$ $`3.7`$ which should be read componentwise . The binary $`x`$-constrained flows of the AKNS hierarchy (3.4) are defined by $$\mathrm{\Phi }_{1,x}=\mathrm{\Lambda }\mathrm{\Phi }_1+q\mathrm{\Phi }_2,\mathrm{\Phi }_{2,x}=r\mathrm{\Phi }_1+\mathrm{\Lambda }\mathrm{\Phi }_2,$$ $`3.8a`$ $$\mathrm{\Psi }_{1,x}=\mathrm{\Lambda }\mathrm{\Psi }_1r\mathrm{\Psi }_2,\mathrm{\Psi }_{2,x}=q\mathrm{\Psi }_1\mathrm{\Lambda }\mathrm{\Psi }_2,$$ $`3.8b`$ $$\frac{\delta H_{k_0+1}}{\delta u}\underset{j=1}{\overset{N}{}}\frac{\delta \lambda _j}{\delta u}=\left(\genfrac{}{}{0pt}{}{c_{k_0+1}}{b_{k_0+1}}\right)\beta \left(\genfrac{}{}{0pt}{}{<\mathrm{\Psi }_1,\mathrm{\Phi }_2>}{<\mathrm{\Psi }_2,\mathrm{\Phi }_1>}\right)=0.$$ $`3.8c`$ (1) For $`k_0=0,\beta =1`$, we have $$\left(\genfrac{}{}{0pt}{}{c_1}{b_1}\right)=\left(\genfrac{}{}{0pt}{}{r}{q}\right)=\left(\genfrac{}{}{0pt}{}{<\mathrm{\Psi }_1,\mathrm{\Phi }_2>}{<\mathrm{\Psi }_2,\mathrm{\Phi }_1>}\right).$$ $`3.9`$ By substituting (3.9) into (3.8a) and (3.8b), one gets a $`x`$-FDHS $$\mathrm{\Phi }_{1x}=\frac{F_1}{\mathrm{\Psi }_1},\mathrm{\Phi }_{2x}=\frac{F_1}{\mathrm{\Psi }_2},\mathrm{\Psi }_{1x}=\frac{F_1}{\mathrm{\Phi }_1},\mathrm{\Psi }_{2x}=\frac{F_1}{\mathrm{\Phi }_2},$$ $`3.10`$ $$F_1=<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_1>+<\mathrm{\Psi }_2,\mathrm{\Phi }_1><\mathrm{\Psi }_1,\mathrm{\Phi }_2>.$$ Under the constraint (3.9) and the FDHS (3.10), the binary $`t_2`$-constrained flow obtained from (3.2) with $`V^{(2)}`$ given by (3.5) and its adjoint equation for $`N`$ distinct real number $`\lambda _j`$ can also be written as a $`t_2`$-FDHS $$\mathrm{\Phi }_{1,t_2}=\frac{F_2}{\mathrm{\Psi }_1},\mathrm{\Phi }_{2,t_2}=\frac{F_2}{\mathrm{\Psi }_2},\mathrm{\Psi }_{1,t_2}=\frac{F_2}{\mathrm{\Phi }_1},\mathrm{\Psi }_{2,t_2}=\frac{F_2}{\mathrm{\Phi }_2},$$ $`3.11`$ $$F_2=<\mathrm{\Lambda }^2\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Lambda }^2\mathrm{\Psi }_1,\mathrm{\Phi }_1>+<\mathrm{\Psi }_2,\mathrm{\Phi }_1><\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>$$ $$+<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1><\mathrm{\Psi }_1,\mathrm{\Phi }_2>\frac{1}{2}(<\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Psi }_1,\mathrm{\Phi }_1>)<\mathrm{\Psi }_2,\mathrm{\Phi }_1><\mathrm{\Psi }_1,\mathrm{\Phi }_2>.$$ The Lax representation for the FDHSs (3.10) and (3.11) which can also be deduced from the adjoint representation of (3.1) and (3.2) are presented by (2.13) with the entries of the Lax matrix $`M`$ given by $$A(\lambda )=1+\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j}}{\lambda \lambda _j},$$ $`3.12a`$ $$B(\lambda )=\underset{j=1}{\overset{N}{}}\frac{\psi _{2j}\varphi _{1j}}{\lambda \lambda _j},C(\lambda )=\underset{j=1}{\overset{N}{}}\frac{\psi _{1j}\varphi _{2j}}{\lambda \lambda _j}.$$ $`3.12b`$ A straightforward calculation yields $$A^2(\lambda )+B(\lambda )C(\lambda )P(\lambda )=1+\underset{j=1}{\overset{N}{}}[\frac{P_j}{\lambda \lambda _j}+\frac{P_{N+j}^2}{(\lambda \lambda _j)^2}],$$ $`3.13`$ where $`P_1,\mathrm{},P_{2N}`$ are independent integrals of motion for the FDHSs (3.10) and (3.11) $$P_j=\psi _{2j}\varphi _{2j}\psi _{1j}\varphi _{1j}$$ $$+\frac{1}{2}\underset{kj}{}\frac{1}{\lambda _j\lambda _k}[(\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j})(\psi _{1k}\varphi _{1k}\psi _{2k}\varphi _{2k})+4\psi _{1j}\varphi _{2j}\psi _{2k}\varphi _{1k}],$$ $`3.14a`$ $$P_{N+j}=\frac{1}{2}(\psi _{1j}\varphi _{1j}+\psi _{2j}\varphi _{2j}),j=1,\mathrm{},N.$$ $`3.14b`$ It is easy to verify that $$F_1=\underset{j=1}{\overset{N}{}}(\lambda _jP_j+P_{N+j}^2)\frac{1}{4}(\underset{j=1}{\overset{N}{}}P_j)^2,$$ $`3.15a`$ $$F_2=\underset{j=1}{\overset{N}{}}(\lambda _j^2P_j+2\lambda _jP_{N+j}^2)\frac{1}{2}(\underset{j=1}{\overset{N}{}}P_j)\underset{j=1}{\overset{N}{}}(\lambda _jP_j+P_{N+j}^2)+\frac{1}{8}(\underset{j=1}{\overset{N}{}}P_j)^3.$$ $`3.15b`$ Similarly, it is easy to show that $`P_1,\mathrm{},P_{2N}`$ are in involution, (3.10) and (3.11) are FDIHSs. The AKNS equation (3.6) is factorized by the $`x`$-FDIHS (3.10) and the $`t_2`$-FDIHS (3.11), namely, if $`(\mathrm{\Psi }_1,\mathrm{\Psi }_2,\mathrm{\Phi }_1,\mathrm{\Phi }_2)`$ satisfies the FDIHSs (3.10) and (3.11) simultaneously, then $`(q,r)`$ given by (3.9) solves the AKNS equation (3.6). In general, the factorization of the n-th AKNS equations (3.4) will be presented in the end of section 3.2. (2) For $`k_0=1,\beta =\frac{1}{2}`$, (3.8c) yields $$r_x=<\mathrm{\Psi }_1,\mathrm{\Phi }_2>,q_x=<\mathrm{\Psi }_2,\mathrm{\Phi }_1>.$$ $`3.16`$ The equations (3.8a), (3.8b) and (3.16) can be written as a $`x`$-FDIHS $$\mathrm{\Phi }_{ix}=\frac{F_1}{\mathrm{\Psi }_1},r_x=\frac{F_1}{q},\mathrm{\Psi }_{ix}=\frac{F_1}{\mathrm{\Phi }_i},q_x=\frac{F_1}{r},i=1,2,$$ $`3.17`$ $$F_1=<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_1>+r<\mathrm{\Psi }_2,\mathrm{\Phi }_1>+q<\mathrm{\Psi }_1,\mathrm{\Phi }_2>.$$ Under the constraint (3.16) and the FDIHS (3.17), $`V^{(2)}`$ becomes $$\stackrel{~}{V}^{(2)}=\left(\begin{array}{cc}\lambda ^2+\frac{1}{2}qr& q\lambda +\frac{1}{2}<\mathrm{\Psi }_2,\mathrm{\Phi }_1>\\ r\lambda +\frac{1}{2}<\mathrm{\Psi }_1,\mathrm{\Phi }_2>& \lambda ^2\frac{1}{2}qr\end{array}\right).$$ $`3.18`$ Then under the constraint (3.16) and the FDIHS (3.17), the binary $`t_2`$-constrained consisting of replicas (3.5) and its adjoint system for $`N`$ distinct real number $`\lambda _j`$ as well as (3.6) can also be written as a $`t_2`$-FDIHS $$\mathrm{\Phi }_{i,t_2}=\frac{F_2}{\mathrm{\Psi }_i},r_{t_2}=\frac{F_2}{q},\mathrm{\Psi }_{i,t_2}=\frac{F_2}{\mathrm{\Phi }_i},q_{t_2}=\frac{F_2}{r}i=1,2,$$ $`3.19`$ $$F_2=<\mathrm{\Lambda }^2\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Lambda }^2\mathrm{\Psi }_1,\mathrm{\Phi }_1>+q<\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>+r<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1>$$ $$\frac{1}{2}qr(<\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Psi }_1,\mathrm{\Phi }_1>)+\frac{1}{2}<\mathrm{\Psi }_2,\mathrm{\Phi }_1><\mathrm{\Psi }_1,\mathrm{\Phi }_2>\frac{1}{2}q^2r^2.$$ The Lax matrix $`M`$ for FDIHS (3.17) and (3.19) is given by $$A(\lambda )=\lambda +\frac{1}{4}\underset{j=1}{\overset{N}{}}\frac{\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j}}{\lambda \lambda _j},$$ $`3.20a`$ $$B(\lambda )=q+\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{\psi _{2j}\varphi _{1j}}{\lambda \lambda _j},C(\lambda )=r+\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{\psi _{1j}\varphi _{2j}}{\lambda \lambda _j}.$$ $`3.20b`$ A straightforward calculation yields $$A^2(\lambda )+B(\lambda )C(\lambda )P(\lambda )=\lambda ^2+P_0+\underset{j=1}{\overset{N}{}}[\frac{P_j}{\lambda \lambda _j}+\frac{P_{N+j}^2}{(\lambda \lambda _j)^2}],$$ $`3.21`$ where $`P_0,\mathrm{},P_{2N}`$ are independent integrals of motion in involution for the FDIHSs (3.17) and (3.19) $$P_0=\frac{1}{2}(<\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Psi }_1,\mathrm{\Phi }_1>)+qr,$$ $$P_j=\frac{1}{2}[\lambda _j\psi _{2j}\varphi _{2j}\lambda _j\psi _{1j}\varphi _{1j}+q\psi _{1j}\varphi _{2j}+r\psi _{2j}\varphi _{1j}]$$ $$+\frac{1}{8}\underset{kj}{}\frac{1}{\lambda _j\lambda _k}[(\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j})(\psi _{1k}\varphi _{1k}\psi _{2k}\varphi _{2k})+4\psi _{1j}\varphi _{2j}\psi _{2k}\varphi _{1k}],$$ $$P_{N+j}=\frac{1}{4}(\psi _{1j}\varphi _{1j}+\psi _{2j}\varphi _{2j}),j=1,\mathrm{},N.$$ It is easy to verify that $$F_1=2\underset{j=1}{\overset{N}{}}P_j,F_2=2\underset{j=1}{\overset{N}{}}(\lambda _jP_j+P_{N+j}^2)\frac{1}{2}P_0^2.$$ $`3.22`$ It is easy to show that $`P_1,\mathrm{},P_{2N}`$ are in involution, (3.17) and (3.18) are FDIHSs and commute each other. The AKNS equation (3.6) is factorized by the $`x`$-FDIHS (3.17) and the $`t_2`$-FDIHS (3.19), namely, if $`(\mathrm{\Psi }_1,\mathrm{\Psi }_2,q,\mathrm{\Phi }_1,\mathrm{\Phi }_2,r)`$ satisfies the FDIHSs (3.17) and (3.19) simultaneously, then $`(q,r)`$ solves the AKNS equation (3.6). ## 3.2 The separation of variables for the AKNS equations (1) For $`k_0=0`$ case, we present the Jacobi inversion problem for (3.10) and (3.11) as well as for (3.6). With respect to the standard Poisson bracket, $`A(\lambda )`$ and $`B(\lambda )`$ given by (3.12) satisfy $$\{A(\lambda ),A(\mu )\}=\{B(\lambda ),B(\mu )\}=0,\{A(\lambda ),B(\mu )\}=\frac{1}{\lambda \mu }[B(\mu )B(\lambda )].$$ $`3.23`$ In contrast with the $`B(\lambda )`$ for the constrained KdV flows, the $`B(\lambda )`$ given by (3.12b) has only $`N1`$ zeros, one has to define the canonical variables $`u_k,v_k,k=1,\mathrm{},N,`$ in a little different way: $$B(\lambda )=\underset{j=1}{\overset{N}{}}\frac{\psi _{2j}\varphi _{1j}}{\lambda \lambda _j}=e^{u_N}\frac{R(\lambda )}{K(\lambda )},R(\lambda )=\underset{k=1}{\overset{N1}{}}(\lambda u_k),K(\lambda )=\underset{k=1}{\overset{N}{}}(\lambda \lambda _k),$$ $`3.24a`$ $$v_k=A(u_k),k=1,\mathrm{},N1,v_N=\frac{1}{2}(<\mathrm{\Psi }_1,\mathrm{\Phi }_1><\mathrm{\Psi }_2,\mathrm{\Phi }_2>).$$ $`3.24b`$ The equation (3.24a) yields $$u_N=\text{ln}<\mathrm{\Psi }_2,\mathrm{\Phi }_1>.$$ $`3.24c`$ Then it is easy to verify that $$\{u_N,B(\mu )\}=\{v_N,A(\mu )\}=0,\{v_N,u_N\}=1,$$ $`3.25a`$ $$\{u_N,A(\mu )\}=\frac{B(\mu )}{<\mathrm{\Psi }_2,\mathrm{\Phi }_1>},\{v_N,B(\mu )\}=B(\mu ).$$ $`3.25b`$ The commutator relations (3.23) and (3.25) guarantee that $`u_1,\mathrm{},u_N,`$ $`v_1,\mathrm{},v_N`$ satisfy the canonical conditions (1.1). Similarly, we define $$v_{N+j}=P_{N+j},u_{N+j}=\text{ln}\frac{\varphi _{1j}}{\psi _{2j}},j=1,\mathrm{},N.$$ $`3.26`$ In the same way we can show the following proposition. ###### Proposition 5 Assume that $`\lambda _j,\varphi _{ij},\psi _{ij}\text{R},i=1,2,j=1,\mathrm{},N`$. Introduce the separated variables $`u_1,\mathrm{},u_{2N}`$ and $`v_1,\mathrm{},v_{2N}`$ by (3.24) and (3.26). If $`u_1,\mathrm{},u_{N1},`$ are single zeros of $`B(\lambda )`$, then $`v_1,\mathrm{},v_{2N}`$ and $`u_1,\mathrm{},u_{2N}`$ are canonically conjugated, i.e., they satisfy (1.1). It follows from (3.24) that $$q=e^{u_N},$$ $`3.27`$ $$\psi _{2j}\varphi _{1j}=e^{u_N}\frac{R(\lambda _j)}{K^{}(\lambda _j)},\frac{\varphi _{1j}}{\psi _{2j}}=e^{u_{N+j}},j=1,\mathrm{},N,$$ or $$\varphi _{1j}=\sqrt{\frac{e^{u_N+u_{N+j}}R(\lambda _j)}{K^{}(\lambda _j)}},\psi _{2j}=\sqrt{\frac{e^{u_Nu_{N+j}}R(\lambda _j)}{K^{}(\lambda _j)}},j=1,\mathrm{},N.$$ $`3.28`$ It is easy to see from (3.13) that $$v_N=\frac{1}{2}(<\mathrm{\Psi }_1,\mathrm{\Phi }_1><\mathrm{\Psi }_2,\mathrm{\Phi }_2>)=\frac{1}{2}\underset{i=1}{\overset{N}{}}P_i.$$ $`3.29`$ Then the separated equations obtained by substituting $`u_k`$ into (3.13) and using (3.24) and the separated equations (3.26) and (3.29) may be integrated to give the generating function of the canonical transformation $$S(u_1,\mathrm{},u_{2N})=\underset{k=1}{\overset{N1}{}}^{u_k}\sqrt{P(\lambda )}𝑑\lambda \frac{1}{2}\underset{i=1}{\overset{N}{}}P_iu_N+\underset{i=1}{\overset{N}{}}P_{N+i}u_{N+i}.$$ $`3.30`$ The linearizing coordinates are then $$Q_i=\frac{S}{P_i}=\frac{1}{2}\underset{k=1}{\overset{N1}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda \frac{1}{2}u_N,i=1,\mathrm{},N,$$ $`3.31a`$ $$Q_{N+i}=\frac{S}{P_{N+i}}=\underset{k=1}{\overset{N1}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+i},i=1,\mathrm{},N.$$ $`3.31b`$ By using (3.15a), the linear flow induced by the FDIHS (3.10) together with the equations (3.31) leads to the Jacobi inversion problem for the FDIHS (3.10) $$\underset{k=1}{\overset{N1}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda u_N=\gamma _i+(2\lambda _i\underset{k=1}{\overset{N}{}}P_k)x,i=1,\mathrm{},N,$$ $`3.32a`$ $$\underset{k=1}{\overset{N1}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+i}=\gamma _{N+i}+2P_{N+i}x,i=1,\mathrm{},N.$$ $`3.32b`$ By using (3.15b), the linear flow induced by the FDIHS (3.11) and the equations (3.31) result in the Jacobi inversion problem for the FDIHS (3.11) $$\underset{k=1}{\overset{N1}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda u_N$$ $$=\overline{\gamma }_i+[2\lambda _i^2\underset{k=1}{\overset{N}{}}(\lambda _kP_k+\lambda _iP_k+P_{N+k}^2)+\frac{3}{4}(\underset{k=1}{\overset{N}{}}P_k)^2]t_2,i=1,\mathrm{},N,$$ $`3.33a`$ $$\underset{k=1}{\overset{N1}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+i}=\overline{\gamma }_{N+i}+P_{N+i}(4\lambda _i\underset{k=1}{\overset{N}{}}P_k)t_2,i=1,\mathrm{},N.$$ $`3.33b`$ Then, since the AKNS equations (3.6) are factorized by the FDIHSs (3.10) and (3.11), combining the equations (3.32) with the equations (3.33) gives rise to the Jacobi inversion problem for the AKNS equations (3.6) $$\underset{k=1}{\overset{N1}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda u_N$$ $$=\gamma _i+(2\lambda _i\underset{k=1}{\overset{N}{}}P_k)x+[2\lambda _i^2\underset{k=1}{\overset{N}{}}(\lambda _kP_k+\lambda _iP_k+P_{N+k}^2)+\frac{3}{4}(\underset{k=1}{\overset{N}{}}P_k)^2]t_2,$$ $`3.34a`$ $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+i}=\gamma _{N+i}+2P_{N+i}x+P_{N+i}(4\lambda _i\underset{k=1}{\overset{N}{}}P_k)t_2,$$ $$i=1,\mathrm{},N.$$ $`3.34b`$ If $`\varphi _{1j},\psi _{2j},q`$ defined by (3.27) and (3.28) can be solved from (3.34) by using the Jacobi inversion technique, then $`\varphi _{2j},\psi _{1j}`$ can be obtained from the first equation and the last equation in (3.10) by an algebraic calculation, respectively. Finally $`q`$ and $`r=<\mathrm{\Psi }_1,\mathrm{\Phi }_2>`$ provides the solution to the AKNS equations (3.6). Comparing (2.20) with (3.13), one gets $$\stackrel{~}{F}_0=1,\stackrel{~}{F}_k=\underset{j=1}{\overset{N}{}}[\lambda _j^{k1}P_j+(k1)\lambda _j^{k2}P_{N+j}^2],k=1,2,\mathrm{},$$ $`3.35`$ where $`\stackrel{~}{F}_k,k=1,2,\mathrm{},`$ are also integrals of motion for both the FDIHS (3.10) and the $`t_n`$-binary constrained flow. The n-th AKNS equations (3.4) are factorized by the $`x`$-FDIHS (3.10) and the $`t_n`$-FDIHS with the Hamiltonian $`F_n`$ given by $$F_n=2\underset{m=0}{\overset{n}{}}(\frac{1}{2})^m\frac{\alpha _m}{m+1}\underset{l_1+\mathrm{}+l_{m+1}=n+1}{}\stackrel{~}{F}_{l_1}\mathrm{}\stackrel{~}{F}_{l_{m+1}},$$ $`3.36`$ where $`l_11,\mathrm{},l_{m+1}1,\alpha _m`$ are given by (2.22c). We have the proposition: ###### Proposition 6 The Jacobi inversion problem for the n-th AKNS equations (3.4) is $$\underset{k=1}{\overset{N1}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda u_N=\gamma _i+(2\lambda _i\underset{k=1}{\overset{N}{}}P_k)x$$ $$+2t_n\underset{m=0}{\overset{n}{}}(\frac{1}{2})^m\alpha _m\underset{l_1+\mathrm{}+l_{m+1}=n+1}{}\lambda _i^{l_{m+1}1}\stackrel{~}{F}_{l_1}\mathrm{}\stackrel{~}{F}_{l_m},i=1,\mathrm{},N,$$ $$\underset{k=1}{\overset{N}{}}[^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}d\lambda +u_{N+i}=\gamma _{N+i}+2P_{N+i}x$$ $$+4t_n\underset{m=0}{\overset{n}{}}(\frac{1}{2})^m\alpha _m\underset{l_1+\mathrm{}+l_{m+1}=n+1}{}(l_{m+1}1)\lambda _i^{l_{m+1}2}P_{N+i}\stackrel{~}{F}_{l_1}\mathrm{}\stackrel{~}{F}_{l_m},i=1,\mathrm{},N,$$ where $`l_11,\mathrm{},l_{m+1}1,`$ and $`\stackrel{~}{F}_{l_1},\mathrm{}\stackrel{~}{F}_{l_m},`$ are given by (3.35). (2) For $`k_0=1`$ case, with respect to the standard Poisson bracket, $`A(\lambda )`$ and $`B(\lambda )`$ given by (3.20) also satisfy the commutator relation (2.31). One define the first $`N+1`$ pair of canonical variables $`u_k,v_k,k=1,\mathrm{},N+1,`$ in the following way: $$B(\lambda )=q+\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{\psi _{2j}\varphi _{1j}}{\lambda \lambda _j}=e^{u_N+1}\frac{R(\lambda )}{K(\lambda )},$$ $`3.37a`$ with $$R(\lambda )=\underset{k=1}{\overset{N}{}}(\lambda u_k),K(\lambda )=\underset{k=1}{\overset{N}{}}(\lambda \lambda _k),$$ and $$v_k=2A(u_k),k=1,\mathrm{},N,$$ $`3.37b`$ $$v_{N+1}=P_0=qr\frac{1}{2}(<\mathrm{\Psi }_1,\mathrm{\Phi }_1><\mathrm{\Psi }_2,\mathrm{\Phi }_2>).$$ $`3.37c`$ The equation (3.24a) yields $$u_{N+1}=\text{ln}q.$$ $`3.37d`$ Then it is easy to verify that $$\{u_{N+1},B(\mu )\}=\{v_{N+1},A(\mu )\}=0,\{v_{N+1},u_{N+1}\}=1,$$ $$\{u_{N+1},A(\mu )\}=0,\{v_{N+1},B(\mu )\}=B(\mu ).$$ $`3.38`$ Similarly, we define $$v_{N+1+j}=2P_{N+j},j=1,\mathrm{},N,$$ $`3.39a`$ $$u_{N+1+j}=\text{ln}\frac{\varphi _{1j}}{\psi _{2j}},j=1,\mathrm{},N.$$ $`3.39b`$ In the same way we can show the following proposition. ###### Proposition 7 Assume that $`\lambda _j,\varphi _{ij},\psi _{ij}\text{R},i=1,2,j=1,\mathrm{},N`$. Introduce the separated variables $`u_1,\mathrm{},u_{2N+1}`$ and $`v_1,\mathrm{},v_{2N+1}`$ by (3.37) and (3.39). If $`u_1,\mathrm{},u_N,`$ are single zeros of $`B(\lambda )`$, then $`v_1,\mathrm{},v_{2N+1}`$ and $`u_1,\mathrm{},u_{2N+1}`$ are canonically conjugated, i.e., they satisfy (1.1). It follows from (3.37) that $$q=e^{u_{N+1}},$$ $`3.40a`$ $$\varphi _{1j}=\sqrt{\frac{2e^{u_{N+1}+u_{N+1+j}}R(\lambda _j)}{K^{}(\lambda _j)}},\psi _{2j}=\sqrt{\frac{2e^{u_{N+1}u_{N+1+j}}R(\lambda _j)}{K^{}(\lambda _j)}},j=1,\mathrm{},N.$$ $`3.40b`$ The first $`N`$ separated equations can be found by substituting $`u_k`$ into (3.21) and using (3.37b), the last $`N+1`$ separated equations are given by (3.37c) and (3.39a). They may be integrated to give $$S(u_1,\mathrm{},u_{2N+1})=\underset{k=1}{\overset{N}{}}(2^{u_k}\sqrt{P(\lambda )}𝑑\lambda +2P_{N+k}u_{N+1+k})+P_0u_{N+1},$$ $`3.41`$ with $`P(\lambda )`$ given by (3.21). Then the Jacobi inversion problem for the FDIHS (3.17) is $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{1}{\sqrt{P(\lambda )}}𝑑\lambda +u_{N+1}=\gamma _0,$$ $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda =\gamma _i+2x,i=1,\mathrm{},N,$$ $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+1+i}=\gamma _{N+i},i=1,\mathrm{},N.$$ $`3.42`$ The Jacobi inversion problem for the FDIHS (3.19) is $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{1}{\sqrt{P(\lambda )}}𝑑\lambda +u_{N+1}=\gamma _0P_0t_2,$$ $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda =\gamma _i+2\lambda _it_2,i=1,\mathrm{},N,$$ $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+1+i}=\gamma _{N+i}+2P_{N+i}t_2,i=1,\mathrm{},N.$$ $`3.43`$ Finally we have ###### Proposition 8 The Jacobi inversion problem for the AKNS equation (3.6) is $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{1}{\sqrt{P(\lambda )}}𝑑\lambda +u_{N+1}=\gamma _0P_0t_2,$$ $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{1}{(\lambda \lambda _i)\sqrt{P(\lambda )}}𝑑\lambda =\gamma _i+2(x+\lambda _it_2),i=1,\mathrm{},N,$$ $$\underset{k=1}{\overset{N}{}}^{u_k}\frac{P_{N+i}}{(\lambda \lambda _i)^2\sqrt{P(\lambda )}}𝑑\lambda +u_{N+1+i}=\gamma _{N+i}+2P_{N+i}t_2,i=1,\mathrm{},N.$$ $`3.44`$ If $`\varphi _{1j},\psi _{2j},q`$ defined by (3.40) can be solved from (3.44) by using the Jacobi inversion technique, then $`\varphi _{2j},\psi _{1j}`$ and $`r`$ can be obtained from the equations in (3.17) by an algebraic calculation, respectively. Finally $`(q,r)`$ provides the solution to the AKNS equations (3.6). (3) The above procedure can be applied to all high-order binary constrained flows (3.8) and and whole AKNS hierarchy (3.4). ## 4. The separation of variables for the Kaup-Newell equations ## 4.1 Binary constrained flows of the Kaup-Newell hierarchy For the Kaup-Newell spectral problem $$\varphi _x=U(u,\lambda )\varphi ,U(u,\lambda )=\left(\begin{array}{cc}\lambda ^2& q\lambda \\ r\lambda & \lambda ^2\end{array}\right),\varphi =\left(\genfrac{}{}{0pt}{}{\varphi _1}{\varphi _2}\right),u=\left(\genfrac{}{}{0pt}{}{q}{r}\right),$$ $`4.1`$ take $$\varphi _{t_n}=V^{(n)}(u,\lambda )\varphi ,V^{(n)}(u,\lambda )=\underset{i=0}{\overset{n1}{}}\left(\begin{array}{cc}a_{2i}\lambda ^{2n2i}& b_{2i+1}\lambda ^{2n2i1}\\ c_{2i+1}\lambda ^{2n2i1}& a_{2i}\lambda ^{2n2i}\end{array}\right)$$ $`4.2`$ where $$a_0=1,a_2=\frac{1}{2}qr,b_1=q,c_1=r,b_3=\frac{1}{2}(q^2r+q_x),c_3=\frac{1}{2}(qr^2r_x),\mathrm{},$$ and in general $`a_{2k+1}=b_{2k}=c_{2k}=0`$ $$\left(\genfrac{}{}{0pt}{}{c_{2k+1}}{b_{2k+1}}\right)=L\left(\genfrac{}{}{0pt}{}{c_{2k1}}{b_{2k1}}\right),a_{2k}=\frac{1}{2}^1(qc_{2k1,x}+rb_{2k1,x}),k=1,2,\mathrm{},$$ $`4.3`$ $$L=\frac{1}{2}\left(\begin{array}{cc}r^1q& r^1r\\ q^1q& q^1r\end{array}\right).$$ Then the compatibility condition of equations (4.1) and (4.2) gives rise to the Kaup-Newell hierarchy $$u_{t_n}=\left(\genfrac{}{}{0pt}{}{q}{r}\right)_{t_n}=J\left(\genfrac{}{}{0pt}{}{c_{2n1}}{b_{2n1}}\right)=J\frac{\delta H_{2n2}}{\delta u},n=1,2,\mathrm{},$$ $`4.4`$ where the Hamiltonian $`H_n`$ and the Hamiltonian operator $`J`$ are given by $$J=\left(\begin{array}{cc}0& \\ & 0\end{array}\right),H_{2n}=\frac{4a_{2n+2}rc_{2n+1}qb_{2n+1}}{2n},\left(\genfrac{}{}{0pt}{}{c_{2n+1}}{b_{2n+1}}\right)=\frac{\delta H_{2n}}{\delta u}.$$ For $`n=2`$ we have $$\varphi _{t_2}=V^{(2)}(u,\lambda )\varphi ,V^{(2)}=\left(\begin{array}{cc}\lambda ^4\frac{1}{2}qr\lambda ^2& q\lambda ^3+\frac{1}{2}(q^2r+q_x)\lambda \\ r\lambda ^3+\frac{1}{2}(qr^2r_x)\lambda & \lambda ^4+\frac{1}{2}qr\lambda ^2\end{array}\right)$$ $`4.5`$ and the coupled derivative nonlinear Schr$`\ddot{\text{o}}`$dinger (CDNS) equations obtained from the equation (4.4) for $`n=2`$ read $$q_{t_2}=\frac{1}{2}q_{xx}+\frac{1}{2}(q^2r)_x,r_{t_2}=\frac{1}{2}r_{xx}+\frac{1}{2}(r^2q)_x.$$ $`4.6`$ The adjoint Kaup-Newell spectral problem is the equation (2.7) with $`U`$ given by (4.1). We have $$\frac{\delta \lambda }{\delta u}=\left(\genfrac{}{}{0pt}{}{\frac{\delta \lambda }{\delta q}}{\frac{\delta \lambda }{\delta r}}\right)=Tr[\left(\begin{array}{cc}\varphi _1\psi _1& \varphi _1\psi _2\\ \varphi _2\psi _1& \varphi _2\psi _2\end{array}\right)\frac{U(u,\lambda )}{u}]=\left(\genfrac{}{}{0pt}{}{\lambda \psi _1\varphi _2}{\lambda \psi _2\varphi _1}\right).$$ $`4.7`$ The binary $`x`$-constrained flows of the Kaup-Newell hierarchy (4.4) are defined by $$\mathrm{\Phi }_{1,x}=\mathrm{\Lambda }^2\mathrm{\Phi }_1+q\mathrm{\Lambda }\mathrm{\Phi }_2,\mathrm{\Phi }_{2,x}=r\mathrm{\Lambda }\mathrm{\Phi }_1+\mathrm{\Lambda }^2\mathrm{\Phi }_2,$$ $`4.8a`$ $$\mathrm{\Psi }_{1,x}=\mathrm{\Lambda }^2\mathrm{\Psi }_1r\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Psi }_{2,x}=q\mathrm{\Lambda }\mathrm{\Psi }_1\mathrm{\Lambda }^2\mathrm{\Psi }_2,$$ $`4.8b`$ $$\frac{\delta H_{k_0}}{\delta u}\underset{j=1}{\overset{N}{}}\frac{\delta \lambda _j}{\delta u}=\left(\genfrac{}{}{0pt}{}{c_{2k_0+1}}{b_{2k_0+1}}\right)\frac{1}{2}\left(\genfrac{}{}{0pt}{}{<\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>}{<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1>}\right)=0.$$ $`4.8c`$ For $`k_0=0`$, we have $$\left(\genfrac{}{}{0pt}{}{c_1}{b_1}\right)=\left(\genfrac{}{}{0pt}{}{r}{q}\right)=\frac{1}{2}\left(\genfrac{}{}{0pt}{}{<\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>}{<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1>}\right).$$ $`4.9`$ By substituting (4.9) into (4.8a) and (4.8b), the first binary $`x`$-constrained flow becomes a FDHS $$\mathrm{\Phi }_{1x}=\frac{F_1}{\mathrm{\Psi }_1},\mathrm{\Phi }_{2x}=\frac{F_1}{\mathrm{\Psi }_2},\mathrm{\Psi }_{1x}=\frac{F_1}{\mathrm{\Phi }_1},\mathrm{\Psi }_{2x}=\frac{F_1}{\mathrm{\Phi }_2},$$ $`4.10`$ with the Hamiltonian $$F_1=<\mathrm{\Lambda }^2\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Lambda }^2\mathrm{\Psi }_1,\mathrm{\Phi }_1>\frac{1}{2}<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1><\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>.$$ Under the constraint (4.9) and the FDHS (4.10), the binary $`t_2`$-constrained flow obtained from (4.5) and its adjoint equation for $`N`$ distinct reral numbers $`\lambda _j`$ can also be written as a FDHS $$\mathrm{\Phi }_{1,t_2}=\frac{F_2}{\mathrm{\Psi }_1},\mathrm{\Phi }_{2,t_2}=\frac{F_2}{\mathrm{\Psi }_2},\mathrm{\Psi }_{1,t_2}=\frac{F_2}{\mathrm{\Phi }_1},\mathrm{\Psi }_{2,t_2}=\frac{F_2}{\mathrm{\Phi }_2},$$ $`4.11`$ with the Hamiltonian $$F_2=<\mathrm{\Lambda }^4\mathrm{\Psi }_2,\mathrm{\Phi }_2>+<\mathrm{\Lambda }^4\mathrm{\Psi }_1,\mathrm{\Phi }_1>+\frac{1}{2}<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1><\mathrm{\Lambda }^3\mathrm{\Psi }_1,\mathrm{\Phi }_2>$$ $$+\frac{1}{2}<\mathrm{\Lambda }^3\mathrm{\Psi }_2,\mathrm{\Phi }_1><\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>\frac{1}{32}<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1>^2<\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>^2$$ $$+\frac{1}{8}(<\mathrm{\Lambda }^2\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Lambda }^2\mathrm{\Psi }_1,\mathrm{\Phi }_1>)<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1><\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>.$$ The Lax representation for the FDHSs (4.10) and (4.11) are presented by (2.13) with the entries of the Lax matrix $`M`$ given by $$A(\lambda )=1+\frac{1}{4}\underset{j=1}{\overset{N}{}}\frac{\lambda _j^2(\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j})}{\lambda ^2\lambda _j^2},$$ $`4.12a`$ $$B(\lambda )=\frac{1}{2}\lambda \underset{j=1}{\overset{N}{}}\frac{\lambda _j\psi _{2j}\varphi _{1j}}{\lambda ^2\lambda _j^2},C(\lambda )=\frac{1}{2}\lambda \underset{j=1}{\overset{N}{}}\frac{\lambda _j\psi _{1j}\varphi _{2j}}{\lambda ^2\lambda _j^2}.$$ $`4.12b`$ A straightforward calculation yields $$A^2(\lambda )+B(\lambda )C(\lambda )P(\lambda )=1+\underset{j=1}{\overset{N}{}}[\frac{P_j}{\lambda ^2\lambda _j^2}+\frac{\lambda _j^4P_{N+j}^2}{(\lambda ^2\lambda _j^2)^2}],$$ $`4.13`$ where $`P_j,j=1,\mathrm{},2N,`$ are $`2N`$ independent integrals of motion for the FDHSs (4.10) and (4.11) $$P_j=\frac{1}{2}\lambda _j^2(\psi _{2j}\varphi _{2j}\psi _{1j}\varphi _{1j})+\frac{1}{8}<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1>\lambda _j\psi _{1j}\varphi _{2j}+\frac{1}{8}<\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>\lambda _j\psi _{2j}\varphi _{1j}$$ $$+\frac{1}{8}\underset{kj}{}\frac{1}{\lambda _j^2\lambda _k^2}[\lambda _j^2\lambda _k^2(\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j})(\psi _{1k}\varphi _{1k}\psi _{2k}\varphi _{2k})+2\lambda _j\lambda _k(\lambda _j^2+\lambda _k^2)\psi _{1j}\varphi _{2j}\psi _{2k}\varphi _{1k}],$$ $$j=1,\mathrm{},N$$ $`4.14a`$ $$P_{N+j}=\frac{1}{4}(\psi _{1j}\varphi _{1j}+\psi _{2j}\varphi _{2j}),j=1,\mathrm{},N.$$ $`4.14b`$ It is easy to varify that $$F_1=2\underset{j=1}{\overset{N}{}}P_j,F_2=2\underset{j=1}{\overset{N}{}}(\lambda _j^2P_j+\lambda _j^4P_{N+j}^2)\frac{1}{2}(\underset{j=1}{\overset{N}{}}P_j)^2,$$ $`4.15a`$ $$<\mathrm{\Psi }_2,\mathrm{\Phi }_2>+<\mathrm{\Psi }_1,\mathrm{\Phi }_1>=4\underset{j=1}{\overset{N}{}}P_{N+j}.$$ $`4.15b`$ By inserting $`\lambda =0`$, (4.13) leads to $$1+\frac{1}{4}(<\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Psi }_1,\mathrm{\Phi }_1>)=\sqrt{P(0)}=\sqrt{1+\underset{j=1}{\overset{N}{}}[P_j\lambda _j^2+P_{N+j}^2]}.$$ $`4.16`$ With respect to the standard Poisson bracket it is found that $$\{A(\lambda ),A(\mu )\}=\{B(\lambda ),B(\mu )\},$$ $`4.17a`$ $$\{A(\lambda ),B(\mu )\}=\frac{\mu }{2(\lambda ^2\mu ^2)}[\mu B(\mu )\lambda B(\lambda )].$$ $`4.17b`$ Then $`\{A^2(\lambda )+B(\lambda )C(\lambda ),A^2(\mu )+B(\mu )C(\mu )\}=0`$ implies that $`P_j,j=1,\mathrm{},2N,`$ are in involution. The CDNS equations (4.6) are factorized by the $`x`$-FDIHS (4.10) and the $`t_2`$-FDIHS (4.11), namely, if $`(\mathrm{\Psi }_1,\mathrm{\Psi }_2,\mathrm{\Phi }_1,\mathrm{\Phi }_2)`$ satisfies the FDIHSs (4.10) and (4.11) simultaneously, then $`(q,r)`$ given by (4.9) solves the CDNS equations (4.6). The factorization of the n-th Kaup-Newell ewuations (4.4) will be presented in the end of section 4.2. ## 4.2 The separation of variables for the Kaup-Newell equations Since the commutator relations (4.17) are quite different from (2.31) and (3.23), we have to modify a little bit of the method presented in sections 2 and 3. Let us denote $`\stackrel{~}{\lambda }=\lambda ^2,\stackrel{~}{\lambda }_j=\lambda _j^2`$. The entries of the Lax matrix $`M`$ given by (4.12) can be rewritten as $$A(\stackrel{~}{\lambda })=1+\frac{1}{4}(<\mathrm{\Psi }_2,\mathrm{\Phi }_2><\mathrm{\Psi }_1,\mathrm{\Phi }_1>)+\frac{1}{2}\stackrel{~}{\lambda }\overline{A}(\stackrel{~}{\lambda }),B(\stackrel{~}{\lambda })=\frac{1}{2}\sqrt{\stackrel{~}{\lambda }}\overline{B}(\stackrel{~}{\lambda }),$$ $`4.18a`$ where $$\overline{A}(\stackrel{~}{\lambda })=\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{\psi _{1j}\varphi _{1j}\psi _{2j}\varphi _{2j}}{\stackrel{~}{\lambda }\stackrel{~}{\lambda }_j},\overline{B}(\stackrel{~}{\lambda })=\underset{j=1}{\overset{N}{}}\frac{\sqrt{\stackrel{~}{\lambda }_j}\psi _{2j}\varphi _{1j}}{\stackrel{~}{\lambda }\stackrel{~}{\lambda }_j}.$$ $`4.18b`$ It is easy to see that $$\{\overline{A}(\stackrel{~}{\lambda }),\overline{A}(\stackrel{~}{\mu })\}=\{\overline{B}(\stackrel{~}{\lambda }),\overline{B}(\stackrel{~}{\mu })\}=0,$$ $`4.19a`$ $$\{\overline{A}(\stackrel{~}{\lambda }),\overline{B}(\stackrel{~}{\mu })\}=\frac{1}{\stackrel{~}{\lambda }\stackrel{~}{\mu }}[\overline{B}(\stackrel{~}{\mu })\overline{B}(\stackrel{~}{\lambda })].$$ $`4.19b`$ It follows from (4.16) and (4.18a) that $$A(\stackrel{~}{\lambda })=\sqrt{1+\underset{j=1}{\overset{N}{}}[P_j\stackrel{~}{\lambda }_j^1+P_{N+j}^2]}+\frac{1}{2}\stackrel{~}{\lambda }\overline{A}(\stackrel{~}{\lambda }).$$ $`4.19c`$ The commutator relations (4.19) and the generating function of integrals of motion (4.13) enable us to introduce $`u_1,\mathrm{},u_N`$ in the following way: $$\overline{B}(\stackrel{~}{\lambda })=\underset{j=1}{\overset{N}{}}\frac{\sqrt{\stackrel{~}{\lambda }_j}\psi _{2j}\varphi _{1j}}{\stackrel{~}{\lambda }\stackrel{~}{\lambda }_j}=e^{u_N}\frac{R(\stackrel{~}{\lambda })}{K(\stackrel{~}{\lambda })},$$ $`4.20a`$ with $$R(\stackrel{~}{\lambda })=\underset{k=1}{\overset{N1}{}}(\stackrel{~}{\lambda }u_k),K(\stackrel{~}{\lambda })=\underset{k=1}{\overset{N}{}}(\stackrel{~}{\lambda }\stackrel{~}{\lambda }_k),$$ and $`v_1,\mathrm{},v_N`$ by $`\overline{A}(\stackrel{~}{\lambda })`$: $$v_k=\overline{A}(u_k),k=1,\mathrm{},N1,$$ $`4.20b`$ $$v_N=<\mathrm{\Psi }_2,\mathrm{\Phi }_2>.$$ $`4.20c`$ The eq. (4.20a) yields $$u_N=\text{ln}<\mathrm{\Lambda }\mathrm{\Psi }_2,\mathrm{\Phi }_1>.$$ $`4.20d`$ Similarly we define $$v_{N+j}=2P_{N+j},j=1,\mathrm{},N,$$ $`4.21a`$ $$u_{N+j}=\text{ln}\frac{\varphi _{1j}}{\psi _{2j}},j=1,\mathrm{},N.$$ $`4.21b`$ Then we have ###### Proposition 9 Assume that $`\lambda _j,\varphi _{ij},\psi _{ij}\text{R},i=1,2,j=1,\mathrm{},N`$. Introduce the separated variables $`u_1,\mathrm{},u_{2N}`$ and $`v_1,\mathrm{},v_{2N}`$ by (4.20) and (4.21). If $`u_1,\mathrm{},u_{N1},`$ are single zeros of $`\overline{B}(\lambda )`$, then $`v_1,\mathrm{},v_{2N}`$ and $`u_1,\mathrm{},u_{2N}`$ are canonically conjugated, i.e., they satisfy (1.1). It follows from (4.9), (4.20a), (4.20d) and (4.21b) that $$q=\frac{1}{2}e^{u_N},$$ $`4.22a`$ $$\varphi _{1j}=\sqrt{\frac{e^{u_N+u_{N+j}}R(\lambda _j^2)}{\lambda _jK^{}(\lambda _j^2)}},\psi _{2j}=\sqrt{\frac{e^{u_Nu_{N+j}}R(\lambda _j^2)}{\lambda _jK^{}(\lambda _j^2)}},,j=1,\mathrm{},N.$$ $`4.22b`$ By substituting $`u_k`$ into (4.13) and using (4.16) and (4.19c), one gets the first $`N1`$ separated equations $$v_k=\overline{A}(u_k)=\frac{2}{u_k}[\sqrt{\stackrel{~}{P}(u_k)}\sqrt{P(0)}],k=1,\mathrm{},N1,$$ $`4.23a`$ where $`P(0)`$ are given by (4.16) and $$\stackrel{~}{P}(\stackrel{~}{\lambda })=1+\underset{j=1}{\overset{N}{}}[\frac{P_j}{\stackrel{~}{\lambda }\lambda _j^2}+\frac{\lambda _j^4P_{N+j}^2}{(\stackrel{~}{\lambda }\lambda _j^2)^2}].$$ It follows from (4.15b), (4.16) and (4.20c) that $$v_N=22\sqrt{P(0)}2\underset{i=1}{\overset{N}{}}P_{N+i}.$$ $`4.23b`$ The separated equations (4.23) and (4.21a) may be integrated to give the generating function of the canonical transformation $$S(u_1,\mathrm{},u_{2N})=\underset{k=1}{\overset{N1}{}}[^{u_k}\frac{2}{\stackrel{~}{\lambda }}\sqrt{\stackrel{~}{P}(\stackrel{~}{\lambda })}𝑑\stackrel{~}{\lambda }2\sqrt{P(0)}ln|u_k|]$$ $$+(22\sqrt{P(0)}2\underset{i=1}{\overset{N}{}}P_{N+i})u_N+2\underset{i=1}{\overset{N}{}}P_{N+i}u_{N+i}.$$ $`4.24`$ The Jacobi inversion problem for the FDIHS (4.10) is $$\underset{k=1}{\overset{N1}{}}[^{u_k}\frac{1}{\stackrel{~}{\lambda }(\stackrel{~}{\lambda }\lambda _i^2)\sqrt{\stackrel{~}{P}(\stackrel{~}{\lambda })}}𝑑\stackrel{~}{\lambda }+\frac{1}{\lambda _i^2\sqrt{P(0)}}ln|u_k|]+\frac{1}{\lambda _i^2\sqrt{P(0)}}u_N=\gamma _i2x,$$ $$\underset{k=1}{\overset{N1}{}}[^{u_k}\frac{\lambda _i^4P_{N+i}}{\stackrel{~}{\lambda }(\stackrel{~}{\lambda }\lambda _i^2)^2\sqrt{\stackrel{~}{P}(\stackrel{~}{\lambda })}}𝑑\stackrel{~}{\lambda }\frac{P_{N+i}}{\sqrt{P(0)}}ln|u_k|]$$ $$(\frac{P_{N+i}}{\sqrt{P(0)}}+1)u_N+u_{N+i}=\gamma _{N+i},i=1,\mathrm{},N.$$ $`4.25`$ The Jacobi inversion problem for the FDIHS (4.11) is $$\underset{k=1}{\overset{N1}{}}[^{u_k}\frac{1}{\stackrel{~}{\lambda }(\stackrel{~}{\lambda }\lambda _i^2)\sqrt{\stackrel{~}{P}(\stackrel{~}{\lambda })}}𝑑\stackrel{~}{\lambda }+\frac{1}{\lambda _i^2\sqrt{P(0)}}ln|u_k|]+\frac{1}{\lambda _i^2\sqrt{P(0)}}u_N$$ $$=\overline{\gamma }_i+(2\lambda _i^2\underset{k=1}{\overset{N}{}}P_k)t_2,$$ $$\underset{k=1}{\overset{N1}{}}[^{u_k}\frac{\lambda _i^4P_{N+i}}{\stackrel{~}{\lambda }(\stackrel{~}{\lambda }\lambda _i^2)^2\sqrt{\stackrel{~}{P}(\stackrel{~}{\lambda })}}𝑑\stackrel{~}{\lambda }\frac{P_{N+i}}{\sqrt{P(0)}}ln|u_k|]$$ $$(\frac{P_{N+i}}{\sqrt{P(0)}}+1)u_N+u_{N+i}=\overline{\gamma }_{N+i}+2\lambda _i^4P_{N+i}t_2,i=1,\mathrm{},N.$$ $`4.26`$ Finally, since the CDNS equations (4.6) are factorized by the FDIHS (4.10) and (4.11), combining the equation (4.25) with the equation (4.26) gives rise to the Jacobi inversion problem for the CDNS equations (4.6) $$\underset{k=1}{\overset{N1}{}}[^{u_k}\frac{1}{\stackrel{~}{\lambda }(\stackrel{~}{\lambda }\lambda _i^2)\sqrt{\stackrel{~}{P}(\stackrel{~}{\lambda })}}𝑑\stackrel{~}{\lambda }+\frac{1}{\lambda _i^2\sqrt{P(0)}}ln|u_k|]+\frac{1}{\lambda _i^2\sqrt{P(0)}}u_N$$ $$=\gamma _i2x+(2\lambda _i^2\underset{k=1}{\overset{N}{}}P_k)t_2,i=1,\mathrm{},N,$$ $`4.27a`$ $$\underset{k=1}{\overset{N1}{}}[^{u_k}\frac{\lambda _i^4P_{N+i}}{\stackrel{~}{\lambda }(\stackrel{~}{\lambda }\lambda _i^2)^2\sqrt{\stackrel{~}{P}(\stackrel{~}{\lambda })}}𝑑\stackrel{~}{\lambda }\frac{P_{N+i}}{\sqrt{P(0)}}ln|u_k|]$$ $$(\frac{P_{N+i}}{\sqrt{P(0)}}+1)u_N+u_{N+i}=\gamma _{N+i}+2\lambda _i^4P_{N+i}t_2,i=1,\mathrm{},N.$$ $`4.27b`$ If $`\varphi _{1j},\psi _{2j},q`$ defined by (4.22) can be solved from (4.36) by using the Jacobi inversion technique, then $`\varphi _{2j},\psi _{1j}`$ can be obtained from the first equation and the last equation in (4.10), respectively. Finally $`q`$ and $`r=<\mathrm{\Lambda }\mathrm{\Psi }_1,\mathrm{\Phi }_2>`$ provides the solution to the CDNS equations (4.6). In general, the above procedure can be applied to the whole Kaup-Newell hierarchy (4.4). Set $$A^2(\lambda )+B(\lambda )C(\lambda )=\underset{k=0}{\overset{\mathrm{}}{}}\stackrel{~}{F}_k\lambda ^{2k},$$ $`4.28a`$ where $`\stackrel{~}{F}_k,k=1,2,\mathrm{},`$ are also integrals of motion for both the $`x`$-FDHSs (4.10) and the $`t_n`$-binary constrained flows (2.16). Comparing (4.28a) with (4.13), one gets $$\stackrel{~}{F}_0=1,\stackrel{~}{F}_k=\underset{j=1}{\overset{N}{}}[\lambda _j^{2k2}P_j+(k1)\lambda _j^{2k}P_{N+j}^2],k=1,2,\mathrm{}.$$ $`4.28b`$ By employing the method in , the $`t_n`$-FDIHS obtained from the $`t_n`$-constrained flow is of the form $$\mathrm{\Phi }_{1,t_n}=\frac{F_n}{\mathrm{\Psi }_1},\mathrm{\Phi }_{2,t_n}=\frac{F_n}{\mathrm{\Psi }_2},\mathrm{\Psi }_{1,t_n}=\frac{F_n}{\mathrm{\Phi }_1},\mathrm{\Psi }_{2,t_n}=\frac{F_n}{\mathrm{\Phi }_2},$$ $`4.29a`$ with the Hamiltonian $$F_n=2\underset{m=0}{\overset{n1}{}}(\frac{1}{2})^m\frac{\alpha _m}{m+1}\underset{l_1+\mathrm{}+l_{m+1}=n}{}\stackrel{~}{F}_{l_1}\mathrm{}\stackrel{~}{F}_{l_{m+1}},$$ $`4.29b`$ where $`l_11,\mathrm{},l_{m+1}1,`$ and $`\alpha _m`$ are given by (2.22). Since the n-th Kaup-Newell equations (4.4) is factorized by the $`x`$-FDIHS (4.10) and the $`t_n`$-FDIHS (4.29). We have the following proposition. ###### Proposition 10 The Jacobi inversion problem for the n-th Kaup-Newell equations (4.4) is given by $$\underset{k=1}{\overset{N1}{}}[^{u_k}\frac{1}{\stackrel{~}{\lambda }(\stackrel{~}{\lambda }\lambda _i^2)\sqrt{\stackrel{~}{P}(\stackrel{~}{\lambda })}}𝑑\stackrel{~}{\lambda }+\frac{1}{\lambda _i^2\sqrt{P(0)}}ln|u_k|]+\frac{1}{\lambda _i^2\sqrt{P(0)}}u_N=\gamma _i2x$$ $$+2t_n\underset{m=0}{\overset{n1}{}}(\frac{1}{2})^m\alpha _m\underset{l_1+\mathrm{}+l_{m+1}=n}{}\lambda _i^{2l_{m+1}2}\stackrel{~}{F}_{l_1}\mathrm{}\stackrel{~}{F}_{l_m},i=1,\mathrm{},N,$$ $`4.30a`$ $$\underset{k=1}{\overset{N1}{}}[^{u_k}\frac{\lambda _i^4P_{N+i}}{\stackrel{~}{\lambda }(\stackrel{~}{\lambda }\lambda _i^2)^2\sqrt{\stackrel{~}{P}(\stackrel{~}{\lambda })}}𝑑\stackrel{~}{\lambda }\frac{P_{N+i}}{\sqrt{P(0)}}ln|u_k|](\frac{P_{N+i}}{\sqrt{P(0)}}+1)u_N+u_{N+i}=\gamma _{N+i}$$ $$+2t_n\underset{m=0}{\overset{n1}{}}(\frac{1}{2})^m\alpha _m\underset{l_1+\mathrm{}+l_{m+1}=n}{}(l_{m+1}1)\lambda _i^{2l_{m+1}}P_{N+i}\stackrel{~}{F}_{l_1}\mathrm{}\stackrel{~}{F}_{l_m},i=1,\mathrm{},N,$$ $`4.30b`$ where $`l_11,\mathrm{},l_{m+1}1,`$ and $`\stackrel{~}{F}_{l_1},\mathrm{}\stackrel{~}{F}_{l_m},`$ are given by (4.28b). For example, the third equations in the Kaup-Newell hierarchy with $`n=3`$ are of the form $$q_{t_3}=\frac{1}{4}q_{xxx}\frac{3}{8}(q^3r^2+2qrq_x)_x,r_{t_3}=\frac{1}{4}r_{xxx}\frac{3}{8}(r^3q^22qrr_x)_x.$$ $`4.31`$ The Kaup-Newell equations (4.31) can be factorized by the $`x`$-FDIHS (4.10) and $`t_3`$-FDIHS with the Hamiltonian $`F_3`$ defined by $$F_3=\underset{j=1}{\overset{N}{}}(2\lambda _j^4P_j+4\lambda _j^6P_{N+j}^2)[\underset{j=1}{\overset{N}{}}(\lambda _j^2P_j+\lambda _j^4P_{N+j}^2)]\underset{j=1}{\overset{N}{}}P_j+\frac{1}{4}(\underset{j=1}{\overset{N}{}}P_j)^3.$$ $`4.32`$ The Jacobi inversion problem for the equations (4.31) is given by $$\underset{k=1}{\overset{N1}{}}[^{u_k}\frac{1}{\stackrel{~}{\lambda }(\stackrel{~}{\lambda }\lambda _i^2)\sqrt{\stackrel{~}{P}(\stackrel{~}{\lambda })}}𝑑\stackrel{~}{\lambda }+\frac{1}{\lambda _i^2\sqrt{P(0)}}ln|u_k|]+\frac{1}{\lambda _i^2\sqrt{P(0)}}u_N$$ $$=\gamma _i2x+[2\lambda _i^4\underset{j=1}{\overset{N}{}}(\lambda _j^2P_j+\lambda _i^2P_j+\lambda _j^4P_{N+j}^2)+\frac{3}{4}(\underset{j=1}{\overset{N}{}}P_j)^2]t_3,i=1,\mathrm{},N,$$ $$\underset{k=1}{\overset{N1}{}}[^{u_k}\frac{\lambda _i^4P_{N+i}}{\stackrel{~}{\lambda }(\stackrel{~}{\lambda }\lambda _i^2)^2\sqrt{\stackrel{~}{P}(\stackrel{~}{\lambda })}}𝑑\stackrel{~}{\lambda }\frac{P_{N+i}}{\sqrt{P(0)}}ln|u_k|](\frac{P_{N+i}}{\sqrt{P(0)}}+1)u_N$$ $$+u_{N+i}=\gamma _{N+i}+[4\lambda _i^6P_{N+i}\lambda _i^4P_{N+j}\underset{j=1}{\overset{N}{}}P_j]t_3,i=1,\mathrm{},N.$$ In general, the method can be applied to all high-order binary constrained flows (4.8) and whole KN hierarchy (4.4) in the exactly same way. ## 4. Concluding remarks For high-order binary constrained flows, the method in \[1-6\] allows us to directly introduce $`N+k_0`$ pairs of canonical separated variables and $`N+k_0`$ separated equations via the Lax matrices and generating function of integrals of motion. In this paper we propose a new method for determining additional $`N`$ pairs of canonical separated variables and $`N`$ additional separated equations for high-order binary constrained flows by directly using $`N`$ additional integrals of motion. This method is completely different from that proposed in and can be applied to all high-order binary constrained flows and other soliton hierarchies admitting $`2\times 2`$ Lax pairs. ## Acknowledgments The author is grateful to Professor V.B. Kuznetsov for useful discussions. This work was supported by the Chinese Basic Research Project “Nonlinear Science”. ## References 1. Sklyanin E.K. 1989 J. Soviet. Math. 47, 2473. 2. Kuznetsov V.B. 1992 J. Math. Phys. 33, 3240. 3. Sklyanin E.K. 1995 Prog. Theor. Phys. Suppl. 118, 35 4. Eilbeck J.C., Enol’skii V.Z., Kuznetsov V.B. and Tsiganov A.V. 1994 J. Phys. A: Math. Gen. 27, 567. 5. Kalnins E.G., Kuznetsov V.B. and Willard Miller Jr 1994 J. Math. Phys. 35, 1710. 6. Harnad J. and Winternitz P. 1995 Commun. Math. Phys. 172, 263. 7. Dubrovin B.A. 1981 Russian Math. Survey 36, 11. 8. Krichever I.M. and Novikov S.P. 1980 Russian Math. Surveys 32, 53. 9. Adler M. and van Moerbeke P. 1980 Adv. Math. 38, 267. 10. Kulish P.P., Rauch-Wojciechowski S. and Tsiganov A.V. 1996 J. Math. Phys. 37, 3463. 11. Zeng Yunbo 1996 Phys. Lett. A 216, 26. 12. Zeng Yunbo 1997 J. Math. Phys. 38, 321. 13. Zeng Yunbo 1997 J. Phys. A: Math. Gen. 30, 3719. 14. Zeng Yunbo 1997 J. Phys. Soc. Jpn. 66, 2277. 15. Ragnisco O. and Rauch-Wojciechowski S. 1992 Inverse Problems 8, 245. 16. Zeng Yunbo 1992 Chinese Science Bulletin 37, 1937. 17. Ma W.X. and Strampp W. 1994 Phys. Lett. A 185, 277. 18. Ma W.X. 1995 J. Phys. Soc. Jpn. 64, 1085. 19. Ma W.X., Fuchssteiner B. and Oevel W. 1996 Physica A 233, 331. 20. Ma W.X., Ding Q., Zhang W.G. and Lu B.Q. 1996 IL Nuovo Cimento B 111, 1135. 21. Li Yishen and Ma W.X. Binary nonlinearization of AKNS spectral problem under higher-order symmetry constraints, to appear in Chaos, Solitons and Fractals. 22. Ma W.X. and Fuchssteiner B. 1996 in: Nonlinear Physics, ed. E. Alfinito, M. Boiti, L. Martina and F. Pempinelli, (Singapore: World Scientific) p217. 23. Zeng Yunbo and Ma W.X. The construction of canonical separated variables for binary constrained AKNS flow, to appear in Physica A. 24. Zeng Yunbo and Ma W.X. Seaparation of variables for soliton equations via their binary constrained flows, to appear in J. Math. Phys. 25. Newell A.C. 1985 Solitons in mathematics and physics (Philadelphia: SIAM). 26. Tu Guizhang, Gradients and coadjoint representation equations of isospectral problem, preprint. 27. Zeng Yunbo and Li Yishen 1993 J. Phys. A: Math. Gen. 26, L273. 28. Zeng Yunbo 1994 Physica D 73, 171. 29. Zeng Yunbo 1991 Phys. Lett. A 160, 541. 30. Ablowitz M. and Segur H. 1981 Solitons and the inverse scattering transform (Philadelphia: SIAM). 31. Kaup D.J. and Newell A.C. 1978 J. Math. Phys. 19, 798.
warning/0001/cond-mat0001259.html
ar5iv
text
# Critical phenomena near the antiferromagnetic quantum critical point of Heavy-Fermions ## I <br>Introduction A central issue in the thermal properties of heavy-fermion compounds is the discovery of a non-Fermi liquid behavior in systems close to an antiferromagnetic quantum phase transition . Such a behavior has been discovered in a series of compounds containing cerium or uranium, for example, $`CeCu_{6x}R_x`$ ($`R`$=$`Au`$, $`Ag`$) , $`CeIn_3`$, $`CePd_2Si_2`$ , $`CeNi_2Ge_2`$ , $`U_{1x}Y_xPd_3`$ and $`Ce_xLa_{1x}Ru_2Si_2`$ . In $`CeCu_{5.9}Au_{0.1}`$, the specific heat $`C`$ depends on T as $`C/Tln(T/T_0)`$, the magnetic susceptibility as $`\chi 1\alpha \sqrt{T}`$, and the T-dependent part of the resistivity as $`\mathrm{\Delta }\rho T`$ (instead of $`C/T\chi Const`$ and $`\mathrm{\Delta }\rho T^2`$ in the Fermi liquid state). Most interestingly, the breakdown of the Fermi liquid behavior can be tuned by alloying (chemical pressure) or by applying a hydrostatic pressure or a magnetic field. The origin of this NFL behavior is a problem of current debate. Different scenarios have been proposed in order to elucidate the non-Fermi liquid (NFL) behavior at the quantum critical point (QCP). On the one hand, there are scenarios involving strong disorder to account for the non-Fermi liquid properties, including (i) a distribution of the Kondo temperature $`P(T_K)`$ induced by disorder following a distribution of the local density of states, and (ii) the possibility of the formation of a Griffiths phase at the quantum critical point . On the other hand, starting from a clean or rather weakly disordered system, there is a scenario (iii) refering to the proximity of a quantum critical point in a theory of itinerant antiferromagnetism , in which the Fermi surface is coupled to critical collective antiferromagnetic excitations. An alternative possibility has been recently put forward on the basis of an analysis of the neutron scattering measurements in $`CeCu_{5.9}Au_{0.1}`$. The unusual spin dynamics that has been observed suggests (iv) the existence of critically screened local magnetic excitations due to the formation of local moments. Finally, it has been proposed (v) a variant of the quadropular Kondo effect in which the interactions of the rare earth (actinide) internal degrees of freedom with the conduction electrons are reduced to a single impurity multichannel Kondo effect. As it stands, the situation remains highly controversial. In this paper, we adopt the point of view (iii) of the proximity of a quantum critical point. The effect of a non zero temperature on quantum critical points is a long-standing problem encountered in itinerant magnetism. We refer to the papers of Hertz and Millis using renormalization group techniques, and of Moriya introducing a self-consistent renormalization (SCR) theory of spin-fluctuations. These theories yield a rich phase diagram with, in addition to the ordering temperature $`T_N`$, a series of crossover temperatures separating different regimes of behaviors. They are well adapted for the treatment of spin fluctuations as present in the single-band Hubbard model but essentially do not account for the Kondo effect. Indeed, the model which is believed to describe the physics of heavy fermions is the Kondo lattice model (KLM) derived from the periodic Anderson model (PAM) in a given limit. The large $`N`$ expansions which have been carried out for these models (where $`N`$ represents simultaneously the degeneracy of the conduction electrons and of the spin channels) give a good description of the Kondo effect. But, unlike the above-mentioned spin-fluctuation theories, they fail to account for the spin-fluctuations since the RKKY interactions only appear at the order $`1/N^2`$ . Hence our motivation to set up an approach of the $`S=1/2`$ Kondo lattice model ($`N=2`$) that enlarges the standard $`1/N`$ expansion theories up on the spin-fluctuation effects. Our previous paper , presented such an approach by performing a generalized Hubbard-Stratonovich transformation on the Kondo interaction term of the Hamiltonian which makes three types of fields (magnetization densities and Kondo-type) appear on an equal footing. The dynamical spin susceptibility was derived in a one-loop expansion associated with the gaussian fluctuations of the magnetization density fields around their saddle-point values. With the aim to describe the critical phenomena around the quantum critical point of the heavy-fermion systems, we propose to push the treatment of the $`S=1/2`$ Kondo lattice model in a self-consistent one-loop approximation in complete analogy with the SCR theory of spin fluctuations developed for the Hubbard model . As in itinerant magnetism, we will show how the system displays a quantum-classical crossover at finite temperature depending whether the temperature is lower or higher than the characteristic energy-scale of the collective mode associated with the magnetic instability. The low temperature behavior in the different regimes essentially depends on the value of the dynamic exponent $`z`$ of the mode and on the dimensionality $`d`$ of the problem. In the case of an antiferromagnetic instability, $`z`$ is found to be equal to $`2`$. Heavy-fermion systems are usually believed to be three-dimensional. However, a recent proposal based on the neutron scattering data obtained in $`CeCu_{6x}Au_x`$ stipulates that the critical magnetic fluctuations which the quasiparticles are coupled to, are simply two-dimensional. So both situations $`d=2`$ and $`d=3`$ will be considered. The rest of the paper is organized as follows. In part II, we introduce the self-consistent one-loop approximation method for the Kondo lattice model using a generalized Hubbard-Stratonovich transformation on the Kondo interaction term. The approach combines both aspects of the $`1/N`$ expansion and of the self-consistent renormalization theory of spin fluctuations. The calculations, presented in the functional integral formalism, closely follow the presentation made by Hertz and Klenin for the Hubbard model. In part III, we extract the general expression of the dynamical spin susceptibility extending the results of the previous paper to a self-consistent treatment. Part IV gives a discussion on the nature of the damped collective mode associated to the proximity of the magnetic transition with a dynamic exponent $`z=2`$ in the antiferromagnetic case. In part V, we derive the resulting quantum-classical crossover. This occurs at a finite temperature T, which depends on how the energy of the mode, on the scale of the magnetic correlation length $`\xi `$ compares to $`k_BT`$. The low temperature behavior of the system in the different regimes, separated by the crossover temperatures, is then discussed in part VI. Concluding remarks follow in section VII. ## II Self-consistent one-loop approximation to the Kondo lattice model We consider the Kondo lattice model (KLM) as given by a periodic array of Kondo impurities with an average number of conduction electrons per site $`n_c<1`$. In the grand canonical ensemble, the Hamiltonian is written as $$H=\underset{k\sigma }{}\epsilon _kc_{k\sigma }^{}c_{k\sigma }+J\underset{i}{}𝐒_{fi}\underset{\sigma \sigma ^{}}{}c_{i\sigma }^{}𝝉_{\sigma \sigma ^{}}c_{i\sigma ^{}}\mu N_S(\frac{1}{N_S}\underset{k\sigma }{}c_{k\sigma }^{}c_{k\sigma }n_c),$$ (1) in which $`𝐒_{fi}`$ represents the spin ($`S=1/2`$) of the impurities distributed on the sites (in number $`N_S`$) of a periodic lattice; $`c_{k\sigma }^{}`$ is the creation operator of the conduction electron of momentum $`𝐤`$ and spin quantum number $`\sigma `$ characterized by the energy $`ϵ_k=\underset{<i,j>}{}t_{ij}\mathrm{exp}(i𝐤.𝐑_{ij})`$ and the chemical potential $`\mu `$; $`𝝉`$ are the Pauli matrices $`(𝝉^x,𝝉^y,𝝉^z)`$ and $`𝝉^0`$ the unit matrix; J is the antiferromagnetic Kondo interaction $`\left(J>0\right)`$. We use the Abrikosov pseudofermion representation of the spin $`𝐒_{fi}=\underset{\sigma \sigma ^{}}{}f_{i\sigma }^{}𝝉_{\sigma \sigma ^{}}f_{i\sigma ^{}}`$. The projection into the physical subspace is implemented by a local constraint $$Q_i=\underset{\sigma }{}f_{i\sigma }^+f_{i\sigma }1=0.$$ (2) A Lagrange multiplier $`\lambda _i`$ is introduced to enforce the local constraint $`Q_i`$. Since $`[Q_i,H]=0`$, $`\lambda _i`$ is time-independent. In this representation, the partition function of the KLM can be expressed as a functional integral over the coherent states of the fermion fields $$Z=𝒟c_{i\sigma }𝒟f_{i\sigma }𝑑\lambda _i\mathrm{exp}\left[_0^\beta (\tau )𝑑\tau \right],$$ (3) where the Lagrangian $`(\tau )`$ is given by $`(\tau )=_0(\tau )+H_0(\tau )+H_J(\tau )`$ $`_0(\tau )={\displaystyle \underset{i\sigma }{}}c_{i\sigma }^{}_\tau c_{i\sigma }+f_{i\sigma }^{}_\tau f_{i\sigma }`$ $`H_0(\tau )={\displaystyle \underset{k\sigma }{}}ϵ_kc_{k\sigma }^{}c_{k\sigma }\mu N_S\left({\displaystyle \frac{1}{N_S}}{\displaystyle \underset{k\sigma }{}}c_{k\sigma }^{}c_{k\sigma }n_c\right)+{\displaystyle \underset{i}{}}\lambda _iQ_i`$ $`H_J(\tau )=J{\displaystyle \underset{i}{}}𝐒_{fi}𝐒_{ci},`$with $`𝐒_{c_i}=\underset{\sigma \sigma ^{}}{}c_{i\sigma }^{}𝝉_{\sigma \sigma ^{}}c_{i\sigma ^{}}`$. Following , we perform a Hubbard-Stratonovich transformation on the Kondo interaction term $`H_J(\tau )`$ which makes the fields $`\mathrm{\Phi }_i`$, $`\mathrm{\Phi }_i^{}`$ and $`𝝃_{f_i},`$ $`𝝃_{c_i}`$ appear. We get $$Z=𝑑\mathrm{\Phi }_i𝑑\mathrm{\Phi }_i^{}𝑑𝝃_{f_i}𝑑𝝃_{c_i}Z(\mathrm{\Phi }_i,\mathrm{\Phi }_i^{},𝝃_{f_i},𝝃_{c_i}),$$ (4) with $`Z(\mathrm{\Phi }_i,\mathrm{\Phi }_i^{},𝝃_{f_i},𝝃_{c_i})={\displaystyle 𝒟c_{i\sigma }𝒟f_{i\sigma }𝑑\lambda _i\mathrm{exp}\left[_0^\beta ^{}(\tau )𝑑\tau \right]}`$ $`^{}(\tau )=_0(\tau )+H_0(\tau )+H_J^{}(\tau )`$ $`H_J^{}(\tau )={\displaystyle \underset{i\sigma \sigma ^{}}{}}\left(\begin{array}{cc}c_{i\sigma }^{}& f_{i\sigma }^{}\end{array}\right)\left(\begin{array}{cc}J_Si𝝃_{f_i}𝝉_{\sigma \sigma ^{}}& J_C\mathrm{\Phi }_i^{}\tau _{\sigma \sigma ^{}}^0\\ J_C\mathrm{\Phi }_i\tau _{\sigma \sigma ^{}}^0& J_Si𝝃_{c_i}𝝉_{\sigma \sigma ^{}}\end{array}\right)\left(\begin{array}{c}c_{i\sigma ^{}}\\ f_{i\sigma ^{}}\end{array}\right)+J_C{\displaystyle \underset{i}{}}\mathrm{\Phi }_i^{}\mathrm{\Phi }_i+J_S{\displaystyle \underset{i}{}}𝝃_{f_i}.𝝃_{c_i},`$with $`J_S=J/4`$ and $`J_C=J/3`$. The saddle-point solution is obtained when neglecting the space and time dependence of the different fields. Details of the solution are given in Appendix A. It gives rise to the band structure illustrated in Figure 1 with the formation of a Kondo or Abrikosov-Suhl resonance pinned at the Fermi level and split by a hybridization gap defining two bands $`\alpha `$ and $`\beta `$. The saddle-point solution transposes to N=2 the large-N results obtained within the slave-boson mean-field theories (SBMFT). Then we consider the self-consistent one-loop approximation beyond the saddle-point solution in the magnetically-disordered phase $`𝝃_{f_0}=𝝃_{c_0}=0`$. Following Read and Newns , we take advantage of the local U(1) gauge transformation of the lagrangian $`^{}(\tau )`$: $`\mathrm{\Phi }_ir_i\mathrm{exp}(i\theta _i)`$; $`f_if_i^{}\mathrm{exp}(i\theta _i)`$; $`\lambda _i\lambda _i^{^{}}+i\theta _i/\tau `$. We use the radial gauge in which the modulus of both fields $`\mathrm{\Phi }_i`$ and $`\mathrm{\Phi }_i^{}`$ are the radial field $`r_i`$, and their phase $`\theta _i`$ (via its time derivative) is incorporated into the Lagrange multiplier $`\lambda _i`$ which turns out to be a field. Use of the radial instead of the cartesian gauge bypasses the familiar complications of infrared divergences associated with unphysical Goldstone bosons. We let the fields fluctuate away from their saddle-point values : $`r_i=r_0+\delta r_i`$, $`\lambda _i=\lambda _0+\delta \lambda _i`$, $`\xi _{𝐟_𝐢}=\delta 𝝃_{𝐟_𝐢}`$ and $`\xi _{𝐜_𝐢}=\delta 𝝃_{𝐜_𝐢}`$. After integrating out the Grassmann variables in the partition function in Equation (3), we express $`Z`$ as a functional integral over the bosonic fields $`(r_i,\lambda _i,𝝃_{f_i},𝝃_{c_i})`$ $$Z=𝒟r_i𝒟\lambda _i𝒟𝝃_{f_i}𝒟𝝃_{c_i}\mathrm{exp}[S_{eff}(r_i,\lambda _i,𝝃_{f_i},𝝃_{c_i})],$$ (5) where the effective action is $`S_{eff}(r,\lambda ,𝝃_f,𝝃_c)`$ $`=`$ $`{\displaystyle \underset{k,i\omega _n}{}}LnDet𝐆^1(𝐤,i\omega _n,r_i,\lambda _i,𝝃_{f_i},𝝃_{c_i})`$ (6) $`+`$ $`\beta [J_C{\displaystyle \underset{i}{}}r_i^2+J_S{\displaystyle \underset{i}{}}𝝃_{f_i}𝝃_{c_i}+N_S(\mu n_c\lambda _0)],`$ (7) with : $`\left[𝐆^1(i\omega _n,r,\lambda ,𝝃_f,𝝃_c)\right]_{ij}^{\sigma \sigma ^{}}=`$ (8) $`\left(\begin{array}{cc}[(i\omega _n\mu )\delta _{ij}t_{ij}]\delta _{\sigma \sigma ^{}}J_Si𝝃_{f_i}.𝝉_{\sigma \sigma ^{}}\delta _{ij}& (\sigma _0+J_C\delta r_i)\delta _{\sigma \sigma ^{}}\delta _{ij}\\ (\sigma _0+J_C\delta r_i)\delta _{\sigma \sigma ^{}}\delta _{ij}& [i\omega _n+\epsilon _f+\delta \lambda _i]\delta _{\sigma \sigma ^{}}\delta _{ij}J_Si𝝃_{c_i}.𝝉_{\sigma \sigma ^{}}\delta _{ij}\end{array}\right).`$ (11) As in the conventional approaches to critical phenomena, we have integrated out the fermions and thereby reduced the problem to the study of an effective bosonic theory describing fluctuations of the ordering fields. But contrary to Hertz and Millis , who considered a single field corresponding to the magnetization density of one type of electron, our approach, starting from a different microscopic model, extends the description to the case of several fields i.e. the Kondo fields $`(r_i,\lambda _i)`$ on the one hand, and the $`f`$\- and $`c`$\- magnetization densities on the other. We then propose to solve this field theory in a variational way. We replace the full expression $`S_{eff}(r_i,\lambda _i,𝝃_{f_i},𝝃_{c_i})S_{eff}^{(0)}`$ (where $`S_{eff}^{(0)}`$ and $`F^{(0)}`$ are the zeroth order effective action and the related free energy) by a trial quadratic form $`\mathrm{\Delta }\stackrel{~}{S}_{eff}^{(2)}(r,\lambda ,𝝃_f,𝝃_c)=S_{eff}^{(2)}(r_i,\lambda _i,𝝃_{f_i},𝝃_{c_i})S_{eff}^{(0)}`$ in the fields $`\delta r_i`$ , $`\delta \lambda _i`$, $`\delta 𝝃_{f_i}`$ and $`\delta 𝝃_{c_i}`$ $`\mathrm{\Delta }\stackrel{~}{S}_{eff}^{(2)}(r,\lambda ,𝝃_f,𝝃_c)`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{𝐪,i\omega _\nu }{}}[\left(\begin{array}{cc}\delta r_{q,\nu }& \delta \lambda _{q,\nu }\end{array}\right)\stackrel{~}{𝐃}_C^1(𝐪,i\omega _\nu )\left(\begin{array}{c}\delta r_{q,\nu }\\ \delta \lambda _{q,\nu }\end{array}\right)`$ (15) $`+`$ $`\left(\begin{array}{cc}\delta \xi _{fq,\nu }^z& \delta \xi _{cq,\nu }^z\end{array}\right)\stackrel{~}{𝐃}_S^1(𝐪,i\omega _\nu )\left(\begin{array}{c}\delta \$xi_{fq,\nu }^z\\ \delta \xi _{cq,\nu }^z\end{array}\right)`$ (19) $`+`$ $`\left(\begin{array}{cc}\delta \xi _{fq,\nu }^+& \delta \xi _{cq,\nu }^+\end{array}\right)\stackrel{~}{𝐃}_S^1(𝐪,i\omega _\nu )\left(\begin{array}{c}\delta \xi _{fq,\nu }^{}\\ \delta \xi _{cq,\nu }^{}\end{array}\right)`$ (23) $`+`$ $`\left(\begin{array}{cc}\delta \xi _{fq,\nu }^{}& \delta \xi _{cq,\nu }^{}\end{array}\right)\stackrel{~}{𝐃}_S^1(𝐪,i\omega _\nu )\left(\begin{array}{c}\delta \xi _{fq,\nu }^+\\ \delta \xi _{cq,\nu }^+\end{array}\right)].`$ (27) where the indices $`q,\nu `$ stand for $`(𝐪,i\omega _\nu )`$. The elements of the matrices $`\stackrel{~}{𝐃}_C^1(𝐪,i\omega _\nu )`$, $`\stackrel{~}{𝐃}_S^1(𝐪,i\omega _\nu )`$ and $`\stackrel{~}{𝐃}_S^1(𝐪,i\omega _\nu )`$ are determined variationally by minimizing the free energy or, more precisely, the upper bound to the free energy according to the Feynman variational principle $`F\stackrel{~}{F}_{eff}^{(2)}+1/\beta S_{eff}(r,\lambda ,𝝃_f,𝝃_c)\stackrel{~}{S}_{eff}^{(2)}(r,\lambda ,𝝃_f,𝝃_c)`$ where $`\stackrel{~}{F}_{eff}^{(2)}`$ represents the free energy related to $`\stackrel{~}{S}_{eff}^{(2)}(r,\lambda ,𝝃_f,𝝃_c)`$ and $`A`$ means the average of the quantity $`A(r_i,\lambda _i,𝝃_{f_i},𝝃_{c_i})`$ over the distribution of fields $`\mathrm{exp}[\mathrm{\Delta }\stackrel{~}{S}_{eff}^{(2)}(r,\lambda ,𝝃_f,𝝃_c)]`$ i.e. $`A={\displaystyle \frac{(𝒟r_i𝒟\lambda _i𝒟𝝃_{f_i}𝒟𝝃_{c_i}A(r_i,\lambda _i,𝝃_{f_i},𝝃_{c_i})\mathrm{exp}[\mathrm{\Delta }\stackrel{~}{S}_{eff}^{(2)}(r,\lambda ,𝝃_f,𝝃_c)])}{(𝒟r_i𝒟\lambda _i𝒟𝝃_{f_i}𝒟𝝃_{c_i})\mathrm{exp}[\mathrm{\Delta }\stackrel{~}{S}_{eff}^{(2)}(r,\lambda ,𝝃_f,𝝃_c)])}}.`$ After minimization, one can show that $`[𝐃_C^1(𝐪,i\omega _\nu )]_{r\lambda }`$ is equal to $`(1/2)^2S_{eff}(r_i,\lambda _i,𝝃_{f_i},𝝃_{c_i})/(r_{q,\nu }\lambda _{q,\nu })`$ averaged over the mentioned distribution of fields (and equivalent expressions for the other elements). Explicitly we obtain, $$\stackrel{~}{𝐃}_C^1(𝐪,i\omega _\nu )=\left(\begin{array}{cc}J_C[1J_C(\overline{\phi }_2(𝐪,i\omega _\nu )+\overline{\phi }_m(𝐪,i\omega _\nu ))]& J_C\overline{\phi }_1(𝐪,i\omega _\nu )\\ J_C\overline{\phi }_1(𝐪,i\omega _\nu )& \overline{\phi }_{ff}(𝐪,i\omega _\nu )\end{array}\right)$$ (28) $`\stackrel{~}{𝐃}_S^1(𝐪,i\omega _\nu )=\left(\begin{array}{cc}J_S^2\phi _{cc}^{}(𝐪,i\omega _\nu )& J_S[1+J_S\phi _{fc}^{}(𝐪,i\omega _\nu )]\\ J_S[1+J_S\phi _{cf}^{}(𝐪,i\omega _\nu )]& J_S^2\phi _{ff}^{}(𝐪,i\omega _\nu )\end{array}\right),`$and equivalent expression for the transverse spin part $`\stackrel{~}{𝐃}_S^1(𝐪,i\omega _\nu )`$. The different bubbles $`\phi (𝐪,i\omega _\nu )`$ in the latter expressions represent the susceptibilities $`\phi (𝐪,i\omega _\nu )`$ (cf. appendix B) averaged over the distribution of fields $`\mathrm{exp}[\mathrm{\Delta }\stackrel{~}{S}_{eff}^{(2)}(r,\lambda ,𝝃_f,𝝃_c)]`$. At that point, the problem is highly complex and cannot be solved exactly. Therefore, we introduce a local and instantaneous approximation which is equivalent to averaging over gaussian distributions of local fields $`(r_i,\lambda _i,𝝃_{f_i},𝝃_{c_i})`$ with variances $`\sigma _r`$, $`\sigma _\lambda `$, $`\sigma _f`$ and $`\sigma _c`$ defined as $$\sigma _f^2=3\underset{𝐪,i\omega _\nu }{}[\stackrel{~}{𝐃}_S^{}(𝐪,i\omega _\nu )_{ff}],$$ (29) and similar expressions for the three other variances. Eq.(29) is obtained from Eq.(15) by letting the summation over $`(q,i\omega _\nu )`$ act on the matrix elements only, hence defining the local and instantaneous fluctuations of the various degrees of freedom. One expects that at low temperatures, the fluctuations of $`r_i`$, $`\lambda _i`$ and $`𝝃_{c_i}`$ only bring $`T^2`$ corrections to their saddle-point values. Those corrections are negligible compared to the $`T`$-dependence brought by the fluctuations of $`𝝃_{f_i}`$ and will not be considered here. So we will neglect in the following the fluctuations of $`r_i`$, $`\lambda _i`$ and $`𝝃_{c_i}`$ and focus on the fluctuations of $`𝝃_{f_i}`$ only. Then $`\phi (𝐪,i\omega _\nu )`$ is simply given by $$\phi (𝐪,i\omega _\nu )=\frac{(𝒟\xi _{f_i}\phi (𝐪,i\omega _\nu ,r_0,\lambda _0,\xi _{f_i},\xi _{c_0})\mathrm{exp}[\xi _{f_i}^2/(2\sigma _f^2)])}{(𝒟\xi _{f_i}\mathrm{exp}[\xi _{f_i}^2/(2\sigma _f^2)])}.$$ (30) Eqs. (29-30) are the basic equations of the paper from which the dynamical spin susceptibility is extracted and the critical phenomena around the quantum critical point (QCP) are studied. Let us note that, when the averages acting on the various bubbles $`\phi (𝐪,i\omega _\nu )`$ in Eq.(28) are omitted, one recovers the standard results of the random phase approximation. As pointed out by Lonzarich and Taillefer , the corresponding self-consistent renormalized spin fluctuation SCR-SF procedure can be analyzed within the Ginzburg-Landau formalism in which the local free energy is expanded in terms of a small and slowly varying varying order parameter $`𝐦(𝐫)`$ as $`f(𝐫)=f_0+\frac{1}{2}a\left|𝐦(𝐫)\right|^2+\frac{1}{4}b\left|𝐦(𝐫)\right|^4+\frac{1}{2}c\underset{i}{}\left|m^i(𝐫)\right|^2+O(6)`$ and the following approximation is made $`\left|𝐦(𝐫)\right|^4=\left|𝐦(𝐫)\right|^2\left|𝐦(𝐫)\right|^2`$. Hence $`\left|𝐦(𝐫)\right|^2`$ is evaluated via the fluctuation-dissipation theorem. ## III <br>Dynamical spin susceptibility Let us now derive the expression of the dynamical spin susceptibility. For that purpose, we study the linear response $`M_f`$ to the source-term $`2𝐒_f.𝐁`$ (we consider $`𝐁`$ colinear to the $`𝐳`$ -axis). The effect on the partition function expressed in Equation (4 ) is to change the $`H_J^{}(\tau )`$ to $`H_J^B(\tau )`$ $`H_J^B(\tau )=`$ $`{\displaystyle \underset{i\sigma \sigma ^{}}{}}\left(\begin{array}{cc}c_{i\sigma }^{}& f_{i\sigma }^{}\end{array}\right)\left(\begin{array}{cc}J_Si𝝃_{f_i}𝝉_{\sigma \sigma ^{}}& J_C\mathrm{\Phi }_i^{}\tau _{\sigma \sigma ^{}}^0\\ J_C\mathrm{\Phi }_i\tau _{\sigma \sigma ^{}}^0& \underset{\alpha =x,y,z}{}(J_Si\xi _{c_i}^\alpha B\delta _{\alpha z}).\tau _{\sigma \sigma ^{}}^\alpha \end{array}\right)\left(\begin{array}{c}c_{i\sigma ^{}}\\ f_{i\sigma ^{}}\end{array}\right)`$ (36) $`+J_C{\displaystyle \underset{i}{}}\mathrm{\Phi }_i^{}\mathrm{\Phi }_i+J_S{\displaystyle \underset{i}{}}𝝃_{f_i}.𝝃_{c_i}.`$ (37) Introducing the change of variables $`\xi _{c_i}^\alpha =\xi _{c_i}^\alpha iB/J_S`$, we connect the $`f`$\- magnetization and the $`ff`$ \- dynamical spin susceptibility to the Hubbard Stratonovich field $`𝝃_{f_i}`$ $`M_f^z={\displaystyle \frac{1}{\beta }}{\displaystyle \frac{LnZ}{B_z}}=i\xi _{f_i}^z`$ $$\chi _{ff}^{\alpha \beta }=\frac{1}{\beta }\frac{^2LnZ}{B^\alpha B^\beta }=\xi _{f_i}^\alpha \xi _{f_i}^\beta +\xi _{f_i}^\alpha \xi _{f_i}^\beta .$$ (38) Using the expression (28) for the boson propagator $`𝐃_S^1(𝐪)`$, we get for the Matsubara spin susceptibility $$\chi _{ff}(𝐪,i\omega _\nu )=\frac{\phi _{ff}(𝐪,i\omega _\nu )}{1J_S^2[\phi _{ff}(𝐪,i\omega _\nu )\phi _{cc}(𝐪,i\omega _\nu )\phi _{fc}(𝐪,i\omega _\nu )^2\frac{2}{J_S}\phi _{fc}(𝐪,i\omega _\nu )]}.$$ (39) The diagrammatic representation of Equation (39) is represented in Fig. 2 when the averaging is omitted. In the low frequency limit, one can easily check that the dynamical spin susceptibility may be written in terms of intra- and inter-band suceptibilities corresponding respectively to particle-hole pair excitations within the lower $`\alpha `$ band and from the lower $`\alpha `$ to the upper $`\beta `$ band. $$\chi _{ff}(𝐪,i\omega _\nu )=\frac{\overline{\chi }_{\alpha \alpha }(𝐪,i\omega _\nu )+\overline{\chi }_{\alpha \beta }(𝐪,i\omega _\nu )}{1J_S^2\chi _{\alpha \alpha }(𝐪,i\omega _\nu )\overline{\chi }_{\alpha \beta }(𝐪,i\omega _\nu )}.$$ (40) The expressions of the susceptibilities $`\chi _{\alpha \alpha }(𝐪,i\omega _\nu )`$ and $`\overline{\chi }_{\alpha \beta }(𝐪,i\omega _\nu )`$ corresponding to intra- and inter-band particle-hole excitations are given in Appendix B. The latter expression is reminiscent of the behaviour proposed by Bernhoeft and Lonzarich to explain the neutron scattering observed in $`UPt_3`$ with the existence of both a ”slow” and a ”fast” component in $`\chi ^\mathrm{"}(𝐪,\omega )/\omega `$ due to spin-orbit coupling. Also in a phenomenological way, the same type of feature has been suggested in the duality model developed by Kuramoto and Miyake . To our knowledge, the proposed approach provides the first microscopic derivation from the Kondo lattice model of such a behaviour. Expanding the various susceptibilities $`\phi (𝐪,i\omega _\nu ,r_0,\lambda _0,\xi _{f_i},\xi _{c_0})`$ up to the second order in $`\xi _{f_i}`$, and making use of Eq. (30), one can draw from Eq. (40) the following expression of the dynamical spin susceptibility (taking the analytical continuation $`i\omega _\nu \omega +i\delta )`$ $$\chi _{ff}(𝐪,\omega )=\frac{\chi _{\alpha \alpha }(𝐪,\omega )+\overline{\chi }_{\alpha \beta }(𝐪,\omega )}{1J_S^2\chi _{\alpha \alpha }(𝐪,\omega )\overline{\chi }_{\alpha \beta }(𝐪,\omega )+\lambda \sigma _f^2}.$$ (41) where $`\lambda `$ is a constant which can be evaluated from the expansion of $`\chi _{\alpha \alpha }(𝐪,i\omega _\nu )`$ and $`\overline{\chi }_{\alpha \beta }(𝐪,i\omega _\nu )`$ up to the second order in $`\sigma _f^2`$. The variance $`\sigma _f^2`$ defined by Eq. (29) can as well be expressed as a function of $`\chi _{ff}(𝐪,i\omega _\nu )`$ $$\sigma _f^2=\underset{𝐪,i\omega _\nu }{}\chi _{ff}(𝐪,i\omega _\nu )=\frac{3}{\pi }\underset{𝐪}{}_0^+\mathrm{}\mathrm{coth}\frac{\beta \omega }{2}\chi _{ff}^\mathrm{"}(𝐪,\omega )𝑑\omega .$$ (42) The two last equations Eqs. (41-42) provide a self-consistent determination of $`\sigma _f^2`$ and hence of the spin susceptibility. We will successively present now (i) in section IV, the result for the dynamical spin susceptibility in the random phase approximation (RPA) giving information on the nature of the collective mode that is involved; (ii) then in section V, the implications of that mode in the critical phenomena around the quantum critical point. ## IV RPA-dynamical spin susceptibility near the antiferromagnetic wave vector : discussion on the nature of the collective mode As pointed out at the end of section II, the random phase approximation (RPA) corresponds to the absence of any averaging in the different equations. Thereby the expression of the dynamical spin susceptibility at the RPA level is given by Eq. (41) taking $`\sigma _f^2=0`$. We now discuss the $`𝐪`$\- and $`\omega `$\- dependence of the RPA dynamical spin susceptibility around the antiferromagnetic wave-vector $`𝐐`$. The bare intraband susceptibility $`\chi _{_{\alpha \alpha }}(𝐐+𝐪,\omega )`$ is well approximated at $`\left|𝐪\right|\left|𝐐\right|`$ by a lorentzian $`\chi _{\alpha \alpha }(𝐐+𝐪,\omega )=\rho _{\alpha \alpha }/(1i\omega /\mathrm{\Gamma }_0+bq^2)`$ where $`\rho _{\alpha \alpha }=\chi _{\alpha \alpha }^{^{}}(𝐐,0)`$ and $`\mathrm{\Gamma }_0`$ is the relaxation rate of order $`\left|y_F\right|=T_K`$. As expected for the antiferromagnetic case, the relaxation rate remains finite when $`q`$ goes to zero reflecting the fact that the fluctuations of the order parameter are not conserved. The bare interband susceptibility $`\overline{\chi }_{\alpha \beta }(𝐐+𝐪,\omega )`$ can be considered as purely real and frequency independent equal to $`\rho _{\alpha \beta }`$ in the low frequency limit. Fig.3 gives the continuum of the intraband and interband particle-hole excitations $`\chi _{\alpha \alpha }^{\prime \prime }0`$ and $`\overline{\chi }_{\alpha \beta }^{\prime \prime }0`$. Due to the presence of the hybridization gap in the density of states, the latter continuum displays a gap ranging from $`2\sigma _0`$, the value of the direct gap for the wave-vector $`\mathrm{𝟎}`$, to $`2\left|y_F\right|`$, the value of the indirect gap at $`𝐐`$. The RPA dynamical spin susceptibility is given by $$\chi _{ff}^{{}_{RPA}{}^{}\text{ }\prime \prime }(𝐐+𝐪,\omega )=\omega \frac{\chi _{ff}^{}(𝐐+𝐪)\mathrm{\Gamma }(q)}{\omega ^2+\mathrm{\Gamma }(q)^2}$$ (43) with $`\chi _{ff}^{}(𝐐+𝐪)={\displaystyle \frac{\rho _{\alpha \alpha }+\rho _{\alpha \beta }}{(1I_q)}}`$ $`\mathrm{\Gamma }(q)=\mathrm{\Gamma }_0(1I_q)`$where $`I_q=Ibq^2`$ and $`I=J_S^2\rho _{\alpha \alpha }\rho _{\alpha \beta }`$ (of order $`1`$ near the antiferromagnetic instability). The bulk staggered susceptibility $`\chi _{ff}^{^{}}(𝐐)`$ diverges at $`I=1`$. The frequency dependence of $`\chi _{ff}^{{}_{RPA}{}^{}\text{ }\prime \prime }(𝐐+𝐪,\omega )`$ is a lorentzian with a vanishing relaxation rate $`\mathrm{\Gamma }(0)`$ at the antiferromagnetic transition. The corresponding excitation can be analyzed as an antiferromagnetic paramagnon mode that softens at the magnetic transition. Eq. (43) provides us with the dispersion of that mode. The corresponding dynamical exponent $`z`$ is found to be equal to $`2`$ due to the overdamping of the mode when enters the continuum of the $`\alpha `$-$`\alpha `$ particle-hole excitations. We will show how the dynamic exponent $`z`$ strongly affects the static critical behavior. This is due to the fact that, contrary to the $`T>0`$ phase transitions, statics and dynamics are intimately linked in quantum ($`T=0`$) phase transitions. We turn now to the resulting quantum-classical crossover occuring at finite temperature. ## V Critical phenomena around the antiferromagnetic quantum critical point (AF-QCP) The parameter which controls the temperature dependence of the thermodynamical variables near the AF-QCP is $`\sigma _f^2`$ which expresses the thermal local fluctuations of the staggered magnetization. Using Eqs. (42) and (43) $$\sigma _f^2=\underset{𝐪}{}S(𝐐+𝐪)S(𝐐+𝐪)q^{d1}𝑑q,$$ (44) where $`S(𝐐+𝐪)`$ is the static form factor $$S(𝐐+𝐪)=\frac{3}{\pi }_0^{\mathrm{}}\mathrm{coth}(\beta \omega /2)\chi _{ff}^{\text{ }\prime \prime }(𝐐+𝐪,\omega )𝑑\omega .$$ (45) Depending on the temperature-scale considered, quantum or classical behavior is observed. In general, the thermal fluctuations $`\sigma _f^2`$ of the magnetization should depend on the total self-consistent renormalized dynamical spin susceptibility $`\chi _{ff}\mathrm{"}(𝐐+𝐪,\omega )`$ as defined in Eq.(40). However, in practice, we will content with its truncated form $`\chi _{ff}^{{}_{RPA}{}^{}\text{ }\prime \prime }(𝐐+𝐪,\omega )`$ derived at the RPA level. The temperature dependence of the static form factor $`S(𝐐+𝐪)`$ is given by $$S(𝐐+𝐪)=\frac{6T}{\pi }\chi _{ff}^{}(𝐐+𝐪)\mathrm{tan}^1(\frac{T}{\mathrm{\Gamma }(q)}).$$ (46) A first cross-over temperature $`T_I=2\mathrm{\Gamma }_0(1I)`$ appears which separates the quantum from the classical regime (i) $`T<T_I`$ : quantum regime. One can show that the static form factor $`S(𝐐+𝐪)`$ exhibits a quadratic temperature dependence $$S(𝐐+𝐪)=\frac{6}{\pi }\frac{\chi _{ff}^{}(𝐐+𝐪)}{\mathrm{\Gamma }_0(1I_q)}T^2,$$ (47) The thermal local fluctuations $`\sigma _f^2`$ of the magnetization are also quadratic as a function of the temperature. They can be easily evaluated from Eq.(46). Taking advantage of the fact that the integrand $`S(𝐐+𝐪)q^{d1}`$ is peaked at $`q_1=\sqrt{(1I)/b}`$, the temperature dependence of $`\sigma _f^2`$ is also found to be quadratic $$\sigma _f{}_{}{}^{2}=\frac{3}{2\pi }\frac{\chi _{ff}^{}(𝐐)}{\mathrm{\Gamma }_0(1I)}T^2,$$ (48) The physics is quantum in the sense that the fluctuations on the scale of the magnetic correlation length $`\xi `$ have energy much greater that the temperature $`k_BT`$. (ii) $`T>T_I`$ : classical regime. In this regime, the static form factor $`S(𝐐+𝐪)`$ shows a linear temperature dependence. As long as $`q`$ is smaller than the thermal cut-off $`q^{}`$ defined by $`\omega _{\mathrm{max}}(q^{})=T`$, $`S(𝐐+𝐪)`$ takes the following form $$S(𝐐+𝐪)=3\chi _{ff}^{}(𝐐+𝐪)T.$$ (49) The thermal local fluctuations $`\sigma _f^2`$ of the magnetization are deduced from Eq. (46). The main contribution arises from the integration over $`[q_1,q^{}]`$. The result depends on the dimensionality $`d`$. One finds $`\sigma _f{}_{}{}^{2}T^{3/2}`$ at $`d=3`$ and $`\sigma _f^2T\mathrm{ln}T`$ at $`d=2`$. This classical regime corresponds to the case where the energy of the mode on the scale of $`\xi `$ becomes less than $`k_BT`$. Table I recapitulates the expressions that we get for the thermal local fluctuations $`\sigma _f^2`$ of the staggered magnetization depending on the temperature scale and on the dimensionality. Following Eq. (41), the staggered spin susceptibility is given by $$\chi _Q^{^{}}=\frac{\chi _{ff}^0^{}(Q)}{1I+\lambda \sigma _f^2},$$ (50) where $`\chi _{ff}^0^{}(Q)=\rho _{\alpha \alpha }+\rho _{\alpha \beta }`$. Incorporating the values of the thermal local fluctuations $`\sigma _f^2`$ reported in Table I, we are able to identify a second cross-over temperature $`T_{II\text{ }}`$. Above $`T_{II\text{ }}`$, $`\sigma _f^2`$ becomes larger than $`(1I)`$ and all the physical quantities are controlled by the temperature only. $`T_{II\text{ }}(1I)^{2/3}`$ at $`d=3`$ and $`T_{II\text{ }}(1I)`$ at $`d=2`$. As expected in the two-dimensional case, the $`T=0`$ transition is at its upper critical dimension since $`z+d=4`$ in the antiferromagnetic case characterized by a dynamic exponent $`z=2`$. That corresponds to the marginal case for which the intermediate regime $`T_I<T<T_{II}`$ is squeezed out of existence. The results for the $`d=3`$ and $`d=2`$ cases are respectively summarized in Figures 4 and 5 which picture the different regimes of behaviors reached depending on the values of the temperature $`T`$ and of the control parameter $`I=J^2\rho _{\alpha \alpha }\rho _{\alpha \beta }`$. When $`I>1`$, long-range antiferromagnetic order occurs when $`T`$ is smaller than the Néel temperature $`T_N`$. $`T_N(I1)^{2/3}`$ at $`d=3`$ and $`T_N(I1)`$ at $`d=2`$. ## VI Discussion The low temperature physics is dominated by the presence of an antiferromagnetic paramagnon mode that softens at the magnetic transition. The results depend crucially on the value of $`d+z`$ where $`d`$ is the spatial dimension and $`z`$ is the dynamic exponent associated to that mode. Due to the overdamping of the mode when it enters the continuum of intraband particle-hole excitations, the dynamic exponent $`z`$ is found to be equal to $`2`$ in the antiferromagnetic case. As far as the dimensionality is concerned, there is no doubt about the dimensionality associated with the heavy quasiparticles in those systems that is clearly of 3. However the question about the dimensionality of the critical magnetic fluctuations which the quasiparticles are coupled to, is still an open problem. Up to now, it has been assumed that the critical fluctuations of the QPT are dominated by the three-dimensional antiferromagnetic correlations. This would impose a description by a quantum critical theory with $`d=3`$, $`z=2`$. Recently Rosch et al proposed, on the basis of the neutron scattering data in $`CeCu_{6x}Au_x`$, that the critical magnetic fluctuations are two-dimensional which leads to a QCP with $`d=2`$, $`z=2`$. Let us summarize the different regimes of behaviors that we get in each of those two cases ($`z=2`$ , $`d=3`$ or $`2`$) depending on the values of the temperature $`T`$ and of the control parameter $`I=J^2\rho _{\alpha \alpha }\rho _{\alpha \beta }`$. As one can see, the results that we obtain are very similar to those established by Hertz and Millis using renormalization group approaches in the spin-fluctuation theory. $`(i)`$ Case $`d=3`$. A long-range antiferromagnetic phase occurs when $`I>1`$ below the Néel temperature $`T_N(I1)^{2/3}`$. A first cross-over temperature $`T_I(1I)`$ separates the quantum from the classical regime. In the regime $`I`$ $`(I<1)`$ and $`T<T_I)`$, the physics is quantum in the sense that the energy of the relevant mode on the scale of $`\xi `$ is greater than the temperature $`k_BT`$. Since $`d+z>4`$, the $`T=0`$ phase transition is above its upper critical dimension and the various physical quantities depend upon the parameter $`I`$. The magnetic correlation length diverges at the magnetic transition $`\xi 1/\sqrt{(1I)}`$ and the staggered spin susceptibility behaves as $`\chi _Q^{^{}}=\chi _{ff}^0^{}(Q)/(1I+aT^2)`$. Regime $`II`$ and $`III`$ are both classical regimes characterized by large thermal effects since the fluctuations on the scale of $`\xi `$ have energy much smaller than $`k_BT`$. In Regime $`II`$ $`(T_I<T<T_{II}`$ with $`T_{II}(1I)^{2/3})`$, the magnetic correlation length $`\xi 1/\sqrt{(1I)}`$ is still controlled by $`(1I)`$ even though modes at the scale of $`\xi `$ have energies less than $`k_BT`$. On the contrary, the staggered spin susceptibility is sensitive to the thermal fluctuations $`\chi _Q^{^{}}=\chi _{ff}^0^{}(Q)/(1I+aT^{3/2})`$. In Regime $`III`$ $`(T>T_{II})`$, the thermal dependence of physical quantities becomes universal and both $`\xi `$ and $`\chi _Q^{^{}}`$ are controlled by the temperature : $`\xi 1/T^{3/4}`$ and $`\chi _Q^{^{}}=\chi _{ff}^0^{}(Q)/T^{3/2}.`$ $`(ii)`$ Case $`d=2`$. Then, since $`z=2`$, $`z+d=4`$ and the $`T=0`$ phase transition is at its upper critical dimension. The physics is qualitatively similar to the case $`d=3`$ with stronger fluctuation effects particularly in the classical regime. A long-range antiferromagnetic phase occurs when $`I>1`$ below the Néel temperature $`T_N(I1)`$. For $`d=2`$, the two cross-over temperatures $`T_I`$ and $`T_{II}`$ coincide so that Regime $`II`$ of Figure 4 is squeezed out of existence. Regime $`I`$ ($`T<T_I`$ with $`T_I(1I))`$ is the quantum regime in which the thermal fluctuations are negligible : $`\xi 1/\sqrt{(1I)}`$ and $`\chi _Q^{^{}}=\chi _{ff}^0^{}(Q)/(1I+aT^2)`$. Regime $`III`$ $`(T>T_I)`$ is the unique classical regime characterized by very strong fluctuation effects since $`k_BT`$ is larger than the energy of the relevant mode on the scale of $`\xi `$ . Both the magnetic correlation length and the staggered spin susceptibility are then controlled by the temperature : $`\xi 1/\sqrt{T\mathrm{ln}T}`$ and $`\chi _Q^{^{}}=\chi _{ff}^0^{}(Q)/(T\mathrm{ln}T)`$. ## VII Concluding remarks We considered the $`S=1/2`$ Kondo lattice model in a self-consistent one-loop approximation starting from a generalized Hubbard-Stratonovich transformation of the Kondo interaction term. The model exhibits a zero temperature magnetic phase transition at a critical value of the Kondo coupling. The transition is usually antiferromagnetic but it may be incommensurate depending on the bandstructure considered. The low temperature physics is controlled by a collective mode that softens at the antiferromagnetic transition with a dynamic exponent $`z`$ equal to $`2`$. A quantum-classical crossover occurs at a temperature $`T_I`$ related to the characteristic energy-scale of that mode. Heavy-fermion systems are usually believed to be $`3`$-dimensional. However since some recent inelastic neutron scattering experiments performed in $`CeCu_{6x}Au_x`$ show that the critical magnetic fluctuations which the quasiparticles are coupled to, are $`2`$-dimensional, we both considered the cases $`d=2`$ and $`d=3`$ with $`z=2`$. The low-temperature behavior of the system is deduced, with predictions for the temperature dependence of the physical quantities such as the magnetic correlation length $`\xi `$ and the staggered susceptibility $`\chi _Q^{}`$. A second crossover temperature $`T_{II}`$ appears above which the thermal dependence of $`\xi `$ and $`\chi _Q^{}`$ becomes universal and is uniquely controlled by the temperature. It would be very interesting now to study the temperature behavior of other physical quantities as the specific heat or the transport properties. A number of significant issues remain to be addressed. Among others, we will mention the possibility of getting a different value of the dynamic exponent for instance $`z=1`$ if nesting effects are considered corresponding to an absence of damping of the mode when located in the gap of excitations around the antiferromagnetic vector. Disorder is expected to play a crucial role in this problem. In that direction, some recent works pointed out the formation of some ”hot lines”, i.e., points at the Fermi surface linked by the magnetic Q vector. On the hot lines, the quasiparticle scattering rate is linear in temperature while, away from the hot lines, it acquires the standard Fermi liquid form in $`T^2`$. At low temperature, the ”cold” regions are shown to short-circuit the scattering near those lines, and induce a $`T^2`$ behavior of the resistivity. The disorder is found to emphasize the contribution of the hot lines, eventually leading to a non-Fermi liquid behavior. It would be worthy to study the effects of the ”hot lines” in presence of disorder within the critical phenomena description presented in this paper. We would like to thank Nick Bernhoeft, Piers Coleman, Jacques Flouquet, Patrick A. Lee and Gilbert Lonzarich for very useful discussions on this work. Also at the Centre National de la Recherche Scientifique (CNRS) ### A Appendix A: Saddle-Point The saddle-point solution is obtained from Eq. (4) for space and time independent fields $`\mathrm{\Phi }_0`$, $`\lambda _0`$, $`\xi _{f_0}`$ and $`\xi _{c_0}`$. In the magnetically-disordered regime ($`\xi _{f_0}=\xi _{c_0}=0)`$, it leads to renormalized bands $`\alpha `$ and $`\beta `$ as schematized in Figure 1. Noting $`\sigma _0^{()}=J_C\mathrm{\Phi }_0^{()}`$ and $`\epsilon _f=\lambda _0`$, $`\alpha _{k\sigma }^{}|0>`$ and $`\beta _{k\sigma }^{}|0>`$ are the eigenstates of $$𝐆_0^{1\sigma }(𝐤,\tau )=\left(\begin{array}{cc}_\tau +\epsilon _𝐤& \sigma _\mathrm{𝟎}^{}\\ \sigma _\mathrm{𝟎}& _\tau +\epsilon _𝐟\end{array}\right),$$ (51) with respectively the eigenenergies $`\left(_\tau +E_k^{}\right)`$ and $`\left(_\tau +E_k^+\right)`$. In the notations: $`x_k=\epsilon _k\epsilon _f`$, $`y_k^\pm =E_k^\pm \epsilon _f`$ and $`\mathrm{\Delta }_k=\sqrt{x_k^2+4\sigma _0^2}`$, we get $$y_k^\pm =\left(x_k\pm \mathrm{\Delta }_k\right)/2.$$ (52) Let us note $`U_{k\sigma }^{}`$ the matrix transforming the initial basis $`(c_{k\sigma }^{}`$ $`f_{k\sigma }^{})`$ to the eigenbasis $`(\alpha _{k\sigma }^{}`$ $`\beta _{k\sigma }^{})`$. The Hamiltonian being hermitian, the matrix $`U_{k\sigma }`$ is unitary : $`U_{k\sigma }U_{k\sigma }^{}=U_{k\sigma }^{}U_{k\sigma }=1`$. In the notation $`U_{k\sigma }^{}=\left(\begin{array}{cc}v_k& u_k\\ u_k& v_k\end{array}\right)`$, we have $$u_k=\frac{\sigma _0/y_k^{}}{\sqrt{1+(\sigma _0/y_k^{})^2}}=\frac{1}{2}\left[1+\frac{x_k}{\mathrm{\Delta }_k}\right]$$ (53) $`v_k={\displaystyle \frac{1}{\sqrt{1+(\sigma _0/y_k^{})^2}}}={\displaystyle \frac{1}{2}}\left[1{\displaystyle \frac{x_k}{\mathrm{\Delta }_k}}\right].`$ The saddle-point equations together with the conservation of the number of conduction electrons are written as $$\sigma _0=\frac{1}{N_S}J_C\underset{k\sigma }{}u_kv_kn_F(E_k^{})$$ (54) $`1={\displaystyle \frac{1}{N_S}}{\displaystyle \underset{k\sigma }{}}u_k^2n_F(E_k^{})`$ $`n_c={\displaystyle \frac{1}{N_S}}{\displaystyle \underset{k\sigma }{}}v_k^2n_F(E_k^{}).`$Their resolution at zero temperature and for $`n_c`$ close to $`1`$ leads to $$\left|y_F\right|=D\mathrm{exp}\left[2/\left(\rho _0J_C\right)\right]$$ (55) $`2\rho _0\sigma _0^2/\left|y_F\right|=1`$ $`\mu =0,`$where $`y_F=\mu \epsilon _F`$ and $`\rho _0`$ is the bare density of states of conduction electrons ($`\rho _0=1/2D`$ for a flat band). Noting $`y=E\epsilon _F`$, the density of states at the energy E is $`\rho \left(E\right)=\rho _0\left(1+\sigma _0^2/y^2\right)`$. If $`n_c<1`$, the chemical potential is located just below the upper edge of the $`\alpha `$ -band. The system is metallic. The density of states at the Fermi level is strongly enhanced compared with the bare density of states of conduction electrons : $`\rho (E_F)/\rho _0=(1+\sigma _0^2/y_F^2)1/(2\rho _0\left|y_F\right|)`$. This corresponds to the flat part of the $`\alpha `$-band in Figure 1. It is associated to the formation of a Kondo or Abrikosov-Suhl resonance pinned at the Fermi level resulting of the Kondo effect. The low-lying excitations are quasiparticles of large effective mass $`m^{}`$ as observed in heavy-Fermion systems. Also note the presence of a hybridization gap between the $`\alpha `$ and the $`\beta `$ bands. The direct gap of value $`2\sigma _0`$ is much larger than the indirect gap equal to 2$`\left|y_F\right|`$. The saddle-point solution transposes to N=2 the large-N results obtained within the slave-boson mean-field theories (SBMFT). ### B Appendix B : Expressions of the different bubbles The expressions of the different bubbles appearing in the expression of the boson propagators (cf. Eq.28) are given here (with i=1, 2, m or ff) $$\overline{\phi }_i(𝐪,i\omega _\nu )=\phi _i(𝐪,i\omega _\nu )+\phi _i(𝐪,i\omega _\nu )$$ (56) $`\phi _1(𝐪,i\omega _\nu )`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{k\sigma ,i\omega _n}{}}G_{cf_0}^\sigma (𝐤+𝐪,i\omega _n+i\omega _\nu )G_{ff_0}^\sigma (𝐤,i\omega _n)`$ $`\phi _2(𝐪,i\omega _\nu )`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{k\sigma ,i\omega _n}{}}G_{cc_0}^\sigma (𝐤+𝐪,i\omega _n+i\omega _\nu )G_{ff_0}^\sigma (𝐤,i\omega _n)`$ $`\phi _m(𝐪,i\omega _\nu )`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{k\sigma ,i\omega _n}{}}G_{cf_0}^\sigma (𝐤+𝐪,i\omega _n+i\omega _\nu )G_{cf_0}^\sigma (𝐤,i\omega _n)`$ $`\phi _{ff}^{}(𝐪,i\omega _\nu )`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{k\sigma ,i\omega _n}{}}G_{ff_0}^\sigma (𝐤+𝐪,i\omega _n+i\omega _\nu )G_{ff_0}^\sigma (𝐤,i\omega _n)`$ $`\phi _{cc}^{}(𝐪,i\omega _\nu )`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{k\sigma ,i\omega _n}{}}G_{cc_0}^\sigma (𝐤+𝐪,i\omega _n+i\omega _\nu )G_{cc_0}^\sigma (𝐤,i\omega _n)`$ $`\phi _{fc}^{}(𝐪,i\omega _\nu )`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{k\sigma ,i\omega _n}{}}G_{fc_0}^\sigma (𝐤+𝐪,i\omega _n+i\omega _\nu )G_{fc_0}^\sigma (𝐤,i\omega _n)`$ $`\phi _{ff}^{}(𝐪,i\omega _\nu )`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{k\sigma ,i\omega _n}{}}G_{ff_0}^{}(𝐤+𝐪,i\omega _n+i\omega _\nu )G_{ff_0}^{}(𝐤,i\omega _n)`$ $`\phi _{cc}^{}(𝐪,i\omega _\nu )`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{k\sigma ,i\omega _n}{}}G_{cc_0}^{}(𝐤+𝐪,i\omega _n+i\omega _\nu )G_{cc_0}^{}(𝐤,i\omega _n)`$ $`\phi _{fc}^{}(𝐪,i\omega _\nu )`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{k\sigma ,i\omega _n}{}}G_{fc_0}^{}(𝐤+𝐪,i\omega _n+i\omega _\nu )G_{fc_0}^{}(𝐤,i\omega _n),`$ where $`G_{cc_0}^\sigma (𝐤,i\omega _n)`$, $`G_{ff_0}^\sigma (𝐤,i\omega _n)`$ and $`G_{fc_0}^\sigma (𝐤,i\omega _n)`$ are the Green’s functions at the saddle-point level obtained by inversing the matrix $`G_0^\sigma (𝐤,\tau )`$ defined in Equation (5). The different bubbles $`\phi _{ff}(𝐪,i\omega _\nu )`$, $`\phi _{cc}(𝐪,i\omega _\nu )`$ and $`\phi _{fc}(𝐪,i\omega _\nu )`$ can be expressed as a function of the Green’s functions associated with the eigenoperators $`\alpha _{k\sigma }^{}`$ and $`\beta _{k\sigma }^{}`$ $$G_{ff}(𝐤,i\omega _n)=u_k^2G_{\alpha \alpha }(𝐤,i\omega _n)+v_k^2G_{\beta \beta }(𝐤,i\omega _n)$$ (57) $`G_{cc}(𝐤,i\omega _n)=v_k^2G_{\alpha \alpha }(𝐤,i\omega _n)+u_k^2G_{\beta \beta }(𝐤,i\omega _n)`$ $`G_{cf}(𝐤,i\omega _n)=G_{fc}(𝐤,i\omega _n)=u_kv_k[G_{\alpha \alpha }(𝐤,i\omega _n)G_{\beta \beta }(𝐤,i\omega _n)],`$ where $`G_{\alpha \alpha }(𝐤,i\omega _n)`$ and $`G_{\beta \beta }(𝐤,i\omega _n)`$ are the Green’s functions associated to the eigenstates $`\alpha _{k\sigma }^{}|0>`$and $`\beta _{k\sigma }^{}|0>`$. In the low frequency limit, one can easily check that the dynamical spin susceptibility may be written as $$\chi _{ff}(𝐪,i\omega _\nu )=\frac{\chi _{\alpha \alpha }(𝐪,i\omega _\nu )+\overline{\chi }_{\alpha \beta }(𝐪,i\omega _\nu )}{1J_S^2\chi _{\alpha \alpha }(𝐪,i\omega _\nu )\overline{\chi }_{\alpha \beta }(𝐪,i\omega _\nu )},$$ (58) for both the longitudinal and the transverse parts. $`\chi _{\alpha \alpha }(𝐪,i\omega _\nu )={\displaystyle \frac{1}{\beta }}{\displaystyle \underset{k}{}}{\displaystyle \frac{n_F(E_k^{})n_F(E_{k+q}^{})}{i\omega _\nu E_{k+q}^{}+E_k^{}}}`$ $`\overline{\chi }_{\alpha \beta }(𝐪,i\omega _\nu )={\displaystyle \frac{1}{\beta }}{\displaystyle \underset{k}{}}(u_k^2v_{k+q}^2+v_k^2u_{k+q}^2){\displaystyle \frac{n_F(E_k^{})n_F(E_{k+q}^+)}{i\omega _\nu E_{k+q}^++E_k^{}}}.`$ TABLE CAPTIONS Table I : Predictions for thermal local fluctuations $`\sigma _f^2`$ of the staggered magnetization depending on the temperature $`T`$ and on the dimensionality $`d`$. $`T_I(1I)`$ represents the first cross-over temperature separating the quantum from the classical regime in the vicinity of the antiferromagnetic-quantum critical point (AF-QCP). | | quantum regime $`T<T_I`$ | classical regime $`T>T_I`$ | | --- | --- | --- | | $`d=3`$ | $`\sigma _f^2\chi _{ff}^{}(Q)T^2/\mathrm{\Gamma }_0`$ | $`\sigma _f^2T^{3/2}`$ | | $`d=2`$ | $`\sigma _f^2\chi _{ff}^{}(Q)T^2/\mathrm{\Gamma }_0`$ | $`\sigma _f^2T\mathrm{ln}T`$ | Figure 1: Energy versus wave-vector k for the two bands $`\alpha `$ and $`\beta `$. Note the presence of a direct gap of value $`2\sigma _0`$ and of an indirect gap of value $`2\left|y_F\right|`$. Figure 2: Diagrammatic representation of Equation (39) for the dynamical spin susceptibility $`\chi _{ff}(𝐪,\omega )`$. Figure 3: Continuum of the intra- and interband electron-hole pair excitations $`\chi _{_{\alpha \alpha }}^\mathrm{"}(q,\omega )0`$ and $`\chi _{_{\alpha \beta }}^\mathrm{"}(q,\omega )0`$. Note the presence of a gap in the interband transitions equal to the indirect gap of value $`2\left|y_F\right|`$ at $`q=k_F`$, and to the direct gap of value $`2\sigma _0`$ at $`q=0`$. Figure 4 : Phase diagram in the plane $`(T,I)`$ for dimension equal to $`3`$. The shaded region represents the long-range antiferromagnetic phase bordered by the Néel temperature $`T_N`$. The unshaded region marks the magnetically-disordered regimes $`I,`$ $`II`$ and $`III`$ associated to different behaviors of the system. Regime $`I`$ is the quantum regime in which the energy of the relevant mode on the scale of $`\xi `$ is much greater than $`k_BT`$. The magnetic correlation length is $`\xi 1/\sqrt{(1I)}`$ and the staggered spin susceptibility is $`\chi _Q^{^{}}=\chi _{ff}^0^{}(Q)/(1I+aT^2)`$. Regime $`II`$ and $`III`$ are both classical regimes in which the thermal effects are important since the fluctuations on the scale of $`\xi `$ have energy much smaller than $`k_BT`$. In Regime $`II`$, $`\xi 1/\sqrt{(1I)}`$ is still controlled by $`(1I)`$ but the staggered spin susceptibility is sensitive to the thermal fluctuations : $`\chi _Q^{^{}}=\chi _{ff}^0^{}(Q)/(1I+aT^{3/2})`$. In Regime $`III`$, both $`\xi `$ and $`\chi _Q^{^{}}`$ are controlled by the temperature : $`\xi 1/T^{3/4}`$ and $`\chi _Q^{^{}}=\chi _{ff}^0^{}(Q)/T^{3/2}.`$ Figure 5 : Phase diagram in the plane $`(T,I)`$ for dimension equal to $`2`$. The shaded region represents the long-range antiferromagnetic phase bordered by the Néel temperature $`T_N`$. The unshaded region marks the magnetically-disordered regimes $`I`$ and $`III`$. In that $`d=2`$ case, the equivalent of Regime $`II`$ in Figure 3 is squeezed out of existence since $`T_I=T_{II}`$. Regime $`I`$ is the quantum regime in which the thermal fluctuations are negligible : $`\xi 1/\sqrt{(1I)}`$ and $`\chi _Q^{^{}}=\chi _{ff}^0^{}(Q)/(1I+aT^2)`$. Regime $`III`$ is the unique classical regime in which both the magnetic correlation length and the staggered spin susceptibility are controlled by the temperature : $`\xi 1/\sqrt{T\mathrm{ln}T}`$ and $`\chi _Q^{^{}}=\chi _{ff}^0^{}(Q)/(T\mathrm{ln}T)`$.
warning/0001/cond-mat0001174.html
ar5iv
text
# Many-body spin related phenomena in ultra-low-disorder quantum wires \[ ## Abstract Zero length quantum wires (or point contacts) exhibit unexplained conductance structure close to $`0.7\times 2e^2/h`$ in the absence of an applied magnetic field. We have studied the density- and temperature-dependent conductance of ultra-low-disorder GaAs/AlGaAs quantum wires with nominal lengths $`l`$=0 and $`2\mu m`$, fabricated from structures free of the disorder associated with modulation doping. In a direct comparison we observe structure near $`0.7\times 2e^2/h`$ for $`l`$=0 whereas the $`l=2\mu m`$ wires show structure evolving with increasing electron density to $`0.5\times 2e^2/h`$ in zero magnetic field, the value expected for an ideal spin-split sub-band. Our results suggest the dominant mechanism through which electrons interact can be strongly affected by the length of the 1D region. \] Quantum wires have been used extensively to study ballistic transport in one dimension (1D) where the conductance is quantised in units of $`2e^2/h`$ . This result is well explained by considering the allowed energies of a non-interacting electron gas confined to 1D, where the factor of 2 is due to spin degeneracy. Electron interaction effects in 1D have been considered for some time, involving models which go beyond the conventional Fermi liquid picture. Such correlated electron models have been applied to quantum wire systems and recent experimental studies have investigated their predictions. Although recent theories have considered the effect of weak disorder on correlation effects , it is generally accepted that low-disorder nanostructures are necessary for such investigations. Low-disorder quantum point contacts (which are quantum wires of length $`l`$ = 0) formed in GaAs/AlGaAs heterostructures exhibit unexplained conductance structure close to $`0.7\times 2e^2/h`$ in the absence of a magnetic field . Studies by Thomas et al. suggest that the structure is a manifestation of electron–electron interactions involving spin. The continuous evolution of the $`0.7\times 2e^2/h`$ structure into a Zeeman spin–split conductance plateau with the application of an in-plane magnetic field, together with enhancement of the $`g`$-factor for lower 1D channels, is consistent with this interpretation . In this article we present transport data for 1D systems free from the disorder associated with modulation doped heterostructures, including strong evidence for spin related many-body effects in long 1D regions. We find conductance structure comparable to Thomas’ in our zero length wires, while our 2 $`\mu m`$ quantum wire exhibits plateau–like structure near $`0.5\times 2e^2/h`$ in zero magnetic field, the value expected for an ideal spin-split level. Our results are suggestive of an interpretation in which spin splitting in zero magnetic field is only fully resolved in long 1D regions, perhaps above a critical length scale. This does not explain why structure in short constrictions consistently occurs near $`0.7\times 2e^2/h`$. A clue may be found in recent theories which consider the possibility of a feature at $`0.75\times 2e^2/h`$ due to a splitting between the one singlet and three triplet states where electron pairs (attractive interaction) dominate transport. As the length of the 1D region is increased it is suggested that the dominant many-body interaction can alter and if spontaneous spin polarisation occurs, a principal feature at $`0.5\times 2e^2/h`$ would be observed, perhaps with some remanent weak structure close to $`0.75\times 2e^2/h`$. The study of correlated electron states requires devices with ultra-low-disorder since such states are expected to be easily destroyed by disorder and may be masked by other effects associated with localisation. We have developed a novel GaAs/AlGaAs layer structure which avoids the major random potential present in conventional HEMT devices by using epitaxially grown gates to produce an enhancement mode FET . These devices are advantageous for the study of 1D interacting systems because they eliminate the need for a dopant layer in the AlGaAs adjacent to the 2DEG, thus greatly reducing disorder while allowing the electron density in the 2DEG to be varied over a large range. The electron mobility in the 2DEG is typically $`46\times 10^6`$ cm<sup>2</sup>V<sup>-1</sup>s<sup>-1</sup> at 4.2K and increases further at lower temperatures. At 100mK the 2D ballistic mean free paths exceed $`160\mu m`$ which is greater than our sample dimensions. These devices are comparable with the highest mobility electron systems yet produced. Ballistic conductance plateaus have been demonstrated in quantum wires up to $`5\mu m`$ in length, with the data exhibiting more than 15 plateaus . To investigate the sensitivity of many-body effects to the length of the 1D region, we have measured the conductance of quantum wires of nominal length $`l`$=$`0`$, $`2\mu m`$ and $`5\mu m`$. The devices were patterned from ultra-high-mobility heterostructures, comprising a 75 nm layer of Al<sub>0.3</sub>Ga<sub>0.7</sub>As on top of GaAs to produce the 2DEG interface. A 25 nm GaAs spacer separated the epitaxial conducting top gate from the AlGaAs. NiAuGe ohmic contacts were made to the 2DEG using a self-aligned technique. Electron beam lithography and shallow wet etching were used to selectively remove the top gate to form the quantum wires. The top gate was sectioned into three separately controllable gates. The center (top gate) was biased positively relative to the contacts to induce a 2DEG at the GaAs/AlGaAs interface. This positive bias $`V_T`$ determined the carrier density in the 2DEG reservoirs which was tunable from $`0.66\times 10^{11}`$cm<sup>-2</sup> corresponding to $`V_T`$=0.05V to 0.8V. A negative voltage $`V_s`$ was then applied to the side gates to produce electrostatic 1D confinement in addition to the geometric confinement already present (see Figure 1). Low frequency four-terminal conductance measurements were made with an excitation voltage below $`10\mu V`$ using two lock-in amplifiers to monitor both current and voltage. We stress that the results presented here are raw data as no equivalent series resistance has been subtracted and no attempt has been made to adjust the plateau heights to fit with quantised units of $`2e^2/h`$. Figure 1: a) Cross-sectional schematic of a quantum wire, showing the positively biased top gate and the side gates biased negatively. b) SEM micrograph of a quantum wire with length $`l=5\mu m`$. c) SEM micrograph of a quantum wire with length $`l=0\mu m`$. The conductance $`G`$ of a zero-length quantum wire is shown in Figure 2 as a function of the side gate voltage $`V_s`$ at a temperature T = 50mK. Data were taken at a series of top gate voltages corresponding to different 1D densities. The 1D electron density $`n_{1D}`$ may be controlled using both the top and side gates to vary the shape of the potential well perpendicular to the channel. When both the top and side gates are strongly (weakly) biased positive and negative respectively the confining potential is steep (shallow), leading to a large (small) 1D sub-band spacing and a corresponding high (low) 1D electron density. In this way it is possible to maintain a constant 1D occupancy, and hence conductance, while varying $`n_{1D}`$. For $`G<2e^2/h`$ an additional feature is observed close to $`0.7\times 2e^2/h`$, as seen by others . A similar feature is also observed in a second identical quantum wire with length $`l=0`$ (not shown). As with other workers we find that this feature is robust to cryogenic cycling, indicating that it is unlikely to be related to an impurity state. Although the data in Figure 2 implies a small enhancement of the $`0.7`$ structure with increasing $`n_{1D}`$, the trend is not fully monotonic and is less so for the second $`l=0`$ wire measured. The inset of Figure 2 shows the temperature dependence of the conductance for the wire with length $`l=0`$. These temperature measurements were made with an average electron density $`(n_{2D}3\times 10^{11}`$cm<sup>-2</sup>, $`V_T=0.4`$V). Similar behavior with temperature is seen at high and low electron densities in both of the zero length quantum wires studied. This temperature dependence deviates from the expected single particle result with little thermal smearing below $`0.7\times 2e^2/h`$. Such puzzling behavior is consistent with measurements made by others . Figure 2: Conductance measurements of a $`l=0`$ quantum wire as a function of side gate voltage for top gate voltages, $`V_T`$ = 172mV - 300mV (right to left) in steps of 4mV. Inset: Temperature dependence of the conductance at $`V_T`$=0.4V. The curves are for temperatures 0.5K, 1.0K, 1.5K & 2.9K. The results in Figure 2 demonstrate that these epitaxially gated nanostructures produce ultra-low-disorder quantum wires for $`l=0`$ which exhibit the $`0.7\times 2e^2/h`$ conductance feature comparable with the strongest so far observed. When we extend to longer quantum wires, new and unexpected results are seen. Figure 3 shows the conductance G of a quantum wire with $`l=2\mu m`$ as a function of side gate voltage $`V_S`$. The density $`n_{1D}`$ increases from right to left as the confining potential is steepened. Data were obtained at temperatures $`T`$=1K and $`T`$=50mK. Clear conductance quantisation is seen near integer multiples of $`2e^2/h`$ with up to 15 platueas evident, indicating ballistic transport along the full length of the $`2\mu m`$ wire, as previously reported . The data collected at $`T`$=1K show a clear plateau-like feature which becomes more pronounced and evolves downwards in $`G`$ towards $`0.5\times 2e^2/h`$ as $`V_T`$ is increased. A much weaker inflection is also present near $`0.7\times 2e^2/h`$ across the full density range. Further evidence of many-body phenomena is seen evolving in the range $`1.51.7\times 2e^2/h`$ where the structure is strong enough to resemble the conductance feature seen near $`0.7\times 2e^2/h`$ in quantum wires with $`l=0`$. Figure 3: Conductance of a $`2\mu m`$ quantum wire. Conductance as a function of side gate voltage for top gate voltages, $`V_T`$ = 300mV - 620mV (right to left). The data at the top was obtained close to 1K and the bottom section was obtained at approximately 50mK. Conductance measurements of quantum wires with $`l=5\mu m`$ exhibit similar plateau-like features near $`0.5\times 2e^2/h`$ to wires with $`l`$=$`2\mu m`$, as noted in our previous work . However, for $`l`$=$`5\mu m`$ the weak disorder which is present leads to a distortion of the conductance plateaus, making interpretation more difficult and here we focus on wires with $`l=2\mu m`$ where the single particle plateaus are as clear as those seen in $`l=0`$ devices. As the $`2\mu m`$ wire is cooled to $`T`$ = 50mK the feature near $`0.5\times 2e^2/h`$ remains, however, rich evolving structure is also revealed. Conductance inflections occur below each of the integer plateaus (within $`e^2/h`$) which predominantly evolve downwards in $`G`$ with increasing $`n_{1D}`$. One explanation within a single-particle picture is weak disorder, leading to interference of electron waves along the quantum wire. However, against this, remnants of the strongest features survive at $`T`$=1K, in particular the feature below $`2\times 2e^2/h`$ is reminiscent of the $`0.7\times 2e^2/h`$ feature seen in low-disorder $`l`$=0 wires, implying a possible many-body origin. Figure 4: Evolution of the conductance features seen in $`l=0`$ and $`2\mu m`$ quantum wires as a function of $`n_{1D}`$. The evolution of the corresponding plateau at $`2e^2/h`$ is also shown. Single conductance traces are included for both devices to facilitate interpretation. Open circles are $`2\mu m`$ data and closed circles are $`l=0`$ data. Figure 4 details the evolution of the conductance features seen in quantum wires of length $`l=0`$ and $`2\mu m`$ with varying $`n_{1D}.`$ We define the position of the feature seen in $`l=0`$ devices near $`0.7\times 2e^2/h`$ as the conductance $`G`$ at which $`dG/dV_s`$ is a local minimum. In a similar manner, for the $`l=2\mu m`$ quantum wire we define the position of the plateau-like feature as the first local minimum in the $`dG/dV_s`$ curve for $`T`$=1K. Note that the plateau at $`2e^2/h`$ remains almost constant for the $`l=0`$ wire but for $`l=2\mu m`$ the plateau falls in $`G`$ (by up to 8%) as $`n_{1D}`$ is increased. Suppression of plateaus below the ideal quantised values has been observed in previous studies on quantum wires , and considered theoretically in a number of many-body treatments . In our case the suppression cannot be explained by a simple increase of the effective series resistance associated with the 2D contact regions, since the 2D sheet resistance decreases with increasing $`V_T`$. Abrupt coupling of the 2D reservoirs to the low density 1D region could result in a reduction of the transmission coefficient as the 2D electron density is increased. We note that the density mismatch is larger for the longer wire, since the top-gate voltage threshold for conduction is almost twice as large for $`l=2\mu m`$ as for $`l=0`$. Turning now to the non-integer plateaus we see that the feature near $`0.7\times 2e^2/h`$ in $`l=0`$ devices becomes slightly more pronounced with increasing $`n_{1D}`$ (Figure 2) but the variation in conductance is small. This is in contrast to the plateau-like feature seen in the $`l=2\mu m`$ wire data, which evolves towards $`0.5\times 2e^2/h`$ with increasing $`n_{1D}`$. It is significant that if the single particle plateau for $`l=2\mu m`$ is normalised to equal $`2e^2/h`$ then this feature still evolves downwards in $`G`$ but never falls below $`0.5\times 2e^2/h`$, the position expected for a spin-split 1D plateau. Conductance data suggestive of many-body effects in 1D have now been observed in a variety of high mobility structures including split-gated HEMTs , gate metallised structures and our undoped enhancement mode FETs considered here. Some evidence for this effect has also been seen in low mobility quantum wires based on ion-beam defined GaAs transistors and other material systems such as GaInAs/InP and n-PbTe . The diverse number of experimental systems that have been examined would seem to establish the feature as an intrinsic property of a 1D correlated system. In particular the temperature dependence, described as activated by and detailed in our $`l=0`$ wires, remains consistent between devices of different design . Some important exceptions do exist, however, as in measurements of narrow wires by Yacoby et al. and Tarucha et al. there appears to be no strong feature present even though clear quantisation is seen. The absence of the feature in reference may be associated with a large 1D sub-band spacing made possible in that case due to a novel epitaxial confinement technique. The most commonly invoked explanation for additional conductance structure near $`0.7\times 2e^2/h`$ has been some form of spontaneous spin polarisation mediated through the exchange interaction, as detailed in references . The possibility of a ferromagnetic instability below a critical electron concentration has also been considered . It has so far remained a mystery as to why measurements show structure near $`0.7\times 2e^2/h`$, rather than $`0.5\times 2e^2/h`$, the value expected for a fully spin-polarised 1D level. The fact that we see structure near $`0.7\times 2e^2/h`$ in $`l=0`$ wires and structure evolving towards $`0.5\times 2e^2/h`$ in longer wires (with $`l=2\mu m`$ and $`l=5\mu m`$ ) leads to a possible scenario in which spin-splitting is only fully resolved in wires above some critical length scale. The additional structure we observe near $`1.7\times 2e^2/h`$, and in higher sub-bands below 1K, also suggest that many-body effects become enhanced in longer 1D regions. We note that conductance anomalies in higher sub-bands have been predicted in reference . An alternative explanation for the $`0.7\times 2e^2/h`$ feature has been argued in two recent theories which consider a scenario in which two- (or more) body processes dominate. In the proposed case where electron pairs dominate transport and $`l=0`$, the three triplet states are lower in energy than the one singlet state, leading to a plateau at (slightly less than) $`0.75\times 2e^2/h`$, since the triplet states are transmitted (with transmission coefficient not precisely 1) while the singlet is reflected by the constriction. Within this model the observation in our $`l=2\mu m`$ wire of a dominant feature near $`0.5\times 2e^2/h`$ together with remanent weaker structure in the vicinity of $`0.75\times 2e^2/h`$, which is still noticeable at $`T`$=1K (see Figure 3), would be interpreted as the dominance of spin polarisation in the finite length wire with increasing $`n_{1D}`$ with some remanent signature associated with the singlet/triplet mechanism. However this observation could also be compatible with 1D Wigner crystallisation (dominant repulsive interaction) providing the Landauer-Büttiker framework can be extended to this regime . In conclusion, we have studied ultra-low-disorder quantum wires utilising a novel GaAs/AlGaAs layer structure which avoids the random impurity potential associated with modulation doping, making these devices ideal for the study of electron correlations in 1D. In common with other workers we find structure near $`0.7\times 2e^2/h`$ in wires with $`l=0`$, whereas in longer wires the dominant structure evolves towards $`0.5\times 2e^2/h`$ at high 1D carrier concentrations. It is not possible to be conclusive as to whether these effects are related to a many-electron spin polarisation, or to a more complex explanation (for example 1D Wigner crystalisation). However it is clear that both the length over which interactions occur and the 1D density play an important role in determining the effect of these mechanisms upon electrical transport. We are grateful to B. Altshuler, V. Flambaum and A. R. Hamilton for valuable discussions and R. P. Starrett for technical support. This work was funded by the Australian Research Council grant A69700583.
warning/0001/astro-ph0001190.html
ar5iv
text
# THE MODULATION EFFECT FOR SUPERSYMMETRIC DARK MATTER DETECTION WITH ASYMMETRIC VELOCITY DISPERSION. ## 1 Introduction It is known that that dark matter is needed to close the Universe $`^\mathrm{?}`$, $`^\mathrm{?}`$. It is also known that one needs two kinds of dark matter. One composed of particles which were relativistic at the time of structure formation. These constitute the hot dark matter component (HDM). The other is made up of particles which were non-relativistic at the time of freeze out. This is the cold dark matter component (CDM). The COBE data $`^\mathrm{?}`$ suggest that CDM is at least $`60\%`$ $`^\mathrm{?}`$. On the other hand recent data from the Supernova Cosmology Project suggest $`^\mathrm{?}`$$`^\mathrm{?}`$ that there is no need for HDM and the situation can be adequately described by $`\mathrm{\Omega }<1`$, e.g. $`\mathrm{\Omega }_{CDM}=0.3`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }=0.6`$. In a more recent analysis Turner $`^\mathrm{?}`$ gives $`\mathrm{\Omega }_m=0.4`$. Since the non exotic component cannot exceed $`40\%`$ of the CDM $`^\mathrm{?}`$<sup>,</sup> $`^\mathrm{?}`$, there is room for the exotic WIMP’s (Interacting Massive Particles). Recently the DAMA experiment $`^\mathrm{?}`$ has claimed the observation of one signal in direct detection of a WIMP, which with better statistics has subsequently been interpreted as a modulation signal $`^\mathrm{?}`$. In the currently favored supersymmetric extensions of the standard model the most natural WIMP candidate is the LSP, i.e. the lightest supersymmetric particle, whose nature can be described in most supersymmetric (SUSY) models to be a Majorana fermion, a linear combination of the neutral components of the gauginos and Higgsinos $`^\mathrm{?}`$<sup>-</sup>$`^\mathrm{?}`$. Since this particle is expected to be very massive, $`m_\chi 30GeV`$, and extremely non relativistic with average kinetic energy $`T100KeV`$, it can be directly detected $`^\mathrm{?}`$<sup>-</sup>$`^\mathrm{?}`$ only via the recoiling of a nucleus (A,Z) in the elastic scattering process: $$\chi +(A,Z)\chi +(A,Z)^{}$$ (1) ($`\chi `$ denotes the LSP). In order to compute the event rate one proceeds with the following steps: 1) Write down the effective Lagrangian at the elementary particle (quark) level obtained in the framework of supersymmetry as described in Refs. $`^\mathrm{?}`$, Bottino et al. $`^\mathrm{?}`$ and $`^\mathrm{?}`$. 2) Go from the quark to the nucleon level using an appropriate quark model for the nucleon. Special attention in this step is paid to the scalar couplings, which dominate the coherent part of the cross section and the isoscalar axial current, which, as we will see, strongly depend on the assumed quark model $`^{\mathrm{?},\mathrm{?},\mathrm{?}}`$ 3) Compute the relevant nuclear matrix elements $`^\mathrm{?}`$<sup>-</sup>$`^\mathrm{?}`$ using as reliable as possible many body nuclear wave functions hoping that, by putting as accurate nuclear physics input as possible, one will be able to constrain the SUSY parameters as much as possible. 4) Calculate the modulation of the cross sections due to the earth’s revolution around the sun by a folding procedure assuming some distribution $`^{\mathrm{?},\mathrm{?}}`$ of velocities for LSP. The purpose of our present review is to focus on the last point of our above list along the lines suggested by our recent letter $`^\mathrm{?}`$, expanding our previous results and giving some of the missing calculational details. For the reader’s convenience, however, we will give a brief description on the basic ingredients on how to calculate LSP-nucleus scattering cross section, without elaborating on how one gets the needed parameters from supersymmetry. For the calculation of these parameters from representative input in the restricted SUSY parameter space, we refer the reader to the literature, e.g. Bottino et al. $`^\mathrm{?}`$, Kane et al. , Castano et al. and Arnowitt et al. $`^\mathrm{?}`$. Then we will specialize our study in the case of the nucleus $`{}_{}{}^{127}I`$ which is one of the most popular targets $`^\mathrm{?}`$<sup>-</sup>$`^\mathrm{?}`$. To this end we will include the effect of the nuclear form factors. We will consider both a symmetric Maxwell-Boltzmann distribution $`^\mathrm{?}`$ as well as asymmetric distributions like the one suggested by Drukier $`^\mathrm{?}`$. We will examine the effect modulation in the directional as well as the non directional detection, both in the differential as well as the total event rates. We will present our results a function of the LSP mass, $`m_\chi `$, for various detector energy thresholds, in a way which can be easily understood by the experimentalists. ## 2 The Basic Ingredients for LSP Nucleus Scattering Because of lack of space we are not going to elaborate here further on the construction of the effective Lagrangian derived from supersymmetry, but refer the reader to the literature $`^{\mathrm{?},\mathrm{?},\mathrm{?},\mathrm{?},\mathrm{?}}`$. The effective Lagrangian can be obtained in first order via Higgs exchange, s-quark exchange and Z-exchange. We will use a formalism which is familiar from the theory of weak interactions, i.e. $$L_{eff}=\frac{G_F}{\sqrt{2}}\{(\overline{\chi }_1\gamma ^\lambda \gamma _5\chi _1)J_\lambda +(\overline{\chi }_1\chi _1)J\}$$ (2) where $$J_\lambda =\overline{N}\gamma _\lambda (f_V^0+f_V^1\tau _3+f_A^0\gamma _5+f_A^1\gamma _5\tau _3)N$$ (3) and $$J=\overline{N}(f_s^0+f_s^1\tau _3)N$$ (4) We have neglected the uninteresting pseudoscalar and tensor currents. Note that, due to the Majorana nature of the LSP, $`\overline{\chi }_1\gamma ^\lambda \chi _1=0`$ (identically). The parameters $`f_V^0,f_V^1,f_A^0,f_A^1,f_S^0,f_S^1`$ depend on the SUSY model employed. In SUSY models derived from minimal SUGRA the allowed parameter space is characterized at the GUT scale by five parameters, two universal mass parameters, one for the scalars, $`m_0`$, and one for the fermions, $`m_{1/2}`$, as well as the parameters $`tan\beta `$, one of $`A_0`$ and $`m_t^{pole}`$ and the sign of $`\mu `$ $`^\mathrm{?}`$. Deviations from universality at the GUT scale have also been considered and found useful $`^\mathrm{?}`$. We will not elaborate further on this point since the above parameters involving universal masses have already been computed in some models $`^{\mathrm{?},\mathrm{?}}`$ and effects resulting from deviations from universality will be published elsewhere $`^\mathrm{?}`$ (see also Arnowitt et al in Ref. $`^\mathrm{?}`$ and Bottino et al in Ref. $`^\mathrm{?}`$). For some choices in the allowed parameter space the obtained couplings can be found in a previous paper $`^\mathrm{?}`$. The invariant amplitude in the case of non-relativistic LSP can be cast $`^\mathrm{?}`$ in the form $`||^2`$ $`=`$ $`{\displaystyle \frac{E_fE_im_x^2+𝐩_i𝐩_f}{m_x^2}}|J_0|^2+|𝐉|^2+|J|^2`$ (5) $``$ $`\beta ^2|J_0|^2+|𝐉|^2+|J|^2`$ where $`m_x`$ is the LSP mass, $`|J_0|`$ and $`|𝐉|`$ indicate the matrix elements of the time and space components of the current $`J_\lambda `$ of Eq. (3), respectively, and $`J`$ represents the matrix element of the scalar current J of Eq. (4). Notice that $`|J_0|^2`$ is multiplied by $`\beta ^2`$ (the suppression due to the Majorana nature of LSP mentioned above). It is straightforward to show that $$|J_0|^2=A^2|F(𝐪^2)|^2\left(f_V^0f_V^1\frac{A2Z}{A}\right)^2$$ (6) $$J^2=A^2|F(𝐪^2)|^2\left(f_S^0f_S^1\frac{A2Z}{A}\right)^2$$ (7) $$|𝐉|^2=\frac{1}{2J_i+1}|J_i||[f_A^0𝛀_0(𝐪)+f_A^1𝛀_1(𝐪)]||J_i|^2$$ (8) with $`F(𝐪^2)`$ the nuclear form factor and $$𝛀_0(𝐪)=\underset{j=1}{\overset{A}{}}\sigma (j)e^{i𝐪𝐱_j},𝛀_1(𝐪)=\underset{j=1}{\overset{A}{}}\sigma (j)\tau _3(j)e^{i𝐪𝐱_j}$$ (9) where $`\sigma (j)`$, $`\tau _3(j)`$, $`𝐱_j`$ are the spin, third component of isospin ($`\tau _3|p=|p`$) and coordinate of the j-th nucleon and $`𝐪`$ is the momentum transferred to the nucleus. The differential cross section in the laboratory frame takes the form $`^\mathrm{?}`$ $`{\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}`$ $`=`$ $`{\displaystyle \frac{\sigma _0}{\pi }}({\displaystyle \frac{\mu _r}{m_N}})^2\xi \{\beta ^2|J_0|^2[1{\displaystyle \frac{2\eta +1}{(1+\eta )^2}}\xi ^2]+|𝐉|^2+|J|^2\}`$ (10) where $`m_N`$ is the proton mass, $`\eta =m_x/m_NA`$, $`\xi =\widehat{𝐩}_i\widehat{𝐪}0`$ (forward scattering) and $$\sigma _0=\frac{1}{2\pi }(G_Fm_N)^20.77\times 10^{38}cm^2$$ (11) The reduced mass $`\mu _r`$ is given by $$\mu _r=\frac{m_\chi }{1+\eta }$$ (12) For the evaluation of the differential rate, which is the main subject of the present work, it will be more convenient to use the variables $`(\upsilon ,u)`$ instead of the variables $`(\upsilon ,\xi )`$. Thus integrating the differential cross section, Eq. (10), with respect to the azimuthal angle we obtain $$d\sigma (u,\upsilon )=\frac{du}{2(\mu _rb\upsilon )^2}[(\overline{\mathrm{\Sigma }}_S+\overline{\mathrm{\Sigma }}_V\frac{\upsilon ^2}{c^2}F^2(u))+\overline{\mathrm{\Sigma }}_{spin}F_{11}]$$ (13) with $$\overline{\mathrm{\Sigma }}_S=\sigma _0(\frac{\mu _r}{m_N})^2\{A^2[(f_S^0f_S^1\frac{A2Z}{A})^2]$$ (14) $$\overline{\mathrm{\Sigma }}_{spin}=\sigma _0(\frac{\mu _r}{m_N})^2[f_A^0\mathrm{\Omega }_0(0))^2\frac{F_{00}(u)}{F_{11}(u)}+2f_A^0f_A^1\mathrm{\Omega }_0(0)\mathrm{\Omega }_1(0)\frac{F_{01}(u)}{F_{11}(u)}+(f_A^1\mathrm{\Omega }_1(0))^2]$$ (15) $$\overline{\mathrm{\Sigma }}_V=\sigma _0(\frac{\mu _r}{m_N})^2A^2(f_V^0f_V^1\frac{A2Z}{A})^2[1\frac{1}{(2\mu _rb)^2}\frac{2\eta +1}{(1+\eta )^2}\frac{2u}{\upsilon ^2}]$$ (16) We should remark that even though the quantity $`\overline{\mathrm{\Sigma }}_{spin}`$ can be a function of u, in actual practice it is indepenent of u. The same is true of te less important term $`\overline{\mathrm{\Sigma }}_V`$ In the above expressions $`F(u)`$ is the nuclear form factor and $$F_{\rho \rho ^{}}(u)=\underset{\lambda ,\kappa }{}\frac{\mathrm{\Omega }_\rho ^{(\lambda ,\kappa )}(u)}{\mathrm{\Omega }_\rho (0)}\frac{\mathrm{\Omega }_\rho ^{}^{(\lambda ,\kappa )}(u)}{\mathrm{\Omega }_\rho ^{}(0)},\rho ,\rho ^{}=0,1$$ (17) are the spin form factors with $$u=q^2b^2/2$$ (18) b being the harmonic oscillator size parameter and q the momentum trasfer to the nucleus. The quantity u is also related to the experimentally measurable energy transfer Q via the relations $$Q=Q_0u,Q_0=\frac{1}{Am_Nb^2}$$ (19) The detection rate for a particle with velocity $`\upsilon `$ and a target with mass $`m`$ detecting in the direction $`𝐞`$ will be denoted by $`R(𝐞)`$. Then one defines the undirectiona rate $`R_{undir}`$ via the equations via the equations $$R_{undir}=\frac{dN}{dt}=\frac{\rho (0)}{m_\chi }\frac{m}{Am_N}\sigma (u,\upsilon )[\upsilon .\widehat{e}_x+\upsilon .\widehat{e}_y+\upsilon .\widehat{e}_z]$$ (20) $`\rho (0)=0.3GeV/cm^3`$ is the LSP density in our vicinity. This density has to be consistent with the LSP velocity distribution (see next section). The differential undirectional rate can be written as $$dR_{undir}=\frac{\rho (0)}{m_\chi }\frac{m}{Am_N}d\sigma (u,\upsilon )[\upsilon .\widehat{e}_x+\upsilon .\widehat{e}_y+\upsilon .\widehat{e}_z]$$ (21) where $`d\sigma (u,\upsilon )`$ is given by Eq. ( 13) The directional rate in the direction $`\widehat{e}`$ takes the form: $$R_{dir}=R(e)R(e)=\frac{\rho (0)}{m_\chi }\frac{m}{Am_N}\upsilon .e\sigma (u,\upsilon )$$ (22) and the corresponding differential rate is given by $$dR_{dir}=\frac{\rho (0)}{m_\chi }\frac{m}{Am_N}\upsilon .ed\sigma (u,\upsilon )$$ (23) ## 3 Convolution of the Event Rate We have seen that the event rate for LSP-nucleus scattering depends on the relative LSP-target velocity. In this section we will examine the consequences of the earth’s revolution around the sun (the effect of its rotation around its axis is expected to be negligible) i.e. the modulation effect. This can be accomplished by convoluting the rate with the LSP velocity distribution.Hitherto such a consistent choice can be a Maxwell distribution $`^\mathrm{?}`$ $$f(\upsilon ^{})=(\sqrt{\pi }\upsilon _0)^3e^{(\upsilon ^{}/\upsilon _0)^2}$$ (24) $$v_0=\sqrt{(2/3)v^2}=220Km/s$$ (25) i.e. $`v_0`$ is the velocity of the sun around the center of the galaxy. In the present paper following the work of Drukier, see Ref. $`^\mathrm{?}`$, we will assume that the velocity distribution is only axially symmetric, i.e. of the form $$f(\upsilon ^{},\lambda )=N(y_{esc},\lambda )(\sqrt{\pi }\upsilon _0)^3)[f_1(\upsilon ^{},\lambda )f_2(\upsilon ^{},\upsilon _{esc},\lambda )]$$ (26) with $$f_1(\upsilon ^{},\lambda )=exp[(\frac{(\upsilon _x^{})^2+(1+\lambda )((\upsilon _y^{})^2+(\upsilon _z^{})^2)}{\upsilon _0^2}]$$ (27) $$f_2(\upsilon ^{},\upsilon _{esc},\lambda )=exp[\frac{\upsilon _{esc}^2+\lambda ((\upsilon _y^{})^2+(\upsilon _z^{})^2)}{\upsilon _0^2}]$$ (28) where $`\upsilon _{esc}`$ is the escape velocity in the gravitational field of the galaxy, $`\upsilon _{esc}=625Km/s`$ $`^\mathrm{?}`$. In the above expressions $`\lambda `$ is a parameter, which describes the asymmetry and takes values between 0 and 1 and N is a proper normalization constant given by $`{\displaystyle \frac{1}{N(\lambda ,y_{esc})}}`$ $`=`$ $`{\displaystyle \frac{1}{\lambda +1}}[erf(y_{esc})e^{(\lambda +1)y_{esc}^2}{\displaystyle \frac{erf(i\sqrt{\lambda }y_{esc})}{i\sqrt{\lambda }}}`$ (29) $``$ $`{\displaystyle \frac{e^{y_{esc}^2}}{\lambda }}[{\displaystyle \frac{2}{\sqrt{\pi }}}y_{esc}e^{\lambda y_{esc}^2}{\displaystyle \frac{erf(i\sqrt{\lambda }y_{esc})}{i\sqrt{\lambda }}}]`$ with $`y_{esc}=\frac{\upsilon _{esc}}{\upsilon _0}`$ and erf(x) the error function given by $$erf(x)=\frac{2}{\sqrt{\pi }}_0^x𝑑te^{t^2}$$ (30) For $`y_{esc}\mathrm{}`$ we get the simple expression $`N^1=\lambda +1`$ The z-axis is chosen in the direction of the disc’s rotation, i.e. in the direction of the motion of the the sun, the y-axis is perpendicular to the plane of the galaxy and the x-axis is in the radial direction. In the above frame we find that the position of the axis of the ecliptic is determined by the angle $`\gamma 29.80`$ (galactic latitude) and the azimuthal angle $`\omega =186.3^0`$ measured on the galactic plane from the $`\widehat{𝐳}`$ axis $`^\mathrm{?}`$. Thus, the axis of the ecliptic lies very close to the $`y,z`$ plane and the velocity of the earth around the sun is $$\upsilon _E=\upsilon _0+\upsilon _1=\upsilon _0+\upsilon _1(sin\alpha \widehat{𝐱}cos\alpha cos\gamma \widehat{𝐲}+cos\alpha sin\gamma \widehat{𝐳})$$ (31) where $`\alpha `$ is the phase of the earth’s orbital motion, $`\alpha =2\pi (tt_1)/T_E`$, where $`t_1`$ is around second of June and $`T_E=1year`$. We are now in a position to express the above distribution in the laboratory frame, i.e. $$f(\upsilon ,\lambda ,\upsilon _E)=f_3(\upsilon ,\upsilon _E,\lambda )f_4(\upsilon ,\upsilon _{esc},\lambda )$$ (32) with $`f_3(\upsilon ,\upsilon _E,\lambda )`$ $`=`$ $`exp[{\displaystyle \frac{(\upsilon _x+\upsilon _1sin\alpha )^2}{\upsilon _0^2}}]`$ $`\times `$ $`exp[{\displaystyle \frac{(1+\lambda )((\upsilon _y+\upsilon _1cos\gamma sin\alpha )^2+(\upsilon _z+\upsilon _0+\upsilon _1sin\gamma cos\alpha )^2)}{\upsilon _0^2}}]`$ $$f_4(\upsilon ,\upsilon _{esc},\lambda )=exp[\frac{\upsilon _{esc}^2+\lambda ((\upsilon _y+\upsilon _1cos\gamma sin\alpha )^2+(\upsilon _z+\upsilon _0+\upsilon _1sin\gamma cos\alpha )^2)}{\upsilon _0^2}]$$ (34) ## 4 Expressions for the Differential Event Rate in the Presence of Velocity Dispersion We will begin with the undirectional rate. ### 4.1 Expressions for the Undirectional Differential Event Rate The mean value of the undirectional event rate of Eq. (22), is given by $$<\frac{dR_{undir}}{du}>=\frac{\rho (0)}{m_\chi }\frac{m}{Am_N}f(\upsilon ,\upsilon _E)[\upsilon .\widehat{e}_x+\upsilon .\widehat{e}_y+\upsilon .\widehat{e}_z]\frac{d\sigma (u,\upsilon )}{du}d^3\upsilon $$ (35) From now on we will omit the subscript $`undir`$ in the case of the undirectional rate. The above expression can be more conveniently written as $$<\frac{dR}{du}>=\frac{\rho (0)}{m_\chi }\frac{m}{Am_N}\sqrt{\upsilon ^2}\frac{d\mathrm{\Sigma }}{du}$$ (36) where $$\frac{d\mathrm{\Sigma }}{du}=\frac{[\upsilon .\widehat{e}_x+\upsilon .\widehat{e}_y+\upsilon .\widehat{e}_z]}{\sqrt{\upsilon ^2}}f(\upsilon ,\upsilon _E)\frac{d\sigma (u,\upsilon )}{du}d^3\upsilon $$ (37) It is convenient to work in spherical coordinates. But even then the angular sun is small. Thus introducing the parameter $$\delta =\frac{2\upsilon _1}{\upsilon _0}=\mathrm{\hspace{0.17em}0.27},$$ (38) expanding in powers of $`\delta `$ and keeping terms up to linear in it we can manage to perform the $`\varphi `$ integration using standard contour integral techniques and express the result in terms of the two modified Bessel functions $`I_m(\frac{\lambda \upsilon ^2}{2\upsilon _0^2}(1t^2))`$ with $`t=cos\theta `$ and m=0,1. Thus the angular integration of Eq. 32 yields $`\stackrel{~}{M}_i(\lambda ,y)`$ $`=`$ $`2\pi `$ $`\times `$ $`exp[(y^2+1)(1+\lambda )]\stackrel{~}{\mathrm{\Lambda }}_i(\lambda ,y)exp[(y_{esc}^2+\lambda y^2)]\stackrel{~}{\mathrm{\Lambda }}_i^{^{}}(\lambda ,y),i=1,2`$ where $`\stackrel{~}{\mathrm{\Lambda }}_i,\stackrel{~}{\mathrm{\Lambda }}_i^{^{}}`$ come from $`f_3,f_4`$ respectively. We find $$\stackrel{~}{\mathrm{\Lambda }}_1(\lambda ,y)=\stackrel{~}{\mathrm{\Lambda }}_{1,1}(\lambda ,y)+\stackrel{~}{\mathrm{\Lambda }}_{1,2}(\lambda ,y)+\stackrel{~}{\mathrm{\Lambda }}_{1,3}(\lambda ,y)$$ (40) with $$\stackrel{~}{\mathrm{\Lambda }}_{1,3}(\lambda ,y)=_1^1𝑑texp[(\zeta ^2/22(\lambda +1)yt)]tI_0(\zeta ^2/2)$$ (41) $$\stackrel{~}{\mathrm{\Lambda }}_{1,2}(\lambda ,y)=\frac{1}{\sqrt{\pi }}_1^1𝑑texp[(2(\lambda +1)yt)](1t^2)^{1/2}\frac{erf(i\zeta )}{i\zeta }$$ (42) $$\stackrel{~}{\mathrm{\Lambda }}_{1,1}(\lambda ,y)=\frac{1}{\sqrt{\pi }}_1^1𝑑texp[(2(\lambda +1)yt)](1t^2)^{1/2}\frac{exp(\zeta ^2)erf(\zeta )}{\zeta }$$ (43) where $`\zeta =y(\lambda (1t^2))^{1/2}`$ and the function $`erf(i\zeta )`$ given by Eq. 30. Furthermore $$\stackrel{~}{\mathrm{\Lambda }}_2(\lambda ,y)=(\lambda +1)\stackrel{~}{\mathrm{\Lambda }}_1^{^{\prime \prime }}(\lambda ,y)$$ (44) where $`\stackrel{~}{\mathrm{\Lambda }}_1^{^{\prime \prime }}(\lambda ,y)`$ is obtained from $`\stackrel{~}{\mathrm{\Lambda }}_1(\lambda ,y)`$ by adding in the integrands the extra factor $`ty+1`$. We should mention that in the last integral we have ommitted the numerical factor $`\delta cos\alpha sin\gamma `$. Note that in the case $`\stackrel{~}{\mathrm{\Lambda }}_2(\lambda ,y)`$, which is asociated with the modulation amplitude, only the z-component of the velocity in the exponential contributes. Hence the dependence on the earth’s phase is $`cos\alpha `$. The formulas for the second term in Eq. LABEL:3.14 for $`\stackrel{~}{\mathrm{\Lambda }}_i^{^{}}(\lambda ,y)`$ are obtained by a mere replacement of the expression $`\lambda +1`$ by $`\lambda `$. In all the above expressions $`y=(\upsilon /\upsilon _0)`$ (not to be confused with the y-coordinate). It is convenient to separate out the asymmetric contribution from the usual one by writing $$2y^2\stackrel{~}{\mathrm{\Lambda }}_1(\lambda ,y)=\stackrel{~}{F}_0(\lambda ,(\lambda +1)2y)+\stackrel{~}{G}_0(\lambda ,y)$$ (45) $$2y^2\stackrel{~}{\mathrm{\Lambda }}_2(\lambda ,y)=\stackrel{~}{F}_1(\lambda ,(\lambda +1)2y)+\stackrel{~}{G}_1(\lambda ,y)$$ (46) The functions $`\stackrel{~}{F}_i`$ have been obtained by considering the leading non vanishing term in the zeta expansion of the integrands of the expressions (41)-(43). Thus $$\stackrel{~}{G}_0(0,y)=0,\stackrel{~}{G}_1(0,y)=0$$ (47) $$\stackrel{~}{F}_0(\lambda ,x)=(\lambda +1)^2[xsinh(x)cos(x)+1+xI_1(x)]$$ (48) $`\stackrel{~}{F}_1(\lambda ,x)`$ $`=`$ $`(1+\lambda )^2[(2+\lambda )((x^2/(2(2+\lambda ))+1)cosh(x)xsinh(x)1)`$ (49) $`+`$ $`x^2I_2[x](\lambda +1)xI_1(x)]`$ note that here $`x=(\lambda +1)2y`$. $`I_m(x)`$ is the modified bessel function of order m. The funcions $`\stackrel{~}{G}`$ cannot bo obtained analytically, but they can easily be expressed as a rapidly convergent series in $`y=\frac{\upsilon }{\upsilon _0}`$, which will not be given here. Similarly $$\stackrel{~}{G}_i^{^{}}(\lambda ,y)=2y^2\stackrel{~}{\mathrm{\Lambda }}_i^{^{}}(\lambda ,y),i=1,2$$ (50) Thus the folded non-directional event rate takes the form $$\frac{d\mathrm{\Sigma }}{du}=\overline{\mathrm{\Sigma }}_S\overline{F}_0(u)+\frac{\upsilon ^2}{c^2}\overline{\mathrm{\Sigma }}_V\overline{F}_1(u)+\overline{\mathrm{\Sigma }}_{spin}\overline{F}_{spin}(u)$$ (51) where the $`\overline{\mathrm{\Sigma }}_i,i=S,V,spin`$ are given by Eqs. (14)- (16). The quantities $`\overline{F}_0,\overline{F}_1,\overline{F}_{spin}`$ are obtained from the corresponding form factors via the equations $$\overline{F}_k(u)=F^2(u)\mathrm{\Psi }_k(u)\frac{(1+k)a^2}{2k+1},k=0,1$$ (52) $$\overline{F}_{spin}(u)=F_{11}(u)\mathrm{\Psi }_0(u)a^2$$ (53) (54) $$\stackrel{~}{\mathrm{\Psi }}_k(u)=[\stackrel{~}{\psi }_{(0),k}(a\sqrt{u})+0.135\mathrm{cos}\alpha \stackrel{~}{\psi }_{(1),k}(a\sqrt{u})]$$ (55) with $$a=\frac{1}{\sqrt{2}\mu _rb\upsilon _0}$$ (56) and $$\stackrel{~}{\psi }_{(l),k}(x)=N(y_{esc},\lambda )e^\lambda (e^1\stackrel{~}{\mathrm{\Phi }}_{(l),k}(x)exp[y_{esc}^2]\stackrel{~}{\mathrm{\Phi }}_{(l),k}^{^{}}(x))$$ (57) $$\stackrel{~}{\mathrm{\Phi }}_{(l),k}(x)=\frac{2}{\sqrt{6\pi }}_x^{y_{esc}}dyy^{2k1}exp((1+\lambda )y^2))(\stackrel{~}{F}_l(\lambda ,(\lambda +1)2y)+\stackrel{~}{G}_l(\lambda ,y)))$$ (58) $$\stackrel{~}{\mathrm{\Phi }}_{(l)_,k}^{^{}}(x)=\frac{2}{\sqrt{6\pi }}_x^{y_{esc}}dyy^{2k1}exp(\lambda y^2))\stackrel{~}{G}_l^{^{}}(\lambda ,y))$$ (59) The undirectional differential rate takes the form $$\frac{dR}{du}=\overline{R}tT(u)[(1+\mathrm{cos}\alpha H(u))]$$ (60) In the above expressions $`\overline{R}`$ is the rate obtained in the conventional approach $`^\mathrm{?}`$ by neglecting the folding with the LSP velocity and the momentum transfer dependence of the differential cross section, i.e. by $$\overline{R}=\frac{\rho (0)}{m_\chi }\frac{m}{Am_N}\sqrt{v^2}[\overline{\mathrm{\Sigma }}_S+\overline{\mathrm{\Sigma }}_{spin}+\frac{\upsilon ^2}{c^2}\overline{\mathrm{\Sigma }}_V]$$ (61) where $`\overline{\mathrm{\Sigma }}_i,i=S,V,spin`$ have been defined above, see Eqs (14) - (16). The factor $`T(u)`$ takes care of the u-dependence of the unmodulated differential rate. It is defined so that $$_{u_{min}}^{u_{max}}𝑑uT(u)=1.$$ (62) i.e. it is the relative diffrential rate. $`u_{min}`$ is determined by the energy cutoff due to the performance of the detector. $`u_{max}`$ is determined by the escape velocity $`\upsilon _{esc}`$ via the relations: $$u_{max}=\frac{y_{esc}^2}{a^2}$$ (63) On the other hand $`H(u)`$ gives the energy tranfer dependent modulation amplitude. The quantity $`t`$ takes care of the modification of the total rate due to the nuclear form factor and the folding with the LSP velocity distribution. Since the functions $`\overline{F}_0(u),\overline{F}_1`$ and $`\overline{F}_{spin}`$ in principle have a different dependence on u, the funcions $`T(u),H(u)`$ and $`t`$ in priciple depend on the SUSY parameters. If, however, we ignore the small vector contribution and assume (i) the scalar and axial (spin) dependence on u is the same or (ii) only one mechanism (S, V, spin) dominates the, parameter $`\overline{R}`$ contains the dependence on all SUSY parameters. The other factors depend only on the LSP mass and the nuclear parameters. More specifically considering only the scalar interaction we get $`\overline{R}\overline{R}_S`$ and $$tT(u)=a^2F^2(u)\stackrel{~}{\psi }_{(0),0}(a\sqrt{u})$$ (64) For the spin interaction we get a similar expression except that $`\overline{R}\overline{R}_{spin}`$ and $`F^2F_{\rho ,\rho ^{}}`$. Finally for completeness we will consider the less important vector contribution. We get $`\overline{R}\overline{R}_V`$ and $$tT(u)=F^2(u)[\stackrel{~}{\psi }_{(0),1}(a\sqrt{u})\frac{1}{(2\mu _rb)^2}\frac{2\eta +1}{(1+\eta )^2}u\stackrel{~}{\psi }_{(0),0}(a\sqrt{u})]\frac{2a^2}{3}$$ (65) The quantity $`T(u)`$ depends on nuclear physics through the form factors or the spin response functions and the parameter $`a`$. The modulation amplitude takes the form $$H(u)=0.135\frac{\stackrel{~}{\psi }_{(1),k}(a\sqrt{u})}{\stackrel{~}{\psi }_{(0),k}(a\sqrt{u})},l=1,3$$ (66) Thus in this case the $`H(u)`$ depends only on $`a\sqrt{u}`$, which coincides with the parameter x of Ref. $`^\mathrm{?}`$, i.e. only on the momentum transfer, the reduced mass and the size of the nucleus. Returning to the differential rate it is sometimes convenient to use the quantity $`T(u)H(u)`$ rather than $`H`$, since H(u) may appear artificially increasing function of u due to the decrease of $`T(u)`$ ( in obtaining H(u) we have divided T(u)) Before concluding this subsection we should mention that the above angular integrations can also be done even if the velocity distribution is triaxial. We will not explore this further since one has too many parameters. ### 4.2 Expressions for the Directional Differential Event Rate The mean value of the directional differential event rate of Eq. (23), is defined by $$<\frac{dR}{du}>_{dir}=\frac{\rho (0)}{m_\chi }\frac{m}{Am_N}f(\upsilon ,\upsilon _E)\upsilon .e\frac{d\sigma (u,\upsilon )}{du}d^3\upsilon $$ (67) where $`\widehat{𝐞}`$ is the unit vector in the direction of observation. It can be more conveniently expressed as $$<\frac{dR}{du}>_{dir}=\frac{\rho (0)}{m_\chi }\frac{m}{Am_N}\sqrt{\upsilon ^2}\frac{d\mathrm{\Sigma }}{du}_{dir}$$ (68) where $$\frac{d\mathrm{\Sigma }}{du}_{dir}=\frac{\upsilon .e}{\sqrt{\upsilon ^2}}f(\upsilon ,\upsilon _E)\frac{d\sigma (u,\upsilon )}{du}d^3\upsilon $$ (69) Working as in the previous subsection, i.e by expanding in powers of $`\delta `$ and keeping terms up to linear in it we can manage to perform the $`\varphi `$ integration using standard contour integral techniques and express the result in terms of the two modified Bessel functions $`I_m(\frac{\lambda \upsilon ^2}{2\upsilon _0^2}(1t^2))`$ with $`t=cos\theta `$ and m=0,1. Thus the angular integration of Eq. 32 yields $`M_i(\lambda ,y)`$ $`=`$ $`2\pi `$ $`\times `$ $`exp[(y^2+1)(1+\lambda )]\mathrm{\Lambda }_i(\lambda ,y)exp[(y_{esc}^2+\lambda y^2)]\mathrm{\Lambda }_i^{^{}}(\lambda ,y),i=1,4`$ where $`\mathrm{\Lambda }_i,\stackrel{~}{\mathrm{\Lambda }}i`$ come from $`f_3,f_4`$ respectively and are given by $$\mathrm{\Lambda }_1(\lambda ,y)=_1^1𝑑texp[((\lambda /2)y^2(1t^2)+2(\lambda +1)yt)]tI_0((\lambda /2)y^2(1t^2))$$ (71) $$\mathrm{\Lambda }_2(\lambda ,y)=_1^1𝑑texp[((\lambda /2)y^2(1t^2)+2(\lambda +1)yt)]t^2I_0((\lambda /2)y^2(1t^2))$$ (72) $$\mathrm{\Lambda }_3(\lambda ,y)=_1^1𝑑texp[((\lambda /2)y^2(1t^2)+2(\lambda +1)yt)](1t^2)I_0((\lambda /2)y^2(1t^2))$$ (73) $$\mathrm{\Lambda }_4(\lambda ,y)=_1^1𝑑texp[((\lambda /2)y^2(1t^2)+2(\lambda +1)yt)](1t^2)I_1((\lambda /2)y^2(1t^2))$$ (74) and analogous expressions for $`\mathrm{\Lambda }^{^{}}`$ with the mere replacement of the expression $`\lambda +1`$ by $`\lambda `$. Again in the above expressions y=$`(\upsilon /\upsilon _0)`$ (not to be confused with the y-coordinate). The above integrals can be expressed in terms of hypergeometric functions as follows $$\mathrm{\Lambda }_1(\lambda ,y)=\underset{k=0}{\overset{\mathrm{}}{}}\frac{2(2(\lambda +1)y)^{(2k+1)}}{(2k+3)((2k+1)!)}F_1^1(\frac{1}{2},\frac{2k+5}{2},\lambda y^2)$$ (75) $$\mathrm{\Lambda }_2(\lambda ,y)=(\lambda +1)\underset{k=0}{\overset{\mathrm{}}{}}\frac{2(2(\lambda +1)y)^{(2k)}}{(2k+3)((2k)!)}F_1^1(\frac{1}{2},\frac{2k+5}{2},\lambda y^2)$$ (76) $$\mathrm{\Lambda }_3(\lambda ,y)=(\lambda +1)\underset{k=0}{\overset{\mathrm{}}{}}\frac{2(2(\lambda +1)y)^{(2k)}}{(2k+1)(2k+3)((2k)!)}F_2^2(\frac{1}{2},2,1,\frac{2k+5}{2},\lambda y^2)$$ (77) $$\mathrm{\Lambda }_4(\lambda ,y)=(\lambda +1)\lambda y^2\underset{k=0}{\overset{\mathrm{}}{}}\frac{2(2(\lambda +1)y)^{(2k)}}{(2k+1)(2k+3)(2k+5)((2k)!)}F_1^1(\frac{3}{2},\frac{2k+7}{2},\lambda y^2)$$ (78) $$\mathrm{\Lambda }_1^{^{}}(\lambda ,y)=\underset{k=0}{\overset{\mathrm{}}{}}\frac{2(2\lambda y)^{(2k+1)}}{(2k+3)((2k+1)!)}F_1^1(\frac{1}{2},\frac{2k+5}{2},\lambda y^2)$$ (79) $$\mathrm{\Lambda }_2^{^{}}(\lambda ,y)=\lambda \underset{k=0}{\overset{\mathrm{}}{}}\frac{2(2\lambda y)^{(2k)}}{(2k+3)((2k)!)}F_1^1(\frac{1}{2},\frac{2k+5}{2},\lambda y^2)$$ (80) $$\mathrm{\Lambda }_3^{^{}}(\lambda ,y)=\lambda \underset{k=0}{\overset{\mathrm{}}{}}\frac{2(2\lambda y)^{(2k)}}{(2k+1)(2k+3)((2k)!)}F_2^2(\frac{1}{2},2,1,\frac{2k+5}{2},\lambda y^2)$$ (81) $$\stackrel{~}{\mathrm{\Lambda }}_4(\lambda ,y)=(\lambda y)^2\underset{k=0}{\overset{\mathrm{}}{}}\frac{2(2\lambda y)^{(2k)}}{(2k+1)(2k+3)(2k+5)((2k)!)}F_1^1(\frac{3}{2},\frac{2k+7}{2},\lambda y^2)$$ (82) It is more convenient to define the functions $`F_i`$ ang $`G_i`$, $`i=0,4`$ as follows $$2y^2\mathrm{\Lambda }_1(\lambda ,y)=F_0(2(\lambda +1)y)+G_0(\lambda ,y)$$ (83) $$2y^2(\mathrm{\Lambda }_1(\lambda ,y)+y\mathrm{\Lambda }_2(\lambda ,y))=F_1(\lambda ,2(\lambda +1))+G_1(\lambda ,y)$$ (84) $$4y^3(\mathrm{\Lambda }_3(\lambda ,y)\mathrm{\Lambda }_4(\lambda ,y))=F_2(2(\lambda +1))+G_2(\lambda ,y)$$ (85) $$4y^3(\mathrm{\Lambda }_3(\lambda ,y)+\mathrm{\Lambda }_4(\lambda ,y))=F_3(2(\lambda +1))+G_3(\lambda ,y)$$ (86) The functions $`F_i`$ are obtained by keeping the leading terms in the expansion of the confluent hypergeommetric functions of Eqs. (75)- (78). Thus we find $$F_i(\chi )=\chi cosh\chi sinh\chi ,i=0,2,3$$ (87) $$F_1(\lambda ,\chi )=\mathrm{\hspace{0.17em}2}(1\lambda )\left[(\frac{(\lambda +1)\chi ^2}{4(1\lambda )}+1)sinh\chi \chi cosh\chi \right]$$ (88) The purely asymmetric quantities $`G_i`$ satisfy $$G_i(0,y)=0,i=0,4$$ (89) They are expressed as rapidly convergent series in $`y`$, but they are not going to be given here. Similarly we define the functions $`G^{^{}}`$ via the equations $$G_0^{^{}}(\lambda ,y)=2y^2\mathrm{\Lambda }_1^{^{}}(\lambda ,y)$$ (90) $$G_1^{^{}}(\lambda ,y)=2y^2(\mathrm{\Lambda }_1^{^{}}(\lambda ,y)+y\mathrm{\Lambda }_2^{^{}}(\lambda ,y))$$ (91) $$G_2(\lambda ,y)=4y^3(\mathrm{\Lambda }_3^{^{}}(\lambda ,y)\mathrm{\Lambda }{}_{}{}^{}{}_{4}{}^{}(\lambda ,y))$$ (92) $$G_3^{^{}}(\lambda ,y)=0$$ (93) Thus the folded directional event rate takes the form $$\frac{d\mathrm{\Sigma }}{du}_{dir}=\frac{1}{2}a^2[\overline{\mathrm{\Sigma }}_SF_0(u)+\frac{\upsilon ^2}{c^2}\overline{\mathrm{\Sigma }}_VF_1(u)+\overline{\mathrm{\Sigma }}_{spin}F_{spin}(u)]$$ (94) where the $`\overline{\mathrm{\Sigma }}_i,i=S,V,spin`$ are given by Eqs. (14)- (16). The quantities $`F_0,F_1,F_{spin}`$ are now obtained from the corresponding form factors via the equations $$F_k(u)=F^2(u)\mathrm{\Psi }_k(u)\frac{(1+k)a^2}{2k+1},k=0,1$$ (95) $$F_{spin}(u)=F_{11}(u)\mathrm{\Psi }_0(u)a^2$$ (96) $`\mathrm{\Psi }_k(u)`$ $`=`$ $`{\displaystyle \frac{1}{2}}[(\psi _{(0),k}(a\sqrt{u})+0.135\mathrm{cos}\alpha \psi _{(1),k}(a\sqrt{u}))𝐞_z.𝐞`$ (97) $``$ $`0.117\mathrm{cos}\alpha \psi _{(2),k}(a\sqrt{u})𝐞_y.𝐞+0.135\mathrm{sin}\alpha \psi _{(3),k}(a\sqrt{u})𝐞_x.𝐞]`$ with $$\psi _{(l),k}(x)=N(y_{esc},\lambda )e^\lambda (e^1\mathrm{\Phi }_{(l),k}(x)exp[y_{esc}^2]\mathrm{\Phi }{}_{}{}^{}{}_{(l),k}{}^{}(x))$$ (98) $$\mathrm{\Phi }_{(l),k}(x)=\frac{2}{\sqrt{6\pi }}_x^{y_{esc}}dyy^{2k1}exp((1+\lambda )y^2))(F_l(2y)+G_l(\lambda ,y)))$$ (99) $$\mathrm{\Phi }_{(l),k}^{^{}}(x)=\frac{2}{\sqrt{6\pi }}_x^{y_{esc}}dyy^{2k1}exp(\lambda y^2))G{}_{}{}^{}{}_{l}{}^{}(\lambda ,y))$$ (100) If we consider each mode (scalar, spin vector) separately the directional rate takes the form $$\frac{dR}{du}_{dir}=\frac{\overline{R}}{2}t^0R^0[(1+\mathrm{cos}\alpha H_1(u))𝐞_z.𝐞\mathrm{cos}\alpha H_2(u)𝐞_y.𝐞+\mathrm{sin}\alpha H_3(u)𝐞_x.𝐞]$$ (101) In other words the non directional differential modulated amplitude is described in terms of the three parameters, $`H_l(u)`$, l=1,2 and 3. The unmodulated one is $`R^0(u)`$, which is again normalized to unity. It is the relative differential rate, i.e. the differential rate divided by the total rate, in the absence of modulation, i.e. The parameter $`t^0`$ entering Eq. (101) takes care of whatever modifications are needed due to the convolution of the non modulated total rate with the LSP velocity distribution in the presence of the nuclear form factors. From Eqs. (93) - (101) we see that if we consider each mode separately the differential modulation amplitudes $`H(l)`$ take the form $$H_l(u)=0.135\frac{\psi _k^{(l)}(a\sqrt{u})}{\psi _k^{(0)}(a\sqrt{u})},l=1,3;H_2(u)=0.117\frac{\psi _k^{(2)}(a\sqrt{u})}{\psi _k^{(0)}(a\sqrt{u})}$$ (102) Thus in this case the $`H_l`$ depend only on $`a\sqrt{u}`$, which coincides with the parameter x of Ref. $`^\mathrm{?}`$. We note that in the case $`\lambda =0`$ we have $`H_2=0.117`$ and $`H_3=0.135`$, so that there remains This means that, if we neglect the coherent vector contribution, which, as we have mentioned, is small, $`H_l`$ essentially depends only on the momentum transfer, the reduced mass and the size of the nucleus. Returning to the differential rate it is sometimes convenient to use the quantity $`R_l`$ rather than $`H_l`$ defined by $$R_l=R^0H_l,l=1,2,3.$$ (103) The reason is that $`H_l`$, being the ratio of two quantities, may appear superficially large due to the denominator becoming small. once again if one mechanism dominates the parameters $`R_0`$ and $`R_l`$ are independent of the particular SUSY model considered, except the LSP mass. In fact we find for the scalar interaction we get $`\overline{R}\overline{R}_S`$ and $$t^0R^0(u)=a^2F^2(u)\psi _0^{(0)}(a\sqrt{u})$$ (104) For the spin interaction we get a similar expression except that $`\overline{R}\overline{R}_{spin}`$ and $`F^2F_{\rho ,\rho ^{}}`$. Finally for completeness we will consider the less important vector contribution. We get $`\overline{R}\overline{R}_V`$ and $$t^0R^0(u)=F^2(u)[\psi _1^{(0)}(a\sqrt{u})\frac{1}{(2\mu _rb)^2}\frac{2\eta +1}{(1+\eta )^2}u\psi _0^{(0)}(a\sqrt{u})]\frac{2a^2}{3}$$ (105) The quantity $`R_0`$ depends on nuclear physics through the form factors or the spin response functions. Equation (101) deviates from the simple trigonometric expression.a The dependence on the phase of the earth is complicated. If we imagine, however, that one can sum up all three directonal rates, with the inclusion of $`H_2`$ and $`H_3`$, the maximum does not occur at $`\alpha =0`$, but at $`\alpha =\alpha _H`$ with $$\alpha _H=tan^1[\frac{H_3(u)}{H_1(u)+H_2(u)}]$$ (106) and the modulation at this value of the phase of the earth takes the value $$H_{max}=[(H_1+H_2)^2+H_3^2]^{\frac{1}{2}}$$ (107) There exists one minimum at $`\alpha =\pi `$, i.e. around Dec.2 and takes the value $$H_{min}=H_2H_1$$ (108) Whenever $`H_1>H_2`$ there exist two more minima $`\alpha =\pi /2`$ and $`3(\pi /2)`$, equal to $`H_3`$ and two secondary maxima. In all cases considered in this work $`H_3>H_2H_1`$ so that the interesting quantities are given by Eqs(107 \- 108). In any case it is useful to know the difference between the maximum and the minimum, which takes the form $$H_m=[(H_1+H_2)^2+H_3^2]^{\frac{1}{2}}Min(H_1H_2,H_3)$$ (109) ## 5 The Total Modulated Event Rates Once again we will distinguish two possibilities, namely the directional and the non directional case. Integrating Eq. (101) we obtain for the total undirectional rate $$R=\overline{R}t[(1+h(a,Q_{min})cos\alpha )]$$ (110) where $`Q_{min}`$ is the energy transfer cutoff imposed by the detector. The modulation of the non-directional total event rate can be described in terms of the parameter $`h`$. The effect of folding with LSP velocity on the total rate is taken into account via the quantity $`t`$. The SUSY parameters have been absorbed in $`\overline{R}`$. From our discussion in the case of differential rate it is clear that strictly speaking the quantities $`t`$ and $`h`$ also depend on the SUSY parameters. They do not depend on them, however, if one considers the scalar,spin etc. modes separately. Let us now examine the directional rate. Integrating Eq. (95) we obtain $`R_{dir}`$ $`=`$ $`\overline{R}(t^0/2)[(1+h_1(a,Q_{min})cos\alpha )𝐞{}_{z}{}^{}.𝐞`$ (111) $``$ $`h_2(a,Q_{min})cos\alpha 𝐞{}_{y}{}^{}.𝐞+h_3(a,Q_{min})sin\alpha 𝐞{}_{x}{}^{}.𝐞]`$ Furthermore if we somehow manage to measure the directional rate in all directions we obtain: $$R_{dir,all}=\overline{R}(t^0/2)[1+h_1(a,Q_{min})cos\alpha +h_2(a,Q_{min})|cos\alpha |+h_3(a,Q_{min})|sin\alpha |]$$ (112) We see that the modulation of the directional total event rate can be described in terms of three parameters $`h_l`$ l=1,2,3. In the special case of $`\lambda =0`$ we essentially have one parameter, namely $`h_1`$, since then we have $`h_2=0.117`$ and $`h_3=0.135`$. The effect of folding with LSP velocity on the total rate is taken into account via the quantity $`t^0`$. All other SUSY parameters have been absorbed in $`\overline{R}`$, under the same asumptions discussed above in the case of undirectional rates. Given the functions$`h_l(a,Q_{min})`$ one can plot the the expression in Eq. (112) as a function of the phase of the earth $`\alpha `$. For a gross description one can follow the procedure outlined above making the substitution $`Hh`$. Thus the maximum occurs at $`\pm \alpha _h`$ with $$\alpha _h=tan^1[\frac{h_3(a,Q_{min})}{h_1(a,Q_{min})+h_2(a,Q_{min})}]$$ (113) The difference between the the maximum and the minimum is now given by $$h_m=[(h_1+h_2)^2+h_3^2]^{\frac{1}{2}}Min(h_1h_2,h_3)|$$ (114) In all cases considered here $`h_3>|h_1h_2|`$ ## 6 Discussion of the Results We have calculated the differential as well as the total event rates (directional and non directional) for elastic LSP-nucleus scattering for the target $`{}_{}{}^{127}I`$, including realistic form factors. Only the coherent mode due to the scalar interaction was considered. The spin contribution will appear elsewhere. Special attention was paid to the modulation effect due to the annual motion of the earth. To this end we included not only the component of the earth’s velocity in the direction of the sun’s motion, as it has been done so far, but all of its components. In addition both spherically symmetric $`^\mathrm{?}`$ as well only axially symmetric $`^\mathrm{?}`$ LSP velocity distributions were examined. Furthermor we considered the effects of the detector energy cutoffs, by studying two typical cases $`Q_{min}=10`$ and 20 KeV both on the modulated and the unmodulated aplitudes. We focused our attention on those aspects which do not depend on the parameters of supersymmetry other then the LSP mass. The parameter $`\overline{R}`$, normally calculated in SUSY theories, was not considered in this work. The interested reader is referred to the literature $`^\mathrm{?}`$ <sup>,</sup> $`^\mathrm{?}`$ and, in our notation, to our previous work $`^\mathrm{?}`$ <sup>,</sup> $`^\mathrm{?}`$ <sup>,</sup>$`^\mathrm{?}`$. ### 6.1 The Undirectional Rates Let us begin with the total rates, i.e. the quantities $`t`$ and $`h`$. In Table 1 we show the dependence of $`h`$ on the components of the earth’s velocity, in the symmetric case($`\lambda =0`$). We see that the modulation amplitude increases for about $`50\%`$ when all components of the earth’s velocity are included. In Table 2 we show how the quantities $`t`$ and $`h`$ depend on the detector energy cutoff and the LSP mass for the symmetric case. In tables 3 and 4 we show the same quantities for $`\lambda =0.5`$ and $`\lambda =1.0`$ respectively. From these tables we see a dramatic incease of the modulation when the realistic axially symmetric velocity distribution is turned on. This means that the modulation amplitude can be exploited by the experimentalists. We further notice that the modulation amplitude increases somewhat with cutoff enery. This is due to the fact that the modulation amplitude decreases less rapidly with the cutoff energy $`Q_{min}`$ than the unmodulated amplitude. This effect may be of use to the experimentalists, even though it occurs at the expense of the total rate. Let us now examine the differential rates, which are described by the functions $`T(u)`$, $`H(u)`$ and $`T(u)\times H(u)`$. These are shown for various LSP masses and $`Q_{min}`$ in Fig. 1 ($`\lambda =0.0`$), Fig. 2 ($`\lambda =0.5`$) and Fig. 3 ($`\lambda =1.0`$) We remind the reader that the dimensionless quantity u is related to the energy transfer Q via Eq. (20) with $`Q_0=60KeV`$ for $`{}_{}{}^{127}I`$. The curves shown correspond to LSP masses as follows: i) Solid line $`m_\chi =30`$ GeV. ii) Dotted line $`m_\chi =50`$ GeV. iii) Dashed line $`m_\chi =80`$ GeV. v) Intermediate dashed line $`m_\chi =100`$ GeV. vi) Fine solid line $`m_\chi =125`$ GeV. vi) Long dashed line $`m_\chi =250`$ GeV. If some curves of the above list seem to have been omitted, it is understood that they fall on top of vi). Note that, due to our normalization of T, the area under the corresponding curve is unity. This normalzation was adopted to bring the various graphs on scale since the absolute values may change much faster as a function of the LSP mass. In order to understand the dependence of the total and differential rates on $`\lambda `$, we will examine the functions $`f(y)`$, which are equal to the quantity $`N(\lambda ,y)\frac{2}{\sqrt{6\pi }}`$ multiplied by the integrand of Eq. (58). THe latter crucially depends on the functions $`\stackrel{~}{F}_i(\lambda ,2(\lambda +1)y)`$ and $`\stackrel{~}{G}_i(\lambda ,y)`$ , $`i=1,2`$. These functions $`f(y)`$ are shown in Fig. 4 for $`\lambda =0,1`$. We see that in the case of $`\lambda =1.0`$ the positive section of the function is enhanced. ### 6.2 The Directional Rates Once again we distinguish two cases, the total and the differential rates. 6.2.1 The Directional Total Event Rates The directional total event rates, which arize by summing the directional rates in all three directions is beyond the goals of the present experiments. We will, however, include it in the present discussion. The unmodulated rates can be can be parameterized in terms of the parameter $`t^0`$. This describes the modification of the total directional non modulated event rate due to the convolution with the velocity distribution. The modulation is now described by the three parameters $`h_1,h_2,h_3`$. These are shown in tables 5-7. It is clear from Eq. (112) that the modulation of the total rate is no longer given by a simple sinusoidal function. For some interesting cases the situation is shown in Fig. 5. An idea about what is happening can be given by $`h_m`$ and $`\alpha _m`$. The first gives the difference between the maximum and the minimum values of modulated amplitude. The second involves the phase shift from the second of June, which is no longer the the date of the maximum. The second of June gives the location of the maximum, when only the component of the earth‘s velocity along the sun‘s direction of motion is considered or when $`h_3`$ is neglected. In almost all cases considered in this work, however, $`h_3`$ is important and in fact the obtained shift is on the average about $`\pm `$ 35 days from the second of June. From tables 5-7 we see that without detector energy cutoff, $`Q_{min}=0`$, $`t^0`$ is a decreasing value of the LSP mass. It takes values of about 2 for low LSP mass and is decreased by an order of magnitude as we go to higher LSP masses. This is due to the nuclear form factor effects, which are present but not so severe for this intermediate mass nucleus. It is not a sensitive function of the asymmetry parameter $`\lambda `$. As expected, in the presence of energy cut off, $`t^0`$ is greatly reduced and becomes unobservable for light LSP. As the LSP mass increases $`t^0`$ increases. (see tables 5-7. It reaches a maximum at about 80 GeV and it starts decreasing. But even for the heaviest LSP this reduction is not much larger than 1/3, even for $`Q_{min}=20KeV`$. In other words a heavy LSP can cause sufficient energy transfer to partially compensate for the loss of phase space. We also see that in all cases $`h_m`$ is much larger than $`2h_1`$, suggesting that when it comes to directional detection all components of the earth’s velocity are important. We also notice that for those LSP masses, which give a detectable total rate, the modulation amplitude does not appreciably change with the LSP mass. It is not greatly affected by the energy cutoff $`Q_{min}`$. It is enhanced, however, by about a factor of two in going from $`\lambda =0`$, no asymmetry, to $`\lambda =1`$, maximum asymmetry. From the above discussion it is clear that one needs all three modulation parameters, $`h_1,h_2,h_3`$. We remind the reader that in Eq (110) the z-axis has been chosen in the direction of the sun’s velocity, the y-axis is perpendicular to the plane of the galaxy and the x-axis in the radial (galactocentric) direction. We mention again that $`h_2`$ and $`h_3`$ are constant, 0.117 and 0.135 respectively, in the symmetric case. On the other hand $`h_1`$ and $`h_3`$ substantially increase in the presence of asymmetry. The precise value of the directional rate depends on the direction of observation. One can find optimal orientations, but we are not going to elaborate further. 6.2.2 The Directional Differential Event Rates The directional differential rate is also very hard to detect, but perhaps a bit more practical than the total rates described in the previous subsection. It can be in terms of four functions of u, namely $`R_0(u)`$ and $`H_i(u),i=1,2,3`$. The situation is rather complicated and following our discussion of the previous section we will give gross description of the modulation using the functions $`a_H(u)`$ and $`H_m(u)`$. The phase shift$`\alpha _H`$ has been found to be a constant and about 0.7, which corresponds to a shift about $`\pm `$35 days from June the second. Since , however, $`H_m`$ is defined as the ratio of two quantities, it can appear large because the denominator (non modulated rate) becomes small. Thus, following the strategy of the previous subsection we also present the quantity $`R_m=R_0H_m`$. These functions are shown in Fig. 6 for LSP masses in the range 30-250 GeV, $`\lambda =0`$ and $`Q_{min}=0`$, 10KeV and 20KeV. Note that the quantity $`H_m`$ is itself independent of the cutoff except that one should look at the u relevant to the allowed energy transfer interval. The curves shown correspond to LSP masses as in the undirectional case. Again due to the normalization of R0, the area under the corresponding curve is unity. The above quantities for $`\lambda =0.5`$ and 1.0 are shown in Fig. 7, but only for $`Q_{min}`$=0. Their dependence on the energy transfer cutoff shows behavior similar to that of Fig. 1. In any case for $`Q_{min}`$=10, 20 KeV, the functions $`R_0`$ and $`R_m`$ show a behavior similar to that for $`Q_{min}=0`$, except that they start from higher energy transfer.We should remind the reader, however, that that in all cases $`R_0`$ represents the relative differential rate, i.e. it is normalized so that the area under the corresponding curve is unity for all LSP masses. One, therefore, should take into account the factor $`t^0`$ of tables 5-7. From the plots shown one can see that $`R_m`$ can be quite large, about 20 $`\%`$ of $`R_0`$, but it falls slower as a function of the energy transfer. For this reason the modulation amplitude $`H_m`$ is increasing as a function of u. It is interesting to note that the modulation amplitude $`H_m`$ is increased by more than a factor of two, as the asymmetry parameter $`\lambda `$ changes from zero to one, for all energy transfers. Thus, even at zero energy transfer, for $`\lambda =1`$ the variation in the amplitude due to the earth’s motion can increase by about $`40\%`$ between the minimum (around December 2) and the maximum (around July 10 or the end of may), a big effect indeed. It can become even larger if one can restrict oneself to only part of the phase space, i.e. if one is satisfied with fewer counts. In the case of the directional differential rate one clearly needs, in addition to $`R_0`$ , the functions $`H_l(u)`$, l=1,2,3, which are plotted in Fig.4. In the case of $`\lambda =0`$ only $`H_1`$ is plotted, since the other two are in this case constant ($`H_2=0.117`$, $`H_3=0.135`$). We see that in the presence of asymmetry, e.g. $`\lambda =0.5`$ and 1.0, all functions, but especially $`H_1`$ and $`H_3`$, are substantially increased. ## 7 Conclusions In the present paper we have expanded the the results obtained in of our recent letter $`^\mathrm{?}`$. We have calculated all the parameters, which can describe the modulation of the direct detection rate for supersymmetric dark matter. The differential as well as the total event rates were obtained both for the non directional as well as directional experiments. All components of the earth’s velocity were taken into account, not just its component along the sun’s direction of motion. Realistic axially symmetric velocity distributions, with enhanced dispersion in the galactocentric direction, were considered. The obtained results were compared to the up to now employed Maxwell-Boltzmann distribution. We presented our results in a suitable fashion so that they do not depend on parameters of supersymmetry other than the the LSP mass. Strickly speaking the obtained results describe the coherent process in the case of $`{}_{}{}^{127}I`$, but we do not expect large changes, if the axial current is considered. Recall that the dependence on supersymmetry is contained in the parameter $`\overline{R}`$ not discussed in the present paper. The nuclear form factor was taken into account and the effects of the detector energy cut off were also considered. Our results, in particular the parameters $`t`$ and $`t_0`$ (see tables 2-4 and 5-8) indicate that for large reduced mass, the advantage of $`\mu _r`$ (see Eqs. (14)- (16) is lost when the nuclear form factor and the convolution with the velocity distribution are taken into account. In the case of the undirectional total event rates we find that in the symmetric case the modulation amplitude for zero energy cutoff is less than 0.07. It gets substantially increased in the case of asymmetric velocity distribution with largest asymmetry ($`\lambda =1`$). It can reach values up t0 $`.31`$. In the presence of the detector energy cutoff it can increase even further up to $`0.46`$, but this occurs at the expense of the total number of counts. The modulation amplitude in the case of the differential rate is shifted by the asymmetry at higher energy transfers and, for maximum asymmetry $`\lambda =1`$, gets about doubled compared to the symmetric case ($`\lambda =0`$). This amplitude does not depend on the the energy cutoff, but the lower energy transfers, will, of course, be excluded if such a cutoff exists. Analogous conclusions can be drawn about the directional differential event rate. The presence of asymmetry more than triples the differential modulation amplitude (from about $`10\%`$ to about $`35\%`$). There exist now regions of the energy transfer such that the modulation amplitude can become as large as $`50\%`$. Finally it is important that one should consider all components of the earth’s motion, not just its velocity along the sun’s motion, especially if the directional signals are to be measured. ## Acknowledgments The author would like to acknowledge partial support of this work by $`\mathrm{\Pi }`$ENE$`\mathrm{\Delta }`$ 1895/95 of the Greek Secretariat for research, TMR Nos ERB FMAX-CT96-0090 and ERBCHRXCT93-0323 of the European Union. He would like also to thank the Humboldt Foundation for their award that provided support during the final stages of this work and Professor Faessler for his hospitality in Tuebingen. ## References
warning/0001/cond-mat0001442.html
ar5iv
text
# A numerical study of the formation of magnetisation plateaus in quasi one-dimensional spin-1/2 Heisenberg models ## 1 Introduction The formation of gaps and plateaus in the magnetisation process of one-dimensional (1D) spin-1/2 antiferromagnetic Heisenberg models has been studied intensively during the last years OYA97 ; Hida94 ; Tots98 ; CHP97 ; TNK98 ; CHP98 ; FKM98 ; CG99 ; FKM99 ; FGK+99 . Oshikawa, Yamanaka and Affleck OYA97 pointed out the crucial role which play the soft modes, predicted by the Lieb-Schultz-Mattis (LSM) construction LSM61 . For example in the case of the 1D spin-1/2 Hamiltonian with nearest neighbour couplings and a homogeneous external field $`B`$: $`𝐇(B)`$ $``$ $`𝐇_1B𝐒_3(0),`$ (1) $`𝐇_j`$ $``$ $`2{\displaystyle \underset{l=1}{\overset{N}{}}}𝐒_l𝐒_{l+j},n=1,2,\mathrm{},`$ (2) $`𝐒_a(q)`$ $``$ $`{\displaystyle \underset{l=1}{\overset{N}{}}}e^{ilq}S_l^a,a=1,2,3,`$ (3) the ground state $`|0`$ has momentum $`p_s=0,\pi `$ and total spin $`S_T^3=S_T=NM(B)`$, where $`M(B)`$ is the magnetisation. The LSM construction LSM61 ; AL86 leads to gap-less excited states $`|k`$: $$|k=𝐔^k|0,𝐔\mathrm{exp}\left(i\frac{2\pi }{N}\underset{l=1}{\overset{N}{}}lS_l^3\right),$$ (4) with momenta $$p_k=p_s+kq_3(M),q_3(M)\pi (12M),$$ (5) e.g. for $`M=1/4`$ one finds a four fold degeneracy of the ground state with momenta $`p_k=p_s+k\pi /2,k=0,1,2,3`$. The magnetisation curve for the system (1) has no plateaus. The latter appear, if translational invariance is broken, by adding to (1) a periodic perturbation with wave vector $`q`$, e.g.: $`\overline{𝐃}_j(q)`$ $``$ $`{\displaystyle \frac{1}{2}}\left[𝐃_j(q)+𝐃_j(q)\right],`$ (6) $`𝐃_j(q)`$ $``$ $`2{\displaystyle \underset{l=1}{\overset{N}{}}}e^{iql}𝐒_l𝐒_{l+j},j=1,2,\mathrm{}.`$ (7) So far the case $`j=1`$ has been studied in detail FGK+99 . A pronounced plateau has been found in the magnetisation curve if the period ($`q/\pi =1/2,1/6,1/3`$) coincides with the first soft mode ($`k=1`$), $`q=q_3(M)=\pi (12M)`$, i.e. a perturbation $`\overline{𝐃}_1(\pi /2)`$ generates a plateau at $`M=1/4`$. However, adding to (1) a perturbation $`\overline{𝐃}_1(\pi )`$ no plateau has been observed at $`M=1/4`$. In the latter case the second soft mode at $`q=\pi `$ ($`k=2`$) coincides with the period of the perturbation. Therefore, the question arises, why certain possibilities for the formation of plateaus, which are allowed according to the quantisation rule of Oshikawa et al. OYA97 , are nevertheless not realized with a given perturbation. One possible answer to this question has been given in Ref. FGK+99 : The efficiency of the mechanism to generate a plateau, by means of a periodic perturbation $`𝐃_1(q)`$, crucially depends on the magnitude of transition matrix elements $`n|𝐃_1(q)|0`$ from the ground state $`|0`$ to the low-lying excited states $`|n`$. For example for $`M=1/4`$ the transition matrix elements $`n|𝐃_1(\pi /2)|0`$ turn out to be large whereas $`n|𝐃_1(\pi )|0`$ are small. The effect can be seen directly in the static dimer-dimer structure factor $`0|𝐃_1(q)𝐃_1(q)|0`$ for $`M=1/4`$ (cf. Fig. 4 in Ref. FGK+99 ), which has a pronounced peak at $`q=\pi /2`$ but no peak for $`q=\pi `$. Of course all these statements only hold for the nearest neighbour Hamiltonian (1). Indeed, Totsuka Tots98 has recently observed a magnetisation plateau at $`M=1/4`$, which was created by adding to the Hamiltonian (1) a perturbation $`\overline{𝐃}_1(\pi )`$ and a strong next-to-nearest neighbour coupling. In this paper we show that the magnetisation plateau at $`M=1/4`$ can be realized as well with perturbations of period $`q=\pi `$, if the perturbation operator is chosen properly. We will discuss in detail the situation with the operator $`\overline{𝐃}_2(q)`$ defined in (6). Note, that both operators $`\overline{𝐃}_1(q)`$ and $`\overline{𝐃}_2(q)`$ have the same momentum- and spin-symmetry properties; they change the momentum by $`\pm q`$ and do not change the total spin. They only differ in the isotropic spin-spin couplings, which extend over nearest neighbours in $`\overline{𝐃}_1(q)`$ and over next-nearest-neighbours in $`\overline{𝐃}_2(q)`$. In Sec. 2 we compare the static structure factors, $`S_j0|𝐃_j(q)𝐃_j(q)|0,j=1,2`$, in the presence of a magnetic field with magnetisation $`M=1/4`$. In Sec. 3 we study the magnetisation plateaus induced by the periodic perturbation $`\overline{𝐃}_2(q)`$ at $`q=\pi `$. The zig zag ladder with two (and three) legs – recently investigated in Ref. CHP99 – can be mapped on a 1D system with translation invariant coupling over one and two (three) lattice spacings and a translation invariance breaking coupling of the type $`\overline{𝐃}_2(q)`$. In Sec. 4 we analyse the sequence of magnetisation plateaus, which appears in the zig-zag ladder system. It should finally be added that the operators and Hamiltonians refer to periodic boundary-conditions in leg directions. The DMRG results given in Sec. 4, however, have been obtained using open boundary-conditions along the legs. ## 2 Signals of soft modes in static structure factors In this section we discuss some properties of the static structure factors $`0|𝐃_j(q)𝐃_j(q)|0,j=1,2`$, of the Hamiltonian (1). In order to compute static structure factors we use periodic boundary conditions and exact Lanczos diagonalizations up to $`N=24`$ sites. Soft modes, as predicted by the LSM construction, can be seen directly as zeros in the dispersion curve FGK+99 : $$\omega (q,M)=E(p_s+q,S+1)E(p_s,S),M=S_T^z/N,$$ (8) where $`E(p,S)`$ are the lowest energy eigenvalues with total spin $`S=S_T`$ and momentum $`p`$. The ground-state momentum $`p_s`$ in the sector with total spin $`S`$ is known to be 0 or $`\pi `$YY66a . The zeros of (8) – in the limit $`N\mathrm{}`$ – appear at the soft mode momenta $`q=q^{(k)}(M)`$. For example, at $`M=1/4`$ three zeros at $`q/\pi =0,1/2,1`$ emerge in the dispersion curve for the Hamiltonian (1) (cf. Fig. 3 of Ref. FGK+99 ). The operators $`𝐃_j(q),j=1,2`$, defined in Eq. (7), commute with the total spin squared $`𝐒_T^2`$ and the dispersion curve (8) describes the lowest-lying excitations, which can be reached with these operators. The transition matrix element $$n|𝐃_j(q)|0,j=1,2,$$ (9) from the ground state $`|0`$ with total spin $`S=S_T`$ and momentum $`p_s`$ to the excited states $`|n`$ with momentum $`p=p_s+q`$ enter in the corresponding static structure factor: $$S_j(q,M)\frac{1}{N}\underset{n}{}(1\delta _{n0})|n|𝐃_j(q)|0|^2$$ (10) The signals of the soft modes in the static structure factor therefore measure the magnitude of the transition matrix elements (9). In Figs. 1(a) and 1(b) we compare the $`q`$-dependence of the static structure factors $`S_j(q,M=1/4)`$. Indeed we observe remarkable differences. We find a pronounced peak in both structure factors $`S_j(q,1/4),j=1,2`$ at the soft mode $`q^{(1)}(1/4)=\pi /2`$. The size of the peaks, however, differ by an order of magnitude. There is no peak at the second soft mode $`q^{(2)}(1/4)=\pi `$. Note the different behaviour of the structure factor. For $`q0`$: $`S_1(q,1/4)`$ converges to zero whereas $`S_2(q,1/4)`$ approaches a maximum $`0.3`$. This feature will play a crucial role, if we add to (1) a periodic perturbation $`\delta \overline{𝐃}_j(\pi ),j=1,2`$ of strength $`\delta `$. The perturbation is invariant under translations by two lattice spacings. Eigenstates of the perturbed Hamiltonian are constructed by a superposition of momentum eigenstates with $`p=0`$ and $`p=\pi `$. The reduction of the Brioullin zone is taken into account in a modification of the dimer operators: $$𝐃_j^\kappa (q)2\underset{l=0}{\overset{N/21}{}}e^{iq2l}𝐒_{2l+\kappa }𝐒_{2l+j+\kappa },\{\begin{array}{cc}j=1,2\hfill & \\ \kappa =0,1.\hfill & \end{array}$$ (11) The corresponding structure factors, defined by: $`S_j^\kappa (q,M)`$ $``$ $`{\displaystyle \frac{1}{N}}\left[0|𝐃_j^\kappa (q)𝐃_j^\kappa (q)|0\left|0|𝐃_j^\kappa (q)|0\right|^2\right],`$ are symmetric under the mapping $`q\pi q`$. In Fig. 2(a) we present the static structure factor $`S_1^\kappa (q,1/4),\kappa =0,1`$ – obtained from the ground state of the Hamiltonian (1) with perturbation $`\delta \overline{𝐃}_1(\pi ),\delta =0.5`$. In both structure factors we only find a peak at $`q=\pi /2`$ but no peak at $`q=0,\pi `$. The situation for the static structure factors $`S_2^\kappa (q,1/4),\kappa =0,1`$ is shown in Fig. 2(b). Here the ground state has been computed for Hamiltonian (1) with a perturbation $`\delta \overline{𝐃}_2(\pi ),\delta =0.5`$. Note the different behaviour of the two structure factors $`S_2^\kappa (q,1/4),\kappa =0,1`$. $`S_2^1(q,1/4)`$ has its maximum at $`q=0,\pi `$, whereas $`S_2^2(q,1/4)`$ has a maximum at $`q=\pi /2`$. We therefore expect that the operator $`\overline{𝐃}_2(\pi )`$ generates a plateau at $`M=1/4`$, whereas the operator $`\overline{𝐃}_1(\pi )`$ does not. ## 3 Magnetisation plateaus induced by periodic perturbations Periodic perturbations of the type $`\overline{𝐃}_j(q)`$ – added to the nearest neighbour Hamiltonian (1) – generate a characteristic sequence of plateaus in the magnetisation curve. We dicsuss the same Hamiltonian for which we have computed the static structure factor in Sec. 2.2. For this reason we also apply periodic boundary conditions in this section. The possible position of the plateaus is given by the quantisation rule of Oshikawa et al. OYA97 . However, whether or not a plateau really appears, crucially depends on the type of the perturbation operator. As an example we compare in Fig. 3 the evolution of the magnetisation plateaus generated by the operators $`\overline{𝐃}_1(\pi )`$ and $`\overline{𝐃}_2(\pi )`$. The magnetisation curves – generated with $`\overline{𝐃}_1(\pi )`$, cf. Fig. 3 column (a) – show a plateau at $`M=0`$, rapidly increasing with the strength $`\delta `$ of the perturbation. This is the well known gap induced by dimerisation. There is no plateau at $`M=1/4`$ in the whole $`\delta `$ range ($`0<\delta 0.7`$). In contrast the magnetisation curves – generated with $`\overline{𝐃}_2(\pi )`$, \[cf. Fig. 3 column (b)\] – have no plateau at $`M=0`$. For $`\delta 0.4`$ a plateau appears at $`M=1/4`$. A finite-size analysis of the plateau width yields the $`\delta `$-evolution shown in Fig. 4. Note in particular, the drastic change in the plateau width at $`\delta =0.7`$. This feature is associated with a change in the ground-state quantum numbers in the fixed $`S_T^z`$-sectors. For $`\delta <1/2`$ all ground states ($`0M1/2`$) have the standard momentum $`0,\pi `$. The first change happens in the sector $`M=1/21/N`$ at $`\delta =1/2`$, where the ground state is degenerate with momentum $`p=0,\pi `$ and $`p=\pm \pi /2`$. For larger values of $`\delta `$ $`(\delta >1/2)`$ we observe a different ground-state behaviour in the sectors with: $$0MM_0(\delta ),\text{and}M_0(\delta )M1/2.$$ (13) In the first regime of (13) the ground state has still momentum $`p=0,\pi `$. In contrast, in the second regime of (13) the ground-state momentum alternates between $`p=0,\pi `$ and $`p=\pm \pi /2`$. The magnetisation $`M_0(\delta )`$, which separates the two regimes passes the plateau $`M=1/4`$ exactly at $`\delta =0.7`$. The plateau structures in the magnetization curves of chains with periodic perturbations can be seen already on rather small systems ($`N24`$). ## 4 Magnetisation plateaus in zig-zag ladders The magnetic properties of zig-zag ladders have been recently investigated by Cabra, Honecker and Pujol CHP99 by means of bosonization techniques and numerical analysis. In order to understand the appearance of plateaus in the magnetisation curve, a mapping of the ladder system onto a 1D spin chain with couplings over short range distances is useful. For normal spin ladders with $`l`$ legs this mapping leads to couplings over nearest neighbour and to couplings over $`l`$ lattice sites FKM99 . The couplings over $`l`$ lattice sites appear along the legs, whereas nearest neighbour couplings form the rungs of the ladder. At the endpoints of the rungs nearest neighbour bonds have to vanish to avoid the appearance of diagonal couplings in the ladder. A Fourier analysis of the translation invariance breaking terms: $$\underset{q}{}\delta _q\overline{D}_1(q),$$ (14) leads to a prediction of magnetisation plateaus at: $$M=M_l^Z\frac{1}{2}\frac{Z}{l},$$ (15) where $`Z`$ is integer and runs over the sequence $$Z=\{\begin{array}{cc}1,2,3,\mathrm{},l/2\hfill & :l\text{ even}\hfill \\ 1,2,3,\mathrm{},(l1)/2\hfill & :l\text{ odd}.\hfill \end{array}$$ (16) The prediction is based on the assumption that a plateau only appears if one of the wave vectors $`q`$ in the Fourier analysis (14) of the perturbation coincides with the first soft mode $`(k=1)`$. In the following subsections we consider different realizations of zig-zag spin ladders, which have been discussed recently by several authors CHP99 ; WSME99 ; Kole99 . We map those systems to 1D systems with appropriate couplings over $`j=1,2,3`$ neighbours. For some of the following systems we expect more than one plateau in the magnetization curve, therfore we use in this section DMRG calculations, which also means we apply open boundary conditions, to get a finer resolution, i.e. more steps in the finite-size magnetization curves. ### 4.1 The two-leg zig-zag ladder The mapping of the two leg zig-zag ladder onto a 1D system with short range couplings is shown in Fig. 5. The Hamiltonian can be written as: $`𝐇`$ $`=`$ $`J_1𝐇_1+J_2(1+\delta )𝐇_{+\delta }+J_2(1\delta )𝐇_\delta ,`$ (17) $`𝐇_{\pm \delta }`$ $``$ $`2{\displaystyle \underset{l=1}{\overset{N}{}}}\left[𝐒_{4l}𝐒_{4l\pm 2}+𝐒_{4l+1}𝐒_{4l\pm 2+1}\right].`$ (18) The Fourier analysis (14) of the translation invariance breaking terms yields in this case: $`𝐇`$ $`=`$ $`{\displaystyle \underset{j=1,2}{}}J_j𝐇_j+\delta J_2\sqrt{8}{\displaystyle \underset{l=1}{\overset{N}{}}}\mathrm{cos}\left({\displaystyle \frac{\pi l}{2}}{\displaystyle \frac{\pi }{4}}\right)𝐒_l𝐒_{l+2}.`$ Therefore we expect magnetisation plateaus, if the wave vector $`q/\pi =1/2,3/2`$ meets the first and second soft mode: $`q={\displaystyle \frac{\pi }{2}}`$ $`:`$ $`M_{4k}^1={\displaystyle \frac{1}{2}}{\displaystyle \frac{1}{4k}},\text{ for }k=1,2,`$ (20) $`q={\displaystyle \frac{3\pi }{2}}`$ $`:`$ $`M_{4k}^3={\displaystyle \frac{1}{2}}{\displaystyle \frac{3}{4k}},\text{ for }k=2.`$ (21) We have looked in particular for plateaus at $`M_8^3=1/8`$ and $`M_8^1=3/8`$ induced by the second soft mode $`k=2`$. The situation at $`J_1=1,J_2=2`$ and $`\delta =0.6,0.8,1.0`$ is shown in Fig. 6, where the emergence of a plateau at $`M=1/8`$ is visible. The effect disappears if we change the ratio $`\alpha =J_1/J_2`$ in both directions $`\alpha <1/2`$ and $`\alpha >1/2`$. ### 4.2 Three-leg zig-zag ladder with periodic rung couplings The mapping of the three-leg zig-zag ladder onto a 1D system with short range couplings is shown in Fig. 7. The corresponding Hamiltonian can be rewritten as: $`𝐇`$ $`=`$ $`{\displaystyle \underset{j=1,3}{}}J_j𝐇_j+{\displaystyle \frac{2}{3}}J_2\left[𝐇_22{\displaystyle \underset{l=1}{\overset{N}{}}}\mathrm{cos}\left({\displaystyle \frac{2\pi }{3}}l\right)𝐒_l𝐒_{l+2}\right].`$ The last term on the right-hand side breaks the translation invariance of the 1D system and we therefore expect magnetisation plateaus for $$M_3^1=\frac{1}{6},M_6^1=\frac{1}{3},M_9^2=\frac{7}{18},$$ (23) if the wave vector $`q=2\pi /3`$ meets the first ($`k=1`$), second ($`k=2`$) and third ($`k=3`$) soft mode, respectively. We have computed the magnetisation curves with periodic rung couplings and open boundary-conditions along the legs for the following values: $$J_1=\frac{3}{2}J_3,J_2=\frac{3}{2}J_3,$$ (24) and system sizes: $`N=3\times L,L=8,12,16,20,24`$. It turns out, that the midpoint extrapolation à la Bonner and Fisher BF64 still shows a surprisingly large finite-size effects for the open boundary conditions. However, as can be seen from Fig. 8, there is still a clean signal for two plateaus at $`M=1/6`$ and $`M=1/3`$. In Ref. CHP99 , the magnetisation curve has been computed for the same set of couplings (24), but with different boundary-conditions along the rungs, which they call periodic boundary-conditions of type A, B or C. No plateau at all is found for type B, one plateau at $`M=1/6`$ is found for type A and C. This is a first indication that the formation of plateaus critically depends on the boundary-conditions along the rungs. ### 4.3 Three-leg ladder with open rung couplings The mapping of this system onto a 1D system with short range couplings can be seen again from Fig. 7. We only have to remove the dotted nearest neighbour bonds, which implement the periodic rung couplings. This changes the Hamiltonian in the following manner: $`𝐇`$ $`=`$ $`{\displaystyle \frac{2}{3}}{\displaystyle \underset{j=1,2}{}}J_j\left[𝐇_j2{\displaystyle \underset{l=1}{\overset{N}{}}}\mathrm{cos}\left[{\displaystyle \frac{2\pi }{3}}(l+2j)\right]𝐒_l𝐒_{l+j}\right]+`$ (25) $`+J_3𝐇_3.`$ The two Hamiltonians (4.2) and (25) differ in the relative weight of the translation symmetry breaking terms. It was found already in Ref. CHP99 that for the couplings (24) $`(J_1=3/2,J_2=3/2,J_3=1)`$ the magnetisation curve for the Hamiltonian (25) evolves two pronounced plateaus at $`M=1/6`$ and at $`M=1/3`$ generated by the first and second soft mode, respectively. This is confirmed in our computation. Switching from antiferromagnetic to ferromagnetic leg coupling $`(J_3=1)`$ we find spontaneous magnetisation at $`M=1/6`$ whereas the plateau at $`M=1/3`$ disappears (Fig. 9). ### 4.4 The Kagomé like three spin ladders This system has been studied recently in Ref. WSME99 for reasons which we explain below. Its couplings are defined by removing from Fig. 7 the middle leg and the dotted vertical lines. The mapping onto the 1D system with short range couplings then leads to the Hamiltonian: $`𝐇`$ $`=`$ $`{\displaystyle \frac{2}{3}}{\displaystyle \underset{r=1}{\overset{3}{}}}J_r𝐇_r`$ (26) $`{\displaystyle \frac{4}{3}}(J_1{\displaystyle \underset{l=1}{\overset{N}{}}}\mathrm{cos}\left[{\displaystyle \frac{2\pi }{3}}(l+1)\right]𝐒_l𝐒_{l+1}+`$ $`J_2{\displaystyle \underset{l=1}{\overset{N}{}}}\mathrm{cos}\left[{\displaystyle \frac{2\pi }{3}}l\right]𝐒_l𝐒_{l+2}+`$ $`J_3{\displaystyle \underset{l=1}{\overset{N}{}}}\mathrm{cos}\left[{\displaystyle \frac{2\pi }{3}}(l1)\right]𝐒_l𝐒_{l+3},).`$ Equation (26) differs from Eq. (25) in the additional translation symmetry breaking term over three lattice spacings. The Kagomé like three leg ladder is interesting because it shares with the two dimensional Kagomé lattice the property that there is a high density of low-lying singlets. A different Kagomé like three leg ladder system has been analysed recently by Azaria et al. ALN98 . The Kagomé lattice itself seems to have a singlet-triplet gap. The authors of Ref. WSME99 tried to find out whether such a gap (i.e. a plateau at $`M=0`$) exists as well for the Kagomé like three spin ladder. Extrapolation from DMRG results for systems up to 120 sites show that the system is gap-less for values of the leg spin coupling $`J`$ in the interval $`0.5<2J_3<1.25,J_1=1/2`$. Note, that the translation invariance breaking terms in (26) do not generate a plateau at $`M=0`$ but at $`M=1/6`$ and $`M=1/3`$. Our results for fixed couplings $`J_1=J_2=1.0`$ and increasing couplings $`J_3=0.4,0.6,1.0`$ are shown in Fig. 10. For $`J_3=0.4`$ we find spontaneous magnetisation at $`M=1/6`$. This phenomenon is a consequence of the Lieb-Mattis theorem LM62 , as was pointed out by the authors of Ref. WSME99 . In condensed matter physics this phenomenon is called ferrimagnetism. The magnetisation curve for $`J_3=0.6`$ starts with a steep increase at $`B=0`$ and reaches quickly the plateau at $`M=1/6`$ for a small value of $`B0.1`$. Going to larger values of the coupling $`J_3=1.0`$, the slope of the magnetisation curve and the plateau width at $`M=1/6`$ are reduced. At the same time, we expect the appearance of a small plateau at $`M=1/3`$. Beyond this point the system jumps into the saturation magnetisation. This phenomenon is called meta-magnetism GMK98 . We finally want to give some comments on the strong finite-size behavior of the presented DMRG calculations of magnetization curves especially appearing at the upper critical fields of the shown magnetization plateaus. Reanalyzing the 3-leg zig-zag and Kagomé like ladders of this section for system sizes $`N=12,18,24`$ and periodic leg boundary conditions applying standard Lanczos techniques led to a straight affirmation of the presented extrapolations. It moreover showed that the strong finite-size effects at the upper plateau edges are no genuine additional features but a simple consequence of the open leg boundary conditions used for the DMRG calculations (see e.g. plateau boundaries additionally given in Fig. 10). In addition, the evaluations with periodic leg boundary conditions could make clear the existence of an additional plateau ($`m=1/3`$) for the Kagomé like 3-leg ladder shown in Fig. 10 ($`c`$), i.e. $`J_3=1.0`$. Again, at the upper critical field strong finite-size corrections hinder the identification of the $`m=1/3`$ plateau on the basis of the shown DMRG results. For the two other cases –$`J_3=0.4,\mathrm{\hspace{0.17em}0.6}`$– additional plateaus could be excluded. It remains to be summarized that evaluations with both types of leg boundary conditions agree –if possible to obtain– in the thermodynamic limits of the considered quantities. Periodic boundary conditions, however, show much smaller and more controlled behavior of finite-size effects while the open boundary conditions require considerably larger system sizes for an equal quality of extrapolation. ## 5 Discussion and Conclusions The Lieb-Schultz-Mattis construction of gap-less excited states (4) in quasi one-dimensional spin-1/2 quantum spin systems demands translation invariance and short range couplings. To our knowledge systems which satisfy these conditions have no plateaus in their magnetisation curve for $`M>0`$. In this paper we have studied the effect of a modulation of the couplings over $`j=1,2,3`$ sites \[cf. Eq. (6)\] on the magnetisation process. According to the quantisation rule of Oshikawa, Yamanaka and Affleck, we expect plateaus if the wave number $`q`$ of the modulation coincides with one of the momenta of the soft modes $`q^{(k)}=k\pi (12M),k=1,2,\mathrm{}`$. We have found, that the modulation of the nearest neighbour coupling \[with the operator $`\overline{𝐃}_1(q)`$\] generates one plateau at $`M=1/2q/\pi `$, i.e. if $`q`$ coincides with the first soft mode $`q^{(1)}(M)`$. For $`q=\pi `$, the modulation of the nearest neighbour coupling with the operator $`\overline{𝐃}_1(\pi )`$ leads to the well-known singlet-triplet gap, i.e. a plateau at $`M=0`$ – which opens with the strength $`\delta `$ of the perturbation as $`\delta ^{2/3}`$CF79 ; FKM98 . The modulation of the next-nearest-neighbour coupling with $`\overline{𝐃}_2(\pi )`$ does not affect substantially the magnetisation process if $`\delta <0.4`$. There is neither a plateau at $`M=0`$ nor at $`M=1/4`$. For $`\delta >0.4`$, however, a plateau opens rapidly at $`M=1/4`$ and shrinks again for $`\delta >0.7`$. We have found that this effect is correlated with a change in the quantum numbers of the ground state. Spin ladders with $`l`$ legs can be mapped on one dimensional systems with modulated short range couplings. Of course these mappings are not unique. There are many possibilities to put a ring with nearest neighbour couplings on a ladder in such a way, that each site is passed once. The links which do not lie on the ring, define further reaching and modulated couplings. However, these mappings become unique, if we postulate that the range of the couplings is minimal. This means for normal ladders with $`l`$-legs, that the corresponding 1D system only contains modulated nearest neighbour couplings $`[\overline{𝐃}_1(q)]`$ and (translation invariant) couplings over $`l`$ lattice sites. Here magnetisation plateaus appear if the wave vectors $`q`$ of the magnetisation plateaus meet the first soft mode. The situation is different if we map zig-zag ladders on quasi 1D systems. Modulations of the next-nearest-neighbour couplings $`[\overline{𝐃}_2(q)]`$ emerge, which generate magnetisation plateaus if the wave vector $`q`$ of the modulation meets either the first or the second soft mode. Finally, we studied the magnetisation process of three leg zig-zag ladders with various boundary-conditions along the rungs. The boundary-conditions change the weight of the terms which modulate the couplings over one, two and three lattice spacings in the 1D Hamiltonian. This again affects the formation of plateaus at $`M=1/6`$ and $`M=1/3`$, respectively.
warning/0001/astro-ph0001292.html
ar5iv
text
# VSL theories and the Doppler peak ## I Introduction Over the past year or so there has been a growing effort to establish some sort of link between ‘fundamental’ high-energy physics and the more standard cosmological scenario . Although this important task is still at a very early stage (perhaps analogous to the state of cosmology itself some 20 years ago), some general ideas are already emerging. Among these are the concepts of dimensional reduction and compactification . Since high-energy theories are best formulated in 10 or 11 dimensions and our own low-energy world is four-dimensional, it is clear that some sort of dimensional reduction mechanism (whose precise details are still unknown) will be involved. When one goes through such a procedure, one usually finds that one or more of the ‘constants’ of Nature are time and/or space-varying quantities. This normally arises because the fundamental coupling constants are associated with the radii of additional dimensions, which are usually variable. Typical examples of this can be found in multi-dimensional Kaluza-Klein models , superstring theories , and in the so-called ‘brane world’ . This fact is in part responsible for the recent sharp increase in the interest in theories with ‘varying constants’, and in particular in theories where the speed of light $`c`$ is time-dependent . These are commonly called ‘varying speed of light’ (VSL) theories, although this is a bit of a misnomer . It should be said that all of these are essentially toy models, but they are interesting because in certain (but arguably rather specific) circumstances they can solve the standard cosmological problems . We should also point out that on the observational side there are some very tentative hints of a time variation (at redshifts of order unity) of the fine-structure constant , but these still require further confirmation. As has been noted before , there is no unique way to generalise General Relativity to accommodate variations in $`c`$ and/or other constants. At the ‘toy-model’ level, one must therefore choose some particular set of postulates which will characterise the theory. Most notably, a number of standard invariance principles and conservation laws (such as covariance, Lorentz invariance, mass and particle number conservation) may or may not hold in the generalised theory. In a previous paper we have shown that solutions to the standard cosmological problems are not a general feature of varying speed of light (or indeed any ‘varying constants’) theories. The fact that the theory proposed by Albrecht, Barrow and Magueijo solves the cosmological problems should be attributed to it breaking covariance and Lorentz invariance. We have also proposed an arguably more ‘natural’ generalisation of General Relativity which allows for variations in $`c`$ and $`G`$ and preserves (at least in some cases) all of the above invariances, and argued that no such generalisation can solve the standard cosmological problems. In this paper we discuss some simple constraints on a varying speed of light in the early universe. In particular, we investigate the constraints imposed on the theory of Albrecht, Barrow and Magueijo by the recent determination of the position of the first Doppler peak in the CMB angular power spectrum. We should point out that there has been some recent work on constraining variations of $`\alpha `$ on cosmological timescales ; these effectively interpret it as a variation in the electric charge $`e`$ (No similar work has been done for variations of the speed of light, although some possible tests have been proposed ). The main point made by these authors is that a variation of $`\alpha `$ alters the ionization history of the universe, and hence changes the pattern of CMB fluctuations. The dominant effect is a change in the redshift of recombination, due to a shift in the energy levels (and in particular the binding energy) of Hydrogen. The Thomson scattering cross section is also changed for all particles (being proportional to $`\alpha ^2`$). Increasing $`\alpha `$ increases the redshift of last scattering (hence the position of the Doppler peak, $`l_p`$) and decreases the high-l damping. A smaller effect (which these authors ignore) is expected to come from a change in the Helium abundance. It is claimed that the next generation of CMB experiments can set constraints of the order of $$\frac{\mathrm{\Delta }\alpha }{\alpha }10^3,$$ (1) or equivalently $$\frac{\dot{\alpha }}{\alpha }5\times 10^{13}y^1$$ (2) at a redshift $`z10^3`$. Note that the claimed detection of Webb et al. is $$\frac{\mathrm{\Delta }\alpha }{\alpha }(1.5\pm 0.3)\times 10^5,$$ (3) at redshifts $`0.5<z<1.6`$, although previous work at higher redshifts (up to $`z3`$) only finds the constraint $$\frac{\mathrm{\Delta }\alpha }{\alpha }1.6\times 10^4.$$ (4) There is also a well-known nucleosynthesis constraint $$\frac{\dot{\alpha }}{\alpha }2\times 10^{14}y^1$$ (5) at redshifts $`z10^910^{10}`$ although this is a model-dependent result, since it relies on a particular (and arguably incorrect) model for the dependence of the neutron to proton mass difference on $`\alpha `$. More recent work finds a weaker bound by two orders of magnitude. Finally, for reference we also add that the best laboratory limit is $$\frac{\dot{\alpha }}{\alpha }3.7\times 10^{14}y^1,$$ (6) while the well-known Oklo constraint is $$\frac{\dot{\alpha }}{\alpha }5\times 10^{17}y^1.$$ (7) The plan of this paper is as follows. In the following section we briefly review the VSL model of Albrecht, Barrow and Magueijo, being particularly careful with the choices of fundamental units. Section III contains the main results of the paper—after a short review of the relevant CMB physics, we derive constraints on the speed of light at the epoch of recombination, and briefly discuss nucleosynthesis. Finally, in section IV we summarize our main results and discuss future work. ## II The model In the VSL theory proposed by Albrecht, Barrow and Magueijo , one postulates minimal coupling at the level of Einstein’s equations. Physically, the most obvious consequence of this choice is that there should be no $`\dot{c}`$ term in the Friedmann and Raychaudhuri equations which can be written in the usual way given by $`\left({\displaystyle \frac{\dot{a}}{a}}\right)^2`$ $`=`$ $`{\displaystyle \frac{8\pi G}{3}}\rho {\displaystyle \frac{Kc^2}{a^2}}`$ (8) $`{\displaystyle \frac{\ddot{a}}{a}}`$ $`=`$ $`{\displaystyle \frac{4\pi G}{3}}\left(\rho +3{\displaystyle \frac{p}{c^2}}\right).`$ (9) However, this choice also implies the breaking of covariance and Lorentz invariance, as well as of mass and particle number conservation if the curvature is non-zero. In this case the conservation equation is given by $$\dot{\rho }+3\frac{\dot{a}}{a}\left(\rho +\frac{p}{c^2}\right)=\rho \frac{\dot{G}}{G}+\frac{3Kc^2}{4\pi Ga^2}\frac{\dot{c}}{c}.$$ (10) Note that this also allows for a time-variation of Newton’s constant $`G`$. As we pointed out above, there is no unique way to generalise General Relativity in order to include variations of the speed of light or other fundamental quantities. Even after imposing the above postulate, there is a lot of freedom remaining. Physically, the reason for this is that one can only measure dimensionless combinations of the fundamental parameters. Hence, different choices of units will lead to theories that, although in the same class, will have different varying dimensional quantities. In the theory of Albrecht and Magueijo the quantity $$Q=\mathrm{}/c$$ (11) is a constant with $$c\mathrm{}\alpha ^{1/2},$$ (12) where $`\alpha `$ is the fine structure constant. They further assume that the mass of elementary particles, such as the electron mass, $`m_e`$, and the electron charge, $`e`$, are constants. It is straightforward to show from dimensional analysis that there is a unique set of standard units of mass, length and time (M,L,T) which can be constructed using combinations of the fundamental parameters $`\mathrm{}`$, $`m_e`$, $`e^{}=e/\sqrt{4\pi ϵ_0}`$ and $`c`$ which satisfy these criteria. These are $$M=m_e,$$ (13) $$L=Q/m_e,$$ (14) $$T=Q^{3/2}/m_ee^{}.$$ (15) Hence, by making this choice of units we are able to interpret a variation in the fine structure constant $`\alpha =e_{}^{}{}_{}{}^{2}/\mathrm{}c`$ as being associated to a change in the speed of light, $`c`$, and the Planck constant $`\mathrm{}`$. A very important aspect of this theory is the way in which the energy of elementary particles scales with $`c`$. Given that the mass of elementary particles is conserved in this theory their energy scales as $`Ec^2`$ which means that for the same mass density a larger speed of light will imply a hotter universe—with $`Tc^2`$. Another crucial aspect of this theory is the way in which the Rydberg energy $$E_R=m_ee_{}^{}{}_{}{}^{4}/2\mathrm{}^2,$$ (16) which represents the dependence of the atomic levels on the fundamental parameters, is going to vary in this cosmology. Given that $`m_e`$ and $`e^{}`$ are both constants, we have that $`E_Rc^2`$. Hence, a larger speed of light also implies a smaller ionisation temperature. Note that, if the Rydberg energy was instead $$E_R=m_ee_{}^{}{}_{}{}^{4}\alpha _+^2/2\mathrm{}^2\alpha ^2,$$ (17) with $`\alpha _+`$ being the present value of the fine structure constant then $`E_R`$ would scale with $`c`$ in the same way as the energy of the elementary particles. This would imply major modifications to quantum mechanics. We will study the implications of these scenarios for the spectrum of CMB anisotropies in the next section. ## III The CMB anisotropy The description of the CMB anisotropies in terms of the angular power spectrum, $`C_l`$, has proved to be an invaluable method and has become a standard procedure for treatment of the temperature fluctuations of the CMB radiation. Contemporary cosmological models with adiabatic fluctuations predict a sequence of peaks on the power spectrum which are generated by acoustic oscillations of the photon-baryon fluid at recombination. Photon pressure resists compression of the fluid by gravitational infall and sets up acoustic oscillations. The fluctuations as a function of the wavenumber $`k`$ go as $`\mathrm{cos}(kc_s\eta _{ls})`$ at last scattering, where $`c_s`$ is the sound speed and $`\eta _{ls}`$ is the conformal time at recombination. This will produce a harmonic series of temperature fluctuation peaks, with the $`m`$th peak corresponding to $$k_m=m\pi /c_s\eta _{ls}.$$ (18) in the case of primordial adiabatic fluctuations. The critical scale is essentially the sound horizon $`c_s\eta _{ls}`$ at last scattering . Of particular interest is the height and position of the main acoustic peak—the so called Doppler peak. In the standard cosmological model with adiabatic fluctuations produced during inflation the height depends on quantities like the baryonic content of the universe ($`\mathrm{\Omega }_b`$) and the Hubble constant ($`H_0`$), whilst its position depends on the total density of the universe ($`\mathrm{\Omega }_0`$, including the contribution of a cosmological constant), and is expected to occur on an angular scale $`1^o`$ if $`\mathrm{\Omega }_01`$. The precise form of the Doppler peak depends on the nature of the dark matter and the values of $`\mathrm{\Omega }_0`$, $`\mathrm{\Omega }_b`$ and $`H_0`$. The scale $`l_p`$ of the main peak reflects the size of the horizon at last scattering of the CMB photons and thus depends almost entirely on the total density of the universe according to $$l_p1/\sqrt{\mathrm{\Omega }_0}.$$ (19) This relation comes from the conversion of a spatial fluctuation on a distant surface to an anisotropy on the sky. Hence, in an open universe a given scale subtends a smaller angle on the sky than in the flat universe, due to the fact that photons curve in their geodesics. In a flat $`\mathrm{\Lambda }`$CDM universe the main acoustic peak is located at around $`l220`$ with a small dependence on $`\mathrm{\Omega }_0`$ and $`h`$ . In the case of topological defect models the situation is not entirely clear yet . Here we study the shift in the Doppler peak position which may be induced by a phase transition after recombination, at a red-shift $`z_{}`$ when the speed of light changed from $`c^{}`$ to $`c^+`$ (in the following we shall assume that $`c^{}>c^+`$). From the discussion in the previous section we see that the recombination temperature in this theory is given by $$T_{ls}=T_{ls}^S(c^+/c^{})^2=T_{ls}^S(\alpha _{}/\alpha _+)$$ (20) if the Rydberg energy has the usual form (see eqn. (16)). The alternative choice given in eqn. (17) would imply that $$T_{ls}=T_{ls}^S(c^{}/c^+)^2=T_{ls}^S(\alpha _+/\alpha _{}).$$ (21) Here $`T_{ls}^S`$ and $`T_{ls}`$ represent respectively the recombination temperature in the standard and VSL cosmological models. On the other hand, the temperature of the photons at a given redshift, $`z>z_{}`$, is given by $$T(z)=T_\gamma ^0(1+z)(c^{}/c^+)^2=T_\gamma ^0(1+z)(\alpha _+/\alpha _{}),$$ (22) where $`T_\gamma ^0`$ is the temperature of the photons today. These two effects combined mean that the redshift of recombination is smaller than the standard one, $$1+z_{ls}=(1+z_{ls}^S)(c^+/c^{})^4=(\alpha _{}/\alpha _+)^2,$$ (23) for the usual form of Rydberg energy, $`E_R`$ (see eqn. (16)). If the alternative choice for the Rydberg Energy (see eqn. (17)) is used then $`z_{ls}^S`$ and $`z_{ls}`$ would be equal. There are two main effects which can modify the CMB spectrum, shifting the main acoustic peak position. Firstly the comoving distance to the last scattering surface: $$d=_{t_{ls}}^t_{}c^{}\frac{dt}{a(t)}+_t_{}^{t_0}c^+\frac{dt}{a(t)},$$ (24) will be altered due to the larger value of the speed of light prior to the phase transition. This effect will tend to shift the peak position to smaller angular scales but it will be negligible if the redshift of the phase transition, $`z_{}`$, is large enough. However, this will be the only effect altering the Doppler peak position in the case of the ‘unusual’ choice of the Rydberg energy (see eqn. (17)). If we want the shift in the main acoustic peak position due to this effect to be small then the following constraint is easily obtained: $$1+z_{}(c^{}/c^+)^2=(\alpha _+/\alpha _{})^2.$$ (25) The second effect which shifts the position of the main acoustic peak if the Rydberg energy has the usual form (see eqn. (16)) is the change in the sound horizon at last scattering. This is a consequence of two competing effects: the larger Hubble radius and the smaller value of the sound speed which are due to a smaller value of the red-shift of last scattering (see eqn. (23)). For this purpose let us consider, without loss of generality, a flat background Universe. In the matter era the conformal time $`\eta `$ will be proportional to $`a^{1/2}`$. On the other hand the adiabatic sound speed will be $`c_s=1/\sqrt{3},a<a_{B\gamma }`$ (26) $`c_sa^{1/2},a_{B\gamma }<a<a_{ls},`$ (27) with $`a_{B\gamma }`$ corresponding to the epoch at which $`\rho _\gamma =\rho _B`$ (and $`B`$ standing for baryons). Therefore the sound horizon before recombination will evolve as $`c_s\eta a^{1/2},a<a_{B\gamma }`$ (28) $`c_s\eta =\mathrm{const}.,a>a_{B\gamma },`$ (29) Here we should note that although in the standard cosmological model $`a_{B\gamma }>a_{ls}`$ this does not need to be the case for our model due to the change of the redshift of last-scattering. Thus we see that a smaller value of the redshift of recombination will imply a larger sound horizon at last scattering. The angle subtended by the sound horizon at last scattering, $`\theta _{ls}`$, will be larger and consequently the scale, $`l_p1/\theta _{ls}`$, of the main peak will be smaller. In conclusion, the main effect induced by the change in the red-shift of last scattering is an increase in the size of the sound horizon at last scattering, shifting the position of the main acoustic peak toward lower values of the multipole $`l`$. It is straightforward to include curvature effects in the previous considerations. If $`z_{}1`$ then $`\mathrm{\Omega }(z_{})`$ will be close to unity and the above statements will still apply. Hence the effect of curvature will be the same as in the standard scenario, namely, to shift the position of the peak towards higher values of the multipole $`l`$ (the effect being proportional to $`1/\sqrt{\mathrm{\Omega }_0}`$). In refs. the effect of a varying fine-structure constant $`\alpha `$ on the CMBR was studied interpreting this variation as being due to a change in the electron charge. In this case the Rydberg energy (see equation (16)) is proportional to $`\alpha ^2`$ and consequently the recombination temperature in this theory is given by: $$T_{ls}=T_{ls}^S(\alpha _{}/\alpha _+)^2,$$ (30) Given that, in this case, the evolution of the energy of the background photons is not affected, the red-shift of recombination is just given by: $$1+z_{ls}=(1+z_{ls}^S)(\alpha _{}/\alpha _+)^2.$$ (31) We see that the amplitude of this effect is exactly the same as that described by eqn. (23). However, the shift of the peak position due to the change of the distance to the last scattering surface is negligible in this case (if $`z_{ls}1`$). We can use the recent determination of the main peak position by the BOOMERanG experiment of $$l_p=215\pm 15$$ (32) to obtain constraints to our model. In everything that follows we will use the usual choice of the Rydberg energy, as discussed above. We will also assume that the effect of the change in the distance to the last-scattering surface is small. For $`a_{B\gamma }a_{ls}`$, or equivalently $`\mathrm{\Omega }_Bh^20.05(c^+/c^{})^4`$, we can express the scale $`l_p`$ as follows: $$l_p\frac{220}{\sqrt{\mathrm{\Omega }_0}}\sqrt{\frac{1+z_{ls}}{1+z_{ls}^S}}=\frac{220}{\sqrt{\mathrm{\Omega }_0}}\left(\frac{c^+}{c^{}}\right)^2$$ (33) Using equation 32 we obtain: $$\frac{c^+}{c^{}}=(1.00\pm 0.04)\mathrm{\Omega }_0^{1/4}.$$ (34) For $`a_{B\gamma }a_{ls}`$, or equivalently $`\mathrm{\Omega }_Bh^20.05(c^+/c^{})^4`$ we obtain the following constraint: $$\frac{\mathrm{\Omega }_Bh^2}{\mathrm{\Omega }_0}=0.05\pm 0.007.$$ (35) Thus we see that in our model the maximum allowed value for $`\mathrm{\Omega }_Bh^2/\mathrm{\Omega }_0`$ is $`0.06`$. This is because the Doppler peak constrains the epoch when baryons start to dominate over radiation (through the effect of the sound horizon, as discussed above), and this obviously depends on how many baryons there are. We should note that there is also a possible contribution from a third effect. It has been suggested (see for example ) that a sudden pase transition will affect the amplitude of any existing density fluctuations, and that sub-horizon and super-horizon scales may be affected differently. The precise meaning of these statements is somewhat unclear, as they can easily be changed depending on the choice of a ‘preferred gauge’ in the theory. This is a manifestation of the lack of covariance of this formulation of the model, and is indeed the most serious problem it faces (and nobody has addressed it so far). In any case it is clear, simply at an intuitive level, that the density fluctuations might be affected, perhaps in a scale-dependent way. If this is true, this might introduce a further shift in the position of the Doppler peak (which has not been addressed here). Even though we use the toy-model of Albrecht, Barrow and Magueijo, our results should be valid (perhaps with small modifications) for other VSL prescriptions. Moreover, since we are discussing variations of speed of light at recent times, we expect these to be at most of the order of a few percent (as was confirmed a posteriori). Any variations in the density fluctuations are expected to be of this order of magnitude. Since the current error in the determination of the position of the Doppler peak is itself of order $`10\%`$, such effects are not crucial at this stage, although they should of course be accounted for when more precise data is available. ## IV Discussion and conclusions In this paper we investigated the implications of a change in the value of the speed of light at some time between last scattering and today in the framework of the theory proposed by Albrecht, Barrow and Magueijo. We have shown that this may alter the spectrum of the CMBR in particular by shifting the position of the main Doppler peak to larger scales. This effect may on one hand complicate cosmological parameter determinations with CMB but on the other hand could bring structure formation models with isocurvature fluctuations (including topological defects) in better agreement with the data. We find that for a ‘standard’ cosmological scenario, the best currently available determination of the position of the Doppler peak (that of the BOOMERanG/NA flight) is consistent with a variation of up to $`4\%`$ in the speed of light after the epoch of recombination, relative to the present value. On the other hand, less standard scenarios are also admissible, provided that they obey $`\mathrm{\Omega }_Bh^20.06\mathrm{\Omega }_0`$. Future CMB experiments should dramatically improve these constraints. In ref. stringent constraints on variations of the fine structure constant from nucleosynthesis were derived (interpreting the variation in $`\alpha `$ as being due to a variation of the electron charge). Here we should note that, even if these constraints could be applied to our model there is always the possibility that the variation of the speed of light is not monotonic making it possible to satisfy nucleosynthesis constraints and still affect the CMBR. Moreover, the problem is further complicated by the fact that, assuming that one can successfully include a VSL formulation in the context of some grand-unified theory, one can expect different charges in the theory to vary at different rates. In any case, in VSL models important modifications to standard quantum mechanics may be required and only after these are properly introduced will it be possible to make reliable predictions as far as nucleosynthesis is concerned. ###### Acknowledgements. We would like to thank Paulo Carvalho and Paulo Macedo for enlightening discussions. C.M. and G.R. are funded by FCT (Portugal) under ‘Programa PRAXIS XXI’, grants nos. PRAXIS XXI/BPD/11769/97 and PRAXIS XXI/BPD/9990/96 respectively.
warning/0001/astro-ph0001173.html
ar5iv
text
# Large-scale shocks in the starburst galaxy NGC 253 ## 1 Introduction Among all tracers of the high density gas in spirals, the SiO molecule presents the most peculiar, and so far, poorly understood chemistry. Galactic surveys of SiO clouds show that the relative abundance of this species varies by more than five orders of magnitude (the relative abundance, X(SiO), ranges from $`<`$10<sup>-12</sup> to 10<sup>-7</sup>). SiO emission is practically absent in quiescent dark clouds, indicating a high degree of depletion in grains as X(SiO)$``$a few 10<sup>-12</sup> (Ziurys et al 1989). In contrast, SiO is enhanced in the dense and hot gas of GMCs forming massive stars, in particular, it appears in association with bipolar outflows where X(SiO)$``$10<sup>-7</sup>-10<sup>-8</sup> (Martín-Pintado et al 1992, Schilke et al 1997, Gueth et al 1998). These high abundances have been interpreted as a signature of shock chemistry: SiO comes into the gas phase by the sputtering of grains by fast shocks. At significantly lower abundances (X(SiO)$``$10<sup>-11</sup>-10<sup>-10</sup>), SiO is predicted in Photon-Dominated-Regions (PDRs), complexes of molecular gas heated by UV photons produced by OB associations (Janssen et al 1995, Walmsley et al 1999). Finally, SiO emission has been reported in molecular clouds of the Galactic Center with X(SiO)$``$10<sup>-8</sup>-10<sup>-9</sup> (Martín-Pintado et al 1997, Hüttemeister et al 1998). However these clouds show little evidence of recent star formation. The emission in these clouds might be connected with large-scale shocks rather than with star formation. Shocks are possibly driven by the bar potential or, more locally, by superbubbles and supernova remnants. SiO emission was first detected in NGC 253 by Mauersberger and Henkel 1991. A later study by Sage and Ziurys (1995) reported the detection of J=2–1 SiO emission in several starburst galaxies, including NGC 253. The authors concluded that the ratio I(SiO)/I(N<sub>2</sub>H+) is an indicator of the SFR per unit mass. However, their conclusions on the origin of SiO emission are based on low resolution observations of only one transition. With the aim of studying the physical and the chemical parameters of SiO clouds, we mapped in the J=5–4, 3–2 and 2–1 transitions of SiO the nuclei of three prototypical starbursts: M 82, NGC 253 and IC 342 (Martín-Pintado et al 1999, hereafter called MP99). First results of our survey, made with the 30m telescope, underlined the different influence of star formation in these galaxies. First, the inferred SiO abundances were seen to vary from galaxy to galaxy (from X(SiO)$``$10<sup>-8</sup> in IC 342 to $``$10<sup>-10</sup> in M 82), showing no correlation with changes on the SFR per unit mass. Physical conditions of SiO clouds, derived from the transition ratios, are also markedly different in our sample. However, any further refinement in the interpretation of SiO emission required a better angular resolution. In particular, it is crucial to elucidate which is the exact spatial extent of SiO emission in the nuclei of starburst galaxies. So far it is unclear if SiO emission is exclusively linked with the formation of massive stars, or alternatively, with the occurrence of large-scale shocks, as observations of clouds in our Galaxy seem to indicate. We want to address the question of which is the chemical scenario explaining the abundance of SiO in galaxies. In this paper we study at high-resolution the SiO(v=0,J=2-1) emission in the nucleus of NGC 253, a highly inclined barred Sc spiral representing the archetype of a nuclear starburst. The bulk of its large infrared luminosity L$`{}_{IR}{}^{}`$1.6$`\times `$10<sup>10</sup>L originates in the inner 300 pc (Telesco and Harper 1980). The observed infrared spectrum of the nucleus (Telesco and Harper 1980; Engelbracht et al 1998) as well as the radio continuum observations made at centimeter wavelengths (Antonucci and Ulvestad 1988, Ulvestad and Antonucci 1991) support the starburst hypothesis. There is also evidence of a giant outflow of hot gas powered by the starburst, that comes out of the galaxy plane (Fabbiano and Trinchieri 1984, McCarthy et al 1987, Schulz and Wegner 1992). The high molecular gas mass of the NGC 253’s nucleus ($``$2–3$`\times `$10<sup>8</sup>M; Canzian et al 1988; Mauersberger et al 1996) explains the high SFR. Water maser emission detected by Ho et al (1987) reveals strong star formation activity. The gas reservoir might have been driven inward by the barred potential first identified by Scoville et al 1985 in a K-band image of the galaxy. Several interferometer maps showing the molecular gas distribution of NGC 253’s nucleus have been published (CO: Canzian et al 1988; HCN: Paglione et al 1995; CN: Hüttemeister and Aalto 1998, HCO<sup>+</sup>: Carlstrom et al 1990; OH: Turner 1985; H<sub>2</sub>CO: Baan et al 1997; CS: Peng et al 1996, hereafter P96) However, no high-resolution image of the dense gas distribution (n(H<sub>2</sub>)$`>`$10<sup>5</sup>cm<sup>-3</sup>) has been obtained so far, using an optically thin tracer. To fill this gap, we have made simultaneous observations in the J=1–0 line of H<sup>13</sup>CO<sup>+</sup>, a tracer of dense gas with low opacity. Furthermore all the results for SiO and H<sup>13</sup>CO<sup>+</sup> are discussed in parallel throughout the paper. In particular we study how the ratio of integrated intensities I(SiO)/I(H<sup>13</sup>CO<sup>+</sup>) changes across the nucleus and discuss how this result is used to get the variation of the absolute abundance of SiO. ## 2 Observations Observations of NGC 253 were made with the IRAM array at Plateau de Bure (France) between June and August 1998. We simultaneously observed the (v=0,J=2–1) line of SiO (at 86.847 GHz) and the (J=1–0) line of H<sup>13</sup>CO<sup>+</sup> (at 86.754 GHz) using the standard BC set of 4-antenna configurations. The 55<sup>′′</sup> primary beam field of the interferometer was phase-centered at $`\alpha _{J2000}`$=$`00^h47^m33.0^s`$ and $`\delta _{J2000}`$=-$`25^{}17^{}18.4^{\prime \prime }`$. The SIS receivers were tuned for single side band operation at 86.731 GHz yielding system temperatures of $`100200`$ K. The spectral correlator was adjusted to give a contiguous bandwidth of 420 MHz. This is equivalent to a velocity range of 1500 kms<sup>-1</sup>; the correlator was centered at 86.731 GHz in order to cover both transitions. The effective frequency resolution was set to 2.5 MHz, or equivalently 8.64 kms<sup>-1</sup> at this frequency. Visibilities were obtained using on-source integration times of 20 minutes interspersed with 4 minutes of instrumental amplitude and phase calibrations. These were made by observing 0135-247 and 2345-167. The atmospheric phase noise on the most extended baselines ranged from 20 to 40, consistent with a seeing of 0.8–1.2”, typical for summer weather conditions. The absolute flux density scale was established on the basis of cross-correlations on the radio star MWC349 (with $`S_{87GHz}=950`$ mJy). This is in full agreement with the measured interferometer efficiency and should be accurate to 10%. The receiver passband shape was determined on 3C454.3 and its accuracy is better than 5% throughout the observing run. Cleaned maps were obtained from the visibilities using the standard IRAM package. The maps were $`256\times 256`$ pixels in extent, with a pixel size of $`0.4^{\prime \prime }`$. The synthesized beam, as determined by fitting a Gaussian to the dirty beam, is $`7.5^{\prime \prime }\times 2.6^{\prime \prime }`$, oriented north-south (PA=180). Due to the low declination of the galaxy, the uv plane is unequally sampled with a maximum spacing of 70 m in the north-south direction and of 250 m in the east-west. The corresponding east-west linear scale at the distance of the source, assuming $`D=3.4`$ Mpc (Sandage and Tammann 1975), is 42 pc. A continuum map was obtained by averaging visibilities over a bandwidth chosen to be free of line emission. In practice we used spectral channels with velocities 200 kms<sup>-1</sup> above and $`200`$ kms<sup>-1</sup> below the galaxy’s systemic velocity. For the effective 235 MHz bandwidth used to obtain continuum data, we derived a one $`\sigma `$ point source sensitivity limit of 0.44 mJy/beam. The rms noise level in 5 MHz wide channel maps, as determined from the dirty maps after subtraction of the continuum, is 2 mJy/beam. This corresponds to an rms brightness temperature of 17 mK for the synthesized beam size and is consistent with a total on-source integration time of 15 hours and a mean system temperature of 150 K. We have estimated the zero-spacing flux lacking in our maps. Taking into account that the shortest spacing measured by the interferometer is $``$20m, we expect to filter out scales $``$30$`\mathrm{}`$. Compared to the 30m data of MP99, only $``$20$`\%`$ of the single-dish flux in both SiO and H<sup>13</sup>CO<sup>+</sup> is missing within the primary beam. Therefore, our data offer an unbiased picture of the total SiO and H<sup>13</sup>CO<sup>+</sup> gas content in the nucleus of NGC 253. We assume that the orientation of the NGC 253 disk is defined by the angle PA<sub>disk</sub>=51, and inclination i=78.5. The orientation of the major axis of the stellar bar, along PA<sub>bar</sub>=68, is used on purpose in some figures throughout the paper. ## 3 The interferometer maps ### 3.1 The radio continuum emission The 3mm continuum emission contours are plotted in Figure 1a. Despite its compactness, this continuum source, with a peak flux of 70mJy/beam, is resolved by our beam, with 80$`\%`$ of the flux coming from the central source with a deconvolved size of $``$8$`\mathrm{}\times `$4.3$`\mathrm{}`$, oriented along the disk major axis (PA$``$50). The integrated flux amounts to 0.250$`\pm `$0.008Jy; extrapolating fluxes at 6cm and 2cm(Turner and Ho 1983), we estimate that nonthermal and thermal free-free emission account each for one third of the observed 3mm continuum flux. The remaining third would come from emission by dust. Although our value for the integrated flux agrees within the errors with that measured at 85GHz by Carlstrom et al 1990 (0.300$`\pm `$0.050Jy), it is slightly lower than that derived by P96 at 98GHz (0.320$`\pm `$0.030Jy). The better sensitivity of our maps, compared with prior observations (Carlstrom et al 1990; P96) reveals the presence of two previously undetected sources located along the bar major axis at offsets x=-10$`\mathrm{}`$ and x=10$`\mathrm{}`$. These new sources appear marginally as extensions in the 1.3mm maps of Krügel et al 1990. This suggests that they correspond mostly to dust emission. However the nature of these 3mm new sources is unknown. Comparison of our data with measurements at other wavelengths is not straightforward because of the missing flux in the images at the centimeter range and, on the other hand, owing to the lack of resolution of the existent maps at 1.3mm, 0.8mm and 0.45mm. Before high angular resolution observations at short millimeter and submillimeter wavelengths are obtained, little can be said on the nature of these new sources. Contrary to P96, we see no extension of continuum emission to the southwest of the nucleus, interpreted as a signature of the giant outflow detected in X-rays. Continuum emission of the dust outflow has only been tentatively detected at 0.45mm by Alton et al 1999, but it is absent in the 1.3mm and 0.8mm maps. We have extrapolated the flux measured at 0.45mm to 3mm and corrected for different resolutions, assuming an emissivity law of $`\nu ^2`$. We predict a 3mm counterpart of 0.1mJy/beam for the dust outflow in NGC 253, well beyond the noise level in our map (0.44mJy/beam) and a factor of $``$200 weaker than the 3mm extensions of P96, that reveal as spurious. ### 3.2 The SiO and H<sup>13</sup>CO<sup>+</sup> emission Figure 2 shows the velocity-channel maps obtained in the (v=0,J=2–1) line of SiO at the centre of NGC 253; figure 3 shows the same but in the (J=1–0) line of H<sup>13</sup>CO<sup>+</sup>. Gas emission is unevenly distributed in the velocity range \[120-353kms<sup>-1</sup>\]. This arises from a circumnuclear disk (CND) extending over 600 pc$`\times `$150 pc. The disk, seen at high inclination (i=78.5), is radially resolved and displays an overall rotating pattern. Although the dominant trend is circular rotation, we also find emission at highly non-circular velocities in both tracers. The emission is patchy in both lines, most noticeably, the SiO and H<sup>13</sup>CO<sup>+</sup> clumps show no one-to-one correspondence over the map: as discussed below (section 7), the T<sub>mb</sub>(SiO)/T<sub>mb</sub>(H<sup>13</sup>CO<sup>+</sup>) ratio has a wide range of variation in the nucleus of NGC 253 (i.e. T<sub>mb</sub>(SiO)/T<sub>mb</sub>(H<sup>13</sup>CO$`{}_{}{}^{+})`$0.5–8)). The CND also shows structure along the apparent minor axis. The velocity-integrated intensity maps of SiO (I<sub>SiO</sub>) and H<sup>13</sup>CO<sup>+</sup> (I$`_{H^{13}CO^+}`$), shown in Figures 1b-c, illustrate the CND substructure. Spatial coordinates are arcsecond offsets along the principal axes of the near infrared (NIR) stellar bar (Scoville et al 1985). The bar major axis (denoted x) is at PA=68, with x$`>`$0 eastwards (the bar minor axis, denoted y, lies along PA=158, with y$`>`$0 northwards). Figure 1b shows that the SiO emission mostly comes from two-nested rings centered at (0,0) and roughly oriented parallel to the x axis. A similar morphological description applies to H<sup>13</sup>CO<sup>+</sup> map (Figure 1c), although here the inner ring orientation deviates significantly from the bar major axis. The inner ring (I), of average diameter D$`{}_{I}{}^{}`$8$`\mathrm{}`$(120pc), contains 60-70$`\%`$ of the total flux in the two lines and it corresponds to the region of the massive starburst in NGC 253 seen in different tracers. The emission of the outer ring (II), of average diameter D$`{}_{II}{}^{}`$24$`\mathrm{}`$(360pc), contains 20-30$`\%`$of the total flux. The two nested rings share as common center the peak of the 3mm continuum source. Emission in I is detected within a large velocity range for the two edges of the ring ($``$120kms<sup>-1</sup>). This spread is symmetrical with respect to the continuum source locus (a similar symmetry is seen to hold for II). We therefore identify it as the dynamical center of NGC 253 ($`\alpha _{J2000}`$=$`00^h47^m33.18^s`$ and $`\delta _{J2000}`$=-$`25^{}17^{}17.2^{\prime \prime }`$). This position coincides, within the errors, with the dynamical center determined using optical (Watson et al 1996), infrared (Sams et al 1994, Boker et al 1998) and radio data (Turner and Ho, 1985). Although most of the emission arises from the double ringed source (I-II), we have also detected gas outside the x axis: the CND structure along the minor axis is partly resolved. Although part of the emission can be accounted by the gas distribution in the CND disk seen in projection (NGC 253 is not fully edge-on), at places, the emission lies far away from the NIR bar major axis, and most noticeably, this corresponds to gas clouds showing highly non-circular motions and a high relative abundance of SiO (T<sub>mb</sub>(SiO)/T<sub>mb</sub>(H<sup>13</sup>CO<sup>+</sup>)=3-5)). In particular, the SiO emission channels of Figure 2 in the range \[284-335kms<sup>-1</sup>\] show receding gas (v-v<sub>sys</sub>=80kms<sup>-1</sup>; where v<sub>sys</sub>=235kms<sup>-1</sup>) in the SE quadrant, where it should be approaching. As will be discussed below, the origin of this anomalous component is related with the giant outflow detected in X-ray and optical emission lines. This does not have the same origin as I-II. However, there is a H<sup>13</sup>CO<sup>+</sup> clump at (x,y)$``$(1,8) which shows no SiO counterpart. Contrary to the anomalous component, this might correspond to disk gas seen in projection. This clump has a low ratio T<sub>mb</sub>(SiO)/T<sub>mb</sub>(H<sup>13</sup>CO<sup>+</sup>)$``$0.3, and most important, it lies close to the minor axis of the galaxy appearing at velocities close to v<sub>sys</sub>. ## 4 The CND structure To study the CND structure we have obtained the peak brightness flux distributions in the two lines (B<sup>peak</sup>(x,y)=max\[T<sub>mb</sub>\]<sub>v</sub>(x,y)). This allows us to filter out the lower spatial frequencies enhancing the most intense features in the maps. Then we deprojected B<sup>peak</sup> onto the galaxy plane, in order to get a sharp picture of the CND, viewed face-on. The result is shown in Figures 4a-b. The peaks of emission of ring I are rather aligned with the major axis of the disk, tilted 20 clockwise with respect to the bar major axis. The tilt is more pronounced for H<sup>13</sup>CO<sup>+</sup> (Figure 4a). On the contrary, the peaks of emission of ring II open out, go across the line of the bar major axis and delineate a spiral-like arc. The face-on picture of the CND of Figure 4a consists of a gaseous spiral-like ridge (II) ending up in an unresolved ring I viewed edge-on and aligned with the major axis of the disk. Interferometer observations by Canzian et al (1988) have shown that the bulk of molecular gas in the CND, traced by the 1–0 line of <sup>12</sup>CO, is distributed in a bar-like source, oriented in the direction of the NIR bar. The recent <sup>13</sup>CO map of Hüttemeister and Aalto (1998) confirms this view. In contrast, the distribution of dense gas (n(H<sub>2</sub>)$`>`$10<sup>5-6</sup>cm<sup>-3</sup>) is similar to the B<sup>peak</sup> maps of SiO and H<sup>13</sup>CO<sup>+</sup>. The CN(1–0) and HNC(1–0) maps of Hüttemeister and Aalto (1998) show a centrally condensed inner disk of diameter $``$8$`\mathrm{}`$, oriented along PA=45-50, while the outer disk tilts farther towards PA=70, delineating a spiral-like arc that goes across the major axis of the stellar bar. The same applies for the HCN(1–0) map of Paglione et al 1995: the inner disk is better aligned with the disk major axis PA=45, whereas the outer ring tilts progressively towards PA=70. There is little doubt that NGC 253 contains a stellar bar oriented along PA=68 and extended 300$`\mathrm{}`$ on the sky, which sets a lower limit for its corotation radius at R<sub>COR</sub>=150$`\mathrm{}`$ (Scoville et al 1985). P96 first explained the distribution of the CS(2–1) emission in terms of a bar potential. However the parameters of the bar were chosen arbitrarily: the bar corotation is placed at 40$`\mathrm{}`$, a factor of $``$4 too short, compared with the estimated bar semi-major axis ($``$150$`\mathrm{}`$). The NIR image of the bar is distorted by dust extinction. There are two dust lane ridges offset with respect to the bar major axis. Assuming that the NW is the near side of NGC 253 (De Vaucouleurs 1958), the two offset dust lanes inside the bar represent a trailing spiral wave. This pattern is an indication of the existence of two Inner Lindblad Resonances (outer=oILR and inner=iILR; see Athanassoula (1992)). The morphology of the gas response in the presence of two ILRs has been analysed by several authors (see Buta and Combes 1996 and references therein). Owing to dissipation by cloud-cloud collisions, the gas cannot follow the periodic orbits of the stars in a bar (x<sub>1</sub>=parallel, between corotation and oILR or x<sub>2</sub>: antiparallel, between the oILR and the iILR). The gas response always leads the stellar orbits, giving a trailing spiral inside the bar corotation that goes across the oILR. It is tempting to identify the spiral-like II as the oILR. The gas compression along the spiral ridges should be best traced, as observed, by the higher density tracers of molecular gas. In contrast, the nuclear spiral across the oILR is washed out in the <sup>13</sup>CO and <sup>12</sup>CO maps tracing the low or moderately dense molecular gas (n(H<sub>2</sub>)=10<sup>3-4</sup>cm<sup>-3</sup>). Moreover, the proposed resonance loci of the bar naturally explain the double-ringed appearance of NGC 253 (Figures 1bc), and how the nuclear starburst might have been onset. When a barred galaxy has two ILRs, as NGC 253, gas tends to accumulate at the iILR, as a consequence of secular evolution. The engine behind the radial inflow of gas is double. Gravity and viscous torques cooperate in bringing the gas closer to the nucleus. Gravity torques act on the gaseous spiral expected to form between the iILR and oILR. The spiral is a superposition of leading and trailing waves. The leading wave dominates at the crossing of the iILR, whereas the trailing wave develops at the crossing of the oILR (Buta and Combes 1996). The gravity torques change their sign at the crossing of the resonances, and also when we move from the trailing to the leading spiral (Combes 1994a, 1994b). The net result would be the formation of two pseudo-rings at the iILR and the oILR, with similar time-scales. However, viscous torques being always negative, may break this symmetry making the gas migrate towards the iILR in a few dynamical times ($``$10<sup>7-8</sup>years) and therefore progressively depopulate the oILR. The oILR may end up as a trailing spiral-like pseudo-ring (ring II in NGC 253), whereas the leading spiral vanishes into the inner ring (similar to ring I). It seems that the morphology of the SiO/H<sup>13</sup>CO<sup>+</sup> maps and how they compare to other molecular gas tracers can be explained by gas accumulation at the ILRs of the bar in NGC 253. Arnaboldi et al 1995, used the H$`\alpha `$ major axis kinematics to make an independent estimation of the bar resonances loci. Within the epicyclic approximation, they identified two ILRs in the nucleus of NGC 253. Assuming a lower limit for the bar corotation, R<sub>cor</sub>=3.9kpc, they derived an oILR at 300pc (20$`\mathrm{}`$) and the iILR near the center, in agreement with what is proposed in this work. Studying the gas kinematics in the SiO/H<sup>13</sup>CO<sup>+</sup> CND will allow to test the location of the principal resonances and understand the relation between the starburst and the bar. ## 5 Kinematics of the CND There is a wealth of published work on the dynamics of the ISM in the nucleus of NGC 253. The determination of a rotation curve (v<sub>rot</sub>) for the nucleus of NGC 253 has been very controversial because of extinction (Ulrich 1978). The high-sensitivity H$`\alpha `$ data of Arnaboldi et al 1995, already showed a velocity gradient more symmetrical and steeper than seen in previous optical surveys. Figures 5a-b show the SiO and H<sup>13</sup>CO<sup>+</sup> position-velocity diagrams taken along the kinematical major axis of the disk of NGC 253 (x’ axis). In view of the high inclination of NGC 253, the terminal velocity method would be well suited to derive v<sub>rot</sub> from the major axis kinematics. However the patchiness of the SiO/H<sup>13</sup>CO<sup>+</sup> emission makes any fit uncertain. There are strong oscillations of the terminal velocities as a function of radius, revealing an uneven gas distribution and/or the presence of non-circular motions. As a zero-order approach we will adopt v<sub>rot</sub> derived by Canzian et al(1988) from CO. This species is the best tracer of molecular gas at a large-scale, from densities as low as n(H<sub>2</sub>)$``$10<sup>3</sup>cm<sup>-3</sup>, and although non-circular motions should also affect the observed kinematics, the estimated v<sub>rot</sub> would be closer to the axisymmetric value. The centroid of SiO/H<sup>13</sup>CO<sup>+</sup> emission in the major axis p-v plots follows closely v<sub>rot</sub>(CO), but most noticeably, the emission shows a large velocity spread around ($`\mathrm{\Delta }_v`$=100-120kms<sup>-1</sup>; see Figures 5a-b). The observed linewidths could be partly due to the smearing of v<sub>rot</sub> within our beam. Assuming i=78.5, we expect to get emission from clouds at y<sub>depr</sub>=$`\pm `$9-10$`\mathrm{}`$ (y<sub>depr</sub> is the deprojected distance from which emission is seen by our beam along y’). Along the major axis we get emission from x’=$`\pm `$1.6$`\mathrm{}`$. v<sub>rot</sub>(CO) follows a rigid body law with a velocity gradient of 6kms<sup>-1</sup>/$`\mathrm{}`$, from which we derive an upper limit for the total velocity broadening of $``$60kms<sup>-1</sup> within our beam, i.e., a factor of $``$2 lower than the observed velocity spread. The existence of a large velocity dispersion in the CND gas cannot be excluded (we would need $`\sigma _v`$40-50kms<sup>-1</sup> to explain the observed linewidths). However, an explanation of the SiO/H<sup>13</sup>CO<sup>+</sup> linewidths in terms of unresolved non-circular motions is more plausible as it is supported by additional observational evidence. Firstly, a part of the emission is detected at velocities that cannot be accounted by any circular rotation law. These regions in the major-axis p–v plots correspond to the (x’,v)-quadrants (x’$`>`$0,v$`>`$v<sub>sys</sub>) and (x’$`<`$0,v$`<`$v<sub>sys</sub>)). SiO/H<sup>13</sup>CO<sup>+</sup> gas emission fills unevenly a romboid-like parallelogram in Figures 5a-b. The symmetry of the above pattern is remarkable and it reminds of the signature of bar-driven elliptical orbits developed inside the ILR domain (Binney et al 1991; García-Burillo and Guélin 1995, Fux 1998, García-Burillo et al 1999). The parallelogram boundary would correspond to the inner non-selfintersecting x<sub>1</sub> orbit (called the cusped orbit). Inside, gas clouds lie along precessing x<sub>2</sub> orbits whose envelope delineate a spiral across the oILR; these orbits are highly elliptical and produce the observed non-circular motions. The velocity gradient is reversed for a SiO component in the major axis p-v plot between x’=-4$`\mathrm{}`$ and x’=4$`\mathrm{}`$. The latter was also reported by P96 in the CS p-v plot and and by Ananthamaraiah and Goss(1996) using recombination-line data. Apparent counter-rotation can also be explained by bar orbits (see Figure 8 of García-Burillo and Guélin 1995). The SiO minor axis position-velocity diagram, displayed in Figure 6a, suggests also that gas distribution in the nucleus deviates from axisymmetry. The cut along the minor axis is markedly asymmetrical: the centroid of redshifted gas lies 1$`\mathrm{}`$-2$`\mathrm{}`$ above the major axis locus, whereas blue-shifted gas lies 1$`\mathrm{}`$-2$`\mathrm{}`$ below. Ananthamaraiah and Goss(1996) reported the existence of a complex kinematic subsystem in the inner 150pc of NGC 253, using recombination-line data. The nucleus would host three gaseous disks; among the anomalous components, one exhibits rotation in a plane perpendicular to the galactic disk, and the inner disk is apparently counter-rotating. These components are introduced to explain deviations from circular motions similar to what is shown in our data. On the basis of this somewhat ad hoc decomposition, the authors find evidence of a past merger involving a small mass companion. However, the merger hypothesis is unable to explain the undisturbed morphology and kinematics of the outer disk (Olson and Kwan, 1990). The distorted kinematics of the NGC 253 nucleus, including the existence of strong non-circular motions detected also in the recombination-line data, are more readily explained in terms of a bar-driven gas flow. ## 6 A molecular gas counterpart of the NGC 253 nuclear outflow In the previous section we have argued that the bulk of the SiO/H<sup>13</sup>CO<sup>+</sup> emission in the CND fits within the bar-driven gas flow scenario. However, we reported in section 3 the detection of an anomalous component which might be far from the CND plane. The off-axis gas, mostly detected in the SiO line (T<sub>mb</sub>(SiO)/T<sub>mb</sub>(H<sup>13</sup>CO<sup>+</sup>=3-5), is best seen towards the south (down the plane of the CND). The southern SiO plume starts at the base of the starburst ring I and it reaches a height of $`\delta `$y=10$`\mathrm{}`$ (measured from the major axis of the bar; see Figure 4b). The gas shows also strong deviations from circular motion ($`\mathrm{\Delta }_v`$100-150kms<sup>-1</sup>) and its emission spreads over $``$200kms<sup>-1</sup> (see Figure 6b). There is a much weaker and point-like counterpart of the southern SiO-plume towards the North. The northern gas shows also strong noncircular motions ($`\mathrm{\Delta }_v`$-150kms<sup>-1</sup>) but it is here characterized by a narrow profile ($``$17kms<sup>-1</sup>). SiO spectra towards these positions are displayed in Figure 7. The Northern and Southern plumes (hereafter called N and S) can be connected by a line that goes across the dynamical center; the N-S line has a PA=131 measured from North (see Figure 4b). This orientation places the anomalous component close to the galaxy minor axis (PA=141) where no contribution from rotation curve in the radial velocity gradient is expected. Moreover it is hard to explain non-circular motions of $`\mathrm{\Delta }_v`$100-150kms<sup>-1</sup> due to bar forcing in the plane of the galaxy. There is compelling evidence for the existence of a giant outflow of gas in the nucleus of NGC 253. Demoulin and Burbidge(1970) and Ulrich (1978) were the first to report the presence of an ionized gas outflow, based on the analysis of gas kinematics derived from H$`\alpha `$ spectroscopy. Soft X-ray emission from the giant outflow was later discovered by Fabbiano and Trinchieri (1984). The X-ray nebula is more extended to the southern part of the galaxy minor axis and was interpreted as thermal emission of gas heated by fast shocks, driven by a bipolar wind coming from the starburst. Further support for this scenario is provided by the distribution of ionized gas in the outflow, which adopts the form of filaments that border the X-ray nebula (McCarthy et al 1987). Furthermore, Schulz and Wegner (1992) detected line-splitting in NII, SII and HII, consistent with the ionized gas lying across the conical working surface of the X-ray nebula. The gas entrainment by the hot wind may cause shock-heating and subsequent emission. The observed line-splitting, together with the inferred expansion velocities ($``$390kms<sup>-1</sup>), are the main arguments supporting the shock-heating scenario. Very recently, Alton et al 1999 have reported the tentative detection of submillimeter emission at 450$`\mu `$ coming from one filament associated with the S outflow. The location of the S filament agrees reasonably well with the one detected in SiO/H<sup>13</sup>CO<sup>+</sup> (see above and Figure 4b). The dust mass contained in the filaments, which partly depends on the dust temperature, cannot be easily estimated from the detection at just one wavelength. No counterpart of the dust outflow is seen neither at 1mm nor at 3mm, as expected from the flux detected at 450$`\mu `$ (see discussion of section 3.1; see also Krügel et al, 1990). However we can try to estimate the dust column density towards the S filament from the flux reported by Alton et al 1999, corrected to our beam. We assume a dust emissivity spectral index of $`\beta `$=2, the emissivity law is taken from Chini et al (1997), and adopt a dust temperature in the range T<sub>dust</sub>=13K-37K (similar to the model that fits the M82 dust outflow: Alton et al 1999). We calculate a range of N$`{}_{dust}{}^{}`$0.6–4$`\times `$10<sup>4</sup>M towards the S filament, within our beam. If M<sub>gas</sub>/M<sub>dust</sub>=100, this results implies N(H<sub>2</sub>)$``$0.6–4$`\times `$10<sup>6</sup>M/beam. As discussed in section 7, we can estimate N(H<sub>2</sub>) in the S filament where H<sup>13</sup>CO<sup>+</sup> is detected, using N(H<sup>13</sup>CO<sup>+</sup>) derived from a Large Velocity Gradient (LVG) transfer model and assuming X(H<sup>13</sup>CO<sup>+</sup>)=10<sup>-10</sup> (see discussion of section 7 for details). We calculate N(H<sub>2</sub>)$``$3.5$`\times `$10<sup>6</sup>M/beam (4.7$`\times `$10<sup>22</sup>cm<sup>-2</sup>), roughly in agreement with the value estimated from the dust emissivity. Therefore no contradiction exists between the failed detection of the S outflow in the continuum and the successful detection of the outflow in the two lines of SiO and H<sup>13</sup>CO<sup>+</sup>. If the molecular CND is locally disrupted and it is entrained by the hot wind out of the plane, we can expect that the geometries for the working surfaces of molecular and ionized gas should be alike. The parameters of Schulz and Wegner’s model are an expansion velocity of 390kms<sup>-1</sup> for the conical outflow coming out of the plane, and a half-opening angle of 25. Projection effects along the line of sight would make appear two velocities at each position of the outflow (N or S). The emission from the S(N) filament should appear either at v<sub>rad</sub>=-10kms<sup>-1</sup> (120kms<sup>-1</sup>) or at v<sub>rad</sub>=300kms<sup>-1</sup> (440kms<sup>-1</sup>). The velocities observed in the SiO lines for the N (v<sub>rad</sub>=100kms<sup>-1</sup>) and S filaments (300kms<sup>-1</sup>) confirm the predictions of Schulz and Wegner’s model, originally intended to fit the observed line-splitting in the optical lines. Turner (1985) was the first to report the existence of a molecular gas ejection event in the NGC 253 nucleus. However, the location of the OH plume hardly fits the morphology and kinematics of the ionized or molecular outflow: the OH gas is ejected from the nucleus towards the NE quadrant (along PA=6) and it reaches $``$1kpc height (see Fig 11 of Turner(1985) for details). We conclude that the distribution and kinematics of the off-axis component detected in our maps are easily accounted for if we suppose that the emission comes from molecular gas in an outflow leaving the nucleus, out of the galaxy plane. Furthermore, an estimate of the molecular gas mass involved in the outflow seems to be consistent with the dust emissivity of the filaments. ## 7 The T<sub>mb</sub>(SiO)/T<sub>mb</sub>(H<sup>13</sup>CO<sup>+</sup>) and N(SiO)/N(H<sup>13</sup>CO<sup>+</sup>) ratios The line temperature ratio R<sub>Tmb</sub>=T<sub>mb</sub>(SiO)/T<sub>mb</sub>(H<sup>13</sup>CO<sup>+</sup>) shows large variations (0.5–8) in the nucleus of NGC 253; these variations however show a systematic trend. For the sake of simplicity, we first examine the ratio of velocity-integrated emission R<sub>I</sub>=I(SiO)/I(H<sup>13</sup>CO<sup>+</sup>) averaged across those regions where R<sub>I</sub> shows a characteristic value. We have chosen as cases of study the dynamical center, regions I-II and the outflow components (N–S). Table 1 shows the mean values of R<sub>I</sub> and its range of variation; Figure 7 shows representative spectra for the regions studied. We see a clear trend in the ratios, which increase as one moves away from the vicinity of the starburst. The average ratios are R$`{}_{I}{}^{}`$1.0, for the center, R$`{}_{I}{}^{}`$1.1 for the starburst ring I, R$`{}_{I}{}^{}`$3 for the outer region II and R$`{}_{I}{}^{}`$2 for the outflow. We will analyse the range of variation of R<sub>Tmb</sub> for each region defined above, and derive the column density ratios N(SiO)/N(H<sup>13</sup>CO<sup>+</sup>) via a Large Velocity Gradient (LVG) scheme. Our objective is to study the range of physical parameters and chemical abundances of SiO clouds in the CND. LVG solutions depend on three parameters: the density (n(H<sub>2</sub>)), the kinetic temperature (T<sub>K</sub>) and the abundance of the species. The kinetic temperature of molecular gas in the center of NGC 253 has been derived by several authors using multitransition studies in <sup>12</sup>CO and its isotopes (<sup>13</sup>CO and C<sup>18</sup>O), CS and HCN; all indicate an average value of T$`{}_{k}{}^{}>`$50K. We have adopted T<sub>k</sub>=50K as representative for the SiO clouds and taken n(H<sub>2</sub>) from the fit to the ratios R<sub>32</sub>=I(SiO(3–2))/I(SiO(2–1)) and R<sub>54</sub>=I(SiO(5–4))/I(SiO(3–2)) derived by MP99, using single-dish 30m spectra. MP99 found little evidence of a density decrease for the SiO clouds, at least within the inner r$``$20$`\mathrm{}`$ of the nucleus. They derived an average n(H<sub>2</sub>)=5$`\times `$10<sup>5</sup> assuming a common filling factor for the three lines. We will take this estimation as representative for SiO and H<sup>13</sup>CO<sup>+</sup> in our model. Note that although there might be a hotter component in the nucleus (T$`{}_{k}{}^{}`$100K; see Mauersberger et al 1990), we must stress that n(H<sub>2</sub>) inferred from the SiO line ratios depend weakly on the adopted kinetic temperature if T$`{}_{k}{}^{}>`$50K. Figure 8 displays the range of LVG solutions that fit R<sub>32</sub> and R<sub>54</sub> within the errors. The assumed density n(H<sub>2</sub>)=5$`\times `$10<sup>5</sup> fits satisfactorily R<sub>32</sub> and R<sub>54</sub>, for a large interval of N(SiO)/$`\mathrm{\Delta }_v`$. Trying to delimit the LVG solution along the N/$`\mathrm{\Delta }_v`$ axis is equivalent to choosing a beam filling factor ($`\eta _{fill}`$) for the clouds. Whereas it is justified to assume that the two lines share a common $`\eta _{fill}`$, its exact value in an external galaxy is unknown a priori. It is worth discussing to what extent our conclusions might depend on $`\eta _{fill}`$. We have run LVG models for n(H<sub>2</sub>)=5$`\times `$10<sup>5</sup>cm<sup>-3</sup> and T<sub>k</sub>=50K, fitting R<sub>Tmb</sub>=T<sub>mb</sub>(SiO)/T<sub>mb</sub>(H<sup>13</sup>CO<sup>+</sup>) in the different regions of the CND. Table 1 fully explores the dependence of N(SiO)/N(H<sup>13</sup>CO<sup>+</sup>) on $`\eta _{fill}`$; the filling factor is varied within its whole range (0.005–1). The main conclusion, independent of $`\eta _{fill}`$, is that the SiO abundance is significantly enhanced in the outer CND (region II), where we derive N(SiO)/N(H<sup>13</sup>CO<sup>+</sup>)=5–15, and contrary to what it might be expected, it is lower at the center and at the starburst ring I, where N(SiO)/N(H<sup>13</sup>CO<sup>+</sup>)=1–2. The SiO abundance rises in the outflow, where we infer N(SiO)/N(H<sup>13</sup>CO<sup>+</sup>)=3–5. A density fall-off with radius has been reported by Wall et al 1991 (and confirmed by Harrison et al 1999). They concluded that n(H<sub>2</sub>) might reach $``$5$`\times `$10<sup>4</sup>cm<sup>-3</sup> for r$`>`$20$`\mathrm{}`$, not far from region II. If we take n(H<sub>2</sub>)$``$5$`\times `$10<sup>4</sup>cm<sup>-3</sup>, we find that N(SiO)/N(H<sup>13</sup>CO<sup>+</sup>)=20-100 for II. Opacity in the 2–1 line of SiO is close to $``$1 for this solution, whereas $`\tau `$0.02 for H<sup>13</sup>CO<sup>+</sup>. To fit the ratio of line temperatures we need imperatively to rise N(SiO)/N(H<sup>13</sup>CO<sup>+</sup>) in region II. Consequently, conclusions on the SiO abundance enhancement of region II remain unchanged, if not clearly favoured, in the lower density scenario. From the ratio N(SiO)/N(H<sup>13</sup>CO<sup>+</sup>), we can infer the absolute abundance of SiO (X(SiO)=N(SiO)/N(H<sub>2</sub>)), if we adopt a canonical value for H<sup>13</sup>CO<sup>+</sup> ($`<`$X(H<sup>13</sup>CO<sup>+</sup>)$`>`$). It is expected that SiO will be more sensitive than H<sup>13</sup>CO<sup>+</sup> to any variation of the chemical environment in NGC 253. This hypothesis is successfully confronted to the measurements of abundances of H<sup>12</sup>CO<sup>+</sup> obtained in a large variety of molecular clouds in our Galaxy where values close to 10<sup>-8</sup> are obtained in all cases (from which we derive X(H<sup>13</sup>CO<sup>+</sup>)$``$10<sup>-10</sup> if we adopt <sup>13</sup>C/<sup>12</sup>C$``$1/90). Though there is observational and theoretical evidence that X(H<sup>12</sup>CO<sup>+</sup>) may be reduced to 10<sup>-10</sup> in the hot cores of GMCs, SiO is lower by a similar factor in these sources (see Blake et al 1987); moreover, very compact sources are not expected to dominate the emissivity of H<sup>13</sup>CO<sup>+</sup>(1–0) at the scale we are observing the nucleus of NGC 253. Contrary to X(H<sup>13</sup>CO<sup>+</sup>), X(SiO) is measured to vary by several orders of magnitude in the Galaxy, depending on the type and the location of molecular clouds in the disk (see discussion). We can estimate $`<`$X(H<sup>13</sup>CO<sup>+</sup>)$`>`$ in the nucleus of NGC 253. From our data we obtain a column density of $`<`$N(H<sup>13</sup>CO<sup>+</sup>)$`>`$=3$`\times `$10<sup>12</sup>, averaged within the CND. Similarly, we can calculate $`<`$N(H<sub>2</sub>)$`>`$ from the <sup>12</sup>CO(1–0) map of Canzian et al (1988), averaged on the same area. If we take the conversion factor x=N(H<sub>2</sub>)/I<sub>CO</sub>=3$`\times `$10<sup>20</sup>cm<sup>2</sup>Kkm<sup>-1</sup>s, it follows $`<`$N(H<sub>2</sub>)$`>`$3$`\times `$10<sup>22</sup>, which implies global abundances of $`<`$X(H<sup>13</sup>CO<sup>+</sup>)$`>`$10<sup>-10</sup> in NGC 253, quite similar to the value inferred for our Galaxy. Column 5 of Table 1 illustrates the variation of X(SiO) in the CND of NGC 253. We derive X(SiO)$``$1–2$`\times `$10<sup>-10</sup> for the center and the starburst ring I, X(SiO)$``$3–15$`\times `$10<sup>-10</sup> for region II, and X(SiO)$``$3-5$`\times `$10<sup>-10</sup> for the outflow. The global abundance of SiO in the CND would be $`<`$X(SiO)$`>`$1.5$`\times `$10<sup>-10</sup>, namely, an order of magnitude above the typical value for a PDR (Janssen et al 1995, Walmsley et al 1999). Most notably, X(SiO) reaches the highest abundance in the outer region II, where the chemical processing of the gas by the nuclear starburst, mostly adscribed to region I, should be minor. The SiO abundance is also significantly enhanced in the outflow, certainly powered by the starburst, but not directly on it. This result emphasizes that SiO emission is not always directly related to the dense gas near sites of recent star formation. Therefore, different mechanisms must be explored to account for the different chemical processing of molecular gas within the nucleus. ## 8 Discussion We have established the existence of three distinct regions in the nuclear region of NGC 253, based on morphological, kinematical and chemical criteria. Region I dominates the global SiO emission and it is associated with the nuclear starburst taking place across $``$150pc. It presents a ring-like pattern, interpreted as the gas response near the iILR of the bar. The SiO abundance in the starburst (X(SiO)$``$1–2$`\times `$10<sup>-10</sup>) is significantly larger than that estimated for a typical PDR (X(SiO-PDR)$``$10<sup>-11</sup>). Region II extending up to the map edges (r$``$300pc(20$`\mathrm{}`$)), displays a spiral-like morphology also present in other high-density tracers. The observed distribution and kinematics of molecular gas are interpreted as the signature of the bar oILR. The orbit crowding across the oILR coincides with the detection of strong non-circular motions ($`\mathrm{\Delta }`$v$``$50–100kms<sup>-1</sup>) and, most notably, with an enhancement in the SiO abundance (X(SiO)$``$a few 10<sup>-9</sup>). The third region constitutes the molecular counterpart of the nuclear gas outflow observed in X-ray and optical lines. This was recently identified as a dust chimney. Two filaments (N and S) come out of the plane of the NGC 253 nucleus from r$`\pm `$60pc, near a working surface where the gas is entrained by the outflow. Again, we have found here a link between the existence of high-velocities ($`\mathrm{\Delta }`$v$``$100–150kms<sup>-1</sup>) and an SiO enrichment (X(SiO)$`>`$3–5$`\times `$10<sup>-10</sup>). Two processes are thought to dominate the chemistry of the molecular material associated with a starburst: the photodissociation of the gas by intense UV field produced by OB associations, and the existence of strong shock waves. Shocks can be generated locally during mass loss episodes in young stars (often identified as bipolar outflows). Alternatively, shocks may take place on a larger scale, either produced by supernova explosions or related to spiral/barred density waves. Contrary to X(H<sup>13</sup>CO<sup>+</sup>), X(SiO) is measured to vary by several orders of magnitude in the Galaxy, depending on the type and the location of molecular clouds in the disk: star forming clouds (X(SiO)$``$10<sup>-7</sup>–10<sup>-8</sup>), galactic center clouds apparently not forming stars (X(SiO)$``$10<sup>-9</sup>) or quiescent clouds (X(SiO)$`<`$10<sup>-12</sup>). It is well established that shocks occurring in bipolar outflows can enhance the abundance of SiO in the gas phase to reach, in the most extreme cases, $``$10<sup>-7</sup> (Martín-Pintado et al 1992). Although the starburst region I might be described as a giant PDR (Carral et al 1994), the relatively large abundances of SiO measured and the high densities of SiO clouds indicate that most of the emission arises in bipolar outflows powered by young massive stars, similarly to what is observed in the galactic complexes W51 and W49. Moreover, the particularly large ratio of I(\[SiII\])/I(\[OI\])$``$1 measured in the inner region of NGC 253 indicates an enhancement of Si in gas phase. This was interpreted by Carral et al 1994 as evidence of grain destruction in the starburst region, most likely by shocks related to massive star formation. The case for shock processing is even stronger in region II and in the outflow where X(SiO)$``$(50–100)$`\times `$X(SiO-PDR). However, the origin of the large abundance of SiO in region II is less clear. SiO emission could arise partly in bipolar outflows powered by young massive stars like in region I. The major drawback to this explanation is that region II is located away from the nuclear starburst. Therefore one should not expect an enhancement of SiO when the chemical processing caused by star formation has declined. Alternatively, the SiO enhancement could be produced by large-scale shocks in the molecular gas, caused by the crossing of clouds orbits near the oILR of the bar. The only evidence for chemical enrichment of gas attributed to large-scale shocks is found in the Galactic Center where values of $``$10<sup>-9</sup> have been measured in SiO clouds (Martín-Pintado et al 1997, Hüttemeister et al 1998). However there is no theoretical consensus whether the shocks can be caused by the action of density waves, especially in molecular clouds. The hierarchy of fragmentation inhibits shocks in molecular clouds; the mean-free paths for cloud-cloud collisions are of the same order of magnitude as the thickness of the disturbing potential well (Casoli and Combes 1982, Combes and Gerin 1985). If the starburst scenario gives rise to SiO emission in bipolar outflows of young massive stars, the differences in the abundance of SiO in regions I and II might be explained if the starbursts are in a markedly different stage of evolution. The older starburst would occur in region I, where a large fraction of the young massive stars are already on the main sequence. The latter explains why the classical tracers of the starburst phenomenon are mostly restricted to I. In contrast, region II would contain a younger and less-evolved starburst, where most of the young massive stars would be still in the mass loss phase, characteristic of the pre-main sequence stage, where energetic bipolar outflows occur. Finally, the interpretation of the morphology and kinematics of the molecular gas filaments give additional support to the outflow scenario. The enhancement of SiO in the gas phase observed in this region can be explained by the chemical processing of grains by the strong high-velocity shocks (v$`>`$200kms<sup>-1</sup>) associated to the outflow. ###### Acknowledgements. This work has been partially supported by the Spanish CICYT under grant number PB96-0104.
warning/0001/astro-ph0001107.html
ar5iv
text
# The variability of the Seyfert galaxy NGC 2992: the case for a revived AGN ## 1 Introduction The Seyfert 1.9 galaxy NGC 2992 is a nearby (z=0.0077), edge on ($`i=70^{}`$) system interacting with NGC 2993. The detection of a broad component of H$`\alpha `$ with no H$`\beta `$ counterpart (Ward et al. 1980; Shuder 1980) implied an intermediate Seyfert type classification, suggesting the existence of an obscured broad line region (BLR). This was confirmed later by the infrared detection of a broad Pa$`\alpha `$ line (Rix et al. 1990; Goodrich, Veilleux & Hill 1994). The optical image of NGC 2992 (Ward et al. 1980) shows a prominent dust lane extending along the major axis of the galaxy and crossing the nucleus. The optical extinction, measured from the Balmer decrement of the narrow H$`\alpha `$ and H$`\beta `$ components, is $`A_V=4.2\pm 0.4`$ (Ward et al. 1980). Finally, evidence for coronal lines of \[FeVII\] and \[FeX\] have been found in the optical spectrum by Shuder (1980) and Winkler (1992). The X–ray flux and spectrum were observed to vary during 16 years of observations performed with different satellites. As noted by Weaver et al. (1996; hereafter We96), since the time of HEAO–1 observations in 1978 (Mushotzky 1982), the 2–10 keV flux of NGC 2992 decreased by a factor of $`20`$ until the ASCA observations in 1994, when it reached the lowest value. A similar decrease was also observed in the soft X–rays (We96). Before ASCA, the spectral shape exhibited some variations but on average it could be described as an absorbed ($`N_\mathrm{H}10^{22}`$ cm<sup>-2</sup>) power law with photon index $`\mathrm{\Gamma }1.7`$ and an iron K$`\alpha `$ line. The ASCA spectrum was however much different, with a very flat slope ($`\mathrm{\Gamma }1.2`$). These spectral variations have been interpreted by We96 as the evidence for radiation reprocessed by an optically and geometrically thick gaseous medium, e.g. the molecular torus invoked by unified models (Antonucci & Miller 1985). We96 measured a time lag between the Fe line and the continuum of 10$`\pm `$4 years, from which they estimate a distance of the torus from the central engine ($`3.2`$ pc, that however is still model-dependent). Flux variability in the infrared was found by Glass (1997), who in particular remarked a fading of the source from 1978 to 1996 apart from a strong outburst in 1988. All of this evidence suggests that before BeppoSAX observed NGC 2992 in 1997 and 1998, the AGN in the center was turning off. In this paper we present an analysis of new BeppoSAX observations of NGC 2992 which show that the AGN is back to its previous high level of activity observed in 1978. The X–ray data are also complemented with optical and near–IR spectra obtained in a followup observation. The results of our analysis are compared with previous data in the literature. We assume $`H_0=50`$ km s<sup>-1</sup> Mpc<sup>-1</sup> yielding a distance of 46 Mpc for NGC 2992. ## 2 X–ray data ### 2.1 New observations and data reduction The four co–aligned Narrow Field Instruments (NFI) on board the Italian–Dutch satellite BeppoSAX (Boella et al. 1997a) span the broad X–ray energy range from 0.1 to 200 keV. The NFI include the Low Energy Concentrator Spectrometer (LECS; Parmar et al. 1997) and the Medium Energy Concentrator Spectrometer (MECS; Boella et al. 1997b), which both consist of a mirror unit plus a gas scintillation proportional counter and have imaging capabilities. The LECS instrument covers the energy range 0.1–10 keV with a spatial resolution of about $`1`$ arcmin and a spectral resolution of 8% at 6 keV. The MECS instrument operates in the 1.8–10 keV band with the same spatial and spectral resolution of the LECS, but is two times more sensitive in the overlapping spectral range. The other NFI are a High Pressure Gas Proportional Counter (HPGSPC; Manzo et al. 1997) and a Phoswich Detector System (PDS; Frontera et al. 1997), which are direct–view detectors with rocking collimators. The HPGSPC (4–120 keV) has better energy resolution but is less sensitive than the PDS (13–200 keV). In this paper we will only take advantage of the LECS, MECS and PDS data. Errors, unless otherwise stated, will be given at the 90% confidence level for one interesting parameter ($`\mathrm{\Delta }\chi ^2=2.71`$). NGC 2992 was observed by BeppoSAX in 1997 (from December 1 to December 3) and in 1998 (from November 25 to November 27). The 1997 data and the 1998 data will be referred to as SAX1 and SAX2, respectively. For the SAX1 observation we will not take into account the LECS data due to poor statistics. Table 1 shows exposure times and net count rates (i.e. background subtracted) for both BeppoSAX observations. The MECS spectra for SAX1 and SAX2, and the LECS spectrum for SAX2 have been extracted with an aperture radius of 4 arcmin. The background subtraction files have been extracted with the same apertures from blank sky fields. The spectral data have been binned in order to have at least 20 counts per channel to validate the use of the $`\chi ^2`$ statistics. A possible source of contamination in the X–ray spectrum detected by BeppoSAX could be NGC 2993, which lies at about 2.9 arcmin from its companion NGC 2992. NGC 2993 has an optical magnitude comparable to that of NGC 2992 ($`\mathrm{\Delta }m=0.03`$) and also shows optical narrow emission lines and radio emission (Ward et al. 1980). On the basis of the optical lines, NGC 2993 has been classified as a LINER (Durret 1990). Due to the uncertain nature of LINERs, which could host both thermal and non-thermal activity, NGC 2993 could contribute to the flux measured by BeppoSAX. However, since at 1415 MHz NGC 2993 is four times less luminous than NGC 2992 (Ward et al. 1980), and radio emission should not suffer from absorption, NGC 2993 should be intrinsically less luminous than NGC 2992 and its contribution to the observed X–ray spectrum should be small in every X–ray energy band from LECS to PDS. Also, the ROSAT PSPC data show that in the soft X–rays the flux of NGC 2993 is less than 5% of the NGC 2992 flux (We96), and an analysis of the ASCA SIS image of the NGC 2992 field has revealed that the emission of NGC 2993 is negligible also in the 2–10 keV band. Therefore, we consider the X–ray photons detected by BeppoSAX as entirely produced by NGC 2992. ### 2.2 Timing analysis The flux of NGC 2992 was constant throughout the SAX1 observation in 1997 (Fig. 1). When a fit with a constant is performed to the SAX1 light curve, the result is $`\chi ^2/dof=9.4/14`$ for a binning time of 10 ks. On the contrary the source exhibited short term variability in 1998 during the SAX2 observation, when it was in the high state: the source varied by a factor of $`1.4`$ on timescales of $`7\times 10^4`$ s. A constant count rate is ruled out at $`>99\%`$ confidence level by the $`\chi ^2`$ test both for the LECS and MECS data. The variability amplitude of NGC 2992 during the SAX2 observation was estimated by means of the normalized “excess variance” $`\sigma _{rms}^2`$ (see Nandra et al. 1997a for the definition of this quantity). In order to allow a comparison with the results of Nandra et al. (1997a) we have calculated $`\sigma _{rms}^2`$ with a binning time of 128 s, which also ensures Gaussian statistics. For the 2–10 keV lightcurve $`\sigma _{rms}^2`$ turned out to be $`9.7(\pm 0.3)\times 10^3`$. This value is typical of Seyfert 1s with the same intrinsic 2–10 keV luminosity of NGC 2992 (Nandra et al. 1997a). We note that the value of $`\sigma _{rms}^2`$ depends on the duration of the observation and that the exposures times for the Seyfert 1s sample of Nandra et al. (1997a) are all shorter than $`60`$ ks, the MECS exposure time in SAX2. Therefore, in order to be consistent with the average exposure of the Nandra et al. (1997a) sample, we have considered different time intervals of $`30`$ ks (first and second half of the observation and a central interval). We obtain a mean $`\sigma _{rms}^2`$ of $`6.7(\pm 0.2)\times 10^3`$, which still agrees with that observed in Seyfert 1s of comparable luminosity. In order to check for spectral variations associated with the SAX2 short term variability, we have analyzed the ratio between the hard and soft X–ray photons (Hardness Ratio, HR) in different bands. When the HR is calculated comparing LECS to MECS data, one could not exclude some degree of spectral variability. However, HR variations are not detected when the HR is calculated only from MECS data. As an example (see Fig. 2), the HR between the MECS data in the 3.5–10.5 keV and 1.8–3.5 keV is constant at a high confidence level ($`\chi ^2/dof=6.1/11`$). We note that the cross calibration between the LECS and MECS data is somewhat uncertain (see the next Section), and it might exhibit time variations. Therefore, we will consider the dispersion in the HR between the LECS and MECS data as an instrumental effect and extract only a time averaged spectrum. ### 2.3 Spectral analysis SAX1 (1997) The spectral analysis was performed by using the XSPEC 10.0 package. The MECS data in the 1.8–10.5 keV band and the PDS data in the 14–200 keV band were fitted simultaneously. A relative normalization factor was introduced between the PDS and MECS spectrum since the BeppoSAX instruments show some mismatches in the absolute flux calibration (the MECS is the best calibrated instrument). The energy ranges covered by the MECS and PDS do not overlap and so we cannot determine a cross calibration constant from the data. Instead, for the fitting procedure, we assumed a fiducial PDS to MECS normalization factor of 0.8 (released by the BeppoSAX Science Data Center). When fitting with an absorbed power law, strong residuals emerge at 6–7 keV suggesting a prominent iron line in the NGC 2992 spectrum. With the addition of a Gaussian iron line to the absorbed power law model we obtain a good fit to the data with $`\chi ^2/dof=149/144`$. The fit is shown in Fig. 3 and spectral parameters are quoted in Table 2 together with the hard X–ray flux and luminosity derived by the best fitting model. Spectral parameters are quoted in the rest frame. The power law spectral index ($`\mathrm{\Gamma }=1.72\pm 0.12`$) is consistent with the typical value 1.7–1.9 for Seyfert galaxies (Nandra & Pounds 1994; Nandra et al. 1997b). The line broadening ($`\sigma _{\mathrm{K}\alpha }=0.18_{0.14}^{+0.10}`$ keV) is not highly significant (it is detected at $`\stackrel{>}{_{}}90\%`$ confidence level, $`\mathrm{\Delta }\chi ^2=3.21`$ with respect to a narrow fit with $`\sigma _{\mathrm{K}\alpha }=0`$ keV). On the other hand, the energy of the Fe line, $`E_{\mathrm{K}\alpha }=6.62\pm 0.07`$ keV, is significantly higher ($`>99.9\%`$ confidence level, $`\mathrm{\Delta }\chi ^2>21`$) than the value of 6.4 keV commonly observed in Seyfert galaxies (e.g. Nandra et al. 1997b; Turner et al. 1997) and expected from neutral or weakly ionized iron (below FeXVI). We note, however, that both the MECS2 and the MECS3 units (the MECS1 failed in May 1997) might be affected by an instrumental energy shift towards the blue of 0.05 and 0.08 keV at 6.6 keV respectively (SAX Cookbook: $`\mathrm{http}://\mathrm{www}.\mathrm{sdc}.\mathrm{asi}.\mathrm{it}/\mathrm{software}/\mathrm{cookbook}/\mathrm{cross}\underset{¯}{}\mathrm{cal}.\mathrm{html}`$). We have verified the significance of our results by calculating the confidence contours for the line energy versus the line width (shown in Fig. 4) and versus the photon index (not shown). Even applying a red shift of 0.05 or 0.08 keV to correct for the instrumental shift we find that $`E_{\mathrm{K}\alpha }>6.4`$ keV at $`>99\%`$ confidence level. Therefore, evidence for iron emission from a highly ionized medium is provided by the data. The equivalent width of the line is very high, $`EW_{\mathrm{K}\alpha }700`$ eV and cannot be easily explained with current emission line models for a direct view of the nucleus (see Section 4). We note that the PDS data are very noisy and do not allow to constraint significantly any reflection component in addition to the power law continuum. SAX2 (1998) The spectral data from the LECS, MECS and PDS were fitted simultaneously in the 0.1–3.5 keV, 1.8–10.5 keV and 14–200 keV bands, respectively. As for the SAX1 observation, the PDS to MECS normalization was fixed to 0.8. The LECS to MECS normalization was determined by fitting simultaneously the LECS and MECS data in the overlapping band 1.8–3.5 keV; the obtained value was subsequently assumed for the spectral analysis. The procedure provides a model–independent LECS to MECS normalization of $`0.64\pm 0.02`$. The typical values observed so far range from 0.65 (Malaguti et al. 1999) to 0.76 (Haardt et al. 1998; Guainazzi et al. 1999). Fit with a power law and a Gaussian line A simple absorbed power law model is unacceptable at $`>97\%`$ confidence level ($`\chi ^2/dof=431/374`$) mostly because of the strong residuals at 6–7 keV due to the iron emission. The addition of a broad Gaussian iron line improves significantly the fit ($`\mathrm{\Delta }\chi ^241`$). The spectral parameters and fit are shown in Table 2 and Fig. 5, respectively. Some residuals at 15–20 keV could suggest that a reflection component is present. However, including a reflection continuum, obtained by means of the pexrav model in XSPEC (Magdziarz & Zdziarski 1995), is not strongly required by the data ($`\mathrm{\Delta }\chi ^21`$ with the addition of one $`dof`$). The relative normalization $`R`$ between the reflected and the direct continuum turns out to be low, $`R=0.12`$ ($`R1`$ for a reflecting material covering $`2\pi `$ of the source) and the steepening of the spectral index, which should be observed when adding a reflection continuum, is only $`1\%`$.<sup>1</sup><sup>1</sup>1We verified that the relative normalization of the reflection component is low ($`R=0.17`$) also in the case of a ionized reflector, obtained by means of the pexriv model in XSPEC (Magdziarz & Zdziarski 1995). Similar to the SAX1 observation, the iron line peak is higher than 6.4 keV (although with a lower significance), while the line is definitely broad, ($`\sigma _{\mathrm{K}\alpha }=0.32_{0.06}^{+0.18}`$ keV) suggesting an emission from a relativistic accretion disc (Fabian et al. 1989). The confidence contours between the line energy and width for the SAX2 data are shown in Fig. 6. The line equivalent width is $`EW_{\mathrm{K}\alpha }=147\pm 37`$ eV, which may be easily explained by accretion disc models where the matter is mildly ionized (Matt et al. 1992). The observed line energy peak above the value of 6.4 keV expected from neutral matter could be due to the Doppler shift in a relativistic disc inclined with respect to the line of sight. Disc line model Because of the above mentioned considerations, we performed fits with line profiles derived from relativistic disc models, under the assumptions of both a Schwarzschild and a Kerr metrics, i.e. non rotating and rotating black hole, respectively (see Tab. 3). Since the number of line parameters is relatively high and most of them would be poorly constrained in the fits, we fixed all the line parameters but the inclination angle $`\theta `$ of the disc and the line normalization. The line was assumed to be cold ($`E_{\mathrm{K}\alpha }=6.4`$ keV). In the Schwarzschild metric we assumed for the inner disc radius $`R_i=6R_G`$ (where $`R_G=GM/c^2`$ is the gravitational radius and $`M`$ is the black hole mass), which corresponds to the last stable orbit for an accretion disc around a Schwarzschild black hole. The outer disc radius was fixed to $`R_o=1000R_G`$ and the line emissivity was parameterized by $`R^q`$, where $`R`$ is the radius. The power law index $`q`$ was assumed equal to 2.5, the mean value derived by Nandra et al. (1997b) for a sample of Seyfert 1 galaxies in the case of a Schwarzschild black hole. In the Kerr metric we assumed $`R_i=1.23R_G`$ (the last stable orbit in a rotating black hole metric), $`R_o=400R_G`$, and $`q=2.8`$, the average value derived from Nandra et al. (1997b) for a Kerr black hole. Fitting the iron line with relativistic disc models gives $`\chi ^2/dof=389/372`$ for the Schwarzschild metric and $`\chi ^2/dof=390/372`$ for the Kerr metric. Both models provide an adequate description of the data, comparable with the analytical broad Gaussian model. Remarkably, the inclination angle of the disc turns out to be the same both from the fits with the Schwarzschild and the Kerr metric: $`\theta =46^o\pm 7`$. Therefore, the disc seems to be inclined and the energy of the line centroid can be explained by the relativistic Doppler shift of a cold line. The estimate of disc inclination depends on the choice of inner radius and so we tested different values of $`R_i`$ for the Schwarzschild and the Kerr metric. In both cases, the best fit is found when the inner radius correspond to the last stable orbit. The 90% upper limits on $`R_i`$ are $`30R_G`$ and $`5.3R_G`$ for the Schwarzschild and the Kerr metric, respectively. If we calculate the values of $`\theta `$ when $`R_i`$ is fixed to its 90% upper limit, we find $`\theta =65_{21}^{+25}`$ for the Schwarzschild metric and $`\theta =44_5^{+4}`$ for the Kerr metric. Also, if the disc is ionized, the line centroid could be higher than 6.4 keV. Then, the Doppler boosting effect could be lower and the inclination estimate could be incorrect. We have therefore performed disc line fits fixing the line energy peak at 6.62 keV, i.e. the SAX1 line energy. The resulting values for $`\theta `$ are $`36_{10}^{o+7}`$ and $`38_7^{o+6}`$ for a Schwarzschild and a Kerr black hole, respectively, fully consistent with $`46^o\pm 7`$ derived previously for neutral disc models. Line blend model An alternate representation of the broad iron line feature could be a blend of lines emitted by iron at different ionization stages. We performed a fit to the iron complex with two narrow Gaussians at 6.4 keV and 6.96 keV as expected from FeI–XVI and FeXXVI, respectively. This model provides an equally good fit to the data as for the broad Gaussian line fit ($`\chi ^2/dof=391/372`$). Adding a 6.96 keV line describes the data better than having only the Fe K$`\beta `$ component at 7.06 keV. The latter is not sufficient to explain the data at $`7`$ keV. The addition of a third line at 6.68 keV expected from FeXXV is not significant. ### 2.4 Comparison with previous results Before the BeppoSAX observations, the X–ray flux of NGC 2992 decreased quite systematically from 1978 to 1994, when it was observed by ASCA. The flux measured in SAX1 shows a modest increase with respect to the ASCA value, while a factor of $`12`$ increase is detected between SAX1 and SAX2. The 2–10 keV flux measured by SAX2 is similar to that measured by HEAO–1 in 1978. The long term flux variability is shown in Fig. 7. The spectrum of NGC 2992 measured by SAX1 and SAX2 is an absorbed power law with $`\mathrm{\Gamma }1.7`$ and $`N_\mathrm{H}=10^{22}`$ cm<sup>-2</sup> and is on average consistent with the results previously obtained by HEAO–1, Einstein, EXOSAT and Ginga, but not with the very flat ($`\mathrm{\Gamma }1.2`$) ASCA spectrum in 1994. The peculiar spectrum detected by ASCA has been interpreted by We96 as evidence for a strong reflection continuum superposed to a weak power law continuum with $`\mathrm{\Gamma }1.7`$. The reflection continuum should represent the echo of the past brightness of the source. No significant evidence for short term variability was observed in the ASCA data (We96). An example of a similar flux and spectral variation has been observed in NGC 4051 (Guainazzi et al. 1998), where a drop out of a factor of $`20`$ in the 2–10 keV flux was accompanied by a spectral change from a power law with $`\mathrm{\Gamma }1.9`$ to a pure reflection continuum. The iron line in the BeppoSAX spectrum is significantly different from that detected by ASCA. We96 found that the line in the ASCA spectrum is cold ($`E_{\mathrm{K}\alpha }=6.42\pm 0.04`$ keV), narrow ($`\sigma _{\mathrm{K}\alpha }<0.06`$ keV) and has an equivalent width of $`EW_{\mathrm{K}\alpha }=514\pm 190`$ eV. Because of the line narrowness and its high equivalent width, We96 argued in favor of iron emission from a distant torus, that must be characterized by a large column density to produce the observed iron line and the cold reflection that flattens the ASCA spectrum. The line energy observed by Ginga in 1990 (Nandra & Pounds 1994) is $`6.43\pm 0.16`$ keV, consistent with neutral iron emission. Because of the poor energy resolution of the Ginga LAC the line width was not determined. No useful information about the line energy and width can be extracted by HEAO–1 data (Weaver, Arnaud & Mushotzky 1995). A summary of the variations of the iron line parameters along all the X–ray observations of NGC 2992 is shown in in Fig. 7. ## 3 Optical and near–IR data As soon as we received the last set of BeppoSAX data and realized the enhanced X–ray activity with respect to the previous BeppoSAX observation, we submitted a target of opportunity proposal to the European Southern Observatory to observe the behavior of various nuclear components such as the broad line region and the hot dust emission. As discussed in this Section the observation were performed in early January 1999, i.e. about 1.5 months after the BeppoSAX observation. ### 3.1 New observations and data reduction Near Infrared and optical observations of NGC 2992 were performed at the ESO New Technology Telescope (NTT). The infrared data were collected on January 7, 1999 with photometric conditions using the imaging and spectroscopic modes of SOFI, equipped with a Hawaii HgCdTe 1024x1024 detector. Images in the J, H and Ks filters were obtained with the small field objective yielding a 0$`\stackrel{}{.}`$144/pix scale. The seeing during the observations was $`0\stackrel{}{.}7`$. Observations consisted of several on-source exposures interleaved with sky exposures while stepping the object on the detector by $`20\mathrm{}`$. The total on-source integration time was of 200 s in J and 120 s in H and Ks. Flat-field correction was performed using differential dome exposures and the sky was subtracted from each on-source frame. Images were then co-aligned by using stars present in the field of view and coadded. Flux calibration was performed using the standard star SJ9138 (Persson et al. 1998). Spectral observations were performed with the Red and Blue grism yielding a dispersion of 10 Å/pix in the 1.52–2.52 $`\mu `$m range and 7 Å/pix in the 0.95–1.64 $`\mu `$m range. The pixels subtended 0$`\stackrel{}{.}`$276 along the 1″ slit and the resolving power was $`600`$ at 2.2 $`\mu `$m and $`500`$ at 1.25 $`\mu `$m. Observations consisted of several couples of exposures in which the object was moved at two different positions along the slit, in order to perform sky subtraction. On–chip integration time was 120 s resulting in a total, on-source integration time of 960 s. Data were flat fielded with spectroscopic dome exposures and sky subtracted. Correction for optical distorsion of the slit direction and wavelength calibration were performed with Xenon and Neon arc exposures. The zero point of the wavelength calibration was then checked using the OH sky lines (Oliva & Origlia 1992). Residual sky emission was removed by fitting a linear polynomium along the slit. Correction for instrumental response and telluric absorption, and flux calibration were performed by using the G1V star HR3916. The intrinsic stellar features were then canceled by dividing for the sun spectrum according to the prescription described by Maiolino, Rieke & Rieke (1995). The FWHM (1″) of the point spread function along the slit was estimated by using the wings of the broad lines which are emitted from a spatially unresolved region. The final spectrum was then extracted in a $`1\mathrm{}\times 2\mathrm{}`$ aperture centered on the nucleus which corresponds to the slit width by 2xFWHM. The optical spectra were collected on January 12, 1999 with photometric conditions using EMMI in the Low Resolution Spectroscopy (RILD) mode with grating $`\mathrm{}6`$. The projected pixel size of the Tektronic 2048x2048 CCD was 0$`\stackrel{}{.}`$27 along the slit and 1.2 Å/pix along dispersion. The slit width was 2″ yielding a resolution of $`R700`$ at 6500 Å. The slit was centered on the galaxy nucleus with a position angle of 120 deg. The 6000-8400 Å spectrum was obtained by merging two different exposures of 1200 s in order to remove cosmic rays. Data reduction was performed using standard procedures for CCD long–slit spectra, i.e. bias subtraction, flat field correction, rectification of the slit direction and wavelength calibration. Correction for instrumental response and flux calibration were performed using the spectrophotometric standard star LTT3864 (Hamuy et al. 1992, 1994). The final spectrum of the nuclear region was extracted, with the same procedure adopted for the IR spectra, in a $`2\mathrm{}\times 1\stackrel{}{.}4`$ aperture. ### 3.2 Broad lines analysis Fig. 8a shows the optical spectrum of NGC 2992 obtained with EMMI. Fig. 8b shows the continuum subtracted zoom of the spectrum around H$`\alpha `$, that clearly shows the presence of a prominent broad component of H$`\alpha `$. For comparison Fig. 8b also shows the nuclear spectrum obtained by Allen et al. (1999) in March 1994, i.e. at the minimum activity ever observed in the X–rays. The latter spectrum was scaled to match the intensity of the \[SII\] doublet in our spectrum. Although Allen’s spectra have slightly lower spectral resolution ($`550`$) their spectrum clearly shows a much weaker, if any, broad component of H$`\alpha `$. In contrast, our spectrum is remarkably similar to the spectrum obtained by Veron et al. (1980) prior to 1979<sup>2</sup><sup>2</sup>2The date of the observation is not given in the Veron et al. paper., i.e. probably when the nucleus was at the same level of X–ray activity as in late 1998 (Sect. 2). We deblended the broad and narrow components of H$`\alpha `$ by using a multiple Gaussian fit. The narrow components are unresolved at our spectral resolution. Some relevant results of the fit are reported in Tab. 4 (additional information is given in the Appendix) and are compared with the values obtained by Veron et al. (1980) in the former high activity phase. The ratio between broad and narrow H$`\alpha `$ are nearly identical for the two observations. Also Allen et al. spectrum seems to require a broad H$`\alpha `$ component of about the same width as that found in our spectrum, but about 10 times weaker. The optical spectrum also shows a variety of other lines that are marked in Fig. 8a. The analysis of flux variation of some coronal lines (such as \[FeVII\]$`\lambda `$6087Å, \[FeX\]$`\lambda `$6374Å, \[FeXI\]$`\lambda `$7892Å and \[SiVI\]$`\lambda `$1.96$`\mu `$m) with respect to the past would provide interesting constraints on the size of the Coronal Line Region, that is expected to be intermediate between the BLR and the NLR. Unfortunately, the quality of previous spectra is too poor to provide useful constraints to the flux of such coronal lines. However, in the Appendix we provide the fluxes of all the lines that are detected and identified in our spectrum and that could be useful for comparison with future monitoring. Fig. 9 shows the infrared spectrum taken with SOFI at ESO, where we combined the blue and red spectra. The spectrum shows a wealth of features, most of which are observed for the first time in this galaxy thanks to the much improved sensitivity of this observation with respect to previous data. In this paper we only discuss those features that are relevant to the issue of the burst observed in the X–rays. However, in the Appendix we report the list of the lines, along with their fluxes, that we detected and identified in the spectrum. Pa$`\beta `$ is the most suitable line for studying the profile of near–IR hydrogen lines, since it is relatively strong, isolated, and in a good atmospheric window. We fitted the Pa$`\beta `$ with two Gaussians (broad and narrow, unresolved) whose parameters were frozen to the same values obtained for H$`\alpha `$ except for their normalization. As reported in Tab. 4 the observed Pa$`\beta `$ can be almost fully accounted for by the broad component alone, while the narrow component contributes for only $``$7%, much lower than in the case of H$`\alpha `$. In Tab. 4 we also compare our result with those obtained by Rix et al. (1990) in 1989, when the nucleus was mildly active, and with the result of Goodrich et al. (1994) in May 1992, i.e. when the nucleus was close to its minimum activity. Although this comparison should be considered with caution because of different instrumental setup and lower sensitivity of past observations, there is an apparent correlation between the Pa$`\beta `$(broad)/Pa$`\beta `$(narrow) ratio and the level of nuclear activity as traced by the hard X–rays. ### 3.3 The hot dust emission The most interesting feature of the near infrared images (not shown) is a powerful nuclear emission characterized by very red colors. More specifically, within the central 2 arcseconds we measure $`\mathrm{H}\mathrm{K}=0.9`$ and $`\mathrm{J}\mathrm{K}=1.9`$, while typical galaxy colors are $`\mathrm{H}\mathrm{K}=0.2`$ and $`\mathrm{J}\mathrm{K}0.6`$. Nuclear dust absorption is probably partly responsible for such red colors, but very likely hot dust emission in the H and K bands also contributes significantly (see Oliva et al. 1995, Maiolino et al. 1998, Thatte et al. 1997 for similar cases in other objects). This interpretation is supported, in the specific case of our observations, by the near–IR spectrum: the stellar CO absorption bands above 2.29 $`\mu `$m and in the H band (e.g. at 1.62 $`\mu `$m) are much weaker than generally observed in normal galaxies. In particular, the equivalent width of the $`{}_{}{}^{12}\mathrm{CO}(20)`$ band is only 7.5 Å, while in normal galaxies this quantity generally ranges between 12 and 14 Å. This effect has been observed in other AGNs and is generally ascribed to hot dust emission diluting the stellar features in the near–IR (Oliva et al. 1995, Oliva et al. 1999). The burst observed by Glass (1997) in 1988 during his near–IR monitoring also had features typical of hot dust emission, indeed the burst had much higher power in the K and L bands than in H and J. In Tab. 5 we compare the magnitudes observed by us with the magnitudes measured by Glass at the maximum, in the same aperture, during the 1988 burst, and the background galaxy magnitudes that corresponds to the minimum values observed during the near–IR monitoring.<sup>3</sup><sup>3</sup>3With regard to the K band magnitudes, note that we used a filter (K-short) slightly different from that used by Glass, that is less sensitive to the hot dust emission at long wavelengths. The magnitudes measured by us are intermediate between the maximum and the minimum value observed. The physical implications are discussed in the next Section. ## 4 Discussion ### 4.1 Long and short term variability in the X–rays The data collected by BeppoSAX suggest that an AGN has revived in the center of NGC 2992. The 2–10 keV flux of the source systematically decreased by a factor of 19–20 between 1978 and 1994 and then increased by the same amount from 1994 to 1998. During the same time intervals the iron line flux changed by a factor $`3.5`$. Therefore, in 1998 both the continuum and the iron K$`\alpha `$ line intensity have settled back to the same values observed in 1978. This coincidence of fluxes could indicate that in the high state the luminosity of NGC 2992 is close to the maximum luminosity achievable, i.e. its Eddington limit. However further observations would be needed to confirm this hypothesis. The rebuilding of the AGN is also supported by the analysis of the short term variability of NGC 2992. In 1998, during the high state SAX2 observation, the source exhibited significant flux variations with an amplitude typical of Seyfert 1s (Nandra et al. 1997a), while in the low states observed in 1994 and 1997 no evidence for short term variability was found. The SAX2 variability is consistent with a scenario where we are observing the AGN directly through an absorbing medium. This situation is frequently observed in other objects with an intermediate Seyfert type classification in the optical and a column density of $`N_\mathrm{H}10^{22}`$ cm<sup>-2</sup> in the X–rays (e.g. NGC 526A, NGC 7314 and MCG –5-23-16; Turner et al. 1997). The lack of variability thoughout the SAX1 observation could suggest that during the rebuilding phase of the AGN the geometry of the central region and the accretion processes were different from those observed in SAX2. Ptak et al. (1998) found that Low Luminosity AGNs (LLAGNs) do not follow the Nandra et al. (1997a) relation, i.e. LLAGNs do not show time variability. These authors ascribed the behaviour of LLAGNs to the presence of an Advection Dominated Accretion Flow (ADAF; see Narayan, Mahadevan & Quataert 1998 for a review) in their nuclei. We note, however, that the X-ray luminosity of NGC 2992 during the SAX1 observation ($`1.8\times 10^{42}`$ erg s<sup>-1</sup>) is rather high compared with the typical luminosity of LLAGNs ($`10^{40}10^{41}`$ erg s<sup>-1</sup>). Furthermore, the ADAF scenario would hardly explain the prominent iron line detected in SAX1. Therefore, we favor the interpretation that ascribes the lack of variability in SAX1 to a distant and extended forming disc (and, possibly, an extended corona) rather than to an ADAF. Another possibility is that during the SAX1 observation the nucleus was in a transition state where both a central ADAF and a distant accretion disc were present. Since the long term variations amplitude of the iron line flux is different from that of the continuum, it is very likely that an accretion disc is not the only source of the line photons. Indeed, the disc is close to the X–ray primary source and it should reprocess the primary radiation on very short timescales of minutes or hours. Thus, one should expect long term variations of the disc line flux identical to those of the continuum. An iron emission from a distant medium, such as the obscuring torus postulated by the unified model, which lags the direct flux by timescales of years is therefore very plausible. We96 derived that the torus must have a geometry and distance from the nuclear source such that it produces a time lag of $`10`$ yr in the variation of the Fe line with respect to the continuum. Although this is likely to be the case for the narrow line detected by ASCA, no constraints are present on the line width detected by HEAO–1 in 1978. We note that if a disc component is present in the HEAO–1 data, the estimate of We96 is likely to be a lower limit on the torus distance. ### 4.2 Iron line and reflection continuum diagnostics in SAX1 The SAX1 observation could have caught NGC 2992 during the rebuilding phase of the AGN. The spectral continuum is similar to that of normal AGNs, while the iron feature is peculiar and cannot be easily explained in the framework of iron emission from an accretion disc or a molecular torus (Ghisellini, Haardt & Matt 1994). Indeed, the line is narrow in contrast with standard disc models. Furthermore, the line energy ($`E_{\mathrm{K}\alpha }=6.62\pm 0.07`$ keV) is not consistent with 6.4 keV as expected from the neutral iron in the torus. Such a narrow and warm line could be produced by ionized iron in a warm scattering material via resonant scattering (Matt, Brandt & Fabian 1996). A problem with this model is related to the observed line equivalent width of $`700`$ eV. The equivalent width of the resonant lines predicted by Matt et al. (1996) can be as high as a few keV if the nuclear source is obscured and only the scattered continuum observed. This is the case of NGC 1068 (Ueno et al. 1994; Matt et al. 1997) which is a Compton thick AGN. However, when the nuclear source is directly viewed, as in the case of NGC 2992, the equivalent width of the resonant lines is predicted to be at most $`100`$ eV (Krolik & Kallman 1987). Another possible explanation of the line properties in SAX1 is to assume a blending between a narrow cold component from the molecular torus and a narrow hot component from a distant ionized disc which is forming. The narrowness of the line can be explained by assuming that the accreting material which is going to rebuild the disc is still far from the central black hole and so the line does not suffer from strong Doppler or gravitational broadening. We have therefore fitted the SAX1 line data with two narrow Gaussians at 6.4 keV and 6.7 keV, also adding a reflection continuum. In order to compare the intensity of the 6.4 keV line and of the reflection continuum with the ASCA values, we have fixed the spectral index to $`\mathrm{\Gamma }=1.7`$, as adopted by We96 for the ASCA data. We obtain a fit as good as the single Gaussian line fit; the equivalent widths and intensities of the two lines are respectively: $`A_{6.4\mathrm{keV}}=1.4(\pm 1)\times 10^5`$ photons cm<sup>-2</sup> s<sup>-1</sup>, $`EW_{6.4\mathrm{keV}}=120_{90}^{+100}`$ eV, $`A_{6.7\mathrm{keV}}=3(\pm 1)\times 10^5`$ photons cm<sup>-2</sup> s<sup>-1</sup>, $`EW_{6.7\mathrm{keV}}=350_{130}^{+160}`$ eV. The 6.4 keV line intensity is lower than the ASCA value, qualitatively in agreement with the idea that this is a fading reprocessed component (quantitatively, the fading of the Fe cold line depends on the detailed geometry of the torus). The comparison between the ASCA and SAX1 reflection continua is less straightforward since a disc contribution is expected in SAX1 in addition to the torus one. However, we note that in SAX1 data there is scope for including a torus reflection normalization lower than the ASCA value. Both the centroid energy and equivalent width of the 6.7 keV component could be explained by means of ionized disc models (Matt, Fabian & Ross 1993). The ionization structure of the disc is basically determined by the ionization parameter $`\xi `$ defined as following: $$\xi =4\pi F_\mathrm{h}/n_\mathrm{H},$$ (1) where $`n_\mathrm{H}`$ is the hydrogen number density of the disc and $`F_\mathrm{h}`$ is the hard X–ray flux impinging on the disc surface. When $`\xi `$ is of the order of $``$ a few hundreds (in units of erg cm s<sup>-1</sup>), the iron ionization state is mostly below FeXVI: the line energy is at $`6.4`$ keV and the equivalent width is $`150`$ eV as expected for cold discs. For $`\xi 2000`$ the emission is dominated by FeXXV: the line energy is close to 6.7 keV and the equivalent width is as high as 350–400 eV since most of the elements lighter than iron are fully stripped and do not absorb the iron line photons. Finally, when $`\xi 500010000`$ the bulk of the line is emitted by FeXXVI: the line energy is at 6.96 keV and the equivalent width has decreased to $`100`$ eV, since a large fraction of iron is now completely bare. Therefore, the 6.7 keV line in the SAX1 spectrum is consistent with emission from a ionized disc with $`\xi 2000`$. Following Matt et al. (1992), if we assume that the hard X–ray source (possibly an extended corona of relativistic electrons, Haardt & Maraschi 1991) is located at $`50R_G`$ above the disc plane, we can estimate a hydrogen density of $`n_\mathrm{H}10^{14}`$ cm<sup>-3</sup> for these stages of the rebuilding disc. ### 4.3 Iron line and reflection continuum diagnostics in SAX2 The SAX2 spectrum of the high state is consistent with the typical X–ray spectrum of AGNs, suggesting that the rebuilding of the AGN phenomenon is almost completed. The iron line profile in SAX2 is consistent with that expected from cold accretion disc models. The hard X–ray luminosity has increased by an order of magnitude with respect to the SAX1 data, and $`\xi `$ is expected to be of the order of $`200`$ as for cold discs. Therefore, the full-developed accretion disc in SAX2 should be about two order of magnitude denser than the forming disc observed in SAX1. If we assume that at the time of the ASCA observation the AGN was in a completely quiescent phase, we can consider the time interval between the ASCA and SAX2 observations as un upper limit to the timescale for the rebuilding of the disc. Also, the time interval between SAX1 and SAX2 provide a lower limit to the rebuilding timescale. As a result, the rebuilding timescale turns out to range from 1 to 5 years. Since the obscuring torus should reside at a distance of $`3.2`$ pc (i.e. 10 ly) from the central source (We96), in SAX2 (1998) any narrow iron line from the torus, in addition to the disc line, is expected to be weaker than that detected by We96 in 1994. We have verified this to be the case. Also, the reflection component produced by the torus is now completely diluted by the nuclear and disc continua. The best fit inclination angle for the disc, as estimated from the iron line profile, is $`\theta =46^o\pm 7`$. Provided that the accretion disc and the molecular torus are coplanar, the inclination angle measured from the disc line profile is in agreement with the unified schemes where intermediate Seyfert types could be active nuclei observed through the rim of the torus. The moderate X–ray absorption of $`N_\mathrm{H}10^{22}`$ cm<sup>-2</sup> could be indeed produced by the outer boundary of an optically thick torus. Alternately, the X–ray absorption could be produced by the gas associated to the dust lane observed in the host galaxy and crossing the nucleus. In line with this interpretation is the fact that the optical extinction inferred from the optical and infrared narrow lines is close to that expected from the gaseous column density derived from the X–rays (for a Galactic gas-to-dust ratio). The fact that that the extinction measured toward the BLR is similar to the extinction affecting the NLR is also in favor of this intepretation. The nature of the absorber will be discussed more in detail in the next section. We note that NGC 2992 could be one of the first sources where there is evidence for two distinct absorbers in different spatial regions: the first (along our line of sight) is associated to the interstellar gas extended on the 100 pc scale, and the second (out of the line of sight) is associated to the thick molecular torus on parsec scales. This phenomenology has been recently observed also in the Seyfert 1.8 galaxy NGC 1365 (Risaliti, Maiolino & Bassani 1999). In SAX2 one problem related to the AGN scenario with an accretion disc model is that the reflection continuum produced by the disc and associated to the iron line is weaker than expected. The relative normalization between the direct and reflected continuum is $`R=0.12`$, which is in contrast with the value of 1 predicted by an isotropic source illuminating a plane–parallel semi–infinite slab and suggested by the line equivalent width. If we introduce a high–energy spectral cut off to $`E_c=150`$ keV, the minimum e-folding energy value to maintain a statistically good fit, then $`R`$ increases only to 0.4, which is still a low value. Two possibilities to solve this problem are to introduce an iron overabundance or to assume that we are observing the source through a dual absorber with column densities of $`10^{2223}`$ cm<sup>-2</sup>. The combination of a dual absorber, a canonical power law with $`\mathrm{\Gamma }1.9`$, and a reflection component could indeed mimic a flatter power law with $`\mathrm{\Gamma }1.7`$. In this scenario the line photons could be produced both in reflection by the accretion disc and in transmission by the dual absorber. Examples of complex absorbers are found in the X–ray spectrum of NGC 2110 (Malaguti et al. 1999) and NGC 5252 (Cappi et al. 1996). In Section 2 we showed that another representation of the iron line complex of SAX2 could be a blending between lines emitted by different iron ions. In particular, two narrow lines at 6.4 keV and at 6.96 keV, deriving from the low and highly ionized iron respectively, adequately fit the data. However, this scenario seem to be disfavoured since the intensity of the cold line at 6.4 keV is higher than expected. Indeed, the narrow cold line at 6.4 keV in SAX2 should be the signature of the emission from the torus and, therefore, should be weaker than that measured by ASCA, just because the nuclear continuum had been falling until 1994 and there is a 10 years time lag between narrow-cold Fe line and continuum (We96). Actually, the 90% lower limit to the intensity of the narrow cold line measured in SAX2 (see Table 3) is higher than the 90% upper limit to the ASCA line intensity ($`3.5\times 10^5`$ photons cm<sup>-2</sup> s<sup>-1</sup>) derived by Turner et al. (1997). ### 4.4 The BLR and the hot dust emission Unfortunately, the optical and near-infrared follow up performed by us cannot provide tight constraints to the size of the Broad Line Region and of the hot dust emitting region because we do not know exactly the time of the X–ray burst; we only know that the transition occurred between December 1997 and November 1998. Nonetheless our data provide some interesting information on the geometry of the circumnuclear region. Reverberation studies have shown that the BLR has typical sizes $`\mathrm{R}_{\mathrm{BLR}}=0.02\mathrm{L}_{45}^{1/2}\mathrm{pc}`$ (Netzer 1990). NGC 2992 has a hard X–ray luminosity of $`\mathrm{L}_{210\mathrm{k}\mathrm{e}\mathrm{V}}=2\times 10^{43}\mathrm{erg}\mathrm{s}^1`$, that translates into a bolometric luminosity of about $`\mathrm{L}_{\mathrm{bol}}=2\times 10^{44}\mathrm{erg}\mathrm{s}^1`$ (Mulchaey et al. 1994). Therefore the expected size of the BLR is about 0.01 pc. Our detection of broad lines both in the optical and in the near–IR spectra can only provide an upper limit to the size of the BLR, given by the fact that the burst cannot be older than about 1 yr at the time of the optical-IR observations and, therefore, $`\mathrm{R}_{\mathrm{BLR}}(\mathrm{NGC}2992)<0.3\mathrm{pc}`$ that is fully consistent with the expected value of 0.01 pc derived above. However, we note that reverberation studies are generally more sensitive to the inner region of the BLR, since this is the region where rapid line variations are better detected. Fig. 8 and Tab. 4 show that the broad lines detected by us are at the same level as those observed by Veron et al. (1980) during the former high level state prior to 1978. This implies that the whole BLR has already been fully ionized and, therefore, the upper limit of 0.3 pc does not simply apply to the average BLR radius or to the “cross-correlation” size (that is usually the quantity measured by reverberation mapping), but this upper limit applies also to the outer radius of the BLR. As discussed in the former Section both the IR spectrum and images clearly show evidence for hot dust emission. Although dust close to the sublimation limit (T$``$1500 K) has the highest emissivity at $`2\mu `$m, the increasing volume of emitting dust at lower temperatures implies that a large contribution to the K band emission comes also from regions at lower temperatures (down to 400 K), hence at larger distances. The K and H band fluxes observed by us are not at the maximum level observed by Glass (1997) during the 1988 burst<sup>4</sup><sup>4</sup>4This is not necessarily the maximum near–IR emission that can be achieved by the nucleus, since in 1988 the burst was temporary and did not last very long and, therefore, probably not all of the hot dust emitting region was completely formed., implying that in January 1999 the hot dust emission is still in the raising phase. The region of dust close to the sublimation limit (T$`>`$1000 K) was already heated. Instead, the regions with dust at temperatures around 500 K, that significantly contribute to the K band emission, were not active yet. According to the dust equilibrium temperature obtained by Laor & Draine (1993), dust at temperatures around 500 K should be located at a radius of about $`\mathrm{R}_{\mathrm{pc}}0.36\mathrm{L}_{46}^{1/2}\mathrm{T}_{1500}^{2.5}=0.8`$ pc, given the luminosity of NGC 2992, and therefore it is not surprising that we do not see the whole K band emission yet, considered that the burst was not older than 1 year at the time of the infrared observations. The dust close to the sublimation limit ($`\mathrm{T}1500`$ K) is expected to be located at a distance of $`0.05`$ pc, fully consistent with the upper limit of 0.3 pc that we can infer from the maximum time delay between the X–ray burst and the infrared observation. Generally, the sizes of BLR and hot dust region have been estimated independently for different objects. However, Netzer & Laor (1993) proposed that the relative dimensions of these two regions are related and, more specifically, that the outer boundary of the BLR is located just inside the sublimation radius of the dust grains. This model was proposed to explain the apparent gap between the BLR and the NLR with almost no line emission. Our optical and infrared followup supports this model. Indeed, as discussed above, the BLR appears to be fully formed when the hot dust region is still forming, consistent with the fact that the BLR outer radius is smaller than the hot dust emitting region. ### 4.5 Optical extinction Finally, we can use the ratio between the broad lines to estimate the absorption toward the BLR and compare it with the absorbing column density derived from our X–ray observations. As discussed in the previous Section and as shown in Tab. 4, in the near IR the hydrogen lines are completely dominated by the broad components. By assuming an intrinsic ratio 0.16 between Br$`\gamma `$ and Pa$`\beta `$, and the extinction curve in Rieke & Lebofsky (1985), we infer $`\mathrm{A}_\mathrm{V}(\mathrm{BLR})=2.2\pm 0.6\mathrm{mag}`$. Instead, by comparing Pa$`\beta `$ with the broad component of H$`\alpha `$, and assuming an intrinsic Pa$`\beta `$/H$`\alpha `$ = 0.052, we derive $`\mathrm{A}_\mathrm{V}(\mathrm{BLR})=3.5\pm 0.4\mathrm{mag}`$. However, the latter value is less reliable both because radiation transport and collisional excitation might affect the intrinsic Pa$`\beta `$/H$`\alpha `$ ratio in the BLR, and because problems of intercalibration between the optical and the infrared spectra might affect the observed value. The extinction toward the NLR can be derived by the narrow components of Pa$`\beta `$ and H$`\alpha `$: we obtain $`\mathrm{A}_\mathrm{V}(\mathrm{NLR})=2.0\pm 0.6\mathrm{mag}`$<sup>5</sup><sup>5</sup>5Ward et al. (1980) obtained a value of $`\mathrm{A}_\mathrm{V}(\mathrm{NLR})=4.2`$ from the narrow lines H$`\alpha `$/H$`\beta `$ Balmer decrement. Part to the difference with our result is to ascribe to a different extinction curve and an adopted intrinsic ratio lower than currently assumed for the narrow line region of AGNs (i.e. 2.8 instead of 3.0); if we use their narrow line Balmer decrement with our extinction curve and the “correct” intrinsic ratio we obtain $`\mathrm{A}_\mathrm{V}(\mathrm{NLR})=3.7`$. Moreover the H$`\beta `$ line is probably affected by the corresponding stellar H$`\beta `$ in absorption (Ward et al. spectrum shows evidence for stellar features) that artificially increases the Balmer decrement. Finally the limited spectral resolution might have resulted in problems with the emission lines deblending., consistent with what obtained for the broad line region. This finding suggests that both BLR and NLR are absorbed by the same (large scale) medium, possibly the 100 pc scale dust lane observed in the images. The optical-infrared absorption towards the BLR is lower than that expected from the $`N_H`$ measured from the X–ray spectrum. By assuming a Galactic gas-to-dust ratio we find $`\mathrm{A}_\mathrm{V}(\mathrm{X})=4.5\times 10^{22}\mathrm{N}_\mathrm{H}(\mathrm{cm}^2)=4.5`$. As suggested by Granato et al. (1997), this mismatch between optical and X–ray absorption might indicate the presence of a dust-free region in the X–ray absorber. Alternatively, the obscuring gas might be characterized by a gas–to-dust ratio lower than Galactic or the average dust grain size might be larger than in the interstellar Galactic medium, thus making the extinction curve flatter (Laor and Draine 1993). ## 5 Conclusions In this paper we have shown that an AGN has revived in the center of NGC 2992. The spectral continuum detected both in SAX1 (1997) and SAX2 (1998) can be described with a power law with $`\mathrm{\Gamma }=1.7`$ absorbed by $`N_\mathrm{H}=10^{22}`$ cm<sup>-2</sup>, which is consistent with the mean spectrum of NGC 2992 observed in the past. A prominent iron K$`\alpha `$ line is also detected in both the data sets. The SAX1 observation is located between the ASCA and the SAX2 observations (which correspond to the lowest and highest states of the source, respectively), and seems to have caught the source during the rebuilding phase of the accretion disc. From the properties of the iron line detected in SAX1, the forming disc appears to be constituted by gas which is still far form the central black hole and is highly ionized, possibly because of its low density. During the SAX2 observation the disc appears to be almost completed and the iron line properties fit into the standard (cold) disc models. We also presented the results of optical and near–IR followup observations obtained 1.5 months after the detection of the X–ray burst. The optical and near–IR hydrogen lines are characterized by prominent broad components, with intensities comparable to those observed in the former active phase prior to 1979. The near–IR images and the near–IR spectrum also show evidence for hot dust emission in the H and K bands. However, at the time of the observations the K and H band nuclear emission was still not as high as observed in a previous IR burst, suggesting that hot dust region was still in the process of forming. ###### Acknowledgements. We thank Chris Lidman, the NTT and SOFI staff for performing the near infrared and optical observations. We thank Roberto Della Ceca for helpful comments. We are grateful to M. Allen and M. Dopita for providing their optical spectrum of the nucleus of NGC 2992. The referee is thanked for his comments. This work was partly supported by the Italian Space Agency (ASI) under grant ARS–98–116/22 and by the Italian Ministry for University and Research (MURST) under grant Cofin98–02–32. ## Appendix A List of line fluxes In Tab. A1 we list the lines detected in our optical and near–IR spectra. In the case of permitted lines, the flux given in the last column gives the total line flux, i.e. both broad and narrow components. In the case of blending the line fluxes were obtained by means of a multiple Gaussian fit. In cases for which the blending is excessive to allow a deconvolution we give the total flux of the blend. We could not derive a flux for the HeI $`\lambda `$6678 since it is severely blended with the H$`\alpha `$ wings and with the \[SII\] doublet, resulting in a plateau between the two groups of lines. Some uncertainties exist for the identification of the line at 13305 Å : it could either be fluorescence OI $`\lambda `$13170Å or \[FeII\] $`\lambda `$13205Å. We favor the latter identification both because it matches much better the observed wavelength and because the observed flux is in excellent agreement with that expected from \[FeII\]$`\lambda `$12567Å (the observed ratio between the two lines is 3.26, while the theoretical value is 3.75). The absence of OI $`\lambda `$13170Å strongly suggests that the OI $`\lambda `$11290Å line is produced by Ly$`\beta `$ fluorescence and not by continuum fluorescence (in the latter case both OI lines should have nearly the same intensity). Ly$`\beta `$ fluorescence is also supported by the lack of OI lines at 7002 Å and 7255 Å . We could not find any identification for the line detected at 7431 Å , possibly this might be an instrumental artifact.
warning/0001/hep-th0001073.html
ar5iv
text
# References Non-commutative geometry arises naturally in string theory when the Neveu-Schwarz $`B`$ field is turned on. In a certain limit, the low-energy effective theory of the worldvolume is described by gauge theory in non-commutative space . Motivated by the attempt to investigate localized structures, some researchers have recently studied topological field configuration in non-commutative geometry, namely the non-Abelian monopoles . In this letter, we consider another elementary example, the Dirac monopole . It would be helpful to make some comparisons first. Non-Abelian monopole can be visualized as D-string stretched between branes . In the $`U(2)`$ case, when a background $`B`$ field is turned on along the branes, the string is tilted because the two endpoints carry opposite charges . This leads to a dipole structure in the magnetic field of the monopole. Explicit calculation of this effect to the $`O(\theta )`$ order has been carried out in . However, $`U(1)`$ monopole does not admit such a simple geometric picture. First of all, there is no need to introduce a Higgs field $`\varphi `$. The singularity is put in by hand: one simply adds a source term to the Bianchi identity, then the base manifold becomes $`^3\{0\}`$, which deformation-retracts to $`S^2`$. The Wu-Yang method is applied to yield the quantization of magnetic monopole charge ($`\pi _1(U(1))=`$). By contrast the topological invariant of a non-Abelian monopole is defined by the asymptotic behavior of the Higgs field. Second, the energy of Dirac monopole diverges, while that of ’t Hooft-Polyakov monopole is finite ($`m1/g_{\mathrm{YM}}`$). Although it’s difficult to make topological argument in non-commutative space, it has been shown that the BPS bound still exists . One naturally asks whether non-commutativity will render Dirac monopole a finite mass. This is interesting since all the attempts to find magnetic monopole have failed. We can even pose the question: why should one treat a “particle” with infinite mass as a physical entity? The case should be compared to that of electron. Although the field energy due to a point electric charge diverges in the same way, it’s renormalizable as electron has an experimentally measurable mass about half MeV (one simple way to get rid of the infinity is to replace the point source with a smooth compactly-supported charge, while the same procedure is not applicable to monopole ). If Dirac monopole became finitely heavy in non-commutative space, it would be of great interests to theorists and experimentalists alike. However, our calculation gives a negative answer. Up to the $`O(\theta ^2)`$ order, we show explicitly that there is no correction to the $`U(1)`$ gauge connection $`A`$ (except for $`\theta `$-dependent gauge transformation), therefore the mass remains infinite. We argue the same is true for any higher order. In the following, we will present our calculation by adopting a mixed notation of tensor and differential form. We also discuss another formulation of the $`U(1)`$ monopole in non-commutative space. Finally, we ponder over the nature of non-commutative spacetime and its relation to the cosmological constant. The field strength in non-commutative gauge theory is defined as usual with replacement of the ordinary product by the $``$-product<sup>1</sup><sup>1</sup>1 $`f(x)g(x)=\mathrm{exp}(\frac{i}{2}\theta _{ij}_i_j^{})f(x)g(x^{})|_{x=x^{}}`$ and $`[x_i,x_j]=i\theta _{ij}`$. $$F=dA\frac{i}{2}[A,A]_{}.$$ (1) Up to the second order $`O(\theta ^2)`$, $`F`$ in component form is<sup>2</sup><sup>2</sup>2We only consider static situation and set $`A_0=0`$ since there is no electric source. $$F_{ij}_iA_j_jA_i+\theta _{mn}_mA_i_nA_j,$$ (2) where $`i`$, $`j`$ run from 1 to 3. One notices there is no $`O(\theta ^2)`$ term. In fact, all the even powers of $`\theta `$ vanish because of its antisymmetric nature. This can also be seen from the definition of $`F`$: all the even powers contain an $`i`$, but $`F`$ is a real number. Then what do we mean by the second order correction? Of course presumably the gauge potential $`A`$ itself should receive some corrections due to non-commutativity. Expanding $`A`$ to $`AA^0+A^1+A^2`$ (the upper indices denote the order in $`\theta `$), we rewrite (2) as $`F_{ij}F_{ij}^0+F_{ij}^1+F_{ij}^2`$, where $`F_{ij}^0`$ $`=`$ $`_iA_j^0_jA_i^0`$ $`F_{ij}^1`$ $`=`$ $`_iA_j^1_jA_i^1+\theta _{mn}_mA_i^0_nA_j^0`$ (3) $`F_{ij}^2`$ $`=`$ $`_iA_j^2_jA_i^2+\theta _{mn}_mA_i^0_nA_j^1+\theta _{mn}_mA_i^1_nA_j^0.`$ For later convenience, we define $`f^{1,2}=dA^{1,2}`$. The goal is to find the solution to the modified Bianchi identity (in Gauss units) $$DF=4\pi g\delta ^3(\stackrel{}{r})1$$ (4) where $`g`$ is the monopole charge and $`1`$ is the volume form. The covariant derivative of $`F`$ in non-commutative space is similarly defined $$DF=dFi[A,F]_{}.$$ (5) Order by order in $`\theta `$, (4) is $`dF^0`$ $`=`$ $`4\pi g\delta ^3(\stackrel{}{r})1`$ $`dF^1`$ $`=`$ $`\theta _{mn}_mA^0_nF^0`$ (6) $`dF^2`$ $`=`$ $`\theta _{mn}_mA^1_nF^0\theta _{mn}_mA^0_nF^1.`$ The zeroth order equation is just the familiar $`\stackrel{}{B}^0=4\pi g\delta ^3(\stackrel{}{r})`$, with $`B^0=F^0`$. Its solution is simply $`\stackrel{}{B}^0=g\stackrel{}{r}/r^3`$, and $`A^0`$ has singularities (Dirac string) when expressed in a single coordinate system. One then solves for $`A^{1,2}`$ by plugging $`A^0`$ into the first and the second order equations. In component form, the first order equation reads $`ϵ_{ijk}_if_{jk}^1`$ $`=`$ $`ϵ_{ijk}_i(\theta _{mn}_mA_j^0_nA_k^0)ϵ_{ijk}\theta _{nm}_nA_k^0_mF_{ij}^0`$ (7) $`=`$ $`ϵ_{ijk}\theta _{mn}\left(_m_iA_j^0_nA_k^0+_mA_j^0_n_iA_k^0_nA_k^0_m(_iA_j^0_jA_i^0)\right).`$ The first and the third terms cancel each other; while shuffling the indices of the fourth term ($`ik`$, $`ji`$, $`kj`$ and $`mn`$) makes it cancel the second one. So the right hand side of (7) vanishes, i.e., $`df^1=0`$. This is a source-free equation with solution $`f^1=0`$ (because of the boundary condition at infinity), therefore we find that $`A^1`$ is a pure gauge. We proceed to the second order $`ϵ_{ijk}_if_{jk}^2`$ $`=`$ $`ϵ_{ijk}_i\left(\theta _{mn}(_mA_j^0_nA_k^1_mA_k^0_nA_j^1)\right)`$ (8) $`ϵ_{ijk}\theta _{nm}_nA_k^1_mF_{ij}^0ϵ_{ijk}\theta _{mn}_mA_k^0_nF_{ij}^1.`$ It’s easy to show that most of the terms in (8) cancel out by either shuffling the indices or directly setting $`A^1`$ to be zero. The only non-trivial term comes from the $`A^0`$ part in $`F^1`$. More precisely we are left with $$ϵ_{ijk}_if_{jk}^2=ϵ_{ijk}\theta _{mn}\theta _{pq}_mA_k^0_n(_pA_i^0_qA_j^0).$$ (9) The term on the right hand side is just $`2ϵ_{ijk}(\theta _{mn}\theta _{pq}_mA_k^0_qA_j^0_n_pA_i^0)`$ with the part in brackets symmetric in $`j`$ and $`k`$, so it vanishes identically. Once again, we are led to the usual Bianchi identity $`df^2=0`$, which shows $`A^2`$ is trivial up to a gauge transformation. One may argue that $`f^{1,2}`$ are closed from the previous definition. But we note $`f^{1,2}=dA^{1,2}`$ are only defined *locally*, just like in the commutative case $`F^0=dA^0`$ is not valid *globally* on $`S^2`$. A priori, it is not obvious the source terms due to $`A^0`$ in (7) and (8) should vanish. It is tempting to generalize the above result to higher order: the gauge potential $`A`$ does not receive any corrections at all. Since the mere existence of a non-trivial $`A`$ is due to the monopole at origin, while it’s hard to see why a physical quantity defined at a *single* point should be modified because of the spatial non-commutativity (this is also the reason why we do not put $`gg^0+g^1+g^2`$ in (4)). In consequence, the monopole mass still diverges. Explicitly, $`|A^0|1/r`$, $`|F^0|1/r^2`$, $`|F^1|1/r^4`$ and $`|F^2|=0`$, so $$m=|F|^2_0^{\mathrm{}}r^2𝑑r\left|\frac{1}{r^2}+\frac{1}{r^4}\right|^2=\mathrm{}.$$ (10) While this result is disappointing, there exists an alternative way to address the monopole problem because of the self-interaction of photons in non-commutative space . As already remarked, the singularity at origin is put in by hand and results in the infinite mass. Instead of introducing the singularity beforehand, one can try to seek finite energy solution to the equation of motion. In other words, we are not looking for a point monopole, but an extended field configuration just like in commutative non-Abelian gauge theory. Now one faces the question as whether a Higgs field should be introduced. In the brane picture, a non-trivial $`\varphi `$ describes the shape of the brane. But it’s not clear whether the topological charge depends on $`\varphi `$. A detailed analysis is beyond the scope of this short paper and may appear in a subsequent report. There is the important question about gauge invariance. Due to the $``$-product structure, the $`U(1)`$ field strength is not gauge invariant any more $`\delta F`$ $``$ $`[i\lambda ,F]_{}`$ (11) $``$ $`0\theta _{mn}_m\lambda ^0_nF^0(\theta _{mn}_m\lambda ^1_nF^0+\theta _{mn}_m\lambda ^0_nF^1),`$ where $`\lambda =\lambda ^0+\lambda ^1+\mathrm{}`$ is the gauge parameter. This makes it difficult to interpret what is the magnetic field generated by the monopole. For instance, consider an electron moving in the monopole background. Even on the classical level, one has to modify the Lorentz force formula $`\stackrel{}{f}=q\stackrel{}{v}\times \stackrel{}{B}`$ in order to make sure $`\delta \stackrel{}{f}=0`$ under $`\delta B=\delta F`$ (or equivalently it imposes constraints on the possible form of gauge transformation). This situation differs too much from the familiar physical facts, thus raises a basic question one bears in mind: what is the reality of spacetime non-commutativity? To achieve non-commutativity, one has to give a non-zero expectation value to the Neveu-Schwarz tensor field, hence the associated field energy (we do not consider the so-called *transverse non-commutativity* here). This adds to the total vacuum energy density $`\rho _\mathrm{V}`$ and therefore is related to the long-standing cosmological constant problem (for a review, see ). The non-commutativity scale is determined by the background field $`\theta =1/B`$. If we assume, for a time, the vacuum energy comes solely from $`B`$ field, we can calculate $`\theta `$ by using the observed value of $`\rho _\mathrm{V}`$. Assuming the energy density of $`B`$ takes the same form as that of magnetic field<sup>3</sup><sup>3</sup>3One needs to be a little cautious: although $`B`$ resembles a magnetic field, it is a two-form *potential*, not a *field strength* in string theory., we have $`\rho _\mathrm{V}`$ $`=`$ $`{\displaystyle \frac{1}{8\pi }}B^2={\displaystyle \frac{1}{8\pi \theta ^2}},`$ (12) which gives $$\sqrt{\theta }=\left(\frac{1}{8\pi \rho _\mathrm{V}}\right)^{1/4}\left(\frac{1}{8\pi \times 10^{47}\mathrm{GeV}^4}\right)^{1/4}5.0\times 10^3\mathrm{cm}.$$ (13) This numerical value is hardly believable as otherwise it would be verified easily by experiment (note a portion of $`\rho _\mathrm{V}`$ only makes the result worse). Of course the above simple calculation does not exclude the possibility that spacetime may be non-commutative on the fundamental level, since the cosmological constant itself has a 120 orders of magnitude contrast between theoretical estimation and experimental observation! Nevertheless, it indicates the nature of non-commutative spacetime is far from clear. A plausible test bed of non-commutativity is superconductivity. Unlike the monopoles, which largely remain to be theorists’ fond toys, the magnetic vortices have been observed in type II superconductors and the associated Abelian Higgs model has been well studied (see, for instance, ). Non-commutativity could show its signature via some modifications to the conventional flux lattice. In conclusion, we studied Dirac monopole in non-commutative space and found no non-trivial corrections. A more interesting solution may appear in an alternative formulation of the problem. In the process, we raised more questions than we could answer. While spacetime non-commutativity seems too remote from the real world, we have no doubts about this increasing abstraction in mathematics and physics as remarked in Dirac’s original paper . Acknowledgements We would like to thank Miao Li for bringing to our attention the paper . This work was supported in part by NSF Grant PHY9511632 and the Robert A. Welch Foundation.
warning/0001/math0001098.html
ar5iv
text
# Random unitaries in non-commutative tori, and an asymptotic model for q-circular systems ## Abstract We consider the concept of q-circular system, which is a deformation of the circular system from free probability, taking place in the framework of the so-called “$`q`$-commutation relations”. We show that certain averages of random unitaries in non-commutative tori behave asymptotically like a $`q`$-circular system. More precisely: let $`q`$ be in $`(1,1)`$; let $`s,k`$ be positive integers; let $`(\rho _{ij})_{1i<jks}`$ be independent random variables with values in the unit circle, such that $`\rho _{ij}=q`$, $`1i<jks`$; and let $`U_1,\mathrm{},U_{ks}`$ be random unitaries such that $`U_iU_j=\rho _{ij}U_jU_i`$, $`1i<jks`$. If we set: $$X_r:=\frac{1}{\sqrt{k}}(U_r+U_{r+s}+\mathrm{}+U_{r+(k1)s}),1rs,$$ then the family $`X_1,\mathrm{},X_s`$ behaves for $`k\mathrm{}`$ like a $`q`$-circular system with $`s`$ elements. The above result generalizes to the case when instead of the hypothesis “$`\rho _{ij}=q`$” we start with “$`\rho _{ij}=z`$”, where $`z`$ is a complex number such that $`|z|<1`$. In this case the limit distribution of $`X_1,\mathrm{},X_s`$ is what we call a z-circular system. From the combinatorial point of view, the new feature brought in by a $`z`$-circular system is that its description involves the enumeration of oriented crossings of certain pairings; it is only in the case when $`z=\overline{z}=q`$ that the orientations cancel out, allowing the $`q`$-circular system to be described via non-oriented crossings. As a consequence of the result, one can easily construct families of random matrices which converge in distribution to $`q`$-circular (or more generally $`z`$-circular) systems. 1. Introduction and statement of the results This section is divided into subsections as follows: 1.1 The concept of circular system. 1.2 $`q`$-circular systems. 1.3 An asymptotic model for $`q`$-circular systems. 1.4 Oriented crossings. 1.5 $`z`$-circular systems. 1.6 Refinements of Theorem 1.5.3. 1.7 Approximation with random matrices. 1.1 The concept of circular system was introduced by D. Voiculescu in , and plays an important role in his theory of free probability. The definition goes as follows. Let $`(𝒜,\phi )`$ be a $`C^{}`$-probability space – by which we mean that $`𝒜`$ is a unital $`C^{}`$-algebra and $`\phi `$ is a state of $`𝒜`$ ($`\phi :𝒜\text{C}`$ positive linear functional, such that $`\phi (I)=1`$). The elements $`c_1,\mathrm{},c_s𝒜`$ ($`s1`$) are said to form a circular system in $`(𝒜,\phi )`$ if the family $$\frac{c_1+c_1^{}}{\sqrt{2}},\frac{c_1c_1^{}}{i\sqrt{2}},\mathrm{},\frac{c_s+c_s^{}}{\sqrt{2}},\frac{c_sc_s^{}}{i\sqrt{2}}$$ (1.1) is free in $`(𝒜,\phi )`$, and if each of the selfadjoint elements listed in (1.1) has normalized semicircular distribution with respect to $`\phi `$. The fact that an element $`a=a^{}𝒜`$ has normalized semicircular distribution means by definition that $$\phi (a^n)=\frac{1}{2\pi }_2^2t^n\sqrt{4t^2}𝑑t,n1.$$ (1.2) For the definition of freeness in $`(𝒜,\phi )`$, we refer to , Chapter 2. The definition of a circular system given above can be rephrased in a purely combinatorial way, by indicating the general formula of the joint moments of $`c_1,c_1^{},\mathrm{},c_s,c_s^{}`$, i.e. of the expressions $$\phi (c_{r_1}^{\epsilon (1)}\mathrm{}c_{r_n}^{\epsilon (n)}),n1,r_1,\mathrm{},r_n\{1,\mathrm{},s\},\epsilon (1),\mathrm{},\epsilon (n)\{1,\}.$$ (1.3) Namely, it turns out that every number in (1.3) is a non-negative integer, which “counts a certain family of non-crossing pairings”; this statement will be made precise (and generalized) in Section 1.2 below. An important realization of a circular system, given in , uses creation and annihilation operators on a full Fock space: if $`𝒯`$ is the full Fock space over $`\text{C}^{2s}`$, if $`\xi _1,\xi _2,\mathrm{},\xi _{2s}`$ is an orthonormal basis of $`\text{C}^{2s}`$, and if $`l_1,l_2,\mathrm{},l_{2s}`$ are the creation operators on $`𝒯`$ determined by $`\xi _1,\xi _2,\mathrm{},\xi _{2s}`$, then $$\frac{(l_1+l_1^{})+i(l_2+l_2^{})}{\sqrt{2}},\mathrm{},\frac{(l_{2s1}+l_{2s1}^{})+i(l_{2s}+l_{2s}^{})}{\sqrt{2}}B(𝒯)$$ (1.4) form a circular system with respect to the so-called vacuum-state on $`B(𝒯)`$. By performing an orthogonal transformation on $`\xi _1,\xi _2,\mathrm{},\xi _{2s}`$ it is seen that instead of the family in (1.4) one can also use $$l_1+l_2^{},\mathrm{},l_{2s1}+l_{2s}^{}B(𝒯).$$ (1.5) The precise definitions of the objects involved in this realization of a circular system is reviewed in Section 2.1 below. 1.2 q-circular systems. In work related to $`q`$-deformations of the canonical commutation relations, Bożejko and Speicher obtained a remarkable deformation of the full Fock space, called the $`q`$-Fock space; here $`q`$ is a parameter in $`(1,1)`$, and the actual full Fock space is obtained for $`q=0`$ (see review in Section 2.2 below). In connection to this, Bożejko and Speicher studied the distribution – called by them q-Gaussian – of $`l+l^{}`$, where $`l`$ is an appropriately normalized creation operator on the $`q`$-Fock space. They discovered that the $`q`$-Gaussian distribution is the probability measure associated to an important family of orthogonal polynomials, the $`q`$-continuous Hermite polynomials. If $`q=0`$, then the $`q`$-Gaussian is the semicircular distribution appearing in (1.2), while the usual Gaussian is obtained in the limit $`q1`$. From the work in and its continuation in it is clear that if in the formulas (1.4), (1.5) one makes $`l_1,l_1^{},\mathrm{},l_s,l_s^{}`$ be creation/annihilation operators on the $`q`$-Fock space, then this should provide realizations of what one should call a $`q`$-circular system. But how does one actually define a $`q`$-circular system? It is unfortunate that the definition cannot be made in the same way as in the first paragraph of Section 1.1; this is because of the absence of a notion of “$`q`$-freeness in $`(𝒜,\phi )`$”. However, the combinatorial reformulation mentioned in the second paragraph of Section 1.1 can be extended to the $`q`$-case. Indeed, it is possible to give an explicit combinatorial description of the joint moments of the candidates of $`q`$-circular systems mentioned above (the families (1.4), (1.5), but where the $`l_i`$’s act on the $`q`$-Fock space). We state this formally in the next definition and proposition. Let us first succinctly describe the meaning of the combinatorial terms involved in Definition 1.2.2. 1.2.1 Notations. The fact that $`\pi =\{B_1,\mathrm{},B_p\}`$ is a pairing of $`\{1,\mathrm{},n\}`$ means that $`B_1\mathrm{}B_p=\{1,\mathrm{},n\}`$, disjoint, and each of $`B_1,\mathrm{},B_p`$ has exactly two elements. (Of course, $`n`$ must be even in order for $`\{1,\mathrm{},n\}`$ to have any pairings.) Two blocks $`B_i=\{a_i,b_i\}`$ and $`B_j=\{a_j,b_j\}`$ of a pairing $`\pi =\{B_1,\mathrm{},B_p\}`$ are said to cross if either $`a_i<a_j<b_i<b_j`$ or $`a_j<a_i<b_j<b_i`$; the number of crossings of $`\pi `$ is $$cr(\pi ):=\text{card}\{(i,j)|1i<jp,B_i\text{ and }B_j\text{ cross}\}.$$ (1.6) 1.2.2 Definition. Let $`(𝒜,\phi )`$ be a $`C^{}`$-probability space, and let $`q`$ be in $`(1,1)`$. The elements $`c_1,\mathrm{},c_s𝒜`$ ($`s1`$) are said to form a q-circular system in $`(𝒜,\phi )`$ if for every $`n1`$, $`r_1,\mathrm{},r_n\{1,\mathrm{},s\}`$, $`\epsilon (1),\mathrm{},\epsilon (n)\{1,\}`$ we have: $$\phi (c_{r_1}^{\epsilon (1)}\mathrm{}c_{r_n}^{\epsilon (n)})=\underset{\pi 𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))}{}q^{cr(\pi )},$$ (1.7) where $`𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))`$ denotes the set of all pairings $`\pi `$ = $`\{\{a_1,b_1\},\mathrm{},\{a_p,b_p\}\}`$ of $`\{1,\mathrm{},n\}`$ which have the property that $`r_{a_i}=r_{b_i}`$ and $`\epsilon (a_i)\epsilon (b_i)`$, $`1ip`$. (In the case that $`𝒫(r_1,\mathrm{},r_n;`$ $`\epsilon (1),\mathrm{},\epsilon (n))`$ is the empty set, the right-hand side of (1.7) is taken to be equal to 0.) 1.2.3 Proposition. Let $`q`$ be in $`(1,1)`$, and let $`s`$ be a positive integer. Let $`𝒯_q`$ denote the $`q`$-Fock space over $`\text{C}^{2s}`$; consider an orthonormal basis $`\xi _1,\mathrm{},\xi _{2s}`$ of $`\text{C}^{2s}`$ and let $`l_1,\mathrm{},l_{2s}B(𝒯_q)`$ be the creation operators associated to $`\xi _1,\mathrm{},\xi _{2s}`$. Then the families of operators $$\frac{(l_1+l_1^{})+i(l_2+l_2^{})}{\sqrt{2}},\mathrm{},\frac{(l_{2s1}+l_{2s1}^{})+i(l_{2s}+l_{2s}^{})}{\sqrt{2}}B(𝒯_q)$$ (1.8) and $$l_1+l_2^{},\mathrm{},l_{2s1}+l_{2s}^{}B(𝒯_q)$$ (1.9) are $`q`$-circular systems with respect to the vacuum-state on $`B(𝒯_q)`$. Note that if $`q=0`$, then the right-hand side of (1.7) counts the non-crossing pairings in $`𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))`$; this recaptures (and makes precise) the statement following to Eqn.(1.3). In the particular case $`q=0`$, a proof of Proposition 1.2.3 can be made by using the concept of R-transform; indeed, the R-transform of the family $`c_1,c_1^{},\mathrm{},c_s,c_s^{}`$ has a very simple form (see e.g. Eqn.(1.6) of for the case $`s=1`$), and the joint moments can be calculated from the knowledge of the R-transform. For general $`q(1,1)`$, the statement of Proposition 1.2.3 does not seem to have been previously considered, but can be inferred without difficulty from the results of and (see Sections 2.3-2.5 below). 1.3 An asymptotic model for q-circular systems. We now arrive to the main object of concern of the present paper, which is a certain asymptotic model for a $`q`$-circular system. The idea of using an asymptotic model for a circular system (case $`q=0`$) was brought to fact by Voiculescu in , and then very successfully used in . The asymptotic model observed in this paper is of a different nature than the one in , and is obtained by averaging unitaries in non-commutative tori. 1.3.1 Definition. Let $`q`$ be in $`(1,1)`$ and let $`s`$ be a positive integer. Suppose that for every $`k1`$ we are given a $`C^{}`$-probability space $`(𝒜_k,\phi _k)`$ and a family $`c_{1;k},\mathrm{},c_{s;k}`$ of $`𝒜_k`$. We will say that these families converge in distribution to a $`q`$-circular system if for every $`n1`$, $`r_1,\mathrm{},r_n\{1,\mathrm{},s\}`$, and $`\epsilon (1),\mathrm{},\epsilon (n)\{1,\}`$, the limit $$\underset{k\mathrm{}}{lim}\phi _k(c_{r_1;k}^{\epsilon (1)}\mathrm{}c_{r_n;k}^{\epsilon (n)})$$ exists and is equal to the right-hand side of Equation (1.7). 1.3.2 Definition. Let $`N`$ be a positive integer, and let $`(\rho _{i,j})_{1i<jN}`$ be a family of complex numbers of absolute value 1. Let $`(𝒜,\phi )`$ be a $`C^{}`$-probability space, and let $`u_1,\mathrm{},u_N`$ be elements of $`𝒜`$. We will say that $`(u_i)_{i=1}^N`$ is a $`(\rho _{i,j})_{i,j}`$commuting family of unitaries if: (i) every $`u_i`$ is a unitary, $`1iN`$, and: (ii) we have the relation $`u_iu_j=\rho _{i,j}u_ju_i`$, $`1i<jN`$. Moreover, we will say that $`(u_i)_{i=1}^N`$ is a $`(\rho _{i,j})_{i,j}`$–commuting Haar family of unitaries if in addition to (i) and (ii) we also have: (iii) $`\phi (u_1^{\lambda _1}\mathrm{}u_N^{\lambda _N})=0`$, $`(\lambda _1,\mathrm{}\lambda _N)\text{Z}^N\{(0,\mathrm{},0)\}`$. $`(\rho _{i,j})_{i,j}`$–commuting Haar families of unitaries can be constructed for any choice of the $`\rho _{i,j}`$’s, and live naturally in a class of $`C^{}`$-algebras called “non-commutative tori” – see e.g. . So now, let us fix a parameter $`q(1,1)`$. We will denote: $$\rho :=q+i\sqrt{1q^2}(|\rho |=1).$$ (1.10) We want to average families $`u_1,\mathrm{},u_N`$ of unitaries in a $`C^{}`$-probability space, such that for every $`1i<jN`$: either $`u_i`$ and $`u_j`$ $`\rho `$-commute, or they $`\rho ^1`$-commute. Since there is no canonical way to choose for which pairs $`i<j`$ we want to have $`u_iu_j=\rho u_ju_i`$ and for which ones we want to have $`u_iu_j=\rho ^1u_ju_i`$, we will use a “randomization” of $`u_1,\mathrm{},u_N`$. That is, we will make $`u_1,\mathrm{},u_N`$ be random unitaries in a $`C^{}`$-probability space, such that for every $`1i<jN`$ we have: $$P(u_iu_j=\rho u_ju_i)=\frac{1}{2}=P(u_iu_j=\rho ^1u_ju_i).$$ (1.11) This means that we will need the following version of Definition 1.3.2: 1.3.3 Definition. Let $`(\mathrm{\Omega },,P)`$ be a probability space, let $`N`$ be a positive integer, and let $`(\rho _{i,j})_{1i<jN}`$ be a family of random variables on $`\mathrm{\Omega }`$ with values in $`\{\zeta \text{C}|`$ $`|\zeta |=1\}`$. Let $`(𝒜,\phi )`$ be a $`C^{}`$-probability space, where the $`C^{}`$-algebra $`𝒜`$ is separable, and let $`U_1,\mathrm{},U_N`$ be measurable functions from $`\mathrm{\Omega }`$ to $`𝒜`$. We will say that $`(U_i)_{i=1}^N`$ form a $`(\rho _{i,j})_{i,j}`$-commuting family of random unitaries in $`(𝒜,\phi )`$ if: (j) $`U_i(\omega )𝒜`$ is a unitary, $`1iN`$, $`\omega \mathrm{\Omega }`$, and: (jj) we have the relation $`U_i(\omega )U_j(\omega )`$ = $`\rho _{i,j}(\omega )U_j(\omega )U_i(\omega )`$, $`1i<jN`$, $`\omega \mathrm{\Omega }`$. Moreover, we will say that $`(U_i)_{i=1}^N`$ is a $`(\rho _{i,j})_{i,j}`$–commuting Haar family of random unitaries if in addition to (j) and (jj) we also have: (jjj) $`\phi (U_1(\omega )^{\lambda _1}\mathrm{}U_N(\omega )^{\lambda _N})=0`$, $`(\lambda _1,\mathrm{}\lambda _N)`$ $``$ $`\text{Z}^N\{(0,\mathrm{},0)\}`$, $`\omega \mathrm{\Omega }`$. For the asymptotic model for a $`q`$-circular system it is sufficient to consider $`(\rho _{i,j})_{i,j}`$–commuting families of random unitaries where the random variables $`(\rho _{i,j})_{1i<jN}`$ are independent, and each of them takes finitely many values. For such $`\rho _{i,j}`$’s, the interested reader should have no difficulty to verify that one can construct $`(\rho _{i,j})_{i,j}`$–commuting Haar families of random unitaries which live in a tensor product of non-commutative tori. Since we will deal with random unitaries in a $`C^{}`$-probability $`(𝒜,\phi )`$, we will have to consider the new $`C^{}`$-probability space where these random unitaries belong: 1.3.4 Notation. Let $`(\mathrm{\Omega },,P)`$ be a probability space, and let $`(𝒜,\phi )`$ be a $`C^{}`$-probability space, where the $`C^{}`$-algebra $`𝒜`$ is separable. We will denote by $`(\mathrm{\Omega },𝒜)`$ the set of all bounded measurable functions from $`\mathrm{\Omega }`$ to $`𝒜`$. Then $`(\mathrm{\Omega },𝒜)`$ is a unital $`C^{}`$-algebra, with the operations defined pointwise, and with the norm given by $`f:=sup\{f(\omega )|\omega \mathrm{\Omega }\}`$, $`f(\mathrm{\Omega },𝒜)`$. Moreover, we have a natural state $`E:(\mathrm{\Omega },𝒜)\text{C}`$ given by the formula $$E(f):=_\mathrm{\Omega }\phi (f(\omega ))𝑑P(\omega ),f(\mathrm{\Omega },𝒜).$$ (1.12) It is immediate that $`((\mathrm{\Omega },𝒜),E)`$ is a $`C^{}`$-probability space; also, clearly, the unitaries in $`(\mathrm{\Omega },𝒜)`$ are random unitaries in $`𝒜`$, over the base space $`\mathrm{\Omega }`$. <sup>1</sup><sup>1</sup>1 One could also consider the space $`L^{\mathrm{}}(\mathrm{\Omega },𝒜)`$, which is the quotient of $`(\mathrm{\Omega },𝒜)`$ by the relation of equality almost everywhere with respect to $`P`$. Since the estimates of moments done in this paper are the same (no matter whether $`(\mathrm{\Omega },𝒜)`$ or $`L^{\mathrm{}}(\mathrm{\Omega },𝒜)`$ is used), we prefer to stay with $`(\mathrm{\Omega },𝒜)`$. We can now return to $`q`$ and $`\rho `$ of Equation (1.10), and state precisely how an asymptotic model for the $`q`$-circular system is obtained. 1.3.5 Proposition. Let $`q`$ be in $`(1,1)`$, and let $`s`$ be a positive integer. Denote $`\rho :=q+i\sqrt{1q^2}`$. Suppose that for every $`k1`$ we have: (a) A family $`(\rho _{i,j;k})_{1i<jks}`$ of independent random variables over some probability space $`\mathrm{\Omega }_k`$, such that every $`\rho _{i,j;k}`$ takes only the values $`\rho `$ and $`\rho ^1`$, with $`P(\rho _{i,j;k}=\rho )=1/2=P(\rho _{i,j:k}=\rho ^1)`$. (b) A $`(\rho _{i,j;k})_{i,j}`$–commuting Haar family $`U_{1;k},\mathrm{},U_{ks;k}`$ of random unitaries in some separable $`C^{}`$-probability space $`(𝒜_k,\phi _k)`$. Denote, for every $`k1`$: $$X_{r;k}:=\frac{1}{\sqrt{k}}(U_{r;k}+U_{r+s;k}+\mathrm{}+U_{r+(k1)s;k}),1rs;$$ (1.13) then the family $`(X_{1;k},\mathrm{},X_{s;k})`$ converges in distribution, for $`k\mathrm{}`$, to a $`q`$-circular system. We should note here a similarity with the idea of the non-commutative central limit theorem of : in Proposition 1.3.5 we wrote $`q`$ as a convex combination of $`\rho `$ and $`\rho ^1`$, whereas in Speicher writes $`q`$ as a convex combination of $`1`$ and $`1`$. In Theorem 1.5.3 below we will generalize Proposition 1.3.5 to a case which contains both these situations, and where the only restriction on the $`\rho _{i,j;k}`$’s (besides their independence) concerns the values of their expectations. In order to state this more general result, we will first introduce the concept of orientation for the crossings of a pairing. 1.4 Oriented crossings. 1.4.1 Crossing of two segments. We start from a simple geometric idea. Let $`P,Q,U,V`$ be distinct points in the plane, such that the segments $`PQ`$ and $`UV`$ cross. Consider the vector product $`\stackrel{}{w}=\stackrel{}{PQ}\times \stackrel{}{UV}`$, which is a vector perpendicular to the plane of $`P,Q,U,V`$. If $`\stackrel{}{w}`$ is oriented upwards we will say that $`PQ`$ and $`UV`$ have a positive crossing, while if $`\stackrel{}{w}`$ is oriented downwards we will say that $`PQ`$ and $`UV`$ have a negative crossing. In other words, if we denote the coordinates of $`P`$ by $`(p_1,p_2)`$, the coordinates of $`Q`$ by $`(q_1,q_2)`$, etc, then the sign of the crossing between $`PQ`$ and $`UV`$ is equal to $$\text{sign}\left(\text{det}\left(\begin{array}{cc}q_1p_1& q_2p_2\\ v_1u_1& v_2u_2\end{array}\right)\right).$$ (1.14) Note that the sign of the crossing is sensitive to the order of the points of each segment, also to the order of the two segments; e.g, if $`PQ`$ and $`UV`$ have positive crossing then $`QP`$ and $`UV`$ have negative crossing, also $`UV`$ and $`PQ`$ have negative crossing. 1.4.2 Crossings of a pairing. Let now $`n=2p`$ be an even positive integer, and let $`\pi =\{B_1,\mathrm{},B_p\}`$ be a pairing of $`\{1,\mathrm{},n\}`$. One can obtain a geometric representation of $`\pi `$, by using the following recipe: draw a circle in the plane, and draw $`n`$ points $`P_1,\mathrm{},P_n`$ in counterclockwise order around the circle; then for every block $`B_i=\{a_i,b_i\}`$ of $`\pi `$ draw the line segment with endpoints $`P_{a_i}`$ and $`P_{b_i}`$. It is immediate that the blocks $`B_i`$ and $`B_j`$ cross if and only if the corresponding line segments $`P_{a_i}P_{b_i}`$ and $`P_{a_j}P_{b_j}`$ do so. Thus the geometric representation of $`\pi `$ will display $`p`$ ( = $`n/2`$ ) line segments, which have a total number of $`cr(\pi )`$ points of intersection. At this point, we would like to orient the crossings of the pairing $`\pi `$, by using the considerations from 1.4.1. But in order to do so we need some additional data to be given, namely: $`(\alpha )`$ a direction of running along the segment $`P_{a_i}P_{b_i}`$, $`1ip`$; and $`(\beta )`$ an ordering of the $`p`$ segments $`P_{a_i}P_{b_i}`$, $`1ip`$. It will be convenient to satisfy the above requirement $`(\alpha )`$ by giving a function $`\epsilon :\{1,\mathrm{},n\}\{1,\}`$ with the property that for every block $`B_i=\{a_i,b_i\}`$ of $`\pi `$ we have $`\epsilon (a_i)\epsilon (b_i)`$. In the presence of such $`\epsilon `$, we will make the convention that every segment $`P_{a_i}P_{b_i}`$ is to be run from the point which is mapped by $`\epsilon `$ into $``$ towards the point which is mapped by $`\epsilon `$ into 1. Concerning the requirement $`(\beta )`$, we will do the book-keeping by comparing the ordering of the blocks of $`\pi `$ which is used in the crossing orientation against the “standard” ordering which lists the blocks in increasing order of their minimal elements. More precisely, let us assume that the blocks $`B_1,\mathrm{},B_p`$ were from the beginning listed in standard order, with $`\mathrm{min}(B_1)<\mathrm{min}(B_2)<\mathrm{}<\mathrm{min}(B_p)`$. Then giving an arbitrary ordering of the blocks amounts to giving a permutation $`\sigma `$ of the set $`\{1,\mathrm{},p\}`$: the convention we will use is that in the presence of such a permutation $`\sigma `$, the ordering “$``$” of the blocks of $`\pi `$ is defined such that $`B_{\sigma (1)}B_{\sigma (2)}\mathrm{}B_{\sigma (p)}`$. To summarize: we do not make the orientation of crossings for just the pairing $`\pi `$, but for a triple $`(\pi ,\epsilon ,\sigma )`$, where $`\epsilon :\{1,\mathrm{},n\}\{1,\}`$ has the property that $`\epsilon (a_i)\epsilon (b_i)`$ for every block $`B_i=\{a_i,b_i\}`$ of $`\pi `$, and $`\sigma `$ is a permutation of the set $`\{1,\mathrm{},p\}`$. For such $`(\pi ,\epsilon ,\sigma )`$, the orientation of crossings is achieved by drawing the geometric representation of $`\pi `$, and then by using the method described in Section 1.4.1. Let $`(\pi ,\epsilon ,\sigma )`$ be as in the preceding paragraph. We will denote by $`cr_+(\pi ,\epsilon ,\sigma )`$ and $`cr_{}(\pi ,\epsilon ,\sigma )`$ the number of crossings of $`(\pi ,\epsilon ,\sigma )`$ which have positive, respectively negative, orientation. A distinctive feature of these numbers is of course that: $$cr_+(\pi ,\epsilon ,\sigma )+cr_{}(\pi ,\epsilon ,\sigma )=cr(\pi ),$$ (1.15) the total number of crossings of $`\pi `$. We leave it as an exercise to the reader to check that if $`\pi `$ = $`\{B_1,\mathrm{},B_p\}`$ with $`B_1,\mathrm{},B_p`$ listed in increasing order of their minimal elements, then the explicit formulas for $`cr_\pm (\pi ,\epsilon ,\sigma )`$ are: $$cr_+(\pi ,\epsilon ,\sigma )=\text{card}\left\{(i,j)\begin{array}{cc}|\hfill & 1i<jp,B_i\text{ and }B_j\text{ cross, }\hfill \\ |\hfill & \epsilon (\mathrm{min}(B_i))\epsilon (\mathrm{min}(B_j))=\text{sign}(\sigma (j)\sigma (i))\hfill \end{array}\right\}$$ (1.16) $$cr_{}(\pi ,\epsilon ,\sigma )=\text{card}\left\{(i,j)\begin{array}{cc}|\hfill & 1i<jp,B_i\text{ and }B_j\text{ cross, }\hfill \\ |\hfill & \epsilon (\mathrm{min}(B_i))\epsilon (\mathrm{min}(B_j))=\text{sign}(\sigma (j)\sigma (i))\hfill \end{array}\right\},$$ where in the products “$`\epsilon (\mathrm{min}(B_i))\epsilon (\mathrm{min}(B_j))`$” of (1.16) the following convention is used: if we encounter a product of two symbols out of which at least one is a “$``$” (e.g. $`1`$, or $``$), then $``$ is to be treated like $`1`$. 1.5 z-circular systems. The limit distribution which appears in the generalization of Proposition 1.3.5 is the following: 1.5.1 Definition. Let $`(𝒜,\phi )`$ be a $`C^{}`$-probability space, and let $`z`$ be a complex number such that $`|z|<1`$. The elements $`c_1,\mathrm{},c_s𝒜`$ ($`s1`$) are said to form a z-circular system in $`(𝒜,\phi )`$ if: – for every positive odd integer $`n`$, for every $`r_1,\mathrm{},r_n\{1,\mathrm{},s\}`$, and for every $`\epsilon (1),\mathrm{},\epsilon (n)\{1,\}`$, we have that $`\phi (c_{r_1}^{\epsilon (1)}\mathrm{}c_{r_n}^{\epsilon (n)})=0`$; and – for every positive even integer $`n=2p`$, for every $`r_1,\mathrm{},r_n\{1,\mathrm{},s\}`$, and for every $`\epsilon (1),\mathrm{},\epsilon (n)\{1,\}`$, we have that: $$\phi (c_{r_1}^{\epsilon (1)}\mathrm{}c_{r_n}^{\epsilon (n)})=\frac{1}{p!}\underset{\sigma 𝒮_p}{}\underset{\pi 𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))}{}z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )},$$ (1.17) where $`𝒮_p`$ denotes the set of all permutations of $`\{1,\mathrm{},p\}`$, the index set $`𝒫(r_1,\mathrm{},r_n;`$ $`\epsilon (1),\mathrm{},\epsilon (n))`$ has the same meaning as in Definition 1.2.1, and $`cr_\pm (\pi ,\epsilon ,\sigma )`$ are as discussed in Section 1.4.2. 1.5.2 Remarks. $`1^o`$ If $`z=q(1,1)`$, then the concept of $`z`$-circular system reduces to the one of $`q`$-circular system from Definition 1.2.1. Indeed, in the relevant case of $`n=2p`$ appearing in Equation (1.17) we will now obtain: $$z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )}=q^{cr_+(\pi ,\epsilon ,\sigma )+cr_{}(\pi ,\epsilon ,\sigma )}=q^{cr(\pi )},$$ for every $`\sigma 𝒮_p`$ and $`\pi 𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))`$. So if one performs first the summation over $`𝒮_p`$, then the right-hand side of Equation (1.17) reduces to the right-hand side of (1.7). $`2^o`$ Starting from Definition 1.5.1, one can easily also define what it means that a sequence of families $`(c_{1;k},\mathrm{},c_{s;k})_{k1}`$ converges in distribution to a $`z`$-circular system – this is just an immediate adaptation of Definition 1.3.1. 1.5.3 Theorem. Let $`z`$ be a complex number such that $`|z|<1`$, and let $`s`$ be a positive integer. Suppose that for every $`k1`$ we have: (a) A family $`(\rho _{i,j;k})_{1i<jks}`$ of independent random variables over some probability space $`\mathrm{\Omega }_k`$, such that every $`\rho _{i,j;k}`$ takes values in the unit circle $`\{\zeta \text{C}||\zeta |=1\}`$, and has the property that $`_{\mathrm{\Omega }_k}\rho _{i,j;k}=z`$. (b) A $`(\rho _{i,j;k})_{i,j}`$–commuting Haar family $`U_{1;k},\mathrm{},U_{ks;k}`$ of random unitaries in some separable $`C^{}`$-probability space $`(𝒜_k,\phi _k)`$. Denote, for every $`k1`$: $$X_{r;k}:=\frac{1}{\sqrt{k}}(U_{r;k}+U_{r+s;k}+\mathrm{}+U_{r+(k1)s;k}),1rs;$$ then the family $`(X_{1;k},\mathrm{},X_{s;k})`$ converges in distribution, for $`k\mathrm{}`$, to a $`z`$-circular system. 1.5.4 Remark. $`1^o`$ In the case when the random variables $`\rho _{i,j;k}`$ of Theorem 1.5.3 take values in $`\{1,1\}`$, we obtain a statement which is close to the framework of the central limit theorem of . We note however that even in this case, the Proposition 1.6.2 below – which generalizes Theorem 1.5.3, and is the statement that we really prove – does not follow from the results of . $`2^o`$ There are some natural questions which are raised by the preceding theorem, concerning the possibility of realizing a $`z`$-circular system as a family of operators on some Hilbert space. One approach that can be used is the following. Consider the unital algebra $`\text{C}X_1,Y_1,\mathrm{},`$ $`X_s,Y_s=:𝒜_o`$ of polynomials in $`2s`$ non-commuting indeterminates $`X_1,Y_1,\mathrm{},X_s,Y_s`$, and make $`𝒜_o`$ be a $``$-algebra by introducing on it the (uniquely determined) $``$-operation with the property that $`X_r^{}=Y_r`$, $`1rs`$. Let us moreover consider the linear functional $`\phi _o:𝒜_o\text{C}`$ determined by the fact that $`\phi _o(1)=1`$ and that $`\phi _o(X_{r_1}^{\epsilon (1)}\mathrm{}X_{r_n}^{\epsilon (n)})`$ is equal to the right-hand side of Equation (1.17), for every $`n1`$, $`r_1,\mathrm{},r_n\{1,\mathrm{},s\}`$, $`\epsilon (1),\mathrm{},\epsilon (n)\{1,\}`$. The Theorem 1.5.3 ensures that $`\phi _o`$ is a positive functional $`(\phi _o(P^{}P)0,P𝒜_o)`$; indeed, it is easy to rephrase the theorem in a way which presents $`\phi _o`$ as a pointwise limit of linear functionals $`(\phi _k)_{k=1}^{\mathrm{}}`$, each of the $`\phi _k`$’s being positive. But then one can consider the GNS construction for $`\phi _o`$; this should yield a $``$-representation $`\mathrm{\Phi }_o:𝒜_oB()`$, with a cyclic vector $`\xi _o`$, such that $`\mathrm{\Phi }_o(X_1),\mathrm{},\mathrm{\Phi }_o(X_s)`$ form a $`z`$-circular system with respect to the vector-state on $`B()`$ given by $`\xi _o`$. The point we cannot settle here is whether the operators $`\mathrm{\Phi }_o(X_1),\mathrm{},\mathrm{\Phi }_o(X_s)`$ are indeed bounded on $``$. We believe nevertheless that this is true, and that the condition “$`|z|<1`$” from the definition of a $`z`$-circular system should be essential in proving it. Another approach which can be tried in order to realize $`z`$-circular systems would be by generalizing the Proposition 1.2.3 to the framework of an appropriately defined $`z`$-Fock space. The concept of $`q`$-Fock space which will be reviewed in Section 2.2 below was amply generalized in , ; on the other hand, Fock space constructions related to the framework of (spin systems with mixed commutation and anti-commutation relations) are discussed in . It isn’t however clear if any of these constructions can be tailored to give a $`z`$-Fock space as required by the situation at hand. 1.6 Refinements of Theorem 1.5.3. It is useful (for instance for the approximation with random matrices shown in Section 1.7 below) to note that one can relax some of the hypotheses of Theorem 1.5.3, and still obtain the same conclusion. In the next theorem we weaken the hypotheses on the expectations $`\rho _{i,j}`$, and on the Haar condition (jjj) from Definition 1.3.3. The weakened Haar condition is described as follows: let $`(U_i)_{i=1}^N`$ be a $`(\rho _{i,j})_{i,j}`$–commuting family of random unitaries, in the sense of (j)$`+`$(jj) of Definition 1.3.3, and let $`L`$ be a positive integer. We will say that $`(U_i)_{i=1}^N`$ is an L-mimic of a Haar family if it satisfies: $$\text{(jjj-}L\text{)}\phi (U_1(\omega )^{\lambda _1}\mathrm{}U_N(\omega )^{\lambda _N})=0,(\lambda _1,\mathrm{},\lambda _N)((L,L)\text{Z})^N\{(0,\mathrm{},0)\}.$$ 1.6.1 Proposition. Let $`z`$ be a complex number such that $`|z|<1`$, and let $`s`$ be a positive integer. Let $`(\delta _k)_{k=1}^{\mathrm{}}`$ be a sequence of positive real numbers, and let $`(L_k)_{k=1}^{\mathrm{}}`$ be a sequence of positive integers, such that $`\delta _k0`$ and $`L_k\mathrm{}`$. Suppose that for every $`k1`$ we have: (a) A family $`(\rho _{i,j;k})_{1i<jks}`$ of independent random variables over some probability space $`\mathrm{\Omega }_k`$, such that every $`\rho _{i,j;k}`$ takes values in the unit circle and has the property that $`|z_{\mathrm{\Omega }_k}\rho _{i,j;k}|\delta _k`$. (b) A $`(\rho _{i,j;k})_{i,j}`$–commuting family $`U_{1;k},\mathrm{},U_{ks;k}`$ of random unitaries in some separable $`C^{}`$-probability space $`(𝒜_k,\phi _k)`$, such that $`U_{1;k},\mathrm{},U_{ks;k}`$ is an $`L_k`$-mimic of a Haar family. Denote, for every $`k1`$: $$X_{r;k}:=\frac{1}{\sqrt{k}}(U_{r;k}+U_{r+s;k}+\mathrm{}+U_{r+(k1)s;k}),1rs;$$ then the family $`(X_{1;k},\mathrm{},X_{s;k})`$ converges in distribution, for $`k\mathrm{}`$, to a $`z`$-circular system. It is worth recording that the statement of 1.6.1 follows from an estimate of moments which can be formulated simply, as described in the next proposition. (Since the extra indices “$`k`$” are not necessary in Proposition 1.6.2, we will write in its statement $`\rho _{i,j},U_i,X_r`$ instead of $`\rho _{i,j;k},U_{i;k},X_{r;k}`$, respectively.) 1.6.2 Proposition. Let $`z`$ be a complex number such that $`|z|<1`$, and let $`s`$ be a positive integer. Let $`\delta `$ be a positive real number, and let $`L`$ be a positive integer. Let $`k`$ be a positive integer, and suppose that we have: (a) A family $`(\rho _{i,j})_{1i<jks}`$ of independent random variables over some probability space $`\mathrm{\Omega }`$, such that every $`\rho _{i,j}`$ takes values in the unit circle and has the property that $`|z_\mathrm{\Omega }\rho _{i,j}|\delta `$. (b) A $`(\rho _{i,j})_{i,j}`$–commuting family $`U_1,\mathrm{},U_{ks}`$ of random unitaries in some separable $`C^{}`$-probability space $`(𝒜,\phi )`$, such that $`U_1,\mathrm{},U_{ks}`$ is an $`L`$-mimic of a Haar family. We denote: $$X_r:=\frac{1}{\sqrt{k}}(U_r+U_{r+s}+\mathrm{}+U_{r+(k1)s}),1rs.$$ (1.18) We denote by $`E:(\mathrm{\Omega },𝒜)\text{C}`$ the linear functional defined as in Equation (1.12) of Notation 1.3.4. Then: $`1^o`$ For every odd positive integer $`n<L`$, for every $`r_1,\mathrm{},r_n\{1,\mathrm{},s\}`$, and for every $`\epsilon (1),\mathrm{},\epsilon (n)\{1,\}`$, we have that: $$E(X_{r_1}^{\epsilon (1)}\mathrm{}X_{r_n}^{\epsilon (n)})=0.$$ (1.19) $`2^o`$ For every even positive integer $`n=2p`$ such that $`n<\mathrm{min}(L,2k)`$, for every $`r_1,\mathrm{},r_n\{1,\mathrm{},s\}`$, and for every $`\epsilon (1),\mathrm{},\epsilon (n)\{1,\}`$, we have that: $$|E(X_{r_1}^{\epsilon (1)}\mathrm{}X_{r_n}^{\epsilon (n)})\frac{1}{p!}\underset{\sigma 𝒮_p}{}\underset{\pi 𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))}{}z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )}|$$ $$<(2p+1)!(\frac{1}{k}+\delta ).$$ (1.20) The framework of Theorem 1.5.3 contains in particular the situation when the families of random unitaries $`(U_{1;k},\mathrm{},U_{ks;k})_{k1}`$ extend each other, i.e. when $`U_{i;k+1}=U_{i;k}`$, $`k1`$, $`1iks`$. When moving to the more general framework of 1.6.1, the case of the extending families of random unitaries needs to be discussed separately. One possibility of treating this case is provided by the following proposition. 1.6.3 Proposition. Let $`z`$ be a complex number such that $`|z|<1`$, and let $`s`$ be a positive integer. Suppose that we have a family $`(\rho _{m,n})_{1m<n}`$ of independent random variables with values in the unit circle, and a family $`(U_n)_{n=1}^{\mathrm{}}`$ of random unitaries in a separable $`C^{}`$-probability space $`(𝒜,\phi )`$ (all the $`\rho _{m,n}`$’s and $`U_n`$’s defined on the same probability space $`\mathrm{\Omega }`$), such that the following conditions are satisfied. (a) The commutation relation $$U_m(\omega )U_n(\omega )=\rho _{m,n}(\omega )U_n(\omega )U_m(\omega )$$ (1.21) holds for every $`1m<n`$ and for every $`\omega \mathrm{\Omega }`$. (b) For every $`\delta >0`$ there exists $`m_o1`$ such that: $`m_om<n`$ $``$ $`|z_\mathrm{\Omega }\rho _{m,n}|\delta `$. (c) For every positive integer $`L`$ there exists $`m_o1`$ such that: $`m_om<n`$ $``$ the family $`U_m,\mathrm{},U_n`$ is an $`L`$-mimic of a Haar family. (d) If $`n1`$, $`\lambda _1,\mathrm{},\lambda _n\text{Z}`$, $`\omega \mathrm{\Omega }`$, and if at least one of $`\lambda _1,\mathrm{},\lambda _n`$ is equal to $`\pm 1`$ or to $`\pm 2`$, then $`\phi (U_1(\omega )^{\lambda _1}\mathrm{}U_n(\omega )^{\lambda _n})`$ = 0. For every $`k1`$ we denote: $$X_{r;k}=\frac{1}{\sqrt{k}}(U_r+U_{r+s}+\mathrm{}+U_{r+(k1)s}),1rs;$$ (1.22) then the family $`(X_{1;k},\mathrm{},X_{s;k})`$ converges in distribution, for $`k\mathrm{}`$, to a $`z`$-circular system. 1.7 Approximation with random matrices. We will now point out that, as a consequence of the results presented in Section 1.6, one can easily obtain families of random matrices which converge in distribution to a $`z`$-circular system. In fact, it is nice to realize all these random matrices as random elements in the same $`C^{}`$-algebra, which will be an UHF-algebra (i.e. a certain inductive limit of matrix algebras). So, let us fix a complex number $`z`$ such that $`|z|<1`$. There exist unique $`\rho `$ and $`\gamma `$ such that $`|\rho |=1`$, Im$`(\rho )>0`$, $`\gamma (0,1)`$, and $`z=\gamma \rho +(1\gamma )\overline{\rho }`$. Let us also fix a sequence $`(\theta _n)_{n=1}^{\mathrm{}}`$ of rational numbers in $`(0,1)`$, such that $`lim_n\mathrm{}e^{2\pi i\theta _n}=\rho `$, and such that when we write $`\theta _n=a_n/b_n`$ with $`a_n,b_n`$ relatively prime positive integers, we get that $`3b_1<b_2<\mathrm{}`$ $`<b_n<\mathrm{}`$. For every $`n1`$, let us consider the finite dimensional $`C^{}`$-algebra $$𝒜_n:=M_{b_1}(\text{C})\mathrm{}M_{b_n}(\text{C}),$$ (where $`b_j`$ is the denominator of $`\theta _j`$, as above); on $`𝒜_n`$ we consider the state $`\phi _n`$ which is the tensor product of the normalized trace-functionals on $`M_{b_1}(\text{C}),\mathrm{},M_{b_n}(\text{C})`$. Let us furthermore consider the inductive limit $$𝒜:=\underset{n\mathrm{}}{lim}𝒜_n,$$ where the mapping from $`𝒜_{n1}`$ to $`𝒜_n`$ is $`xx1_{b_n}`$, for every $`n1`$. For notational convenience we shall regard each $`𝒜_n`$ as a unital subalgebra of $`𝒜`$. We will denote by $`\phi `$ the unique state of $`𝒜`$ with the property that $`\phi |𝒜_n=\phi _n`$, $`n1`$. On the other hand, let us consider a probability space $`(\mathrm{\Omega },,P)`$ on which an infinite family $`(\xi _{m,n})_{1m<n}`$ of independent random variables is given, such that every $`\xi _{m,n}`$ takes only the values $`\pm 1`$, with $`P(\xi _{m,n}=1)=\gamma `$, $`P(\xi _{m,n}=1)=1\gamma `$. We denote $$\rho _{m,n}(\omega )=e^{2\pi i\theta _m\xi _{m,n}(\omega )},\omega \mathrm{\Omega },1m<n.$$ (1.23) Then $`(\rho _{m,n})_{1m<n}`$ is also an independent family of random variables on $`\mathrm{\Omega }`$, with values in the unit circle. We construct a sequence of random unitaries $`(U_n)_{n=1}^{\mathrm{}}`$ in the $`C^{}`$-algebra $`𝒜`$, as follows. For every $`n1`$, consider first the $`b_n\times b_n`$-matrices: $$V_n=\left(\begin{array}{ccccc}0& & & & 1\\ 1& 0& & & \\ & 1& 0& & \\ & & \mathrm{}& \mathrm{}& \\ & & & 1& 0\end{array}\right),W_n=\left(\begin{array}{ccccc}e^{2\pi i\theta _n}& & & & \\ & e^{4\pi i\theta _n}& & & \\ & & \mathrm{}& & \\ & & & e^{2(b_n1)\pi i\theta _n}& \\ & & & & 1\end{array}\right);$$ (1.24) then set for every $`\omega \mathrm{\Omega }`$: $$\{\begin{array}{ccc}U_1(\omega )\hfill & =\hfill & V_1\hfill \\ U_2(\omega )\hfill & =\hfill & W_1^{\xi _{1,2}(\omega )}V_2\hfill \\ & \mathrm{}\hfill & \\ U_n(\omega )\hfill & =\hfill & W_1^{\xi _{1,n}(\omega )}W_2^{\xi _{2,n}(\omega )}\mathrm{}W_{n1}^{\xi _{n1,n}(\omega )}V_n\hfill \\ & \mathrm{}\hfill & \end{array}$$ (1.25) Clearly, the random unitary $`U_n`$ takes values in the finite dimensional subalgebra $`𝒜_n𝒜`$ (hence it is in fact a random unitary matrix of size $`b_1b_2\mathrm{}b_n`$). We claim that: 1.7.1 Proposition. The random variables $`(\rho _{m,n})_{1m<n}`$ defined in Equation (1.23), and the random unitaries $`(U_n)_{n=1}^{\mathrm{}}`$ defined in (1.25), satisfy the conditions considered in the Proposition 1.6.3. Indeed, the commutation relations (1.21) follow immediately from the fact that the matrices $`V_n,W_n`$ in (1.24) satisfy the relation $`V_nW_n=e^{2\pi i\theta _n}W_nV_n`$. It is also immediate that: $$\left|z_\mathrm{\Omega }\rho _{m,n}\right|=|(\gamma \rho +(1\gamma )\overline{\rho })(\gamma e^{2\pi i\theta _m}+(1\gamma )e^{2\pi i\theta _m})|$$ $$|e^{2\pi i\theta _m}\rho |,1m<n,$$ and this gives the required behavior for the expectations of the $`\rho _{m,n}`$’s. Concerning the Haar conditions, we leave it as an exercise to the reader to check that for every $`1m<n`$, the family $`W_m,W_{m+1},\mathrm{},W_n`$ is a $`b_m`$-mimic of a Haar family; the verification of both this statement and of the hypothesis (d) in 1.6.3 reduce to the fact that matrices of the form $`V_n^\alpha W_n^\beta `$ with $`0\alpha (b_n,b_n)\text{Z}`$ and $`\beta \text{Z}`$ have only zeros on the diagonal (and therefore have zero trace). Hence the recipe presented in Equation (1.22) of Proposition 1.6.3 will lead to an asymptotic $`z`$-circular system living in $`(𝒜,\phi )`$ (and which consists in fact of random matrices with sizes tending to infinity). Alternatively, one can fabricate random matrices which converge in distribution to a $`z`$-circular system by cutting out disjoint segments of the sequence $`(U_n)_{n=1}^{\mathrm{}}`$, and by invoking the Proposition 1.6.1. For example, in order to produce an asymptotic $`z`$-circular system with $`s=2`$ elements, one can set: $$X_{1;1}=U_1,X_{2;1}=U_2,$$ $$X_{1;2}=\frac{U_3+U_5}{\sqrt{2}},X_{2;2}=\frac{U_4+U_6}{\sqrt{2}},$$ $$X_{1;3}=\frac{U_7+U_9+U_{11}}{\sqrt{3}},X_{2;3}=\frac{U_8+U_{10}+U_{12}}{\sqrt{3}},\mathrm{}$$ (in general, the construction of $`X_{1;k}`$ and $`X_{2;k}`$ will use the segment $`U_{k^2k+1},\mathrm{},U_{k^2+k}`$ of the sequence $`(U_n)_{n=1}^{\mathrm{}}`$ ). The rest of the paper is divided into two sections. In Section 2 we review the $`q`$-Fock space, and prove Proposition 1.2.3. In Section 3 we present the estimates of moments which lead to the theorems presented in the Sections 1.3-1.6. 2. Combinatorics of the joint moments of q-creation/annihilation operators 2.1 Review of the full Fock space. In this paper we use the full Fock space over $`\text{C}^{2s}`$ ($`s`$ a fixed positive integer), which is: $$𝒯:=\text{C}\left(_{n=1}^{\mathrm{}}(\text{C}^{2s})^n\right)$$ (2.1) (orthogonal direct sum of Hilbert spaces). The number 1 in the first summand C on the right-hand side of Eqn.(2.1) is called the vacuum-vector, and is denoted by $`\mathrm{\Omega }`$. The vector-state determined by $`\mathrm{\Omega }`$ on $`B(𝒯)`$ is called the vacuum-state, and will be denoted by $`\phi _{vac}`$ $`(\phi _{vac}(X)`$ = $`X\mathrm{\Omega }|\mathrm{\Omega }`$, $`XB(𝒯))`$. For every $`\xi \text{C}^{2s}`$ we denote by $`l(\xi )`$ the creation operator determined by $`\xi `$ on $`𝒯`$, which is described by: $$\{\begin{array}{c}l(\xi )\mathrm{\Omega }=\xi \hfill \\ \\ \begin{array}{c}\hfill l(\xi )(\eta _1\mathrm{}\eta _n)=\xi \eta _1\mathrm{}\eta _n\\ \hfill n1,\eta _1,\mathrm{},\eta _n\text{C}^{2s}.\end{array}\hfill \end{array}$$ (2.2) The adjoint of $`l(\xi )`$ is called the annihilation operator determined by $`\xi `$, and acts by: $$\{\begin{array}{c}l(\xi )^{}\mathrm{\Omega }=0\hfill \\ \\ \begin{array}{c}\hfill l(\xi )^{}(\eta _1\mathrm{}\eta _n)=\eta _1|\xi \eta _2\mathrm{}\eta _n\\ \hfill n1,\eta _1,\mathrm{},\eta _n\text{C}^{2s}.\end{array}\hfill \end{array}$$ (2.3) It is immediately verified that we have: $$l(\xi )^{}l(\eta )=\eta |\xi I,\xi ,\eta \text{C}^{2s}.$$ (2.4) It is occasionally convenient to fix an orthonormal basis $`\xi _1,\mathrm{},\xi _{2s}`$ of $`\text{C}^{2s}`$, and denote $`l_i:=l(\xi _i)`$, $`1i2s`$. Then (2.4) gives us that $$l_i^{}l_j=\delta _{ij}I,1i,j2s,$$ (2.5) i.e. that $`l_1,\mathrm{},l_{2s}`$ form a family of Cuntz isometries (isometries with mutually orthogonal ranges). It is such a family which was used in (1.4) and (1.5) of Section 1.1, presenting realizations of a circular system. 2.2 Review of the q-Fock space. Besides the positive integer $`s`$, we now also fix a parameter $`q(1,1)`$. For every $`n1`$ we introduce an inner product $`,_q`$ on $`(\text{C}^{2s})^n`$, determined by the formula: $$\xi _1\mathrm{}\xi _n,\eta _1\mathrm{}\eta _n_q:=\underset{\sigma 𝒮_n}{}q^{inv(\sigma )}\xi _1|\eta _{\sigma (1)}\mathrm{}\xi _n|\eta _{\sigma (n)},$$ (2.6) for $`\xi _1,\mathrm{},\xi _n,\eta _1,\mathrm{},\eta _n\text{C}^{2s}`$, where $`𝒮_n`$ denotes the set of all permutations of $`\{1,\mathrm{},n\}`$, and where $`inv(\sigma )`$ stands for the number of inversions of the permutation $`\sigma `$ ( $`inv(\sigma ):=|\{(i,j)|1i<jn,\sigma (i)>\sigma (j)\}|`$ ). The fact that (2.6) gives indeed an inner product was shown in . We view $`,_q`$ as a “deformation” of the usual inner product on $`(\text{C}^{2s})^n`$ (which would correspond to the case when $`q=0`$). The $`q`$-Fock space over $`\text{C}^{2s}`$ is defined as $$𝒯_q:=\text{C}\left(_{n=1}^{\mathrm{}}((\text{C}^{2s})^n,,_q)\right)$$ (2.7) (orthogonal direct sum of Hilbert spaces). The vacuum-vector of $`𝒯_q`$ and the vacuum-state on $`B(𝒯_q)`$ are defined in exactly the same way as for the full Fock space (cf. Section 2.1). For every $`\xi \text{C}^{2s}`$ there exists a unique operator in $`B(𝒯_q)`$, denoted by $`l_q(\xi )`$, such that: $$\{\begin{array}{c}l_q(\xi )\mathrm{\Omega }=\xi \hfill \\ \\ \begin{array}{c}\hfill l_q(\xi )(\eta _1\mathrm{}\eta _n)=\xi \eta _1\mathrm{}\eta _n\\ \hfill n1,\eta _1,\mathrm{},\eta _n\text{C}^{2s};\end{array}\hfill \end{array}$$ (2.8) its adjoint acts by the formulas: $$\{\begin{array}{c}l_q(\xi )^{}\mathrm{\Omega }=0\hfill \\ \\ \begin{array}{c}\hfill l_q(\xi )^{}(\eta _1\mathrm{}\eta _n)=_{m=1}^nq^{m1}\eta _m|\xi \eta _1\mathrm{}\eta _{m1}\eta _{m+1}\mathrm{}\eta _n\\ \hfill n1,\eta _1,\mathrm{},\eta _n\text{C}^{2s}.\end{array}\hfill \end{array}$$ (2.9) $`l_q(\xi )`$ and $`l_q(\xi )^{}`$ are called the q-creation and respectively the q-annihilation operator determined by $`\xi `$. Note that the formulas describing $`l_q(\xi )`$ are identical to those for $`l(\xi )`$ in Section 2.1, but that the situation is not the same concerning the adjoints. (This is possible because $`𝒯_q`$ has an inner product which is a deformation of the one on $`𝒯`$.) Instead of (2.4), we now get that the $`q`$-creation and $`q`$-annihilation operators satisfy: $$l_q(\xi )^{}l_q(\eta )=ql_q(\eta )l_q(\xi )^{}+\eta |\xi I,\xi ,\eta \text{C}^{2s};$$ (2.10) these are called “the $`q`$-commutation relations”. It is occasionally convenient to fix an orthonormal basis $`\xi _1,\mathrm{},\xi _{2s}`$ of $`\text{C}^{2s}`$, and denote $`l_i:=l_q(\xi _i)`$, $`1i2s`$. The Eqn.(2.10) then gives us that $$l_i^{}l_j=ql_jl_i^{}+\delta _{ij}I,1i,j2s.$$ (2.11) It is such a family of operators in $`B(𝒯_q)`$ which was used in (1.8) and (1.9) of Section 1.2, presenting realizations of a $`q`$-circular system. We now turn to the proof of Proposition 1.2.3. We will stick to the framework introduced in Section 2.2, including an orthonormal basis $`\xi _1,\mathrm{},\xi _{2s}`$ of $`\text{C}^{2s}`$ which is fixed until the end of the Section 2, and for which we denote $`l_i:=l_q(\xi _i)`$, $`1i2s`$. The argument will rely on a combinatorial formula established in for the joint moments of $`l_1,l_1^{},\mathrm{},l_{2s},l_{2s}^{}`$ with respect to the vacuum-state $`\phi _{vac}`$ on $`B(𝒯_q)`$. This formula is stated as follows: 2.3 Proposition (cf. Part I, Proposition 2 on page 529). For every $`n1`$, $`t_1,\mathrm{},t_n\{1,\mathrm{},2s\}`$, $`\theta _1,\mathrm{},\theta _n\{1,\}`$, we have that $$\phi _{vac}(l_{t_1}^{\theta _1}l_{t_2}^{\theta _2}\mathrm{}l_{t_n}^{\theta _n})=\underset{\pi 𝒬(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)}{}q^{cr(\pi )},$$ (2.12) where $`𝒬(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)`$ denotes the set of all pairings $`\pi `$ = $`\{\{a_1,b_1\},\mathrm{},\{a_p,b_p\}\}`$ of $`\{1,\mathrm{},n\}`$ which have the property that $`t_{a_i}=t_{b_i}`$, $`\theta _{a_i}=`$, and $`\theta _{b_i}=1`$, $`1ip`$. We will first discuss the family of elements appearing in the formula (1.9) of Proposition 1.2.3. 2.4 Proposition. If $`c_1:=l_1+l_2^{},\mathrm{},c_s:=l_{2s1}+l_{2s}^{}`$, then $`c_1,\mathrm{},c_s`$ is a $`q`$-circular system with respect to the vacuum-state on $`B(𝒯_q)`$. Proof. We fix $`n1`$, $`r_1,\mathrm{},r_n\{1,\mathrm{},s\}`$, $`\epsilon (1),\mathrm{},\epsilon (n)\{1,\}`$ for which we will verify that Eqn.(1.7) holds. In this proof it will be convenient to use the following notation: given $`t_1,\mathrm{},t_n\{1,\mathrm{},2s\}`$, $`\theta _1,\mathrm{},\theta _n\{1,\}`$, we will write $$(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))$$ (2.13) to mean that for every $`1mn`$ the operator $`l_{t_m}^{\theta _m}`$ is one of the two terms which form $`c_{r_m}^{\epsilon (m)}`$. (For example: if $`c_{r_m}^{\epsilon (m)}`$ = $`c_3^{}=(l_5+l_6^{})^{}`$, then $`l_{t_m}^{\theta _m}`$ has to be either $`l_5^{}`$ or $`l_6`$; i.e., if $`r_m=3`$ and $`\epsilon _m=`$, then it is part of (2.13) that we have either $`t_m=5`$ and $`\theta _m=`$, or $`t_m=6`$ and $`\theta _m=1`$.) It is clear that: $$c_{r_1}^{\epsilon (1)}\mathrm{}c_{r_n}^{\epsilon (n)}=\underset{\begin{array}{c}\hfill (t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)\\ \hfill (r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))\end{array}}{}l_{t_1}^{\theta _1}\mathrm{}l_{t_n}^{\theta _n},$$ hence $$\phi _{vac}(c_{r_1}^{\epsilon (1)}\mathrm{}c_{r_n}^{\epsilon (n)})=\underset{\begin{array}{c}\hfill (t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)\\ \hfill (r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))\end{array}}{}\phi _{vac}(l_{t_1}^{\theta _1}\mathrm{}l_{t_n}^{\theta _n})$$ $$=\underset{\begin{array}{c}\hfill (t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)\\ \hfill (r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))\end{array}}{}\underset{\pi 𝒬(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)}{}q^{cr(\pi )}\text{( by (2.12) ).}$$ (2.14) We will next prove that: $$_{\begin{array}{c}\hfill (t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)\\ \hfill (r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))\end{array}}𝒬(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)=𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n)),$$ (2.15) disjoint union. In order to verify (2.15), let us first observe that: $$𝒬(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n)),$$ (2.16) whenever $`(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))`$. This is immediately seen by comparing the definition of $`𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))`$ (see Definition 1.2.2) with the one of $`𝒬(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)`$, and by taking into account how “$``$” works. The inclusion (2.16) gives the “$``$” part of (2.15). We now pass to “$``$” of (2.15). We pick a partition $`\pi 𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))`$, and we will construct $`t_1,\mathrm{},t_n\{1,\mathrm{},2s\}`$, $`\theta _1,\mathrm{},\theta _n\{1,\}`$ such that $$\{\begin{array}{c}(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n)),\text{and}\hfill \\ \pi 𝒬(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n).\hfill \end{array}$$ (2.17) Let $`B=\{a,b\}`$, with $`a<b`$, be an arbitrary block of $`\pi `$. From the fact that $`\pi 𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))`$, we get that $`r_a=r_b=:r`$, and $`\epsilon (a)\epsilon (b)`$. If $`\epsilon (a)=1`$ and $`\epsilon (b)=`$, this means that $`c_{r_a}^{\epsilon (a)}=c_r=l_{2r1}+l_{2r}^{}`$, $`c_{r_b}^{\epsilon (b)}=c_r^{}=l_{2r1}^{}+l_{2r}`$, and we choose: $`t_a=t_b=2r`$, $`\theta _a=`$, $`\theta _b=1`$ (such that $`l_{t_a}^{\theta _a}`$ is a term of $`c_r`$, and $`l_{t_b}^{\theta _b}`$ is a term of $`c_r^{}`$). If $`\epsilon (a)=`$ and $`\epsilon (b)=1`$, this means that $`c_{r_a}^{\epsilon (a)}=c_r^{}=l_{2r1}^{}+l_{2r}`$, $`c_{r_b}^{\epsilon (b)}=c_r=l_{2r1}+l_{2r}^{}`$, and we choose: $`t_a=t_b=2r1`$, $`\theta _a=`$, $`\theta _b=1`$. When we make the choices for $`t_a,t_b,\theta _a,\theta _b`$ as described in the preceding paragraph, and for every block of $`\pi `$, we obtain some $`t_1,\mathrm{},t_n\{1,\mathrm{},2s\}`$ and $`\theta _1,\mathrm{},\theta _n\{1,\}`$ such that (2.17) holds. This completes the proof of “$``$” in (2.15). It is also immediate (by inspecting again, one by one, the blocks of the partition $`\pi `$ considered above) that the choices for $`t_1,\mathrm{},t_n,\theta _1,\mathrm{},\theta _n`$ such that $`(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))`$ and at the same time $`𝒬(t_1,\mathrm{},t_n;\theta _1,\mathrm{},\theta _n)\pi `$ are uniquely determined; this proves the disjointness of the union in (2.15). Finally, from (2.15) it follows that the expression in (2.14) is $$\underset{\pi 𝒫(r_1,\mathrm{},r_n;\epsilon (1),\mathrm{},\epsilon (n))}{}q^{cr(\pi )},$$ which is exactly the desired expression for $`\phi _{vac}(c_{r_1}^{\epsilon (1)}\mathrm{}c_{r_n}^{\epsilon (n)})`$. QED It only remains that we prove the $`q`$-circularity of the family appearing in (1.8) of Proposition 1.2.3. By using arguments from , this can in fact be reduced to the $`q`$-circularity of (1.9), which was shown above. 2.5 Proposition. If we denote: $$c_1^{}:=\frac{(l_1+l_1^{})+i(l_2+l_2^{})}{\sqrt{2}},\mathrm{},c_s^{}:=\frac{(l_{2s1}+l_{2s1}^{})+i(l_{2s}+l_{2s}^{})}{\sqrt{2}},$$ then $`c_1^{},\mathrm{},c_s^{}`$ is a $`q`$-circular system with respect to the vacuum-state on $`B(𝒯_q)`$. Proof. Recall that “$`l_k`$” stands here for “$`l_q(\xi _k)`$”, $`1k2s`$, where $`\xi _1,\mathrm{},\xi _{2s}`$ is an orthonormal basis of $`\text{C}^{2s}`$ which was fixed prior to the Proposition 2.3. Consider the vectors $`\eta _1,\mathrm{},\eta _{2s}\text{C}^{2s}`$ defined by: $$\eta _{2r1}=\frac{\xi _{2r1}+\xi _{2r}}{\sqrt{2}},\eta _{2r}=\frac{\xi _{2r1}\xi _{2r}}{i\sqrt{2}},1rs,$$ (2.18) and let $`T`$ denote the unique operator in $`B(\text{C}^{2s})`$ such that $`T\xi _k=\eta _k`$, $`1k2s`$. It is immediate that $`\eta _1,\mathrm{},\eta _{2s}`$ is an orthonormal basis of $`\text{C}^{2s}`$, hence that $`T`$ is an orthogonal transformation. Let $`𝒱\text{C}^{2s}`$ be the real vector space spanned by $`\xi _1,\mathrm{},\xi _{2s}`$ (i.e. the set of vectors of the form $`_{k=1}^{2s}\lambda _k\xi _k`$, with $`\lambda _1,\mathrm{},\lambda _{2s}\text{R}`$), and let $`𝒲\text{C}^{2s}`$ be the real vector space spanned by $`\eta _1,\mathrm{},\eta _{2s}`$. Moreover, let $`,𝒩B(𝒯_q)`$ denote the von Neumann algebras generated by $`\{l_q(\xi )+l_q(\xi )^{}|\xi 𝒱\}`$, and respectively by $`\{l_q(\eta )+l_q(\eta )^{}|\eta 𝒲\}`$. The Theorem 2.11 of gives us the existence of a unital $``$-homomorphism $`\mathrm{\Phi }:𝒩`$, which preserves the vacuum-state (i.e. $`\phi _{vac}(\mathrm{\Phi }(x))=\phi _{vac}(x)`$, $`x`$), and such that: $$\mathrm{\Phi }(l_q(\xi )+l_q(\xi )^{})=l_q(T\xi )+l_q(T\xi )^{},\xi 𝒱.$$ (2.19) It is obvious that the operators $`c_1^{},\mathrm{},c_s^{}`$ defined in the statement of the proposition belong to $``$, and an immediate calculation which uses (2.18), (2.19), and the linearity of $`l_q()`$ gives that: $$\mathrm{\Phi }(c_r^{})=l_{2r1}+l_{2r}^{},1rs.$$ (2.20) Denoting $`c_r:=l_{2r1}+l_{2r}^{}`$, $`1rs`$, we thus obtain that $`c_1,\mathrm{},c_s𝒩`$ and also (since $`\mathrm{\Phi }`$ is a $``$-homomorphism which preserves $`\phi _{vac}`$) that: $$\phi _{vac}(c_{r_1}^{\epsilon (1)}\mathrm{}c_{r_n}^{\epsilon (n)})=\phi _{vac}(\mathrm{\Phi }((c_{r_1}^{})^{\epsilon (1)}\mathrm{}(c_{r_n}^{})^{\epsilon (n)}))$$ $$=\phi _{vac}((c_{r_1}^{})^{\epsilon (1)}\mathrm{}(c_{r_n}^{})^{\epsilon (n)}),$$ (2.21) for every $`n1`$ and $`r_1,\mathrm{},r_n\{1,\mathrm{},s\}`$, $`\epsilon (1),\mathrm{},\epsilon (n)\{1,\}`$. But then the conclusion of the current proposition follows from (2.21) and Proposition 2.4. QED 3. Moment estimates leading to asymptotic z-circular systems In this section we prove the results stated in the Sections 1.3-1.6 of the Introduction. It is clear that in fact only the Propositions 1.6.2 and 1.6.3 need to be proved (then 1.6.1, 1.5.3, 1.3.5 will follow). The bulk of the section will be devoted to the estimates of moments presented in Proposition 1.6.2. We fix, from this moment on and until the end of Section 3.6, the framework described in 1.6.2. We will first dispose of the easy case appearing in the part $`1^o`$ of the proposition. 3.1 Proof of part $`1^o`$ in Proposition 1.6.2. By substituting $`X_1,\mathrm{},X_s`$ from their definition in Eqn.(1.18), then by expanding the sums and by using the definition of $`E`$, we get: $$E(X_{r_1}^{\epsilon (1)}\mathrm{}X_{r_n}^{\epsilon (n)})=$$ $$=\frac{1}{k^{n/2}}\underset{\begin{array}{c}1i_1,\mathrm{},i_nkssuchthat\\ i_1=r_1(mods),\mathrm{},i_n=r_n(mods)\end{array}}{}E(U_{i_1}^{\epsilon (1)}\mathrm{}U_{i_n}^{\epsilon (n)})$$ $$=\frac{1}{k^{n/2}}\underset{\begin{array}{c}1i_1,\mathrm{},i_nkssuchthat\\ i_1=r_1(mods),\mathrm{},i_n=r_n(mods)\end{array}}{}_\mathrm{\Omega }\phi (U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_n}(\omega )^{\epsilon (n)})𝑑P(\omega ).$$ (3.1) We will show that: $$\phi (U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_n}(\omega )^{\epsilon (n)})=0,$$ (3.2) for every $`\omega \mathrm{\Omega }`$ and every $`1i_1,\mathrm{},i_nks`$ such that $`i_1=r_1(mods),\mathrm{},i_n=r_n(mods)`$. This and (3.1) clearly imply the conclusion of the lemma. So let us fix $`\omega \mathrm{\Omega }`$ and $`1i_1,\mathrm{},i_nks`$ such that $`i_1=r_1(mods),\mathrm{},i_n=r_n(mods)`$. The commutation relations satisfied by $`U_1,\mathrm{},U_{ks}`$ (see condition (jj) in Definition 1.3.3) give us that $$U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_n}(\omega )^{\epsilon (n)}=cU_1(\omega )^{\lambda _1}\mathrm{}U_{ks}(\omega )^{\lambda _{ks}},$$ (3.3) where $`c`$ is a constant of absolute value 1, and where $`\lambda _1,\mathrm{},\lambda _{ks}[n,n]\text{Z}(L,L)\text{Z}`$. It cannot be true that $`\lambda _1=\mathrm{}=\lambda _{ks}=0`$, because: $$\lambda _1+\mathrm{}+\lambda _{ks}=|\{1mn|\epsilon (m)=1\}||\{1mn|\epsilon (m)=\}|,$$ which is an odd number (indeed, $`|\{m|\epsilon (m)=1\}|`$ and $`|\{m|\epsilon (m)=\}|`$ must have different parities, since their sum is the odd number $`n`$). But then the condition (jjj-$`L`$) introduced in Section 1.6 gives us that $`\phi (U_1(\omega )^{\lambda _1}\mathrm{}U_{ks}(\omega )^{\lambda _{ks}})`$ = 0, and (3.2) is obtained by applying $`\phi `$ to both sides of (3.3). QED We now move towards the sensibly harder case discussed in part $`2^o`$ of Proposition 1.6.2. We will start by making a number of preliminary considerations. Unlike in the preceding proof, where we did not need to know what was the constant $`c`$ in Equation (3.3), the arguments in the sequel will require some information about such constants which arise from commutations. The next lemma will be used for that. 3.2 Lemma. Let $`p`$ be a positive integer and let $`\pi =\{B_1,\mathrm{},B_p\}`$ be a pairing of $`\{1,\mathrm{},2p\}`$, where the blocks $`B_1,\mathrm{},B_p`$ of $`\pi `$ are listed in increasing order of their minimal elements. Let $`𝒞`$ be a unital algebra and let $`V_1,\mathrm{},V_p`$ be invertible elements of $`𝒞`$ which satisfy the commutation relations $$V_lV_m=\gamma _{lm}V_mV_l,1l<mp,$$ (3.4) where the $`\gamma _{lm}`$’s are some complex numbers. Define $`W_1,\mathrm{},W_{2p}`$ according to the formula: $$W_i=\{\begin{array}{cccc}V_l\hfill & \text{if}& i=\text{min}(B_l)& \text{(for some }1lp)\hfill \\ & & & \\ V_m^1\hfill & \text{if}& i=\text{max}(B_m)& \text{(for some }1mp\text{).}\hfill \end{array}$$ (3.5) Then we have $$W_1W_2\mathrm{}W_{2p}=\left(\underset{\begin{array}{c}1l<mpsuch\\ thatB_lcrossesB_m\end{array}}{}\gamma _{lm}\right)I.$$ (3.6) Proof. By induction on $`p`$. The case $`p=1`$ is obvious (both sides of (3.6) are equal to $`I`$). Let us assume the lemma true for $`p1`$ and prove it for $`p2`$. Let $`\pi =\{B_1,\mathrm{},B_p\}`$, $`V_1,\mathrm{},V_p`$ and $`W_1,\mathrm{},W_{2p}`$ be as in the statement of the lemma. We write explicitly $`B_p=\{a,b\}`$, $`a<b`$ (recall that $`B_p`$ is the block of $`\pi `$ with the largest minimal element). Note that $`\{a+1,\mathrm{},b1\}`$ coincides with the set of maximal elements of the blocks $`B_l`$ $`(1lp1)`$ which cross $`B_p`$. By using this observation, the rule (3.5) for defining $`W_{a+1},\mathrm{},W_{b1}`$, and the commutation relations (3.4), we obtain that: $$(W_{a+1}\mathrm{}W_{b1})V_p^1=\left(\underset{\begin{array}{c}1lp1such\\ thatB_lcrossesB_p\end{array}}{}\gamma _{lp}\right)V_p^1(W_{a+1}\mathrm{}W_{b1}).$$ (3.7) On the other hand, let us denote by $`\pi _o`$ the pairing which is obtained from $`\pi `$ by deleting the block $`B_p`$ and by redenoting the elements of $`\{1,\mathrm{},2p\}B_p`$ as $`1,2,\mathrm{},2p2`$, in increasing order. The induction hypothesis applied to $`\pi _o`$ and $`V_1,\mathrm{},V_{p1}`$ gives us that: $$W_1\mathrm{}W_{a1}W_{a+1}\mathrm{}W_{b1}W_{b+1}\mathrm{}W_{2p}=\left(\underset{\begin{array}{c}1l<mp1such\\ thatB_lcrossesB_m\end{array}}{}\gamma _{lm}\right)I.$$ (3.8) But then: $$W_1W_2\mathrm{}W_{2p}=(W_1\mathrm{}W_{a1})V_p(W_{a+1}\mathrm{}W_{b1})V_p^1(W_{b+1}\mathrm{}W_{2p})$$ $$=\left(\underset{\begin{array}{c}1lp1such\\ thatB_lcrossesB_p\end{array}}{}\gamma _{lp}\right)(W_1\mathrm{}W_{a1})V_pV_p^1(W_{a+1}\mathrm{}W_{b1})(W_{b+1}\mathrm{}W_{2p})$$ ( by Equation (3.7) ) $$=\left(\underset{\begin{array}{c}1l<mpsuch\\ thatB_lcrossesB_m\end{array}}{}\gamma _{lm}\right)I\text{( by Equation (3.8) ).}$$ QED In the estimates of moments which will be presented below, we will also use the following notation and lemma. The positive integers $`p,s,k`$ appearing in 3.3 and 3.4 are the ones given in the statement of Proposition 1.6.2. 3.3 Notation. Let $`j_1,\mathrm{},j_p`$ be distinct numbers in $`\{1,\mathrm{},ks\}`$. We will denote by $`ord(j_1,\mathrm{},j_p)`$ the permutation $`\sigma `$ of $`\{1,\mathrm{},p\}`$ which keeps track of the order of $`j_1,\mathrm{},j_p`$; that is, $`\sigma `$ is the unique bijection from $`\{1,\mathrm{},p\}`$ to itself which has the property that $$\sigma (l)<\sigma (m)j_l<j_m,lm\text{ in }\{1,\mathrm{},p\}.$$ (3.9) 3.4 Lemma. Let $`\sigma `$ be a permutation of $`\{1,\mathrm{},p\}`$, and let $`t_1,\mathrm{},t_p`$ be in $`\{1,\mathrm{},s\}`$. Consider the number: $$N(\sigma ;t_1,\mathrm{},t_p)=|\left\{(j_1,\mathrm{},j_p)\begin{array}{cc}|\hfill & 1j_1,\mathrm{},j_pks,\hfill \\ |\hfill & ord(j_1,\mathrm{},j_p)=\sigma ,\hfill \\ |\hfill & j_1=t_1(mods),\mathrm{},j_p=t_p(mods)\hfill \end{array}\right\}|.$$ (3.10) Then: $$\left(\begin{array}{cc}k& \\ p& \end{array}\right)N(\sigma ;t_1,\mathrm{},t_p)\left(\begin{array}{cc}k+p& \\ p& \end{array}\right).$$ (3.11) Proof. It is immediate that $$N(\sigma ;t_1,\mathrm{},t_p)=N(id;t_{\sigma ^1(1)},\mathrm{},t_{\sigma ^1(p)}),$$ where $`id`$ denotes the identity permutation. Due to this fact, it suffices to verify (3.11) in the case when $`\sigma =id`$; i.e, it suffices to verify that for any choice of $`t_1,\mathrm{},t_p\{1,\mathrm{},s\}`$, the set $$𝒮:=\left\{(j_1,\mathrm{},j_p)\begin{array}{cc}|\hfill & 1j_1<\mathrm{}<j_pks,\hfill \\ |\hfill & j_1=t_1(mods),\mathrm{},j_p=t_p(mods)\hfill \end{array}\right\}$$ (3.12) has cardinality between $`\left(\begin{array}{cc}k& \\ p& \end{array}\right)`$ and $`\left(\begin{array}{cc}k+p& \\ p& \end{array}\right)`$. Let us denote $`I_1=\{1,\mathrm{},s\}`$, $`I_2=\{s+1,\mathrm{},2s\},\mathrm{},`$ $`I_k=\{(k1)s+1,\mathrm{},ks\}`$. To every $`(j_1,\mathrm{},j_p)`$ in the set $`𝒮`$ of (3.12) we can associate the $`p`$-tuple $`(m_1,\mathrm{},m_p)`$, where $`1m_1m_2\mathrm{}m_pk`$ are determined by the conditions: $$j_1I_{m_1},j_2I_{m_2},\mathrm{},j_pI_{m_p}.$$ Then the map $`(j_1,\mathrm{},j_m)(m_1,\mathrm{},m_p)`$ is one-to-one; this is immediately implied by the fact that every $`(j_1,\mathrm{},j_p)`$ in the set $`𝒮`$ of (3.12) has to satisfy the conditions $`j_1=t_1(mods),\mathrm{},j_p=t_p(mods)`$. We hence obtain that the cardinality of $`𝒮`$ is bounded above by $$|\left\{(m_1,\mathrm{},m_p)\right|1m_1m_2\mathrm{}m_pk\}|=\left(\begin{array}{cc}k+p1& \\ p& \end{array}\right)\left(\begin{array}{cc}k+p& \\ p& \end{array}\right).$$ On the other hand, the range of the map $`(j_1,\mathrm{},j_m)(m_1,\mathrm{},m_p)`$ considered in the preceding paragraph contains all the $`p`$-tuples $`(m_1,\mathrm{},m_p)`$ with the property that $`m_1<m_2<\mathrm{}<m_p`$. Indeed, if $`m_1<m_2<\mathrm{}<m_p`$, then there are unique $`j_1,\mathrm{},j_p\{1,\mathrm{},ks\}`$ such that: $`j_1I_{m_1}`$ and $`j_1=t_1(mods)`$; $`j_2I_{m_2}`$ and $`j_2=t_2(mods);\mathrm{},`$ $`j_pI_{m_p}`$ and $`j_p=t_p(mods)`$. These $`j_1,\mathrm{},j_p`$ form an element of the set $`𝒮`$ of (3.12), which is mapped to $`(m_1,\mathrm{},m_p)`$. So we obtain that the cardinality of $`𝒮`$ is bounded below by: $$|\left\{(m_1,\mathrm{},m_p)\right|1m_1<m_2<\mathrm{}<m_pk\}|=\left(\begin{array}{cc}k& \\ p& \end{array}\right).$$ QED We are now ready to attack the proof of part $`2^o`$ of Proposition 1.6.2. Before starting on this task, let us list some conventions of notation which will be used during the proof. 3.5 Notations. $`1^o`$ We will use the following conventions: – For $`1i<jks`$, the complex-conjugate of the random variable $`\rho _{i,j}`$ given in Proposition 1.6.2 will be denoted by $`\rho _{j,i}`$. (Thus $`\rho _{j,i}`$ is also a random variable on $`\mathrm{\Omega }`$, with values in the unit circle.) – In the $`2p`$-tuple $`\epsilon (1),\mathrm{},\epsilon (2p)`$ which appears in the statement of Proposition 1.6.2, the $`\epsilon (m)`$’s which are equal to $``$ will be treated in algebraic expressions as if they were equal to $`1`$. (For instance “$`_{bB}\epsilon (b)=0`$”, for $`B`$ a subset of $`\{1,\mathrm{},2p\}`$, will actually mean that $`|\{bB|\epsilon (b)=1\}|`$ = $`|\{bB|\epsilon (b)=\}|`$. ) $`2^o`$ Combinatorial notations: $`𝒫(2p)`$ will denote the set of pairings of $`\{1,\mathrm{},2p\}`$, where a pairing of $`\{1,\mathrm{},2p\}`$ is as defined in Notations 1.2.1. The set of all partitions of $`\{1,\mathrm{},2p\}`$ will be denoted by $`Part(2p)`$. (A partition $`\pi =\{B_1,\mathrm{},B_m\}`$ of $`\{1,\mathrm{},2p\}`$ is defined in the same way as a pairing, but without any restriction on the cardinalities of $`B_1,\mathrm{},B_m`$.) – Let $`\pi `$ be in $`Part(2p)`$. We will say that $`\pi `$ is r-stable if $`r_a=r_b`$ whenever $`a,b\{1,\mathrm{},2p\}`$ belong to the same block of $`\pi `$; and we will say that $`\pi `$ is $`\epsilon `$-null if $`_{bB}\epsilon (b)=0`$ for every block $`B`$ of $`\pi `$. (Here $`r_a,r_b`$ are extracted out of the $`2p`$-tuple $`r_1,r_2,\mathrm{},r_{2p}`$ appearing in the statement of Proposition 1.6.2, and similarly for the $`\epsilon (b)`$’s.) Note that the index set $`𝒫(r_1,\mathrm{},r_{2p};\epsilon (1),\mathrm{},\epsilon (2p))`$ appearing in Equation (1.20) of Proposition 1.6.2 can be presented as $$𝒫(r_1,\mathrm{},r_{2p};\epsilon (1),\mathrm{},\epsilon (2p))=\{\pi 𝒫(2p)|\pi \text{ is }r\text{-stable and }\epsilon \text{-null }\}.$$ (3.13) – If $`1i_1,\mathrm{},i_{2p}ks`$, then we will denote by $`ker(i_1,\mathrm{},i_{2p})Part(2p)`$ the partition $`\pi `$ determined as follows: $`a,b\{1,\mathrm{},2p\}`$ lie in the same block of $`\pi `$ if and only if $`i_a=i_b`$. 3.6 Proof of part $`2^o`$ in Proposition 1.6.2. The presentation of this fairly lengthy proof will be divided into several steps. Step 1. The evaluation of $`E(X_{r_1}^{\epsilon (1)}\mathrm{}X_{r_{2p}}^{\epsilon (2p)})`$ starts in the same way as the one for $`E(X_{r_1}^{\epsilon (1)}\mathrm{}X_{r_n}^{\epsilon (n)})`$ which was made in Section 3.1. We obtain the analogue of the Equation (3.1) of that proof: $$E(X_{r_1}^{\epsilon (1)}\mathrm{}X_{r_{2p}}^{\epsilon (2p)})=$$ (3.14) $$\frac{1}{k^p}\underset{\begin{array}{c}1i_1,\mathrm{},i_{2p}kssuchthat\\ i_1=r_1(mods),\mathrm{},i_{2p}=r_{2p}(mods)\end{array}}{}_\mathrm{\Omega }\phi (U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_{2p}}(\omega )^{\epsilon (2p)})𝑑P(\omega ).$$ We then write the right-hand side of (3.14) as a double summation, as follows: $$\underset{\pi Part(2p)}{}(\frac{1}{k^p}\underset{\begin{array}{c}1i_1,\mathrm{},i_{2p}ks\\ suchthatker(i_1,\mathrm{},i_{2p})=\pi and\\ i_1=r_1(mods),\mathrm{},i_{2p}=r_{2p}(mods)\end{array}}{}_\mathrm{\Omega }\phi (U_{i_1}(\omega )^{\epsilon (1)}\times $$ $$\times \mathrm{}U_{i_{2p}}(\omega )^{\epsilon (2p)})dP(\omega )).$$ In other words we write $$E(X_{r_1}^{\epsilon (1)}\mathrm{}X_{r_{2p}}^{\epsilon (2p)})=\underset{\pi Part(2p)}{}T_\pi ,$$ (3.15) where for $`\pi Part(2p)`$ we set: $$T_\pi :=\frac{1}{k^p}\underset{\begin{array}{c}1i_1,\mathrm{},i_{2p}ks\\ suchthatker(i_1,\mathrm{},i_{2p})=\pi and\\ i_1=r_1(mods),\mathrm{},i_{2p}=r_{2p}(mods)\end{array}}{}_\mathrm{\Omega }\phi (U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_{2p}}(\omega )^{\epsilon (2p)})𝑑P(\omega ).$$ (3.16) Our strategy will be to analyze, in the following few steps of the proof, the quantities $`T_\pi `$, $`\pi Part(2p)`$. Step 2. In this step we observe that if $`\pi Part(2p)`$ is not $`r`$-stable (in the sense defined in Notations 3.5), then the index set of the summation in (3.16) is void, and hence $`T_\pi `$ = 0 (in a vacuous way). Proof of Step 2. Suppose that $`\pi Part(2p)`$ is such that the index set of the summation in (3.16) is non-void. This means that there exist $`1i_1,\mathrm{},i_{2p}ks`$ such that $`ker(i_1,\mathrm{},i_{2p})=\pi `$ and such that $`i_1=r_1(mods),\mathrm{},i_{2p}=r_{2p}(mods)`$. Then for every $`a,b`$ belonging to the same block of $`\pi `$ we have: $`i_a=i_b`$ $``$ $`r_a=r_b(mods)`$ (because $`r_a=i_a(mods)`$, $`r_b=i_b(mods)`$ ) $``$ $`r_a=r_b`$ (because $`1r_a,r_bs`$), and we conclude that $`\pi `$ is $`r`$-stable. Step 3. Consider now a partition $`\pi Part(2p)`$ which is $`r`$-stable but is not $`\epsilon `$-null. We show that $`T_\pi `$ = 0. Proof of Step 3. We can prove in fact a stronger statement than $`T_\pi =0`$, namely that: $$\{\begin{array}{c}\phi (U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_{2p}}(\omega )^{\epsilon (2p)})=0,\hfill \\ \omega \mathrm{\Omega },1i_1,\mathrm{},i_{2p}ks\text{ such that }ker(i_1,\mathrm{},i_{2p})=\pi .\hfill \end{array}$$ (3.17) The proof of of (3.17) is similar to the proof of part $`1^o`$ of Proposition 1.6.2 (compare to Equation (3.2) in Section 3.1). Let $`B`$ be a block of $`\pi `$ such that $`_{bB}\epsilon (b)0`$. If $`1i_1,\mathrm{},i_{2p}ks`$ are such that $`ker(i_1,\mathrm{},i_{2p})=\pi `$, then $`i_a=i_b`$ for every $`a,bB`$, and it makes sense to denote by $`i\{1,\mathrm{},ks\}`$ the common value of the $`i_b`$’s with $`bB`$. The commutation relations satisfied by the unitaries $`U_1,\mathrm{},U_{ks}`$ give us, for an arbitrary $`\omega \mathrm{\Omega }`$, an equality of the form $$U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_{2p}}(\omega )^{\epsilon (2p)}=cU_1(\omega )^{\lambda _1}\mathrm{}U_{ks}(\omega )^{\lambda _{ks}},$$ (3.18) where $`c`$ is a constant of absolute value 1 and $`\lambda _1,\mathrm{},\lambda _{ks}[2p,2p]\text{Z}(L,L)\text{Z}`$. The point is that $`\lambda _i=_{bB}\epsilon (b)0`$; hence when we apply $`\phi `$ in (3.18), we obtain 0 because of the condition (jjj-$`L`$) introduced in Section 1.6. Step 4. We consider next a partition $`\pi Part(2p)`$ which is $`r`$-stable and $`\epsilon `$-null, but is not a pairing (i.e. not all the blocks of $`\pi `$ have exactly two elements). For such a $`\pi `$ we prove the inequality $$|T_\pi |<1/k.$$ (3.19) Proof of Step 4. Observe first that the number of terms in the sum defining $`T_\pi `$ in (3.16) is bounded above by $`k^m`$, where $`m`$ is the number of blocks of $`\pi `$. Indeed, constructing a $`2p`$-tuple $`(i_1,\mathrm{},i_{2p})`$ such that $`ker(i_1,\mathrm{},i_{2p})=\pi `$ amounts to constructing an injective function from the set of the blocks of $`\pi `$ to the set $`\{1,\mathrm{},ks\}`$; but the requirements $`i_1=r_1(mods)\mathrm{},i_{2p}=r_{2p}(mods)`$ allow only $`k`$ possible values for each of the values taken by this function – so even if the injectivity requirement is ignored, there still are at most $`k^m`$ such functions which can be constructed. On the other hand, it is obvious that every term of the sum on the right-hand side of (3.16) is less or equal 1 in absolute value (contractive linear functional applied to a unitary). We thus obtain that the quantity in (3.16) is bounded in absolute value by $`k^{mp}`$. But the facts that $`\pi `$ is $`\epsilon `$-null and is not a pairing imply $`mp1`$. (Indeed, every block of $`\pi `$ has an even number of elements, because $`\pi `$ is $`\epsilon `$-null; this implies $`mp`$, with equality holding if and only if every block of $`\pi `$ has exactly two elements – which we supposed is not the case.) Hence $`k^{mp}1/k`$, and (3.19) is obtained. Step 5. It is now the moment to consider a pairing $`\pi 𝒫(2p)`$, which is both $`r`$-stable and $`\epsilon `$-null – or in other words, an element $`\pi 𝒫(r_1,\mathrm{},r_{2p};\epsilon (1),\mathrm{},\epsilon (2p))`$. In this step of the proof we also fix some indices $`1i_1,\mathrm{},i_{2p}ks`$ such that $`ker(i_1,\mathrm{},i_{2p})=\pi `$ and such that $`i_1=r_1(mods),\mathrm{},i_{2p}=r_{2p}(mods)`$. The goal of the step is to give a good approximation for the integral $$_\mathrm{\Omega }\phi (U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_{2p}}(\omega )^{\epsilon (2p)})𝑑P(\omega ).$$ Let us write explicitly $`\pi =\{B_1,\mathrm{},B_p\}`$ where the blocks $`B_1,\mathrm{},B_p`$ are listed in increasing order of their minimal elements. The values $`i_{min(B_1)},`$ $`\mathrm{},i_{min(B_p)}\{1,\mathrm{},ks\}`$ are distinct, hence it makes sense to consider the permutation $$\sigma :=ord(i_{min(B_1)},\mathrm{},i_{min(B_p)})$$ of $`\{1,\mathrm{},p\}`$, which keeps track of their order ($`\sigma `$ defined as in Notation 3.3). We will show that: $$|_\mathrm{\Omega }\phi (U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_{2p}}(\omega )^{\epsilon (2p)})𝑑P(\omega )z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )}|$$ $$\frac{p(p1)}{2}\delta .$$ (3.20) Proof of Step 5. Let us consider the unitaries $$V_l(\omega ):=\left(U_{i_{min(B_l)}}(\omega )\right)^{\epsilon (min(B_l))},1lp,\omega \mathrm{\Omega }.$$ Note that: $$\left(U_{i_{max(B_l)}}(\omega )\right)^{\epsilon (max(B_l))}=V_l(\omega )^1,1lp,\omega \mathrm{\Omega };$$ this is because $`\epsilon (\text{max}(B_l))=\epsilon (\text{min}(B_l))`$ (which happens because $`\pi `$ is $`\epsilon `$-null), and $`i_{max(B_l)}=i_{min(B_l)}`$ (which comes from the fact that $`ker(i_1,\mathrm{},i_{2p})=\pi `$). On the other hand for every $`1l<mp`$ and every $`\omega \mathrm{\Omega }`$ we have the commutation relation $$V_l(\omega )V_m(\omega )=\left(\rho _{i_{min(B_l)},i_{min(B_m)}}(\omega )\right)^{\epsilon (min(B_l))\epsilon (min(B_m))}V_m(\omega )V_l(\omega ),$$ which is implied by the commutation relations known for the unitaries $`U_i(\omega )`$. But then the commutation Lemma 3.2 applies, and gives us that $$U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_{2p}}(\omega )^{\epsilon (2p)}$$ $$=\underset{\begin{array}{c}1l<mp\\ suchthat\\ B_lcrossesB_m\end{array}}{}\left(\rho _{i_{min(B_l)},i_{min(B_m)}}(\omega )\right)^{\epsilon (min(B_l))\epsilon (min(B_m))}I.$$ Hence we obtain: $$_\mathrm{\Omega }\phi (U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_{2p}}(\omega )^{\epsilon (2p)})𝑑P(\omega )$$ $$=_\mathrm{\Omega }\underset{\begin{array}{c}1l<mp\\ suchthat\\ B_lcrossesB_m\end{array}}{}\left(\rho _{i_{min(B_l)},i_{min(B_m)}}(\omega )\right)^{\epsilon (min(B_l))\epsilon (min(B_m))}dP(\omega )$$ $$=\underset{\begin{array}{c}1l<mp\\ suchthat\\ B_lcrossesB_m\end{array}}{}_\mathrm{\Omega }\left(\rho _{i_{min(B_l)},i_{min(B_m)}}(\omega )\right)^{\epsilon (min(B_l))\epsilon (min(B_m))}𝑑P(\omega );$$ (3.21) the product and the integration could be interchanged at the last equality sign because the random variables $`(\rho _{ij})_{1i<jks}`$ are independent (which immediately implies that the random variables $`(\rho _{i_{min(B_l)},i_{min(B_m)}})_{1l<mp}`$ are also independent). In the product (3.21), every factor is either within $`\delta `$ from $`z`$, or within $`\delta `$ from $`\overline{z}`$. In fact, one sees by direct inspection that: – if the crossing between $`B_l`$ and $`B_m`$ has positive orientation in $`(\pi ,\epsilon ,\sigma )`$, then $$|_\mathrm{\Omega }\left(\rho _{i_{min(B_l)},i_{min(B_m)}}(\omega )\right)^{\epsilon (min(B_l))\epsilon (min(B_m))}𝑑P(\omega )z|\delta ;$$ (3.22) – if the crossing between $`B_l`$ and $`B_m`$ has negative orientation in $`(\pi ,\epsilon ,\sigma )`$, then $$|_\mathrm{\Omega }\left(\rho _{i_{min(B_l)},i_{min(B_m)}}(\omega )\right)^{\epsilon (min(B_l))\epsilon (min(B_m))}𝑑P(\omega )\overline{z}|\delta .$$ (3.23) In order to check (3.22-23), there are four possible cases to discuss, according to whether $`i_{min(B_l)}`$ is bigger or smaller than $`i_{min(B_m)}`$, and whether $`\epsilon (\text{min}(B_l))\epsilon (\text{min}(B_m))`$ is 1 or $`1`$. We show one of them, say when $`i_{min(B_l)}>i_{min(B_m)}`$ and $`\epsilon (\text{min}(B_l))\epsilon (\text{min}(B_m))=1`$. The inequality $`i_{min(B_l)}>i_{min(B_m)}`$ is equivalent to $`\sigma (l)>\sigma (m)`$ (by the definition of the permutation $`\sigma `$ – see (3.9) in Notation 3.3); comparing this against the formulas (1.16), we see that $`B_l`$ and $`B_m`$ have a negative crossing. But on the other hand: $$_\mathrm{\Omega }\left(\rho _{i_{min(B_l)},i_{min(B_m)}}(\omega )\right)^{\epsilon (min(B_l))\epsilon (min(B_m))}𝑑P(\omega )$$ $$=\left(_\mathrm{\Omega }\overline{\rho _{i_{min(B_m)},i_{min(B_l)}}(\omega )}𝑑P(\omega )\right)$$ with $`i_{min(B_m)}<i_{min(B_l)}`$; this integral is within $`\delta `$ of $`\overline{z}`$, by one of the hypotheses of Proposition 1.6.2. Finally, (3.22) and (3.23) imply (3.20), via the well-known fact (easily checked by induction) that if $`\xi _1,\mathrm{},\xi _N,\eta _1,\mathrm{},\eta _N`$ are complex numbers of value not exceeding 1, and if $`|\xi _1\eta _1|\delta ,\mathrm{},|\xi _N\eta _N|\delta `$ then $`|\xi _1\mathrm{}\xi _N\eta _1\mathrm{}\eta _N|N\delta `$. (Here $`N`$ is the number of crossings of $`\pi `$, which cannot exceed $`p(p1)/2`$.) Step 6. In this step we fix again a pairing $`\pi 𝒫(r_1,\mathrm{},r_{2p};\epsilon (1),\mathrm{},\epsilon (2p))`$. We will prove the inequality: $$|T_\pi \frac{1}{p!}\underset{\sigma 𝒮_p}{}z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )}|<2^{p1}p^2\delta +\frac{(p+1)^p}{k}.$$ (3.24) Proof of Step 6. Let us write explicitly the partition $`\pi `$ fixed in this step as $`\{B_1,\mathrm{},B_p\}`$, where the blocks $`B_1,\mathrm{},B_p`$ are listed in increasing order of their minimal elements. Also, let us denote: $$𝒥=\left\{(i_1,\mathrm{},i_{2p})\begin{array}{cc}|\hfill & 1i_1,\mathrm{},i_{2p}ks,\hfill \\ |\hfill & ker(i_1,\mathrm{},i_{2p})=\pi ,\hfill \\ |\hfill & i_1=r_1(mods),\mathrm{},i_{2p}=r_{2p}(mods)\hfill \end{array}\right\};$$ i.e, $`𝒥`$ is the index set of the summation defining $`T_\pi `$ in Equation (3.16). For every $`2p`$-tuple $`(i_1,\mathrm{}i_{2p})𝒥`$ we write the inequality (3.20) obtained in Step 5; then we sum all these inequalities. The integrals from (3.20) will add up to $`k^pT_\pi `$. The terms “$`z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )}`$” from (3.20) will add up to: $$\underset{\sigma 𝒮_p}{}N(\sigma )z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )},$$ where for every $`\sigma 𝒮_p`$ we denoted $$N(\sigma )=|\{(i_1,\mathrm{},i_{2p})𝒥|ord(i_{min(B_1)},\mathrm{},i_{min(B_p)})=\sigma \}|.$$ We thus obtain, after also dividing by $`k^p`$: $$|T_\pi \frac{1}{k^p}\underset{\sigma 𝒮_p}{}N(\sigma )z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )}|\frac{1}{k^p}\underset{\sigma 𝒮_p}{}N(\sigma )\frac{p(p1)}{2}\delta .$$ (3.25) Now, the Lemma 3.4 gives us that: $$\left(\begin{array}{c}k\\ p\end{array}\right)N(\sigma )\left(\begin{array}{c}k+p\\ p\end{array}\right),\sigma 𝒮_p.$$ (3.26) One consequence of (3.26) is that the right-hand side of (3.25) is bounded above by: $$\frac{1}{k^p}p!\left(\begin{array}{c}k+p\\ p\end{array}\right)\frac{p(p1)}{2}\delta =\frac{(k+1)\mathrm{}(k+p)}{k^p}\frac{p(p1)}{2}\delta <2^{p1}p^2\delta $$ (3.27) (where at the last equality sign we used the fact that $`pk`$). Another consequence of (3.26) is that $$|\frac{1}{k^p}\underset{\sigma 𝒮_p}{}N(\sigma )z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )}$$ (3.28) $$\frac{1}{p!}\underset{\sigma 𝒮_p}{}z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )}|<\frac{(p+1)^p}{k}.$$ Indeed, the left-hand side of (3.28) can be written as: $$|\frac{1}{k^pp!}\underset{\sigma 𝒮_p}{}(p!N(\sigma )k^p)z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )}|.$$ (3.29) But for every $`\sigma 𝒮_p`$: $$|p!N(\sigma )k^p|\text{max}(|p!\left(\begin{array}{c}k\\ p\end{array}\right)k^p|,|p!\left(\begin{array}{c}k+p\\ p\end{array}\right)k^p|)$$ $$<\text{max}(k^p(kp)^p,(k+p)^pk^p)<k^{p1}(p+1)^p,$$ hence the the quantity in (3.29) is dominated by $`\frac{1}{k^pp!}p!k^{p1}(p+1)^p`$ = $`\frac{(p+1)^p}{k}`$. The inequality (3.24) (which is the goal of Step 6) is immediately obtained from (3.27) and (3.28). Step 7. In this final part of the proof, we combine the results of the previous steps in order to obtain the inequality (1.20) stated in Proposition 1.6.2. We first claim that: $$|E(X_{r_1}^{\epsilon (1)}\mathrm{}X_{r_{2p}}^{\epsilon (2p)})\underset{\pi 𝒫(r_1,\mathrm{},r_{2p};\epsilon (1),\mathrm{},\epsilon (2p))}{}T_\pi |<\frac{(2p)!}{k}.$$ (3.30) Indeed, we know that $$E(X_{r_1}^{\epsilon (1)}\mathrm{}X_{r_{2p}}^{\epsilon (2p)})=\underset{\pi Part(2p)}{}T_\pi \text{(by Step 1)}$$ $$=\underset{\begin{array}{c}\pi Part(2p),\\ \pi rstable\\ and\epsilon null\end{array}}{}T_\pi \text{(by Steps 2 and 3).}$$ By taking into account the Equation (3.13) of Notations 3.5, we see that the left-hand side of (3.30) is hence equal to $$|\underset{\begin{array}{c}\pi Part(2p)𝒫(2p),\\ \pi rstable\\ and\epsilon null\end{array}}{}T_\pi |;$$ but by the Step 4, this is bounded above by $`|Part(2p)𝒫(2p)|/k`$, which in turn is dominated by $`(2p)!/k`$ (we used the rough estimates $`|Part(2p)𝒫(2p)|<|Part(2p)|<(2p)!`$ ). Hence (3.30) is obtained. We next claim that $$|\underset{\pi 𝒫(r_1,\mathrm{},r_{2p};\epsilon (1),\mathrm{},\epsilon (2p))}{}T_\pi \underset{\pi 𝒫(r_1,\mathrm{},r_{2p};\epsilon (1),\mathrm{},\epsilon (2p))}{}\frac{1}{p!}\underset{\sigma 𝒮_p}{}z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )}|$$ $$<p!(2^{p1}p^2\delta +\frac{(p+1)^p}{k}).$$ (3.31) Indeed, if we write the inequality (3.24) obtained in Step 6 for every $`\pi 𝒫(r_1,\mathrm{},r_{2p};`$ $`\epsilon (1),\mathrm{},\epsilon (2p))`$, and if we sum over $`\pi `$, we obtain that the left-hand side of (3.31) is bounded above by $$|𝒫(r_1,\mathrm{},r_{2p};\epsilon (1),\mathrm{},\epsilon (2p))|(2^{p1}p^2\delta +\frac{(p+1)^p}{k}).$$ The latter quantity is in turn dominated by $`p!(2^{p1}p^2\delta +\frac{(p+1)^p}{k})`$, because the number of pairings in $`𝒫(2p)`$ which are $`\epsilon `$-null (but not necessarily $`r`$-stable) is exactly $`p!`$ . The desired inequality (1.20) immediately follows from (3.30), (3.31), and the rough estimates $`(2p)!+p!(p+1)^p<(2p+1)!`$, $`p!2^{p1}p^2<(2p+1)!`$. QED For the rest of the section we move to the framework of Proposition 1.6.3. The proof of 1.6.3 is in many respects similar to the one of 1.6.2. For this reason we will not write the arguments in the same detail, and occasionally we will leave it as an exercise to the reader to check that parts of the proof of 1.6.2 can be trivially adjusted to the current situation. 3.7 Proof of Proposition 1.6.3. Let $`E:(\mathrm{\Omega },𝒜)\text{C}`$ be the linear functional defined as in Equation (1.12) of Notation 3.4. We fix $`n1`$, and $`r_1,\mathrm{},r_n\{1,\mathrm{},s\}`$, $`\epsilon (1),\mathrm{},\epsilon (n)\{1,\}`$, about which we will show that the limit $$\underset{k\mathrm{}}{lim}E(X_{r_1;k}^{\epsilon (1)}\mathrm{}X_{r_n;k}^{\epsilon (n)})$$ (3.32) exists and is equal to the right-hand side of Equation (1.17). In connection to these $`n`$, $`r_1,\mathrm{},r_n`$, $`\epsilon (1),\mathrm{},\epsilon (n)`$ that are fixed, we will use combinatorial notations similar to some of those set in Notations $`\mathrm{3.5.2}^o`$: $`Part(n)`$ will denote the set of all the partitions of $`\{1,\mathrm{},n\}`$; and we will say that $`\pi Part(n)`$ is “$`r`$-stable” if $`r_a=r_b`$ whenever $`a,b\{1,\mathrm{},n\}`$ belong to the same block of $`\pi `$. We leave it as an exercise to the reader to verify that the Step 1 of the proof in Section 3.6 can be performed in the current situation, and leads to the following analogue of the Equations (3.15-16): $$E(X_{r_1;k}^{\epsilon (1)}\mathrm{}X_{r_n;k}^{\epsilon (n)})=\underset{\pi Part(n)}{}T_{\pi ,k},k1,$$ (3.33) where for $`\pi Part(n)`$ we set: $$T_{\pi ,k}:=\frac{1}{k^{n/2}}\underset{\begin{array}{c}1i_1,\mathrm{},i_nks\\ suchthatker(i_1,\mathrm{},i_n)=\pi and\\ i_1=r_1(mods),\mathrm{},i_n=r_n(mods)\end{array}}{}_\mathrm{\Omega }\phi (U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_n}(\omega )^{\epsilon (n)})𝑑P(\omega ).$$ (3.34) It is also clear that the Step 2 of the proof in Section 3.6 can be repeated identically, and leads to the conclusion that $`T_{\pi ,k}=0`$ for every $`k1`$ and every $`\pi Part(n)`$ which is not $`r`$-stable. Thus the partitions which are not $`r`$-stable can be ignored in the summation on the right-hand side of (3.33). We next observe that for every $`k1`$ and every $`\pi Part(n)`$ which is not $`r`$-stable, we have the inequality: $$|T_{\pi ,k}|k^{(\frac{n}{2}|\pi |)},$$ (3.35) where $`|\pi |`$ stands for the number of blocks of the partition $`\pi `$. The verification of this inequality is very similar to the argument shown in Step 4 of the proof in 3.6, and is left to the reader. Due to the fact that in (3.33) we are actually interested only in what happens when $`k\mathrm{}`$, the inequality (3.35) shows that in the summation on the right-hand side of (3.33) we can also safely ignore all the $`r`$-stable partitions $`\pi `$ such that $`|\pi |<n/2`$. Now let us remark that $`T_{\pi ,k}=0`$ for every $`k1`$ and for every $`\pi Part(n)`$ which has at least one singleton (i.e. a block with one element). Indeed, let us suppose that the partition $`\pi `$ has a one-element block $`B=\{b\}`$, $`1bn`$. Then for every $`1i_1,\mathrm{},i_nks`$ such that $`ker(i_1,\mathrm{},i_n)=\pi `$, the monomial $`U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_n}(\omega )^{\epsilon (n)}`$ is brought by the commutation relations (1.21) to the form $`cU_1(\omega )^{\lambda _1}\mathrm{}U_{ks}(\omega )^{\lambda _{ks}}`$, where $`|c|=1`$, $`\lambda _1,\mathrm{},\lambda _{ks}\text{Z}`$, and – most importantly here – $`\lambda _{i_b}=\pm 1`$. But then the hypothesis (d) of Proposition 1.6.3 gives that $`\phi (U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_n}(\omega )^{\epsilon (n)})`$ = 0, and the equality $`T_{\pi ,k}=0`$ follows. The conclusion of the preceding three paragraphs is that in the summation on the right-hand side of (3.33) we may keep (without affecting what happens when $`k\mathrm{}`$) only the terms which correspond to partitions $`\pi Part(n)`$ that are $`r`$-stable, satisfy $`|\pi |n/2`$, and have no singletons. However, if $`n`$ is odd, then there are no partitions at all which satisfy $`|\pi |n/2`$ and at the same time have no singletons. This simply means that if $`n`$ is odd, then the limit in (3.32) exists and is equal to 0 (and the case of odd $`n`$ is thus settled). If $`n`$ is even, it is immediate that a partition $`\pi Part(n)`$ satisfies $`|\pi |n/2`$ and has no singletons if and only if it is a pairing. Thus in the case of even $`n`$, the summation on the right-hand side of (3.33) can be restricted to the set of $`r`$-stable pairings of $`\{1,\mathrm{},n\}`$. From now on and until the end of the proof we will assume that $`n`$ is even, $`n=2p`$ with $`p`$ positive integer. Similarly to the terminology introduced in the Notations $`\mathrm{3.5.2}^o`$, we will say that a pairing $`\pi `$ = $`\{\{a_1,b_1\},\mathrm{},\{a_p,b_p\}\}`$ of $`\{1,\mathrm{},n\}`$ is $`\epsilon `$-null if $`\epsilon (a_i)\epsilon (b_i)`$, $`1ip`$ (where $`\epsilon (1),\mathrm{},\epsilon (n)\{1,\}`$ are as fixed at the beginning of the proof). By taking into account the conclusion of the preceding paragraph, and by examining at the same time the right-hand side of Equation (1.17), we see that the proof will be completed if we can show that: $$\underset{k\mathrm{}}{lim}T_{\pi ,k}=0$$ (3.36) for every pairing $`\pi `$ of $`\{1,\mathrm{},2p\}`$ which is $`r`$-stable but not $`\epsilon `$-null; and $$\underset{k\mathrm{}}{lim}T_{\pi ,k}=\frac{1}{p!}\underset{\sigma 𝒮_p}{}\underset{\pi 𝒫(r_1,\mathrm{},r_{2p};\epsilon (1),\mathrm{},\epsilon (2p))}{}z^{cr_+(\pi ,\epsilon ,\sigma )}\overline{z}^{cr_{}(\pi ,\epsilon ,\sigma )}$$ (3.37) for every pairing $`\pi `$ of $`\{1,\mathrm{},2p\}`$ which is both $`r`$-stable and $`\epsilon `$-null. The limit in (3.36) holds trivially: $`T_{\pi ,k}=0`$ for every $`k1`$ and every pairing $`\pi `$ which is $`r`$-stable but not $`\epsilon `$-null. This is a direct application of the hypothesis (d) in Proposition 1.6.3, and is left to the reader. (The discussion is similar to the one which ruled out the partitions with singletons, but this time one uses the case when there exists a $`\lambda _j`$ equal to $`\pm 2`$.) So it suffices if from now on we fix a pairing $`\pi `$ of $`\{1,\mathrm{},2p\}`$ which is both $`r`$-stable and $`\epsilon `$-null, and we prove that the limit (3.37) holds. We denote the quantity on the right-hand side of (3.37) by $`Q_\pi `$. We will also fix a number $`\beta >0`$, and we will show that $`|T_{\pi ,k}Q_\pi |<\beta `$ if $`k`$ is sufficiently large. Denote $`\delta :=\beta /(2^pp^2)`$ and $`L:=n+1`$. By the hypotheses (b) and (c) of Proposition 1.6.3, there exists $`m_o1`$ such that for every $`m_om<n`$ we have that $`|z_\mathrm{\Omega }\rho _{m,n}|\delta `$, and that $`U_m,U_{m+1},\mathrm{},U_n`$ is an $`L`$-mimic of a Haar family. We fix $`k_o`$ such that $`k_os+1m_o`$. For every $`k>k_o`$ we will write $`T_{\pi ,k}`$ as a sum, $$T_{\pi ,k}=T_{\pi ,k}^{}+T_{\pi ,k}^{\prime \prime },$$ (3.38) by splitting the index set of the sum in (3.34), which defines $`T_{\pi ,k}`$, into two disjoint parts: in $`T_{\pi ,k}^{}`$ we take the terms indexed by $`n`$-tuples $`(i_1,\mathrm{},i_n)`$ such that $`k_os+1i_1,\mathrm{},i_nks`$, and in $`T_{\pi ,k}^{\prime \prime }`$ we take the rest of the terms (indexed by $`n`$-tuples $`(i_1,\mathrm{},i_n)`$ such that $`\mathrm{min}(i_1,\mathrm{},i_n)k_os`$). Note that for every $`k>k_o`$, the random variables $`(\rho _{i,j})_{k_os+1i<jks}`$ and the random unitaries $`(U_i)_{i=k_os+1}^{ks}`$ fall under the hypotheses of Proposition 1.6.2 (for the chosen values of $`\delta `$ and $`L`$); thus the estimates found in the proof of Proposition 1.6.2 apply to this situation. Out of these estimates, the one which we need here is the inequality (3.24) established in the Step 6 of Section 3.6. When reporting to the current notations, “$`T_\pi `$” of (3.24) has to be replaced by: $$\frac{1}{(kk_o)^p}\underset{\begin{array}{c}k_os+1i_1,\mathrm{},i_{2p}ks\\ suchthatker(i_1,\mathrm{},i_{2p})=\pi and\\ i_1=r_1(mods),\mathrm{},i_{2p}=r_{2p}(mods)\end{array}}{}_\mathrm{\Omega }\phi (U_{i_1}(\omega )^{\epsilon (1)}\mathrm{}U_{i_{2p}}(\omega )^{\epsilon (2p)})𝑑P(\omega );$$ but this is exactly $`k^p/(kk_o)^pT_{\pi ,k}^{}`$, with $`T_{\pi ,k}^{}`$ taken from (3.38). So the inequality (3.24) becomes in this situation: $$\left|\left(\frac{k}{kk_o}\right)^pT_{\pi ,k}^{}Q_\pi \right|<2^{p1}p^2\delta +\frac{(p+1)^p}{kk_o};$$ or after multiplication with $`(kk_o)^p/k^p<1`$, and after taking into account the relation between $`\beta `$ and $`\delta `$: $$\left|T_{\pi ,k}^{}\left(\frac{kk_o}{k}\right)^pQ_\pi \right|<\frac{\beta }{2}+\frac{(p+1)^p}{kk_o},k>k_o.$$ (3.39) But on the other hand, a counting argument very similar to the one shown in Step 4 of Section 3.6 shows that for $`k>k_o+p`$ there are less than $`k^p`$ terms in the summation defining $`T_{\pi ,k}`$, and there are more than $`(kk_op)^p`$ terms in the summation defining $`T_{\pi ,k}^{}`$; this implies that there are less than $`k^p(kk_op)^p`$ terms in the summation defining $`T_{\pi ,k}^{\prime \prime }`$, and consequently that: $$|T_{\pi ,k}^{\prime \prime }|<\frac{k^p(kk_op)^p}{k^p},k>k_o+p.$$ (3.40) Finally, for $`k>k_o+p`$ we can write: $$|T_{\pi ,k}Q_\pi ||T_{\pi ,k}^{}\left(\frac{kk_o}{k}\right)^pQ_\pi |+|1\left(\frac{kk_o}{k}\right)^p|\left|Q_\pi \right|+\left|T_{\pi ,k}^{\prime \prime }\right|$$ $$<\frac{\beta }{2}+\frac{(p+1)^p}{kk_o}+|1\left(\frac{kk_o}{k}\right)^p|\left|Q_\pi \right|+1\left(\frac{kk_op}{k}\right)^p$$ ( by Equations (3.39) and (3.40) ), and the latter expression is clearly smaller than $`\beta `$ if $`k`$ is large enough. QED Acknowledgement: The second-named author acknowledges the hospitality of the Henri Poincaré Institute (Centre Emile Borel – UMS 839 IHP CNRS/UPMC) in Paris, where he visited during the final stage of the preparation of this paper.
warning/0001/cond-mat0001190.html
ar5iv
text
# Density Fluctuations in Molten Lithium: Inelastic X-Ray Scattering Study. ## I INTRODUCTION The dynamics of liquid metals has been extensively investigated in the recent past with the main purpose of ascertaining the role of the mechanisms underlying both collective and single-particle motions at the microscopic level. In the special case of collective density fluctuations, it is known that well-defined oscillatory modes can be supported even outside the strict hydrodynamic region. In molten alkali metals, this feature is found to persist down to wavelengths of one or two interparticle distances, making these systems excellent candidates to test the various theoretical approaches developed so far for the microdynamics of the liquid state. As a consequence, since the pioneering inelastic neutron scattering (INS) study by Copley and Rowe in liquid rubidium the interest in performing more and more accurate experiments is continuously renewed: INS investigations have been devoted to liquid cesium , sodium , lithium , potassium and again rubidium . Up to a few years ago the only experimental probes adequate to access the interparticle distance region in collective dynamics were thermal neutrons. With this technique foundamental results have been achieved in the field of condensed matter structure and dynamics. Unfortunately, in several systems their use for the determination of the dynamic structure factor $`S(Q,E)`$ becomes extremely difficult (if not impossible) for two reasons. The first one reflects the presence of an incoherent contribution to the total neutron scattering cross section. In liquid sodium, for example, the incoherent cross section dominates; even in more favorable cases (Li, K) at small $`Q`$ the intensity of the collective contribution is low, and its extraction requires a detailed knowledge of the single particle dynamics. The second reason is dictated by the need of satisfying both the energy and momentum conservation laws which define the $`(QE)`$ region accessible to the probe . Roughly speaking, when the sound speed of the system exceeds the velocity of the probing neutrons ($`1500`$ m/s for thermal neutrons) collective excitations cannot be detected for $`Q`$ values below $`Q_m,`$ the position of the first sharp diffraction peak of the sample, namely just in the most significant region for the collective properties. By virtue of the $`m^{1/2}`$ dependence of the isothermal sound speed $`c_0`$ (see section II A) the higher the atomic number of the system, the wider is the kinematic region accessible to neutrons, so that accurate INS data are available for rubidium ($`c_01260`$ m/s) and cesium ($`c_0970`$ m/s) , while more difficulties are met in the case of lighter metals. In particular, lithium represents the most critical case due to its high sound speed ($`c_04500`$ m/s) and to the equality of the coherent and incoherent cross sections: for this reason INS aiming to the study of collective properties of Li represents a very hard challenge . From a general point of view the main outcome of all these INS experiments, as far as collective properties are concerned, is the evidence of inelastic excitations in $`S(Q,E)`$ that, in the specific cases of rubidium and cesium, exhibit a positive dispersion, i.e. an increase of $`\omega _m(Q)`$ (the position of the $`S(Q,\mathrm{}\omega )`$ peaks) with respect to the values implied by the hydrodynamic sound speed. About the details of the relaxations driving such dispersion not much can be inferred at this level: the experimental results have been necessarily analyzed within simple models such as the damped harmonic oscillator or the kinetic model , suitable to extract reliable and resolution-corrected information on the peak positions but not about the lineshape features. Some extra efforts have been done in the case of cesium where informations about an average relaxation time have been extracted . A very useful tool, complementary to the ”traditional” experimental facilities is the numerical simulation technique, in particular Molecular Dynamics (MD): the choice of a realistic interatomic potential - i.e. a potential model able to reproduce structural properties - allows the determination of the dynamics of the system via the integration of the classical Newton equations. In this framework the single particle and the collective properties can easily be investigated within technical restrictions due to the finite box size (defining the minimum accessible wavevector) and computation time (related to the statistical quality and to the energy resolution of the calculated spectra). Broadly speaking the features of the atomic collective motion i.e. the details of the $`S(Q,E)`$ lineshape, as outcome of MD run, turns out to be less noisy and more straightforward than the correspondent INS results: no absolute normalizations are required, no mixing between coherent/incoherent dynamics occurs and, above all, basically no resolution corrections are needed. After the first experiments on rubidium, a considerable number of MD works have been published in this field, giving a valuable support to the experimental measurements . On the specific case of liquid lithium, the performed MD simulations show features common to all the other molten alkali metals, like the presence of positive dispersion - a particularly relevant result that could have not be achieved with neutrons for the previously mentioned reasons. At temperature around $`T_m`$ the increase of the sound velocity is of the order of 20%, and not much is known about its microscopic origin. In the recent past, the development of new synchrotron radiation facilities opened the possibility of using X-rays to measure the $`S(Q,\omega )`$ (which is proportional to the scattered intensity) in the non-hydrodynamic region; in this case the photon speed is obviously much larger than the excitations velocity, and no kinematic restriction occurs. Moreover, in a monatomic system as lithium, the scattering cross section is purely coherent and so it is directly associated with the dynamic structure factor. Some experiments have been performed on liquid lithium with progressively increasing resolution; it has been possible to show the existence of propagating collective excitations, but due to resolution limitations (never below $`\mathrm{\Delta }E_{FWHM}12`$ meV), no detailed information about the lineshape could be extracted. In this work we report the results of Inelastic X-ray Scattering (IXS) experiments performed on liquid lithium with very high energy resolution, adequate to probe accurately the detailed features of $`S(Q,E)`$ \- this allows to investigate whether the lineshape dependence on momentum transfer can be interpreted within the phenomenology of one or more relaxation processes. In particular in Section II we review the basic theoretical framework adopted for the interpretation of the IXS results. The latter are reported in Section III together with a brief account of the experimental setup. Section IV is devoted to the data analysis and discussion. The main outcome of the present paper is finally summarized in Section V along with some concluding remarks. ## II BASIC THEORY ### A Definitions In the classical limit, the basic time correlation probing collective dynamics in a monatomic fluid ($`N`$ particles with mass $`m`$) is the intermediate scattering function $$F(Q,t)=(1/N)\underset{i,j}{}e^{i𝐐𝐫_i(0)}e^{i𝐐𝐫_j(t)}$$ (1) where $`𝐫_i(t)`$ denotes the position of the j-th particle at time $`t`$. The dynamic structure factor $`S(Q,\omega )`$ is the frequency spectrum of $`F(Q,t)`$, while the structural features are accounted for the initial value $`F(Q,t=0)=S(Q)`$. The quantity $`F(Q,t)`$ then obeys the equation $`\stackrel{..}{F}(Q,t)`$ $`+\omega _0^2(Q)F(Q,t)`$ (3) $`+{\displaystyle _0^t}M(Q,tt^{})\stackrel{.}{F}(Q,t^{})dt^{}=0`$ where $$\omega _0^2(Q)=K_BTQ^2/mS(Q)$$ (4) is the second classical normalized frequency moment of $`S(Q,\omega ),`$ while $`M(Q,t)`$ is the so called memory function of the system, related in some way to the details of the Hamiltonian . In the limit $`Q0,`$ $`\omega _0^2(Q)c_0^2Q^2`$, where $`c_0`$ is the isothermal sound velocity; hence the quantity $`c_0`$ $`(Q)\omega _0(Q)/Q`$ can be interpreted as a suitable generalization of $`c_0`$ to finite wavevectors. From Eq.(3) a formally exact representation of the Laplace transform of $`F(Q,t)`$ can be written as $`\stackrel{~}{F}(Q,z)`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}𝑑te^{zt}F(Q,t)`$ (5) $`=`$ $`S(Q)\left\{z+{\displaystyle \frac{\omega _0^2(Q)}{[z+\stackrel{~}{M}(Q,z)]}}\right\}^1`$ (6) From the knowledge of $`\stackrel{~}{F}(Q,z)`$ one straightforwardly obtains $`S(Q,\omega )=(1/\pi )Re\stackrel{~}{F}(Q,z=i\omega )`$ in terms of the real ($`M^{}`$) and imaginary ($`M^{\prime \prime }`$) parts of the Fourier-Laplace transform of the memory function: $$S(Q,\omega )=\frac{S(Q)}{\pi }\frac{\omega _0^2(Q)M^{}(Q,\omega )}{\left[\omega ^2\omega _0^2+\omega M^{\prime \prime }(Q,\omega )\right]^2+\left[\omega M^{}(Q,\omega )\right]^2}$$ (7) The spectral features of the dynamic structure factor can be characterized by its frequency moments $`\mathrm{\Omega }_S^{(n)}(Q)\omega ^nS(Q,\omega )𝑑\omega `$, where, for a classical system, only the even frequency moments (such as $`\mathrm{\Omega }_S^{(0)}(Q)=S(Q)`$ and $`\mathrm{\Omega }_S^{(2)}(Q)=S(Q)\omega _0^2(Q)`$) are different from zero. In the following, we shall also find convenient to consider the ”longitudinal current spectrum” defined as $`C_L(Q,\omega )=\left(\omega ^2/Q^2\right)S(Q,\omega )`$. The presence of the factor $`\omega ^2`$ wipes out the low frequency portion of the dynamic structure factor, and consequently emphasizes the genuine inelastic features of $`S(Q,\omega )`$. After its definition and Eq.(5), it is readily seen that the Laplace transform $`\stackrel{~}{C}_L(Q,z)`$ satisfies $`\stackrel{~}{C}_L(Q,z)`$ $`=`$ $`z[z\stackrel{~}{F}(Q,z)S(Q)]`$ (8) $`=`$ $`\left\{z+[\omega _0^2(Q)/z]+\stackrel{~}{M}(Q,z)\right\}^1`$ (9) Again, the spectrum $`C_L(Q,\omega )`$ can be expressed as $`(1/\pi )Re\stackrel{~}{C}_L(Q,z=i\omega )`$. Then the position and the width of the inelastic peaks in $`C_L(Q,\omega )`$ are determined by the poles of $`\stackrel{~}{C}_L(Q,z)`$. The above definitions are valid for a classical system. In the case of our interest, the main effect of quantum-mechanical corrections stems from the well-known inequality of the positive and negative-frequency parts of the spectra, connected by the detailed balance factor $`e^{\beta \mathrm{}\omega }`$. Additional sources of non-classical behavior, such as those associated with a finite value of the deBoer wavelength $`\mathrm{\Lambda }=\left(2\pi \mathrm{}^2/mK_BT\right)^{1/2}`$, are small (in the explored lithium states $`\mathrm{\Lambda }`$ is only $`0.11`$ times the average interparticle distance) and can safely be neglected. Since the effects of the detailed balance are clearly visible in the experimental IXS spectra, we briefly discuss a possible procedure to account for this quantum feature in a consistent way, while preserving the inherent advantages of the classical description. In doing this, for the sake of clarity we shall denote all the previous classical quantities with the subscript cl, while the notation q will refer to the quantum case. The natural theoretical counterpart of the classical density correlation function is the so called Kubo canonical relaxation function $$K_q(Q,t)=\frac{1}{\beta N}\underset{i,j}{}_0^\beta 𝑑\lambda e^{i𝐐\widehat{𝐫}_i(0)}e^{\lambda \widehat{H}}e^{i𝐐\widehat{𝐫}_j(t)}e^{\lambda \widehat{H}}$$ (10) where $`\beta =1/K_BT`$ and the angular brackets denote a quantum statistical average. In the classical limit ($`\beta 0,`$ $`\mathrm{}0`$) the operators $`\widehat{A}`$ become classical commuting dynamical variables and $`K_q(Q,t)F_{cl}(Q,t).`$ It can be shown that $`K_q(Q,t)`$ is a real even function of time, so that its spectrum $`K_q(Q,\omega )`$ is an even function of frequency. On the other hand, the experimental scattering cross section involves the Fourier transform $`S_q(Q,\omega )`$ of the quantum density correlator $`F_q(Q,t)=(1/N)_{i,j}e^{i𝐐\widehat{𝐫}_i(0)}e^{i𝐐\widehat{𝐫}_j(t)}.`$ The relation between $`S_q(Q,\omega )`$ and $`K_q(Q,\omega )`$ reads $`S_q(Q,\omega )={\displaystyle \frac{\beta \mathrm{}\omega }{1e^{\beta \mathrm{}\omega }}}K_q(Q,\omega )`$ and satisfies the detailed balance condition. Moreover the condition $$\mathrm{\Omega }_K^{(2n)}=\frac{2}{\beta \mathrm{}}\mathrm{\Omega }_S^{(2n1)}$$ (11) relates the even frequency moments of $`K_q`$ with the odd ones of $`S_q.`$ The same memory function framework of Eq. (5) can be phrased for the Kubo relaxation function and for its Laplace transform $`\stackrel{~}{K}_q(Q,z)`$. By virtue of all these properties, in a situation where the quantum aspects not associated with detailed balance are marginal, it is reasonable (although not strictly rigorous) to identify the spectrum $`K_q(Q,\omega )`$ with the classical quantity $`S_{cl}(Q,\omega )`$ so that $$S_q(Q,\omega )\frac{\beta \mathrm{}\omega }{1e^{\beta \mathrm{}\omega }}S_{cl}(Q,\omega )$$ (12) Having established such a correspondence, from now on we will drop out the subscript cl and refer to the classical quantities as in fact done at the beginning of this section. ### B The memory function features The memory function $`M(Q,t)`$ accounts for all the relaxation mechanisms affecting collective dynamics, and consequently is the central quantity in most theoretical approaches. From straightforward algebra, the initial value of $`M(Q,t)`$ is related to the spectral moments of $`S(Q,\omega )`$ by: $$M(Q,t=0)=\frac{\mathrm{\Omega }_S^{(4)}(Q)}{\mathrm{\Omega }_S^{(2)}(Q)}\mathrm{\Omega }_S^{(2)}(Q)$$ (13) An exact expression of $`\mathrm{\Omega }_S^{(4)}(Q)`$ exists, but involves both the derivatives of the interparticle potential and the pair distribution function . It is usual to define $`\omega _L^2(Q)=\mathrm{\Omega }_S^{(4)}(Q)/\mathrm{\Omega }_S^{(2)}(Q)`$ and $`\mathrm{\Delta }^2(Q)=\omega _L^2(Q)\omega _0^2(Q)`$, their meaning being evident from the following argument. For sufficiently large $`\left|z\right|`$, $`\stackrel{~}{M}(Q,z)M(Q,t=0)/z`$ and Eq. (8) is seen to have poles at $`z=\pm i\sqrt{\omega _0^2(Q)+\mathrm{\Delta }^2(Q)}=\pm i\omega _L(Q)`$, showing that the frequency $`\omega _L(Q)`$ characterizes the instantaneous collective response of the liquid at the wavevector $`Q`$. Similar remarks can be made for the generalized infinite-frequency velocity $`c_{\mathrm{}}(Q)\omega _L(Q)/Q.`$ In liquid systems, in the long time limit, one expects that $`M(Q,t\mathrm{})`$ approach zero value, therefore, regardless the details of its shape, $`M(Q,t)`$ it is expected to decay over a certain timescale $`\tau (Q)`$ that, for the sake of simplicity, can be defined as $`\tau (Q)=\mathrm{M}(Q)/\mathrm{\Delta }^2(Q),`$ being $`\mathrm{M}(Q)=_0^{\mathrm{}}M(Q,t)𝑑t.`$ It can be of considerable interest to point out the asymptotic behaviors of Eq. (3) in the opposite regimes $`\tau (Q)\omega _0(Q)<<`$ or $`>>1`$. In the first limit ($`M(Q,t)2\mathrm{M}(Q)\delta (t)`$) Eq. (3) reduces to the equation for a damped harmonic oscillator, and it can be easily proved that $`M^{}(Q,\omega )=`$ $`\mathrm{M}(Q)`$ and $`M^{\prime \prime }(Q,\omega )=0,`$ which means two inelastic peaks in the $`S(Q,\omega )`$ (in the current spectra they are centered at $`\pm \omega _0(Q)`$) damped with a factor $`\mathrm{M}(Q).`$ In the opposite limit, $`\tau (Q)\omega _0(Q)>>1,`$ the decay of $`F(Q,t)`$ is much faster than $`\tau (Q)`$ and $`M(Q,t)`$ appearing in the convolution integral of Eq. (3) can be considered constant, $`M(Q,t)`$ $`\mathrm{\Delta }^2(Q)`$. Therefore Eq. (3) becomes non homogeneous and the solution of Eq. (7) reduces to a harmonic oscillator of frequency $`\omega _L(Q)`$ with no damping, plus a sharp elastic line. In this extreme case, the intensity ratio between the elastic and inelastic lines (the Debye Waller factor) in $`S(Q,\omega )`$ is $`f(Q)=1\omega _0^2(Q)/\omega _L^2(Q)`$. The true time dependence of the memory function is a priori unknown, and from the very start it is convenient to separate in $`M(Q,t)`$ the decay channels which explicitly involve couplings to thermal fluctuations ( $`M_{th}(Q,t)`$ ) from those directly associated with longitudinal density modes ( $`M_L(Q,t)`$ ). A convenient way to perform this splitting is by the use of the generalized hydrodynamics theory . In the simplest version of this approach, one writes $`M(Q,t)`$ $`=`$ $`M_L(Q,t)+M_{th}(Q,t)`$ (14) $`=`$ $`\mathrm{\Delta }_L^2(Q)m_L(Q,t)+\mathrm{\Delta }_{th}^2(Q)m_{th}(Q,t)`$ (15) with $`\mathrm{\Delta }_L^2(Q)\omega _L^2(Q)\gamma (Q)\omega _0^2(Q)`$, $`\mathrm{\Delta }_{th}^2(Q)=[\gamma (Q)1]\omega _0^2(Q)`$. Here $`\gamma (Q)`$ is a generalization of the specific heat ratio $`\gamma =C_P/C_V`$ to finite wavevectors. The relaxation processes underlying the dynamics are accounted for by the normalized quantities $`m_L(Q,t)`$ and $`m_{th}(Q,t)`$, defined in such a way that $`m_L(Q,0)=m_{th}(Q,0)=1`$. A straightforward generalization of ordinary hydrodynamics suggests for the thermal contribution the following form $$m_{th}(Q,t)exp[a(Q)Q^2t]$$ (16) where $`a(Q)`$ can be viewed as a finite $`Q`$ generalization of the quantity $`D_T=\kappa /nC_v,`$ being $`\kappa `$ the thermal conductivity. The role of the thermal contribution in the dynamic structure factor of liquid metals will be discussed in the following section. In contrast with the thermal decay channel, no guidance for $`M_L(Q,t)`$ is provided by ordinary hydrodynamics. In the latter, one implicitly assumes that $$M_L(Q0,t)2(\eta _L/nm)Q^2\delta (t)$$ (17) where the longitudinal viscosity coefficient $`\eta _L`$ is related to the ordinary shear and bulk viscosities by $`\eta _L=(4/3)\eta +\eta _B`$. Clearly in the limit $`Q0`$ the ratio $`\stackrel{~}{M}_L(Q,z=0)/Q^2`$ approaches $`\eta _L/nm`$. The simplest way to go beyond the hydrodynamic result (17) is to allow for a finite decay rate of $`M_L(Q,t)`$: $$M_L(Q,t)=\mathrm{\Delta }_L^2(Q)e^{t/\tau (Q)}$$ (18) where it has been additionally assumed an exponential lineshape for the memory function. Although this has the advantage of analytical simplicity when dealing with Fourier transform, a drawback of this ansatz lies in the violation of some basic short-time features of the memory function (such as a zero derivative at $`t=0`$), causing the divergency of $`\mathrm{\Omega }_S^{(n)}`$ for $`n6.`$ Eq. (18) yields the so-called viscoelastic model for $`S(Q,\omega )`$ . Since as $`Q0`$ $`\stackrel{~}{m}_L(Q,z=0)/Q^2`$ can written as $`[c_{\mathrm{}}^2c_0^2]\tau (Q0)`$, the requirement that this coincides with $`\eta _L/nm`$ shows that the time $`\tau (Q)`$ must be finite as $`Q0`$. Such a connection with viscous effects justifies the physical interpretation of the rate $`1/\tau (Q)`$ as a parameter giving an overall account of all relaxation processes by which the longitudinal response of the liquid is affected by time-dependent disturbances. In particular, for slow perturbations developing over a timescale $`t\tau (Q)`$ the system can adjust itself to attain local equilibrium and the usual viscous behavior. In contrast, for disturbances fast enough that $`t\tau (Q)`$ the liquid responds instantaneously, with a solid-like (elastic) behavior. The crossover between these limiting situations (times $`t\tau (Q)`$, or frequencies $`\omega `$ such that $`\omega \tau (Q)1`$) is ultimately responsible for the gradual changes often detected in the sound dispersion of several liquids at increasing wavevectors. Similar considerations can be applied to the $`M_{th}(Q,t)`$ contribution which involves the timescale $`\tau _{th}=D_T^1(Q)Q^2;`$ in this case the ”elastic response ” is achieved in the hydrodynamic regime: $`\omega _0(Q0)\tau (Q0)c_0/Q`$ $`\mathrm{},`$ while the ”viscous response ” emerges at short wavelengths. The crossover between the two limits embodies the transition between the adiabatic and isothermal regimes, and reflects the actual possibility of density fluctuations to decay by the thermal channel: at high frequencies there is no time for such conversion, and the fluctuations evolve without heat transfer (higher sound speed, no attenuation) while in the low frequency limit there is enough time for the system to equilibrate so that an isothermal dynamics takes place (lower sound speed, damped excitation). These qualitative considerations are supported by the evolution of the poles of $`\stackrel{~}{C}_L(Q,z)`$ at increasing wavevectors. Eqs. (14), (16) and (18) in fact imply that $$\stackrel{~}{M}(Q,z)=\frac{\mathrm{\Delta }_L^2(Q)}{z+1/\tau (Q)}+\frac{\mathrm{\Delta }_{th}^2(Q)}{z+a(Q)Q^2}$$ (19) In the hydrodynamic regime ($`Q0`$) both the inequalities $`\left|z\right|1/\tau (0)`$ and $`\left|z\right|`$ $`a(Q)Q^2`$ are satisfied, and the poles are approximately located at $`z\pm i\sqrt{\omega _0^2(Q0)+\mathrm{\Delta }_{th}^2(Q0)}=\pm i`$ $`c_s`$ $`Q`$, where $`c_s\sqrt{\gamma }`$ $`c_0`$ is the adiabatic sound velocity. At larger wavevectors different situations may arise: the one of our interest is $`\left|z\right|a(Q)Q^2`$ , which is readily seen to cause a shift of the imaginary part of the poles toward an isothermal behavior, i.e. $`\omega \pm \omega _0(Q)`$. The last limiting case occurs whenever the frequency $`\omega `$ becomes distinctly larger than $`\left[1/\tau (Q)\right]`$, which eventually yields poles at $`z=\pm \omega _L(Q)`$, as already noticed. Although appealing, the simplicity of the viscoelastic model can be deceptive. First of all, the model itself provides no clue for the physical origin of the decay mechanisms leading to the rate $`1/\tau (Q)`$. Secondly, even ignoring this aspect and treating $`1/\tau (Q)`$ as a fitting parameter, the practical results of the model are rather unsatisfactory (see Fig. 5 in the following). These drawbacks of the simple viscoelastic model have been theoretically predicted, and somehow reported in a number of liquids through a detailed analysis of MD spectra . However, to our knowledge, no cutting edge analysis exists in the case of experimental measurements of the coherent dynamic structure factor, probably because of the afore-mentioned limits of the neutron technique. The obvious remedy is to modify the simple ansatz (18) by allowing a more sophisticated decay of $`M_L(Q,t)`$. We adopted the following two-exponential ansatz: $$M_L(Q,t)=\mathrm{\Delta }_L^2(Q)\left[(1\alpha (Q))e^{\gamma _1(Q)t}+\alpha (Q)e^{\gamma _2(Q)t}\right]$$ (20) where the rate $`\gamma _1(Q)`$ is chosen to be larger than $`\gamma _2(Q)`$, so that the dimensionless factor $`\alpha (Q)`$ measures the relative weight of the ”slow” decay channel. Besides being more flexible than the viscoelastic model, we shall see that the ansatz (20) has the much more important merit that the presence of two different timescale does in fact have a definite physical interpretation. The previous ansatz has been proposed phenomenologically several years ago and tested against the MD results in LJ fluids , it is interesting to note that an expression analogous to Eq. (20) has been implicitly introduced in the viscoelastic analysis of Brillouin Light Scattering (BLS) spectra of glass forming materials (see for example ref. ). In fact, in these works, the general expression of $`M_L(Q,t)`$ for a two times exponential decay is always expressed as $$M_L(Q,t)=\mathrm{\Delta }_\mu ^2(Q)e^{t/\tau _\mu }+\mathrm{\Delta }_\alpha ^2(Q)e^{t/\tau _\alpha }$$ (21) with explicit reference to the so-called $`\alpha `$\- (structural) relaxation process as responsible for the long lasting tail, and to the $`\mu `$ microscopic process as additional, faster, relaxation dominant over a very short timescale. To be more precise, in the BLS spectral window the condition $`\omega \tau _\mu <<1`$ holds, so that the approximation $`M_L(Q,t)=2\mathrm{\Delta }_\mu ^2(Q)\tau _\mu \delta (t)+\mathrm{\Delta }_\alpha ^2(Q)e^{t/\tau _\alpha }`$ is customarily adopted as a fitting model. On the contrary, we will show in this work how such an approximation is no longer tenable in the case of liquid Lithium at the IXS frequencies. The origin, at the atomic level, of this fast decay channel is still an open issue: one of the purposes of the present paper is to point out the features of such relaxation and to inspect its nature. The rapidly decaying portion of $`M_L(Q,t)`$ is customarily attributed to largely uncorrelated collisional events, similar to those occurring in a dilute fluid. In addition, at the high densities typical of the liquid state, non-negligible correlations among the collisions can be expected, making no longer valid an interpretation only in terms of ”binary” collisions. Albeit the magnitude of the correlation effects is relatively small and their buildup slow, once established their decay is even slower, and for $`t>1/\gamma _1(Q)\tau _\mu `$ this relaxation channel dominates the decay of $`M_L(Q,t)`$, which consequently may exhibit a small but long-lasting ”tail”. The ansatz (20) can incorporate most of this physics: on the basis of the latter, one may reasonably anticipate that $`\alpha (Q)1`$, and that the time $`1/\gamma _2(Q)\tau _\alpha `$ is distinctly longer than $`1/\gamma _1(Q)\tau _\mu `$. In this picture, the best fitted values of the viscoelastic rate $`1/\tau (Q)`$ clearly represent some sort of ”weighted average” between $`\gamma _1(Q)`$ and $`\gamma _2(Q)`$. Finally, we may argue that at increasing $`Q`$ (namely, over a shrinking length scale) the magnitude $`\alpha (Q)`$ of correlation effects should decrease, and that at higher temperatures the value of $`\alpha (Q)`$ at a given wavevector should equally decrease. On a general basis, the requirement that $`lim_{Q0}M_L(Q,z=0)/Q^2\eta _L/nm`$ now takes the form $$(c_{\mathrm{}}^2c_0^2)\left[\frac{(1\alpha (Q0))}{\gamma _1(Q0)}+\frac{\alpha (Q0)}{\gamma _2(Q0)}\right]\eta _L/nm$$ (22) The refined model (20) yields a dynamic structure factor given by $`S(Q,\omega )={\displaystyle \frac{S(Q)}{\pi }}`$ (23) $`\times Re\left\{{\displaystyle \frac{\omega _0^2(Q)}{i\omega +\frac{\mathrm{\Delta }_\mu ^2(Q)}{i\omega +\gamma _1(Q)}+\frac{\mathrm{\Delta }_\alpha ^2(Q)}{i\omega +\gamma _2(Q)}+\frac{\mathrm{\Delta }_{th}^2(Q)}{i\omega +a(Q)Q^2}}}\right\}^1.`$ (24) ## III THE EXPERIMENTS In this work we report the determination by IXS of the dynamic structure factor of liquid lithium in the wavevector range from $`1.4`$ to $`110`$ nm<sup>-1</sup> , corresponding to $`Q/Q_m510^2÷5.`$ The scanned energy range has been settled at each $`Q`$ in order to detect all the scattered signal up to the tails region, where it becomes comparable to the background. The experiment has been performed at the very high energy resolution beamline ID16-BL21 at the European Synchrotron Radiation Facility. A monochromatic beam of $`10^{9\text{ }}`$photons/s is obtained from a cryogenically cooled Si(111) double crystal followed by a high energy resolution monochromator operating in back-scattering geometry at selectable Bragg reflections . The scattered beam is collected by perfect spherical silicon crystal analyzers operating in back-scattering and Rowland circle geometry at the same reflection order of the monochromator. They were obtained by gluing $`12000`$ perfect crystals of $`0.6\times 0.6\times 2`$ mm<sup>3</sup> on a spherical blank . The overall energy resolution has been measured using a plexiglass scatterer at a $`Q`$ value corresponding to the maximum of its structure factor where the diffusion is dominated by the elastic contribution, and it turns out to have approximately a Lorenzian shape. Different configurations have been adopted depending on the explored $`Q`$-range. Below $`Q_m`$ ($`25`$ nm<sup>-1</sup>) we used simultaneously five-analyzers mounted on a horizontally rotating arm and with a fixed offset of $`1.5^0`$. Either the Si(9 9 9) or Si(11 11 11) reflections, corresponding to energy resolutions of $`3.0`$ meV and $`1.5`$ meV (FWHM) respectively, have been selected. At higher $`Q`$ values, where the spectra start to broaden, we used a vertical scattering geometry based on a single analyzer housed on a rotating arm that can reach more than 150<sup>0</sup> scattering angle. In this configuration we selected the Si (7 7 7) reflection with an overall energy resolution of $`8.5`$ meV FWHM. The $`Q`$ resolution was $`\pm 0.2`$ nm<sup>-1</sup> and $`\pm 0.35`$ nm<sup>-1</sup> in the horizontal and vertical geometries respectively. The momentum transfer is related to the scattering angle by the relation $`Q=2k_0sin(\theta _s/2)`$. Energy scans have been done by varying the temperature of the monochromator with respect to that of the analyzer crystals. The absolute energy calibration between successive scans is better than $`1`$ meV. Each scan took about 3 hours, and each $`Q`$-point spectrum has been obtained from the average of 2 to 8 scans depending on the Bragg-reflection order and on the $`Q`$-transfer. The data have been normalized to the intensity of the incident beam. In the $`QE`$ region of interest, empty vacuum chamber measurements gave either the flat electronic detector background of 0.6 counts/min or, in the angular region corresponding to $`9<Q<13`$ nm<sup>-1</sup>, a small elastic line due to spurious reflection of the beam entering the analyzers. These background signals have been subtracted from the data before any other data treatement. The chamber kapton windows (each 50 $`\mu `$m thick) gave no detectable contribution to the scattered flux. The liquid lithium uncapped container was made out of austenitic stainless steel in contact with a resistive heater, used to kept the liquid at constant temperature. We selected two different values: one slightly above the melting point, $`T=475`$ K, and the second at $`T=600`$ K; we have not been able to reach higher temperatures with our experimental setup due to the lack of windows. The $`20`$ mm long sample, kept together by surface tension, was maintained in a 10<sup>-6</sup> bar vacuum. The lithium has been loaded in an argon glove box. As far as experimental aspects are concerned, the energy transfer $`E`$ between the probe (X-rays) and the sample (excitations) has always been measured in meV. To make contact with the traditional notations adopted in computer simulations we will refer from now on only to the frequency $`\omega =2\pi \upsilon =E/\mathrm{}`$ expressed in ps<sup>-1</sup> (a numerical factor of $`0.659`$ ps<sup>-1</sup>/meV accounts for such a conversion). The IXS spectra of liquid lithium below $`Q_m`$ (horizontal geometry) are reported in Fig. 1 at the indicated $`Q`$ values. The low $`Q`$ data show the Brillouin triplet structure with the energy of the inelastic peaks increasing with $`Q`$ up to a $`Q`$ value of $`12`$ nm<sup>-1</sup>. This value is close to $`Q_m/2`$, as deduced from the $`S(Q)`$ reported in Fig. 2. The dispersion up to $`Q_m/2`$ can therefore be interpreted as that of the longitudinal acoustic branch in a pseudo first Brillouin zone (BZ). Furthermore, similarly to what is found in the second BZ of a crystal, we observe that, also in liquid lithium, the energy of the acoustic modes decreases with increasing $`Q`$ from $`Q_m/2`$ to $`Q_m`$. Increasing $`Q`$ above $`Q_m`$, i. e. in the ”third” or higher BZs, the spectrum increasingly broaden and distinct peaks are no longer observable. At the highest $`Q`$-values one finds that the dynamic structure factor becomes a symmetric peak centered at frequencies larger than $`\omega =0,`$ the behavior expected for a quantum free particle (in the classical limit no recoil energy is expected). Such evolution can be clearly followed in Fig. 3. Beside the observation of dispersion in a first and second pseudo-BZs, it is also important to note that the broadening of the excitations increases monotonically with $`Q`$, to the extent that near the end of the second pseudo-BZ a well defined inelastic peak is no longer visible. In order to extract quantitative information from the experimental intensity, i.e. to perform measurements of $`S_q(Q,\omega )`$ on an absolute scale, the most direct way would be to use a reference scatterer as customarily done in neutron experiments. In IXS such a procedure is extremely difficult because of the $`Q`$-dependence of the form factor and, in our case, for the different efficiencies of the analyzers. For these reasons we preferred to use an alternative indirect method based on the knowledge of the sum rules of $`S_q(Q,\omega )`$: in particular for the first two frequency moments we have $`\mathrm{\Omega }_S^{(0)}`$ $`=`$ $`{\displaystyle S_q(Q,\omega )𝑑\omega }=S(Q),`$ $`\mathrm{\Omega }_S^{(1)}`$ $`=`$ $`{\displaystyle \omega S_q(Q,\omega )𝑑\omega }=\mathrm{}Q^2/2m.`$ where the second equality follows from Eq. (11) applied for $`n=1`$. The measured raw intensity is related to the dynamic structure factor through $$I(Q,\omega )=A(Q)𝑑\omega ^{}S_q(Q,\omega ^{})R(\omega \omega ^{})$$ (25) where $`R(\omega )`$ is the experimental resolution function and $`A(Q)`$ is a factor taking into account the scattering geometries, the experimental setup and the lithium atomic form factor. The first moments of the experimental data, $`\mathrm{\Omega }_I^{(0)}`$ and $`\mathrm{\Omega }_I^{(1)}`$, and those of the resolution function, $`\mathrm{\Omega }_R^{(0)}`$ and $`\mathrm{\Omega }_R^{(1)}`$, are related to $`\mathrm{\Omega }_S^{(0)}`$ and $`\mathrm{\Omega }_S^{(1)}`$ by: $`\mathrm{\Omega }_I^{(0)}`$ $`=`$ $`A(Q)\mathrm{\Omega }_S^{(0)}\mathrm{\Omega }_R^{(0)},`$ $`\mathrm{\Omega }_I^{(1)}`$ $`=`$ $`A(Q)(\mathrm{\Omega }_S^{(0)}\mathrm{\Omega }_R^{(1)}+\mathrm{\Omega }_S^{(1)}\mathrm{\Omega }_R^{(0)}).`$ From the previous equation one derives that $$S_q(Q)=\frac{\mathrm{}Q^2}{2M}(\mathrm{\Omega }_I^{(1)}/\mathrm{\Omega }_I^{(0)}\mathrm{\Omega }_R^{(1)}/\mathrm{\Omega }_R^{(0)})^1.$$ (26) This procedure has been adopted to establish an absolute scale for $`S_q(Q,\omega )`$ using the experimentally determined $`I(Q,\omega )`$ and $`R(\omega )`$. Its reliability is shown in Fig. 2 where we obtain an excellent agreement between the $`S_q(Q)`$ values obtained by Eq. (26) and those derived by a MD simulation at $`T=475`$ K . The reported MD data are the outcame of a simulation made using the Price-Tosi interaction pseudopotential already tested against neutron diffraction data (see ref for further details). ## IV DATA ANALYSIS Lithium has been chosen because, among the simple monatomic liquids, is the one that is best suited to IXS. Indeed, its low mass gives recoil energies observable in the considered $`Q`$-range, and its low atomic number and large sound velocity give optimal signal within the available energy resolution, compensating for the large form factor decrease at high $`Q`$ values. The $`I(Q,\omega )`$ spectra, reported to their absolute scale exploiting the zeroth and first moments sum rules, show the transition from a triplet to a Gaussian. The maxima, $`\omega _m(Q)`$, of the longitudinal current spectra ($`\omega ^2S(Q,\omega )/Q^2`$) show an almost linear dispersion relation at low $`Q`$ \- typical of a sound wave - and a completely different dependence in the high $`Q`$ limit, where they approach the parabolic dispersion of the quantum free particle. Between these two regions, $`\omega _m(Q)`$ exhibits oscillations which are in phase with the structural correlations as observed in $`S(Q)`$. The low $`Q`$ region is of particular interest for its large number of informations. The so called positive dispersion, a fingerprint of the relaxation dynamics of disordered systems, either liquid or glassy , reflects the host of mechanisms driving such dynamics, their timescale and, at a careful inspection, their nature. We will start our analysis from here. ### A The ”low Q region” Looking at the spectra in Figs. 1,3, as also confirmed by the results reported in this section, the memory function approach can profitably be used approximately up to $`Q_m.`$ We adopted the following procedure. The classical result of Eq. (7) has been modified according to Eq. (12) to take into account of the detailed balance effects. The resulting $`S_q(Q,\omega )`$ has then been convoluted with the instrumental resolution according to (25). The $`\chi ^2`$ function arising from the difference between the outcoming function and the experimental data has finally been minimized by a standard Levenberg-Marquardt routine. Among the fitting parameters, $`S(Q)`$ and $`\omega _0^2(Q)`$ are independent of the specific model for the memory function, and their value is basically known: the first has been calculated from the first moment sum rule as discussed in Sec. III (we have already discussed the approximation $`S_q(Q)S(Q)`$), so that it has been kept fixed during the iterations. Then the second quantity $`\omega _0^2(Q)`$ is simply deduced from its definition (see Eq. (4)). It is worth to point out that $`\omega _0^2(Q)`$ cannot be taken as the second frequency moment of $`I(Q,\omega )`$ for two main reasons. The first one, related to the effects of the resolution, may be in principle overcome by a procedure similar to the one leading to the determination of $`S(Q),`$ but with the further difficulties arising from the almost-Lorenzian shape of the resolution that leads to a diverging $`\mathrm{\Omega }_R^{(2)}`$. Most severe is the non-invariance of the second moment under the transformation (12). For the sake of clarity we report in Fig. 4 the theoretical value $`\omega _0^2(Q)`$ compared to the second normalized moment of the experimental $`I(Q,\omega )`$ at $`T=475`$ K: the latter is always larger (this may be due also to the small but non-vanishing background in the experimental data) and the agreement gets worse at decreasing wavevectors, when the resolution effects become dominant. All the other fitting parameters obviously depend on the specific model for $`M(Q,t)`$. Let us firstly examine the spectral relevance in liquid Li of the coupling to thermal fluctuations, accounted for by the contribution $`M_{th}(Q,t)`$ to the total memory function. At not too high temperatures in the molten phase, all liquid alkali metals are characterized by a specific heat ratio $`\gamma 1,`$ the typical values giving $`\gamma 10.1`$. Outside the hydrodynamic ($`Q0`$) regime (and, in particular, in the wavevector range probed in this work) there are several indications that the difference $`\gamma (Q)1`$ is even smaller . Despite its modest strength in the present $`Q`$-range, the role of thermal relaxation in liquid metals is strikingly different from the one played in ordinary non-conducting liquids. A simple estimate of the decay rate embodied in the $`M_{th}(Q,t)`$ reported in Eq. (16) based on the experimental values of the thermal conductivity and of the specific heat shows that, in liquid lithium, $`\omega (\kappa /nC_V)Q^2`$ for most of the spectral range of interest here, and in particular for the range characterized by the presence of inelastic peaks (see Fig. 1). This indicates that in liquid Li (but a similar argument is expected to hold in general for molten metals) the afore-mentioned transition from the adiabatic to the isothermal response occurs at wavevectors well below those probed in the present experiments. Indeed, for wavevector-independent thermal parameters (a reasonable assumption at so low $`Q`$) we find for the threshold wavevector $`\omega _0(Q^{})\tau _{th}(Q^{})=1`$ a value of $`Q^{}0.2`$ nm<sup>-1</sup> at $`T=475`$ K, varying only of few percents at $`T=600`$ K. Of course, this does not mean that the presence of thermal fluctuations is not influent: as shown for example by Eq. (19), and as can be inferred by the simple arguments of Sec. II A, they may still give some contribution to the real part of the memory function, i.e. ultimately to Brillouin peaks widths. Such linewidth contribution gives $`\delta \omega _{th}=\mathrm{\Delta }_{th}^2\tau _{th}=\left(\gamma (Q)1\right)c_0(Q)/D_T(Q)0.056`$ ps<sup>-1</sup> ($`0.07`$ ps<sup>-1</sup>) for $`T=475`$ K ($`T=600`$ K) (in the $`Q0`$ limit) and is not explicitly dependent on $`Q`$. As already remarked, such contribution may be of some relevance only at the first two or three $`Q`$ values investigated here, but the strongly $`Q`$-dependent broadening of the inelastic components has to be ascribed to some other dynamical process. On the basis of the same argument, it is worth pointing out that the hydrodynamic value of the sound speed is not yet reached at the lower Q values of this work: as far as the positive dispersion is concerned, in the present case, the lower edge of the transition is the isothermal rather than the adiabatic value. As previously noted, a completely different scenario characterizes instead non-metallic fluids: due to the much lower thermal diffusivity, the opposite thermal regime is experienced all over the collective excitation region and the final low $`Q`$ result is an adiabatic response without any damping of the sound waves and with a narrow elastic line. Having ascertained the role of thermal effects, we now come to the more important implications of the fitting analysis for the dynamics of $`m_L(Q,t)`$. In the following we will refer to Fig. 5 where the specific situation of the spectrum at $`Q=7`$ nm<sup>-1</sup> has been chosen to enlighten the lineshape details associated with the different contributions to the memory function. As a zero-level approximation we consider the lineshape obtained by the simple hydrodynamic expression which only accounts for thermal decaying processes, embodying all the other relaxation channels as instantaneous: beside the use of Eqs. (14,16), this is simply done assuming that $`m_L(Q,t)\delta (t)`$. As in our case even the thermal relaxation is quite rapid on the X-rays timescale, the net result is a lineshape close to the one of the damped harmonic oscillator model. It is therefore not surprising (see Fig. 5) that the agreement with experimental data is very poor (no elastic peak is reproduced). Consequently, the assumption of an instantaneous decay of $`m_L(Q,t)`$ is untenable. Following section II B, as a next step, we generalize the memory function by the expression (19) i.e. the viscoelastic model. All the merits and the drawbacks of this model are again illustrated in Fig. 5. Leaving the time $`\tau (Q)`$ and the strength $`\mathrm{\Delta }_L^2(Q)`$ as fitting parameters, it is possible to obtain a rather good agreement for the position of the inelastic peaks, and consequently for the main features of the dispersion of sound-like excitations. However, one notes clear discrepancies in the quasi- elastic region at small $`\omega `$. The fitting procedure yields with comparable $`\chi ^2`$ values two different sets of parameters, corresponding to two different timescale driving the dynamics of the system. More precisely, in one case the broader part of the quasielastic and the Brillouin lineshape are well reproduced but the narrow quasi-elastic is missing, while in the other one the narrow elastic peak is well fitted but the Brillouin lines are too sharp. In other words, the viscoelastic model cannot account for the double-sloped shape of the quasielastic region. Although in Fig. 5 we refer to $`Q=7`$ nm$`^1,`$ the situation at most wavevectors is similar. The inescapable conclusion is that there are at least two different timescale, each giving a contribution to the width of the ”quasi-elastic” region, and accounting at the same time for the inelastic lineshape. On these grounds, the more flexible phenomenological model (20), based on two exponential decays, appears more promising. In this case the best fitted lineshapes (quantum corrected and resolution convoluted), including the thermal relaxation (using the $`Q0`$ values of specific heat ratio and thermal diffusivity from ) through Eq. (23), are again reported in Figs. 1 and 5, and a marked improvement of the agreement is clearly visible. In Fig. 6 we also report the classical and resolution deconvoluted lineshapes of $`S(Q,\omega )`$ and $`C_L(Q,\omega )`$ as builded from the fitting parameters, in the region $`Q<14`$ nm<sup>-1</sup>, the most significant as far as the positive dispersion task is concerned. The maxima of the latter function have been utilized to determine the apparent sound velocity of the system. Let us now discuss the different parameters entering Eq. (20), starting from the relaxation times: in Fig. 7 we report the wavevector dependence of such quantities at both temperatures. The presence of the double timescale is confirmed by the difference by about one order of magnitude between the two parameters $`\tau _\alpha `$ and $`\tau _\mu .`$ Concerning $`\tau _\alpha `$ it is worth to point out the non negligible effect of resolution: the value of $`1/HWHM`$ is comparable to the relaxation time itself so that it is expected to affect this fitting parameter, particularly at low $`Q`$. As a consequence it is unfortunately not possible to inspect any temperature dependence of $`\tau _\alpha (Q,T).`$ Completely different is the situation as far as the fast timescale is concerned. The determination of $`\tau _\mu `$ is expected to be definitely more reliable and on the basis of our analysis the fast dynamics turns out to be temperature independent at the investigated $`Q`$-values. The influence of these relaxation mechanisms on the dynamics can be inferred looking at the $`\omega (Q)\tau (Q)`$ values as done before for the thermal channel: as far as the slow mechanism is concerned the system response turns out to be ”solid like” across the full first Brillouin pseudo-zone (see the insets in Fig. 7a). The fast relaxation accounts instead for a moderately viscous response: $`\omega (Q)\tau (Q)1`$ around the peak value at $`Q8`$ nm<sup>-1</sup>and remains quite lower all around (insets in Fig. 7b). Following the arguments of Sec. II B, we expect the slow relaxation to be responsible for the narrow quasi elastic contribution and for the position of the inelastic features of $`S(Q,\omega )`$ while the fast process should drive mainly the acoustic damping (Brillouin linewidth) also accounting for the broader part of the Mountain peak . To go further and point out the real quantitative influence of such two channels is necessary to look at the strength parameters, or better at the rate $`\alpha (Q).`$ Such quantity, defined as the ratio $`\mathrm{\Delta }_\alpha ^2/\left(\mathrm{\Delta }_\mu ^2+\mathrm{\Delta }_\alpha ^2\right),`$ has its maximum of $`0.20`$ at low $`Q`$ and decreases at increasing the wavevector as shown in Fig. 8. This means that the fastest process plays the major role in characterizing the dynamics of liquid lithium all over the investigated first Brillouin pseudo-zone. Such statement can be better understood looking at the apparent sound speed behavior as plotted in Fig. 9. The sound velocity moves from the low $`Q`$ value (isothermal in the IXS window) to its high frequency value $`c_{\mathrm{}}(Q).`$ As previously noted, such behavior has been observed in several MD simulations of alkali metals and in some cases experimentally, even if with some ambiguity due to kinematic restrictions/incoherent contribution on neutron side and resolution broadening in the pioneer IXS experiments . What is less clear is the microscopic origin of such positive dispersion. On the basis of our analysis, once unambiguously found out an experimental evidence of a double timescale, we are able to ascribe the positive dispersion of the sound velocity to the faster process. We reported in fact in Fig. 9 the value of the solid-like response associated to the single slow process and to both slow and fast relaxation processes. As expected the ”jump” in sound velocity associated to the slow $`\alpha `$-process ($`\left[\sqrt{\omega _0^2(Q)+\mathrm{\Delta }_\alpha ^2(Q)}\omega _0(Q)\right]/Q`$) has already occurred at the lower IXS $`Q`$ values, but it is quantitatively negligible as compared to the full positive dispersion of the system. Despite of the fact that the condition $`\omega _l(Q)\tau _\mu (Q)1`$ is never satisfied, the faster process, due to its dominant strength, accounts almost entirely for the viscous-elastic transition. As can be seen in Fig. 9, the apparent sound velocity $`c_{app}=\omega _l(Q)/Q`$, never reaches the infinite frequency sound velocity deduced by the fit, $`c_{\mathrm{}}(Q)=\sqrt{\omega _0^2(Q)+\mathrm{\Delta }_\alpha ^2(Q)+\mathrm{\Delta }_\mu ^2(Q)}/Q.`$ At first sight this discrepancy may be explained on the fact that $`\omega _l(Q)\tau _\mu (Q)`$ is always $`1`$. However $`c_{\mathrm{}}(Q)`$ can be also evaluated theoretically starting from the knowledge of the interatomic potential and the pair distribution function (see for example ref. ). In Fig. 9 we also report the values of the unrelaxed sound speed $`c_{\mathrm{}}^{th}(Q)`$ computed in this way, they turns out to be lower than those determined from the fitting of the spectra at both the temperatures. Moreover, $`c_{\mathrm{}}^{th}(Q)`$ is actually reached by the apparent sound speed at $`Q9`$ nm$`^1,`$ i.e. when $`\omega _l(Q)\tau _\mu (Q)1.`$ A possible explanation of the inconsistency between $`c_{\mathrm{}}^{th}(Q)`$ and $`c_{\mathrm{}}(Q)=\sqrt{\omega _0^2(Q)+\mathrm{\Delta }_\alpha ^2(Q)+\mathrm{\Delta }_\mu ^2(Q)}/Q`$ can be related to the arbitrary choice of the memory function details. Even though the presence of a two timescale dynamics is directly seen in the raw experimental data (particularly in the $`7<Q<14`$ nm<sup>-1</sup> region, the quasi elastic peak reflects the superposition of two peaks of different widths), the detailed shape of $`M_L(Q,t)`$ can have in principle more complicated features than the double exponential ansatz of Eq. (20). In particular we already mentioned the major drawback of such an assumption, namely the cusp at $`t=0`$. It is reasonable to think that an exponential decay forced to represent a more complicated time dependence can give an accurate estimate as far as the $`\tau `$ is concerned, while at short times the lack of a zero second derivative inescapably leads to an overestimate of $`M_L(Q,t=0),`$ i.e. just the positive dispersion amplitude: a similar effect, i.e. the overestimation of the $`c_{\mathrm{}}(Q0)`$ deduced by the fit of BLS data using instantaneous approximation for the microscopic process with respect to the experimental IXS current spectra maxima, has been recently shown in polybutadiene . The last important aspect concerns the relative weights of the two processes and how to relate them to the acoustic attenuation. The role of the thermal effects has been already discussed. Looking at the generalized viscosity contributions $`\mathrm{\Delta }_\mu ^2\tau _\mu `$ and $`\mathrm{\Delta }_\alpha ^2\tau _\alpha `$ (see Fig. 10) we find that they are nearly comparable, (the strength differences are nearly balanced by the different $`\tau `$), but once again only the microscopic (faster) relaxation process can be related to the Brillouin linewidth being the only one close to the $`\omega \tau <<1`$ condition in the explored $`Q`$-range. Moreover, in the spirit of the hydrodynamic generalization, the sum of such contributions is expected to have a $`Q^2`$ dependence. Such prediction is quite well reproduced by our fits: only at very low $`Q`$ a slight deviation occurs, ascribable to the slow part; once again the reason may be an excess estimate of the $`\alpha `$relaxation time at small $`Q`$. Finally, in Fig. 11, the longitudinal viscosity as outcome of the fit (the full area under $`M(Q,t)`$) is compared to an experimental determination of shear static ($`Q0`$) viscosity . Noting that in molten alkali metals the shear part usually accounts for a $`3040\%`$ of the longitudinal part, the agreement is satisfactory, except for the very low $`Q`$ points where finite resolution function effects prevent an accurate check. ### B The ”high Q region” Above $`Q_m`$, collective excitations are still evident in $`I(Q,\omega )`$ for nearly one more Brillouin pseudo-zone, even if the analysis of the previous section is not wholly satisfactory because of the worsening of the experimental resolution. Moreover, single particle features gradually emerge giving rise to a crossover between two completely different dynamical regimes. As the density fluctuation wavelength matches the mean interparticle distance, a transition between strongly correlated to incoherent motion occurs. The lineshape associated to a quantum free particle is well known to be Gaussian. Once that $`S_q(Q)`$ is determined as before from the first two moments sum rules, there are no unknown parameters. While in the previous section our sample has been supposed to be single-component, we must now, due to the emergency of additional quantum effect, be more accurate by taking into account isotopic effects. From the natural abundance ratios (92% of <sup>7</sup>Li, 8% of <sup>6</sup>Li) one finds that $$S_q(Q,\omega )=\frac{1}{\sqrt{2\pi }}\underset{i=6,7}{}\frac{C_i}{\sigma _i}e^{(\omega \omega _i)^2/2\sigma _i^2}$$ (27) where $`C_i`$ are the above mentioned isotope concentrations while $`\omega _i=`$ $`\mathrm{}Q^2/2m_i`$ and $`\sigma _i^2=K_BTQ^2/m_i`$ are the first (central) and the second (relative to $`\omega _i`$) frequency moments of the two isotopes. In Fig. 3 the experimental spectra are reported together with the theoretical lineshapes (resolution convoluted) of Eq. (27) in order to emphasize all the crossover features. To be more quantitative we report in Fig. 12 the Stokes current spectra maxima $`\omega _l`$ as function of the exchanged wavevector. Indeed, in this “high $`Q`$” region (above $`Q_m`$=24 nm<sup>-1</sup>) this quantity has been calculated starting from the spectra of Fig. 3 (we expect the effect of the resolution broadening to be negligible at such high $`Q^{}s`$). For this reason, due to the quantum features of the $`S(Q,\omega )`$, the quantity $`\omega _l`$ takes different values in the Stokes/Antistokes sides of the spectra . The asymptotic values are expected to be reached in the very low $`Q`$ (adiabatic) and high $`Q`$ (free particle) . After an initial nearly linear dispersion, structural effects take place suppressing the sound propagation around $`Q_m/2`$ due to strong negative interference. With increasing $`Q`$ values, the points in Fig. 12 show not only a second pseudo-BZ, but also a series of oscillations that damp out with increasing $`Q`$ \- here, $`\omega _l(Q)`$ is approaching the single particle behavior. These oscillations are in anti-phase with those of $`S(Q)`$ (see Fig. 2) and can therefore be associated with the local order in the liquid. ## V CONCLUSION In the present work inelastic X-ray scattering has been utilized to study in detail the main features of the collective dynamics in liquid lithium, a system where the use of inelastic neutron scattering is quite difficult and basically impossible in the low Q region. This IXS experiments have allowed for a precise assessment of the different relaxation processes which affect the spectral shape of the dynamic structure factor. Specifically, the quality of the data required an analysis where one has to invoke the simultaneous presence of three distinct decay channels of the collective memory function. The first one is associated with the coupling between density and temperature fluctuations. As in other alkali metals, although its relevance is rather small, this process cannot be neglected and plays a different role in respect to ordinary non-conducting liquids. Much more important is the unambiguous evidence of two well separated timescales in the decay of the $`M_L(Q,t)`$ \- the portion of the memory function directly associated with generalized viscosity effects. In this respect, our IXS data prove that the traditional use of a single time-scale (viscoelastic model) cannot account for the detailed shape of $`S(Q,\omega )`$, and that it only provides a guidance for a qualitative description of different dynamical regimes. Despite the effect of an additional time-scale to the viscoelastic model has been tested on MD results in the case of Rubidium (see pag. 248 of ref. ), and theoretical interpretations of such framework (in particular devoted to frame the slow process within Mode Coupling theory) have been recently given we give in this work, to the best of our knowledge, the first experimental evidence of the necessity of such an assumption to correctly reproduce the measured $`S(Q,\omega )`$, stressing the role of the contributions from the different relaxations to the $`S(Q,\omega )`$ in the considered $`QE`$ regions. The presence of a two time-scales mechanism leaves open, however, a point of crucial importance: namely, their physical origin and, most important, of the fast one. The slow process, referring to the terminology used to describe glass forming systems, is related to the $`\alpha `$-relaxation responsible for the liquid-glass transition in systems capable to sustain strong supercooling. The same process can be framed within kinetic theory in terms of ”correlated collisions”: in mode coupling approaches, the onset of these correlation effects is traced back to the coupling to slowly relaxing dynamical variables, and specifically in the liquid region, to long lasting density fluctuations. The corresponding decay mechanism (often referred to as ”structural relaxation”), almost negligible in the initial collisional region, controls the dynamics at intermediate and long times. For $`Q0`$ the amplitude of the slow portion of the memory function turns out to vanish as $`Q^2`$, whereas at increasing wavevectors it becomes proportional to $`Q`$ . At sufficiently large $`Q`$, the relevance of these couplings as long lasting decay channel decreases markedly. In fact, in a rather wide portion of the explored range of wavevectors, we find that the amplitude $`\mathrm{\Delta }_L^2(Q)`$ of the total memory function behaves approximately as $`Q^2.`$ Therefore, outside the hydrodynamic region, the normalized weight of the slowly varying contributions to $`M_L(Q,t)`$ approximately decreases proportionally to $`Q/Q^2=1/Q.`$ This weight may reasonably be identified with the dimensionless amplitude factor $`\alpha (Q)`$ in the model of Eq. (20). A quantitative understanding of the ”slow” timescale would require a full evaluation of the mode-coupling contribution at different wavevectors, as in fact done in the above mentioned MD study . The situation is much more involved as far as the fast process is concerned and, in fact, its microscopic interpretation is still matter of debate. In the present work we have given an experimental evidence that, in liquid Li, the fast process is indeed the dominant one, and it mainly controls both the sound speed dispersion and attenuation i.e. the position and the width of the Brillouin component of the spectra. This observation justifies the need for a detailed discussion on the nature of the fast relaxation channel. Within generalized kinetic theory the fast initial decay of both single particle and collective memory function is traced back to collisional events, which are short ranged both in space and in time. In particular, in the collective case, the short timescale $`1/\gamma _1(Q)`$ can be associated with the duration of a rapid structural rearrangement occurring over a spatial range $`2\pi /Q_m.`$ In most liquids, at low and intermediate wavevectors, the typical value of the time $`1/\gamma _1(Q)`$ is of the order of a few $`10^{14}`$ s, with some structural oscillations of decreasing trend . At $`Q>Q_m`$, the structural effects damp out, the decrease becomes more marked, and, when the single-particle aspects start to prevail, $`1/\gamma _1(Q)Q^1.`$ An inspection of the results reported in a recent MD study in liquid lithium near melting basically confirms these qualitative features as well as the actual values of the ”short time” found in our data analysis (see Fig. 7). Although the description of the fast process in term of interparticle collision is a possible way to account for the dynamical features at short times, it does not allow a deep insight into the physics behind it. Although the description of the fast process in terms of local collisional events has been widely used in the past to account for the short-time dynamics, other mechanisms cannot be excluded. We have already mentioned in sec. II that a formally similar ”microscopic mechanism” has been introduced in the analysis of the Brillouin spectra of glassformers ; more recently, an analogous microscopic relaxation process has been detected in a simulated harmonic glass . Generally speaking, the inclusion of a relaxation process in a memory function reflects the existence of decay channels by which the energy stored in a given degree of freedom (”mode”) relaxes toward other modes. The problem is then the physical identification of the modes entering the fast relaxation process, and the ultimate origin of the latter in a widely different class of systems ranging from fluids to glasses. In generalized kinetic theories the mechanism is traced back to a mainfold of non-hydrodynamic, phase-space ”kinetic” modes, whose non-conserved character ensures a rapid decay of the memory function . A different approach to describe the fast part of the memory function in disordered systems relies on the normal modes (”instantaneous” in normal liquids) analysis of the atomic dynamics . In this case one uses a framework (dynamical matrix, etc.) formally similar to the one customarily adopted for harmonic crystals; however, owing to the lack of translational symmetry of the system, it turns out that the eigenstates cannot anymore be represented by planewaves (PW) even at relatively small wavevectors . As a result, when probing the system at a specific wavevector $`Q`$ one effectively detects a transfer from a PW mode toward other PWs of different wavevectors. Assuming that it is actually this energy flow which causes the fast relaxation, we deal with a mechanism whose ultimate origin is the topological disorder, and not a truly dynamical event. Ordinary liquids (such as the one considered here) are certainly ”disordered systems” and at the high frequency considered here ($`\omega \tau _\alpha >>1`$), they can be considered as ”frozen”; thereforethe description used for glasses can be applied as well. At the present stage of the theory our experimental data cannot support either interpretation, or not even ascertain whether the distinction is largely semanthic, or not. However, the posssibility that the phenomenology reported here for a simple liquid can be understood within the same mental framework developed for more complex liquids and glasses, is certainly interesting. Further studies on other simple fluids, as for example neon , seem to suggest that is indeed the case. ## VI ACKNOWLEDGMENTS We gratefully aknowledge the ID-16 (ESRF) beamline staff for the assistence during the experiments preparation and performance. One of us (T. Scopigno) feels to be indebted with Dr. R. Di Leonardo for many fruitful discussions. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al. “Density Fluctuations in Molten Lithium…” by T.Scopigno et al.
warning/0001/gr-qc0001013.html
ar5iv
text
# Transition from inspiral to plunge in binary black hole coalescences ## I Introduction The most promising candidate sources for ground based interferometric gravitational-wave (GW) detectors such as LIGO and VIRGO are binary systems made of massive (stellar) black holes . Such binary black holes (with individual masses in the range, say, $`315M_{}`$) pose special problems , . Let us recall that gravitational radiation damping is efficient at circularizing such binary systems, and then drives, for a long time, a slow inspiraling quasi-circular motion of the binary system. This quasi-circular “adiabatic inspiral phase” is expected to terminate abruptly, and to change to some type of “plunge phase” (leading to final coalescence) when the binary orbit shrinks down to the Last Stable (circular) Orbit (LSO) defined by the conservative part of the nonlinear relativistic force law between two bodies. \[In the test-mass limit, the full nonlinear relativistic force law corresponds to geodesic motion in a Schwarzschild spacetime, and exhibits, as is well known, an LSO located at $`R=6GM`$. One expects that a comparable-mass system will still exhibit such an LSO; see below.\] Now, the signal to noise ratio (in an initial LIGO detector) for inspiral signals from comparable-mass black hole binaries reaches a maximum for $`M28M_{}`$, which corresponds to a GW frequency for the waves emitted at the Last Stable (circular) Orbit (LSO) equal to $`f_{\mathrm{GW}}^{\mathrm{LSO}}170\mathrm{H}\mathrm{z}`$, a value which is (not accidentally) very close to the location $`f_{\mathrm{det}}167\mathrm{H}\mathrm{z}`$ (for initial LIGO) of the minimum of the characteristic detector noise amplitude $`h_n(f)\sqrt{fS_n(f)}`$ (see Fig. 1 of ). Therefore the first detections will probably concern massive systems with $`M30M_{}`$. Moreover, Ref. has shown that when the total mass $`Mm_1+m_2`$ lies in the range $`540M_{}`$ the proximity (within a factor $`2`$) between the observationally most important frequencies<sup>*</sup><sup>*</sup>* We neglect here the very small difference between the optimal frequency $`f_{\mathrm{det}}`$ for generic broad-band bursts, and the optimal frequency $`f_p`$ for inspiral signals (see ). $`f_{\mathrm{det}}`$ and the GW frequency at the LSO, $`f_{\mathrm{GW}}^{\mathrm{LSO}}`$, was calling both for an especially careful treatment of the Fourier transform of the emitted waveform, and for an improved knowledge of the transition between the inspiral phase and the plunge phase. The present paper will attempt to improve our knowledge of the transition between inspiral and plunge by combining two, recently proposed, non-perturbative techniques: Refs. and . Let us first recall that, a few years ago, Will and collaborators , tried to attack the problem of the late-time evolution of compact binaries (including the transition from inspiral to plunge) by a direct use of the Damour-Deruelle equations of motion. These equations of motion are given in the form of a perturbative expansion in powers of a small parameter $`\epsilon =v/c`$ (“post-Newtonian”, or, in short, PN, expansion). In Ref. a direct integration of these perturbative equations of motion (using the method of osculating elements) was used, while, in Ref. it was proposed to improve the straightforward perturbative approach by using “hybrid” equations of motion. The “hybrid” approach is a partial re-summation approach in which the perturbative terms in the equations of motion which survive in the test mass limit ($`\nu m_1m_2/(m_1+m_2)^20`$) are replaced by the known, exact “Schwarzschild terms”, while the $`\nu `$-dependent terms are left as a perturbative expansion. Both the robustness , and the consistency of the hybrid approach of have been questioned. \[In particular, it was pointed out in Ref. that, in this approach, the supposedly small “$`\nu `$-corrections” represent, in several cases, a very large (larger than 100%) modification of the corresponding $`\nu `$-independent terms.\] Another sign of the unreliability of the hybrid approach is the fact that the recent study of the location of the LSO at the third post-Newtonian (3PN) accuracy has qualitatively confirmed the 2PN-level results of the non-perturbative techniques to be discussed below (namely that the LSO is “lower than 6GM”), thereby casting doubt on the most striking prediction of the hybrid approach (an LSO “higher than 6GM”, i.e. with a lower orbital frequency). By contrast with the perturbative approach of and the partially re-summed approach of , the present paper will rely on the systematic use of non-perturbative re-summation techniques. The basic philosophy underlying our approach is the following. We are interested in understanding, in quantitative detail, the combined influence on the inspiral $``$ plunge transition of radiation reaction and of non-linear effects in the force law for comparable-mass binary systems (i.e. for systems for which $`\nu m_1m_2/(m_1+m_2)^2`$ is around Note that because $`\nu `$, considered as a function of the ratio $`m_1/m_2`$, reaches its maximum $`\nu _{\mathrm{max}}=1/4`$ for $`m_1/m_2=1`$ it stays numerically near $`1/4`$ even for mass ratios quite different from 1. E.g., even for $`m_1/m_2=3`$, $`4\nu =0.75`$. $`1/4`$). At present there exists no method for deriving, from first principles, non-perturbative expressions for the two-body equations of motion, especially in the case of interest where $`4\nu `$ is not a small parameter. As a substitute we shall combine two different re-summation techniques that have been recently introduced to deal with two separate aspects of the problem we wish to tackle. The first re-summation technique, introduced in , allows one to get a non-perturbative, $`\nu `$-dependent, estimate of the rate of loss of angular momentum (under gravitational damping) in quasi-circular, comparable-mass binaries. The idea of is three-pronged: (i) to work with an invariant function of an invariant argument, $`F(v)`$, (ii) to inject some plausible information about the meromorphic structure of this function, and, finally, (iii) to use Padé approximants to estimate $`F(v)`$ from the first few known terms in the perturbative (PN) expansion of $`F(v)`$. The second re-summation technique, introduced in , allows one to derive a non-perturbative, $`\nu `$-dependent, estimate of the (conservative part of the) nonlinear force law determining the motion of comparable binaries. The idea of is to map the real two-body problem on a simpler effective one-body problem, i.e. on the problem of the motion of a particle of mass $`\mu m_1m_2/(m_1+m_2)`$ in some “effective” background metric $`g_{\mu \nu }^{\mathrm{eff}}(x^\lambda )`$. The possibility (and uniqueness, given some natural requirement) of such a mapping, real $``$ effective, was proven at the 2PN level in . The extension of this approach at the 3PN level has been recently discussed on the basis of the 3PN dynamics recently derived by Jaranowski and Schäfer . At the 2PN level the $`\nu `$-dependent terms in the effective metric were found to be numerically so small (around the LSO) that the need for a further (Padé-type) re-summing of the effective metric coefficients did not arise. \[However, note that Ref. has introduced, at 2PN and 3PN, the further idea of a specific, Padé-improvement of $`g_{\mu \nu }^{\mathrm{eff}}(x^\lambda )`$.\] In this paper we shall show how one can combine the methods of and to derive a full force law (including radiation reaction) describing the quasi-circular motion of comparable-mass binaries. Our approach is intended to apply to any value of $`\nu `$, but is restricted to considering quasi-circular motions, where the radial velocity $`\dot{R}`$ is much smaller than the circular one $`R\dot{\phi }`$. As we shall see, we shall consistently check that the condition $`\dot{R}R\dot{\phi }`$ holds true not only during the adiabatic inspiral, but also during the transition to the plunge, and even during most of the plunge. We apply our method, in this paper, to deriving two sorts of results which are of direct interest to the ongoing effort to detect gravitational waves. First, we shall give initial dynamical data (i.e. initial positions and momenta) for binary black holes that have just started their plunge motion. The idea here is that numerical relativity will probably not be able, before quite a few years, to accurately evolve binary systems over many (or even $`10`$) orbits. This is why we propose a method for computing accurate initial dynamical data at a moment so late in the evolution that there remains (when $`4\nu 1`$) less than one orbit to evolve. \[In the equal-mass case, $`\nu =1/4`$, we shall compute data $`0.6`$ orbit before “coalescence”.\] Our contention (whose robustness we shall try to establish) is that suitably re-summed versions of analytic (PN) results allow one to push the evolution that far. \[We shall use here 2.5PN-accurate information for angular momentum loss and 2PN-accurate information for the conservative force law. However, as shown in and our method can be pushed to higher accuracy when the correspondingly needed PN results become unambiguously known.\] Note that this attitude is opposite to the one taken in in which it was assumed that “there is little hope, via PN Padé approximants, to evolve” a binary system up to the moment where it can provide initial data for the final coalescence. Let us, however, immediately add that the present paper is still incomplete, in that we give only dynamical data $`(𝒒_1,𝒒_2,𝒑_1,𝒑_2)`$ but we do not solve the remaining problem of constructing the initial gravitational data $`(g_{ij}(x),K_{ij}(x))`$ determined (in principle) by $`(𝒒_a,𝒑_a)`$ (given, say, some no-incoming-radiation condition). We shall leave this (important) issue to future work. The second aim of this work is to provide, for data analysis purposes, some estimate of the complete waveform emitted by the coalescence of two black holes (with negligible spins). We do not claim that this part of the work will be as accurate as the first one. The idea here is to provide a (hopefully $``$ 10% accurate) guess of the complete waveform, with its transition from an inspiral phase to a plunge one, followed by a coalescence ending in a stationary final state. In view of the recent realization of the crucial importance of the details of the transition to the plunge for the construction of faithful GW templates (for massive binaries with $`5M_{}<M<\mathrm{\hspace{0.17em}40}M_{}`$) even an approximate knowledge of the complete waveform will be a valuable information for data analysis (e.g. to test the accuracy of present templates, and/or to propose more accurate or, at least, more robust, templates). While preparing this work for publication, we learned of the existence of an independent work of Ori and Thorne which deals with the transition between the inspiral and the plunge in the test mass limit ( $`\nu 0`$). ## II Conservative part of the two-body force law In this section, we recall the non-perturbative construction of the (conservative) two-body force law given in Ref. . There it was shown that the conservative part (i.e. without radiation damping) of the dynamics of a binary system, represented in ADM phase-space coordinates $`(𝒒_1^{\mathrm{ADM}},𝒒_2^{\mathrm{ADM}},𝒑_1^{\mathrm{ADM}},𝒑_2^{\mathrm{ADM}})`$, could be mapped (at the 2PN level), via the combination of an energy map, $`_{\mathrm{eff}}=f(_{\mathrm{real}})`$, and a canonical transformation, $`(𝒒_a^{\mathrm{ADM}},𝒑_a^{\mathrm{ADM}})(𝒒_a,𝒑_a)`$, $`a=1,2`$, into the simpler dynamics of the geodesic motion of a particle of mass $`\mu =m_1m_2/(m_1+m_2)`$ in some effective background geometry $`g_{\mu \nu }^{\mathrm{eff}}(x)`$: $$ds_{\mathrm{eff}}^2=g_{\mu \nu }^{\mathrm{eff}}(x^\lambda )dx^\mu dx^\nu =A(R)c^2dt^2+B(R)dR^2+C(R)R^2(d\theta ^2+\mathrm{sin}^2\theta d\phi ^2).$$ (1) \[See for the generalization of this approach to the 3PN level.\] Here the coordinates $`(R,\theta ,\phi )`$ are polar coordinates in the effective problem (describing the relative motion). They are related in the standard way ($`Q^x=R\mathrm{sin}\theta \mathrm{cos}\phi `$, $`Q^y=R\mathrm{sin}\theta \mathrm{sin}\phi `$, $`Q^z=R\mathrm{cos}\theta `$) to the (relative) effective Cartesian coordinates $`𝑸=𝒒_1𝒒_2`$, where $`𝒒_1`$ and $`𝒒_2`$ are the effective coordinates of each body. One works in the center-of-mass frame of the binary system, i.e. $`𝒑_1+𝒑_2=0=𝒑_1^{\mathrm{ADM}}+𝒑_2^{\mathrm{ADM}}`$. The canonical conjugate of the relative position $`𝑸`$ is the relative momentum $`𝑷=𝒑_1=𝒑_2`$. In most of this paper we shall work with the effective phase-space coordinates $`(𝑸,𝑷)`$ \[or rather with scaled versions of their polar Note that we have the usual relations, such as, $`P_R=n^iP_i`$ with $`n^i=Q^i/R`$, and $`P_\phi =Q^xP_yQ^yP_x`$. counterparts $`(R,\theta ,\phi ;P_R,P_\theta ,P_\phi )`$\]. We shall only discuss at the end how to construct the more physically relevant ADM phase space coordinates $`(𝒒_a^{\mathrm{ADM}},𝒑_a^{\mathrm{ADM}})`$ from $`(𝑸,𝑷)`$. In absence of damping (to be added later), the evolution (with respect to the real ADM time coordinate $`t_{\mathrm{real}}`$) of $`(𝑸,𝑷)`$ is given by Hamilton’s equations $`{\displaystyle \frac{dQ^i}{dt_{\mathrm{real}}}}{\displaystyle \frac{H_{\mathrm{real}}^{\mathrm{improved}}(𝑸,𝑷)}{P_i}}=0,`$ (2) $`{\displaystyle \frac{dP_i}{dt_{\mathrm{real}}}}+{\displaystyle \frac{H_{\mathrm{real}}^{\mathrm{improved}}(𝑸,𝑷)}{Q^i}}=0,`$ (3) where the real (i.e. giving the $`t_{\mathrm{real}}`$-evolution, and the real two-body energy) improved (i.e. representing a non-perturbative re-summed estimate of the real PN Hamiltonian) Hamiltonian reads $$H_{\mathrm{real}}^{\mathrm{improved}}(𝑸,𝑷)=Mc^2\sqrt{1+2\nu \left(\frac{H_{\mathrm{eff}}(𝑸,𝑷)\mu c^2}{\mu c^2}\right)},$$ (4) and $$H_{\mathrm{eff}}(𝑸,𝑷)=\mu c^2\sqrt{A(Q)\left[1+\frac{(𝒏𝑷)^2}{\mu ^2c^2B(Q)}+\frac{(𝒏\times 𝑷)^2}{\mu ^2c^2C(Q)}\right]}.$$ (5) Here $`Q\sqrt{\delta _{ij}Q^iQ^j}=R`$, $`n^i=Q^i/Q`$ is the unit vector in the radial direction, and the scalar and vector products are performed as in Euclidean space. Henceforth, we shall pose $`tt_{\mathrm{real}}`$, $`HH_{\mathrm{real}}^{\mathrm{improved}}`$ and use the following notation: $$Mm_1+m_2,\mu \frac{m_1m_2}{M},\nu \frac{\mu }{M}\frac{m_1m_2}{(m_1+m_2)^2}.$$ (6) In polar coordinates, restricting ourselves to planar motion in the equatorial plane $`\theta =\pi /2`$ and to the Schwarzschild gauge ($`C(Q)=1`$), we get the equations of motion $`{\displaystyle \frac{dR}{dt}}={\displaystyle \frac{H}{P_R}}(R,P_R,P_\phi ),`$ (7) $`{\displaystyle \frac{d\phi }{dt}}={\displaystyle \frac{H}{P_\phi }}(R,P_R,P_\phi ),`$ (10) $`{\displaystyle \frac{dP_R}{dt}}+{\displaystyle \frac{H}{R}}(R,P_R,P_\phi )=0,`$ $`{\displaystyle \frac{dP_\phi }{dt}}=0,`$ with $$H(R,P_R,P_\phi )=Mc^2\sqrt{1+2\nu \left[\sqrt{A(R)\left(1+\frac{P_R^2}{\mu ^2c^2B(R)}+\frac{P_\phi ^2}{\mu ^2c^2R^2}\right)}1\right]}.$$ (11) Like in any (non-degenerate) Hamiltonian system, this conservative dynamics is equivalent to a Lagrangian dynamics $$L_{\mathrm{real}}^{\mathrm{improved}}(𝑸,\dot{𝑸})=P_i\dot{Q}^iH_{\mathrm{real}}^{\mathrm{improved}}(𝑸,𝑷),$$ (12) with $`P_i(\dot{Q})`$ obtained by solving $`\dot{Q}^i=H/P_i`$. The Lagrangian equations of motion read: $$\frac{d}{dt}\frac{L_{\mathrm{real}}^{\mathrm{improved}}}{\dot{Q}^i}\frac{L_{\mathrm{real}}^{\mathrm{improved}}}{Q^i}=0.$$ (13) To ease the notation we denote $`LL_{\mathrm{real}}^{\mathrm{improved}}`$. Finally, the 2PN-accurate metric coefficients $`A(R)`$, $`B(R)`$, Eq. (1), (in the Schwarzschild gauge where $`C(R)1`$) read $$A(R)=1\frac{2GM}{c^2R}+2\nu \left(\frac{GM}{c^2R}\right)^3,$$ (14) $$B(R)D(R)/A(R),$$ (15) with $$D(R)=16\nu \left(\frac{GM}{c^2R}\right)^2.$$ (16) Note that it was recently suggested (because of the slow convergence of the 3PN contributions) to replace the straightforward expression (14) by a suitably Padeed version, namely (at 2PN): $`A_{P_2}(R)=12u(1+\nu u^2)^1`$, where $`uGM/c^2R`$. However, we have checked that this refinement has only a very minor effect on the results to be discussed below. The re-summed (conservative) dynamics defined by the Hamiltonian (11) contains a Last Stable (circular) Orbit (LSO) which is a $`\nu `$-deformed version of the well known Schwarzschild LSO. Let us recall that the radius of the LSO is obtained by imposing the existence of an inflection point in the effective potential $`H(R,P_R=0,𝒥)`$ for the radial motion, $$\frac{H}{R}(R,P_R=0,𝒥)=0=\frac{^2H}{R^2}(R,P_R=0,𝒥),$$ (17) where the total angular momentum $`𝒥P_\phi `$ stays fixed. Eq. (17) has a solution in $`R`$ (for each value of $`\nu `$) only for some specific value of $`𝒥=𝒥^{\mathrm{LSO}}(\nu )`$. In terms of the rescaled variables $`rc^2R/GM`$, $`jc𝒥/(\mu GM)`$, $`\widehat{\omega }GM\dot{\phi }/c^3`$, the LSO quantities defined, in the equal-mass case $`\nu =1/4`$, by the Hamiltonian (11), take the following values $`r_{\mathrm{LSO}}\left(1/4\right)=5.718,j_{\mathrm{LSO}}\left(1/4\right)=3.404,`$ (18) $`\widehat{\omega }_{\mathrm{LSO}}\left(1/4\right)=0.07340,{\displaystyle \frac{_{\mathrm{real}}^{\mathrm{LSO}}\left(1/4\right)Mc^2}{Mc^2}}=0.01501.`$ (19) Note that the comparable-mass LSO is slightly more inwards (both in terms of the coordinate $`R`$ and in the sense of having a higher orbital frequency) than its corresponding rescaled test-mass limit: $`r_{\mathrm{LSO}}(0)=6`$, $`j_{\mathrm{LSO}}(0)=\sqrt{12}=3.4641`$, $`\widehat{\omega }_{\mathrm{LSO}}(0)=6^{3/2}=0.068041`$. As we shall need in the following to refer to the numerical value of $`\widehat{\omega }_{\mathrm{LSO}}(\nu )`$ for arbitrary values of $`\nu `$, we have fitted the result obtained by the (rather intricate) method of Ref. to a simple polynomial in $`\nu `$. We find $`\widehat{\omega }_{\mathrm{LSO}}(\nu )\omega _0[1+\omega _1(4\nu )+\omega _2(4\nu )^2],`$ (20) $`\omega _0=0.0680414,\omega _1=0.0693305,\omega _2=0.00935142.`$ (21) ## III Incorporating radiation reaction effects We wish to augment the conservative dynamics described in the previous section by adding, as accurately as possible, radiation reaction effects. If we were doing it in the Lagrangian formalism we would write (in any coordinate system) $$\frac{d}{dt}\frac{L}{\dot{Q}^i}\frac{L}{Q^i}=_i^{\mathrm{Lag}}(Q,\dot{Q}).$$ (22) This would define the additional damping force $`_i^{\mathrm{Lag}}(Q,\dot{Q})`$ needed in the Lagrangian formalism. In particular, in polar coordinates we would write (for planar motion $`\theta =\pi /2`$): $`{\displaystyle \frac{d}{dt}}{\displaystyle \frac{L}{\dot{R}}}{\displaystyle \frac{L}{R}}=_R^{\mathrm{Lag}}(R,\phi ,\dot{R},\dot{\phi }),`$ (23) $`{\displaystyle \frac{d}{dt}}{\displaystyle \frac{L}{\dot{\phi }}}=_\phi ^{\mathrm{Lag}}(R,\phi ,\dot{R},\dot{\phi }).`$ (24) We want to work in the Hamiltonian framework, hence coming back to the coordinates $`R`$, $`P_R`$, $`\phi `$ and $`P_\phi `$ and imposing the constraint that the usual definition $`P_i=L/\dot{Q}^i`$ holds without corrections (which implies that the other usual relations $`\dot{Q}^i=H/P_i`$, $`H/Q^i=L/Q^i`$ and Eq. (12) hold too) we get $`{\displaystyle \frac{dR}{dt}}{\displaystyle \frac{H}{P_R}}(R,P_R,P_\phi )=0,`$ (25) $`{\displaystyle \frac{d\phi }{dt}}{\displaystyle \frac{H}{P_\phi }}(R,P_R,P_\phi )=0,`$ (26) $`{\displaystyle \frac{dP_R}{dt}}+{\displaystyle \frac{H}{R}}(R,P_R,P_\phi )=_R^{\mathrm{Ham}}(R,\phi ,P_R,P_\phi ),`$ (27) $`{\displaystyle \frac{dP_\phi }{dt}}=_\phi ^{\mathrm{Ham}}(R,\phi ,P_R,P_\phi ),`$ (28) where the Hamiltonian damping force $`_i^{\mathrm{Ham}}(Q^j,P_j)`$ is numerically equal to the Lagrangian one: $`_i^{\mathrm{Ham}}(Q^j,P_j)=_i^{\mathrm{Lag}}(Q^j,\dot{Q}^j=H/P_j)`$ . ### A What do we know about the radiation reaction force? The radiation reaction force $``$ was computed explicitly, at lowest (Newtonian) fractional order, in harmonic Cartesian-like coordinates, as part of the complete 2.5PN equations of motion, by Damour and Deruelle . An equivalent result was also derived within the ADM canonical formalism by Schäfer . At higher post-Newtonian orders one has only an incomplete knowledge of the equations of motion, and one has to rely on the (assumed) balance between energy and angular momentum losses in the system and at infinity . To get an idea of the generic structure of the radiation damping (in various coordinate systems, and at various PN approximations) let us consider the general radiation reaction force written (at 1PN fractional accuracy; and setting $`G=1`$) by Iyer and Will . $`_i^{\mathrm{Lag}}=\mu \left[\alpha (R,v)\dot{R}n^i+\beta (R,v)v^i\right],`$ (29) $`\alpha (R,v)={\displaystyle \frac{8}{5}}\nu {\displaystyle \frac{M}{R^2}}{\displaystyle \frac{M}{R}}(A_{5/2}+A_{7/2}^{}+\mathrm{}),`$ (30) $`\beta (R,v)={\displaystyle \frac{8}{5}}\nu {\displaystyle \frac{M}{R^2}}{\displaystyle \frac{M}{R}}(B_{5/2}+B_{7/2}^{}+\mathrm{}),`$ (31) where $`R`$ is the relative radius and $`v`$ is the velocity. Then, using post-Newtonian expressions for the energy and the angular momentum flux at infinity, and assuming energy and angular momentum balance, they obtained at lowest (Newtonian) fractional order $`A_{5/2}=3(1+\overline{\beta })v^2+{\displaystyle \frac{1}{3}}(23+6\overline{\alpha }9\overline{\beta }){\displaystyle \frac{M}{R}}5\overline{\beta }\dot{R}^2,`$ (32) $`B_{5/2}=(2+\overline{\alpha })v^2+(2\overline{\alpha }){\displaystyle \frac{M}{R}}3(1+\overline{\alpha })\dot{R}^2.`$ (33) See Ref. for the expressions of the 1PN-accurate radiation damping terms $`A_{7/2}`$ and $`B_{7/2}`$ in the equations of motion (equivalent, after some reshuffling, with the Lagrangian contributions $`A_{7/2}^{}`$, $`B_{7/2}^{}`$ in Eqs. (30), (31)). The coefficients $`\overline{\alpha }`$ and $`\overline{\beta }`$ that appear in Eqs. (32) and (33) are two arbitrary gauge parameters that cannot be fixed by the energy balance method. Iyer and Will showed that this gauge freedom is equivalent to shifting the (conservative) coordinate system by small radiative corrections. Let us notice that the gauge dependence is reduced when considering quasi-circular orbits. Indeed, in that case $`\dot{R}^20`$, $`M/Rv^2`$ and Eqs. (30), (31) become (considering only the 5/2PN terms which are sufficient for the point we wish to make) $$\alpha _{\mathrm{circ}}\frac{8}{5}\nu \frac{M}{R^2}\left(\frac{M}{R}\right)^2\left(\frac{32}{3}+2\overline{\alpha }\right),\beta _{\mathrm{circ}}\frac{32}{5}\nu \frac{M}{R^2}\left(\frac{M}{R}\right)^2.$$ (34) Hence, in the quasi-circular case the only gauge-dependence left is in the coefficient $`\alpha (R,v)`$ multiplying the radial component of the damping force $`(n^i)`$. We can use this gauge arbitrariness to set the ratio $$\left(\frac{\alpha }{\beta }\right)_{\mathrm{circ}.}\frac{1}{2}\left(\frac{16}{3}+\overline{\alpha }\right),$$ (35) to any value we like. For example, by choosing $`\overline{\alpha }=16/3`$ we can set $`\alpha _{\mathrm{circ}.}=0`$ or by choosing $`\overline{\alpha }=10/3`$ we can set $`(\alpha +\beta )_{\mathrm{circ}.}=0`$. Having understood the gauge dependence of the coefficient $`\alpha `$ in Eq. (29) let us come back to the general structure (29) (considered at any PN accuracy, with some (unknown) coefficients $`\alpha `$ and $`\beta `$). The polar-coordinate version (for planar motion $`\theta =\pi /2`$) of the Cartesian-like Lagrangian damping force (29) reads (23), (24) with $`_R^{\mathrm{Lag}}=_i^{\mathrm{Lag}}{\displaystyle \frac{Q^i}{R}}=n^i_i^{\mathrm{Lag}},`$ (36) $`_\phi ^{\mathrm{Lag}}=_i^{\mathrm{Lag}}{\displaystyle \frac{Q^i}{\phi }}=Q^x_y^{\mathrm{Lag}}Q^y_x^{\mathrm{Lag}}.`$ (37) This yields $$_\phi ^{\mathrm{Lag}}=\mu \beta R^2\dot{\phi },_R^{\mathrm{Lag}}=\mu (\alpha +\beta )\dot{R}.$$ (38) The important information for our present purpose is the difference between the $`\phi `$-component of the damping force, which contains only $`\beta `$ and is, therefore, gauge-independent <sup>§</sup><sup>§</sup>§The discussion above concerns only the lowest-order term in $`\beta `$, but we shall see below that, to all orders, the crucial combination $`\beta R^2`$ can, for circular orbits, be expressed in terms of invariant quantities., and the $`R`$-component which contains the gauge-dependent combination $`\alpha +\beta `$. Let us note, in particular, the expression of the ratio $$\frac{_R^{\mathrm{Lag}}}{_\phi ^{\mathrm{Lag}}}=\left(\frac{\alpha }{\beta }+1\right)\frac{\dot{R}}{R^2\dot{\phi }}.$$ (39) In the following we shall be interested in quasi-circular motions with $`\dot{R}R\dot{\phi }`$. \[We shall see that this condition remains satisfied even during part of the plunge phase.\] As we see from Eq. (39), for such motions the radial component of the damping force will contain one power of the small dimensionless quantity $`\dot{R}/(R\dot{\phi })`$. But we learned above, from the gauge dependence of the lowest-order damping force, that we can change the definition of the radial coordinate so as to set, for instance, the quantity $`(\alpha /\beta )+1`$ to zero (for circular orbits). This means that the RHS of Eq. (39) can be arranged, in the case of quasi-circular orbits, to contain three powers of the small parameter $`\dot{R}/(R\dot{\phi })`$. (From Eqs. (32), (33) we see that for quasi-circular orbits $`\alpha +\beta \dot{R}^2`$.) We have checked that the reasoning made above, using the lowest-order gauge dependence, can be formally extended to all higher PN orders. The conclusion is that there should exist a special coordinate gauge where, for quasi-circular motions, an excellent approximation to the damping force is obtained by replacing the radial component simply by zero: $$_R^{\mathrm{Lag}}=0=_R^{\mathrm{Ham}}.$$ (40) To test, a posteriori, the robustness of the approximation (40), we shall also consider another special gauge: namely that where $`(\alpha /\beta )_{\mathrm{circ}}=0`$. \[As we said above, this can be achieved at lowest order by a suitable choice of $`\overline{\alpha }`$, and this can be extended to higher PN orders by suitable choices of higher gauge parameters.\] Finally, this means that there exists another coordinate gauge where, to an excellent approximation, the radial damping force is given as $$_R^{\mathrm{Lag}}=_R^{\mathrm{Ham}}=\frac{\dot{R}}{R^2\dot{\phi }}_\phi ^{\mathrm{Ham}}.$$ (41) The results in the two gauges are compared and discussed at the end of Sec. V. What is important for the following is that in both gauges (40) or (41), the knowledge of the full damping force can be deduced from the sole knowledge of $`_\phi `$. ### B Non-perturbative estimate of the angular momentum reaction force along quasi-circular orbits The analysis of the previous subsection has shown that the crucial equation in which one should accurately incorporate radiation reaction effects is $$\frac{dP_\phi }{dt}=_\phi ^{\mathrm{Ham}}(R,\phi ,P_R,P_\phi ).$$ (42) As $`P_\phi `$ is just the total angular momentum of the binary system, Eq. (42) expresses the rate of loss of angular momentum under gravitational radiation reaction. As usual we shall estimate the RHS $`_\phi =_\phi ^{\mathrm{Ham}}=_\phi ^{\mathrm{Lag}}`$ (remember that $`^{\mathrm{Ham}}`$ and $`^{\mathrm{Lag}}`$ differ only in the arguments in which they are expressed) by assuming that there is a balance between the mechanical angular momentum lost by the system, and the flux of angular momentum at infinity in the form of gravitational waves. In the case of interest here of quasi-circular orbits we expect that, to a good approximation, $`_\phi `$ will not depend explicitly on $`\phi `$ and will, therefore, be expressible in terms of the orbit-averaged flux of angular momentum. Moreover, in the case of quasi-circular orbits there is a simple relation between angular-momentum-loss and energy-loss. Indeed, the rate of energy-loss along any orbit, in polar coordinates, is given by $$\frac{d}{dt}=\frac{dH}{dt}=\dot{R}_R+\dot{\phi }_\phi ,$$ (43) and in particular along quasi-circular orbit we have (remembering Eq. (39)) $$\left(\frac{dH}{dt}\right)_{\mathrm{quasi}\mathrm{circ}.}\dot{\phi }_\phi ^{\mathrm{circ}.}+𝒪(\dot{R}^2).$$ (44) Finally, if we know some good estimate of the (averaged) energy-loss along circular orbits, say $$\left(\frac{dH}{dt}\right)_{\mathrm{circ}.}\mathrm{\Phi }_{\mathrm{circ}.}(\dot{\phi }),$$ (45) we can obtain a good estimate of the needed $`\phi `$-reactive force $$_\phi ^{\mathrm{circ}.}\frac{\mathrm{\Phi }_{\mathrm{circ}.}(\dot{\phi })}{\dot{\phi }}.$$ (46) The problem of giving a non-perturbative, re-summed estimate of the energy-loss-rate (or “flux function”) along circular orbits, say $`\mathrm{\Phi }_{\mathrm{circ}}`$, has been recently tackled by Damour, Iyer and Sathyaprakash . By combining several of the non-perturbative techniques recalled above (to work with an invariant function $`F(v)`$, to use some global information about $`F(v)`$ in the complex $`v`$-plane, to use Padé approximants) Ref. came up with the following expression for $`\mathrm{\Phi }_{\mathrm{circ}}`$, considered as a function of the gauge-invariant observable $$v_\omega (GM\omega /c^3)^{1/3};\omega \dot{\phi },$$ (47) namely, $$\mathrm{\Phi }_{\mathrm{circ}.}=F_{\mathrm{DIS}}(v_\omega )=\frac{32}{5G}\nu ^2v_\omega ^{10}\frac{\widehat{f}_{\mathrm{DIS}}(v_\omega ;\nu )}{1v_\omega /v_{\mathrm{pole}}(\nu )}.$$ (48) Here, and in the following, we set $`c=1`$ to simplify formulas. The function $`\widehat{f}_{\mathrm{DIS}}(v_\omega ;\nu )`$ entering Eq. (48) is the “factored flux function” of , scaled to the Newtonian (quadrupole) flux (hence the caret on $`f_{\mathrm{DIS}}`$). It was shown in that the sequence of near-diagonal Padé approximants of $`\widehat{f}_{\mathrm{DIS}}(v)`$ exhibits a very good convergence (at least in the $`\nu =0`$ limit where high-order PN expansions are known ) toward the exact result (numerically known when $`\nu =0`$ ). On this basis, it was argued in that, in the comparable-mass case, $`\nu 0`$, our “best estimate” of $`\widehat{f}`$ is obtained by Padeeing the currently most complete post-Newtonian results, namely the 2.5PN ones . This yields a result of the form $$\widehat{f}_{\mathrm{DIS}}(v;\nu )=\frac{1}{1+\frac{c_1v}{1+\frac{c_2v}{1+\frac{c_3v}{1+\frac{c_4v}{1+c_5v}}}}},$$ (49) where the dimensionless coefficients $`c_i`$ depend only on $`\nu `$. The $`c_k`$’s are some explicit functions of the coefficients $`f_k`$ of the straightforward Taylor expansion of $`\widehat{f}(v)`$. In turn, the $`f_k`$’s, being defined by the identity (where $`T`$ means “Taylor expansion”) $$T[\widehat{f}(v)]T\left[\left(1\frac{v}{v_{\mathrm{pole}}}\right)\widehat{F}(v)\right]=1+f_1v+f_2v^2+\mathrm{}$$ (50) are given by $$f_k=F_kF_{k1}/v_{\mathrm{pole}}$$ (51) in terms of the Taylor coefficients of the usual (Newton-normalized) flux function $$T[\widehat{F}(v)]T\left[\frac{5G}{32\nu ^2v^{10}}F(v)\right]=1+F_2v^2+F_3v^3+\mathrm{}.$$ (52) \[Note that $`F_1=0`$, but that $`f_1=1/v_{\mathrm{pole}}0`$.\] More explicitly we have $$F_2=\frac{1247}{336}\frac{35}{12}\nu ,F_3=4\pi ,$$ (53) $$F_4=\frac{44711}{9072}+\frac{9271}{504}\nu +\frac{65}{18}\nu ^2,F_5=\left(\frac{8191}{672}+\frac{535}{24}\nu \right)\pi ,$$ (54) and $`c_1=f_1,c_2=f_1{\displaystyle \frac{f_2}{f_1}},c_3={\displaystyle \frac{f_1f_3f_2^2}{f_1(f_1^2f_2)}},`$ (55) $`c_4={\displaystyle \frac{f_1(f_2^3+f_3^2+f_1^2f_4f_2(2f_1f_3+f_4))}{(f_1^2f_2)(f_1f_3f_2^2)}},`$ (56) $`c_5={\displaystyle \frac{(f_1^2f_2)(f_3^3+2f_2f_3f_4f_1f_4^2f_2^2f_5+f_1f_3f_5)}{(f_1f_3f_2^2)(f_2^3+f_3^2+f_1^2f_4f_2(2f_1f_3+f_4))}}.`$ (57) As is clear from these expressions, they depend on the definition used for the quantity $`v_{\mathrm{pole}}(\nu )`$ which represents a $`\nu `$-dependent estimate of the location of the “pole” in $`\mathrm{\Phi }_{\mathrm{circ}}`$, which coincides (see the discussion in ) with the location of the “light-ring” or last unstable circular orbit ( $`R_{\mathrm{light}\mathrm{ring}}^{\mathrm{Schw}.}=3GM`$ in the $`\nu 0`$ limit). Actually, as we shall use the Padé representation only above and around the LSO ($`R_{\mathrm{LSO}}^{\mathrm{Schw}.}=6GM`$ when $`\nu =0`$) the precise choice of $`v_{\mathrm{pole}}(\nu )`$ is probably not crucial (as long as it stays near its known $`\nu =0`$ limit: $`v_{\mathrm{pole}}(\nu =0)=1/\sqrt{3}`$). In this work, we shall follow Ref. and use the pole location they obtained from Padeing their “new” energy function $`e(x)`$, namely $$v_{\mathrm{pole}}^{\mathrm{DIS}}=\frac{1}{\sqrt{3}}\sqrt{\frac{1+\frac{1}{3}\nu }{1\frac{35}{36}\nu }}.$$ (58) Then, combining Eqs. (46), (47) and (48) we define our best estimate of the $`\phi `$-component of the radiation reactive force along quasi-circular orbits as: $$_\phi ^{\mathrm{circ}.}\frac{GM}{v_\omega ^3}\mathrm{\Phi }_{\mathrm{DIS}}(v_\omega )=\frac{32}{5}\mu \nu v_\omega ^7\frac{\widehat{f}_{\mathrm{DIS}}(v_\omega ;\nu )}{1v_\omega /v_{\mathrm{pole}}^{\mathrm{DIS}}(\nu )}.$$ (59) To ease the notation we shall work in the following with reduced quantities, that is: $`r{\displaystyle \frac{R}{GM}},p_r{\displaystyle \frac{P_R}{\mu }},p_\phi {\displaystyle \frac{P_\phi }{\mu GM}}={\displaystyle \frac{𝒥}{\mu GM}}j,`$ (60) $`\widehat{t}{\displaystyle \frac{t}{GM}},\widehat{H}{\displaystyle \frac{H_{\mathrm{real}}^{\mathrm{improved}}}{\mu }},\widehat{H}_{\mathrm{eff}}{\displaystyle \frac{H_{\mathrm{eff}}}{\mu }}.`$ (61) Finally, the dynamics, including radiation reaction, in re-scaled coordinates, is explicitly described by the following system of equations (in the “canonical” case where Eq. (40) holds) $`{\displaystyle \frac{dr}{d\widehat{t}}}={\displaystyle \frac{\widehat{H}}{p_r}}(r,p_r,p_\phi ),`$ (62) $`{\displaystyle \frac{d\phi }{d\widehat{t}}}=\widehat{\omega }{\displaystyle \frac{\widehat{H}}{p_\phi }}(r,p_r,p_\phi ),`$ (63) $`{\displaystyle \frac{dp_r}{d\widehat{t}}}+{\displaystyle \frac{\widehat{H}}{r}}(r,p_r,p_\phi )=0,`$ (64) $`{\displaystyle \frac{dp_\phi }{d\widehat{t}}}=\widehat{}_\phi (\widehat{\omega }(r,p_r,p_\phi )),`$ (65) with $$\widehat{H}=\frac{1}{\nu }\sqrt{1+2\nu \left[\sqrt{A(r)\left(1+\frac{p_r^2}{B(r)}+\frac{p_\phi ^2}{r^2}\right)}1\right]},$$ (66) $$\widehat{}_\phi (v_\omega \widehat{\omega }^{1/3})=\frac{_\phi }{\mu }=\frac{32}{5}\nu v_\omega ^7\frac{\widehat{f}_{\mathrm{DIS}}(v_\omega ;\nu )}{1v_\omega /v_{\mathrm{pole}}^{\mathrm{DIS}}(\nu )},$$ (67) and where in Eq. (66) we use the scaled versions of our current best estimate of the effective metric coefficients $`A(r)`$, $`B(r)`$, see and Eqs. (14)–(16) above, that is $$A(r)1\frac{2}{r}+\frac{2\nu }{r^3},B(r)\frac{1}{A(r)}\left(1\frac{6\nu }{r^2}\right).$$ (68) Note that the argument $`v_\omega `$ entering $`\widehat{}_\phi `$, Eq. (67), is simply defined as $`v_\omega \widehat{\omega }^{1/3}`$, where $`\widehat{\omega }\omega (GM)`$ is the function of $`r`$, $`p_r`$ and $`p_\phi `$ defined by Eq. (63), i.e. $`\widehat{\omega }(r,p_r,p_\phi )\widehat{H}(r,p_r,p_\phi )/p_\phi `$. ## IV Transition between inspiral and plunge The first-order evolution system (62)–(65) defines our proposed best estimate for completing the usually considered “adiabatic” inspiral evolution into a system which exhibits a smooth transition between inspiral and plunge. The rest of this paper will be devoted to extracting some of the important information contained in this new evolution system. Before coming to grips with such detailed information, it is useful to have a first visual impression of the physics contained in our system (62)–(65). To do this we plot on the left panel of Fig. 1 the result of a full numerical evolution of Eqs. (62)–(65) in the equal-mass case ($`\nu =1/4`$). We started the evolution at $`r=15`$, $`\phi =0`$ and used as initial values for $`p_\phi `$ and $`p_r`$ the ones provided by the adiabatic approximation (see Eqs. (74) and (81) below). The dashed circle in this plot indicates the radial coordinate location of the LSO defined by the conservative part of the dynamics, i.e. by the Hamiltonian $`\widehat{H}(r,p_r,p_\phi )`$. More precisely this “$`r`$-LSO” is simply defined (for any $`\nu `$) by $`r=r_{\mathrm{LSO}}(\nu )`$, where $`r_{\mathrm{LSO}}(\nu )`$ is the solution of Eq. (17). In particular, $`r_{\mathrm{LSO}}\left(\frac{1}{4}\right)=5.718`$, as recalled in Eq. (19). Note that, in presence of radiation reaction effects, there is an arbitrariness in what one would like to mean by saying: “the system is crossing the LSO”. Indeed, we could define the “LSO-crossing” in several inequivalent ways, notably: (i) $`r`$-LSO: the time when $`r=r_{\mathrm{LSO}}(\nu )`$; (ii) $`j`$-LSO: the time when $`p_\phi j=j_{\mathrm{LSO}}(\nu )`$; (iii) $`\omega `$-LSO: the time when $`d\phi /d\widehat{t}\widehat{\omega }=\widehat{\omega }_{\mathrm{LSO}}(\nu )`$. \[The “LSO” functions of $`\nu `$ being defined by solving Eq. (17); see Eq. (19).\] This arbitrariness is not a problem. Our new evolution system (62)–(65) describes a smooth transition “through” the formally defined “old” LSO, and does not care about old definitions. In other words, when $`\nu `$ is finite, and especially when $`\nu 1/4`$ (which, one should remember, is expected to be an accumulation point of observed values of $`\nu `$; see footnote I above) the smooth transition process blurs the notion of LSO. It is only for $`\nu 1`$ (see below) that one recovers a sharp transition near the $`H`$-defined LSO. On the right panel of Fig. 1 we compare the two kinetic contributions to the Hamiltonian (66): the “azimuthal” contribution $`p_\phi ^2/r^2`$, and the “radial” contribution $`p_r^2/B(r)`$. One sees on this Figure that our basic assumption of quasi-circularity (which, at the level of $`\widehat{H}`$, means $`p_r^2/B(r)p_\phi ^2/r^2`$) is well satisfied throughout the transition. In fact, even down to $`r3.79`$ one has $`p_r^2/B(r)<\mathrm{\hspace{0.17em}0.1}p_\phi ^2/r^2`$. The radial kinetic energy would become equal to the azimuthal one only below $`r=3`$. We shall, anyway, not use, in the following, our system below the (usual) “light-ring” $`r3`$ (where $`p_r^2/B(r)0.30p_\phi ^2/r^2`$). We exhibit more quantitative results on the transition between the inspiral and the plunge in Figs. 2 and 3. These figures plot the values of several physical quantities (energy, angular momentum, radial velocity and radial coordinate) computed at the $`\omega `$-LSO (i.e. when $`\omega =\omega _{\mathrm{LSO}}(\nu )`$) after integration of the system (62)–(65). The energy which is plotted is the reduced non-relativistic real energy, i.e. $`(_{\mathrm{real}}M)/\mu `$. \[In the test-mass limit, this reduced energy equals $`\sqrt{8/9}1=0.057191`$.\] Having obtained, through Figs. 2 and 3, a first impression of the physics of the inspiral $``$ plunge transition, we shall now study in more detail this transition, notably by comparing it with various analytical approximations. The first approximation we shall consider is the current standard one used for dealing with the inspiral phase: the adiabatic approximation. ### A Comparison with the adiabatic approximation Let us compare the exact numerical evolution with the usual adiabatic approximation to inspiral motion. This approximation is defined by saying that the (effective) body follows an adiabatic sequence of exact circular orbits whose energy is slowly drained out by gravitational radiation. It is obtained from Eqs. (62), (65), by neglecting $`p_r^2`$, i.e. by setting $`p_r=0`$. Noticing that $`\widehat{H}/p_r=2p_r\widehat{H}/p_r^2p_r`$ we get that $`dr/d\widehat{t}`$ vanishes linearly with $`p_r`$. The first equation (62) is then formally satisfied with $`p_r=0=\dot{r}`$. Imposing now $`p_r=0`$ in Eqs. (63) and (64) we obtain two further equations: $`{\displaystyle \frac{\widehat{H}_0}{r}}(r,p_\phi )=0,`$ (69) $`\widehat{\omega }={\displaystyle \frac{\widehat{H}_0}{p_\phi }}(r,p_\phi ),`$ (70) where we define $$\widehat{H}_0(r,p_\phi )\widehat{H}(r,p_r=0,p_\phi )=\frac{1}{\nu }\sqrt{1+2\nu \left[\sqrt{A(r)\left(1+\frac{p_\phi ^2}{r^2}\right)}1\right]}.$$ (71) Eq. (69) provides a link between $`r`$ and $`p_\phi j`$ in the adiabatic limit. From the structure (66) of $`\widehat{H}`$, it is easily seen that Eq. (69) is equivalent to looking for the minimum, say (for convenience) in the variable $`u1/r`$, of the “radial potential” $$W_j(u)=A(u)[1+j^2u^2].$$ (72) Solving $`_uW_j(u)=0`$ gives a parametric representation of $`j^2`$ in terms of $`u`$: $$j_{\mathrm{adiab}}^2(u)=\frac{A^{}(u)}{(u^2A(u))^{}},$$ (73) where the prime denote $`d/du`$. In the case where the function $`A`$ is given by Eq. (68), i.e. $`A(u)=12u+2\nu u^3`$, Eq. (73) yields, in term of the orignal (reduced) radial variable $`r=1/u`$ $$j_{\mathrm{adiab}.}^2(r)=\frac{r^2(r^23\nu )}{r^33r^2+5\nu }.$$ (74) Note that there exist real circular orbits (though possibly unstable ones) as long as $`j_{\mathrm{adiab}}^2(r)>0`$, i.e. as long as $`r^33r^2+5\nu >0`$. In fact the positive, real solution in $`r`$ of $$[r^33r^2+5\nu ]_{\mathrm{light}\mathrm{ring}}=0$$ (75) defines the light-ring or last unstable circular orbit (with $`j^2(r_{\mathrm{light}\mathrm{ring}})=+\mathrm{}`$). We find $`r_{\mathrm{light}\mathrm{ring}}2.84563`$ in the case $`\nu =1/4`$. Eq. (70) then gives the parametric representation of $`\widehat{\omega }=\omega (GM)`$ throughout the adiabatic phase for circular orbits: $$\widehat{\omega }_{\mathrm{adiab}.}(r)=\frac{1}{r^{3/2}}\frac{\sqrt{13\nu /r^2}}{\sqrt{1+2\nu (\sqrt{z(r)}1)}},$$ (76) where $`z(r)`$ denotes the following quantity $$z(r)\widehat{H}_{\mathrm{eff}}^2(r,p_r=0,p_\phi =j_{\mathrm{adiab}.})=\frac{r^3A^2(r)}{r^33r^2+5\nu }.$$ (77) Note that the effective one-body description seems to become somewhat unsatisfactory at the light-ring (at least for exactly circular orbits). Indeed, we see from Eqs. (76) and (77) that the blow up of $`z(r)`$, i.e. of the effective energy, at the light-ring, Eq. (75), implies that the real orbital frequency of circular orbits, $`\widehat{\omega }_{\mathrm{circ}.}(r)`$, Eq. (76), tends to zero at the light-ring. This is probably an unphysical behaviour \[from the test-mass limit, one expects the orbital frequency to have a non-zero limit at the light-ring; see, e.g., Ref. where Padé approximants are used to compute a finite value of $`\widehat{\omega }_{\mathrm{light}\mathrm{ring}}(\nu )`$\]. The other factors in Eq. (76) imply, as expected, a regular increase of $`\widehat{\omega }(r)`$ as $`r`$ decreases below the LSO. Pending the construction of an improved version of the effective one-body approach which would be better behaved, we have decided, when dealing with the evolution of the system (62)–(65), to stop the simulation at the light-ring. \[In our simulations of plunging orbits the effective energy stays bounded, but the orbital frequency $`\widehat{\omega }(\widehat{t})`$ levels off very close to the light-ring.\] Finally Eq. (65) becomes in the adiabatic limit $$\frac{dj}{d\widehat{t}}=\widehat{}_\phi \left(\frac{\widehat{H}_0}{p_\phi }(r,j)\right).$$ (78) Then using $`dj/d\widehat{t}=(dj/dr)(dr/d\widehat{t})`$ and $`d\phi =\widehat{\omega }d\widehat{t}`$ we can solve the motion in the adiabatic limit by quadratures: $$d\widehat{t}_{\mathrm{adiab}.}=\left(\frac{dj_{\mathrm{adiab}.}}{dr}\right)\frac{dr}{\widehat{}_\phi (\widehat{\omega }_{\mathrm{adiab}.}(r))},$$ (79) $$d\phi _{\mathrm{adiab}.}=\left(\frac{dj_{\mathrm{adiab}.}}{dr}\right)\frac{\widehat{\omega }_{\mathrm{adiab}.}(r)}{\widehat{}_\phi (\widehat{\omega }_{\mathrm{adiab}.}(r))}dr.$$ (80) The radial velocity $`v_rdr/d\widehat{t}`$, as a function of the parameter $`r`$, in the adiabatic limit, is given by: $$v_r^{\mathrm{adiab}.}=\frac{\widehat{}_\phi (\widehat{\omega }_{\mathrm{adiab}.}(r))}{dj_{\mathrm{adiab}.}/dr}.$$ (81) Note that $`v_r^{\mathrm{adiab}.}`$ formally tends to $`\mathrm{}`$ when $`rr_{\mathrm{LSO}}`$ (indeed, $`j_{\mathrm{adiab}.}(r)`$ reaches, by definition, a minimum at $`r=r_{\mathrm{LSO}}`$). This shows that the adiabatic approximation is meaningful only during the inspiral phase (i.e. “above” the LSO). In Figs. 4, 5 we compare, for $`\nu =1/4`$, the number of gravitational cycles, defined by $`𝒩_{\mathrm{GW}}=\phi _{\mathrm{GW}}/(2\pi )=\phi /\pi `$, the orbital angular frequency $`\omega `$ (or, equivalently, the gravitational wave frequency, $`f_{\mathrm{GW}}=\omega _{\mathrm{GW}}/(2\pi )=\omega /\pi `$), and the radial velocity, computed with the exact equations of motion and in the adiabatic limit, as well as the gravitational waveform. These figures show that, in the equal-mass case $`\nu =1/4`$, the adiabatic approximation starts to significantly deviate from the exact evolution quite before one reaches the LSO. Fig. 4 is normalized so that $`𝒩_{\mathrm{GW}}^{\mathrm{adiab}}`$ and $`𝒩_{\mathrm{GW}}^{\mathrm{exact}}`$ coincide for large values of $`R/GM`$, and that $`𝒩_{\mathrm{GW}}^{\mathrm{adiab}}`$ be zero at the $`r`$-LSO. For instance, we find that the number of GW cycles given by the adiabatic approximation differs from the exact number already by 0.1 when $`r8.8`$, and that $`𝒩_{\mathrm{GW}}^{\mathrm{exact}}(r_{\mathrm{LSO}})=0.9013`$ . The left panel of Fig. 5 contrasts $`\omega /\omega _{\mathrm{LSO}}`$ ($`=f_{\mathrm{GW}}/f_{\mathrm{GW}\mathrm{LSO}}^{\mathrm{Schw}.}`$ where $`f_{\mathrm{GW}\mathrm{LSO}}^{\mathrm{Schw}.}=6^{3/2}/GM\pi `$ is the fiducial Schwarzschild LSO GW frequency), computed with the exact evolution and within the adiabatic approximation, as a function of time. Note that, for the horizontal axis we use $`\widehat{\omega }_{\mathrm{LSO}}(0)(\widehat{t}\widehat{t}_{\mathrm{LSO}})`$, where $`\widehat{\omega }_{\mathrm{LSO}}(0)=\pi \widehat{f}_{\mathrm{GW}\mathrm{LSO}}^{\mathrm{Schw}.}=6^{3/2}`$ (provided by the $`\nu 0`$ limit of Eq. (20)) and $`\widehat{t}_{\mathrm{LSO}}`$ is defined as the time at which the adiabatic solution reaches the $`r`$-LSO position. Finally, on the right panel of Fig. 5 we compare the last few GW cycles of the exact and the adiabatic restricted waveform, i.e. $`h(t)v^2\mathrm{cos}\varphi _{\mathrm{GW}}(t)`$, with $`v=(d\phi /d\widehat{t})^{1/3}`$ and $`\varphi _{\mathrm{GW}}=2\phi `$, in the crucial interesting region around the LSO. By adiabatic restricted waveform we mean the restricted waveform in which $`\phi (\widehat{t})=\phi _{\mathrm{adiab}.}(\widehat{t})`$ is derived by integrating the two equations (79) and (80) (which give a parametric representation of $`\widehat{t}_{\mathrm{adiab}.}(r)`$ and $`\widehat{\phi }_{\mathrm{adiab}.}(r)`$ in terms of the auxiliary parameter $`r`$). Note in Fig. 5 that the dephasing between the two waveforms becomes visible somewhat before the LSO (we shall dwell more on this subject in Section V). Note also that the time when the adiabatic evolution reaches the LSO (“adiabatic LSO”) corresponds to a time when the exact evolution reaches a frequency $`\omega 0.80\omega _{\mathrm{LSO}}(0)`$, i.e. a time significantly before the $`\omega `$-LSO. This is why there are more cycles after the adiabatic LSO in Fig. 5 (more than two cycles), than there will be after the (exact) $`\omega `$-LSO (we shall see below that $`N_{\mathrm{GW}}^{\mathrm{afterLSO}}=2N_{\mathrm{orbit}}^{\mathrm{afterLSO}}=1.2048`$ for $`\nu =1/4`$). ### B The $`\dot{r}`$-linearized approximation The previous subsection has shown the severe shortcomings of the adiabatic approximation. Let us now consider a second analytical approximation which is more accurate than the adiabatic one, and which, in particular, allows one to see analytically what happens during the transition between the inspiral and the plunge. This approximation is based on a simple linearization with respect to the radial velocity $`dr/d\widehat{t}`$, which is small during the inspiral, as well as the beginning of the plunge. As $`\widehat{H}`$ depends quadratically on $`p_r`$ and $`p_r1`$ we pose $$C_r(r,j)\left[\frac{1}{p_r}\frac{\widehat{H}}{p_r}(r,p_r,j)\right]_{p_r0}=\frac{1}{\nu \widehat{H}_0(r,j)}\frac{1}{\widehat{H}_{\mathrm{eff}}^0(r,j)}\frac{A^2(r)}{16\nu /r^2},$$ (82) (note that $`C_r`$ is a positive quantity), where $$\widehat{H}_{\mathrm{eff}}^0(r,j)=\widehat{H}_{\mathrm{eff}}(r,p_r=0,j)=\sqrt{A(r)\left(1+\frac{j^2}{r^2}\right)}.$$ (83) Then, modulo $`p_r^2`$ fractional effects that we neglect, we can write $$\frac{dr}{d\widehat{t}}C_r(r,j)p_r.$$ (84) Differentiating twice the above equation with respect to time, we obtain $$\frac{d^2p_r}{d\widehat{t}^2}\frac{1}{C_r(r,j)}\frac{d^3r}{d\widehat{t}^3},$$ (85) when neglecting some nonlinear terms $`(dr/d\widehat{t})^2`$ and $`(dr/d\widehat{t})(dj/d\widehat{t})`$. On the other hand, taking the derivative with respect to time of Eq. (64) and neglecting fractional corrections of $`𝒪(p_r^2)`$, we end up with $$\frac{d^2p_r}{d\widehat{t}^2}=\frac{d}{d\widehat{t}}\frac{\widehat{H}}{r}(r,p_r,p_\phi )\frac{^2\widehat{H}_0}{r^2}\frac{dr}{d\widehat{t}}\frac{^2\widehat{H}_0}{rj}\widehat{}_\phi .$$ (86) To get an autonomous system we further approximate $`j`$ by solving for $`j`$ in the lowest-order approximation to Eq. (64), obtained by neglecting both $`p_r`$ and $`dp_r/d\widehat{t}`$. In other words, $`j(r)`$ is obtained, as in adiabatic approximation, by solving Eq. (69). Finally, $`jj_{\mathrm{adiab}.}(r)`$, as given by Eq. (74). We define $`\omega _r^2(r)`$ $``$ $`C_r(r,j_{\mathrm{adiab}.}(r)){\displaystyle \frac{^2\widehat{H}_0}{r^2}}(r,j_{\mathrm{adiab}.}(r)),`$ (87) $`=`$ $`{\displaystyle \frac{1}{\nu ^2\widehat{H}_0^2(r,j_{\mathrm{adiab}.})}}{\displaystyle \frac{r^56r^4+3\nu r^3+20\nu r^230\nu ^2}{r^6(r^26\nu )}},`$ (88) $`B_r(r)`$ $``$ $`C_r(r,j_{\mathrm{adiab}.}(r)){\displaystyle \frac{^2\widehat{H}_0}{rj}}(r,j_{\mathrm{adiab}.}(r))\widehat{}_\phi (\widehat{\omega }_{\mathrm{adiab}.}(r)),`$ (89) $`=`$ $`{\displaystyle \frac{2j_{\mathrm{adiab}.}(r)}{\nu ^2\widehat{H}_0^2(r,j_{\mathrm{adiab}.})}}{\displaystyle \frac{(r^33r^2+5\nu )^2}{r^7(r^26\nu )}}\widehat{}_\phi (\widehat{\omega }_{\mathrm{adiab}.}(r)),`$ (90) (where the replacements $`jj_{\mathrm{adiab}.}(r)`$ are done after the partial differentiations). It is easily seen that the quantity $`^2\widehat{H}_0/rj`$ is negative, so that ($`\widehat{}_\phi `$ being also negative) the quantity $`B_r`$ given by Eq. (89) is positive. Combining Eqs. (85) and (86), we finally derive the following third order differential equation in $`r`$: $$\frac{d^3r}{d\widehat{t}^3}+\omega _r^2(r)\frac{dr}{d\widehat{t}}B_r(r).$$ (91) We shall often refer to Eq. (91) as the “linear $`\dot{r}`$-equation” because it was obtained by working linearly in the radial velocity $`\dot{r}=dr/d\widehat{t}`$. \[Note, however, that this is a third-order nonlinear differential equation in $`r`$.\] It is easily seen that the quantity $`\omega _r^2(r)`$ defines the square of the frequency of the radial oscillations. As seen in Eq. (87) it is proportional to the curvature of the effective radial potential $`H_0(r,j)`$ determining the radial motion. Above the LSO, i.e. when $`r>r_{\mathrm{LSO}}(\nu )`$, the radial potential has a minimum (defining the stable circular orbit with angular momentum $`j`$) and, therefore, $`\omega _r^2(r)`$ is positive. When $`r=r_{\mathrm{LSO}}(\nu )`$, the radial potential has an inflection point (see Eq. (17)), and, therefore, $`\omega _r^2(r)`$ vanishes. When $`r<r_{\mathrm{LSO}}(\nu )`$, the radial potential is concave, and $`\omega _r^2(r)`$ becomes negative. \[See, e.g., Fig. 1 of for a plot of the shape of the radial potential.\] Within the same approximation used above (i.e., essentially, neglecting terms which are fractionally of order $`p_r^2`$), we can finally write the angular frequency along our quasi-circular orbits as $$\frac{d\phi }{d\widehat{t}}\frac{\widehat{H}_0}{j}(r,j_{\mathrm{adiab}.}(r)).$$ (92) Note that $`\phi `$ is obtained from this equation by a quadrature, once the radial motion $`r(\widehat{t})`$ is known from the integration of Eq. (91). The conceptually interesting feature of the above “$`\dot{r}`$-linearized” approximation is the structure of Eq. (91). The previously considered “adiabatic” approximation corresponds to neglecting $`d^3r/d\widehat{t}^3`$ in Eq. (91). We now see that this is a good approximation only when the characteristic frequency of variation of the radial motion, defined, say, by $`\omega _{\mathrm{caract}.}^2(d^3r/d\widehat{t}^3)/(dr/d\widehat{t})`$ is much smaller than the frequency of radial oscillations $`\omega _r^2`$ (determined by the restoring radial force ensuring the existence of stable circular orbits). As $`\omega _r^2`$ tends to zero, before changing sign, at the LSO, it is clear that the adiabatic approximation must break down somewhat above the LSO. When it breaks down the “inertia term” $`d^3r/d\widehat{t}^3`$ in Eq. (91) becomes comparable to both the “restoring force” term $`\omega _r^2dr/d\widehat{t}`$ and the “driving force” $`B_r`$ coming from gravitational radiation damping. In Figs. 6 we compare the number of gravitational cycles and the radial velocity evaluated with the exact evolution and the $`\dot{r}`$-linearized equations. We start the evolution at $`r=15`$ and fix the initial values of $`dr/d\widehat{t}`$ and $`d^2r/d\widehat{t}^2`$ in the “adiabatic approximation” defined by neglecting in Eq. (91) the “inertia term” $`d^3r/d\widehat{t}^3`$ (and then by differentiating again the resulting approximate equation). Moreover, we normalize $`𝒩_{\mathrm{GW}}^{\mathrm{linear}}`$ to be zero at the $`r`$-LSO. We derive from the exact evolution $`𝒩_{\mathrm{GW}}^{\mathrm{exact}}(r_{\mathrm{LSO}})=0.04223`$. The main conclusion drawn from Figs. 6 is that the $`\dot{r}`$-linearized approximation is quite good both during the inspiral phase and, more importantly, during the transition to the plunge taking place near the LSO. This is interesting to know because it shows that the crucial physical effect that is lacking in the usually considered adiabatic approximation is the simple “inertia term” $`d^3r/d\widehat{t}^3`$ in Eq. (91). Note, however, that in order to add this inertia term it is necessary to have in hand the Hamiltonian describing at least the slightly non-circular orbits (the normalization of Eq. (91) crucially depends on the knowledge of $`\omega _r^2`$ which depends both on $`\widehat{H}/p_r^2`$ and on $`^2\widehat{H}/r^2`$). This being said, we do not, however, recommend to use in practice the $`\dot{r}`$-linearized approximation. Indeed, we think that the “exact” system (62)–(65) is a more accurate description of the evolution of the system because it keeps all the nonlinear effects in $`p_r^2`$. Numerically speaking, it is essentially as easy to integrate the “exact” system than its $`\dot{r}`$-linearized approximation, so that there would be anyway no practical advantage in downgrading the accuracy of the system (62)–(65). However, we shall see next that the $`\dot{r}`$-linearized system can be further used to lead to a simple analytical approach to the transition to the plunge in the case where $`\nu 1`$. ### C The universal $`\rho `$-equation Until now we have been considering the general case where the symmetric mass ratio $`\nu m_1m_2/(m_1+m_2)^2`$ can be of order of its maximum value $`\nu _{\mathrm{max}}=1/4`$. As is clear from the results above when $`4\nu `$ is of order unity the non-adiabatic aspects of radiation damping effects become important in an extended region of order $`\mathrm{\Delta }(R/GM)1`$, above the standard LSO. On the other hand, we expect that when $`4\nu 1`$ the transition between the adiabatic inspiral and the plunge will be sharply localized around the standard LSO, defined by Eq. (17). Indeed, when $`\nu `$ is a small parameter, the damping force $`\widehat{}_\phi `$, Eq. (67), being proportional to $`\nu `$, can be treated as a perturbatively small quantity in the evolution of the system. Consequently, the “driving force” term, $`B_r`$, in the $`\dot{r}`$-linearized equation (91) contains the small parameter $`\nu `$. It is then clear that all the time derivatives of $`r`$ (being driven by $`B_r`$) will tend to zero with $`\nu `$. If the coefficient $`\omega _r^2`$ in Eq. (91) never vanishes it is easy to see how one would satisfy Eq. (91) by solving for $`dr/d\widehat{t}`$, while considering $`d^3r/d\widehat{t}^3`$ as a fractionally small term (to be evaluated by further differentiating $`dr/d\widehat{t}B_r/\omega _r^2`$). In that case, one sees that $`dr/d\widehat{t}`$ would be $`𝒪(\nu )`$ (and $`d^3r/d\widehat{t}^3=𝒪(\nu ^3)`$) as $`\nu 0`$. However, the fact that $`\omega _r^2(r)`$ vanishes when $`r=r_{\mathrm{LSO}}(\nu )`$ shows that the way $`dr/d\widehat{t}`$ tends to zero with $`\nu `$, near the LSO, is more subtle. Having understood from this reasoning that, when $`\nu 0`$, the interesting transition effects take place very near the LSO, we now turn to a precise analysis of this transition. A first method for dealing (when $`\nu 0`$) with this transition would be (as just sketched) to continue working with the third-order equation (91), considered in the immediate neighbourhood of $`r=r_{\mathrm{LSO}}(\nu )`$. However, it is better (in order not to increase the differential order) to go back to the exact system (62)–(65) and to approximate it directly when $`\nu 0`$ and $`rr_{\mathrm{LSO}}(\nu )`$. Let us see the consequences of the evolution (62)–(65) when $`r`$ is very near $`r_{\mathrm{LSO}}(\nu )`$. To do this it is convenient to introduce some notation. Using, as we did in Sec. IV B, the fact that $`\widehat{H}`$ depends quadratically on $`p_r`$ and that $`p_r1`$, we define: $$C_r^{\mathrm{LSO}}(\nu )\left[\frac{1}{p_r}\frac{\widehat{H}}{p_r}(r,p_r,j)\right]_{p_r0}^{\mathrm{LSO}}.$$ (93) Note that $`C_r^{\mathrm{LSO}}`$ is a number, which depends on $`\nu `$ As we consider $`\nu 1`$, we could further take the limit $`\nu 0`$ in all the quantities which have a finite limit as $`\nu =0`$. However, in order not to unnecessarily loose accuracy we shall not do so. For instance we shall always consider that $`r_{\mathrm{LSO}}(\nu )`$ is computed for $`\nu 0`$, though we shall see later that the direct $`\nu `$-dependence in $`r_{\mathrm{LSO}}(\nu )`$ (which is $`𝒪(\nu )`$) is parametrically small compared to the width $`𝒪(\nu ^{2/5})`$ of the radial axis where the transition takes place.. In terms of the previous definition (82), one has simply $`C_r^{\mathrm{LSO}}(\nu )=C_r(r_{\mathrm{LSO}}(\nu ),j_{\mathrm{LSO}}(\nu ))`$. Explicitly, it reads $$C_r^{\mathrm{LSO}}(\nu )=\left[\frac{A^2(r)}{\nu \widehat{H}_0(r,j)\widehat{H}_{\mathrm{eff}}^0(r,j)(16\nu /r^2)}\right]_{\mathrm{LSO}}.$$ (94) In the $`\nu =0`$ limit this simplifies to $$C_r^{\mathrm{LSO}}(0)=\frac{\sqrt{2}}{3}.$$ (95) The point in having introduced the notation (93) is that Eq. (62) reads simply, when one is very near the LSO: $$p_r\frac{1}{C_r^{\mathrm{LSO}}}\frac{dr}{d\widehat{t}}.$$ (96) This allows us to recast Eq. (64) in the form (after neglecting fractional $`p_r^2`$ terms on the RHS) $$\frac{1}{C_r^{\mathrm{LSO}}}\frac{d^2r}{d\widehat{t}^2}\frac{\widehat{H}_0}{r}(r,j).$$ (97) Here, as above, $`\widehat{H}_0(r,j)\widehat{H}(r,p_r=0,p_\phi j)`$. Then we expand the RHS of the above equation around the LSO, i.e. we write $$r=r_{\mathrm{LSO}}(\nu )+\delta r,j=j_{\mathrm{LSO}}(\nu )+\delta j.$$ (98) Keeping the first nontrivial terms in the expansion in powers of $`\delta r`$ and $`\delta j`$ (and neglecting subleading terms, such as those of order $`𝒪(\delta r\delta j)`$, $`𝒪((\delta j)^2)`$ and $`𝒪((\delta r)^3)`$) one obtains $$\frac{\widehat{H}_0}{r}\frac{1}{2}\left(\frac{^3\widehat{H}_0}{r^3}\right)_{\mathrm{LSO}}(\delta r)^2+\left(\frac{^2\widehat{H}_0}{rj}\right)_{\mathrm{LSO}}(\delta j).$$ (99) Moreover, near the LSO we can write Eq. (65) as: $$\frac{d(\delta j)}{d\widehat{t}}=\frac{dj}{d\widehat{t}}\widehat{}_\phi (\widehat{\omega }_{\mathrm{LSO}}),\text{with}\widehat{\omega }_{\mathrm{LSO}}=\left(\frac{\widehat{H}_0}{j}\right)_{\mathrm{LSO}}.$$ (100) This yields $$\delta j\widehat{}_\phi (\widehat{\omega }_{\mathrm{LSO}})(\widehat{t}\widehat{t}_{\mathrm{LSO}}),$$ (101) where $`\widehat{t}_{\mathrm{LSO}}`$ is the time at which $`j(\widehat{t})=j_{\mathrm{LSO}}(\nu )`$. Let us also define $$A_r^{\mathrm{LSO}}C_r^{\mathrm{LSO}}\left(\frac{^3\widehat{H}_0}{r^3}\right)_{\mathrm{LSO}},B_r^{\mathrm{LSO}}C_r^{\mathrm{LSO}}\left(\frac{^2\widehat{H}_0}{rj}\right)_{\mathrm{LSO}}\widehat{}_\phi (\widehat{\omega }_{\mathrm{LSO}}).$$ (102) The quantity $`B_r^{\mathrm{LSO}}`$ is the LSO value of the quantity $`B_r(r)`$ introduced in Eq. (89) above. The explicit values of these quantities are $`A_r^{\mathrm{LSO}}(\nu )=\left[{\displaystyle \frac{(r^32r^2+2\nu )(210\nu j^260\nu r^2+60j^2r^212j^2r^3+6r^4)}{r^7(r^26\nu )(j^2+r^2)\nu ^2\widehat{H}_0^2(r,j)}}\right]_{\mathrm{LSO}},`$ (103) $`B_r^{\mathrm{LSO}}(\nu )=\left[{\displaystyle \frac{2j(r^32r^2+2\nu )(r^33r^2+5\nu )}{r^5(r^26\nu )(r^2+j^2)\nu ^2\widehat{H}_0^2(r,j)}}\right]_{\mathrm{LSO}}\widehat{}_\phi (\widehat{\omega }_{\mathrm{LSO}}).`$ (104) In the $`\nu =0`$ limit they simplify to $$A_r^{\mathrm{LSO}}(0)=\frac{1}{1296},B_r^{\mathrm{LSO}}(\nu )\stackrel{\nu 0}{=}\frac{1}{72\sqrt{3}}\nu \left[\frac{\widehat{}_\phi (\widehat{\omega }_{\mathrm{LSO}}(\nu );\nu )}{\nu }\right]_{\nu 0}=1.05210^4\nu .$$ (105) Finally, inserting Eq. (101) into Eq. (99), and replacing everything in Eq. (97) yields the simple equation $$\frac{d^2\delta r}{d\widehat{t}^2}+\frac{1}{2}A_r^{\mathrm{LSO}}(\delta r)^2=B_r^{\mathrm{LSO}}(\widehat{t}\widehat{t}_{\mathrm{LSO}}).$$ (106) This equation can be recast in a universal form by re-scaling the variables $`\delta r`$ and $`\delta \widehat{t}=(\widehat{t}\widehat{t}_{\mathrm{LSO}})`$. Indeed, posing $$\delta r=k_r\rho ,\widehat{t}\widehat{t}_{\mathrm{LSO}}=\delta \widehat{t}=k_t\tau ,$$ (107) with $$k_r(B_r^{\mathrm{LSO}})^{2/5}(A_r^{\mathrm{LSO}})^{3/5},k_t(A_r^{\mathrm{LSO}}B_r^{\mathrm{LSO}})^{1/5},$$ (108) it is straightforward to derive the following “universal $`\rho `$-equation” $$\frac{d^2\rho }{d\tau ^2}+\frac{1}{2}\rho ^2=\tau .$$ (109) The explicit values of the scaling coefficients $`k_r`$ and $`k_t`$ are easily derived from our previous results. Let us only quote explicitly their $`\nu =0`$ limit: $$k_r(0)=1.890\nu ^{2/5},k_t(0)=26.19\nu ^{1/5}.$$ (110) Note the interesting fractional scalings $`k_r\nu ^{2/5}`$, $`k_t\nu ^{1/5}`$. Let us also note the autonomous (time-independent) equation obtained by taking the time derivative of Eq. (109): $$\frac{d^3\rho }{d\tau ^3}+\rho \frac{d\rho }{d\tau }=1.$$ (111) Eq. (111) could have been directly derived by considering the $`\dot{r}`$-linearized Eq. (91) close to $`r=r_{\mathrm{LSO}}`$. There is, however, more information in Eq. (109) because its derivation showed that $`\tau =0`$ marks the moment where $`j(t)=j_{\mathrm{LSO}}(\nu )`$. The adiabatic approximation is recovered by neglecting in Eq. (109) the first term on the RHS. This gives $$\rho _{\mathrm{adiab}.}=\sqrt{2\tau },\left(\frac{d\rho }{d\tau }\right)_{\mathrm{adiab}.}=\frac{1}{\sqrt{2\tau }}=\frac{1}{\rho _{\mathrm{adiab}.}}.$$ (112) The universal $`\rho `$ and $`\dot{\rho }`$ curves and their adiabatic approximations are shown in Fig. 7. We have integrated Eq. (109) fixing the initial values (for large, negative $`\tau `$) of $`\rho `$ and $`d\rho /d\tau `$ in the adiabatic limit provided by Eq. (112). We see from Fig. 7 that the adiabatic approximation begins to be unacceptably bad when $`\tau 1`$. From the integration of Eq. (109) we get the important numerical values: $`\tau =0:\rho =0.8339,{\displaystyle \frac{d\rho }{d\tau }}=0.8233,`$ (113) $`\rho =0:\tau =0.8226,{\displaystyle \frac{d\rho }{d\tau }}=1.267.`$ (114) We recall that $`\tau =0`$ marks the moment where $`j(t)=j_{\mathrm{LSO}}(\nu )`$, while $`\rho =0`$ corresponds to $`r(t)=r_{\mathrm{LSO}}`$. The values given by Eqs. (113) and (114) can then be used to compute corresponding values of the physical quantities $`r`$, $`dr/d\widehat{t}`$ and $`j`$ by using the following parametric representations derived from our treatment above: $`r(\tau )=r_{\mathrm{LSO}}(\nu )+k_r\rho (\tau ),\widehat{t}(\tau )=\widehat{t}_{\mathrm{LSO}}+k_t\tau ,`$ (115) $`j(\tau )=j_{\mathrm{LSO}}(\nu )+\widehat{}_\phi (\widehat{\omega }_{\mathrm{LSO}})k_t\tau ,\left({\displaystyle \frac{dr}{d\widehat{t}}}\right)(\tau )={\displaystyle \frac{k_r}{k_t}}{\displaystyle \frac{d\rho }{d\tau }}.`$ (116) Correspondingly to these approximate results for $`r`$, $`\widehat{t}`$, $`j`$ and $`dr/d\widehat{t}`$, one can also write an approximate result for the angular frequency, namely $`\left({\displaystyle \frac{d\phi }{d\widehat{t}}}\right)(\tau )=\widehat{\omega }(\tau )`$ $`={\displaystyle \frac{\widehat{H}_0}{j}}(r(\tau ),j(\tau ))`$ (118) $`\widehat{\omega }_{\mathrm{LSO}}(\nu )+\left({\displaystyle \frac{^2\widehat{H}_0}{rj}}\right)_{\mathrm{LSO}}k_r\rho (\tau )+\left({\displaystyle \frac{^2\widehat{H}_0}{j^2}}\right)_{\mathrm{LSO}}\widehat{}_\phi ^{\mathrm{LSO}}k_t\tau .`$ In the approximation where we replace $`\nu `$ by zero in all quantities which have a finite limit when $`\nu 0`$, the above parametric results give the following explicit numerical links \[except for $`\widehat{t}_{\mathrm{LSO}}`$ which is an arbitrary integration constant\] $`r(\tau )=6+1.890\nu ^{2/5}\rho (\tau )+𝒪(\nu ),\widehat{t}(\tau )=\widehat{t}_{\mathrm{LSO}}+26.19\nu ^{1/5}\tau ,`$ (119) $`j(\tau )=\sqrt{12}0.3436\nu ^{4/5}\tau +𝒪(\nu ),\left({\displaystyle \frac{dr}{d\widehat{t}}}\right)(\tau )0.07216\nu ^{3/5}{\displaystyle \frac{d\rho }{d\tau }},`$ (120) $`\widehat{\omega }(\tau )={\displaystyle \frac{1}{6\sqrt{6}}}0.03214\nu ^{2/5}\rho (\tau )0.005062\nu ^{4/5}\tau +𝒪(\nu ).`$ (121) Note that these explicit results are less accurate than our previous implicit expressions Eqs. (115)–(116) (because of the $`𝒪(\nu )`$ error terms entailed by $`r_{\mathrm{LSO}}(\nu )=r_{\mathrm{LSO}}(0)+𝒪(\nu )`$, etc.). For consistency with the rest of the paper, we have used here (as in Eq. (105)) the $`\nu 0`$ limit of the value of $`\nu ^1_\phi ^{\mathrm{LSO}}`$ defined by the 2.5PN Padé estimate (67), namely $`\nu ^1_\phi ^{\mathrm{LSO}}0.01312`$. Note that a more accurate value of this quantity is, according to Poisson’s numerical results $`\nu ^1\widehat{}_\phi ^{\mathrm{LSO}}0.01376`$, which is $`5\%`$ larger (in modulus). Note the various scalings with $`\nu `$ implied (when considering a point in the transition region parametrized by some fixed numerical values of $`\rho `$ and $`\tau `$) by Eqs. (115)–(116): notably $`\delta r=𝒪(\nu ^{2/5})`$, $`\delta j=𝒪(\nu ^{4/5})`$ and $`p_r\dot{r}=𝒪(\nu ^{3/5})`$. We shall discuss below in more details some of these scalings. Fig. 7 vividly illustrates the fact (mentioned above) that the definition of “LSO-crossing” becomes ambiguous in presence of radiation damping. Indeed, for instance, the time where $`r=r_{\mathrm{LSO}}(\nu )`$ (“$`r`$-LSO”), i.e. the time where $`\rho =0`$, differs from the time where $`j=j_{\mathrm{LSO}}(\nu )`$ (“$`j`$-LSO”), i.e. the time where $`\tau =0`$ (see also Eq. (113)). An important issue is the domain of validity of the universal $`\rho `$-equation, i.e. the range of values of $`\nu `$ for which one can use Eqs. (115)–(116) to approximate the transition between inspiral and plunge. We have investigated this question numerically by comparing the radial velocity computed with the “exact” evolution (62)–(65), and with the $`\rho `$-equation (109). Let us define the practical limit of the domain of validity of the $`\rho `$-approximation by requiring that the fractional error in $`dr/d\widehat{t}`$ at the (say) $`r`$-LSO be 10%. We find that this limit is reached when $`\nu `$ gets as large as $$\nu _{\mathrm{max}}0.05.$$ (122) Therefore, the explicit expressions above can be used to estimate numerically the physical quantities in the transition region only for $`\nu \nu _{\mathrm{max}}`$. Note that the accuracy of the $`\rho `$-results above is, by construction, limited to some small neighbourhood of the LSO. They should not be used (even if $`\nu <\nu _{\mathrm{max}}`$) to estimate, for instance, the radial velocity at a radius which is significantly different from $`r_{\mathrm{LSO}}`$ (say at $`r=5`$ or $`r=7`$). To illustrate this we compare in Figs. 8 and 9 the radial velocity computed with the exact evolution, with that deduced from the $`\rho `$-equation. We examine three cases: $`\nu =\nu _{\mathrm{max}}=0.05`$, $`\nu =10^2`$ and $`\nu =10^4`$. Note that, though the accuracy of the approximation defined by the $`\rho `$-equation increases as $`\nu 0`$, its domain of validity actually shrinks as $`\nu `$ gets small. Indeed, if we keep $`\rho `$ finite we see that $`\delta r1.890\nu ^{2/5}\rho `$, tends to zero with $`\nu `$. Before discussing the scaling predictions made by the $`\rho `$-approximation, let us comment on the various possible definitions of “LSO crossing”. We recall that we define: (i) the “$`r`$-LSO” (by the requirement $`r(t)=r_{\mathrm{LSO}}(\nu )`$), (ii) the “$`j`$-LSO” ($`j(t)=j_{\mathrm{LSO}}(\nu )`$), and (iii) the “$`\omega `$-LSO” ($`\widehat{\omega }(t)=\widehat{\omega }_{\mathrm{LSO}}(\nu )`$). \[In addition, one can also define an “energy-LSO”, and a “naive” LSO such that $`R=6\mathrm{G}\mathrm{M}`$.\] We see from our results above that the $`r`$-LSO corresponds (in the $`\rho `$-approximation) to $`\rho =0`$, while the $`j`$-LSO corresponds to $`\tau =0`$, and the $`\omega `$-LSO to $`\rho +0.1575\nu ^{2/5}\tau =0`$. From these results and the results displayed in Fig. 7 and Eqs. (113) and (114), we have the following ordering between these LSO’s: $`\omega \text{-LSO}<r\text{-LSO}<j\text{-LSO}`$, where the order symbols refer to the location on the radial axis. We see also that when $`\nu ^{2/5}1`$ the $`\omega `$-LSO nearly coincides with the $`r`$-LSO. When discussing scaling relations it would be essentially equivalent to use any definition of LSO-crossing. For definiteness, and for consistency with the rest of this paper where we shall use it, we shall consider the $`\omega `$-LSO (because it is more invariantly defined than the $`r`$-LSO). To sufficient approximation for determining the leading scaling with $`\nu `$, we shall consider that the $`\omega `$-LSO corresponds to $`\rho 0`$. One of the most useful scaling law to consider is that concerning the radial momentum at the $`\omega `$-LSO. Combining Eqs. (96) and (116) we get: $$p_r=\frac{1}{C_r^{\mathrm{LSO}}}\frac{dr}{d\widehat{t}}=\frac{1}{C_r^{\mathrm{LSO}}}(A_r^{\mathrm{LSO}})^{2/5}(B_r^{\mathrm{LSO}})^{3/5}\frac{d\rho }{d\tau }.$$ (123) From Eq. (114) the value of $`d\rho /d\tau `$ at the $`\omega `$-LSO (i.e. $`\rho 0`$) is $`d\rho /d\tau 0.8233`$. Using also the numerical values (taken when $`\nu 0`$) of the coefficients entering Eq. (123), we get the predicted scaling $$(p_r)_{\omega \mathrm{LSO}}0.0844(4\nu )^{3/5}.$$ (124) In the left panel of Fig. 10 we compare the analytical scaling prediction, $`(p_r)_{\omega \mathrm{LSO}}(4\nu )^{3/5}`$, with the numerical results obtained by integrating the full evolution system (62)–(65) down to the $`\omega `$-LSO. We have also computed the best fits to the data using either a formula with one free parameter, of the type $`p_r=a(4\nu )^{3/5}`$ or with two free parameters, $`p_r=a(4\nu )^b`$. Note that the predicted scaling is a surprisingly good fit to the exact results, even for values of $`\nu `$ much larger than the domain of validity of the $`\rho `$-equation. In fact, it is numerically quite accurate even for $`\nu =1/4`$. \[In the one-parameter fit, note that the best-fit coefficient $`a=0.0750`$ is $`11\%`$ smaller than the calculated one, Eq. (124). This is because the best-fit one takes into account the values of $`p_r`$ for larger values of $`\nu `$ than the test-mass-limit result (124).\] Another useful scaling law concerns the number of orbits remaining “after LSO-crossing”. Let us define the number of orbits after LSO-crossing as $`\mathrm{\Delta }\phi /2\pi `$ where $`\mathrm{\Delta }\phi `$ is the difference in orbital phase between the “light-ring” $`r=r_{\mathrm{light}\mathrm{ring}}(\nu )`$ (obtained from Eq. (75)) and the $`\omega `$-LSO, $`\omega =\omega _{\mathrm{LSO}}(\nu )`$. This quantity cannot be really estimated within the $`\rho `$-approximation, because this approximation assumes that $`\delta r1`$. However, we can formally say that, within the $`\rho `$-approximation, we wish to consider the asymptotic limit where $`\rho `$ tends to $`\mathrm{}`$ proportionally to $`\nu ^{2/5}`$ (so that $`\delta r`$ is finite). The question is therefore: what is the asymptotic behaviour of the solution $`\rho =\rho (\tau )`$ of Eq. (109) when $`\rho \mathrm{}`$? It seems that in this limit the “source term” $`\tau `$ on the RHS of Eq. (109) is relatively negligible. Indeed, let us neglect it and solve the approximate equation $`\frac{d^2\rho }{d\tau ^2}+\frac{1}{2}\rho ^2=0`$. This equation describes the motion of a particle ($`\ddot{\rho }=V(\rho )/\rho `$) with potential energy $`V(\rho )=\rho ^3/6`$. This potential energy (which represents the effective radial potential near the inflection point corresponding to the LSO) is unboundedly negative when $`\rho \mathrm{}`$. Writing the conservation of “energy”, $`\frac{1}{2}\dot{\rho }^2+V(\rho )=`$ const, one finds that, as $`\rho \mathrm{}`$, the kinetic energy grows without bound and approximately satisfy $`\frac{1}{2}\dot{\rho }^2V(\rho )`$ whose solution is $$\rho =12(\tau _{\mathrm{}}\tau )^2$$ (125) for some constant $`\tau _{\mathrm{}}`$. We conclude that, as $`\rho \mathrm{}`$, the variable $`\tau `$ tends to a finite limit $`\tau _{\mathrm{}}`$. \[We find $`\tau _{\mathrm{}}3.9`$. The corresponding curve is shown in the left panel of Fig. 7.\] Therefore, from Eq. (125), the total time elapsed after the LSO, $`\widehat{t}_{\mathrm{}}\widehat{t}_{\mathrm{LSO}}`$, scales like $`\nu ^{1/5}`$. Correspondingly, within the $`\rho `$-approximation, the leading approximation to the orbital phase (obtained by integrating the zeroth order term in Eq. (125)) reads $$\frac{\mathrm{\Delta }\phi }{2\pi }=_{\widehat{t}_{\mathrm{LSO}}}^{\widehat{t}_{\mathrm{}}}\frac{\widehat{\omega }}{2\pi }𝑑\widehat{t}\frac{\widehat{\omega }_{\mathrm{LSO}}}{2\pi }(\widehat{t}_{\mathrm{}}\widehat{t}_{\mathrm{LSO}})=\frac{\widehat{\omega }_{\mathrm{LSO}}}{2\pi }(A_r^{\mathrm{LSO}}B_r^{\mathrm{LSO}})^{1/5}\tau _{\mathrm{}}.$$ (126) As $`\widehat{\omega }_{\mathrm{LSO}}`$ admits a finite limit as $`\nu 0`$, we expect from Eq. (126) the scaling law $$\frac{\mathrm{\Delta }\phi }{2\pi }(4\nu )^{1/5}.$$ (127) This prediction is compared in Fig. 10 with the numerical results obtained by integrating the full system (62)–(65). As expected from the necessity to inconsistently consider parametrically large values of $`\rho \nu ^{2/5}`$, this prediction is less accurate than that obtained for the radial momentum at the $`\omega `$-LSO. We have indicated both the best fit to a formula of the type $`\mathrm{\Delta }\phi /2\pi =a(4\nu )^{1/5}`$, and the best fit to $`\mathrm{\Delta }\phi /2\pi =a(4\nu )^b`$. Note that, both fits have been evaluated including values of $`\nu `$ only up to $`\nu _{\mathrm{max}}=0.05`$. Indeed, as discussed above, beyond this value the fractional error in the radial velocity at the $`r_{\mathrm{LSO}}`$ is $`10\%`$. Some comments are in order concerning these results. First, we note that although $`N_{\mathrm{LSO}}^{\mathrm{after}}=\mathrm{\Delta }\phi /2\pi `$ tends to infinity when $`\nu 0`$, it does so very slowly so that the total number of orbits after the LSO is always quite small compared to the number of orbits “just before and around the LSO”. Let us define the latter number as $`N_{\mathrm{LSO}}^{\mathrm{around}}f_{\mathrm{orbit}.}^2/\dot{f}_{\mathrm{orbit}.}=\frac{1}{2}f_{\mathrm{GW}}^2/\dot{f}_{\mathrm{GW}}`$ where $`f_{\mathrm{orbit}.}=\frac{1}{2}f_{\mathrm{GW}}=\omega /2\pi `$ denotes the orbital frequency, and $`\dot{f}_{\mathrm{orbit}.}`$ the time derivative of the orbital frequency caused by GW damping. In the adiabatic approximation, combined with a Newtonian approximation for both the orbital energy and the GW flux, this number reads (see, e.g., ) $$N_{\mathrm{LSO}}^{\mathrm{around}}\frac{2.924}{4\nu }.$$ (128) The ratio $`N_{\mathrm{LSO}}^{\mathrm{after}}/N_{\mathrm{LSO}}^{\mathrm{around}}0.3446(4\nu )^{4/5}`$ (derived using the result of the fit, i.e. $`N_{\mathrm{LSO}}^{\mathrm{after}}=1.0075(4\nu )^{1/5}`$) is therefore parametrically small as $`\nu 0`$. This suggests that, when $`4\nu 1`$, the existence of even a formally parametrically large $`(\nu ^{1/5})`$ absolute number of cycles left after the LSO will have only a fractionally negligible effect on the extraction of a GW signal from the noise by means of relativistic filters built on the adiabatic approximation, and terminated at the LSO , . On the other hand, when $`4\nu 1`$ the ratio $`N_{\mathrm{LSO}}^{\mathrm{after}}/N_{\mathrm{LSO}}^{\mathrm{around}}`$ is not very small. In particular, when $`\nu =1/4`$ the number of orbits after the $`\omega `$-LSO is equal to $`N_{\mathrm{LSO}}^{\mathrm{after}}(\nu =1/4)=0.6024`$ (computed from the exact evolution), while $`N_{\mathrm{LSO}}^{\mathrm{around}}(\nu =1/4)=2.924`$. The ratio between the two is $`N_{\mathrm{LSO}}^{\mathrm{after}}/N_{\mathrm{LSO}}^{\mathrm{around}}=0.2060`$. As recently emphasized in Ref. , the fact that $`N_{\mathrm{LSO}}^{\mathrm{around}}`$ is not large means that the filtering of such a signal out of the noise is a delicate matter which sensitively depends on the modeling of the phase evolution near the LSO, and on the modeling of what happens to the signal after LSO crossing. In Ref. it was assumed that the signal is abruptly terminated at the LSO. In a later section we shall use the tools introduced here to go beyond such an approximation and study the part of the waveform which is emitted after LSO crossing. ## V Initial data for numerical relativity One of the main aims of this paper is to use the improved approach to the transition from the inspiral to the plunge introduced above to compute initial dynamical data (i.e. initial positions and momenta) for binary black holes that have just started their plunge motion. Ideally, we wish to give dynamical data for two black holes $`(𝒒_1,𝒒_2,𝒑_1,𝒑_2)`$ such that the coordinate distance $`|𝒒_1𝒒_2|`$ is: (i) large enough that one can trust the re-summed non-perturbative technique allowing one to compute these data; (ii) large enough to allow one to hope to complete the present work by constructing the initial gravitational data $`(g_{ij}(x),K_{ij}(x))`$ determined (in principle) by $`(𝒒_a,𝒑_a)`$; and, finally (iii) small enough to leave only less than one orbit (at least when $`\nu 1/4`$) to evolve by means of a full 3D numerical relativity code. We think that the point (i) is satisfied if we use the Padé-type plus effective-one-body methods we have combined above, and if we stop the evolution of quasi-circular orbits anywhere around the LSO. We shall leave the point (ii), i.e. the important task of completing the present work by constructing gravitational data to future work. However, in preparation for this task we shall show how one can compute the dynamical data $`(𝒒_a,𝒑_a)`$ in the convenient ADM coordinates. Indeed, the coordinate conditions introduced by Arnowitt, Deser and Misner have the double advantage: (a) to be linked to the $`3+1`$ formulation which is used in numerical relativity, and (b) to be linked to explicit, high-order post-Newtonian calculations . Concerning the point (iii), the work above shows that if we stop the inspiral $`+`$ plunge evolution at the (invariantly defined) $`\omega `$-LSO (i.e. when $`d\phi /dt=\omega _{\mathrm{LSO}}(\nu )`$) there indeed remains (when $`4\nu 1`$) less than one orbit to go before reaching the light-ring (see next section for a discussion of the importance of the light-ring). Note that there is nothing sacred about giving data precisely at the $`\omega `$-LSO. Because of the points (i) and (ii) above we wish to stay “as high as possible”. Because of point (iii) we must, however, be just after LSO crossing. As was already discussed, there are several possible definitions of “LSO crossing”. The $`\omega `$-LSO is the innermost LSO (see below) and is therefore a convenient choice (however, there would be nothing wrong in giving data at a slightly different place; in fact we recommend to do it to check the robustness of the numerical spacetimes evolved from our data). As just recalled we wish to (numerically) compute complete dynamical data at the $`\omega `$-LSO, and in ADM coordinates. The evolution system (62)–(65) given above allows one to compute dynamical data $`(r,\phi ,p_r,p_\phi )`$ for the relative motion described in (reduced) effective coordinates (i.e. the coordinates used in the effective-one-body description). In Ref. we have shown how to map the ADM positions and momenta $`(q^{\mathrm{ADM}},p^{\mathrm{ADM}})`$ onto the effective positions and momenta $`(q,p)`$ by means of a generating function $`G(q^{\mathrm{ADM}},p)`$. Let us first recall, in order to avoid any confusion, the trivial transformations linking Cartesian-like to polar-like coordinates, as well as those linking the original to the scaled coordinates. We recall that we work in the center of mass frame and that we consider planar motion in the equatorial plane $`\theta =\pi /2`$: $$Q^i=q_1^iq_2^i,P_i=p_{1i}=p_{2i},$$ (129) $$P_R=n^iP_i,P_\phi =Q^xP_yQ^yP_x,$$ (130) $$q^i=\frac{Q^i}{GM},p_i=\frac{P_i}{\mu },$$ (131) $$p_r=\frac{P_R}{\mu }=n^ip_i,p_\phi =\frac{P_\phi }{\mu GM}=q^xp_yq^yp_x.$$ (132) Here $`n^i=Q^i/R=q^i/r`$ is the radial unit vector ($`R=|𝑸|`$, $`r=|𝒒|`$). We have also $`Q^x=R\mathrm{cos}\phi `$, $`Q^y=R\mathrm{sin}\phi `$, $`q^x=r\mathrm{cos}\phi `$, $`q^y=r\mathrm{sin}\phi `$. The relations above hold both in effective coordinates (denoted by $`(q^i,p_i)`$ without extra labels) and in ADM coordinates $`(q_{\mathrm{ADM}}^i,p_i^{\mathrm{ADM}})`$. The link between $`(q^i,p_i)`$ and $`(q_{\mathrm{ADM}}^i,p_i^{\mathrm{ADM}})`$ is defined by a generating function $`G(q_{\mathrm{ADM}}^i,p_i)`$ and reads $`q^i=q_{\mathrm{ADM}}^i+{\displaystyle \frac{G(q^{\mathrm{ADM}},p)}{p_i}},`$ (133) $`p_i^{\mathrm{ADM}}=p_i+{\displaystyle \frac{G(q^{\mathrm{ADM}},p)}{q_{\mathrm{ADM}}^i}}.`$ (134) The generating function $`G`$ has been derived up to 2PN order in (see Ref. for the determination of $`G`$ at the 3PN level) $$G(q^{\mathrm{ADM}},p)=\frac{1}{c^2}G_{1\mathrm{P}\mathrm{N}}(q^{\mathrm{ADM}},p)+\frac{1}{c^4}G_{2\mathrm{P}\mathrm{N}}(q^{\mathrm{ADM}},p).$$ (135) The partial derivatives needed in Eqs. (133), (134) read $`{\displaystyle \frac{G_{1\mathrm{P}\mathrm{N}}(q,p)}{q^i}}`$ $`=`$ $`p_i\left[{\displaystyle \frac{\nu }{2}}𝒑^2+\left(1+{\displaystyle \frac{\nu }{2}}\right){\displaystyle \frac{1}{q}}\right]q_i(𝒒𝒑)\left(1+{\displaystyle \frac{\nu }{2}}\right){\displaystyle \frac{1}{q^3}},`$ (136) $`{\displaystyle \frac{G_{1\mathrm{P}\mathrm{N}}(q,p)}{p_i}}`$ $`=`$ $`q^i\left[{\displaystyle \frac{\nu }{2}}𝒑^2+\left(1+{\displaystyle \frac{\nu }{2}}\right){\displaystyle \frac{1}{q}}\right]p^i(𝒒𝒑)\nu ,`$ (137) $`{\displaystyle \frac{G_{2\mathrm{P}\mathrm{N}}(q,p)}{q^i}}`$ $`=`$ $`p_i[{\displaystyle \frac{1}{8}}\nu (1+3\nu )𝒑^4+{\displaystyle \frac{\nu }{8}}(25\nu ){\displaystyle \frac{𝒑^2}{q}}+{\displaystyle \frac{3}{8}}\nu (8+3\nu ){\displaystyle \frac{(𝒒𝒑)^2}{q^3}}`$ (140) $`+{\displaystyle \frac{1}{4}}(17\nu +\nu ^2){\displaystyle \frac{1}{q^2}}]+q_i(𝒒𝒑)[{\displaystyle \frac{3}{8}}\nu (8+3\nu ){\displaystyle \frac{(𝒒𝒑)^2}{q^5}}`$ $`{\displaystyle \frac{\nu }{8}}(25\nu ){\displaystyle \frac{𝒑^2}{q^3}}{\displaystyle \frac{1}{2}}(17\nu +\nu ^2){\displaystyle \frac{1}{q^4}}],`$ $`{\displaystyle \frac{G_{2\mathrm{P}\mathrm{N}}(q,p)}{p_i}}`$ $`=`$ $`q^i[{\displaystyle \frac{1}{8}}\nu (1+3\nu )𝒑^4+{\displaystyle \frac{\nu }{8}}(25\nu ){\displaystyle \frac{𝒑^2}{q}}+{\displaystyle \frac{3}{8}}\nu (8+3\nu ){\displaystyle \frac{(𝒒𝒑)^2}{q^3}}`$ (142) $`+{\displaystyle \frac{1}{4}}(17\nu +\nu ^2){\displaystyle \frac{1}{q^2}}]+p^i(𝒒𝒑)[{\displaystyle \frac{\nu }{2}}(1+3\nu )𝒑^2+{\displaystyle \frac{\nu }{4}}(25\nu ){\displaystyle \frac{1}{q}}].`$ Given $`q^i`$ and $`p_i`$, we use first Eq. (133), and the values of the partial derivatives (136)–(142), to solve numerically for $`q_{\mathrm{ADM}}^i`$. Then we use Eq. (134) to compute $`p_i^{\mathrm{ADM}}`$: The initial data we start with are the results of the numerical integration of the system (62)–(65), i.e. the values of $`r`$, $`\phi `$, $`p_r`$ and $`p_\phi `$ at some time in the evolution (which we choose to be the time when $`\omega (t)=\omega _{\mathrm{LSO}}(\nu )`$). Actually, the value of $`\phi `$ is without significance and we renormalize it to the convenient value $`\phi _{\mathrm{new}}=0`$ so that we work with Cartesian-like data of the simple form (remember that we work in the $`xy`$ plane, $`q^z=0=p_z`$, and that we simplify the writing by denoting $`q_iq^i`$ when working in Cartesian-like coordinates) $$q_x=r,q_y=0,p_x=p_r,p_y=\frac{p_\phi }{r}.$$ (143) When solving, as indicated above, Eqs. (133), (134) to derive $`q_x^{\mathrm{ADM}}`$, $`q_y^{\mathrm{ADM}}`$ and $`p_x^{\mathrm{ADM}}`$, $`p_y^{\mathrm{ADM}}`$, we get these quantities in a not optimally oriented coordinate system (i.e. though we started with $`q_y=0`$, we end up with $`q_y^{\mathrm{ADM}}0`$ because there is a rotation between the two coordinate systems). As the global orientation is of no physical significance, it is convenient to turn the ADM coordinate system by an angle $`\alpha `$ so that $`\phi _{\mathrm{new}}^{\mathrm{ADM}}=\phi _{\mathrm{old}}^{\mathrm{ADM}}\alpha =0`$. In other words, after this rotation one has, as in Eq. (143) above, $$q_x^{\mathrm{ADM}\mathrm{new}}=r^{\mathrm{ADM}},q_y^{\mathrm{ADM}\mathrm{new}}=0,p_x^{\mathrm{ADM}\mathrm{new}}=p_r^{\mathrm{ADM}},p_y^{\mathrm{ADM}\mathrm{new}}=\frac{p_\phi ^{\mathrm{ADM}}}{r^{\mathrm{ADM}}}.$$ (144) The angle of rotation $`\alpha `$ is determined by $$\mathrm{tan}\alpha =\frac{q_y^{\mathrm{ADM}\mathrm{old}}}{q_x^{\mathrm{ADM}\mathrm{old}}},$$ (145) while the more invariant quantities $`r^{\mathrm{ADM}}`$ and $`p_r^{\mathrm{ADM}}`$ are given by $$r^{\mathrm{ADM}}\sqrt{(q_{x\mathrm{old}}^{\mathrm{ADM}})^2+(q_{y\mathrm{old}}^{\mathrm{ADM}})^2},p_r^{\mathrm{ADM}}\frac{1}{r^{\mathrm{ADM}}}(q_{x\mathrm{old}}^{\mathrm{ADM}}p_{x\mathrm{old}}^{\mathrm{ADM}}+q_{y\mathrm{old}}^{\mathrm{ADM}}p_{y\mathrm{old}}^{\mathrm{ADM}}).$$ (146) Note that (because of the rotational invariance of $`G`$) all the angular momenta coincide: $$p_\phi =p_\phi ^{\mathrm{ADM}}=q_xp_yq_yp_x=q_{x\mathrm{old}}^{\mathrm{ADM}}p_{y\mathrm{old}}^{\mathrm{ADM}}q_{y\mathrm{old}}^{\mathrm{ADM}}p_{x\mathrm{old}}^{\mathrm{ADM}}=q_{x\mathrm{new}}^{\mathrm{ADM}}p_{y\mathrm{new}}^{\mathrm{ADM}}q_{y\mathrm{new}}^{\mathrm{ADM}}p_{x\mathrm{new}}^{\mathrm{ADM}}.$$ (147) This relation is a useful check on the numerical precision of the solution of Eqs. (133), (134). In Tab. I we give initial data in ADM coordinates at the $`\omega `$-LSO for five values of the parameter $`\nu `$. We give the more invariant quantities corresponding to the “new” ADM coordinate system Eq. (144). The quantity $`p_t^{\mathrm{ADM}}`$ denotes the “transverse” momentum, i.e. simply $`p_t^{\mathrm{ADM}}p_\phi ^{\mathrm{ADM}}/r^{\mathrm{ADM}}p_{y\mathrm{new}}^{\mathrm{ADM}}`$. For completeness, we give also the value of the angle $`\alpha `$, Eq. (145). So far all the results we have discussed considered the evolution system (62)–(65) as the “exact” description of the transition through the LSO. However, as discussed in Sec. III this system is more like a convenient fiducial system within a class of systems obtained by shifting (by $`𝒪(v^5/c^5)`$ terms) the coordinate system. To test the robustness of our predictions for physical quantities at the LSO we shall now compare the results of the fiducial system (62)–(65) with the results obtained by the more general system (25)–(28), with a radial force $`_R`$ given (in terms of $`_\phi `$) by Eq. (41). For simplicity, we consider only the (most crucial) equal-mass case, $`\nu =1/4`$. We find that our fiducial system (with $`_R=0`$) yields the following numerical values at the $`\omega `$-LSO (when starting with an orbital phase $`\phi =0`$ at $`r=15`$) $`r=5.639,p_r=0.07432,\dot{r}=0.03563,`$ (148) $`\phi =82.72,j=3.312,{\displaystyle \frac{_{\mathrm{real}}^{\mathrm{NR}}}{M}}=0.01640.`$ (149) On the other hand, the system including the non-zero radial force (41) yields at the $`\omega `$-LSO (still starting with an orbital phase $`\phi =0`$ at $`r=15`$) $`r=5.638,p_r=0.07388,\dot{r}=0.03542,`$ (150) $`\phi =82.77,j=3.311,{\displaystyle \frac{_{\mathrm{real}}^{\mathrm{NR}}}{M}}=0.01643.`$ (151) As we see the differences in the numerical results are quite small. For instance, the fractional change in the (crucial) radial momentum is less than $`6\times 10^3`$. We note also that the dephasing at the LSO is only 0.05 radians. This analysis indicates that the results based on our fiducial system are quite robust, mainly because our basic assumption of “quasi-circularity” ($`\dot{R}R\dot{\phi }`$) is well satisfied during the transition to the plunge. ## VI Gravitational wave-forms from inspiral to ring-down In this section, we provide, for data analysis purposes, an estimate of the complete waveform emitted by the coalescence of two black holes (with negligible spins). This estimate will be less accurate than our results above because we shall extend the integration of our basic system (62)–(65) beyond its range of validity. We think, however, that even a rough estimate of the complete waveform (exhibiting the way the inspiral waveform smoothly transforms itself in a “plunge waveform” and then into a “merger plus ring-down” waveform) is a very valuable information for designing and testing effectual gravitational wave templates. \[See, in particular, the recent work which emphasizes the importance of the details of the transition to the plunge for the construction of faithful GW templates for massive binaries.\] Our (rough) assumptions in this section will be the following: (i) we use the basic evolution system (62)–(65) to describe the dynamics of the binary system from deep into the inspiral phase (say $`r15`$) down to the “light-ring” $`r=r_{\mathrm{light}\mathrm{ring}}(\nu )3`$; (ii) we estimate the waveform emitted during the inspiral and the plunge by means of the usual “restricted waveform” approximation $$\widehat{t}\widehat{t}_{\mathrm{end}}:h_{\mathrm{inspiral}}(\widehat{t})=𝒞v_\omega ^2(\widehat{t})\mathrm{cos}(\varphi _{\mathrm{GW}}(\widehat{t})),v_\omega \left(\frac{d\phi }{d\widehat{t}}\right)^{\frac{1}{3}},\varphi _{\mathrm{GW}}2\phi ,$$ (152) and (iii) we estimate the waveform emitted during the coalescence and ring-down by matching, at a time $`\widehat{t}=\widehat{t}_{\mathrm{end}}`$ where the light-ring is crossed, the inspiral $`+`$ plunge waveform (152) to the least-damped quasi-normal mode of a Kerr black hole with mass and spin equal to the total energy and angular momentum of the plunging binary (at $`\widehat{t}=\widehat{t}_{\mathrm{end}}`$): $$\widehat{t}\widehat{t}_{\mathrm{end}}:h_{\mathrm{merger}}(\widehat{t})=𝒜e^{(\widehat{t}\widehat{t}_{\mathrm{end}})/\tau }\mathrm{cos}(\omega _{\mathrm{qnm}}(\widehat{t}\widehat{t}_{\mathrm{end}})+).$$ (153) For convenience, we shall normalize the waveform by taking $`𝒞=1`$ in Eq. (152). The amplitude $`𝒜`$ and the phase $``$ of the merger waveform (153) are then determined by requiring the continuity of $`h(\widehat{t})`$ and $`dh/d\widehat{t}`$ at the matching point $`\widehat{t}=\widehat{t}_{\mathrm{end}}`$. Before giving technical details let us comment on our assumptions (i)–(iii). First, we recall that Fig. 1 had shown that the quasi-circularity condition $`p_r^2/B(r)p_\phi ^2/r^2`$ (which is the basic condition determining the validity of our evolution system) was satisfied with good accuracy during the inspiral and the beginning of the plunge, and was still satisfied, though with less accuracy ($`p_r^2/B0.3p_\phi ^2/r^2`$ in the worst case $`\nu =1/4`$) down to the light-ring $`r3`$. In other words, our work is showing that the so called “plunge” following the inspiral phase is better thought of as being still a quasi-circular inspiral motion, even down to the light-ring. We therefore expect that the usual restricted waveform (152) (valid for circular motion) will be an acceptable description of the GW emission during the plunge. Note that we consider that the description of the amplitude of the gravitational wave in terms of $`v_\omega ^2\dot{\phi }^{2/3}`$, being simpler and more invariant, has a better chance of being correct than a description in terms of some other Newtonian-like approximation to the “squared velocity” such as $`(r\dot{\phi })^2`$ or $`1/r`$. Some evidence for this faith is given by the fact that the GW flux is surprisingly well approximated (within 10% down to the LSO) by the usual “quadrupole formula” if the velocity used to define the quadrupole formula is the invariant $`v_\omega =\dot{\phi }^{1/3}`$ (see, e.g., Fig. 3 of ). Concerning the choice of the light-ring for shifting the description between a (quasi-circular) binary motion and a deformed Kerr black hole, our motivation is twofold. First, in the test-mass limit, $`\nu 1`$, it has been realized long ago, in the first work which found the existence of a merger signal of the type (153) following a plunge event, that the basic physical reason underlying the presence of a “universal” merger signal was that when a test particle falls below $`R3GM`$, the GW it generates is strongly filtered by the potential barrier, centered around $`R3GM`$, describing the radial propagation of gravitational waves. It was then realized that the peaking of the potential barrier around $`R3GM`$ is itself linked to the presence of an unstable “light storage ring” (i.e. an unstable circular orbit for massless particles) precisely at $`R=R_{\mathrm{light}\mathrm{right}}=3GM`$. A second argument (applying now in the equal-mass case, $`\nu =1/4`$) indicating that $`r_{\mathrm{light}\mathrm{right}}(1/4)2.84563`$ is an acceptable divide between the two-body and the perturbed-black-hole descriptions comes from the works on the, so called, “close limit approximation” . Indeed, recent work (see the review ) suggests a matching between the two-body and the perturbed-black-hole descriptions when the distance modulus $`\mu _02`$. Using the formulas of Ref. one finds that $`\mu _02`$ corresponds to a coordinate distance in isotropic coordinates of $`r^{\mathrm{iso}}\sqrt{2}xGM`$. This corresponds to a Schwarzschild-like radial distance $`Rr(1+GM/2r)^22.59GM`$ which is not very far from $`R_{\mathrm{light}\mathrm{ring}}(1/4)2.84GM`$. In keeping with our prescription of setting the divide between a binary-black-hole description and a perturbed-single-black-hole one, at the time $`\widehat{t}_{\mathrm{end}}`$, when $`rr_{\mathrm{light}\mathrm{ring}}(\nu )`$, it is natural to assume that the final hole formed by the merger is a Kerr hole with mass $`M_{\mathrm{BH}}`$ and angular momentum $`𝒥_{\mathrm{BH}}`$ given by: $$\frac{M_{\mathrm{BH}}}{\mu }\widehat{H}_{\mathrm{end}}=\frac{1}{\nu }\sqrt{1+2\nu (\widehat{H}_{\mathrm{eff}}^{\mathrm{end}}1)},j_{\mathrm{end}}\frac{𝒥_{\mathrm{BH}}}{\mu GM},$$ (154) while the dimensionless rotation parameter $`\widehat{a}`$ is: $$\widehat{a}_{\mathrm{BH}}\frac{𝒥_{\mathrm{BH}}}{GM_{\mathrm{BH}}^2}=\frac{\nu j_{\mathrm{end}}}{1+2\nu (\widehat{H}_{\mathrm{eff}}^{\mathrm{end}}1)}.$$ (155) As the system reaches the stationary Kerr state, the non-linear dynamics of the merger becomes more and more describable in terms of oscillations of the black hole quasi-normal modes . During this phase, often called the ring-down phase, the gravitational signal will be a superposition of exponentially damped sinusoids. The gravitational waveform will be dominated by the $`l=2,m=2`$ quasi-normal mode, which is the most slowly damped mode. As a rough approximation we assume that the full merger $`+`$ ring-down signal (starting when the light-ring is reached) can be represented in terms of this least damped quasi-normal mode. If $`\omega _{qnm}`$ denotes the circular frequency of this mode, and $`\tau `$ its damping time, this leads to the simple description (153). The quantities $`(\omega _{qnm},\tau )`$ are functions of $`(M_{\mathrm{BH}},\widehat{a}_{\mathrm{BH}})`$ which have been investigated numerically , . Using analytic fits the following expressions for the frequency and the decay time of the quasi-normal modes were obtained $`M_{\mathrm{BH}}\omega _{\mathrm{qnm}}=\left[10.63(1\widehat{a})^{3/10}\right]f_f(\widehat{a}),`$ (156) $`\tau \omega _{\mathrm{qnm}}=4\left[1\widehat{a}\right]^{9/20}f_Q(\widehat{a}),`$ (157) where $`f_f(\widehat{a})`$ and $`f_Q(\widehat{a})`$ are correction factors provided by Tab. 2 of . Note that $`f_f=0.9587`$ and $`f_Q=1.0501`$ for $`\widehat{a}=10^4`$. We have numerically studied only the equal-mass case $`\nu =1/4`$. We have chosen the matching point $`\widehat{t}_{\mathrm{end}}`$ such that $`r(\widehat{t}_{\mathrm{end}})=r_{\mathrm{light}\mathrm{ring}}(1/4)=2.84563`$. With this value of $`\widehat{t}_{\mathrm{end}}`$ we obtain the following values for the characteristics of the formed black hole: $`\widehat{a}_{\mathrm{BH}}=0.7952,E_{\mathrm{BH}}=0.9761M,`$ (158) $`M\omega _{\mathrm{qnm}}=0.5976,M/\tau =0.07795.`$ (159) Note the numerical value of the quasi-normal mode frequency $$f_{\mathrm{qnm}}\frac{\omega _{\mathrm{qnm}}}{2\pi }=1885\left(\frac{10M_{}}{M_{\mathrm{BH}}}\right)\mathrm{Hz}.$$ (160) Our results for the waveform are shown in Figs. 11 and 12. In Fig. 11 we compare the inspiral $`+`$ plunge waveform (152) (terminated at the light-ring) to the usually considered adiabatic waveform (terminated at the “adiabatic LSO”). As already discussed in Section IV, by “adiabatic waveform” we mean a restricted waveform (152) (with $`𝒞=1`$) in which $`\phi (\widehat{t})=\phi _{\mathrm{adiab}.}(\widehat{t})`$ is defined by integrating the two equations (79) and (80). This Figure shows that there is a significant dephasing of the adiabatic waveform with respect to the (more) exact one already before the LSO. Moreover, the real inspiral signal continues to increase and oscillate for $`2.35`$ cycles after the adiabatic LSO. In Fig. 12 we plot our estimate of the complete waveform: inspiral and plunge (solid line) followed by merger and ring-down (dashed line). We also indicated the locations of several possible definitions of LSO crossing (see Section IV above). In addition to the definitions mentioned above we also included a “naive LSO” (defined simply by $`r_{\mathrm{LSO}}^{\mathrm{naive}}6`$ as in the Schwarzschild geometry) and an energy-LSO (such that $`_{\mathrm{real}}(t)=_{\mathrm{real}}^{\mathrm{LSO}}(\nu )`$). The corresponding numerical values of the reduced radial coordinate $`r`$ are: $`r_{\mathrm{j}\mathrm{LS0}}=6.631,r_{\mathrm{LSO}}^{\mathrm{naive}}=6.000,r_{\mathrm{LS0}}=6.534,`$ (161) $`r_{\mathrm{r}\mathrm{LS0}}=5.718,r_{\omega \mathrm{LS0}}=5.639.`$ (162) As mentioned above, the fact that the various definitions of the LSO differ significantly is due to the fact that when $`\nu =1/4`$ the GW damping effects are rather large and blur the transition to the plunge. Note that the number of GW cycles left after the (exact) $`\omega `$-LSO (and until the light-ring) is $`N_{\mathrm{GW}}^{\mathrm{after}}=2N_{\mathrm{orbit}}^{\mathrm{after}}=1.2048`$ (for $`\nu =1/4`$). As said above, this is smaller than the (physically less relevant) number of cycles left after the adiabatic LSO (where $`\omega 0.80\omega _{\mathrm{LSO}}`$), which is $`2.35`$. Even if our estimate of the waveform is admittedly rough, we think that it can play an important role for defining better filters for the search of signals in LIGO and VIRGO. In particular, two features of this waveform are striking: (i) the ‘plunge’ part of the waveform looks like a continuation of the inspiral part (this is because the orbital motion remains in fact quasi-circular), and (ii) the adiabatic waveform gets significantly out of phase with the exact waveform before crossing the LSO. We shall come back in future work to the consequences of these results for data analysis, and see how they can be used to improve upon the state-of-the-art filters constructed in Ref. ,. ## VII Discussion In this paper we have extended a methodology introduced in previous papers , , and applied it to the study of the transition from inspiral to plunge in coalescing binary black holes with comparable masses, moving on quasi-circular orbits. Our philosophy is that it is possible to use suitably re-summed versions of post-Newtonian results to write an explicit (analytical) system of ordinary differential equations describing the transition to the plunge. Our explicit proposal is the evolution system (62)–(65) obtained by combining the results of for the re-summation of the gravitational wave damping, and the results of for the re-summation of the conservative part of the dynamics of comparable-mass binaries. The basic reason why we think the simple evolution system (62)–(65) can accurately describe the transition to the plunge is that we have consistently checked that most of the “plunge” motion (at least down to $`R3GM`$) is in fact very much like a quasi-circular inspiral motion (with $`\dot{R}^2(R\dot{\phi })^2`$). In general one needs to numerically integrate the basic evolution system (62)–(65) to get physical results of direct interest. However, we have shown that one can understand the various physical elements entering this system by comparing it to several simple approximations: the adiabatic approximation, the $`\dot{r}`$-linearized one, and the universal $`\rho `$-approximation (valid when $`\nu <\mathrm{\hspace{0.17em}0.05}`$). In particular, the latter approximation allowed us to derive some scaling laws: one scaling law (which is very well satisfied, even up to the maximum value $`\nu =1/4`$) states that the radial momentum at the Last Stable Orbit (LSO) scales like $`\nu ^{3/5}`$, while another scaling law (accurately satisfied only for $`\nu 1`$) states that the number of cycles left after the LSO scales like $`\nu ^{1/5}`$. The two most important consequences of the present approach are: (i) a way to compute initial dynamical data $`(𝒒_1,𝒒_2,𝒑_1,𝒑_2)`$ for a comparable-mass binary black hole system, represented in ADM coordinates, such that only a fraction of an orbit needs to be further evolved by numerical relativity techniques, and (ii) an estimate of the complete waveform emitted by a binary black hole coalescence, smoothly combining an inspiral signal, a plunge signal, a merger signal and a ring-down. However, much work remains to be done to firm up and complete our approach. We checked the robustness of our approach by considering an as-well-justified, slightly different evolution system. But stronger checks are called for. In particular it would be quite important to extend the present work (which used as input the 2.5PN-accurate damping and 2PN-accurate dynamics) to higher PN levels, when they become fully available. We note in this respect the recent work which extended the effective-one-body approach to the 3PN level. \[Note in passing that quasi-static tidal interactions between black holes enter only at the 5PN level .\] It is quite important to complete our determination of initial dynamical data $`(𝒒_a,𝒑_a)`$ by explicitly constructing the initial gravitational data $`(g_{ij}(x),K_{ij}(x))`$ corresponding to $`(𝒒_a,𝒑_a)`$ (and containing no free incoming radiation). When this becomes available it will be possible to further check our method (by numerically evolving spacetimes starting at various stages of the plunge) and to provide more accurate estimates of the merger waveform. Though our “light-ring-matching” approach to estimating the complete waveform is admittedly rough, we think it can play a useful role for data analysis: it can be used to test the accuracy of present templates (based on the adiabatic approximation) and allow one to construct more accurate, or at least, more robust, templates. We will come back to this issue in future work. Finally, let us note that it would be, in principle, important to be able to extend our approach to black holes having significant intrinsic spins. We, however, anticipate that this is a highly non-trivial task. ###### Acknowledgements. A.B.’s research was supported at Caltech by the Richard C. Tolman Fellowship and by NSF Grant AST-9731698 and NASA Grant NAG5-6840. All the numerical results in the present paper were produced using Mathematica.
warning/0001/cond-mat0001314.html
ar5iv
text
# Quantization of Hall Conductance in Double Exchange Systems: Topology and Lattice Gauge Field ## Abstract We study quantization conditions of the Hall conductivity for a two dimensional system described by a double exchange Hamiltonian with and without an external magnetic field. This is obtained by an extension of the topological arguments familiar from the theory of the integer quantum Hall effect. The quantization conditions are related to spontaneous breaking of spin $`O(3)`$, time-reversal, and spin chiral symmetries. Extension to systems with higher dimensions is briefly discussed. The quantum Hall effect (QHE) is a remarkable phenomena in condensed matter physics . After its discovery, there have been many theoretical works which explained the quantization of the Hall conductivity $`\sigma _{xy}=n\frac{e^2}{h}`$ for two dimensional electron gas in a uniform perpendicular magnetic field. An elegant approach, of which we will make use below, is based on topological and symmetry arguments. For Bloch electrons in a magnetic field, several authors, derived an explicit formula for the Hall conductance which is independent on the detailed structure of the periodic potential . The integer $`n`$ is shown to be a topological invariant, the first Chern class of a $`U(1)`$ principal fiber bundle on a torus . (See also .) In a pioneering work by Haldane , the possible occurrence of QHE without an external magnetic field in systems with broken time-reversal symmetry and its relation to chiral anomaly in 2d field theories have been discussed. (See also .) The QHE is mainly concerned with charge degrees of freedom of electrons. An experimentally and theoretically relevant question is, therefore, whether there is a similar effect pertaining to the spin degrees of freedom. Indeed, in transition metal oxides, coupling of spin degrees of freedom between itinerant and localized electrons induces many interesting transport properties . In this letter we show that, in principle, a special variant of the QHE may occur also in transition metal oxides with and without an external magnetic field. As we have reported recently , the basic ingredients are the Double Exchange (DE) model and the theory developed in Refs. . The DE model exhibits rich transport properties . The phase factor in the hopping integral induced by the $`t_{2g}`$ spins leads to an exotic ground state, (referred to as “flux” state), where the Hall conductance is expected to be quantized . The mechanism of stabilization of the “flux” state is similar to that of the flux phase discussed in the Hubbard model and in a generalized Peierls instability . Numerous other facets due to the Berry phase or to the “flux” were studied, such as an anomalous Hall effect , Jhan-Teller effect , and stripe formation . The conditions we study for the quantization of the Hall conductivity are summarized as follows: (i) the Fermi level should lie in a gap of extended states . (ii) the ground state is not degenerate . (iii) time-reversal symmetry is broken. Condition (iii) is evidently realized in the presence of a uniform external magnetic field in the standard QHE. Alternatively, in the DE model one may perceive two possibilities of breaking time-reversal symmetry: (a) Due to the minimum principle of external flux force line, it is natural to conjecture that the ground state of the DE model has a staggered flux structure where the net spontaneous flux vanishes. However, if a state with non-vanishing spontaneous flux exists, it breaks time-reversal symmetry. The pertinent state is characterized by the Wilson loop . The mechanism of stabilization of such a state was discussed in Refs . (b) Even if case (a) is hardly realizable, a system with a proper lattice structure spontaneously breaks the time-reversal symmetry. This kind of symmetry breaking corresponds to that of spin chirality . One such example is the Kagome lattice . In the rest of this letter, we first study the Hall conductivity in its general form within the DE model along the lines of TKNN, and then consider some specific examples. Finally, extension to systems with higher dimensions and the relation with theorems due to Lieb and Elitzur are briefly discussed. The Hamiltonian of the DE model in its minimum version at finite doping reads, $`H={\displaystyle \underset{i,j;\mu }{}}(tc_{i,\mu }^{}c_{j,\mu }^{}+h.c.+J\stackrel{}{S}_i\stackrel{}{S}_j)J_H{\displaystyle \underset{i}{}}\stackrel{}{\sigma }_i\stackrel{}{S}_i,`$ (1) where $`c_{i,\mu }`$ is a fermion (in fact, an $`e_g`$ electron) annihilation operator at site $`i`$ with spin projection $`\mu `$, $`\stackrel{}{\sigma }`$ is the spin operator of the $`e_g`$ electrons, $`\stackrel{}{S}_i`$ are localized ($`t_{2g}`$) spins which are treated as classical vectors directed along $`(\theta _i,\varphi _i)`$ in spherical coordinates. Moreover, $`J_H`$ is the Hund rule coupling constant and $`J`$ is the direct exchange coupling strength between $`t_{2g}`$ spins. There is a local $`SU`$(2) rotation invariance at every site, and, in the limit $`J_H\mathrm{}`$, the original Hamiltonian transforms into , $`\stackrel{~}{H}=t{\displaystyle \underset{i,j}{}}[(\mathrm{cos}{\displaystyle \frac{\theta _i}{2}}\mathrm{cos}{\displaystyle \frac{\theta _j}{2}}+e^{i(\varphi _i\varphi _j)}\mathrm{sin}{\displaystyle \frac{\theta _i}{2}}\mathrm{sin}{\displaystyle \frac{\theta _j}{2}})c_i^{}c_j^{}+h.c.]+J{\displaystyle \underset{i,j}{}}\stackrel{}{S}_i\stackrel{}{S}_j.`$ (2) The current operator on link $`ij`$ is transformed as $`v_{ij}={\displaystyle \frac{i}{\mathrm{}}}\left(c_i^{}c_j^{}c_j^{}c_i^{}\right)\stackrel{~}{v}_{ij}={\displaystyle \frac{i}{\mathrm{}}}\left[\mathrm{cos}{\displaystyle \frac{\theta _i}{2}}\mathrm{cos}{\displaystyle \frac{\theta _j}{2}}\left(c_i^{}c_j^{}c_j^{}c_i^{}\right)+\mathrm{sin}{\displaystyle \frac{\theta _i}{2}}\mathrm{sin}{\displaystyle \frac{\theta _j}{2}}\left(c_i^{}c_j^{}e^{i(\varphi _i\varphi _j)}c_j^{}c_i^{}e^{i(\varphi _i\varphi _j)}\right)\right].`$ (3) The transformation is unitary and does not affect the energy spectrum. However, the link variable which is employed to calculate the Hall conductance is not invariant. Below, we calculate the Hall conductacne, $`\sigma _{xy}`$, associated with the Hamiltonian (2). The phase factor in the second term in (3) is induced from the $`t_{2g}`$ spin configuration, which we denote by $`\{\theta _i,\varphi _i\}`$ below. The structure of the phase factor is related to the non-trivial magnetic ordering pattern $`\{\theta _i,\varphi _i\}`$. In a generic configuration, the $`e_g`$ electron acquires flux which cannot be eliminated by a local gauge transformation, that is, it accumulates a non-zero phase (mod $`2\pi `$) on moving around a closed path. One of the realizations of this scenario is the “flux” state . We now explain the calculation of $`\sigma _{xy}`$. In the absence of an external magnetic field, the contribution from the first term in (3) vanishes, so let us concentrate on the second one. (When the field is switched on, the conductivity is the sum of the two terms. The contribution from the first term is obtained within the TKNN formalism.) In the course of its evaluation, one has to employ a Fourier transformation. However, this turn out to be useless since in generic situations, the spin configuration $`\{\theta _i,\varphi _i\}`$ is not periodic. This is due to the property of the DE model where the energy is a functional of the oriented relative angle of the $`t_{2g}`$ spin . Thus, for a generic spin configuration, there is no simple algorithm for this calculation problem. Below, we assume a periodicity of the $`\{\theta _i,\varphi _i\}`$ configuration. This can be justified at least for the staggered flux state by noting that the flux state is degenerate. We define the respective unit cells for $`\theta `$’s and $`\varphi `$’s, $`U_\theta `$ and $`U_\varphi `$, which are not necessarily identical. These unit cells can be enlarged to the least common multiple (denoted hereafter by $`U`$) of $`U_\theta `$ and $`U_\varphi `$, if their ratio is rational. (An interesting situation occurs when the ratio of $`U_\theta `$ and $`U_\varphi `$ is irrational.) The expression for the Hall conductivity now reads , $`\sigma _{xy}={\displaystyle \underset{j}{}}\left[{\displaystyle \frac{e^2}{h}}{\displaystyle \frac{1}{2\pi i}}{\displaystyle _{MBZ}}𝑑\stackrel{}{k}\stackrel{}{k}|_k|\stackrel{}{k}\right]_j={\displaystyle \frac{e^2}{h}}t_r,`$ (4) where $`\stackrel{}{k}`$ is the crystal momentum and $`|\stackrel{}{k}`$ is a wave function in momentum space defined on the extended magnetic Brillouin zone (MBZ) associated with $`U`$. Note that here, the MBZ refers to the Brillouin zone of the “spin configuration.” Here $`j`$ is the index of the sub-band. Finally, the number $`t_r`$ on the right hand side of equation (4) is the solution of the Diophantine equation . The simplest example for which the above formalism can be worked out in detail is the DE model on a square lattice at half-filling. The staggered $`\pi `$-flux state is spontaneously stabilized , and the ground state has an infinite continuous degeneracy. The dispersion relation near the Fermi level is approximated by that for a gapless Dirac fermion. In the absence of an external magnetic field, the time-reversal symmetry is not broken because the spontaneous flux around a plaquette is $`\pm \pi `$, and the Hall conductance vanishes. Below we assume that the symmetry of the $`t_{2g}`$ spin is broken, thus removing the continuous degeneracy. We also introduce a next-nearest neighbor (NNN) hopping term for staggared plaquette in order to open an energy gap and to break time-reversal symmetry . We further assume that the staggered structure of the $`\pi `$-flux in (2) is robust against perturbation by NNN hopping. The effective Hamiltonian for the $`e_g`$ electron is described by $`H`$ $`=`$ $`t{\displaystyle \underset{m,n}{}}[e^{i\theta _1}c_{m+\frac{1}{2},n+\frac{1}{2}}^{}c_{m,n}^{}+e^{i\theta _2}c_{m+\frac{1}{2},n+\frac{1}{2}}^{}c_{m+1,n}^{}+e^{i\theta _3}c_{m,n+1}^{}c_{m+\frac{1}{2},n+\frac{1}{2}}^{}`$ (6) $`+e^{i\theta _2}c_{m+1,n+1}^{}c_{m+\frac{1}{2},n+\frac{1}{2}}^{}t^{}(e^{i\theta _4}c_{m,n+1}^{}c_{m,n}^{}+e^{i\pi \phi }c_{m+\frac{1}{2},n+\frac{1}{2}}^{}c_{m\frac{1}{2},n+\frac{1}{2}}^{})]+h.c.,`$ where $`\theta _1=\pi [\varphi (m+\frac{1}{4})+i2\phi ]`$, $`\theta _2=\pi \varphi (m+\frac{3}{4})`$, $`\theta _3=\pi \varphi (m+\frac{1}{4})`$, and $`\theta _4=2\pi \varphi m+i\pi \phi `$. Here the effect of the spontaneous staggered $`\pi `$-flux is replaced by $`\phi =1/2`$. Recently, it has been possible to fabricate networks which are described by the Hamiltonian (6) with $`t^{}=0`$ and to investigate the corresponding energy spectrum . In our calculations, we introduce also a uniform external magnetic field $`\varphi `$ (in the Landau gauge) thus taking into account the experimental setup. We now derive the Chern number using the method detailed in Ref. which provides a simple counting technique for the TKNN-vortex. In the absence of the external field, i.e. $`\varphi =0`$, the Hall conductance is quantized as a function of the perturbation, $`t_r=sgn(t^{})`$ and $`t_r=0`$ for $`t^{}=0`$. The value of $`t_r`$ for finite $`\varphi `$ is shown in Fig. 1 where the staggered flux state of the DE model can be interpreted only near $`\varphi 0`$. In the above formalism, we artificially break the time-reversal symmetry and open the energy gap. However, the symmetry would be spontaneously broken away from half-filling because the super-cell structure is expected to stabilize if one employs the analogy between the present model and the generalized Peierls instability . The geometrical structure of the lattice may induce a non-trivial spin configuration . Motivated by the work we study the DE model on a two dimensional pyrochlore like lattice , defined by assigning tetrahedrons on the Kagome lattice shared with one of the four triangles (see Fig.2(a)). In the limit $`J\mathrm{}`$, the ground state has continuous infinite degeneracy. (Its nature is distinct from that of the Kagome lattice where the relative angle between adjacent spins is $`2/3\pi `$. In the pyrochlore like lattice, the corresponding angles need not be identical. Typical configurations are displayed in Fig.2(b)-(e). The spin structure over entire lattice is obtained by combining these tetrahedrons and successively adjusting the relative angles of the shared edges.) Because of the degeneracy, one cannot evaluate (4). Below we assume that one of the spin configurations is selected due to symmetry breaking of the spin and that the periodicity coincides with the Wigner-Seitz cell of the lattice. Employing the method of Ref. we optimize the four spin configuration in the Wigner-Seitz cell in order to minimize the kinetic and exchange energies. We studied the model at fillings $`x=`$ 1/5, 2/5, and 3/5 and confirmed the above picture. Two simple and symmetric realizations are shown in Figs.2(f) and (g). In these examples, the flux through a triangle is exactly $`\pi `$ and the time-reversal symmetry is not broken. Note that in the degenerate ground state, there exists a spin configuration with broken time-reversal symmetry. An example is shown in Fig.2(e). However, here we study the most extreme cases (f) and (g). The effective Hamiltonian associated with these states is $`H=t{\displaystyle \underset{m,n}{}}[`$ $`e^{i\frac{2}{3}\pi |\phi |}c_{m+1,n}^{}c_{m,n}^{}+e^{i\frac{2}{3}\pi \phi }c_{m,n}^{}c_{m1,n}^{}+e^{i2\pi m\varphi i\frac{2}{3}\pi |\phi |}c_{m,n+1}^{}c_{m,n}^{}+e^{i2\pi m\varphi +i\frac{2}{3}\pi \phi }c_{m,n}^{}c_{m,n1}^{}`$ (10) $`+c_{m+1,n+1}^{(1)}c_{m,n}^{}+c_{m,n}^{}c_{m1,n1}^{(2)}+e^{i(2m+\frac{1}{3})\pi \varphi }c_{m+1,n+1}^{(1)}c_{m,n+1}^{}+e^{i(2m\frac{1}{3})\pi \varphi }c_{m,n1}^{}c_{m1,n1}^{(2)}`$ $`+e^{i\frac{1}{3}\varphi \pi }c_{m+1,n+1}^{(1)}c_{m+1,n}^{}+e^{i\frac{1}{3}\varphi \pi }c_{m1,n}^{}c_{m1,n1}^{(2)}`$ $`+e^{i\pi \varphi (2m+1)+i\frac{2}{3}\pi |\phi |}c_{m,n+1}^{}c_{m+1,n}^{}+e^{i\pi \varphi (2m1)i\frac{2}{3}\pi \phi }c_{m1,n}^{}c_{m,n1}^{}]+h.c.`$ Here the effect of the spontaneous staggered $`\pi `$-flux is replaced by $`\phi =\pm 1/2`$ for Figs.2(f) and (g), respectively. We also introduce a uniform external magnetic field $`\varphi `$ in the Landau gauge and study the energy gap structure of the model. For a generic $`\phi `$, the state shown in Fig.2(f) breaks time-reversal, while the state shown in (g) does not. These two states with different behavior under time reversal are (Kramer) degenerate. The pertinent Chern number is calculated by using the method developed in Refs. where, in the present case, the structure of the complex energy surface is classified as a Riemann surface associated with a ten degree algebraic equation which we have solved numerically. The $`t_r`$’s are shown in Fig.1(b). The region $`\varphi 0`$ corresponds to the states depicted in Fig.2(f) and (g). Perturbation by additional hopping matrix elements to the honeycomb cell (which is identical to the one discussed in Ref. ) opens an energy gap and $`\sigma _{xy}`$ is quantized to the value associated with the nearest energy gap. Realizing the quantization conditions (i-iii) is most difficult for the states which we have studied above. Introduction of an anisotropy in the hopping amplitude or exchange energy between in and out of plane bonds induces a tilt from the $`\pi `$-flux, and the time-reversal symmetry is broken. As another approach, we can construct the entire spin configuration from that of the tetrahedron configuration with broken time-reversal symmetry using the unit shown in Fig.2(e). The occurrence of the QHE in 3$`d`$ pyrochlore was speculated . We confirmed that the energy gap G in Fig. 1(b) survives in 3$`d`$ pyrochlore for an anisotropic inter-layer hopping integral. This leads the QHE in 3$`d`$ . Discussion: We have investigated the quantization condition of $`\sigma _{xy}`$ in the ground state of the DE model. In general, it is hard to simultaneously maintain the conditions (i-iii). Especially, the ground state of the DE model have a continuous infinite degeneracy which violates the condition (ii). (The Kramer degeneracy also violates it.) Some different degenerate spin configurations lead the same band structure of the $`e_g`$ electrons, and the other spin configurations might lead to different band structures. (We denote the Bloch wave associated with the band structures by $`|k_i`$ where $`i`$ is the index of degeneracy.) The full wave function is a superposition of Bloch functions and the integral in equation (4) is ill defined except when all $`|k_i`$ have the same Chern number. Therefore, we should stress that in most cases, the Hall conductance either vanishes or it is ill defined, even when the Fermi energy lies in an energy gap. The problem associated with this type of degeneracy is not explicitly addressed in TKNN. On the other hand, the possibility of the quantization cannot be ruled out if symmetry breaking occurs. The global spin O(3) symmetry breaking is allowed at zero temperature in 2$`d`$ or at finite temperature in 3$`d`$. The Chern number is different for each spin configuration and the quantized Hall conductivity depends on spin configuration. Therefore, the value of the quantization may serve as a probe determining the spin configuration. (In three or higher dimensions, the QHE occurs in transition metal oxides if the system satisfies the conditions (i-iii). For 3$`d`$, each band has three topological invariants (the first Chern numbers) on a 2-tori obtained by slicing the three-torus in three different manners. For general $`d`$, every quantized invariant on a $`d`$-dimensional torus $`T^d`$ is a function of the $`d(d1)/2`$ sets of TKNN integers obtained by slicing $`T^d`$ by the $`d(d1)/2`$ distinct $`T^2`$ .) The DE model with classical spin has local symmetry because the ground state energy is a functional of the oriented relative angles between spins. The symmetry breaking is difficult. Its relation to the Elitzur’s theorem is quite interesting. In this case, the Hall conductance again vanishes or undefined. At least, at half-filling, the $`\pi `$–flux state is expected to be stabilized in the DE model with quantum spin (even in three dimensions and/or with interactions), because the system maintains reflection positivity in spin space . In closing, we point out that flux states in the DE model manifest an interesting physics associated with the QHE and transport properties of transition metal oxides, as well as of networks with modulated nano-structures . Acknowledgment: We would like to thank Y. Avishai and M. Oshikawa for stimulating discussions and comments on the manuscript. The authors thank the Supercomputer Center, ISSP, U.Tokyo for the use of the facilities. M.Y. is supported by the Moritani scholarship foundation.
warning/0001/astro-ph0001114.html
ar5iv
text
# Advection–dominated Inflow/Outflows from Evaporating Accretion Disks ## 1. Introduction In recent years X–ray and optical observations provided increasing evidence that advection–dominated flows (ADAFs, Ichimaru ichimaru:1977 (1977); Narayan & Yi 1995b ; Abramowicz et al. abrchen:1995 (1995)) are a ubiquitous feature of accretion at all scales, from black holes (and possibly neutron stars) in X–ray binaries (e.g. Narayan narayan:1996 (1996); Esin, McClintock & Narayan esinmcclnar:1997 (1997); Menou et al. menou:1999 (1999)) to supermassive black holes in the center of galaxies (e.g. Narayan, Yi, & Mahadevan nayima95:1995 (1995), Lasota et al. lasota:1996 (1996), Narayan et al. nar98:1998 (1999); Gammie, Narayan, & Blandford ganarbla:1999 (1999)). Despite their structure being inherently two–dimensional, much theoretical work on ADAFs still relies on a vertically–integrated approach, which, although questionable, may indeed catch some of the essential properties of the model, as its successful application to various sources shows. The standard ADAF picture was, however, challenged in a recent paper by Blandford, & Begelman (blabe:1999 (1999)). Since, at least in self–similar ADAFs, the Bernoulli number is positive (as already noted by Narayan, & Yi narayanyi94:1994 (1994), 1995a ), they suggested that a large outflow may form. If this is the case, they have shown that the mass carried out by the wind must exceed by orders of magnitude that which is crossing the horizon, turning ADAFs into Advection–Dominated Inflow/Outflow Solutions, or ADIOS. Although the positiveness of the Bernoulli number is only a necessary (but not sufficient) condition for starting an outflow, this argument poses a serious problem to the ADAF model for black hole accretion and must be addressed carefully. Goal of this Letter is to investigate how, and to which extent, the inclusion of the source of ADAF material affects the Bernoulli number and the onset of a wind. Advection–dominated flows in black hole X–ray binaries (BHXBs) are most likely produced by the evaporation of a Shakura–Sunyaev disk (SSD, Shakura, & Sunyaev shaksuny:1973 (1973)), as observational and theoretical arguments suggest (e.g. Narayan, Mahadevan, & Quataert namaqua:1999 (1999)). The evaporation process is not fully understood as yet (see e.g. Meyer, & Meyer–Hofmeister meyermeyhof:1994 (1994); Honma honma:1996 (1996); Dullemond dul:1999 (1999); Rózańska roz:1999 (1999)) and definitely needs to be modeled in at least two spatial dimensions. Here we just assume a simple, analytical law for the evaporation rate and limit ourselves to a vertically–integrated description of the hot flow. A quite general argument, based on angular momentum conservation, indicates that purely inflowing solutions can not exist if the accretion rate decreases with radius. Numerical models, computed for several values of the $`\alpha `$–viscosity parameter and of the transition radius $`R_0`$, support this conclusion. In all of them a stagnation radius (where the radial velocity vanishes) separates an inner inflowing region from an outer transsonic wind. We find that the Bernoulli number for the infalling gas is negative if the transition radius is less than $`100`$ Schwarzschild radii. In these solutions about $`1/3`$ of the mass carried inwards from large radii by the thin disk reaches the horizon. More extended inflows, with $`R_0100R_s`$, have a region of positive Bernoulli number, and are likely to be replaced by an ADIOS. ## 2. The Model We consider a highly idealized model for a hybrid accretion flow, in which a hot, advection–dominated phase coexists with a Shakura–Sunyaev disk. All the ADAF gas is assumed to be supplied by the evaporation of the surface layers of the SSD which extends down to the transition radius $`R_0`$. In the following we will not be concerned with the SSD anymore and focus our attention on the hot component. The inner rim is chosen to be at $`R_{in}=3GM/c^2`$ ($`M`$ is the hole mass), and $`R_0>R_{in}`$. The advection–dominated flow is described by the usual stationary, vertically–integrated equations (see e.g. Narayan, Kato, & Honma narkathon:1997 (1997)), which now include the energy and momentum exchange between the hot gas and the evaporating material; the standard $`\alpha `$–prescription for viscosity is retained, the pseudo–Newtonian potential is used to describe the gravitational field and we neglect radiative losses. The physics of the evaporation process is still unclear, so we just assume that matter is lost from the SSD (per unit area and time) according to the simple law $$\dot{\sigma }(R)=\{\begin{array}{ccc}\dot{\sigma }_0(R/R_0)^{\xi 2}& & RR_0\\ 0& & R<R_0.\end{array}$$ (1) The continuity equation reads $$\frac{d(R\mathrm{\Sigma }v_R)}{dR}=R\dot{\sigma }$$ (2) which can be immediately integrated to yield $`2\pi R\mathrm{\Sigma }v_R=\dot{M}_{ADAF}(R)`$. In the previous expressions $`\mathrm{\Sigma }`$ is the surface density and $`v_R`$ the radial velocity, chosen to be negative for matter flowing towards the hole. The ADAF accretion rate now depends on $`R`$ and is given by $$\dot{M}_{ADAF}(R)=\{\begin{array}{cc}\dot{M}(R/R_0)^\xi \dot{M}_{out}& RR_0\\ \dot{M}\dot{M}_{out}& R<R_0\end{array}$$ (3) where $`\dot{M}_{out}=(2\pi R\mathrm{\Sigma }v_R)|_R\mathrm{}`$, $`\dot{M}=2\pi \xi ^1\dot{\sigma }_0R_0^2=\dot{M}_{SSD}(R\mathrm{})`$ is total accretion rate, and we consider only the case $`\xi >0`$, so $`\dot{M}_{ADAF}`$ decreases with $`R`$. If $`\dot{M}_{out}>0`$, $`\dot{M}_{ADAF}`$ becomes negative for $`R>R_{st}=[1+(\dot{M}\dot{M}_{out})/\dot{M}_{out}]^{1/\xi }R_0`$ and the gas crosses the horizon at a rate $`\dot{M}_{in}=\dot{M}\dot{M}_{out}<\dot{M}`$; this implies that also $`v_R`$ has to switch sign at the stagnation radius $`R_{st}`$. In the following we assume that the injected material rotates at the Keplerian angular speed $`\mathrm{\Omega }_K`$ and that the rising gas elements move predominantly in the vertical direction with velocity $`\mathrm{\Omega }_KR`$. Since the ADAF is expected to rotate at $`\mathrm{\Omega }<\mathrm{\Omega }_K`$, the difference in the circular speeds produces a torque. Conservation of angular momentum then implies $$\mathrm{\Sigma }Rv_R\frac{d(R^2\mathrm{\Omega })}{dR}\frac{d}{dR}\left(\mathrm{\Sigma }\nu R^3\frac{d\mathrm{\Omega }}{dR}\right)=\dot{\sigma }(\mathrm{\Omega }_K\mathrm{\Omega })R^3$$ (4) where $`\nu =2/3\alpha c_sH`$ is the viscosity coefficient, $`H`$ the flow half–thickness and $`c_s`$ the isothermal sound speed. Other possible sources of friction between the SSD and the ADAF like, e.g., magnetic stresses, have been neglected. The ADAF has to spend part of its energy to heat up the injected gas which has initially the same temperature of the Shakura–Sunyaev disk. At the same time heat is produced by frictional dissipation. Including both these effects, the local energy balance takes the form $`v_R\left[(\gamma 1){\displaystyle \frac{dc_s^2}{dR}}c_s^2{\displaystyle \frac{d\mathrm{ln}(\mathrm{\Sigma }/H)}{dR}}\right]\nu \left(R{\displaystyle \frac{d\mathrm{\Omega }}{dR}}\right)^2=`$ $`{\displaystyle \frac{\dot{\sigma }}{\mathrm{\Sigma }}}\left[{\displaystyle \frac{1}{2}}v_R^2+{\displaystyle \frac{1}{2}}R^2(\mathrm{\Omega }_K\mathrm{\Omega })^2{\displaystyle \frac{\gamma }{\gamma 1}}c_s^2\right]`$ (5) where $`\gamma `$ is the adiabatic index and we assumed that the rising gas elements do not suffer any energy loss before thermalizing with the hot plasma. Finally, the radial force balance is expressed as $$v_R\frac{dv_R}{dR}+(\mathrm{\Omega }_K^2\mathrm{\Omega }^2)R+\frac{dc_s^2}{dR}+c_s^2\frac{d\mathrm{ln}(\mathrm{\Sigma }/H)}{dR}=\frac{\dot{\sigma }}{\mathrm{\Sigma }}v_R;$$ (6) the last term arises because the injected mass has zero radial momentum and is usually negligible. As expected, Eqs. (2), (4)–(6) reduce to the standard ADAF form for $`\dot{\sigma }=0`$ . ## 3. Global Solutions for the Hot Flow It can be easily shown that, under the usual assumptions, the flow equations admit a self–similar solution. However, at variance with standard advection–dominated models, a self–similar regime is not always allowed for evaporation–fed ADAFs. In fact, at least if $`0<\mathrm{\Omega }<\mathrm{\Omega }_K`$, angular momentum conservation implies $`(12\xi )/v_R<0`$, so inflowing self–similar solutions are ruled out unless $`\xi <1/2`$. The reason for this is as follows. If the amount of mass injected into the ADAF drops off too quickly with increasing $`R`$, viscous stresses can not transport enough angular momentum outwards because the density becomes too low at large radii. The ADAF can not get rid of its angular momentum by transferring it to the SSD either, since the cold disk rotates faster. On a physical basis, it seems therefore unlikely that purely inflowing global solutions could exist. The only possibility for the flow to transport efficiently angular momentum to infinity is to advect it, reversing its motion from a certain radius onwards. Eqs. (2), (4)–(6) have been solved numerically using a Henyey relaxation method (Dullemond, & Turolla dultur:1998 (1998)). Despite several attempts, no purely inflowing solution was found for $`\xi >1/2`$, as the previous argument predicts. In all models both the outflow and the inflow are transsonic. The presence of three critical points (the two sonic radii and the stagnation radius) reduces the number of the boundary condition from five (there are three first order and one second order differential equations) to two, the other three being replaced by regularity conditions at the critical points. Since the viscosity must be well–behaved at both the inner and outer edge (the no–torque condition, see Narayan, Kato, & Honma narkathon:1997 (1997)), we require $$\frac{d\mathrm{log}\mathrm{\Omega }}{d\mathrm{log}R}=2\text{at}R=R_{in}\text{and}R=R_{out}.$$ (7) This choice introduces no further degree of freedom. The solution depends only on the function $`\dot{\sigma }(R)`$, on $`\alpha `$ and $`\gamma `$. Both the mass loss rate at large radii $`\dot{M}_{out}`$, and the stagnation radius follow from the calculation and are found to obey the analytical expression for $`R_{st}`$ derived in §2 to high accuracy. We have computed several series of models with $`0.1<\alpha <1`$, $`1/2<\xi <3/2`$ and $`\gamma =3/2`$, varying $`R_0`$ in the range $`20R_s`$$`1000R_s`$. As Eqs. (4)–(6) show, $`v_R`$, $`\mathrm{\Omega }`$ and $`c_s`$ are independent of $`\dot{\sigma }_0`$ and $`M`$; the same is for the ratio $`\dot{M}_{in}/\dot{M}_{out}`$ of accretion to mass loss rates. The radial dependence of $`v_R`$ and $`c_s`$ for a typical run is plotted in Fig. Advection–dominated Inflow/Outflows from Evaporating Accretion Disks; solutions with different values of $`\alpha `$ and $`\xi `$ show the same general behaviour. Only a limited $`\xi `$–range was considered here, but we have verified that steeper (e.g. exponential) injection laws give quite similar results. The accretion to mass loss ratio goes from $`\dot{M}_{in}/\dot{M}_{out}0.5`$ to $`0.3`$ as $`R_0`$ increases from 20 to 500 $`R_s`$ and is not much dependent of $`\alpha `$ and $`\xi `$. Correspondingly, $`R_{st}`$ is in all cases about $`1.5R_0`$. As expected, for $`R_0500R_s`$ the inflow closely resembles standard ADAFs and, in particular, has a nearly self–similar region at intermediate radii. For smaller values of $`R_0`$ the accretion flow has no room to attain self–similarity, being squeezed in between the sonic and the stagnation radius. ## 4. Discussion Our solutions share with ADIOS the property that a significant (although very different) fraction of the material is expelled in a wind. The two models, however, differ substantially in many respects. Blandford, & Begelman (blabe:1999 (1999)) constructed self–similar advection–dominated solutions for which the Bernoulli number $`Be`$ is negative assuming that $`\dot{M}_{ADAF}R^p`$ with $`p>0`$. They concluded that all the mass which does not reach the horizon escapes in a wind more massive than the ADAF by orders of magnitude. Although no detailed model including the wind has been presented as yet, preliminary hydrodynamical calculations seem indeed to support the original suggestion that the mass inflow rate increases with radius (Stone, Pringle, & Begelman stopribe:1999 (1999)). In addition, 2–D simulations by Igumenshchev, & Abramowicz (iguabra:1999 (1999)) have shown that ADIOS–like solutions are present for large $`\alpha `$. In our model the outflow is a direct consequence of having assumed that the ADAF material is supplied by the SSD. The accreting gas has to transfer angular momentum to larger radii and, since $`\dot{M}_{ADAF}`$ is decreasing with increasing $`R`$ in the evaporation region, this can not be done by viscous stresses alone. From the stagnation radius onwards, the gas is centrifugally accelerated away from the hole, carrying angular momentum with it: no stationary solution would be possible without an outflowing region. This is not related to the positiveness of the Bernoulli number. Moreover, the wind is not produced by the ADAF itself but originates directly from the evaporating SSD material. The basic objection Blandford and Begelman raised to standard ADAFs is very general and concerns also the inflowing part of evaporation–fed models. The Bernoulli number for our solutions is shown in Fig. Advection–dominated Inflow/Outflows from Evaporating Accretion Disks. In the wind region the Bernoulli number becomes positive shortly beyond the stagnation radius, while $`Be`$ can be either positive or negative for the infalling gas, depending on $`R_0`$. Solutions with a small transition radius have always $`Be<0`$. This can be understood considering the energy the hot flow transfers to the cold injected material (last term in eq. ). The evaporation stops at $`R_0`$ but the inflowing gas will stay cooler (with respect to a standard ADAF) a bit further down. For $`R_0100R_s`$ viscous dissipation has no time to heat the flow sufficiently before it crosses the horizon, keeping $`Be`$ negative (see also Nakamura naka:1998 (1998) for a discussion on the sign of the Bernoulli number in non–self–similar ADAF solutions). As $`R_0`$ increases above $`100R_s`$ the Bernoulli number is still negative close to $`R_0`$ but it flips sign as soon as the inflowing gas has become hot enough. An ADIOS–type wind may be expected where $`Be>0`$. The inclusion of a source for the hot gas points towards the existence of a more general class of advection–dominated flows, with somewhat intermediate characteristics between standard ADAFs and ADIOS. In the light of the relation to (and possible inclusion of) ADIOS–type flows, we refer to these models as “Consistent Inflow And Outflows”, or CIAOs. Our vertically–integrated CIAO model has to be supported by detailed 2–D and 3–D hydrodynamical calculations, but the present analysis suggests that for $`R_0100R_s`$ the inflowing matter is gravitationally bound and no ADIOS–like wind is expected to be present. CIAOs can supply the central hole with about 1/3 to 1/4 of the mass originally carried inwards by the thin disk. This may be relevant for applications of advection–dominated flows to BHXBs and AGNs. It has been often proposed (e.g. Narayan narayan:1996 (1996); Esin, McClintock, & Narayan esinmcclnar:1997 (1997); Belloni, et al. belloni:1997 (1997); Narayan, Mahadevan & Quataert namaqua:1999 (1999)) that the different spectral states observed in BHXBs may be explained in terms of a bimodal disk in which the transition radius varies with time. The inferred values of $`R_0`$ are in between few $`R_s`$ and $`100R_s`$. Recent optical/UV observations also indicate that the inner edge of the thin disk in the nucleus of M81 and NGC 4579 lies at $`100R_s`$ (Quataert, et al. quaetal:1999 (1999)). Although a more detailed analysis is definitely required before any firm conclusion can be drawn, it is interesting to note that the maximum value of the transition radius for which our models have $`Be<0`$ is close to the observed limit mentioned above. CIAOs may comfortably provide the hot accretion flow in all the range of $`R_0`$ implied by observations. At the same time, the absence of hot flows with an extent $`100R_s`$ can be interpreted in terms of the onset of a strong wind which, as in ADIOS, blows off the accreting part of the CIAO and reduces dramatically the accretion rate onto the black hole.
warning/0001/math0001166.html
ar5iv
text
# Weyl structures for parabolic geometries ## 1. Introduction Cartan’s generalized spaces are curved analogs of the homogeneous spaces $`G/P`$ defined by means of an absolute parallelism on a principal $`P`$–bundle. This very general framework was originally built in connection with the equivalence problem and Cartan’s general method for its solution, cf. e.g. . Later on, however, these ideas got much more attention. In particular, several well known geometries were shown to allow a canonical object of that type with suitable choice of semisimple $`G`$ and parabolic $`P`$, see e.g. . Cartan’s original approach was generalized and extended for all such groups, cf. , and links to other areas were discovered, see e.g. . The best known examples are the conformal Riemannian, projective, almost quaternionic, and CR structures and the common name adopted is parabolic geometries. The relation to twistor theory renewed the interest in a good calculus for such geometries, which had to improve the techniques in conformal geometry and to extend them to other geometries. Many steps in this direction were done, see for example for classical methods in conformal geometry, and for generalizations. A new approach to this topic, motivated mainly by , was started in . The novelty consists in the combination of Lie algebraic tools with the frame bundle approach to all objects and we continue in this spirit here. Our general setting for Weyl structures and scales has been also inspired by . In Section 2 we first outline some general aspects of parabolic geometries and then we present the basic objects like tangent and cotangent bundles and the curvature of the geometry in a somewhat new perspective. This will pave our way to the Weyl structures in the rest of the paper. Our basic references for Section 2 are and , the reader may also consult . For the classical point of view of over–determined systems, we refer to and the references therein. The Weyl structures are introduced in the beginning of Section 3. Exactly as in the conformal Riemannian case, the class of Weyl structures underlying a parabolic geometry on a manifold $`M`$ is always an affine space modeled on one–forms on $`M`$ and each of them determines a linear connection on $`M`$. Moreover, the difference between the linear connection induced by a Weyl structure and the canonical Cartan connection is encoded in the so called Rho–tensor (used heavily in conformal geometry since the beginning of the century). Next, we define the bundles of scales as certain affine line bundles generalizing the distinguished bundles of conformal metrics, and we describe the correspondence between connections on these line bundles and the Weyl structures, see Theorem 3.12. On the way, we achieve explicit formulae for the deformation of Weyl structures and the related objects in Proposition 3.9, which offers a generalization for the basic ingredients of various calculi. The exact Weyl geometries are given by scales, i.e. by (global) sections of the bundles of scales, thus generalizing the class of Levi–Civita connections for conformal geometries. At the same time, this point of view leads to a new presentation of the canonical Cartan bundle as the bundle of connections on the bundle of scales (pulled back to the defining infinitesimal flag structure, cf. 2.7 and 3.12). In the end of Section 3, we define another class of distinguished local Weyl structures which achieve the best possible approximation of the canonical Cartan connections, see Theorem 3.16. In the conformal case, these normal Weyl structures improve the construction of the Graham’s normal coordinates, cf. . The last section is devoted to characterizations of all the objects related to a choice of a Weyl structure. More explicitly, the ultimate goal is to give a recipe how to decide which soldering forms and linear connections on a manifold $`M`$ equipped with a regular infinitesimal flag structure are obtained from a Weyl–structure and to compute the corresponding Rho–tensor. For this purpose, we define the general Weyl forms and their Weyl curvatures and the main step towards our aim is achieved in Theorem 4.4. Next, we introduce the total curvature of a Weyl form which is easier to interpret on the underlying manifold than the Weyl curvature. The characterization is then obtained by carefully analyzing the relation between these two curvatures. This entire paper focuses on the introduction of new structures and their nice properties. We should like to mention that essential use of these new concepts has appeared already in and . Acknowledgements. The initial ideas for this research evolved during the stay of the second author at the University of Adelaide in 1997, supported by the Australian Research Council. The final work and writing was done at the Erwin Schrödinger Institute for Mathematical Physics in Vienna. The second author also acknowledges the support from GACR, Grant Nr. 201/99/0296. Our thanks are also due to our colleagues for many discussions. ## 2. Some background on parabolic geometries ### 2.1. $`|k|`$–graded Lie algebras Let $`G`$ be a real or complex semisimple Lie group, whose Lie algebra $`𝔤`$ is equipped with a grading of the form $$𝔤=𝔤_k\mathrm{}𝔤_0\mathrm{}𝔤_k.$$ Such algebras $`𝔤`$ are called $`|k|`$–graded Lie algebras. Throughout this paper we shall further assume that no simple ideal of $`𝔤`$ is contained in $`𝔤_0`$ and that the (nilpotent) subalgebra $`𝔤_{}=𝔤_k\mathrm{}𝔤_1`$ is generated by $`𝔤_1`$. Such algebras are sometimes called effective semisimple graded Lie algebras of $`k`$-th type, cf. . By $`𝔭_+`$ we denote the subalgebra $`𝔤_1\mathrm{}𝔤_k`$ and by $`𝔭`$ the subalgebra $`𝔤_0𝔭_+`$. We also write $`𝔤_{}=𝔤_k\mathrm{}𝔤_1`$, and $`𝔤^j=𝔤_j\mathrm{}𝔤_k`$, $`j=k,\mathrm{},k`$. It is well known that then $`𝔭`$ is a parabolic subalgebra of $`𝔤`$, and actually the grading is completely determined by this subalgebra, see e.g. , Section 3. Thus all complex simple $`|k|`$–graded $`𝔤`$ are classified by subsets of simple roots of complex simple Lie algebras (i.e. arbitrary placement of crosses over the Dynkin diagrams in the notation of ), up to conjugation. The real $`|k|`$–graded simple Lie algebras are classified easily by means of Satake diagrams: the $`|k|`$–grading of the complex simple $`𝔤`$ induces a $`|k|`$–grading on a real form if and only if (i) only ‘white’ nodes in the Satake diagram have been crossed out, and, (ii) if a node is crossed out, then all nodes connected to this one by the double arrows in the Satake diagram have to be crossed out too, see or for more details. Very helpful notational conventions and computational recipes can be found in . ### 2.2. Let us recall basic properties of Lie groups $`G`$ with (effective) $`|k|`$–graded Lie algebras $`𝔤`$. First of all, there is always a unique element $`E𝔤_0`$ with the property $`[E,Y]=jY`$ for all $`Y𝔤_j`$, $`j=k,\mathrm{},k`$, the grading element. Of course, $`E`$ belongs to the center $`𝔷`$ of the reductive part $`𝔤_0`$ of $`𝔭𝔤`$. The Killing form provides isomorphisms $`𝔤_i^{}𝔤_i`$ for all $`i=k,\mathrm{},k`$ and, in particular, its restrictions to the center $`𝔷`$ and the semisimple part $`𝔤_0^{ss}`$ of $`𝔤_0`$ are non–degenerate. Now, there is the closed subgroup $`PG`$ of all elements whose adjoint actions leave the $`𝔭`$–submodules $`𝔤^j=𝔤_j\mathrm{}𝔤_k`$ invariant, $`j=k,\mathrm{},k`$. The Lie algebra of $`P`$ is just $`𝔭`$ and there is the subgroup $`G_0P`$ of elements whose adjoint action leaves invariant the grading by $`𝔤_0`$–modules $`𝔤_i`$, $`i=k,\mathrm{},k`$. This is the reductive part of the parabolic Lie subgroup $`P`$, with Lie algebra $`𝔤_0`$. We also define subgroups $`P_+^j=\mathrm{exp}(𝔤_j\mathrm{}𝔤_k)`$, $`j=1,\mathrm{},k`$, and we write $`P_+`$ instead of $`P_+^1`$. Obviously $`P/P_+=G_0`$ and $`P_+`$ is nilpotent. Thus $`P`$ is the semidirect product of $`G_0`$ and the nilpotent part $`P_+`$. More explicitly, we have (cf. , Proposition 2.10, or ) ###### 2.3 Proposition. For each element $`gP`$, there exist unique elements $`g_0G_0`$ and $`Z_i𝔤_i`$, $`i=1,\mathrm{},k`$, such that $$g=g_0\mathrm{exp}Z_1\mathrm{exp}Z_2\mathrm{}\mathrm{exp}Z_k.$$ ### 2.4. Parabolic geometries Following Elie Cartan’s idea of generalized spaces (see for a recent reading), a curved analog of the homogeneous space $`G/P`$ is a right invariant absolute parallelism $`\omega `$ on a principal $`P`$–bundle $`𝒢`$ which reproduces the fundamental vector fields. In our approach, a (real) parabolic geometry $`(𝒢,\omega )`$ of type $`G/P`$ is a principal fiber bundle $`𝒢`$ with structure group $`P`$, equipped with a smooth one–form $`\omega \mathrm{\Omega }^1(𝒢,𝔤)`$ satisfying 1. $`\omega (\zeta _Z)(u)=Z`$ for all $`u𝒢`$ and fundamental fields $`\zeta _Z`$, $`Z𝔭`$ 2. $`(r^b)^{}\omega =\mathrm{Ad}(b^1)\omega `$ for all $`bP`$ 3. $`\omega |_{T_u𝒢}:T_u𝒢𝔤`$ is a linear isomorphism for all $`u𝒢`$. In particular, each $`X𝔤`$ defines the constant vector field $`\omega ^1(X)`$ defined by $`\omega (\omega ^1(X)(u))=X`$, $`u𝒢`$. In this paper, we shall deal with smooth real parabolic geometries only. The one forms with properties (1)–(3) are called Cartan connections, cf. . The morphisms between parabolic geometries $`(𝒢,\omega )`$ and $`(𝒢^{},\omega ^{})`$ are principal fiber bundle morphisms $`\phi `$ which preserve the Cartan connections, i.e. $`\phi :𝒢𝒢^{}`$ and $`\phi ^{}\omega ^{}=\omega `$. ### 2.5. The curvature The structure equations define the horizontal smooth form $`K\mathrm{\Omega }^2(𝒢,𝔤)`$ called the curvature of the Cartan connection $`\omega `$: $$d\omega +\frac{1}{2}[\omega ,\omega ]=K.$$ The curvature function $`\kappa :𝒢^2𝔤_{}^{}𝔤`$ is then defined by means of the parallelism $$\kappa (u)(X,Y)=K(\omega ^1(X)(u),\omega ^1(Y)(u))=[X,Y]\omega ([\omega ^1(X),\omega ^1(Y)]).$$ In particular, the curvature function is valued in the cochains for the second cohomology $`H^2(𝔤_{},𝔤)`$. Moreover, there are two ways how to split $`\kappa `$. We may consider the target components $`\kappa _i`$ according to the values in $`𝔤_i`$. The whole $`𝔤_{}`$–component $`\kappa _{}`$ is called the torsion of the Cartan connection $`\omega `$. The other possibility is to consider the homogeneity of the bilinear maps $`\kappa (u)`$, i.e. $$\kappa =\underset{\mathrm{}=k+2}{\overset{3k}{}}\kappa ^{(\mathrm{})},\kappa ^{(\mathrm{})}:𝔤_i\times 𝔤_j𝔤_{i+j+\mathrm{}}.$$ Since we deal with semisimple algebras only, there is the codifferential $`^{}`$ which is ajoint to the Lie algebra cohomology differential $``$, see e.g. . Consequently, there is the Hodge theory on the cochains which enables to deal very effectively with the curvatures. In particular, we may use several restrictions on the values of the curvature which turn out to be quite useful. ### 2.6. Definition The parabolic geometry $`(𝒢,\omega )`$ with the curvature function $`\kappa `$ is called flat if $`\kappa =0`$, torsion–free if $`\kappa _{}=0`$, normal if $`^{}\kappa =0`$, and regular if it is normal and $`\kappa ^{(j)}=0`$ for all $`j0`$. Obviously, the morphisms of parabolic geometries preserve the above types and so we obtain the corresponding full subcategories of regular, normal, torsion free, and flat parabolic geometries of a fixed type $`G/P`$. See , Section 2, for more details. ### 2.7. Flag structures The homogeneous models for parabolic geometries are the real generalized flag manifolds $`G/P`$. Curved parabolic geometries look like $`G/P`$ infinitesimally. Indeed, the filtration of $`𝔤`$ by the $`𝔭`$–submodules $`𝔤^j`$ is transfered to the right invariant filtration $`T^j𝒢`$ on the tangent space $`T𝒢`$ by the parallelism $`\omega `$. The tangent projection $`Tp:T𝒢TM`$ then provides the filtration $`TM=T^kMT^{k+1}M\mathrm{}T^1M`$ of the tangent space of the underlying manifold $`M`$. Moreover, the structure group of the associated graded tangent space $`\mathrm{Gr}TM=(T^kM/T^{k+1}M)\mathrm{}(T^2M/T^1M)T^1M`$ reduces automatically to $`G_0`$ since $`𝒢_0=𝒢/P_+`$ clearly plays the role of its frame bundle. The following lemma is not difficult to prove, see e.g. , Lemma 2.11. ###### Lemma. Let $`(𝒢,\omega )`$ be a parabolic geometry, $`\kappa `$ its curvature function. Then $`\kappa ^{(j)}=0`$ for all $`j<0`$ if and only if the Lie bracket of vector fields on $`M`$ is compatible with the filtration, i.e. $`[\xi ,\eta ]`$ is a section of $`T^{i+j}M`$ for all sections $`\xi `$ of $`T^iM`$, and $`\eta `$ of $`T^jM`$. Hence it defines an algebraic bracket $`\{,\}_{\text{Lie}}`$ on $`\mathrm{Gr}TM`$. Moreover, this bracket coincides with the algebraic bracket $`\{,\}_{𝔤_0}`$ defined on $`\mathrm{Gr}TM`$ by means of the $`G_0`$–structure if and only if $`\kappa ^{(j)}=0`$ for all $`j0`$. We call the filtrations of $`TM`$ with reduction of $`\mathrm{Gr}TM`$ to $`G_0`$ satisfying the very last condition of the lemma the regular infinitesimal flag structures of type $`𝔤/𝔭`$. In fact, the structures clearly depend on the choice of the Lie group $`G`$ with the given Lie algebra $`𝔤`$. This choice is always encoded already in $`G_0`$. On the other hand, there are always several distinguished choices, e.g. the full automorphism group of $`𝔤`$, the adjoint group, and the unique connected and simply connected group. In the conformal geometries these choices lead to conformal Riemannian manifolds, oriented conformal menifolds, and (oriented) conformal spin manifolds, respectively. Obviously, the various choices of $`G`$ do not matter much locally and we shall not discuss them explicitly in this paper. The $`G_0`$ structures on $`\mathrm{Gr}TM`$ are equivalent to the frame forms of length one defined and used in while the condition $`\kappa ^{(j)}=0`$ for all $`j0`$ is equivalent to the structure equations for these frame forms imposed in the construction of . In view of this relation, we also call our bundles $`𝒢_0`$ equipped with the regular infinitesimal flag structures the $`P`$–frame bundles of degree one. In particular, we obtain (see , Section 3) ###### 2.8 Theorem. There is the bijective correspondence between the isomorphism classes of regular parabolic geometries of type $`G/P`$ and the regular infinitesimal flag structures of type $`𝔤/𝔭`$ on $`M`$, except for one series of one–graded, and one series of two–graded Lie algebras $`𝔤`$ for which $`H^1(𝔤_{},𝔤)`$ is nonzero in homogeneous degree one. Both types of the exceptional geometries from the Theorem will be mentioned in the examples below. ### 2.9. Example The parabolic geometries with $`|1|`$–graded Lie algebras $`𝔤`$ are called irreducible. Their tangent bundles do not carry any nontrivial natural filtration and this irreducibility of $`TM`$ is reflected in the name. The classification of all such simple real Lie algebras is well known (cf. or 2.1 above). We may list all the corresponding geometries, up to the possible choices of the groups $`G_0`$, roughly as follows: * the split form, $`\mathrm{}>2`$ — the almost Grassmannian structures with homogeneous models of $`p`$–planes in $`^{\mathrm{}+1}`$, $`p=1,\mathrm{},\mathrm{}`$. The choice $`p=1`$ yields the projective structures which represent one of the two exceptions in 2.8. * the quaternionic form, $`\mathrm{}=2p+1>2`$ — the almost quaternionic geometries in dimensions $`4p`$, and more general geometries modeled on quaternionic Grassmannians. * one type of geometry for the algebra $`𝔰𝔲(p,p)`$, $`\mathrm{}=2p1`$. * the (pseudo) conformal geometries in all odd dimensions $`2m+13`$. * the split form, $`\mathrm{}>2`$ — the almost Lagrangian geometries modeled on the Grassmann manifold of maximal Lagrangian subspaces in the symplectic $`^2\mathrm{}`$. * another type of geometry corresponding to the algebra $`𝔰𝔭(p,p)`$, $`\mathrm{}=2p`$. * the (pseudo) conformal geometries in all even dimensions $`m4`$. * the real almost spinorial geometries with $`𝔤=𝔰𝔬(p,2\mathrm{}p)`$, $`p=1,\mathrm{},\mathrm{}2`$. * the quaternionic almost spinorial geometries with $`𝔤=𝔲^{}(\mathrm{},)`$. * the split form $`EI`$ — exactly one type with $`𝔤_0=𝔰𝔬(5,5)`$ and $`𝔤_1=^{16}`$. * the real form $`EIV`$ — exactly one type with $`𝔤_0=𝔰𝔬(1,9)`$ and $`𝔤_1=^{16}`$. * the split form $`EV`$ — exactly one type with $`𝔤_0=EI`$ and $`𝔤_1=^{27}`$. * the real form $`EVII`$ – exactly one type with $`𝔤_0=EIV`$ and $`𝔤_1=^{27}`$. ### 2.10. Example The parabolic contact geometries form another important class. They correspond to $`|2|`$–graded Lie algebras $`𝔤`$ with one–dimensional top components $`𝔤_2`$. Thus the regular infinitesimal structures are equivalent to contact geometric structures, together with the reduction of the graded tangent space to the subgroup $`G_0`$ in the group of contact transformations. The only exceptions are the so called projective contact structures ($`C_{\mathrm{}}`$ series of algebras) where more structure has to be added, see e.g. . The general classification scheme allows a simple formulation for the contact cases: The dimension one condition on $`𝔤_2`$ yields the prescription which simple roots have to be crossed while the prescribed length two of the grading gives further restrictions. The outcome may be expressed as (see ): ###### Proposition. Each non–compact real simple Lie algebra $`𝔤`$ admits a unique grading of contact type (up to conjugacy classes), except $`𝔤`$ is one of $`𝔰𝔩(2,)`$, $`𝔰𝔩(\mathrm{},)`$, $`𝔰𝔭(p,q)`$, $`𝔰𝔬(1,q)`$, $`EIV`$, $`FII`$ and in these cases no such gradings exist. The best known examples are the non–degenerate hypersurface type CR geometries (with signature $`(p,q)`$ of the Levi form) which are exactly the torsion free regular parabolic geometries with $`𝔤=𝔰𝔲(p+1,q+1)`$, see e.g. , Section 4.14–4.16. The real split forms of the same complex algebras give rise to the so called almost Lagrangian contact geometries, cf. . ### 2.11. Example The previous two lists of geometries include those with most simple infinitesimal flag structures. The other extreme is provided by the real parabolic geometries with most complicated flags in each tangent space, i.e. those corresponding to the Borel subgroups $`PG`$. Here we need to cross out all nodes in the Satake diagram and so there must not be any black ones. Thus all real split forms, $`𝔰𝔲(p,p)`$, $`𝔰𝔬(\mathrm{}1,\mathrm{}+1)`$, and EII list all real forms which admit the right grading. ### 2.12. Natural bundles Consider a fixed parabolic geometry $`(𝒢,\omega )`$ over a manifold $`M`$. Then each $`P`$–module $`𝕍`$ defines the associated bundle $`VM=𝒢\times _P𝕍`$ over $`M`$. In fact, this is a functorial construction which may be restricted to all subcategories of parabolic geometries mentioned in 2.6. Similarly, we may treat bundles associated to any action $`P\mathrm{Diff}(𝕊)`$ on a manifold $`𝕊`$, the standard fiber for $`SM=𝒢\times _P𝕊`$. We shall meet only natural vector bundles defined by $`P`$–modules in this paper, however. There is a special class of natural (vector) bundles defined by $`G`$–modules $`𝕎`$. Such natural bundles are called tractor bundles, see for historical remarks. We shall distinguish them by the script letters here and often omit the base manifold $`M`$ from the notation. We may view each such tractor bundle $`𝒲M`$ as associated to the extended principal fiber bundle $`\stackrel{~}{𝒢}=𝒢\times _PG`$, i.e. $`𝒲=\stackrel{~}{𝒢}\times _G𝕎`$. Now, the Cartan connection $`\omega `$ on $`𝒢`$ extends uniquely to a principal connection form $`\stackrel{~}{\omega }`$ on $`\stackrel{~}{𝒢}`$, and so there is the induced linear connection on each such $`𝒲`$. With some more careful arguments, this construction may be extended to all $`(𝔤,P)`$–modules $`𝕎`$, i.e. $`P`$–modules with a fixed extension of the induced representation of $`𝔭`$ to a representation of $`𝔤`$ compatible with the $`P`$–action, see , Section 2. One of the achievements of the latter paper is the equivalent treatment of the regular parabolic geometries entirely within the framework of the tractor bundles, inclusive the discussion of the canonical connections. ### 2.13. Adjoint tractors It seems that the most important natural bundle is the adjoint tractor bundle $`𝒜=𝒢\times _P𝔤`$ with respect to the adjoint action $`\mathrm{Ad}`$ of $`G`$ on $`𝔤`$. The $`P`$–submodules $`𝔤^j𝔤`$ give rise to the filtration $$𝒜=𝒜^k𝒜^{k+1}\mathrm{}𝒜^0𝒜^1\mathrm{}𝒜^k$$ by the natural subbundles $`𝒜^j=𝒢\times _P𝔤^j`$. Moreover, the associated graded natural bundle (often denoted by the abuse of notation by the same symbol again) $$\mathrm{Gr}𝒜=𝒜_k\mathrm{}𝒜_1𝒜_0𝒜_1\mathrm{}𝒜_k$$ with $`𝒜_j=𝒜^j/𝒜^{j+1}`$ is available. By the very definition, there is the algebraic bracket on $`𝒜`$ defined by means of the Lie bracket in $`𝔤`$ (since the Lie bracket is $`\mathrm{Ad}`$-equivariant), which shows up on the graded bundle as $$\{,\}:𝒜_i\times 𝒜_j𝒜_{i+j}.$$ For the same reason, the Killing form defines a pairing on $`\mathrm{Gr}𝒜`$ such that $`𝒜_i^{}=𝒜_i`$, and the algebraic codifferential $`^{}`$, cf. 2.5, defines natural algebraic mappings $$^{}:^{k+1}𝒜^1𝒜^k𝒜^1𝒜$$ which are homogeneous of degree zero with respect to the gradings in $`\mathrm{Gr}𝒜`$. Similarly to the notation for $`𝔤`$, we also write $`𝒜_+=𝒜^1`$, $`𝒜_{}=𝒜/𝒜^0`$ for bundles associated either to $`𝒢`$ or $`𝒢_0`$. Thus $`𝒜=𝒜_{}+𝒜_0+𝒜_+`$, understood either as composition series induced by the filtration, or direct sum of invariant subbundles, respectively. ### 2.14. Tangent and cotangent bundles For each parabolic geometry $`(𝒢,\omega )`$, $`p:𝒢M`$, the absolute parallelism defines the identification $$𝒢\times _P(𝔤/𝔭)TM,𝒢\times 𝔤_{}(u,X)Tp(\omega ^1(X)(u)).$$ In other words, the tangent spaces $`TM`$ are natural bundles equipped with the filtrations which correspond to the Lie algebras $`𝔤_{}`$ viewed as the $`P`$–modules $`𝔤/𝔭`$ with the induced $`\mathrm{Ad}`$–actions. Equivalently, the tangent spaces are the quotients $$TM=𝒜/𝒜^0$$ of the adjoint tractor bundles. Therefore, the induced graded tangent spaces $`\mathrm{Gr}TM`$ are exactly the negative parts of the graded adjoint tractor bundles $$\mathrm{Gr}TM=𝒜_k\mathrm{}𝒜_1.$$ Moreover, the definition of the algebraic bracket on $`𝒜`$ implies immediately that the bracket induced by the Lie bracket of vector fields on $`\mathrm{Gr}TM`$ for regular infinitesimal flag structures on $`M`$ coincides with $`\{,\}`$. Now, the cotangent bundles clearly correspond to $$T^{}M=𝒢\times _P𝔭_+𝒜^1$$ and so the graded cotangent space is identified with $$\mathrm{Gr}T^{}M=𝒜_1\mathrm{}𝒜_k.$$ Finally, the pairing of a one–form and a vector field is given exactly by the canonical pairing of $`𝒜/𝒜^1`$ and $`𝒜^1`$ induced by the Killing form. ### 2.15. The first important observation about the adjoint tractors and their links to tangent and cotangent spaces is that the curvature $`K`$ of the parabolic geometry $`(𝒢,\omega )`$ is in fact a section of $`\mathrm{\Lambda }^2(𝒜/𝒜^0)^{}𝒜`$ whose frame form is the curvature function $`\kappa `$. Thus, the curvature is a two–form on the underlying manifold $`M`$ valued in the adjoint tractors and all the conditions on the curvature discussed in 2.6 are expressed by natural algebraic operations on the adjoint tractors. The remarkable relation of both tangent and cotangent spaces to the positive and negative parts of the adjoint tractors is the most important tool in what follows. In particular, let us notice already here that once we are given a reduction of the structure group $`P`$ of $`𝒢`$ to its reductive part $`G_0`$, the adjoint tractor bundles are identified with their graded versions and both tangent and cotangent bundles are embedded inside of $`𝒜`$. ## 3. Weyl–structures ### 3.1. Definition Let $`(p:𝒢M,\omega )`$ be a parabolic geometry on a smooth manifold $`M`$, and consider the underlying principal $`G_0`$–bundle $`p_0:𝒢_0M`$ and the canonical projection $`\pi :𝒢𝒢_0`$. A Weyl–structure for $`(𝒢,\omega )`$ is a global $`G_0`$–equivariant smooth section $`\sigma :𝒢_0𝒢`$ of $`\pi `$. ###### 3.2 Proposition. For any parabolic geometry $`(p:𝒢M,\omega )`$, there exists a Weyl–structure. Moreover, if $`\sigma `$ and $`\widehat{\sigma }`$ are two Weyl–structures, then there is a unique smooth section $`\mathrm{{\rm Y}}=(\mathrm{{\rm Y}}_1,\mathrm{},\mathrm{{\rm Y}}_k)`$ of $`𝒜_1\mathrm{}𝒜_k`$ such that $$\widehat{\sigma }(u)=\sigma (u)\mathrm{exp}(\mathrm{{\rm Y}}_1(u))\mathrm{}\mathrm{exp}(\mathrm{{\rm Y}}_k(u)).$$ Finally, each Weyl-structure $`\sigma `$ and section $`\mathrm{{\rm Y}}`$ define another Weyl-structure $`\widehat{\sigma }`$ by the above formula. ###### Proof. We can choose a finite open covering $`\{U_1,\mathrm{},U_N\}`$ of $`M`$ such that both $`𝒢`$ and $`𝒢_0`$ are trivial over each $`U_i`$. Since by Proposition 2.3 $`P`$ is the semidirect product of $`G_0`$ and $`P_+`$ it follows immediately that there are smooth $`G_0`$–equivariant sections $`\sigma _i:p_0^1(U_i)p^1(U_i)`$. Moreover, we can find open subsets $`V_i`$ such that $`\overline{V}_iU_i`$ and such that $`\{V_1,\mathrm{},V_N\}`$ still is a covering of $`M`$. Now from Proposition 2.3 and the Baker–Campbell–Hausdorff formula it follows that there is a smooth mapping $`\mathrm{\Psi }:p_0^1(U_1U_2)𝔭_+`$ such that $`\sigma _2(u)=\sigma _1(u)\mathrm{exp}(\mathrm{\Psi }(u))`$. Equivariance of $`\sigma _1`$ and $`\sigma _2`$ immediately implies that $`\mathrm{\Psi }(ug)=\mathrm{Ad}(g^1)(\mathrm{\Psi }(u))`$ for all $`gG_0`$. Now let $`f:M[0,1]`$ be a smooth function with support contained in $`U_2`$, which is identically one on $`V_2`$ and define $`\sigma :p_0^1(U_1V_2)p^1(U_1V_2)`$ by $`\sigma (u)=\sigma _1(u)\mathrm{exp}(f(p_0(u))\mathrm{\Psi }(u))`$ for $`uU_1`$ and by $`\sigma (u)=\sigma _2(u)`$ for $`uV_2`$. Then obviously these two definitions coincide on $`U_1V_2`$, so $`\sigma `$ is smooth. Moreover, from the equivariance of the $`\sigma _i`$ and of $`\mathrm{\Psi }`$ one immediately concludes that $`\sigma `$ is equivariant. Similarly, one extends the section next to $`U_1V_2V_3`$ and by induction one reaches a globally defined smooth equivariant section. If $`\widehat{\sigma }`$ and $`\sigma `$ are two global equivariant sections, then applying Proposition 2.3 directly, we see that there are smooth maps $`\mathrm{{\rm Y}}_i:𝒢_0𝔤_i`$ for $`i=1,\mathrm{},k`$ such that $`\widehat{\sigma }(u)=\sigma (u)\mathrm{exp}(\mathrm{{\rm Y}}_1(u))\mathrm{}\mathrm{exp}(\mathrm{{\rm Y}}_k(u))`$. As above, equivariance of $`\widehat{\sigma }`$ and $`\sigma `$ implies that $`\mathrm{{\rm Y}}_i(ug)=\mathrm{Ad}(g^1)(\mathrm{{\rm Y}}_i(u))`$ for all $`gG_0`$. Hence, $`\mathrm{{\rm Y}}_i`$ corresponds to a smooth section of $`𝒜_i`$. The last statement of the Proposition is obvious now. ∎ ### 3.3. Weyl connections We can easily relate a Weyl–structure $`\sigma :𝒢_0𝒢`$ to objects defined on the manifold $`M`$ by considering the pullback $`\sigma ^{}\omega `$ of the Cartan connection $`\omega `$ along the section $`\sigma `$. Clearly, $`\sigma ^{}\omega `$ is a $`𝔤`$–valued one–form on $`𝒢_0`$, which by construction is $`G_0`$–equivariant, i.e. $`(r^g)^{}(\sigma ^{}\omega )=\mathrm{Ad}(g^1)\sigma ^{}\omega `$ for all $`gG_0`$. Since $`\mathrm{Ad}(g^1)`$ preserves the grading of $`𝔤`$, in fact each component $`\sigma ^{}\omega _i`$ of $`\sigma ^{}\omega `$ is a $`G_0`$–equivariant one form with values in $`𝔤_i`$. Now consider a vertical tangent vector on $`𝒢_0`$, i.e. the value $`\zeta _A(u)`$ of a fundamental vector field corresponding to some $`A𝔤_0`$. Since $`\sigma `$ is $`G_0`$–equivariant, we conclude that $`T_u\sigma \zeta _A(u)=\zeta _A(\sigma (u))`$, where the second fundamental vector field is on $`𝒢`$. Consequently, we have $`\sigma ^{}\omega (\zeta _A)=\omega (\zeta _A)=A𝔤_0`$. Thus, for $`i0`$ the form $`\sigma ^{}\omega _i`$ is horizontal, while $`\sigma ^{}\omega _0`$ reproduces the generators of fundamental vector fields. From this observation, it follows immediately, that for $`i0`$, the form $`\sigma ^{}\omega _i`$ descends to a smooth one form on $`M`$ with values in $`𝒜_i`$, which we denote by the same symbol, while $`\sigma ^{}\omega _0`$ defines a principal connection on the bundle $`𝒢_0`$. This connection is called the Weyl connection of the Weyl structure $`\sigma `$. ### 3.4. Soldering forms and Rho-tensors We view the positive components of $`\sigma ^{}\omega `$ as a one–form $$\text{P}=\sigma ^{}(\omega _+)\mathrm{\Omega }^1(M;𝒜_1\mathrm{}𝒜_k)$$ with values in the bundle $`𝒜_1\mathrm{}𝒜_k`$. We call it the Rho–tensor of the Weyl–structure $`\sigma `$. This is a generalization of the tensor $`\text{P}_{ab}`$ well known in conformal geometry. Since $`\omega `$ restricts to a linear isomorphism in each tangent space of $`𝒢`$, we see that the form $$\sigma ^{}\omega _{}=(\sigma ^{}\omega _k,\mathrm{},\sigma ^{}\omega _1)\mathrm{\Omega }^1(M,𝒜_k\mathrm{}𝒜_1)$$ induces an isomorphism $$TM𝒜_k\mathrm{}𝒜_1\mathrm{Gr}TM.$$ We will denote this isomorphism by $$\xi (\xi _k,\mathrm{},\xi _1)𝒜_k\mathrm{}𝒜_1$$ for $`\xi TM`$. In particular, each fixed $`u𝒢_0`$ provides the identification of $`T_{p_0(u)}M𝔤_{}`$ compatible with the grading. Thus, the choice of a Weyl structure $`\sigma `$ provides a reduction of the structure group of $`TM`$ to $`G_0`$ (by means of the soldering form $`\sigma ^{}\omega _{}`$ on $`𝒢_0`$), the linear connection on $`M`$ (the Weyl connection $`\sigma ^{}\omega _0`$), and the Rho–tensor P. ### 3.5. Remarks As discussed in 2.72.8 above, there is the underlying frame form of length one on $`𝒢_0`$ which is the basic structure from which the whole parabolic geometry $`(𝒢,\omega )`$ may be reconstructed, with exceptions mentioned explicitly in 2.9 and 2.10. By definition, for $`i<0`$ and $`\xi T^i𝒢_0`$ this frame form can be computed by choosing any lift of $`\xi `$ to a tangent vector on $`𝒢`$ and then taking the $`𝔤_i`$–component of the value of $`\omega `$ on this lift. In particular, we can use $`T\sigma \xi `$ as the lift, which implies that the restriction of $`\sigma ^{}\omega _i`$ (viewed as a form on $`𝒢_0`$) to $`T^i𝒢_0`$ coincides with the $`𝔤_i`$–component of the frame form of length one. This in turn implies that the restriction of $`\sigma ^{}\omega _i`$ (viewed as a form on $`M`$) to $`T^iM`$ coincides with the canonical projection $`T^iM𝒜_i=T^iM/T^{i+1}M`$. There is also another interpretation of the objects on $`M`$ induced by the choice of a Weyl–structure that will be very useful in the sequel. Namely, consider the form $$\sigma ^{}\omega _0=\sigma ^{}\omega _k\mathrm{}\sigma ^{}\omega _0\mathrm{\Omega }^1(𝒢_0,𝔤_k\mathrm{}𝔤_0).$$ We have seen above that this form is $`G_0`$–equivariant, it reproduces the generators of fundamental vector fields, and restricted to each tangent space, it is a linear isomorphism. Thus $`\sigma ^{}\omega _0`$ defines a Cartan connection on the principal $`G_0`$–bundle $`p_0:𝒢_0M`$. In the case of the irreducible parabolic geometries, these connections are classical affine connections on the tangent space $`TM`$ belonging to its reduced structure group $`G_0`$. ### 3.6. Bundles of scales As we have seen in 3.3, 3.4 above, choosing a Weyl–structure $`\sigma :𝒢_0𝒢`$ leads to several objects on the manifold $`M`$. Now the next step is to show that in fact a small part of these data is sufficient to completely fix the Weyl–structure. More precisely, we shall see below that even the linear connections induced by the Weyl connection $`\sigma ^{}\omega _0`$ on certain oriented line bundles suffice to pin down the Weyl–structure. Equivalently, one can use the corresponding frame bundles, which are principal bundles with structure group $`^+`$. The principal bundles appropriate for this purpose are called bundles of scales. To define these bundles, we have to make a few observations: A principal $`^+`$–bundle associated to $`𝒢_0`$ is determined by a homomorphism $`\lambda :G_0^+`$. The derivative of this homomorphism is a linear map $`\lambda ^{}:𝔤_0`$. Now $`𝔤_0`$ splits as the direct sum $`𝔷(𝔤_0)𝔤_0^{ss}`$ of its center and its semisimple part, and $`\lambda ^{}`$ automatically vanishes on the semisimple part. Moreover, as discussed in 2.2 the restriction of the Killing form $`B`$ of $`𝔤`$ to the subalgebra $`𝔤_0`$ is non–degenerate, and one easily verifies that this restriction respects the above splitting. In particular, the restriction of $`B`$ to $`𝔷(𝔤_0)`$ is still non–degenerate and thus there is a unique element $`E_\lambda 𝔷(𝔤_0)`$ such that $`\lambda ^{}(A)=B(E_\lambda ,A)`$ for all $`A𝔤_0`$. Next, the action of the element $`E_\lambda 𝔷(𝔤_0)`$ on any $`G_0`$–irreducible representation commutes with the action of $`G_0`$, and thus is given by a scalar multiple of the identity by Schur’s lemma. ### Definition An element $`E_\lambda `$ of $`𝔷(𝔤_0)`$ is called a scaling element if and only if $`E_\lambda `$ acts by a nonzero real scalar on each $`G_0`$–irreducible component of $`𝔭_+`$. A bundle of scales is a principal $`^+`$ bundle $`^\lambda M`$ which is associated to $`𝒢_0`$ via a homomorphism $`\lambda :G_0^+`$, whose derivative is given by $`\lambda ^{}(A)=B(E_\lambda ,A)`$ for some scaling element $`E_\lambda 𝔷(𝔤_0)`$. Having given a fixed choice of a bundle $`^\lambda `$ of scales, a (local) scale on $`M`$ is a (local) smooth section of $`^\lambda `$. ###### 3.7 Proposition. Let $`G`$ be a fixed semisimple Lie group, whose Lie algebra $`𝔤`$ is endowed with a $`|k|`$–grading. Then the following holds: (1) There are scaling elements in $`𝔷(𝔤_0)`$. (2) Any scaling element $`E_\lambda 𝔷(𝔤_0)`$ gives rise to a canonical bundle $`^\lambda `$ of scales over each manifold endowed with a parabolic geometry of the given type. (3) Any bundle of scales admits global smooth sections, i.e. there always exist global scales. ###### Proof. (1) The grading element $`E𝔷(𝔤_0)`$, cf. 2.2, acts on $`𝔤_i`$ by multiplication with $`i`$, so it is a scaling element. More generally, one can consider the subspace of $`𝔷(𝔤_0)`$ of all elements which act by real scalars on each irreducible component of $`𝔭_+`$. Then each irreducible component determines a real valued functional and thus a hyperplane in that space, and the complement of these finitely many hyperplanes (which is open and dense) consists entirely of scaling elements. (2) Let $`𝔭_+=𝔭^\alpha `$ be the decomposition of $`𝔭_+`$ into $`G_0`$–irreducible components, and for a fixed grading element $`E_\lambda `$ denote by $`a_\alpha `$ the scalar by which $`E_\lambda `$ acts on $`𝔭^\alpha `$. The adjoint action defines a smooth homomorphism $`G_0_\alpha GL(𝔭^\alpha )`$, whose components we write as $`g\mathrm{Ad}^\alpha (g)`$. Then consider the homomorphism $`\lambda :G_0^+`$ defined by $$\lambda (g):=\underset{\alpha }{}|det(\mathrm{Ad}^\alpha (g))|^{2a_\alpha }.$$ The derivative of this homomorphism is given by $`\lambda ^{}(A)=_\alpha 2a_\alpha \text{tr}(\mathrm{ad}(A)|_{𝔭^\alpha })`$. Now $`𝔤_{}=_\alpha (𝔭^\alpha )^{}`$, and $`E_\lambda `$ acts on $`(𝔭^\alpha )^{}`$ by $`a_\alpha `$ and on $`𝔤_0`$ by zero, and thus $`B(E_\lambda ,A)=\mathrm{tr}(\mathrm{ad}(A)\mathrm{ad}(E_\lambda ))=_\alpha a_\alpha \mathrm{tr}(\mathrm{ad}(A)|_{𝔭^\alpha })_\alpha a_\alpha \mathrm{tr}(\mathrm{ad}(A)|_{(𝔭^\alpha )^{}})=\lambda ^{}(A)`$. (3) This is just due to the fact that orientable real line bundles and thus principal $`^+`$–bundles are automatically trivial and hence admit global smooth sections. ∎ ###### 3.8 Lemma. Let $`\sigma :𝒢_0𝒢`$ be a Weyl–structure for parabolic geometry $`(𝒢M,\omega )`$ and let $`^\lambda `$ be a bundle of scales. (1) The Weyl connection $`\sigma ^{}\omega _0\mathrm{\Omega }_1(𝒢_0,𝔤_0)`$ induces a principal connection on the bundle of scales $`^\lambda `$. (2) $`^\lambda `$ is naturally identified with $`𝒢_0/\mathrm{ker}(\lambda )`$, the orbit space of the free right action of the normal subgroup $`\mathrm{ker}(\lambda )G_0`$ on $`𝒢_0`$. (3) The form $`\lambda ^{}\sigma ^{}\omega _0\mathrm{\Omega }^1(𝒢_0)`$ descends to the connection form of the induced principal connection on $`^\lambda =𝒢_0/\mathrm{ker}(\lambda )`$. (4) The composition of $`\lambda ^{}`$ with the curvature form of $`\sigma ^{}\omega _0`$ descends to the curvature of the induced connection on $`^\lambda `$. ###### Proof. All claims are straightforward consequences of the definitions. ∎ To see that the Weyl–structure $`\sigma `$ is actually uniquely determined by the induced principal connection on $`^\lambda `$ (cf. Theorem 3.12 below), we have to compute how the principal connection $`\sigma ^{}\omega _0`$ changes when we change $`\sigma `$. For later use, we also compute how the other objects induced by $`\sigma `$ change under the change of the Weyl–structures. So let us assume that $`\widehat{\sigma }`$ is another Weyl–structure and $`\mathrm{{\rm Y}}=(\mathrm{{\rm Y}}_1,\mathrm{},\mathrm{{\rm Y}}_k)`$ is the section of $`𝒜_1\mathrm{}𝒜_k`$ characterized by $`\widehat{\sigma }(u)=\sigma (u)\mathrm{exp}(\mathrm{{\rm Y}}_1(u))\mathrm{}\mathrm{exp}(\mathrm{{\rm Y}}_k(u))`$. We shall use the convention that we simply denote quantities corresponding to $`\widehat{\sigma }`$ by hatted symbols and quantities corresponding to $`\sigma `$ by unhatted symbols. Consequently, $`(\xi _k,\mathrm{},\xi _1)`$ and $`(\widehat{\xi }_k,\mathrm{},\widehat{\xi }_1)`$ denote the splitting of $`\xi TM`$ according to $`\sigma `$, respectively $`\widehat{\sigma }`$, and P and $`\widehat{\text{P}}`$ are the Rho–tensors. Finally, let us consider any vector bundle $`E`$ associated to the principal bundle $`𝒢_0`$. Then for any Weyl–structure the corresponding principal connection on $`𝒢_0`$ induces a linear connection on $`E`$, which is denoted by $``$ for $`\sigma `$ and by $`\widehat{}`$ for $`\widehat{\sigma }`$. To write the formulae efficiently, we need some further notation. By $`\underset{¯}{j}`$ we denote a sequence $`(j_1,\mathrm{},j_k)`$ of nonnegative integers, and we put $`\underset{¯}{j}=j_1+2j_2+\mathrm{}+kj_k`$. Moreover, we define $`\underset{¯}{j}!=j_1!\mathrm{}j_k!`$ and $`(1)^{\underset{¯}{j}}=(1)^{j_1+\mathrm{}+j_k}`$, and we define $`(\underset{¯}{j})_m`$ to be the subsequence $`(j_1,\mathrm{},j_m)`$ of $`\underset{¯}{j}`$. By $`0`$ we denote sequences of any length consisting entirely of zeros. ###### 3.9 Proposition. Let $`\sigma `$ and $`\widehat{\sigma }`$ be two Weyl–structures related by $$\widehat{\sigma }(u)=\sigma (u)\mathrm{exp}(\mathrm{{\rm Y}}_1(u))\mathrm{}\mathrm{exp}(\mathrm{{\rm Y}}_k(u)),$$ where $`\mathrm{{\rm Y}}=(\mathrm{{\rm Y}}_1,\mathrm{},\mathrm{{\rm Y}}_k)`$ is a smooth section of $`𝒜_1\mathrm{}𝒜_k`$. Then we have: (1) $`\widehat{\xi }_i`$ $`={\displaystyle \underset{\underset{¯}{j}+\mathrm{}=i}{}}{\displaystyle \frac{(1)^{\underset{¯}{j}}}{\underset{¯}{j}!}}\mathrm{ad}(\mathrm{{\rm Y}}_k)^{j_k}\mathrm{}\mathrm{ad}(\mathrm{{\rm Y}}_1)^{j_1}(\xi _{\mathrm{}}),`$ (2) $`\widehat{\text{P}}_i(\xi )`$ $`={\displaystyle \underset{\underset{¯}{j}+\mathrm{}=i}{}}{\displaystyle \frac{(1)^{\underset{¯}{j}}}{\underset{¯}{j}!}}\mathrm{ad}(\mathrm{{\rm Y}}_k)^{j_k}\mathrm{}\mathrm{ad}(\mathrm{{\rm Y}}_1)^{j_1}(\xi _{\mathrm{}})+`$ $`{\displaystyle \underset{\underset{¯}{j}+\mathrm{}=i}{}}{\displaystyle \frac{(1)^{\underset{¯}{j}}}{\underset{¯}{j}!}}\mathrm{ad}(\mathrm{{\rm Y}}_k)^{j_k}\mathrm{}\mathrm{ad}(\mathrm{{\rm Y}}_1)^{j_1}(\text{P}_{\mathrm{}}(\xi ))+`$ $`{\displaystyle \underset{m=1}{\overset{k}{}}}{\displaystyle \underset{\begin{array}{c}(\underset{¯}{j})_{m1}=0\\ m+\underset{¯}{j}=i\end{array}}{}}{\displaystyle \frac{(1)^{\underset{¯}{j}}}{(j_m+1)\underset{¯}{j}!}}\mathrm{ad}(\mathrm{{\rm Y}}_k)^{j_k}\mathrm{}\mathrm{ad}(\mathrm{{\rm Y}}_m)^{j_m}(_\xi \mathrm{{\rm Y}}_m),`$ (3) $`\widehat{}_\xi s`$ $`=_\xi s+{\displaystyle \underset{\underset{¯}{j}+\mathrm{}=0}{}}{\displaystyle \frac{(1)^{\underset{¯}{j}}}{\underset{¯}{j}!}}(\mathrm{ad}(\mathrm{{\rm Y}}_k)^{j_k}\mathrm{}\mathrm{ad}(\mathrm{{\rm Y}}_1)^{j_1}(\xi _{\mathrm{}}))s,`$ where $``$ denotes the map $`𝒜_0\times EE`$ induced by the action of $`𝔤_0`$ on the standard fiber of $`E`$. ###### Proof. The essential part of the proof is to compute the tangent map $`T_u\widehat{\sigma }`$ in a point $`u𝒢_0`$. By definition, $`\widehat{\sigma }(u)=\sigma (u)\mathrm{exp}(\mathrm{{\rm Y}}_1(u))\mathrm{}\mathrm{exp}(\mathrm{{\rm Y}}_k(u))`$. Thus, we can write the evaluation of the tangent map, $`T_u\widehat{\sigma }\xi `$, as the sum of $`T_{\sigma (u)}r^gT_u\sigma \xi `$, where $`g=\mathrm{exp}(\mathrm{{\rm Y}}_1(u))\mathrm{}\mathrm{exp}(\mathrm{{\rm Y}}_k(u))P_+`$, and the derivative at $`t=0`$ of $$\sigma (u)\mathrm{exp}(\mathrm{{\rm Y}}_1(c(t)))\mathrm{}\mathrm{exp}(\mathrm{{\rm Y}}_k(c(t))),$$ where $`c:𝒢_0`$ is a smooth curve with $`c(0)=u`$ and $`c^{}(0)=\xi `$. By construction, the latter derivative lies in the kernel of $`T\pi `$, where $`\pi :𝒢𝒢_0`$ is the projection, so we can write it as $`\zeta _{\mathrm{\Phi }(\xi )}(\widehat{\sigma }(u))`$ for suitable $`\mathrm{\Phi }(\xi )𝔭_+`$. Now, for $`\xi T_u𝒢_0`$, we have $`\widehat{\sigma }^{}\omega (\xi )=\omega (\widehat{\sigma }(u))(T_u\widehat{\sigma }\xi )`$. By equivariance of the Cartan connection $`\omega `$, we get $`\omega (\sigma (u)g)(Tr^gT\sigma \xi )=\mathrm{Ad}(g^1)(\omega (u)(T\sigma \xi ))`$. Consequently, $$\widehat{\sigma }^{}\omega (\xi )=\mathrm{Ad}(g^1)(\sigma ^{}\omega (\xi ))+\mathrm{\Phi }(\xi ).$$ Since $`\mathrm{\Phi }(\xi )𝔭_+`$, this term affects only the transformation of the Rho–tensor, and does not influence the changes of $`\sigma ^{}\omega _i`$ for $`i0`$. In particular, for the components $`\widehat{\sigma }^{}\omega _i`$ with $`i<0`$, we only have to take the part of the right degree in (4) $$e^{\mathrm{ad}(\mathrm{{\rm Y}}_k(u))}\mathrm{}e^{\mathrm{ad}(\mathrm{{\rm Y}}_1(u))}(\sigma ^{}\omega (u)(\xi )),$$ and expanding the exponentials, this immediately leads to formula (1). To compute the change in the connection, we have to notice that $`\widehat{\sigma }^{}\omega _0(\xi )`$ is the component of degree zero in (4) above. Consequently, if we apply $`\widehat{\sigma }^{}\omega _0`$ to the horizontal lift of a tangent vector on $`M`$, the outcome is just this degree zero part. Otherwise put, the horizontal lift with respect to $`\widehat{\sigma }^{}\omega _0`$ is obtained by subtracting the fundamental vector field corresponding to the degree zero part of (4) from the horizontal lift with respect to $`\sigma ^{}\omega _0`$. Applying such horizontal vector field to a smooth $`G_0`$–equivariant function with values in any $`G_0`$–representation and taking into account that a fundamental vector fields acts on such functions by the negative of its generator acting on the values, this immediately leads to formula (3) by expanding the exponentials. Finally, we have to deal with the change of the Rho–tensor. Recall that we view this as a tensor on the manifold $`M`$, so we can compute $`\widehat{\text{P}}_i(\xi )`$ by applying $`\widehat{\sigma }^{}\omega _i`$ to any lift of $`\xi `$. In particular, we may use the horizontal lift $`\xi ^h`$ with respect to $`\sigma ^{}\omega _0`$, so we may assume $`\sigma ^{}\omega _0(\xi )=0`$. But then expanding the exponentials in (4) and taking the part of degree $`i`$ we see that we exactly get the first two summands in formula (2). Thus we are left with proving that the last summand corresponds to $`\mathrm{\Phi }(\xi )`$. For this aim, let us rewrite the curve that we have to differentiate as $$\widehat{\sigma }(u)\mathrm{exp}(\mathrm{{\rm Y}}_k(u))\mathrm{}\mathrm{exp}(\mathrm{{\rm Y}}_1(u))\mathrm{exp}(\mathrm{{\rm Y}}_1(c(t)))\mathrm{}\mathrm{exp}(\mathrm{{\rm Y}}_k(c(t))).$$ Differentiating this using the product rule we get a sum of terms in which one $`\mathrm{{\rm Y}}_i`$ is differentiated, while all others have to be evaluated at $`t=0`$, i.e. in $`u`$. So each of these terms reads as the derivative at $`t=0`$ of $$\widehat{\sigma }(u)\text{conj}_{\mathrm{exp}(\mathrm{{\rm Y}}_k(u))}\mathrm{}\text{conj}_{\mathrm{exp}(\mathrm{{\rm Y}}_{i+1}(u))}\left(\mathrm{exp}(\mathrm{{\rm Y}}_i(u))\mathrm{exp}(\mathrm{{\rm Y}}_i(c(t)))\right),$$ where $`\text{conj}_g`$ denotes the conjugation by $`g`$, i.e. the map $`hghg^1`$. This expression is just the principal right action by the value of a smooth curve in $`P`$ which maps zero to the unit element, so its result is exactly the value at $`\widehat{\sigma }(u)`$ of the fundamental vector field generated by the derivative at zero of this curve. This derivative is clearly obtained by applying $$e^{\mathrm{ad}(\mathrm{{\rm Y}}_k(u))}\mathrm{}e^{\mathrm{ad}(\mathrm{{\rm Y}}_{i+1}(u))}$$ to the derivative at zero of $`t\mathrm{exp}(\mathrm{{\rm Y}}_i(u))\mathrm{exp}(\mathrm{{\rm Y}}_i(c(t)))`$. By , 4.26, and the chain rule, the latter derivative equals the left logarithmic derivative of $`\mathrm{exp}`$ applied to the derivative at zero of $`t\mathrm{{\rm Y}}_i(c(t))`$. Moreover, the proof of , Lemma 4.27, can be easily adapted to the left logarithmic derivative, showing that this gives $$\underset{p=0}{\overset{\mathrm{}}{}}\frac{(1)^p}{(p+1)!}\mathrm{ad}(\mathrm{{\rm Y}}_i(u))^p(\xi ^h\mathrm{{\rm Y}}_i).$$ Finally, we have to observe that $`\xi ^h\mathrm{{\rm Y}}_i`$ corresponds to $`_\xi \mathrm{{\rm Y}}_i`$ and to sort out the terms of the right degree in order to get the remaining summand in (2). ∎ ### 3.10. Example For all irreducible parabolic geometries, the formulae from Proposition 3.9 become extremely simple. In fact they coincide completely with the known ones in the conformal Riemannian geometry: The grading of $`TM`$ is trivial, the connection transforms as $$\widehat{}_\xi s=_\xi s\{\mathrm{{\rm Y}},\xi \}s,$$ where $`\mathrm{{\rm Y}}`$ is a section of $`𝒜_1=T^{}M`$, and the bracket of $`\mathrm{{\rm Y}}`$ and $`\xi `$ is a field of endomorphisms of $`TM`$ acting on $`s`$ in an obvious way. Indeed, there are no more terms on the right–hand side of 3.9(3) which make sense. Next, the Rho–tensor transforms as $$\widehat{\text{P}}(\xi )=\text{P}(\xi )+_\xi \mathrm{{\rm Y}}+\frac{1}{2}\{\mathrm{{\rm Y}},\{\mathrm{{\rm Y}},\xi \}\}.$$ The formulae for the $`|2|`$–graded examples are a bit more complicated. The splitting of $`TM`$ and the connection and Rho–tensors change as follows $`\widehat{\xi }_2=`$ $`\xi _2`$ $`\widehat{\xi }_1=`$ $`\xi _1\{\mathrm{{\rm Y}}_1,\xi _2\}`$ $`\widehat{}_\xi s=`$ $`_\xi s+(\frac{1}{2}\{\mathrm{{\rm Y}}_1,\{\mathrm{{\rm Y}}_1,\xi _2\}\}\{\mathrm{{\rm Y}}_2,\xi _2\}\{\mathrm{{\rm Y}}_1,\xi _1\})s,`$ $`\widehat{\text{P}}_1(\xi )=`$ $`\text{P}_1(\xi )\frac{1}{6}\{\mathrm{{\rm Y}}_1,\{\mathrm{{\rm Y}}_1,\{\mathrm{{\rm Y}}_1,\xi _2\}\}\}+\{\mathrm{{\rm Y}}_2,\{\mathrm{{\rm Y}}_1,\xi _2\}\}+`$ $`\frac{1}{2}\{\mathrm{{\rm Y}}_1,\{\mathrm{{\rm Y}}_1,\xi _1\}\}\{\mathrm{{\rm Y}}_2,\xi _1\}+_\xi \mathrm{{\rm Y}}_1`$ $`\widehat{\text{P}}_2(\xi )=`$ $`\text{P}_2(\xi )\{\mathrm{{\rm Y}}_1,\text{P}_1(\xi )\}+_\xi \mathrm{{\rm Y}}_2\frac{1}{2}\{\mathrm{{\rm Y}}_1,_\xi \mathrm{{\rm Y}}_1\}+`$ $`\frac{1}{24}\mathrm{ad}(\mathrm{{\rm Y}}_1)^4(\xi _2)\frac{1}{2}\{\mathrm{{\rm Y}}_2,\{\mathrm{{\rm Y}}_1,\{\mathrm{{\rm Y}}_1,\xi _2\}\}\}+\frac{1}{2}\{\mathrm{{\rm Y}}_2,\{\mathrm{{\rm Y}}_2,\xi _2\}\}`$ $`\frac{1}{6}\mathrm{ad}(\mathrm{{\rm Y}}_1)^3(\xi _1)+\{\mathrm{{\rm Y}}_2,\{\mathrm{{\rm Y}}_1,\xi _1\}\}.`$ ### 3.11. Remark In applications, one is often interested in questions about the dependence of some objects on the choice of the Weyl–structures and then the infinitesimal form of the available change of the splittings, Rho’s and connections is important. In our terms, this amounts to sorting out the terms in formulae 3.9(1)–(3) which are linear in upsilons. Thus, the infinitesimal version of Proposition 3.9 for the variations $`\delta \xi _i`$, $`\delta `$, and $`\delta \text{P}_i`$ reads (1) $`\delta \xi _i=`$ $`\{\mathrm{{\rm Y}}_1,\xi _{i1}\}\mathrm{}\{\mathrm{{\rm Y}}_{k+i},\xi _k\}`$ (2) $`\delta \text{P}_i(\xi )=`$ $`_\xi \mathrm{{\rm Y}}_i\{\mathrm{{\rm Y}}_1,\text{P}_{i1}(\xi )\}\mathrm{}\{\mathrm{{\rm Y}}_{i1},\text{P}_1(\xi )\}`$ $`\{\mathrm{{\rm Y}}_{i+1},\xi _1\}\mathrm{}\{\mathrm{{\rm Y}}_k,\xi _{k+i}\}`$ (3) $`\delta _\xi s=`$ $`(\{\mathrm{{\rm Y}}_1,\xi _1\}+\mathrm{}+\{\mathrm{{\rm Y}}_k,\xi _k\})s.`$ ### 3.12. Proposition 3.9 not only allows us to show that a Weyl–structure is uniquely determined by the induced connection on any bundle of scales, but it also leads to a description of the Cartan bundle $`p:𝒢M`$. To get this description, recall that for any principal bundle $`EM`$ there is a bundle $`QEM`$ whose sections are exactly the principal connections on $`E`$, see , 17.4. ###### Theorem. Let $`p:𝒢M`$ be a parabolic geometry on $`M`$, and let $`^\lambda M`$ be a bundle of scales. (1) Each Weyl–structure $`\sigma :𝒢_0𝒢`$ determines the principal connection on $`^\lambda `$ induced by the Weyl connection $`\sigma ^{}\omega _0`$. This defines a bijective correspondence between the set of Weyl–structures and the set of principal connections on $`^\lambda `$. (2) There is a canonical isomorphism $`𝒢p_0^{}Q^\lambda `$, where $`p_0:𝒢_0M`$ is the projection. Under this isomorphism, the choice of a Weyl structure $`\sigma :𝒢_0𝒢`$ is the pullback of the principal connection on the bundle of scales $`^\lambda `$, viewed as a section $`MQ^\lambda `$. Moreover, the principal action of $`G_0`$ is the canonical action on $`p_0^{}Q^\lambda `$ induced from the action on $`𝒢_0`$, while the action of $`P_+`$ is described by equation (3) from Proposition 3.9. ###### Proof. (1) Consider the map $`\lambda ^{}:𝔤_0`$ defining the bundle $`^\lambda `$ of scales. Take elements $`Z𝔭_+`$ and $`X𝔤_{}`$, and consider $`\lambda ^{}([Z,X])`$. By assumption, this is given by $`B(E_\lambda ,[Z,X])=B([E_\lambda ,Z],X)`$ for some scaling element $`E_\lambda 𝔷(𝔤_0)`$. Hence if we assume that $`Z`$ lies in a $`G_0`$–irreducible component of $`𝔭_+`$ this is just a nonzero real multiple of $`B(Z,X)`$. In particular, this implies that for each $`0Z𝔭_i`$, we can find an element $`X𝔤_i`$, such that $`\lambda ^{}([Z,X])0`$. Moreover, since $`E_\lambda 𝔷(𝔤_0)`$ we get $`\mathrm{Ad}(g)(E_\lambda )=E_\lambda `$ for all $`gG_0`$ and this immediately implies that mapping $`Z𝔤_i`$ to $`X\lambda ^{}([Z,X])`$ induces an isomorphism $`𝔤_i𝔤_i^{}`$ of $`G_0`$–modules. To prove (1), we may as well use the induced linear connection on the line bundle $`L^\lambda =^\lambda \times _^+`$ corresponding to the standard representation. For this bundle, the map $``$ from Proposition 3.9 is clearly given by $`(As)(x)=\lambda ^{}(A(x))s(x)`$, where we denote by $`\lambda ^{}:𝒜_0M\times `$ also the mapping induced by $`\lambda ^{}:𝔤_0`$. We first claim that the map from Weyl–structures to linear connections is injective. So assume that $`\sigma `$ and $`\widehat{\sigma }`$ induce the same linear connection on $`L^\lambda `$ and let $`\mathrm{{\rm Y}}`$ be the section of $`𝒜_1\mathrm{}𝒜_k`$ describing the change from $`\sigma `$ to $`\widehat{\sigma }`$. For $`\xi T^1M`$, we have $`\xi _i=0`$ for all $`i<1`$, hence formula (3) of 3.9 reduces to $`\widehat{}_\xi s=_\xi s+\lambda ^{}(\{\mathrm{{\rm Y}}_1,\xi \})s`$ in this case. If $`\mathrm{{\rm Y}}_1`$ would be nonzero, then by the above argument we could find $`\xi `$ such that $`\lambda ^{}(\{\mathrm{{\rm Y}}_1,\xi \})0`$, which would contradict $`\widehat{}=`$, so $`\mathrm{{\rm Y}}_1`$ must be identically zero. But then for $`\xi `$ in $`T^2M`$, the change reduces to $`\widehat{}_\xi s=_\xi s+\lambda ^{}(\{\mathrm{{\rm Y}}_2,\xi _2\})s`$ and as above, we conclude that $`\mathrm{{\rm Y}}_2`$ is identically zero. Inductively, we get $`\mathrm{{\rm Y}}=0`$ and thus $`\widehat{\sigma }=\sigma `$. To see surjectivity, assume that $`\widehat{}`$ is any linear connection on $`L^\lambda `$, and let $`\sigma `$ be any Weyl–structure with induced linear connection $``$ on $`L^\lambda `$. Then there is a one–form $`\tau \mathrm{\Omega }^1(M)`$ such that $`\widehat{}_\xi s=_\xi s+\tau (\xi )s`$. Restricting $`\tau `$ to $`T^1M`$, we can find a unique smooth section $`\mathrm{{\rm Y}}_1`$ of $`𝒜_1`$ such that $`\tau (\xi )=\lambda ^{}(\{\mathrm{{\rm Y}}_1,\xi \})`$ for all $`\xi `$ in $`T^1M`$. Next, consider the map $`T^2MM\times `$ given by $$\xi \tau (\xi )+\lambda ^{}(\{\mathrm{{\rm Y}}_1,\xi _1\})\frac{1}{2}\lambda ^{}(\{\mathrm{{\rm Y}}_1,\{\mathrm{{\rm Y}}_1,\xi _2\}\}),$$ where the $`\xi _i`$ are the components of $`\xi `$ with respect to the Weyl–structure $`\sigma `$. By construction, this vanishes on $`T^1M`$, so it factors to a map defined on $`𝒜_2`$, and thus there is a unique section $`\mathrm{{\rm Y}}_2`$ of $`𝒜_2`$ such that it equals $`\lambda ^{}(\{\mathrm{{\rm Y}}_2,\xi _2\})`$. Inductively, we find a section $`\mathrm{{\rm Y}}`$ such that the Weyl–structure $`\widehat{\sigma }`$ corresponding to $`\sigma `$ and $`\mathrm{{\rm Y}}`$ induces the linear connection $`\widehat{}`$, cf. formula 3.9(3). (2) Consider any point $`u𝒢`$. Proposition 3.2 implies that there is a Weyl–structure $`\sigma :𝒢_0𝒢`$ such that $`u=\sigma (\pi (u))`$. If $``$ is the linear connection on $`L^\lambda `$ induced by $`\sigma `$, then we see from Proposition 3.9 that the value of $`_\xi s(p(u))`$ for a vector field $`\xi `$ on $`M`$ and a section $`s`$ of $`L^\lambda `$ depends only on $`\sigma (p(u))`$, since its change under a change of the Weyl–structure depends only on the value of $`\mathrm{{\rm Y}}`$ in $`p(u)`$. Thus, mapping $`u`$ to the value at $`p(u)`$ of the principal connection on $`^\lambda `$ induced by $`\sigma ^{}\omega _0`$ is independent of the choice of $`\sigma `$, so we get a well defined bundle map from the bundle $`𝒢𝒢_0`$ to the bundle $`Q^\lambda M`$ covering the projection $`p_0:𝒢_0M`$. Moreover, from part (1) of this proof it follows that this map induces isomorphisms in each fiber, so it leads to an isomorphism $`𝒢p_0^{}Q^\lambda `$ of bundles over $`𝒢_0`$. Obviously, the $`G_0`$–equivariant sections of $`p_0^{}Q^\lambda 𝒢_0`$ correspond exactly to the induced principal connections on $`^\lambda `$, i.e. the sections of $`Q^\lambda M`$. In order to describe the principal action of $`P`$ on $`p_0^{}Q^\lambda `$ obtained by the above isomorphism, one just has to note that for $`u𝒢_0`$ and $`gG_0`$ the fibers of $`p_0^{}Q^\lambda `$ over $`u`$ and $`ug`$ are canonically isomorphic since $`p_0(u)=p_0(ug)`$. Thus, the principal right action of $`G_0`$ is simply given by acting on $`𝒢_0`$. On the other hand, fix $`u𝒢_0`$ and an element $`\mathrm{exp}(Z_1)\mathrm{}\mathrm{exp}(Z_k)P_+`$ for $`Z_i𝔤_i`$. Via $`u`$, the element $`Z_i`$ corresponds to an element $`\mathrm{{\rm Y}}_i𝒜_i`$ at the point $`p_0(u)`$. Then the principal right action of $`P_+`$ is described by the formula (3) of Proposition 3.9 as required. ∎ ### 3.13. Closed and exact Weyl–structures Let us fix a bundle of scales $`^\lambda `$ for some parabolic geometry. Then the bijective correspondence between Weyl–structures and principal connections on $`^\lambda `$ immediately leads to two distinguished subclasses of Weyl–structures. Namely, we call a Weyl–structure $`\sigma :𝒢_0𝒢`$ closed, if the induced principal connection on $`^\lambda `$ (or equivalently the induced linear connection $``$ on $`L^\lambda `$) is flat. Moreover, by Proposition 3.7 the bundle $`^\lambda `$ of scales admits global smooth sections, and any such section gives rise to a flat principal connection on $`^\lambda `$ (which in addition has trivial holonomy) and hence to a closed Weyl–structure. The closed Weyl–structures induced by such global sections are called exact. Note that in the case of conformal structures, the canonical choice for the bundle of scales is simply the $`^+`$–bundle whose smooth sections are the metrics in the conformal class. Thus, the exact Weyl–structures in conformal geometry correspond exactly to the Levi–Civita connections of the metrics in the conformal class. The reason for the names “closed” and “exact” becomes apparent, once one studies the affine structures on the sets of closed and exact Weyl–structures. So let us assume that $`\sigma `$ is a closed Weyl–structure, and consider any other Weyl–structure $`\widehat{\sigma }`$ corresponding to the section $`\mathrm{{\rm Y}}=(\mathrm{{\rm Y}}_1,\mathrm{},\mathrm{{\rm Y}}_k)`$ of $`𝒜_1\mathrm{}𝒜_k`$. Now we can reinterpret theorem 3.12(1) together with proposition 3.9 as showing that the set of Weyl–structures is an affine space over $`\mathrm{\Omega }^1(M)`$, in such a way that fixing $`\sigma `$ the section $`\mathrm{{\rm Y}}`$ corresponds to the one–form $`\mathrm{{\rm Y}}^{\sigma ,\lambda }`$ defined by $$\mathrm{{\rm Y}}^{\sigma ,\lambda }(\xi )=\underset{\underset{¯}{j}+\mathrm{}=0}{}\frac{(1)^{\underset{¯}{j}}}{\underset{¯}{j}!}\lambda ^{}(\mathrm{ad}(\mathrm{{\rm Y}}_k)^{j_k}\mathrm{}\mathrm{ad}(\mathrm{{\rm Y}}_1)^{j_1}(\xi _{\mathrm{}})).$$ This identification is obtained simply by pulling back the affine structure on the space of principal connections on $`^\lambda `$ to the space of Weyl–structures. In particular, the change of the principal connections $`\tau `$ and $`\widehat{\tau }`$ on $`^\lambda `$ induced by $`\sigma `$ and $`\widehat{\sigma }`$, respectively, is just given by $`\widehat{\tau }=\tau +\mathrm{{\rm Y}}^{\sigma ,\lambda }`$. But then their curvatures change simply by $`\widehat{\rho }=\rho +d\mathrm{{\rm Y}}^{\sigma ,\lambda }`$, so in particular if $`\sigma `$ is closed then $`\widehat{\sigma }`$ is closed if and only if $`d\mathrm{{\rm Y}}^{\sigma ,\lambda }=0`$. Thus, in the same way as Weyl–structures are affine over all one–forms, closed Weyl–structures are affine over closed one–forms. For exact Weyl–structures, the situation is even simpler. If $`s`$ and $`\widehat{s}`$ are two global sections of $`^\lambda `$, then there is a unique smooth function $`f`$ such that $`\widehat{s}(x)=e^{f(x)}s(x)`$. It is then well known that the associated principal connections simply change by $`\widehat{\tau }=\tau +df`$, so exact Weyl–structures are affine over the space of exact one–forms. ### 3.14. Remark Another useful observation about exact Weyl geometries is related to the identification of $`^\lambda `$ with $`𝒢_0/\mathrm{ker}\lambda `$ from 3.8(2). By the general properties of classical $`G`$–structures, the sections of such bundles are in bijective correspondence with reductions of the structure groups to $`\mathrm{ker}\lambda G_0`$. Thus the holonomy of the Weyl connections given by closed Weyl structures is always at most $`\mathrm{ker}\lambda `$. In particular, in $`|1|`$–graded cases the scaling element is unique up to scalar multiples, and the kernel of $`\lambda `$ is exactly the semisimple part of $`G_0`$. The same observation is then true for the closed Weyl geometries locally. ### 3.15. Normal Weyl–structures Besides the rather obvious closed and exact Weyl–structures discussed above there is a second kind of special Weyl–structures, the so–called normal Weyl–structures. In several respects, they are quite different from closed and exact Weyl–structures. On one hand, they are “more canonical” since their definition does not involve the choice of a bundle of scales. On the other hand, in contrast to closed and exact Weyl–structures, which always exist globally, normal Weyl–structures in general exist only locally (over $`M`$). Their existence is closely related to the existence of normal coordinates for parabolic geometries. This subject will be taken up elsewhere. We would like to point out at this place that the existence of normal Weyl–structures seems to be a new result even in the case of conformal structures, where it significantly improves the result on the existence of Graham normal coordinates, see . Since the Rho tensors give the information about the difference of the covariant derivative with respect to the Weyl connection and the invariant derivative with respect to $`\omega `$ along the image of the chosen Weyl–structure $`\sigma `$, the “normality” we have in mind will be described in terms of certain minimality of P. More explicitly, a Weyl–structure $`\sigma `$ will be called normal at the point $`xM`$ if it satisfies the properties imposed in Theorem 3.16. This Theorem also describes completely the freedom in the choice. Recall that once we have chosen a Weyl–structure, we get an identification of the tangent bundle with its associated graded vector bundle. Thus $`TM`$ is associated to $`𝒢_0`$ and so there is the induced linear Weyl connection on $`TM`$. Since the Weyl–structure induces covariant derivatives on all components of the associated graded of the tangent bundle, the Weyl connection on $`TM`$ preserves the grading. For the same reason, we can form covariant derivatives of the Rho–tensor, viewed as a one–form with values in $`T^{}M𝒜_1\mathrm{}𝒜_k`$, which again preserve the grading. ###### 3.16 Theorem. Let $`p:𝒢M`$ be a parabolic geometry with underlying $`G_0`$–bundle $`p_0:𝒢_0M`$ and let $`\pi :𝒢𝒢_0`$ be the canonical projection. Let $`xM`$ be a point and let $`u_0𝒢_0`$ and $`u𝒢`$ be points such that $`\pi (u)=u_0`$ and $`p_0(u_0)=x`$. Then there exists an open neighborhood $`U`$ of $`x`$ in $`M`$ and a Weyl–structure $`\sigma :p_0^1(U)p^1(U)`$ such that $`\sigma (u_0)=u`$ and the Rho–tensor P of $`\sigma `$ has the property that for all $`k`$ the symmetrization over all $`\xi _i`$ of $`(_{\xi _k}\mathrm{}_{\xi _1}\text{P})(\xi _0)`$ vanishes at $`x`$, so in particular $`\text{P}(x)=0`$. Moreover, this condition uniquely determines the infinite jet of $`\sigma `$ in $`u_0`$. ###### Proof. Consider the Cartan connection $`\omega `$ on $`𝒢`$. Since $`\omega `$ restricts to a linear isomorphism, for each element $`A𝔤`$ we get the constant vector field $`\stackrel{~}{A}𝔛(𝒢)`$ defined by $`\stackrel{~}{A}(v)=\omega (v)^1(A)`$, cf. 2.4. (Note that for $`A𝔭`$ this is just the fundamental vector field.) In particular, we may consider the vector fields $`\stackrel{~}{X}`$ for $`X𝔤_{}`$. Now we can find a neighborhood $`V`$ of zero in $`𝔤_{}`$, such that for all $`XV`$ the flow of $`X`$ in the point $`u`$ exists up to time $`t=1`$. Define $`\phi :V𝒢`$ by $`\phi (X)=\mathrm{Fl}_1^{\stackrel{~}{X}}(u)`$. Since $`T_upT_0\phi :𝔤_{}T_xM`$ is obviously a linear isomorphism, we may assume (possibly shrinking $`V`$) that the maps $`\phi `$, $`\pi \phi `$ and $`p\phi `$ are all diffeomorphisms onto their images, and we put $`U=p(\phi (V))`$. For a point $`v_0p_0^1(U)`$ there clearly exist unique elements $`XV`$ and $`gG_0`$ such that $`v_0=\pi (\phi (X))g`$, and we define $`\sigma (v_0):=\phi (X)g`$. Obviously, this defines a smooth $`G_0`$–equivariant section $`\sigma :p_0^1(U)p^1(U)`$ and $`\sigma (u_0)=u`$. Next, consider a tangent vector $`\xi T_xM`$, and its horizontal lift $`\xi ^hT_{u_0}𝒢_0`$ with respect to the principal connection $`\sigma ^{}\omega _0`$. Since $`\sigma ^{}\omega _0`$ defines a Cartan connection on $`p_0^1(U)`$ (see 3.5) we can extend $`\xi ^h`$ uniquely to a vector field $`\stackrel{~}{\xi }^h`$ such that $`\sigma ^{}\omega _0(\stackrel{~}{\xi }^h)`$ is constantly equal to some $`X𝔤_{}`$. Moreover, $`\stackrel{~}{\xi }^h`$ is projectable to a vector field $`\stackrel{~}{\xi }`$ on $`U`$ and it is exactly the horizontal lift of $`\stackrel{~}{\xi }`$ (which also justifies the notation). Now consider the flow line $`c(t)=\mathrm{Fl}_t^{\stackrel{~}{X}}(u)=\phi (tX)`$ in $`𝒢`$, which is defined for sufficiently small $`t`$. By construction, we have $`\sigma (\pi (c(t)))=c(t)`$ for all $`t`$. But this implies that $`T\sigma (\pi c)^{}(t)=c^{}(t)`$, so $`\sigma ^{}(\omega )((\pi c)^{}(t))`$ is constantly equal to $`X`$ and thus $`(\pi c)(t)=\mathrm{Fl}_t^{\stackrel{~}{\xi }}(u_0)`$. On the other hand, from the construction it is clear that $`\omega (c^{}(t))=X𝔤_{}`$, so if we consider the function $`\text{P}:𝒢_0L(𝔤_{},𝔭_+)`$ describing the Rho–tensor, then $`\text{P}(\pi (c(t)))(X)=0`$ for all $`t`$. Consequently, all derivatives of this curve in $`t=0`$ vanish. But since $`\pi c`$ is an integral curve of $`\stackrel{~}{\xi }^h`$ these iterated derivatives exactly correspond to iterated covariant derivatives of P in direction $`\xi `$ evaluated at $`\xi `$. Thus, we obtain $`(_\xi \mathrm{}_\xi \text{P})(\xi )=0`$ for any number of covariant derivatives. Using polarization, this implies that the symmetrization of $`(_{\xi _k}\mathrm{}_{\xi _1}\text{P})(\xi _0)`$ over all $`\xi _i`$ vanishes at $`x`$. To see that our condition fixes the infinite jet of the Weyl–structure suppose that $`\widehat{\sigma }`$ is another normal Weyl structure with $`\widehat{\sigma }(u_0)=u`$ and let $`\mathrm{{\rm Y}}=(\mathrm{{\rm Y}}_1,\mathrm{},\mathrm{{\rm Y}}_k)`$ be the section of $`𝒜_1\mathrm{}𝒜_k`$ describing the change from $`\sigma `$ to $`\widehat{\sigma }`$. We want to show that the infinite jet of $`\mathrm{{\rm Y}}`$ vanishes at $`x=p(u)`$. Since both Weyl–structures map $`u_0`$ to $`u`$, we must have $`\mathrm{{\rm Y}}(x)=0`$. Next, we know that $`\text{P}(x)=\widehat{\text{P}}(x)=0`$. Since all $`\mathrm{{\rm Y}}_i`$ vanish in $`x`$, formula (2) from Proposition 3.9 immediately shows that this implies $`\mathrm{{\rm Y}}_i(x)=0`$ for all $`i=1,\mathrm{},k`$, so $`\mathrm{{\rm Y}}(x)=0`$. Now, $`\widehat{\text{P}}=0`$ and $`\mathrm{{\rm Y}}=0`$. On one hand, it follows that $`\widehat{}\text{P}(x)=\text{P}(x)`$ and on the other hand that $`(_\eta \widehat{\text{P}})(\xi )(x)=_\eta (\widehat{\text{P}}(\xi ))(x)`$. But hitting formula (2) from Proposition 3.9 with $`_\eta `$ and symmetrizing over $`\xi `$ and $`\eta `$, we always get terms involving some $`\mathrm{{\rm Y}}_i`$ or $`_\eta \mathrm{{\rm Y}}_i`$ or $`_\eta \text{P}(\xi )`$ which all vanish at $`x`$, except for one term in the very last line, in which we get a second covariant derivative of some $`\mathrm{{\rm Y}}_i`$. So we see that the symmetrizations of $`\widehat{}_\eta \widehat{\text{P}}(\xi )`$ and $`_\eta _\xi \mathrm{{\rm Y}}`$ coincide. Thus vanishing of the symmetrization of the first covariant derivative implies that the symmetrized second derivative of $`\mathrm{{\rm Y}}`$ is zero, and thus the two–jet of $`\mathrm{{\rm Y}}`$ at $`x`$ must be zero. Iteratively, one similarly sees that in the expression of an symmetrized iterated covariant derivative of P we always get terms involving symmetrized iterated covariant derivatives of $`\mathrm{{\rm Y}}_i`$’s or P’s except for one term coming from the very last line of the transformation formula. As above, one then concludes that vanishing of the symmetrization of the $`k`$–fold covariant derivative of $`\widehat{\text{P}}`$ is equivalent to vanishing of the symmetrization of the $`(k+1)`$–fold covariant derivative of $`\mathrm{{\rm Y}}`$ and thus to the $`k+1`$–jet of $`\mathrm{{\rm Y}}`$ in $`x`$ being trivial. ∎ ## 4. Characterization of Weyl–structures In the last section, we started with a Weyl–structure for a parabolic geometry $`(𝒢M,\omega )`$ and we constructed several underlying objects on the manifold $`M`$, see Figure 1 for an illustration. Now we are going to characterize when general objects of that type actually come from a Weyl structure. In the final stage, this will mean explicit conditions relating the soldering form, linear connection and its torsion and curvature, together with a procedure building the corresponding Rho–tensors. This is quite simple for irreducible geometries, where the soldering form is fixed, and the whole condition prescribes uniquely the torsion of a $`G_0`$–connection. The Rho–tensor is then given by a simple formula in terms of the curvature, see Example 4.8 below. Of course, the same story gets much more complicated for the general $`|k|`$–graded case. The main step is done in Theorem 4.4 and then a detailed analysis of the curvature fulfills our goal. Throughout this section we restrict to the case of regular parabolic geometries associated to a $`|k|`$–graded semisimple Lie algebra $`𝔤`$ such that $`H^1(𝔤_{},𝔤)`$ is concentrated in homogeneous degrees $`0`$, i.e. such that none of the simple $`|k_i|`$–graded ideals is of one of the two types mentioned in 2.8. In the case that such ideals are present, a similar characterization is possible, but the conditions are more complicated to formulate. ### 4.1. Definition Let $`p_0:𝒢_0M`$ be a regular infinitesimal flag structure, see 2.7. A Weyl–form for $`M`$ is a one–form $`\tau \mathrm{\Omega }^1(𝒢_0,𝔤)`$ which is $`G_0`$–equivariant, i.e. $`(r^g)^{}\tau =\mathrm{Ad}(g^1)\tau `$ for all $`gG_0`$, reproduces the generators of fundamental vector fields, i.e. $`\tau (\zeta _A)=A`$ for all $`A𝔤_0`$ and has the property that for each $`i<0`$ the restriction of $`\tau _i`$ to $`T^i𝒢_0`$ coincides with the $`𝔤_i`$–component of the frame form of degree one on $`𝒢_0`$ induced by the regular infinitesimal flag structure, see 2.7 and 3.5. By 3.3 and 3.4, for any Weyl–structure $`\sigma :𝒢_0𝒢`$, the pullback $`\sigma ^{}\omega `$ is a Weyl–form for $`M`$. As in 3.4, the condition of the restriction of $`\tau _i`$ to $`T^i𝒢_0`$, $`i<0`$, means on $`M`$ exactly that the restriction of $`\tau _i`$ to $`T^iM`$ coincides with the canonical projection $`T^iM𝒜_i`$. In particular, this implies that $`\tau _{}=\tau _k\mathrm{}\tau _1`$ induces a linear isomorphism $`T_u𝒢_0/V_u𝒢_0𝔤_{}`$, and thus $`\tau _0`$ is a Cartan connection on $`𝒢_0`$. Completely parallel to the development in 3.33.5 we can equivalently interpret a Weyl–form for $`M`$ as a one form $`\tau _{}\mathrm{\Omega }^1(M,𝒜_k\mathrm{}𝒜_1)`$ inducing an isomorphism between $`TM`$ and its associated graded bundle, plus a principal connection $`\tau _0\mathrm{\Omega }^1(𝒢_0,𝔤_0)`$ on $`𝒢_0`$, plus a Rho–tensor $`\text{P}=\text{P}^\tau \mathrm{\Omega }^1(M,𝒜_1\mathrm{}𝒜_k)`$, so a Weyl–form essentially consists of objects living on $`M`$. ### 4.2. Weyl–curvature Next, for a Weyl–form $`\tau `$ for $`M`$, we define the Weyl–curvature $`W`$ of $`\tau `$. As a $`𝔤`$–valued two form on $`𝒢_0`$, it is defined by $$W(\xi ,\eta )=d\tau (\xi ,\eta )+[\tau (\xi ),\tau (\eta )].$$ From the fact that $`\tau `$ is $`G_0`$–equivariant and reproduces the generators of fundamental vector fields, one immediately concludes that $`W`$ is horizontal and $`G_0`$–equivariant, so it descends to an $`𝒜`$–valued two form on $`M`$. Taking into account the identification of $`TM`$ with $`𝒜_{}`$, we can also view $`W`$ as a section of $`L(\mathrm{\Lambda }^2𝒜_{},𝒜)`$. Finally note that any section $`\mathrm{\Phi }`$ of $`L(\mathrm{\Lambda }^2𝒜_{},𝒜)`$ can be split according to homogeneous degrees. We denote by $`\mathrm{\Phi }^{(\mathrm{})}`$ the homogeneous part of degree $`\mathrm{}`$, i.e. $`\mathrm{\Phi }^{(\mathrm{})}(\xi ,\eta )𝒜_{i+j+\mathrm{}}`$ for sections $`\xi `$ of $`𝒜_i`$ and $`\eta `$ of $`𝒜_j`$ with $`i,j<0`$. ###### Lemma. Let $`p_0:𝒢_0M`$ be a regular infinitesimal flag structure. Then any Weyl–form $`\tau \mathrm{\Omega }^1(𝒢_0,𝔤)`$ has the property that $`W^{(\mathrm{})}=0`$ for all $`\mathrm{}0`$. ###### Proof. Consider $`\xi \mathrm{\Gamma }(𝒜_i)`$ and $`\eta \mathrm{\Gamma }(𝒜_j)`$, for $`i,j<0`$. Then $`\tau _n(\xi )=0`$ for $`n<i`$ and $`\tau _m(\eta )=0`$ for $`m<j`$, so for $`\mathrm{}<0`$ if $`m+n=i+j+\mathrm{}`$ then $`[\tau _n(\xi ),\tau _m(\eta )]=0`$. Thus, in this case, the definition of $`W^{(\mathrm{})}(\xi ,\eta )`$ can be rewritten as $`W^{(\mathrm{})}(\xi ,\eta )=d\tau _{i+j+\mathrm{}}(\xi ,\eta )=\tau _{i+j+\mathrm{}}([\xi ,\eta ])`$. By definition of a Weyl–form, $`W^{(\mathrm{})}(\xi ,\eta )`$ thus equals the class of the bracket $`[\xi ,\eta ]`$ in $`T^{i+j+\mathrm{}}M/T^{i+j+\mathrm{}+1}M`$. But according to 2.7, we in particular know that the bracket of any section of $`T^iM`$ with a section of $`T^jM`$ lies in $`T^{i+j}M`$, so since $`\mathrm{}<0`$, we must have $`W^{(\mathrm{})}=0`$. Next, for $`\mathrm{}=0`$, we can write $$W^{(0)}(\xi ,\eta )=d\tau _{i+j}(\xi ,\eta )+\{\xi ,\eta \}=\tau _{i+j}([\xi ,\eta ])+\{\xi ,\eta \}.$$ Again, $`\tau _{i+j}([\xi ,\eta ])`$ is just the class of the bracket in $`T^{i+j}M/T^{i+j+1}M`$ and so the vanishing of $`W^{(0)}`$ is just the remaining part of the definition of regular infinitesimal flag structures, see 2.7. ∎ ### 4.3. Definition Let $`p_0:𝒢_0M`$ be a regular infinitesimal flag structure. Then a Weyl–form $`\tau \mathrm{\Omega }^1(𝒢_0,𝔤)`$ is called normal if and only if its Weyl–curvature $`W\mathrm{\Gamma }(L(\mathrm{\Lambda }^2𝒜_{},𝒜))`$ satisfies $`^{}(W)=0`$, where $`^{}:L(\mathrm{\Lambda }^2𝒜_{},𝒜))L(𝒜_{},𝒜)`$ is the bundle map induced by the Lie algebra codifferential, see 2.13. ###### 4.4 Theorem. Let $`(p:𝒢M,\omega )`$ be a regular parabolic geometry and let $`p_0:𝒢_0M`$ be the underlying regular infinitesimal flag structure. Then a Weyl–form $`\tau \mathrm{\Omega }^1(𝒢_0,𝔤)`$ for $`M`$ is coming from some Weyl–structure $`\sigma :𝒢_0𝒢`$, i.e. $`\tau =\sigma ^{}\omega `$, if and only if $`\tau `$ is normal. ###### Proof. First we show that for any Weyl–structure $`\sigma :𝒢_0𝒢`$ the Weyl–form $`\sigma ^{}\omega `$ is normal. By the definition in 4.2 the Weyl–curvature $`W`$ is a $`𝔤`$–valued two–form on $`𝒢_0`$, given by $$W(\xi ,\eta )=d\sigma ^{}\omega (\xi ,\eta )+[\sigma ^{}\omega (\xi ),\sigma ^{}\omega (\eta )]=\sigma ^{}(d\omega +\frac{1}{2}[\omega ,\omega ])(\xi ,\eta ).$$ Thus, $`W`$ is simply the pullback along $`\sigma `$ of the curvature of the Cartan connection $`\omega `$ on $`𝒢_0`$. By definition of a normal parabolic geometry, this curvature is $`^{}`$–closed, so the same is true for $`W`$. Now, let us assume that we have given an arbitrary normal Weyl–form $`\tau \mathrm{\Omega }^1(𝒢_0,𝔤)`$. Moreover, let us choose any bundle $`^\lambda `$ of scales for the parabolic geometry in question. Since $`\tau _0`$ is a principal connection on $`𝒢_0`$, it induces a principal connection on $`^\lambda `$, which by Theorem 3.12 in turn gives rise to a unique Weyl–structure $`\sigma :𝒢_0𝒢`$ such that the connection on $`^\lambda `$ induced by $`\sigma ^{}\omega _0`$ coincides with the connection induced by $`\tau _0`$. We claim that $`\tau =\sigma ^{}\omega `$, which will conclude the proof. Consider the difference $`\tau \sigma ^{}\omega \mathrm{\Omega }^1(𝒢_0,𝔤)`$. For $`i<0`$, we know from our assumptions that both $`\tau _i`$ and $`\sigma ^{}\omega _i`$ coincide on $`T^i𝒢_0`$ with the frame form of degree one. In particular, the difference $`\tau _i\sigma ^{}\omega _i`$ vanishes on $`T^i𝒢_0`$ for all $`i<0`$. Since $`T^0𝒢_0`$ is just the vertical bundle of $`𝒢_0`$ and since both $`\tau _0`$ and $`\sigma ^{}\omega _0`$ are principal connections on $`𝒢_0`$, we see that $`\tau _0\sigma ^{}\omega _0`$ vanishes on $`T^0𝒢_0`$, too. Finally, if we put $`T^i𝒢_0`$ to be the zero section for $`i>0`$, then $`\tau _i\sigma ^{}\omega _i`$ vanishes on $`T^i𝒢_0`$ for all $`i=k,\mathrm{},k`$. Let us inductively assume that $`\tau _i\sigma ^{}\omega _i`$ vanishes on $`T^{in+1}𝒢_0`$ for all $`i`$ and some $`n`$. Then consider the restriction of $`\tau _i\sigma ^{}\omega _i`$ to $`T^{in}𝒢_0`$, which can be viewed as a map $`T^{in}𝒢_0/T^{in+1}𝒢_0𝔤_i`$. For each $`i`$ such that $`in0`$, the forms $`\tau _{in}`$ and $`\sigma ^{}\omega _{in}`$ coincide on $`T^{in}𝒢_0`$ and induce an isomorphism$`T^{in}𝒢_0/T^{in+1}𝒢_0𝒢_0\times 𝔤_{in}`$. Consequently, we get a unique map $`\mathrm{\Phi }:𝒢_0L(𝔤_{},𝔤)`$ which has values in maps homogeneous of degree $`n`$, such that $`(\tau _i\sigma ^{}\omega _i)(\xi )=\mathrm{\Phi }(\tau _{in}(\xi ))`$ for all $`\xi T^{in}𝒢_0`$. Next, let $`W^{(n)}`$ be the homogeneous component of degree $`n`$ of the Weyl–curvature of $`\tau `$ viewed as a function $`𝒢_0L(\mathrm{\Lambda }^2𝔤_{},𝔤)`$ (having values in the maps homogeneous of degree $`n`$), and let $`\stackrel{~}{W}^{(n)}`$ be the corresponding object for $`\sigma ^{}\omega `$. We claim that for all $`X`$, $`Y𝔤_{}`$ (1) $`\stackrel{~}{W}^{(n)}(X,Y)`$ $`=W^{(n)}(X,Y)[X,\mathrm{\Phi }(Y)]+[Y,\mathrm{\Phi }(X)]+\mathrm{\Phi }([X,Y])=`$ $`=W^{(n)}(X,Y)(\mathrm{\Phi })(X,Y).`$ Let us postpone the proof of this claim and assume it is true for a while. Since both $`W^{(n)}`$ and $`\stackrel{~}{W}^{(n)}`$ are $`^{}`$–closed, this implies $`^{}\mathrm{\Phi }=0`$, which implies $`\mathrm{\Phi }=0`$ since $``$ and $`^{}`$ are adjoint, see 2.5. Since $`H^1(𝔤_{},𝔤)`$ is concentrated in non-positive degrees of homogeneity, this implies $`\mathrm{\Phi }=0`$ for $`n>k`$ and $`\mathrm{\Phi }(X)=[Z,X]`$ for some smooth $`Z:𝒢_0𝔤_n`$ for $`nk`$. But in the latter case, the proof of Theorem 3.12(1) shows that since $`\tau _0`$ and $`\sigma ^{}\omega _0`$ induce the same principal connection on $`^\lambda `$, we must have $`Z=0`$, and thus $`\mathrm{\Phi }=0`$. Hence, $`\tau _i`$ and $`\sigma ^{}\omega _i`$ coincide on $`T^{in}𝒢_0`$ for all $`i<n`$, for $`i=n`$ this follows since $`n>0`$ and thus both $`\tau _n`$ and $`\sigma ^{}\omega _n`$ are horizontal, while for $`i>n`$ it is trivially satisfied. Thus the result follows by induction. So we are left with the proof of (1) only. Let us fix $`X𝔤_i`$, $`Y𝔤_j`$, $`i,j<0`$. By definition, $$W^{(n)}(u)(X,Y)=d\tau _{i+j+n}(\tau _0^1(X),\tau _0^1(Y))+[\tau (\tau _0^1(X)),\tau (\tau _0^1(Y))]_{i+j+n},$$ where the index in the bracket means that we just have to take the component in $`𝔤_{i+j+n}`$. For $`\stackrel{~}{W}^{(n)}`$ we get the analogous formula with all $`\tau `$’s replaced by $`\sigma ^{}\omega `$. Next, observe that both $`\tau _0^1(X)`$ and $`\sigma ^{}\omega _0^1(X)`$ lie in $`T^i𝒢_0T^{i+j}𝒢_0`$ and similarly for $`Y`$. From above, we know that $`\sigma ^{}\omega _{i+j+n}(\xi )=\tau _{i+j+n}(\xi )\mathrm{\Phi }(\tau _{i+j}(\xi ))`$ for all $`\xi `$ in $`T^{i+j}𝒢_0`$. Taking the exterior derivative of this equation and keeping in mind that $`\tau _{i+j}`$ vanishes on $`T^i𝒢_0`$ and $`T^j𝒢_0`$, we see that for $`\xi T^i𝒢_0`$ and $`\eta T^j𝒢_0`$ we get $$d\sigma ^{}\omega _{i+j+n}(\xi ,\eta )=d\tau _{i+j+n}(\xi ,\eta )\mathrm{\Phi }(d\tau _{i+j}(\xi ,\eta )).$$ Since $`W^{(0)}=0`$, the second term (including the $``$ sign) can be rewritten as $`\mathrm{\Phi }([\tau _i(\xi ),\tau _j(\eta )])`$, and we may as well replace $`\tau `$ by $`\sigma ^{}\omega `$ in this expression. Thus, we see that $`\stackrel{~}{W}^{(n)}(X,Y)`$ $`=d\tau _{i+j+n}(\sigma ^{}\omega _0^1(X),\sigma ^{}\omega _0^1(Y))+\mathrm{\Phi }([X,Y])+`$ $`+[\sigma ^{}\omega (\sigma ^{}\omega _0^1(X)),\sigma ^{}\omega (\sigma ^{}\omega _0^1(Y))]_{i+j+n}.`$ Now we have to distinguish a few cases: Let us first assume that $`i+n>0`$. Then $`\sigma ^{}\omega _0^1(X)=\tau _0^1(X)`$, and $`\sigma ^{}\omega (\sigma ^{}\omega _0^1(X))=\tau (\tau _0^1(X))\mathrm{\Phi }(X)`$, and $`\mathrm{\Phi }(X)𝔤_{i+n}𝔭_+`$. In particular, this implies that $$\begin{array}{c}[\sigma ^{}\omega (\sigma ^{}\omega _0^1(X)),\sigma ^{}\omega (\sigma ^{}\omega _0^1(Y))]_{i+j+n}=\hfill \\ \hfill =[\tau (\tau _0^1(X)),\sigma ^{}\omega (\sigma ^{}\omega _0^1(Y))]_{i+j+n}[\mathrm{\Phi }(X),Y].\end{array}$$ Secondly, if $`i+n=0`$ then $`\mathrm{\Phi }(X)𝔤_0`$, and thus $`\sigma ^{}\omega _0^1(X)=\tau _0^1(X)+\zeta _{\mathrm{\Phi }(X)}`$. The infinitesimal version of equivariance of $`\tau _{i+j+n}`$ then implies that $$d\tau _{i+j+n}(\sigma ^{}\omega _0^1(X),\sigma ^{}\omega _0^1(Y))=d\tau _{i+j+n}(\tau _0^1(X),\sigma ^{}\omega _0^1(Y))[\mathrm{\Phi }(X),Y],$$ since $`i+j+n=j`$ in this case. On the other hand both $`\sigma ^{}\omega (\sigma ^{}\omega _0^1(X))`$ and $`\tau (\tau _0^1(X))`$ in this case are congruent to $`X`$ modulo $`𝔭_+`$, so $$[\sigma ^{}\omega (\sigma ^{}\omega _0^1(X)),\sigma ^{}\omega (\sigma ^{}\omega _0^1(Y))]_{i+j+n}=[\tau (\tau _0^1(X)),\sigma ^{}\omega (\sigma ^{}\omega _0^1(Y))]_{i+j+n}.$$ Finally, suppose that $`i+n<0`$, so $`\mathrm{\Phi }(X)𝔤_{i+n}𝔤_{}`$. Then $`\sigma ^{}\omega _0^1(X)`$ is congruent to $`\tau _0^1(X+\mathrm{\Phi }(X))`$ modulo $`T^{i+n+1}𝒢_0`$. Since the bracket of a section of this subbundle with a section of $`T^j𝒢_0`$ is a section of $`T^{i+j+n+1}𝒢_0`$ and $`\tau _{i+j+n}`$ vanishes on the latter subbundle, we conclude that $`d\tau _{i+j+n}(\sigma ^{}\omega _0^1(X),\sigma ^{}\omega _0^1(Y))`$ $`=d\tau _{i+j+n}(\tau _0^1(X),\sigma ^{}\omega _0^1(Y))+`$ $`+d\tau _{i+j+n}(\tau _0^1(\mathrm{\Phi }(X)),\sigma ^{}\omega _0^1(Y)).`$ Since $`W^{(0)}=0`$, the last term can be rewritten as $`[\mathrm{\Phi }(X),Y]`$. As above, both $`\sigma ^{}\omega (\sigma ^{}\omega _0^1(X))`$ and $`\tau (\tau _0^1(X))`$ are congruent to $`X`$ modulo $`𝔭_+`$, so again the bracket term makes no problem. Hence we see, that in any case we get $`\stackrel{~}{W}^{(n)}(X,Y)`$ $`=d\tau _{i+j+n}(\tau _0^1(X),\sigma ^{}\omega _0^1(Y))+[\tau (\tau _0^1(X)),\sigma ^{}\omega (\sigma ^{}\omega _0^1(Y))]_{i+j+n}`$ $`[\mathrm{\Phi }(X),Y]+\mathrm{\Phi }([X,Y]).`$ Doing the same changes to $`Y`$ instead of $`X`$ we obtain the required equality (1), and the whole proof of the theorem is finished. ∎ ### 4.5. Remark If one does not assume that $`H^1(𝔤_{},𝔤)`$ is concentrated in non-positive degrees, i.e. if one allows $`𝔤`$ to contain one of the two simple factors mentioned in 2.8, then $`H^1(𝔤_{},𝔤)`$ is concentrated in homogeneous degrees less or equal to one. Thus, the above proof shows that $`\tau =\sigma ^{}\omega `$ if $`\tau `$ is normal and has the property that the restrictions of $`\tau _i`$ and $`\sigma ^{}\omega _i`$ to $`T^{i1}𝒢_0`$ coincide for all $`i`$. This condition is then fairly simple to interpret for any concrete choice of such structure. ### 4.6. In the proof of Theorem 4.4, we observed that for a Weyl–structure $`\sigma :𝒢_0𝒢`$ the Weyl–curvature $`W`$ of the Weyl–form $`\sigma ^{}\omega `$ is exactly the pullback along $`\sigma `$ of the curvature $`\kappa `$ of the normal Cartan connection $`\omega `$ on $`𝒢`$. This allows us to compute the change of the Weyl–curvature under a change of the Weyl–structure. Suppose that $`\widehat{\sigma }`$ is another Weyl–structure and $`\mathrm{{\rm Y}}=(\mathrm{{\rm Y}}_1,\mathrm{},\mathrm{{\rm Y}}_k)`$ is the smooth section of $`𝒜_1\mathrm{}𝒜_k`$ describing the change from $`\sigma `$ to $`\widehat{\sigma }`$, see Proposition 3.2, i.e. $$\widehat{\sigma }(u)=\sigma (u)\mathrm{exp}(\mathrm{{\rm Y}}_1(u))\mathrm{}\mathrm{exp}(\mathrm{{\rm Y}}_k(u)).$$ Equivariance of the Cartan connection $`\omega `$ immediately implies that the curvature $`\kappa `$ is equivariant, i.e. viewing $`\kappa `$ as a two form on $`𝒢`$ with values in $`𝔤`$, we have $`\kappa (vg)(Tr^g\xi ,Tr^g\eta )=\mathrm{Ad}(g^1)(\kappa (v)(\xi ,\eta ))`$ for $`gP`$ and $`\xi ,\eta T_v𝒢`$. Putting $`v=\sigma (u)`$ and $`g=\mathrm{exp}(\mathrm{{\rm Y}}_1(u))\mathrm{}\mathrm{exp}(\mathrm{{\rm Y}}_k(u))`$, we see from the proof of Proposition 3.9 that for $`\xi T_u𝒢_0`$ the element $`T_u\widehat{\sigma }\xi `$ is congruent to $`Tr^gT_u\sigma \xi `$ modulo vertical elements, which are killed by the curvature anyhow. Thus, viewing $`W`$ and $`\widehat{W}`$ as $`𝔤`$–valued two forms on $`𝒢_0`$, we get $`\widehat{W}(\xi ,\eta )=\mathrm{Ad}(g^1)(W(\xi ,\eta ))`$. Moreover, to get the interpretation of our two Weyl curvatures $`W`$ and $`\widehat{W}`$ as $`𝒜`$–valued two forms on $`M`$, we just have to apply the above definition to lifts of vector fields on $`M`$, and the result is independent of the choice of the lifts since $`W`$ is horizontal. Keeping in mind that the Lie–bracket in $`𝔤`$ corresponds to the algebraic bracket of sections of $`𝒜`$ and expanding the exponentials in $`\mathrm{Ad}(g^1)`$ as in the proof of Proposition 3.9 we arrive (with notation as in 3.9) at (1) $$\widehat{W}_i(\xi ,\eta )=\underset{\underset{¯}{j}+\mathrm{}=i}{}\frac{(1)^{\underset{¯}{j}}}{\underset{¯}{j}!}\mathrm{ad}(\mathrm{{\rm Y}}_k)^{j_k}\mathrm{}\mathrm{ad}(\mathrm{{\rm Y}}_1)^{j_1}(W_{\mathrm{}}(\xi ,\eta )).$$ From this formula, one can also derive a formula describing the change of $`W`$ viewed as a section of $`L(\mathrm{\Lambda }^2𝒜_{},𝒜)`$ taking into account the change of the identification of $`TM`$ with $`𝒜_{}`$ described by (1) in Proposition 3.9, and thus a formula for the change of the individual homogeneous components $`W^{(\mathrm{})}`$. The only point that is important for us here is that the homogeneous component $`W^{(1)}`$ of degree one is actually independent of $`\sigma `$. This can be immediately verified from the above formula, taking into account that $`W^{(\mathrm{})}=0`$ for all $`\mathrm{}0`$. ### 4.7. Remark The results obtained so far in principle allow to give a description of the Cartan bundle and the Cartan connection completely in terms of data on the manifold $`M`$. More precisely, if we start from a regular infinitesimal flag structure underlying some parabolic geometry, then we may proceed as follows: Choose a scaling element $`E_\lambda 𝔷(𝔤_0)`$, and consider the corresponding homomorphism $`\lambda :G_0^+`$ described in the proof of Proposition 3.8. Then form $`_\lambda =𝒢_0\times _{G_0}^+`$. From Theorem 3.12(2) we then know that the Cartan bundle $`𝒢`$ is just the pullback of the bundle of principal connections on $`_\lambda `$, and we have a description of the principal action. Moreover, a choice of a principal connection on $`^\lambda `$ is just the choice of a global section of the bundle of connections, so its pullback is a smooth $`G_0`$–equivariant section $`\sigma :𝒢_0𝒢`$. Any Cartan connection $`\omega `$ on $`𝒢`$ is uniquely determined by its pullback $`\sigma ^{}\omega `$ by equivariance. Thus, describing the canonical normal Cartan connection on $`𝒢`$ is equivalent to finding a normal Weyl–form on $`𝒢_0`$ which induces a given connection on $`^\lambda `$. ### 4.8. Example Let us look more closely at the irreducible parabolic geometries. Here the regular infinitesimal flag structures are just $`G_0`$–structures on $`M`$ in the sense of classical G–structures. The Weyl forms are $`\tau =\tau _1+\tau _0+\tau _1`$ where $`\tau _1:T𝒢_0𝔤_1`$ is the fixed soldering form for $`M`$, $`\tau _0`$ is any linear connection on $`M`$ belonging to the fixed $`G_0`$–structure and $`\tau _1`$ is any one–form in $`\mathrm{\Omega }^1(M;T^{}M)`$. Now, $$W_1=d\tau _1+[\tau _1,\tau _0],$$ i.e. the torsion of the connection $`\tau _0`$. The individual components of $`W`$ have homogeneities one, two, and three and so they have to be $`^{}`$–closed separately. The condition $`^{}W_1=0`$ means that the torsion of $`\tau _0`$ is harmonic and this is the part of $`W`$ independent of the choice of the Weyl–structure. Next, $$W_0=d\tau _0+\frac{1}{2}[\tau _0,\tau _0]+[\tau _1,\tau _1]$$ which is the curvature $`R`$ of the connection $`\tau _0`$ plus some additional term. The co–closedness of $`W_0`$ imposes a condition on the choice of $`\tau _1`$, while $`^{}W_1`$ always vanishes since its values are in the trivial vector space. We shall see later that the resulting system of equations for the tensor $`\tau _1`$ is always solvable, except for the projective structures (where the first cohomology is concentrated in degree one). Moreover, we shall prove an explicit algebraic formula for the necessary choice for the Rho–tensor: $`\tau _1=\mathrm{}^1^{}R`$. Expanding this formula in the case of the conformal (pseudo) Riemannian geometry, we obtain the well known Rho–tensor used heavily by many authors since the beginning of this century, while $`d\tau _1`$ happens to be exactly another well known tensor, the Cotton–York tensor. As mentioned above, this computation may be understood as an alternative for the explicit construction of the canonical Cartan connection for all irreducible parabolic geometries. ### 4.9. Total curvature The explicit construction of a normal Weyl–form depends a lot on the structure in question, a detailed treatment in the case of partially integrable almost CR–structures of hypersurface type will appear in . Here we just describe the basic ingredient of this procedure. The upshot of this is that the condition on a Weyl–form $`\tau `$ being normal can be step by step reduced to a condition on $`\tau _0`$ only, at the same time computing step by step the components of the Rho–tensor $`\text{P}=\tau _+`$. The first step in this direction is to replace the Weyl curvature of a Weyl–form $`\tau `$ by 2–forms defined by splitting the structure equations for $`\tau `$. The curvature of the Cartan connection $`\tau _0`$ is the 2–form $`K_0\mathrm{\Omega }^2(𝒢_0,𝔤_{})`$ given by $$K_0(\xi ,\eta )=d\tau _0(\xi ,\eta )+[\tau _0(\xi ),\tau _0(\eta )].$$ On the other hand, we define the 2–form $`K_+\mathrm{\Omega }^2(𝒢_0,𝔭_+)`$ by $$K_+(\xi ,\eta )=d\tau _+(\xi ,\eta )+[\tau _+(\xi ),\tau _+(\eta )].$$ Motivated by conformal geometry, we call $`K_+`$ the Cotton–York–tensor associated to the Weyl–form $`\tau `$. We write $`K=K_0+K_+`$ and we call it the (total) curvature of $`\tau `$. Since $`\tau _0`$ is a Cartan connection, it is well known that its curvature is horizontal and $`G_0`$–equivariant, so it can be viewed as a two form on $`M`$, with values in the bundle $`𝒜_k\mathrm{}𝒜_0`$. On the other hand, since $`\tau _+`$ is by assumption $`G_0`$–equivariant and horizontal, the part $`K_+`$ descends to $`M`$, too. Finally, taking into account the isomorphism $`TM𝒜_{}=𝒜_k\mathrm{}𝒜_1`$, we can finally view $`K`$ as a smooth section of the bundle $`L(\mathrm{\Lambda }^2𝒜_{},𝒜)`$ over $`M`$. The reason for introducing this curvature is that it is more closely related to usual invariants of the Weyl–form than the Weyl–curvature, cf. Example 4.8. On the other hand, we shall see that there still is a simple relation between curvature and Weyl–curvature. To get explicit expressions for the components of $`K`$, recall that the component $`\tau _0`$ of any Weyl–form $`\tau `$ is a principal connection on $`𝒢_0`$, and thus induces a linear connection $``$ on each of the bundles $`𝒜_i`$. Let us also recall that $`\tau _i`$ are identified with forms $`\mathrm{\Omega }^1(M;𝒜_i)`$ for all negative $`i`$. ###### 4.10 Proposition. Let $`p_0:𝒢_0M`$ be a regular infinitesimal flag structure, let $`\tau \mathrm{\Omega }^1(𝒢_0,𝔤)`$ be a Weyl–form for $`M`$, and let $`K`$ be its total curvature, viewed as an $`𝒜`$–valued two form on $`M`$ with $`𝒜_{\mathrm{}}`$–component $`K_{\mathrm{}}`$. Then for all vector fields $`\xi `$ and $`\eta `$ on $`M`$ we have: (1) $`K_{\mathrm{}}(\xi ,\eta )=_\xi (\tau _{\mathrm{}}(\eta ))_\eta (\tau _{\mathrm{}}(\xi ))\tau _{\mathrm{}}([\xi ,\eta ])+_{\begin{array}{c}i,j<0\\ i+j=\mathrm{}\end{array}}\{\tau _i(\xi ),\tau _j(\eta )\}`$, for $`\mathrm{}<0`$. (2) For $`\zeta 𝒜_m`$ we get $`\{K_0(\xi ,\eta ),\zeta \}=R_m(\xi ,\eta )(\zeta )`$, where $`R_m`$ is the curvature of the linear connection $``$ on $`𝒜_m`$. Moreover, if we view $`K`$ as a section of $`L(\mathrm{\Lambda }^2𝒜_{},𝒜)`$ and consider $`\mathrm{}>0`$, then the homogeneous component $`K^{(\mathrm{})}`$ of $`K`$ depends only on the restrictions of $`\tau _i`$ to $`T^i\mathrm{}𝒢_0`$ for all $`i0`$ and on the restrictions of $`\tau _i`$ to $`T^{i\mathrm{}+1}𝒢_0`$ for $`i>0`$. ###### Proof. By definition, for $`\mathrm{}<0`$ the function $`𝒢_0𝔤_{\mathrm{}}`$ corresponding to $`K_{\mathrm{}}(\xi ,\eta )`$ is given by $$\begin{array}{c}d\tau _{\mathrm{}}(\xi ^h,\eta ^h)+\underset{i,j0,i+j=\mathrm{}}{}[\tau _i(\xi ^h),\tau ^j(\eta ^h)]=\hfill \\ \hfill =\xi ^h\tau _{\mathrm{}}(\eta ^h)\eta ^h\tau _{\mathrm{}}(\xi ^h)\tau _{\mathrm{}}([\xi ^h,\eta ^h])+\underset{i,j0,i+j=\mathrm{}}{}[\tau _i(\xi ^h),\tau _j(\eta ^h)],\end{array}$$ where the superscript $`h`$ denotes the horizontal lift with respect to the principal connection $`\tau _0`$. But now $`\tau _{\mathrm{}}(\eta ^h):𝒢_0𝔤_{\mathrm{}}`$ is exactly the smooth function corresponding to the section $`\tau _{\mathrm{}}(\eta )`$ of $`𝒜_{\mathrm{}}`$, so the function $`\xi ^h\tau _{\mathrm{}}(\eta ^h)`$ corresponds to $`_\xi (\tau _{\mathrm{}}(\eta ))`$ and similarly for the second term. On the other hand, $`[\xi ^h,\eta ^h]`$ is a lift of the vector field $`[\xi ,\eta ]`$, so since $`\tau _{\mathrm{}}`$ is horizontal for $`\mathrm{}<0`$, we see that the function $`\tau _{\mathrm{}}([\xi ^h,\eta ^h])`$ corresponds to the section $`\tau _{\mathrm{}}([\xi ,\eta ])`$ of $`𝒜_{\mathrm{}}`$. Finally, for the last sum one only has to take into account that $`\tau _0`$ vanishes on horizontal lifts and the bracket in $`𝔤`$ corresponds to the algebraic bracket on $`𝒜`$. If $`\mathrm{}=0`$, the definition of $`K_0`$ reduces to $`d\tau _0(\xi ^h,\eta ^h)`$ and this exactly represents the curvature of the principal connection $`\tau _0`$, so the result follows immediately, taking into account that the action of $`𝔤_0`$ on $`𝔤_m`$ is given by the Lie bracket in $`𝔤`$ and thus corresponds to the algebraic bracket $`𝒜_0\times 𝒜_m𝒜_m`$. To verify the statements about homogeneous degrees, take sections $`\xi `$ of $`𝒜_i`$ and $`\eta `$ of $`𝒜_j`$, and let $`\stackrel{~}{\xi }`$ be the (unique) section of $`T^iM`$ such that $`\tau _n(\stackrel{~}{\xi })=0`$ for all $`i<n<0`$, $`\tau _i(\stackrel{~}{\xi })=\xi `$, and similarly for $`\stackrel{~}{\eta }`$. Then for $`\mathrm{}>0`$, $`K^{(\mathrm{})}(\xi ,\eta )=K_{i+j+\mathrm{}}(\stackrel{~}{\xi },\stackrel{~}{\eta })`$. If $`i+j+\mathrm{}<0`$, then the above formula just gives us $$\delta _{i+\mathrm{}}^0_{\stackrel{~}{\xi }}\eta \delta _{j+\mathrm{}}^0_{\stackrel{~}{\eta }}\xi \tau _{i+j+\mathrm{}}([\stackrel{~}{\xi },\stackrel{~}{\eta }]).$$ This is completely independent of the components $`\tau _n`$ for $`n>0`$. If we allow a change of $`\tau `$ without changing the restriction of $`\tau _n`$ to $`T^n\mathrm{}`$ for all $`n0`$, then this means that $`\stackrel{~}{\xi }`$ is changed at most by a section of $`T^{i+\mathrm{}+1}M`$. In particular, if the first term in the above expression actually occurs, i.e. $`i+\mathrm{}=0`$ then $`\stackrel{~}{\xi }`$ is fixed, and moreover, since the restriction of $`\tau _0`$ to $`T^{\mathrm{}}𝒢_0=T^i𝒢_0`$ is fixed, also the covariant derivative is fixed. Similarly one analyzes the second term. Finally, the last term depends only on the restriction of $`\tau `$ since the bracket of a section of $`T^{i+\mathrm{}+1}M`$ with a section of $`T^jM`$ is a section of $`T^{i+j+\mathrm{}+1}M`$ and this subbundle lies in the kernel of $`\tau _{i+j+\mathrm{}}`$. If $`i+j+\mathrm{}=0`$, then $`K^{(\mathrm{})}(\xi ,\eta )=d\tau _0((\stackrel{~}{\xi })^h,(\stackrel{~}{\eta })^h)`$, and as above, we see that $`(\stackrel{~}{\xi })^h`$ and $`(\stackrel{~}{\eta })^h`$ depend only on the appropriate restriction of $`\tau `$. Moreover, the bracket $`[(\stackrel{~}{\xi })^h,(\stackrel{~}{\eta })^h]`$ by construction is a section of $`T^{i+j}𝒢_0`$, so the whole expression depends only on the restriction of $`\tau _0`$ to $`T^{i+j}𝒢_0=T^0\mathrm{}𝒢_0`$. Finally, we have to consider the case $`i+j+\mathrm{}>0`$, so we are dealing with a component of $`K`$ having values in $`𝒜_+`$. As before, one verifies that all extensions and horizontal lifts depend only on the appropriate restrictions of $`\tau _0`$, so what remains to be discussed is the dependence on P. But viewing P as a section of $`L(𝒜_{},𝒜_+)`$, the statement to be proved reduces to the fact that a homogeneous component of $`K`$ depends only on homogeneous components of P of strictly smaller degree. But this is obvious from the definition of $`K_+`$. ∎ ### 4.11. Remark The previous Proposition reveals that the $`𝒜_{}`$–components of the total curvature give exactly the torsion of the linear connection $`\tau _0`$ corrected by the algebraic contribution of the Lie bracket in $`𝔤_{}`$, while the component $`K_0`$ is just the standard curvature of $`\tau _0`$. For a normal Weyl form $`\tau `$ this means (using Proposition 4.12 below) that the torsion of $`\tau _0`$ has the algebraic bracket as its homogeneous component of degree zero, no components of negative degrees, and some positive degree components. The torsion component of degree one is an invariant of the parabolic structure in question. The key point in the further analysis is that while the total curvature of a Weyl–form is much easier to relate to the underlying structure than its Weyl–curvature, there is the quite simple relation between them described in the next Proposition. ###### 4.12 Proposition. Let $`\tau \mathrm{\Omega }^1(𝒢_0,𝔤)`$ be a Weyl–form for $`M`$, let $`\text{P}\mathrm{\Gamma }(L(𝒜_{},𝒜_+))`$ be its Rho–tensor, and let $`K,W\mathrm{\Gamma }(L(\mathrm{\Lambda }^2𝒜_{},𝒜))`$ be its total curvature and its Weyl–curvature, respectively. Then $$W(\xi ,\eta )=K(\xi ,\eta )+\{\text{P}(\xi ),\eta \}\{\text{P}(\eta ),\xi \}.$$ In particular, $`W^{(i)}=K^{(i)}`$ for all $`i1`$. ###### Proof. Let $`\xi `$ be a section of $`𝒜_i`$ and $`\eta `$ be a section of $`𝒜_j`$, with $`i,j<0`$. To compute $`W(\xi ,\eta )`$, we first have to view $`\xi `$ and $`\eta `$ as vector fields on $`M`$ via $`\tau _{}:TM𝒜_{}`$. Then, by construction the section $`W(\xi ,\eta )`$ of $`𝒜`$ corresponds to the function $`𝒢_0𝔤`$ given by $$d\tau (\xi ^h,\eta ^h)+[\tau (\xi ^h),\tau (\eta ^h)],$$ where the subscript $`h`$ denotes the horizontal lift with respect to the principal connection $`\tau _0`$. Thus, the $`𝔤_0`$–components of $`\tau (\xi ^h)`$ and $`\tau (\eta ^h)`$ are automatically zero, so we may write $`[\tau (\xi ^h),\tau (\eta ^h)]=`$ $`[\tau _{}(\xi ^h),\tau _{}(\eta ^h)]+[\tau _+(\xi ^h),\tau _{}(\eta ^h)]+`$ $`[\tau _{}(\xi ^h),\tau _+(\eta ^h)]+[\tau _+(\xi ^h),\tau _+(\eta ^h)].`$ On the other hand, from the definition of the curvature it is clear, that the section $`K(\xi ,\eta )`$ corresponds to the function $$d\tau (\xi ^h,\eta ^h)+[\tau _{}(\xi ^h),\tau _{}(\eta ^h)]+[\tau _+(\xi ^h),\tau _+(\eta ^h)].$$ Now $`\tau _+(\xi ^h)`$ is exactly the function corresponding to $`\text{P}(\xi )`$, while $`\tau _{}(\eta ^h)`$ is the function corresponding to $`\eta `$. (Actually, by construction $`\tau _{}(\eta )`$ has values in $`𝔤_j`$ only, but this is not important here.) Since the algebraic bracket $`\{,\}`$ is simply induced by the Lie bracket on $`𝔤`$, the formula for $`W(\xi ,\eta )`$ follows immediately. To see the second statement, one just has to notice that the algebraic bracket is by definition homogeneous of degree zero, while all nonzero homogeneous components of P have degree at least two. ∎ ### 4.13. Remark Note that the latter result, together with the formula (1) for the change of the Weyl–curvature of a Weyl–structure from 4.6 and the formula (2) for the change of the Rho–tensor from 3.9, gives us a formula for the change of the total curvature of a Weyl–structure under the change of the Weyl–structure. ### 4.14. The construction of normal Weyl–forms Now we are ready to describe the procedure of step by step reducing the condition of normality of a Weyl–form $`\tau \mathrm{\Omega }^1(𝒢_0,𝔤)`$ to a condition on $`\tau _0`$ and at the same time computing step by step the Rho–tensor. From Proposition 4.12 we know that $`W^{(1)}=K^{(1)}`$ and from 4.6 we know that this is actually the same expression for any normal Weyl–form. Usually, this can be computed in advance, and thus gives us a condition on the restriction of $`\tau _i`$ to $`T^{i1}𝒢_0`$ for $`i0`$. Next, by Proposition 4.12, we have $`W^{(2)}(\xi ,\eta )`$ $`=K^{(2)}(\xi ,\eta )+\{\text{P}^{(2)}(\xi ),\eta \}\{\text{P}^{(2)}(\eta ),\xi \}`$ $`=(K^{(2)}\text{P}^{(2)})(\xi ,\eta ).`$ If $`W^{(2)}`$ is to be $`^{}`$–closed, then this implies that $`^{}(K^{(2)})=^{}\text{P}^{(2)}`$. On the other hand, since $`H^1(𝔤_{},𝔤)`$ is concentrated in homogeneous degrees less or equal to one and $`H^0(𝔤_{},𝔤)=𝔤_k`$, the Hodge decomposition implies that $`\text{P}^{(2)}=\mathrm{}^1^{}\text{P}\alpha _2`$ for a unique smooth section $`\alpha _2`$ of $`𝒜_2`$. Moreover, since $`\text{P}^{(2)}`$ has to have values in $`𝒜_+`$, it follows that the restriction of $`\mathrm{}^1^{}(K^{(2)})`$ to $`𝒜_k\mathrm{}𝒜_2`$ must be given by $`(\alpha _2)`$, which gives a condition on the restriction of $`\tau _i`$ to $`T^{i2}𝒢_0`$ for $`i0`$. If this is satisfied, then $`\alpha _2`$ is uniquely determined, and we can compute $`\text{P}^{(2)}`$ as $`\mathrm{}^1^{}(K^{(2)})\alpha _2`$. Let us notice, how simple the latter step gets for $`|1|`$–graded examples: then there is no $`\alpha _2`$, the entire forms P and $`K_0`$ are of homogeneous degree two, and so P is simply obtained in the unique way by the formula $`\text{P}=\mathrm{}^1^{}K_0`$ promised in Example 4.8. Now this process can be easily iterated. We next consider $`K^{(3)}`$ which depends only on the (known) component $`\text{P}^{(2)}`$ of the Rho–tensor and on the restrictions of $`\tau _i`$ to $`T^{i3}𝒢_0`$ for $`i0`$. As above, the restriction of $`\mathrm{}^1^{}(K^{(3)})`$ to $`𝒜_k\mathrm{}𝒜_3`$ must be given by $`(\alpha _3)`$ for a section $`\alpha _3`$ of $`𝒜_3`$, which gives conditions on the restrictions of $`\tau _i`$ for $`i0`$. If these are satisfied, $`\alpha _3`$ is uniquely determined, and we can compute $`\text{P}^{(3)}`$. Finally, once we have reached $`K^{(k)}`$, there are no more conditions, since $`\tau _0`$ is already completely determined at this stage, so we only get a way to compute the remaining homogeneous components of the Rho–tensor.
warning/0001/hep-ph0001037.html
ar5iv
text
# The 𝐾_𝐿→𝜋⁰⁢𝜈⁢𝜈̄ Decay in Models of Extended Scalar Sector ## I Introduction Within the standard model (SM) the $`K_L\pi ^0\nu \overline{\nu }`$ decay is known to be a CP violating (CPV) process to a very good approximation , and subject to a clean theoretical interpretation . In the SM, CP conserving (CPC) contributions are chirally suppressed and smaller by many orders of magnitude than the CPV ones . In a previous work , we calculated the CPC contributions that arise when the SM Lagrangian is extended to include neutrino mass terms. These contributions are known to be chirally enhanced, but we found that they are suppressed by a factor of order $`m_\nu ^2/m_W^2`$ which makes them negligibly small. In this work we study models where, in addition to neutrino mass terms, there are new light scalars. The question that we ask is whether the fact that such scalars can have Yukawa couplings of order one to neutrinos allows for a situation where the new contributions are chirally enhanced while avoiding the $`m_\nu ^2/m_W^2`$ suppression factor. Specifically, we consider the following two models: 1. Neutrino masses are related to the vacuum expectation value of an $`SU(2)`$-triplet scalar. We present the model and calculate the new contributions to the $`K_L\pi ^0\nu \overline{\nu }`$ decay in section II. 2. Neutrino masses are related to loop diagrams involving a singly charged $`SU(2)`$-singlet and an additional $`SU(2)`$-doublet scalar. We present the model and calculate the new contributions to the $`K_L\pi ^0\nu \overline{\nu }`$ decay in section III. A summary of our conclusions is given in section IV. ## II An $`SU(2)`$ Triplet Scalar ### A The Model We consider the SM Lagrangian with the addition of an SU(2) triplet Higgs field ($`\mathrm{\Delta }`$). The most general scalar potential is given by : $`V(H,\mathrm{\Delta })`$ $`=`$ $`{\displaystyle \frac{\lambda _H}{2}}(H_i^{}H^i)^2+{\displaystyle \frac{\lambda _\mathrm{\Delta }}{2}}(\mathrm{\Delta }_t^{}\mathrm{\Delta }^t)^2+\lambda H_i^{}H^i\mathrm{\Delta }_t^{}\mathrm{\Delta }^t+\stackrel{~}{\lambda }H_{}^{i}{}_{}{}^{}\sigma _{ij}^qH^j\mathrm{\Delta }_{}^{r}{}_{}{}^{}t_{rs}^q\mathrm{\Delta }^s`$ (1) $`+`$ $`m(\overline{\mathrm{\Delta }}^0H^0H^0+\sqrt{2}\mathrm{\Delta }^{}H^+H^0+\mathrm{\Delta }^{}H^+H^+)+\mu _H^2H_i^{}H^i+\mu _\mathrm{\Delta }^2\mathrm{\Delta }_t^{}\mathrm{\Delta }^t+h.c,`$ (2) where $`\mathrm{\Delta }^t=[\mathrm{\Delta }^{++},\mathrm{\Delta }^+,\mathrm{\Delta }^0]`$, $`H^i=[H^+,H^0]`$ and $`\sigma ,t`$ are are SU(2) generators in the $`J=1/2,1`$ representations, respectively. The $`m`$-term in (2) breaks the global lepton number symmetry ($`L`$) so that the phenomenologically unacceptable Majoron is avoided. For simplicity, we assume that the couplings in (2) conserve CP. As concerns the VEVs, $`H^0\frac{v_H}{\sqrt{2}}`$ and $`\mathrm{\Delta }^0\frac{v_\mathrm{\Delta }}{\sqrt{2}}`$, we assume for simplicity that they are both real and we take into account the constraints from the $`\rho `$ parameter : $$\frac{v_\mathrm{\Delta }}{v_H}<10^2.$$ (3) In order to calculate the contribution to the decay, the scalar mass eigenstates and mixing angles should be identified. With the assumption that the scalar potential $`V(H,\mathrm{\Delta })`$ (2) is real, the imaginary and the real parts of the neutral scalars remain unmixed. For the CP even fields, $`\sqrt{2}e(H^0)`$ and $`\sqrt{2}e(\mathrm{\Delta }^0)`$, we have the following mass matrix (we neglect terms that are higher order in $`v_\mathrm{\Delta }/v_H`$ and assume that the $`\lambda `$’s are positive and of order one): $$M_{}^2=v_H^2\left(\begin{array}{cc}\lambda _H& \frac{\sqrt{2}}{v_H}(\widehat{\lambda }\frac{v_\mathrm{\Delta }}{\sqrt{2}}+m)\\ \frac{\sqrt{2}}{v_H}(\widehat{\lambda }\frac{v_\mathrm{\Delta }}{\sqrt{2}}+m)& \frac{m}{\sqrt{2}v_\mathrm{\Delta }}\end{array}\right),$$ (4) where $`\widehat{\lambda }\lambda +\stackrel{~}{\lambda }`$. The eigenvectors of $`M_{}^2`$ are: $`\xi _1`$ $`=e(H^0)\mathrm{cos}\gamma +e(\mathrm{\Delta }^0)\mathrm{sin}\gamma `$ (5) $`\xi _2`$ $`=e(H^0)\mathrm{sin}\gamma +e(\mathrm{\Delta }^0)\mathrm{cos}\gamma ,`$ (6) where $`\mathrm{tan}\gamma `$ $`=`$ $`{\displaystyle \frac{v_H}{2v_\mathrm{\Delta }}}\left(\widehat{\lambda }+{\displaystyle \frac{\sqrt{2}m}{v_\mathrm{\Delta }}}\right)^1\left[\lambda _H+{\displaystyle \frac{m}{\sqrt{2}v_\mathrm{\Delta }}}\sqrt{\left(\lambda _H+{\displaystyle \frac{m}{\sqrt{2}v_\mathrm{\Delta }}}\right)^2+4{\displaystyle \frac{v_\mathrm{\Delta }^2}{v_H^2}}\left(\widehat{\lambda }+{\displaystyle \frac{\sqrt{2}m}{v_\mathrm{\Delta }}}\right)^2}\right]`$ (7) $``$ $`2{\displaystyle \frac{v_\mathrm{\Delta }}{v_H}}\left({\displaystyle \frac{\widehat{\lambda }+\frac{\sqrt{2}m}{v_\mathrm{\Delta }}}{2\lambda _H+\frac{\sqrt{2}m}{v_\mathrm{\Delta }}}}\right)1.`$ (8) In deriving the inequality in eq. (8) and below we assume that there are no fine-tuned cancellations among the independent parameters of the scalar sector. The eigenvectors correspond to the following mass eigenvalues: $`m_{\xi _{1,2}}^2`$ $`=`$ $`{\displaystyle \frac{v_H^2}{2}}\left[\lambda _H{\displaystyle \frac{m}{\sqrt{2}v_\mathrm{\Delta }}}\pm \sqrt{\left(\lambda _H+{\displaystyle \frac{m}{\sqrt{2}v_\mathrm{\Delta }}}\right)^2+4{\displaystyle \frac{v_\mathrm{\Delta }^2}{v_H^2}}\left(\widehat{\lambda }+{\displaystyle \frac{\sqrt{2}m}{v_\mathrm{\Delta }}}\right)^2}\right]`$ (9) $``$ $`v_H^2[\lambda _H+{\displaystyle \frac{v_\mathrm{\Delta }^2}{v_H^2}}{\displaystyle \frac{\left(\widehat{\lambda }+\frac{\sqrt{2}m}{v_\mathrm{\Delta }}\right)^2}{\lambda _H+\frac{m}{\sqrt{2}v_\mathrm{\Delta }}}},{\displaystyle \frac{m}{\sqrt{2}v_\mathrm{\Delta }}}+O\left(m{\displaystyle \frac{v_\mathrm{\Delta }^2}{v_H^2}}\right)].`$ (10) For the CP odd fields, $`\sqrt{2}m(H^0)`$ and $`\sqrt{2}m(\mathrm{\Delta }^0)`$, we have the following mass matrix: $$M_{}^2=\sqrt{2}m\left(\begin{array}{cc}2v_\mathrm{\Delta }& v_H\\ v_H& \frac{v_H^2}{2v_\mathrm{\Delta }}\end{array}\right).$$ (11) The eigenvectors of $`M_{}^2`$ are: $`G^0`$ $`=`$ $`m(H^0)\mathrm{cos}\eta +m(\mathrm{\Delta }^0)\mathrm{sin}\eta `$ (12) $`J^0`$ $`=`$ $`m(H^0)\mathrm{sin}\eta +m(\mathrm{\Delta }^0)\mathrm{cos}\eta ,`$ (13) where $$|\mathrm{tan}\eta |=\left|\frac{2v_\mathrm{\Delta }}{v_H}\right|1.$$ (14) We learn that mixing in the CP-odd sector is very small and we neglect it from here on. The eigenvectors correspond to the following mass eigenvalues : $$m_{G^0,J^0}^2=[0,\frac{mv_H^2}{\sqrt{2}v_\mathrm{\Delta }}],$$ (15) where the massless component $`G^0\sqrt{2}m(H^0)`$ corresponds to the unphysical Goldstone boson which is eaten by the $`Z`$ boson, while $`J^0\sqrt{2}m(\mathrm{\Delta }^0)`$ corresponds to the would-be Majoron. As expected its squared mass is proportional to the explicit lepton number violating parameter $`m`$. For the singly charged scalars, $`H^\pm `$ and $`\mathrm{\Delta }^\pm `$, we have the following mass matrix: $$M_C^2=\frac{v_H}{\sqrt{2}}(\stackrel{~}{\lambda }\frac{v_\mathrm{\Delta }}{\sqrt{2}}+m)\left(\begin{array}{cc}2\frac{v_\mathrm{\Delta }}{v_H}& \sqrt{2}\\ \sqrt{2}& \frac{v_H}{v_\mathrm{\Delta }}\end{array}\right).$$ (16) The eigenvectors of $`M_C^2`$ are: $`G^\pm `$ $`=H^\pm \mathrm{cos}\eta ^{}+\mathrm{\Delta }^\pm \mathrm{sin}\eta ^{}`$ (17) $`\xi ^\pm `$ $`=H^\pm \mathrm{sin}\eta ^{}+\mathrm{\Delta }^\pm \mathrm{cos}\eta ^{},`$ (18) where $$|\mathrm{tan}\eta ^{}|=\left|\frac{\sqrt{2}v_\mathrm{\Delta }}{v_H}\right|1.$$ (19) We learn that mixing in the charged sector is very small and we neglect it from here on. The eigenvectors correspond to the following mass eigenvalues: $$m_{G^\pm ,\xi ^\pm }^2=[0,\frac{v_H^2}{\sqrt{2}}\left(\frac{\stackrel{~}{\lambda }}{\sqrt{2}}+\frac{m}{v_\mathrm{\Delta }}\right)],$$ (20) where the massless component $`G^\pm H^\pm `$ corresponds to the unphysical charged Goldstone boson which is eaten by the $`W^\pm `$ bosons, while $`\xi ^\pm \mathrm{\Delta }^\pm `$ corresponds to a new physical charged scalar. Neutrino Majorana masses are induced via the following interaction terms: $$_{\nu \nu }=f_{mn}(L_{Li}^m)_\alpha \tau _{\alpha \beta }^2\sigma _{ir}^2\sigma _{rj}^t(L_{Lj}^n)_\beta \mathrm{\Delta }^t,$$ (21) where $`\tau `$ is an SU(2) generator in the $`J=1/2`$ spinor representation. For simplicity, we take $`f_{mn}=f_m\delta _{mn}`$ with $`f_m`$ real. In this way, we avoid unnecessary complications related to flavor mixing and CP violation in the neutrino sector (see discussion in ref. ). The interaction term (21) induces neutrino masses: $$(m_\nu ^{Maj})_i=f_i\mathrm{\Delta }^0$$ (22) with $`i=1,2,3`$. In addition, it generates new contributions to the $`K_L\pi ^0\nu \overline{\nu }`$ decay. ### B The $`K_L\pi ^0\nu \overline{\nu }`$ Decay Rate The dominant new contributions to the $`K_L\pi ^0\nu \overline{\nu }`$ decay come from the diagrams presented in fig. 1, which generate the following effective Hamiltonian: $$_{eff}^\mathrm{\Delta }=\underset{\mathrm{}}{}\frac{G_F}{\sqrt{2}}\frac{\alpha }{2\pi \mathrm{sin}^2\mathrm{\Theta }_W}(\overline{d}s)(\nu _{\mathrm{}}^Ti\tau ^2\nu _{\mathrm{}})(X_{\mathrm{}}^{(a)}+X_{\mathrm{}}^{(b)}+X_{\mathrm{}}^{(c)}),$$ (23) where the $`X_{\mathrm{}}^{(i)}`$ function corresponds to the diagram in fig 1(i). In the $`X_{\mathrm{}}^{(a)}`$ and $`X_{\mathrm{}}^{(b)}`$ terms, the L breaking effects enter through, respectively, the $`\mathrm{\Delta }WW`$ and $`\mathrm{\Delta }HH`$ couplings and is related to the spontaneous breaking ($`\mathrm{\Delta }^00`$): $`X_{\mathrm{}}^{(a)}+X_{\mathrm{}}^{}{}_{}{}^{(b)}`$ $`=`$ $`m_\nu ^{\mathrm{}}m_s{\displaystyle \underset{i}{}}\lambda _i\left({\displaystyle \frac{\mathrm{sin}^2\gamma }{m_{\xi _1}^2}}+{\displaystyle \frac{\mathrm{cos}^2\gamma }{m_{\xi _2}^2}}\right)\left({\displaystyle \frac{1}{2}}+2x_i{\displaystyle \frac{\lambda \stackrel{~}{\lambda }}{g^2}}\right)`$ (24) $`\times `$ $`\left[1+{\displaystyle \frac{x_i}{(x_i1)^3}}\left(2x_i\mathrm{ln}(x_i)+1x_i^2\right)\right],`$ (25) where $`x_i=(m_i/M_W)^2`$ and $`\lambda _i=V_{is}^{}V_{id}`$. Since the top quark contribution is dominant, (25) can be simplified: $`X_{\mathrm{}}^{(a)}+X_{\mathrm{}}^{(b)}`$ $``$ $`\lambda _tm_\nu ^{\mathrm{}}m_s\left({\displaystyle \frac{\mathrm{sin}^2\gamma }{m_{\xi _1}^2}}+{\displaystyle \frac{\mathrm{cos}^2\gamma }{m_{\xi _2}^2}}\right)`$ (26) $`\times `$ $`x_t\left[\left({\displaystyle \frac{1}{2}}+2x_t{\displaystyle \frac{\lambda \stackrel{~}{\lambda }}{g^2}}\right){\displaystyle \frac{\left(2x_t\mathrm{ln}(x_t)+1x_t^2\right)}{(x_t1)^3}}+2{\displaystyle \frac{\lambda \stackrel{~}{\lambda }}{g^2}}\right].`$ (27) In the $`X_{\mathrm{}}^{(c)}`$ term, the L breaking effect enters through the $`\mathrm{\Delta }H`$ mixing and is related to both the soft breaking ($`m0`$) and the nonzero VEV of $`\mathrm{\Delta }^0`$. The calculation of $`X_{\mathrm{}}^{(c)}`$ is simplified by the use of the $`sdH`$ effective coupling $`\mathrm{\Gamma }_{sdH}`$, \[represented by a square in fig. 1(c)\] which was calculated within the SM in ref. . For our model $`\mathrm{\Gamma }_{sdH}`$ should be expressed in terms of the appropriate masses and mixing angles. The mixing of the charged scalar can, however, be neglected \[see eq. (19)\] and therefore we use directly the SM calculation: $$\mathrm{\Gamma }_{sdH}=\lambda _t\frac{g^3}{128\pi ^2}\frac{m_t^2m_s}{M_W^3}\left(\frac{3}{2}+\frac{4\lambda _H}{g^2}f_2(x_t)\right)(1+\gamma ^5),$$ (28) where $$f_2(x)=\frac{x}{2(1x)^2}\left(\frac{x}{1x}\mathrm{ln}x+\frac{2}{1x}\mathrm{ln}x\frac{1}{2}\frac{3}{2x}\right).$$ (29) Then $`X_{\mathrm{}}^{(c)}`$ is given by $`X_{\mathrm{}}^{(c)}=\lambda _tm_\nu ^{\mathrm{}}m_s{\displaystyle \frac{\mathrm{sin}2\gamma }{8}}{\displaystyle \frac{v_H}{v_\mathrm{\Delta }}}\left({\displaystyle \frac{1}{m_{\xi _1}^2}}{\displaystyle \frac{1}{m_{\xi _2}^2}}\right)x_t\left({\displaystyle \frac{3}{2}}+{\displaystyle \frac{4\lambda _H}{g^2}}f_2(x_t)\right).`$ (30) Note that since the lepton and quark operators in the effective Hamiltonian $`_{eff}^\mathrm{\Delta }`$ in eq. (23) are scalar operators, the new contributions are CPC. We are interested in finding the upper bound on the new contributions. We therefore focus on the region in parameter space that maximizes them. For all scalar masses, we will use a lower bound of 45 $`GeV`$, thus avoiding any conflict with constraints from the invisible width of the $`Z`$ boson. From eqs. (15) and (20) we then find: $$\mathrm{sign}(m/v_\mathrm{\Delta })=1,|m/v_\mathrm{\Delta }|>1.$$ (31) Substituting the proper values for the masses and angles (8), (10) into the functions $`X_{\mathrm{}}^{(a)},X_{\mathrm{}}^{(b)}`$ (25) and $`X_{\mathrm{}}^{(c)}`$ (30) we find: $`X_{\mathrm{}}^{(a)}+X_{\mathrm{}}^{(b)}\lambda _t{\displaystyle \frac{m_\nu ^{\mathrm{}}m_s}{v_H^2}}\left({\displaystyle \frac{v_\mathrm{\Delta }}{m}}+O\left(v_\mathrm{\Delta }^2/v_H^2\right)\right)`$ (32) $`X_{\mathrm{}}^{(c)}\lambda _t{\displaystyle \frac{m_\nu ^{\mathrm{}}m_s}{v_H^2}}\left({\displaystyle \frac{v_H^2}{\lambda _Hv_H^2+mv_\mathrm{\Delta }}}+O\left(v_\mathrm{\Delta }^2/v_H^2\right)\right)`$ (33) For $`mv_H`$ $`X_{\mathrm{}}^{(a)}+X_{\mathrm{}}^{(b)}`$ and $`X_{\mathrm{}}^{(c)}`$ are all highly suppressed, together with the total rate. Larger contributions are found with $`m`$ in the intermediate regime $`v_\mathrm{\Delta }<m<v_H`$, where $`X_{\mathrm{}}^{(a)}+X_{\mathrm{}}^{(b)}`$ and $`X_{\mathrm{}}^{(c)}`$ are all of the order of $`\frac{m_\nu ^{\mathrm{}}m_s}{v_H^2}\frac{m_\nu ^{\mathrm{}}m_s}{m_{\xi _{1,2}}^2}`$. The $`\nu _m\nu _m\mathrm{\Delta }`$ coupling, that is $`f_m`$ of eq. (21), is proportional to $`m_{\nu _m}/v_\mathrm{\Delta }`$. Naively, one may think that this mechanism of inducing neutrino masses can induce a contribution to the $`K_L\pi \nu \nu `$ rate that is enhanced by a factor of order $`v_H/v_\mathrm{\Delta }`$ compared to the mechanisms of . We find that this is not the case. For the diagrams in fig. 1(a,b), there is a $`v_\mathrm{\Delta }`$-factor in the $`WW\mathrm{\Delta }`$ and $`HH\mathrm{\Delta }`$ couplings. For the diagram in fig. 1(c), there is a factor of $`\mathrm{sin}2\gamma v_H(1/m_{\xi _1}^21/m_{\xi _2}^2)v_\mathrm{\Delta }/v_H^2`$. In either case, the final result is proportional to $`f_mv_\mathrm{\Delta }m_{\nu _m}`$ and there is no enhancement. We are now in a position to compare the contribution of the triplet scalar to the leading, CPV one : $$R_{CPV}^\mathrm{\Delta }=\frac{\mathrm{\Gamma }_\mathrm{\Delta }(K_L\pi ^0\nu \overline{\nu })}{\mathrm{\Gamma }_{CPV}(K_L\pi ^0\nu \overline{\nu })}<\left(\frac{M_K}{m_{\xi _{light}}}\frac{m_\nu }{m_{\xi _{light}}}\right)^210^{11}\left[\frac{m_\nu }{10MeV}\right]^2\left[\frac{45GeV}{m_{\xi _{light}}}\right]^4,$$ (34) where $`m_{\xi _{light}}`$ stands for the lighter between $`m_{\xi _1}`$ and $`m_{\xi _2}`$. While the direct bound is $`m_{\nu _\tau }18.2MeV`$, there is a significantly stronger bound from cosmology, $`m_\nu <10eV`$ . Therefore, very likely, $`R_{CPV}^\mathrm{\Delta }<10^{23}`$. ## III The Zee Model ### A The Model The Zee model enlarges the SM scalar sector by a charged $`SU(2)`$ singlet $`\varphi ^+`$ and an $`SU(2)`$ doublet $`H_2`$. The SM doublet is denoted by $`H_1`$. The Lagrangian of the model is : $`_{Zee}_{SM}+f_{mn}ϵ_{\alpha \beta }(L_{}^{m\alpha }{}_{}{}^{T}CL_{}^{n\beta }{}_{L}{}^{})\varphi ^++\mu (H_1^{}H_2+H_2^{}H_1)+V(H_1,H_2,\varphi ^+),`$ (35) where $`L`$ is a lepton doublet, $`\alpha ,\beta `$ are $`SU(2)`$ indices, $`m,n`$ are flavor indices, $`C`$ is the Dirac charge conjugation matrix and $`V(H_1,H_2,\varphi ^+)`$ is the most general scalar potential which respects the SM gauge symmetries and $`L`$ : $`V(H_1,H_2,\varphi ^+)`$ $``$ $`\lambda _1\left(\left|H_1\right|^2{\displaystyle \frac{a_1^2}{2}}\right)^2+\lambda _2\left(\left|H_2\right|^2{\displaystyle \frac{a_2^2}{2}}\right)^2+\lambda _3\left(\left|H_1\right|^2{\displaystyle \frac{a_1^2}{2}}\right)\left(\left|H_2\right|^2{\displaystyle \frac{a_2^2}{2}}\right)`$ (36) $`+`$ $`\lambda _4(\left|H_1\right|^2\left|H_2\right|^2H_{1}^{}{}_{}{}^{}H_2H_{2}^{}{}_{}{}^{}H_1)+\lambda _5\left|H_{1}^{}{}_{}{}^{T}i\tau ^2H_2\right|^2+\lambda _6m^2\left|\varphi ^+\right|^2`$ (37) $`+`$ $`\lambda _7|\varphi ^+|^4+\lambda _8|\varphi ^+|^2|H_1|^2+\lambda _9|\varphi ^+|^2|H_2|^2+M(H_{}^{1}{}_{}{}^{T}i\tau ^2H^2\varphi ^{}+h.c).`$ (38) Both Higgs doublets develop nonzero VEVs: $$H_{}^{0}{}_{1,2}{}^{}\frac{v_{1,2}}{\sqrt{2}}.$$ (39) For $`\mu =0`$, the parameters $`a_{1,2}`$ would be equal to $`v_{1,2}`$ respectively. With a nonzero $`\mu `$ term, $`v_{1,2}`$ are complicated functions of $`a_i`$, $`\lambda _i`$ and $`\mu `$. The $`L`$-charges of the scalars fields are: $$L(H_1)=0,L(H_2)=2,L(\varphi ^+)=2.$$ (40) Then we see that $`L`$ is broken spontaneously via the nonzero VEV of $`H_2`$. The $`\mu `$-term in eq. (35) breaks $`L`$ explicitly so that the phenomenologically unacceptable Majoron is avoided. For simplicity, we assume that CP is conserved in the lepton sector and take all the dimensionless scalar couplings to be of order one. In order to calculate the new contribution to the decay, the scalar mass eigenstates and mixing angles should be identified. There are seven physical and three unphysical combinations (eaten by the $`Z`$ and the $`W^\pm `$ bosons). Out of the seven physical combinations, four are charged and three are neutral. The neutral ones are the would-be Majoron \[which gains mass due to the $`\mu `$ term in eq. (35)\] and two other real fields. As long as $`\mu `$ is real, the real and imaginary parts of $`H_{1,2}^0`$ remain unmixed (as in the cases in which L is spontaneously broken ), thus the imaginary physical combination corresponds to the would-be Majoron which is irrelevant to our calculation. The real neutral mass matrix is read from the scalar potential of eq. (35): $$M_{}=\frac{1}{2}\left(\begin{array}{cc}2\lambda _1(3v_{1}^{}{}_{}{}^{2}a_{1}^{}{}_{}{}^{2})+\lambda _3(v_{2}^{}{}_{}{}^{2}a_{2}^{}{}_{}{}^{2})& 2\lambda _3v_1v_2+\mu \\ 2\lambda _3v_1v_2+\mu & 2\lambda _2(3v_{2}^{}{}_{}{}^{2}a_{2}^{}{}_{}{}^{2})+\lambda _3(v_{1}^{}{}_{}{}^{2}a_{1}^{}{}_{}{}^{2})\end{array}\right).$$ (41) The eigenvectors of $`M_{}`$ are: $`\rho _1`$ $`=`$ $`e(H_1^0)\mathrm{cos}\theta +e(H_2^0)\mathrm{sin}\theta `$ (42) $`\rho _2`$ $`=`$ $`e(H_1^0)\mathrm{sin}\theta +e(H_2^0)\mathrm{cos}\theta .`$ (43) Each eigenvector corresponds to an eigenvalue denoted by $`m_{\rho _{1,2}}`$. The masses $`m_{\rho _{1,2}}`$ and $`\mathrm{tan}\theta `$ are complicated functions of $`a_i`$, $`\lambda _i`$ and $`\mu `$. The charged mass matrix is: $$M_C=\frac{1}{2}\left(\begin{array}{ccc}\overline{\lambda }v_2^22\mu \frac{v_2}{v_1}& \overline{\lambda }v_1v_2+2\mu & \sqrt{2}Mv_2\\ \overline{\lambda }v_1v_2+2\mu & \overline{\lambda }v_1^22\mu \frac{v_1}{v_2}& \sqrt{2}Mv_1\\ \sqrt{2}Mv_2& \sqrt{2}Mv_1& 2\lambda _6m^2+\lambda _8v_1^2+\lambda _9v_2^2\end{array}\right),$$ (44) where $`\overline{\lambda }\lambda _4+\lambda _5`$. The massless combination of the charged fields (see e.g refs. ) is: $$G^\pm =H_1^\pm \mathrm{cos}\delta +H_2^\pm \mathrm{sin}\delta ,$$ (45) with $$\mathrm{tan}\delta =\frac{v_2}{v_1}.$$ (46) The massless combination is not affected by the addition of the singlet field $`\varphi ^+`$, nor by the $`\mu `$ term, since it is determined (according to the Goldstone theorem) only by the broken $`SU(2)_L`$ generators. The physical charged fields are given by the following linear combinations of $`H_{1,2}^\pm `$ and $`\varphi ^\pm `$ (which must be orthogonal to $`G^\pm `$): $`\chi _1^\pm `$ $`=`$ $`(H_1^\pm \mathrm{sin}\delta +H_2^\pm \mathrm{cos}\delta )\mathrm{cos}\beta +\varphi ^\pm \mathrm{sin}\beta `$ (47) $`\chi _2^\pm `$ $`=`$ $`(H_1^\pm \mathrm{sin}\delta +H_2^\pm \mathrm{cos}\delta )\mathrm{sin}\beta +\varphi ^\pm \mathrm{cos}\beta ,`$ (48) with $`\mathrm{tan}\beta `$ being a complicated function of the scalar potential parameters. The new interactions in eq. (35) induced a neutrino Majorana mass matrix $`M_m\mathrm{}^\nu `$. For simplicity we assume the dominance of the one loop induced mass : $$M_m\mathrm{}^\nu =\frac{2\sqrt{2}}{(4\pi )^2}f_m\mathrm{}\mathrm{tan}\delta \mathrm{sin}2\beta \frac{(m_{\mathrm{}}^2m_m^2)}{M_W}\mathrm{ln}\left(\frac{m_{\chi _2}}{m_{\chi _1}}\right),$$ (49) with $`m,\mathrm{}`$ flavor indices. ### B The $`K_L\pi ^0\nu \overline{\nu }`$ Decay Rate The new interactions also generate new contributions to the $`K_L\pi ^0\nu \overline{\nu }`$ decay. Note that the one loop induced mass and the new contributions to the decay (shown in fig. 2) are related to mixing between the charged scalars and therefore vanish in the limit $`\delta ,\beta 0`$. Below we concentrate on the contributions which are dominant when $`\mathrm{sin}\delta `$ is small. Contributions that are of higher order do not modify significantly the results, even with $`\mathrm{sin}\delta 1`$, and thus they are omitted. Note that, since $`v_1`$ induces the top mass, we always have $`\mathrm{tan}\delta <1`$. The scalar operator is induced by the neutral Higgs mediated penguin diagram shown in fig. 2 . The square in the figures represents an effective $`sdH_{1,2}`$ vertex denoted by $`\mathrm{\Gamma }_{sdH_{1,2}}`$, similar to the one we encountered in section II.B . The effective vertices $`\mathrm{\Gamma }_{sdH_{1,2}}`$ induced by the Zee model fields differ from the one calculated within the SM . Since, however, we are only interested in finding an upper bound on the decay rate, we can simply set the charged mixing angles factors to one, and replace the boson masses in the propagator (as explained in section II.B) with $`m_{\chi _{light}}45GeV`$. Then we get: $$\mathrm{\Gamma }_{sdH_{1,2}}<\frac{g^3\lambda _t}{128\pi ^2}\frac{m_t^2m_s}{m_{\chi _{light}}^3}(1+\gamma ^5)\times [C_{sdH_1},C_{sdH_2}],$$ (50) where $`C_{sdH_i}`$ is a constant of $`O(10)`$. Neglecting subdominant contributions of $`O[(m_{\mathrm{}}/m_{\rho _i,\chi _i})^4]`$, the diagrams in fig. 2 generate the following CPC effective Hamiltonian: $$_{eff}^\varphi =\frac{G_F}{\sqrt{2}}\frac{\alpha }{2\pi \mathrm{sin}^2\mathrm{\Theta }_W}\lambda _t(\overline{d}s)(\nu _m^Ti\sigma ^2\nu _L\mathrm{})(Y_\mathrm{}m^{(a)}+Y_\mathrm{}m^{(b)}),$$ (51) where the $`Y_\mathrm{}m^{(i)}`$ function corresponds to the diagram in fig. 2(i). $`Y_\mathrm{}m^{(a)}`$ is given by: $`Y_\mathrm{}m^{(a)}`$ $`=`$ $`2{\displaystyle \frac{M_\mathrm{}m^\nu M_W}{\mathrm{cos}\delta }}{\displaystyle \frac{m_t^2m_s}{m_{\chi _{light}}^3}}\left[C_{sdH_1}\left({\displaystyle \frac{\mathrm{cos}^2\theta }{m_{\rho _1}^2}}+{\displaystyle \frac{\mathrm{sin}^2\theta }{m_{\rho _2}^2}}\right)+C_{sdH_2}{\displaystyle \frac{\mathrm{sin}2\theta }{2}}\left({\displaystyle \frac{1}{m_{\rho _1}^2}}{\displaystyle \frac{1}{m_{\rho _2}^2}}\right)\right].`$ (52) The calculation of the diagram in fig. 2(b) is much more involved. This is due to the fact that each of the trilinear couplings $`H_i^0H_j^\pm H_k^{}`$ and $`H_i^0H_j^\pm \varphi ^{}`$ ($`i,j,k=1,2`$) translates into a set of eight couplings in the mass basis. In order to estimate the upper bound on the rate we set (again) all mixing factors to unity and replace the charged scalar propagators by their maximal values, i.e. $`\frac{1}{k^2m_{\chi _1}^2}\frac{1}{k^2m_{\chi _2}^2}\frac{1}{k^2m_{\chi _{light}}^2}`$ and $`\frac{1}{k^2m_{\chi _1}^2}+\frac{1}{k^2m_{\chi _2}^2}\frac{2}{k^2m_{\chi _{light}}^2}`$ . We then find: $`Y_\mathrm{}m^{(b)}`$ $`<`$$``$ $`{\displaystyle \frac{\sqrt{2}}{g64\pi ^2}}f_\mathrm{}m(m_{\mathrm{}}^2m_m^2){\displaystyle \frac{M_W^2}{m_{\rho _1}^2}}{\displaystyle \frac{m_t^2m_s}{m_{\chi _{light}}^5}}\left[F_1(\theta ,m_{\rho _i},m_{\chi _i},M)+F_2(\theta ,m_{\rho _i},m_{\chi _i},M)\right]`$ (53) $`=`$ $`{\displaystyle \frac{1}{8g}}{\displaystyle \frac{\mathrm{cot}\delta }{\mathrm{sin}2\beta }}\left(\mathrm{ln}{\displaystyle \frac{m_{\chi _2}}{m_{\chi _1}}}\right)^1{\displaystyle \frac{M_m\mathrm{}^\nu m_s}{m_{\rho _1}^2}}{\displaystyle \frac{m_t^2M_W^3}{m_{\chi _{light}}^5}}`$ (54) $`\times `$ $`\left[F_1(\theta ,m_{\rho _i},m_{\chi _i},M)+F_2(\theta ,m_{\rho _i},m_{\chi _i},M)\right],`$ (55) where $`F_1(\theta ,m_{\rho _i},m_{\chi _i},M)`$ $`=`$ $`\left[C_{sdH_1}\left(\mathrm{cos}^2\theta +\mathrm{sin}^2\theta {\displaystyle \frac{m_{\rho _1}^2}{m_{\rho _2}^2}}\right)+C_{sdH_2}{\displaystyle \frac{\mathrm{sin}2\theta }{2}}\left(1+{\displaystyle \frac{m_{\rho _1}^2}{m_{\rho _2}^2}}\right)\right]`$ (56) $`\times `$ $`\left(2\lambda _1+{\displaystyle \frac{1}{2}}\overline{\lambda }+2\lambda _3+2\lambda _8+3{\displaystyle \frac{M}{v_1}}\right),`$ (57) and $`F_2(\theta ,m_{\rho _i},m_{\chi _i},M)`$ $`=`$ $`\left[C_{sdH_2}\left(\mathrm{sin}^2\theta +\mathrm{cos}^2\theta {\displaystyle \frac{m_{\rho _1}^2}{m_{\rho _2}^2}}\right)C_{sdH_1}{\displaystyle \frac{\mathrm{sin}2\theta }{2}}\left(1+{\displaystyle \frac{m_{\rho _1}^2}{m_{\rho _2}^2}}\right)\right]`$ (58) $`\times `$ $`\left(2\lambda _2+2\lambda _3+{\displaystyle \frac{1}{2}}\overline{\lambda }+2\lambda _95{\displaystyle \frac{M}{v_1}}\right).`$ (59) The $`M`$-term in the scalar potential (38) does not break any symmetry. One might think then that the ratio $`\frac{M}{v_1}`$ that appears in eqs. (57) can be arbitrarily large and enhance the decay rate. This is however not the case. For very large $`M`$, the scalar potential would spontaneously break $`U(1)_{EM}`$. An even stronger upper bound on $`M`$ comes from two loop contributions to the neutrino masses. We assumed, for simplicity, that the neutrino masses are dominated by the one loop contributions. This assumption requires that $`\frac{M}{v_1}`$ is not much larger than the various $`\lambda `$ couplings. Adding (52) to (55), and applying the approximation described above eq. (55) we find: $`Y_\mathrm{}m^{(a)}+Y_\mathrm{}m^{(b)}`$ $`<`$$``$ $`{\displaystyle \frac{m_{\nu _3}M_W}{m_{\rho _{light}}^2}}{\displaystyle \frac{m_t^2m_s}{m_{\chi _{light}}^3}}\left[C_{sdH_1}\left(4+3{\displaystyle \frac{M_W^2}{m_{\chi _{light}}^2}}\right)+C_{sdH_2}\left(1+3{\displaystyle \frac{M_W^2}{m_{\chi _{light}}^2}}\right)\right]`$ (60) $`<`$$``$ $`10^2{\displaystyle \frac{m_{\nu _3}M_W}{m_{\rho _{light}}^2}}{\displaystyle \frac{m_t^2m_s}{m_{\chi _{light}}^3}},`$ (61) where $`m_{\nu _3}`$ is the largest eigenvalue of $`M_\mathrm{}m^\nu `$. We can now compare the CPC rate of the Zee model with the leading CPV one : $`R_{CPV}^{Zee}={\displaystyle \frac{\mathrm{\Gamma }_{Zee}(K_L\pi ^0\nu \overline{\nu })}{\mathrm{\Gamma }_{CPV}(K_L\pi ^0\nu \overline{\nu })}}`$ $`<`$$``$ $`\left(10^2{\displaystyle \frac{m_\nu M_W}{m_{\rho _{light}}^2}}{\displaystyle \frac{m_t^2m_K}{m_{\chi _{light}}^3}}\right)^2`$ (62) $``$ $`10^5\left[{\displaystyle \frac{m_\nu }{10MeV}}\left({\displaystyle \frac{45GeV}{m_{\rho _{light}}}}\right)^2\left({\displaystyle \frac{45GeV}{m_{\chi _{light}}}}\right)^3\right]^2.`$ (63) Since, very likely, $`m_{\nu _\tau }10eV`$ , we expect $`R_{CPV}^\mathrm{\Delta }<10^{17}`$. Recall that this upper bound was obtained using some crude approximations and is expected to be even smaller for an exact calculation. ## IV Final Conclusions In this work we examined the question of whether the SM CP violating contribution to the $`K_L\pi ^0\nu \overline{\nu }`$ decay is still dominant in the presence of new scalars that induce Majorana masses for neutrinos. We found the following unambiguous answers: * For CPC contributions induced by an $`SU(2)`$-triplet scalar, we get: $$\frac{\mathrm{\Gamma }_\mathrm{\Delta }(K_L\pi ^0\nu \overline{\nu })}{\mathrm{\Gamma }_{SM}(K_L\pi ^0\nu \overline{\nu })}<\left(\frac{M_Km_\nu }{m_{\xi _{light}}^2}\right)^2<10^{11}.$$ (64) * For CPC contributions that are generated in the Zee model, we get: $$\frac{\mathrm{\Gamma }_{Zee}(K_L\pi ^0\nu \overline{\nu })}{\mathrm{\Gamma }_{SM}(K_L\pi ^0\nu \overline{\nu })}<\left(10^2\frac{m_\nu M_W}{m_{\rho _{light}}^2}\frac{m_t^2m_K}{m_{\chi _{light}}^3}\right)^2<10^5.$$ (65) In obtaining the final bounds in eqs. (64) and (65) we used the direct upper bound on $`m_{\nu _\tau }`$ of order 10 MeV. If we use the cosmological bound, $`m_\nu <10eV,`$ then the bounds become stronger by twelve orders of magnitude. It is clear then that the $`K_L\pi ^0\nu \overline{\nu }`$ decay provides a very clean measurement of fundamental, CP violating properties and that it does not probe neutrino masses. Acknowledgments I thank Yossi Nir for his guidance in this work. I also thank Sven Bergman and Galit Eyal for fruitful discussions.
warning/0001/hep-th0001012.html
ar5iv
text
# 1 Liouville field theory ## 1 Liouville field theory During last 20 years the Liouvlle field theory permanently attracts much attention mainly due to its relevance in the quantization of strings in non-critical space-time dimensions (see also refs.). It is also applied as a field theory of the 2D quantum gravity. E.g., the results of the Liouville field theory (LFT) approach can be compared with the calculations in the matrix models of two-dimensional gravity and this comparison shows that when the LFT central charge $`c_L25`$ this field theory describes the same continuous gravity as was found in the critical region of the matrix models. Although there are still no known applications of LFT with $`c_L<25`$, the theory is interesting on its own footing as an example of non-rational 2D conformal field theory. In the bulk the Liouville field theory is defined by the Lagrangian density $$=\frac{1}{4\pi }(_a\varphi )^2+\mu e^{2b\varphi }$$ (1.1) where $`\varphi `$ is the two-dimenesional scalar field, $`b`$ is the dimensionless Liouville coupling constant and the scale parameter $`\mu `$ is called the cosmological constant. This expression implies a trivial background metric $`g_{ab}=\delta _{ab}`$. In more general background the action reads $$A_{\mathrm{bulk}}=\frac{1}{4\pi }\underset{\mathrm{\Gamma }}{}\left[g^{ab}_a\varphi _b\varphi +QR\varphi +4\pi \mu e^{2b\varphi }\right]\sqrt{g}d^2x$$ (1.2) Here $`R`$ is the scalar curvature associated with the background metric $`g`$ while $`Q`$ is an important quantity in the Liouville field theory called the background charge $$Q=b+1/b$$ (1.3) It determines in particular the central charge of the theory $$c_L=1+6Q^2$$ (1.4) In what follows we always will consider only the simplest topologies like sphere or disk which can be described by a trivial background. For example, a sphere can be represented as a flat projective plane where the flat Liouville lagrangian (1.1) is valid if we put away all the curvature to the spacial infinity where it is seen as a special boundary condition on the Liouville field $`\varphi `$ $$\varphi (z,\overline{z})=Q\mathrm{log}(z\overline{z})+O(1)\mathrm{at}|z|\mathrm{}$$ (1.5) called the background charge at infinity. The basic objects of LFT are the exponential fields $`V_\alpha (x)=\mathrm{exp}(2\alpha \varphi (x))`$ which are conformal primaries w.r.t. the stress tensor $`T(z)`$ $`=(\varphi )^2+Q^2\varphi `$ (1.6) $`\overline{T}(\overline{z})`$ $`=(\overline{}\varphi )^2+Q\overline{}^2\varphi `$ The field $`V_\alpha `$ has the dimension $$\mathrm{\Delta }_\alpha =\alpha (Q\alpha )$$ (1.7) In fact not all of these operators are independent. One has to identify the operators $`V_\alpha `$ and $`V_{Q\alpha }`$ so that the whole set of local LFT fields is obtained by the “folding” of the complex $`\alpha `$-plane w.r.t. this reflection. The only exception is the line $`\alpha =Q/2+iP`$ with $`P`$ real where these exponential fields, if interpreted in terms of quantum gravity, seem not to correspond to local operators. E.g. in the classical theory they appear as hyperbolic solutions to the Lioville equation and “create holes” in the surface . Instead, these values of $`\alpha `$ are attributed to the normalizable states. The LFT space of states $`𝒜`$ consists of all conformal families $`[v_P]`$ corresponding to the primary states $`|v_P`$ with real $`0P<\mathrm{}`$, i.e., $$𝒜=\underset{P[0,\mathrm{})}{}[v_P]$$ (1.8) The primary states $`|v_P`$ are related to the values $`\alpha =Q/2+iP`$ and have dimensions $`Q^2/4+P^2`$ while other values of $`\alpha `$ are mapped onto non-normalizable states. This is a peculiarity of the operator-state correspondence of the Liouville field theory which differs it from conventional CFT with discret spectra of dimensions but make it similar to some conformal $`\sigma `$-models with non-compact target spaces. In what follows the primary physical states are normalized as $$v_P^{}|v_P=\pi \delta (PP^{})$$ (1.9) The solution of the spherical LFT amounts to constructing all multipoint correlation functions of these fields, $$G_{\alpha _1\mathrm{}\alpha _N}(x_1,\mathrm{},x_N)=V_{\alpha _1}(x_1)\mathrm{}V_{\alpha _N}(x_N)$$ (1.10) In principle these quantities are completely determined by the structure of the operator product expansion (OPE) algebra of the exponential operators, i.e. can be completely restored from the two-point function $$V_a(x)V_a(0)=\frac{D(\alpha )}{(x\overline{x})^{2\mathrm{\Delta }_\alpha }}$$ (1.11) which determines the normalization of the basic operators and the three-point function $$G_{\alpha _1,\alpha _2,\alpha _3}(x_1,x_2,x_3)=\left|x_{12}\right|^{2\gamma _3}\left|x_{23}\right|^{2\gamma _1}\left|x_{31}\right|^{2\gamma _2}C(\alpha _1,\alpha _2,\alpha _3)$$ (1.12) with $`\gamma _1=\mathrm{\Delta }_{\alpha _1}\mathrm{\Delta }_{\alpha _2}\mathrm{\Delta }_{\alpha _3}`$, $`\gamma _2=\mathrm{\Delta }_{\alpha _2}\mathrm{\Delta }_{\alpha _3}\mathrm{\Delta }_{\alpha _1}`$, $`\gamma _3=\mathrm{\Delta }_{\alpha _3}\mathrm{\Delta }_{\alpha _1}\mathrm{\Delta }_{\alpha _2}`$ . Once these quantities are known, the multipoint functions can be in principle reconstructed by the purely “kinematic” calculations relied on the conformal symmetry only. Although these calculations present a separate rather complicated technical problem, conceptually one can say that a CFT (on a sphere) is constructed if these basic objects are found. For LFT these quantities were first obtained by Dorn and Otto in 1992 (see also ). We will present here the derivation of the simplest of them, the two-point function $`D(\alpha )`$, to illustrate a different approach to this problem proposed more recently by J.Techner which seems more effecticient. Close ideas are also developed in the studies of LFT by Gervais and collaborators . We will use similar approach shortly in the discussion of the boundary Liouville problem. Among the exponential operators $`V_\alpha `$ there is a series of fields $`V_{nb/2}`$, $`n=0,1,\mathrm{}`$ which are degenerate w.r.t. the conformal symmetry algebra and therefore satisfy certain linear differential equations. For example, the first non-trivial operator $`V_{b/2}`$ satisfies the following second order equation $$\left(\frac{1}{b^2}^2+T(z)\right)V_{b/2}=0$$ (1.13) and the same with the complex conjugate differentiation in $`\overline{z}`$ and $`\overline{T}(\overline{z})`$ instead of $`T`$. In the classical limit of LFT the existence of this degenerate operator can be traced back to the well known relation between the ordinary second-order linear differential equation and the classical partial-derivative Liouville equation . The next operator $`V_b`$ satisfies two complex conjugate third-order differential equations $$\left(\frac{1}{2b^2}^3+2T(z)+(1+2b^2)T(z)\right)V_b=0$$ (1.14) and so on. It follows from these equations that the operator product expansion of these degenerate operators with any primary field, in the present case with our basic exponential fields $`V_\alpha `$, is of very special form and contains in the r.h.s only finite number of primary fields. For example for the first one there are only two representations $$V_{b/2}V_\alpha =C_+\left[V_{\alpha b/2}\right]+C_{}\left[V_{\alpha +b/2}\right]$$ (1.15) where $`C_\pm `$ are the special structure constants. What is important to remark about these special structure constants is that the general CFT and Coulomb gas experience suggests that they can be considered as “perturbative”, i.e. are obtained as certain Coulomb gas (or “screening”) integrals . For example in our case in the first term of (1.15) there is no need of screening insertion and therefore one can set $`C_+=1`$. The second term requires a first order insertion of the Liouville interaction $`\mu \mathrm{exp}(2b\varphi )d^2x`$ and $`C_{}`$ $`=\mu {\displaystyle d^2xV_\alpha (0)V_{b/2}(1)e^{2b\varphi (x)}V_{Q\alpha b/2}(\mathrm{})}`$ $`=\mu {\displaystyle \frac{\pi \gamma (2b\alpha 1b^2)}{\gamma (b^2)\gamma (2b\alpha )}}`$ (1.16) where as usual $`\gamma (x)=\mathrm{\Gamma }(x)/\mathrm{\Gamma }(1x)`$. It is remarkable that all the special structure constants entering the special truncated OPE’s with the degenerate fields can be obtained in this way. Now let us take the two-point function $`D(\alpha )`$ and consider the auxiliary three-point function $$V_\alpha (x_1)V_{\alpha +b/2}(x_2)V_{b/2}(z)$$ Then, tending $`zx_1`$ we see that in the OPE only the second term survives and in fact our auxiliary function is $`C_{}D(\alpha +b/2)`$. Instead tending $`zx_2`$ we can “lower” the parameter of the second operator down to $`\alpha `$ which results in $`C_+D(\alpha )`$. Equating these two things we arrive at the functional equation for the two-point function $$\frac{D(\alpha +b/2)}{D(\alpha )}=C_{}^1(\alpha )$$ (1.17) This equation can be easily solved in terms of gamma-functions $$D(\alpha )=\left(\pi \mu \gamma (b^2)\right)^{(Q2\alpha )/b}\frac{\gamma (2b\alpha b^2)}{b^2\gamma (22\alpha /b+1/b^2)}$$ (1.18) which coincides precisely with what was obtained for this quantity in the original studies. In fact there are many solutions to the above functional equation. It is relevant for the moment to stop at the remarkable duality property of LFT. Besides the abovementioned series of degenerate operators $`V_{nb/2}`$ there is a “dual” series with $`b`$ replaced by $`1/b`$. This results in another “dual” functional equation for $`D(\alpha )`$ with the shift by $`1/b`$ instead of $`b`$. The solution becomes unique (at least if these two shifts are uncomparable) . Note that these two equations are compatible only if in the dual equation the cosmological constant $`\mu `$ is replaced by the “dual cosmological constant” $`\stackrel{~}{\mu }`$ related to $`\mu `$ as follows $$\pi \stackrel{~}{\mu }\gamma (1/b^2)=\left(\pi \mu \gamma (b^2)\right)^{1/b^2}$$ (1.19) With this definition of $`\stackrel{~}{\mu }`$ the duality property, which turn out to hold exactly in LFT, can be formulated as the symmetry of all observables w.r.t. the substitution $`b1/b`$ and $`\mu \stackrel{~}{\mu }`$. The same way one can readily obtain and solve the functional equations for the three-point function which reads $`C(\alpha _1,\alpha _2,\alpha _3)=\left[\pi \mu \gamma (b^2)b^{22b^2}\right]{}_{}{}^{(Q{\scriptscriptstyle \alpha _i})/b}\times `$ (1.20) $`{\displaystyle \frac{\mathrm{{\rm Y}}_0\mathrm{{\rm Y}}(2\alpha _1)\mathrm{{\rm Y}}(2\alpha _2)\mathrm{{\rm Y}}(2\alpha _3)}{\mathrm{{\rm Y}}(\alpha _1+\alpha _2+\alpha _3Q)\mathrm{{\rm Y}}(\alpha _1+\alpha _2\alpha _3)\mathrm{{\rm Y}}(\alpha _2+\alpha _3\alpha _1)\mathrm{{\rm Y}}(\alpha _3+\alpha _1\alpha _2)}}`$ where a special function $`\mathrm{{\rm Y}}(x)`$ has to be introduced $$\mathrm{log}\mathrm{{\rm Y}}(x)=_0^{\mathrm{}}\frac{dt}{t}\left[\left(Q/2x\right)^2e^{2t}\frac{\mathrm{sinh}^2\left(Q/2x\right)t}{\mathrm{sinh}(bt)\mathrm{sinh}(t/b)}\right]$$ (1.21) This integral representation is convergent only in the strip $`0<\mathrm{Re}`$ $`x<Q`$, otherwise it is an analytic continuation. In fact $`\mathrm{{\rm Y}}(x)`$ is an entire function of $`x`$ with zeroes at $`x=nbm/b`$ and $`x=Q+nb+m/b`$ with $`n`$ and $`m`$ non-negative integers. In the sense mentioned above the explicit results (1.11) and (1.20) constitute the exact construction of the Liouville field theory on a sphere. For example, the four-point function can be explicitly expressed in terms of the three-point function $$G_{\alpha _1,\alpha _2,\alpha _3,\alpha _4}(x_i)=\frac{1}{\pi }\underset{0}{\overset{\mathrm{}}{}}C(\alpha _1,\alpha _2,Q/2+iP)C(\alpha _3,\alpha _4,Q/2iP)|𝐅(\mathrm{\Delta }_{\alpha _i},\mathrm{\Delta },x_i)|^2𝑑P$$ (1.22) where the intergration is over the variety of physical states $`|v_P`$ and $`𝐅(\mathrm{\Delta }_{\alpha _i},\mathrm{\Delta },x_i)`$ is the four-point conformal block, determined completely by the conformal symmetry <sup>3</sup><sup>3</sup>3Strictly speaking, (1.22) is literally correct only if Re$`\alpha _i`$ are sufficiently close to $`Q/2`$. Otherwise additional discrete terms in the r.h.s of (1.22) can appear due to certain poles of $`C`$ braeking through the integration contour, see .. In the four-point case, which we are considering now, the latter can be further reduced to a function of one variable, e.g., $`𝐅(\mathrm{\Delta }_{\alpha _i},\mathrm{\Delta },x_i)`$ $`=`$ (1.23) $`(x_4x_1)^{2\mathrm{\Delta }_1}(x_4x_2)^{\mathrm{\Delta }_1+\mathrm{\Delta }_3\mathrm{\Delta }_2\mathrm{\Delta }_4}(x_4x_3)^{\mathrm{\Delta }_1+\mathrm{\Delta }_2\mathrm{\Delta }_3\mathrm{\Delta }_4}(x_3x_2)^{\mathrm{\Delta }_4\mathrm{\Delta }_1\mathrm{\Delta }_2\mathrm{\Delta }_3}\times `$ $`\left(\begin{array}{cc}\alpha _1& \alpha _3\\ \alpha _2& \alpha _4\end{array}P\eta \right)`$ (1.26) where $$\eta =\frac{(x_1x_2)(x_3x_4)}{(x_1x_4)(x_3x_2)}$$ Parameters $`\alpha _i`$ are related to $`\mathrm{\Delta }_{\alpha _i}`$ as in eq.(1.7) and in the intermediate dimension $`\mathrm{\Delta }=Q^2/4+P^2`$. ## 2 The boundary Liouville problem The basic ideas of 2D conformal field theory with conformally invariant boundary were developed long ago mostly by J.Cardy who also applied them successfully to rational CFT’s, in particular to the minimal series . Here we’ll try to apply these ideas to the Liouville CFT with boundary. A conformally invariant boundary condition in LFT can be introduced through the following boundary interaction $$A_{\mathrm{bound}}=A_{\mathrm{bulk}}+\underset{\mathrm{\Gamma }}{}\left(\frac{QK}{2\pi }\varphi +\mu _Be^{b\varphi }\right)g^{1/4}𝑑\xi $$ (2.1) where the integration in $`\xi `$ is along the boundary while $`K`$ is the curvature of the boundary in the background geometry $`g`$. In what follows we consider only the geometry of a disk which can be represented as a simply connected domain $`\mathrm{\Gamma }`$ in the complex plane with a flat background metric $`g_{ab}=\delta _{ab}`$ inside. The action is simplified as $$A_{\mathrm{bound}}=\underset{\mathrm{\Gamma }}{}\left(\frac{1}{4\pi }(_a\varphi )^2+\mu e^{2b\varphi }\right)d^2x+\underset{\mathrm{\Gamma }}{}\left(\frac{Qk}{2\pi }\varphi +\mu _Be^{b\varphi }\right)𝑑\xi $$ (2.2) where $`k`$ is the curvature of the boundary in the complex plane. Typically the most convenient domain is either a unit circle or the upper half-plane. In the last case the boundary $`\mathrm{\Gamma }`$ is the real axis and one can omit the term linear in $`\varphi `$ in the boundary action (2.2). The price is again a “background charge at infinity”, i.e., the same boundary condition on the field $`\varphi `$ at infinity in the upper half plane $$\varphi (z,\overline{z})=Q\mathrm{log}(z\overline{z})+O(1)\mathrm{at}|z|\mathrm{}$$ (2.3) as in the case of the sphere. It seems natural to call the additional parameter $`\mu _B`$ the boundary cosmological constant. We see that in fact there is a one-parameter family of conformally invariant boundary conditions characterized by different values of the boundary cosmological constant $`\mu _B`$. Contrary to the pure bulk situation where the cosmological constant enters only as a scale parameter, the observables in the boundary case actually depend on the scale invariant ratio $`\mu /\mu _B^2`$. For example, a disk correlation function with the bulk operators $`V_{\alpha _1},V_{\alpha _2}\mathrm{}V_{\alpha _n}`$ and the boundary operators (see below) $`B_{\beta _1},B_{\beta _2}\mathrm{}B_{\beta _m}`$ scales as follows $$𝒢(\alpha _{1,}\mathrm{}\alpha _n,\beta _1,\mathrm{}\beta _m)\mu ^{(Q2_i\alpha _i_j\beta _j)/2b}F\left(\frac{\mu _B^2}{\mu }\right)$$ (2.4) where $`F`$ is some scaling function and we indicate only the dependence on the scale parameters $`\mu `$ and $`\mu _B`$<sup>4</sup><sup>4</sup>4In the presence of boundary operators it is possible to impose different boundary conditions at different pieces of the boundary, each being characterised by its own value of $`\mu _B`$. In this case the scaling function in (2.4) may depend on several invariant ratios, see below.. Our present purpose is to study this dependence. In the boundary case we have to introduce the boundary operators. In LFT the basic boundary primaries are again the exponential in $`\varphi `$ boundary fields $`B_\beta =\mathrm{exp}\beta \varphi `$. Their dimensions are $$\mathrm{\Delta }_\beta =\beta \left(Q\beta \right)$$ (2.5) To avoid any confusions we shall always use parameter $`\alpha `$ for the bulk exponentials and parameter $`\beta `$ in relation with the boundary operators. In general a boundary operator is not characterized completely by its dimension, because the conformal boundary conditions at both sides of the location of the boundary operator may be in general different. One has to specify which boundary condition it joins. Therefore in general we are talking about a juxtaposition boundary operator between, in our case, two boundary conditions with the parameters $`\mu _{B_1}=\mu _1`$ and $`\mu _{B_2}=\mu _2`$ and denote it $`B_\beta ^{\mu _1\mu _2}(x)`$. To define completely the boundary LFT on the disk, i.e, to be able to construct an arbitrary multipoint correlation function including bulk and boundary operators, we have to reveal few more basic objects in addition to the bulk two- and three-point functions (1.18) and (1.20) we already have. 1. First is the bulk one-point function (we imply almost constantly the upper half-plane geometry) $$V_\alpha (x)=\frac{U(\alpha |\mu _B)}{\left|z\overline{z}\right|^{2\mathrm{\Delta }_\alpha }}$$ (2.6) In fig.1a it is drawn however as the one-point function in the unit disk. 2. Second, one needs the boundary two-point function $$B_\beta ^{\mu _1\mu _2}(x)B_\beta ^{\mu _2\mu _1}(0)=\frac{d(\beta |\mu _1,\mu _2)}{\left|x\right|^{2\mathrm{\Delta }_\beta }}$$ (2.7) which in general depends on two boundary cosmological constants $`\mu _1`$and $`\mu _2`$ (see fig.1b). 3. The bulk-boundary structure constant, which determines the fusion of a bulk operator $`V_\alpha `$ with the boundary resulting in the boundary operator $`B_\beta ^{\mu _B\mu _B}`$. This is basically the same as the bulk-boundary two-point function (fig.1c) $$V_\alpha (z)B_\beta (x)=\frac{R(\alpha ,\beta |\mu _B)}{\left|z\overline{z}\right|^{2\mathrm{\Delta }_\alpha \mathrm{\Delta }_\beta }\left|zx\right|^{2\mathrm{\Delta }_\beta }}$$ (2.8) In fact the one-point function (2.6) is a particular case of this quantity with $`\beta =0`$ so that its introduction, however convenient, is redundant. 4. Finally, there is a boundary three-point function $$B_{\beta _1}^{\mu _2\mu _3}(x_1)B_{\beta _2}^{\mu _3\mu _1}(x_2)B_{\beta _3}^{\mu _1\mu _2}(x_3)=\frac{c(\beta _1,\beta _2,\beta _3|\mu _1,\mu _2,\mu _3)}{\left|x_{12}\right|^{\mathrm{\Delta }_1+\mathrm{\Delta }_2\mathrm{\Delta }_3}\left|x_{23}\right|^{\mathrm{\Delta }_2+\mathrm{\Delta }_3\mathrm{\Delta }_1}\left|x_{31}\right|^{\mathrm{\Delta }_3+\mathrm{\Delta }_1\mathrm{\Delta }_2}}$$ (2.9) which in fact depends now on three different boundary parameters, $`\mu _1`$, $`\mu _2`$ and $`\mu _3`$ related to the corresponding sides of the triangle as shown in fig.1d. These three basic boundary structure constants, together with the bulk structure constants, allow in principle to write down an intermediate state expansions for any multipoint function. An instructive example is the bulk two-point function $$G_{\alpha _1\alpha _2}(z_1,z_2)=V_{\alpha _1}(z_1)V_{\alpha _2}(z_2)$$ (2.10) Joining these two operators together with the bulk structure constant we can reduce this quantity to the one-point bulk function and write down the following expansion $$G_{\alpha _1\alpha _2}=\frac{\left|z_2\overline{z}_2\right|^{2\mathrm{\Delta }_12\mathrm{\Delta }_2}}{\left|z_1\overline{z}_2\right|^{4\mathrm{\Delta }_1}}\frac{dP}{\pi }C(\alpha _1,\alpha _2,Q/2+iP)U(Q/2iP)\left(\begin{array}{cc}\alpha _1& \alpha _1\\ \alpha _2& \alpha _2\end{array}\left|P\right|\eta \right)$$ (2.11) where $$\eta =\frac{(z_1z_2)(\overline{z}_1\overline{z}_2)}{(z_1\overline{z}_2)(\overline{z}_1z_2)}$$ (2.12) is the projective invariant of the four points $`z_1`$, $`z_2`$, $`\overline{z}_1`$ and $`\overline{z}_2`$ and $``$ is the same four-point conformal block as enters the expansion of the four-point bulk function, see (1.22,1.23). Notice that while in that case it entered in a sesquilinear combination (1.22), here it appears linearly (J.Cardy ). Expansion (2.11) is suitable appropriate if the bulk operators are close to each other, i.e., $`\eta 0`$. Another representation is suitable the limit $`\eta 1`$ where the points $`z_1`$ and $`z_2`$ approach boundary and the bulk operators can be expanded in the boundary ones. This gives $$G_{\alpha _1\alpha _2}=\frac{\left|z_2\overline{z}_2\right|^{2\mathrm{\Delta }_12\mathrm{\Delta }_2}}{\left|z_1\overline{z}_2\right|^{4\mathrm{\Delta }_1}}\frac{dP}{\pi }R(\alpha _1,Q/2+iP)R(\alpha _2,Q/2iP)\left(\begin{array}{cc}\alpha _1& \alpha _2\\ \alpha _1& \alpha _2\end{array}\left|P\right|1\eta \right)$$ (2.13) Equating these two expressions we see that the basic boundary quantities also must satisfy some bootstrap relations analogous to that in the bulk case. It is interesting to note, that there is another application of this relation. The conformal block itself, although being completely determined by the conformal symmetry, is it fact a complicated function which is not in general known explicitly. On the other hand it is important since it explicitely enters the conformal bootstrap equations . Besides, one might expect that it encodes some information about the structure of the representations of the conformal symmetry. In particular conformal block must satisfy the following cross-relation $$\left(\begin{array}{cc}\alpha _1& \alpha _3\\ \alpha _2& \alpha _4\end{array}\left|P\right|\eta \right)=\frac{dP^{}}{2\pi }K\left(\begin{array}{cc}\alpha _1& \alpha _3\\ \alpha _2& \alpha _4\end{array}P,P^{}\right)\left(\begin{array}{cc}\alpha _1& \alpha _2\\ \alpha _3& \alpha _4\end{array}\left|P^{}\right|1\eta \right)$$ (2.14) with some cross-matrix $`K`$ which determines the monodromy properties of the conformal block. Suppose now we’ve managed to find the basic quantities of the boundary Liouville problem, in particular the one-point function $`U(\alpha )`$ and the bulk-boundary structure constant $`R(\alpha ,\beta )`$. Then the crossing relation becomes a linear equation for the cross-matrix of the symmetric (i.e., $`\alpha _3=\alpha _1`$ and $`\alpha _4=\alpha _2`$) conformal block, from where this matrix can be figured out<sup>5</sup><sup>5</sup>5In a recent paper an explicit expression for this matrix has been proposed on the basis of completely different approach. ### 2.1 Bulk one-point function We start with the calculation of the bulk one point function $`U(\alpha |\mu _B)`$. For this we apply the degenerate operator insertion, like above for the bulk two-point function. Consider the auxiliary bulk two-point function with the additional degenerate bulk field $`V_{b/2}(z)`$ $$G_{\alpha ,b/2}(x,z)=V_\alpha (x)V_{b/2}(z)$$ (2.15) Apply first the OPE at $`zx`$ where the degenerate operator $`V_{b/2}`$ generates only two primary fields so that $$G_{\alpha ,b/2}=C_+(\alpha )U(\alpha b/2)𝒢_+(x,z)+C_{}(\alpha )U(\alpha +b/2)𝒢_{}(x,z)$$ (2.16) where $`C_\pm (\alpha )`$ are the special structure constants as given by the screening integrals and $`𝒢_\pm (x,z)`$ are expressed through the special conformal blocks $`_\pm (x,z)`$ related to these special values of parameters $$𝒢_\pm (x,z)=\frac{\left|x\overline{x}\right|^{2\mathrm{\Delta }_\alpha 2\mathrm{\Delta }_{b/2}}}{\left|z\overline{x}\right|^{4\mathrm{\Delta }_\alpha }}_\pm (\eta )$$ (2.17) where $`\mathrm{\Delta }_{b/2}=1/23b^2/4`$ and $$\eta =\frac{(zx)(\overline{z}\overline{x})}{(z\overline{x})(\overline{z}x)}$$ (2.18) In fact $`V_{b/2}`$ satisfies the second order differential equation. Therefore these special conformal blocks are solution to a second order linear differential equation and can be expressed in terms of the hypergeometric functions $`_+(\eta )=`$ $`\left(\begin{array}{cc}b/2& b/2\\ \alpha & \alpha \end{array}|\alpha {\displaystyle \frac{b}{2}}|\eta \right)=`$ (2.21) $`\eta ^{\alpha b}(1\eta )^{b^2/2}F(2\alpha b12b^2,b^2,2\alpha bb^2,\eta )`$ (2.22) $`_{}(\eta )=`$ $`\left(\begin{array}{cc}b/2& b/2\\ \alpha & \alpha \end{array}|\alpha +{\displaystyle \frac{b}{2}}|\eta \right)=`$ (2.25) $`\eta ^{1+b^2\alpha b}(1\eta )^{b^2/2}F(b^2,12\alpha b,2+b^22\alpha b,\eta )`$ This is a particular case of more general conformal block with a degenerate operator $`V_{b/2}`$ $`\left(\begin{array}{cc}b/2& \alpha _3\\ \alpha _2& \alpha _4\end{array}|\alpha _2{\displaystyle \frac{b}{2}}|\eta \right)`$ $`=\eta ^{b\alpha _2}(1\eta )^{\alpha _3b}F(A,B,C,\eta )`$ (2.28) $`\left(\begin{array}{cc}b/2& \alpha _3\\ \alpha _2& \alpha _4\end{array}|\alpha _2+{\displaystyle \frac{b}{2}}|\eta \right)`$ $`=\eta ^{1+b^2b\alpha _2}(1\eta )^{b\alpha _3}F(AC+1,BC+1,2C,\eta )`$ (2.31) where $`A`$ $`=1+b(\alpha _2+\alpha _3+\alpha _43b/2)`$ $`B`$ $`=b(\alpha _2+\alpha _3\alpha _4b/2)`$ (2.32) $`C`$ $`=b(2\alpha _2b)`$ Now, as both operators approach the boundary, they are expanded in the boundary operators. It turns out that the degenerate bulk operator $`V_{b/2}`$ near the boundary gives rise to only two primary boundary families $`B_0`$ and $`B_b`$. The simplest thing is to find the contribution of $`B_0=I`$. The fusion of $`V_\alpha `$ to the unity boundary operator is described by the quantity $`R(\alpha ,0)=v(\alpha )`$ while the fusion of the field $`V_{b/2}`$ ($`V_{b/2}`$ boundary) is described by a special bulk-boundary structure constant $`R(b/2,Q).`$ It can be computed as the following boundary screening integral with one insertion of the boundary interaction $`\mu _BB_b(x)𝑑x`$ $`R(b/2,Q)`$ $`=2^{2\mathrm{\Delta }_{12}}\mu _B{\displaystyle V_{b/2}(i)B_b(x)B_Q(\mathrm{})𝑑x}`$ (2.33) $`={\displaystyle \frac{2\pi \mu _B\mathrm{\Gamma }(12b^2)}{\mathrm{\Gamma }^2(b^2)}}`$ Comparing this with the behavior predicted by the bulk expansion (2.16) we find the following functional equation for the one-point function $$\frac{2\pi \mu _B}{\mathrm{\Gamma }(b^2)}U(\alpha )=\frac{\mathrm{\Gamma }(b^2+2b\alpha )}{\mathrm{\Gamma }(12b^2+2b\alpha )}U(\alpha b/2)\frac{\pi \mu \mathrm{\Gamma }(1b^2+2b\alpha )}{\gamma (b^2)\mathrm{\Gamma }(2b\alpha )}U(\alpha +b/2)$$ (2.34) (in the last term we used the bulk special structure constant $`C_{}(\alpha )`$ from(1.16)). Equation is solved by the following simple expression $$U(\alpha )=\frac{2}{b}\left(\pi \mu \gamma (b^2)\right)^{(Q2\alpha )/2b}\mathrm{\Gamma }(2b\alpha b^2)\mathrm{\Gamma }\left(\frac{2\alpha }{b}\frac{1}{b^2}1\right)\mathrm{cosh}(2\alpha Q)\pi s$$ (2.35) where the parameter $`s`$ is related to the scale invariant ratio of the cosmological constants $$\mathrm{cosh}^2\pi bs=\frac{\mu _B^2}{\mu }\mathrm{sin}\pi b^2$$ (2.36) Also this expression satisfies the dual functional equation provided the dual bulk cosmological constant $`\stackrel{~}{\mu }`$ is related to $`\mu `$ as before in (1.19) while the parameter $`s`$ is self-dual, i.e. the dual boundary cosmological constant $`\stackrel{~}{\mu }_B`$ is defined as follows $$\mathrm{cosh}^2\frac{\pi s}{b}=\frac{\stackrel{~}{\mu }_B^2}{\stackrel{~}{\mu }}\mathrm{sin}\frac{\pi }{b^2}$$ (2.37) It is remarkable enough that the expression (2.35) automatically satisfies the “reflection relation” for the operator $`V_\alpha `$ $$U(\alpha )=U(Q\alpha )D(\alpha )$$ (2.38) with the bulk Liouville two-point function (1.18). If $`\alpha `$ corresponds to a physical state, i.e., $`\alpha =Q/2+iP`$ with $`P`$ real, expression (2.35) reads $$U(P)=\left(\pi \mu \gamma (b^2)\right)^{iP/b}\mathrm{\Gamma }\left(1+2ibP\right)\mathrm{\Gamma }\left(1+2iP/b\right)\frac{\mathrm{cos}\left(2\pi sP\right)}{iP}$$ (2.39) This quantity is interpreted as the matrix element between a primary physical state $`|v_P`$ from (1.8) and the boundary state $`B_s|`$ created by the boundary action (2.2) $$U(P)=B_s|v_P$$ (2.40) It is natural that this matrix element satisfies the reflection relation $$U(P)=S(P)U(P)$$ (2.41) with the Liouville reflection amplitude $$S_\mathrm{L}(P)=\left(\pi \mu \gamma (b^2)\right)^{2iP/b}\frac{\mathrm{\Gamma }(1+2iP/b)\mathrm{\Gamma }(1+2iPb)}{\mathrm{\Gamma }(12iP/b)\mathrm{\Gamma }(12iPb)}$$ (2.42) Of course the functional relation does not fix the overall constant so that it can be multiplied by any (self-dual) factor $`U_0(b)=U_0(1/b)`$. In (2.35) this factor is chosen in the way that all the residues in the “on-mass-shell” poles at $`2\alpha =Qnb`$, with $`n=1,2,3,\mathrm{}`$ are *equal precisely* to the corresponding perturbative integrals appearing in expansions in $`\mu `$ and $`\mu _B`$, i.e. $`\underset{2\alpha =Qnb}{res}`$ $`{\displaystyle \frac{U(\alpha )}{2^{2\mathrm{\Delta }_\alpha }}}={\displaystyle \underset{k=0}{\overset{[n/2]}{}}}{\displaystyle \frac{(\mu )^k(\mu _B)^{n2k}}{k!(n2k)!}}\times `$ (2.43) $`{\displaystyle \underset{\mathrm{Im}z_i>0}{}}d^2z_1\mathrm{}d^2z_k{\displaystyle \underset{\mathrm{}}{\overset{\mathrm{}}{}}}𝑑x_1\mathrm{}𝑑x_{n2k}V_{Qnb}(i)V_b(z_1)\mathrm{}V_b(z_k)B_b(x_1)\mathrm{}B_b(x_{n2k})_0`$ where $`\mathrm{}_0`$ is the correlation function w.r.t the upper half plane free field with $`\mu =\mu _B=0`$ i.e., with free boundary conditions. Explicitely $$V_{\alpha _1}(z_1)\mathrm{}V_{\alpha _n}(z_n)B_{\beta _1}(x_1)\mathrm{}B_{\beta _m}(x_m)_0=\frac{\underset{i=1}{\overset{n}{}}\left|z_i\overline{z}_i\right|^{2\alpha _i^2}\underset{i,j}{}\left|z_ix_j\right|^{4\alpha _i\beta _j}}{\underset{i>j}{\overset{m}{}}\left|x_ix_j\right|^{2\beta _i\beta _j}\underset{i>j}{\overset{n}{}}\left|(z_iz_j)(z_i\overline{z}_j)\right|^{4\alpha _i\alpha _j}}$$ (2.44) In particular, the pure boundary perturbations in $`\mu _B`$ reproduce the Dyson integrals over a unit circle $$\frac{_i^ndu_i}{_{i>j}^n\left|u_iu_j\right|^{2b^2}}=\left(\frac{2\pi }{\mathrm{\Gamma }(1b^2)}\right)^n\mathrm{\Gamma }(1nb^2)$$ (2.45) Several remarks are in order in connection with the expression (2.35) presented. 1. Semiclassical tests. Consider the limit $`b0`$ while $`P`$ in eq.(2.39) is of order $`b`$ and $`s`$ is of order $`b^1`$. In this limit the minisuperspace approximation is expected to work. Take the geometry of semi-infinite cylinder of circumference $`2\pi `$ and consider the states on the circle. In the minisuperspace approximation one takes into account the dynamics of the zero mode $$\varphi _0=\frac{1}{2\pi }_0^{2\pi }\varphi (x)𝑑x$$ (2.46) neglecting completely all the oscillator modes of field $`\varphi (x)`$. The primary state $`|v_P`$ is represented now by the wave function $$\psi _P(\varphi _0)=\frac{2\left(\pi \mu /b^2\right)^{iP/b}}{\mathrm{\Gamma }\left(2iP/b\right)}K_{2iP/b}\left(2\sqrt{\pi \mu /b^2}e^{b\varphi _0}\right)$$ (2.47) ($`K_v(z)`$ is the modified Bessel function) which satisfies the minisuperspace Shrödinger equation $$\left(\frac{1}{2}\frac{d^2}{d\varphi _0^2}+2\pi \mu e^{2b\varphi _0}2P^2\right)\psi _P(\varphi _0)=0$$ (2.48) and has the following asymptotic at $`\varphi _0\mathrm{}`$ $$\psi _P(\varphi _0)=e^{2iP\varphi _0}+S_\text{L}^{(\mathrm{cl})}(P)e^{2iP\varphi _0}$$ (2.49) where $`S_\text{L}^{(\mathrm{cl})}(P)`$ is the classical limit of the Liouville reflection amplitude (2.42). Also it meets the normalization (1.9) $$_{\mathrm{}}^{\mathrm{}}\psi _P^{}^{}(\varphi _0)\psi _P(\varphi _0)=\pi \delta (PP^{})$$ (2.50) In the approximation under consideration the boundary state $`B_s|`$ wave function is simply related to the boundary Lagrangian $$\mathrm{\Psi }_{B_s}(\varphi _0)=\mathrm{exp}(2\pi \mu _Be^{b\varphi _0})$$ (2.51) The matrix element $`B_s|v_P`$ can be carried out explicitely $$_{\mathrm{}}^{\mathrm{}}\mathrm{\Psi }_{B_s}(\varphi _0)\psi _P(\varphi _0)𝑑\varphi _0=\frac{2}{b}\left(\pi \mu /b^2\right)^{iP/b}\mathrm{\Gamma }\left(2iP/b\right)\mathrm{cos}(2\pi Ps)$$ (2.52) and agrees precisely with the corresponding limit of (2.39). Note, that this calculation is sensitive to the prefactor in eq.(2.35) and confirms our choice $`U_0(b)=1`$, at least in the limit $`b0`$. 2. Boundary length distribution. From the point of view of 2D gravity one can interprete the quantity $$l=_\mathrm{\Gamma }\mathrm{exp}b\varphi $$ (2.53) as the lenth of the boundary. Let $`W_\alpha (l)`$ be the boundary length distribution for the fluctuating disk with the bulk cosmological constant $`\mu `$ and an insertion of the operator $`V_\alpha `$ somewhere inside the disk. Then $$U(\alpha |\mu _B)=_0^{\mathrm{}}\frac{dl}{l}e^{\mu _Bl}W_\alpha (l)$$ (2.54) Form the result (2.35) one finds explicitely $$W_\alpha (l)=\frac{2}{b}\left(\pi \mu \gamma (b^2)\right)^{(Q2\alpha )/2b}\frac{\mathrm{\Gamma }(2\alpha bb^2)}{\mathrm{\Gamma }(1+1/b^22\alpha /b)}K_{(Q2\alpha )/b}(\kappa l)$$ (2.55) where $$\kappa ^2=\frac{\mu }{\mathrm{sin}\pi b^2}$$ (2.56) Compare (2.55) with the minisuperspace distribution (2.47). This result implies that the Shrödinger equation (2.48) in the logarithm of the scale $`\varphi _0=b^1\mathrm{log}(l/2\pi )`$ (which is called sometimes the Wheeler-deWitt equation) does not hold only in semiclassical limit but is in fact exact with a suitable renormalizations of constants (see in this relation the paper where this equation first appeared in the context of the Liouville field theory, and also where similar expressions are obtained in the framework of random surface models). Let us present also the double distribution in the length (2.53) and area $`A`$ defined as $$W_\alpha (l,\mu )=_0^{\mathrm{}}\frac{dA}{A}e^{\mu A}Z_\alpha (A,l)$$ (2.57) It is given by a rather simple expression $$Z_\alpha (A,l)=\frac{1}{b}\frac{\mathrm{\Gamma }(2\alpha bb^2)}{\mathrm{\Gamma }(1+1/b^22\alpha /b)}\left(\frac{l\mathrm{\Gamma }(b^2)}{2A}\right)^{(Q2\alpha )/b}\mathrm{exp}\left(\frac{l^2}{4A\mathrm{sin}\pi b^2}\right)$$ (2.58) 3. “Heavy” $`\alpha `$ semiclassics. Consider again the limit $`b0`$ but with large value of $`\alpha =\eta /b`$ not nessesserily close to $`Q/2`$. Exact expression (2.58) gives in this limit $$Z_{\eta /b}(A,l)=\frac{l\mathrm{\Gamma }(2\eta )e^{(\eta 1/2)C}}{2b^2A\sqrt{2\pi (12\eta }}\mathrm{exp}\left(\frac{1}{b^2}S_\eta (A,l)\right)$$ (2.59) where $`C`$ is the Euler’s constant and $$S_\eta (A,l)=\frac{l^2}{4\pi A}+(12\eta )\left(\mathrm{log}\left(2A/l\right)+\mathrm{log}(12\eta )1\right)$$ (2.60) On the other hand the corresponding classical solution with the area $`A`$ and the boundary lenght $`l`$ reads for the classical field $`\phi =2b\varphi `$ (we imply here the geometry of the disk $`\left|z\right|1`$ with the unit circle as the boundary) $$e^{\phi (z)}=\frac{l^2(a^1a)^2}{\left[2\pi (a^1(z\overline{z})^\eta a(z\overline{z})^{1\eta })\right]^2}$$ (2.61) where $`a`$ is related to the area as follows $$a^2=1\frac{4\pi A}{l^2}(12\eta )$$ (2.62) (we imply here that $`\eta 1/2`$ and $`l^2>4\pi A(12\eta )`$ so that a real classical solution exists). The classical Liouville action for this solution is readily carried out $$S_{\mathrm{cl}}(A,l)=\frac{l^2}{4\pi A}+(12\eta )\left(\mathrm{log}\left(2A/l\right)+\mathrm{log}(12\eta )1\right)$$ (2.63) and coincides with (2.60). In principle it might be possible to check the prefactor in (2.59) performing the one-loop correction. This is not yet done however. 4.“Light” $`\alpha `$ semiclassics. Direct semiclassical calculation of the one-point function (2.35) is possible also in the case $`\alpha =b\sigma `$, with $`b0`$ and $`\sigma `$ fixed. In particular, one can calculate the semiclassical approximation to the function (2.58) by taking the saddle-point contribution to the corresponding functional integral over $`\varphi `$ with fixed area $`A`$ and boundary length $`l`$. In the present case $`\alpha b`$ the exponential insertion does not affect the saddle-point configurations. The nature of these classical solutions depends on the relative value of $`A`$ and $`l^2`$. Here we consider explicitely only the negative-curvature situation $`4\pi A<l^2`$, in which case the classical configurations form an orbit under the action of $`SL(2,R)`$. To be specific we adopt the upper half-plane geometry with the boundary at the real axis. Then generic classical solution $`\varphi _G`$ is obtained from the “standard” solution $$e^{2b\varphi _I}(z,\overline{z})dzd\overline{z}=\frac{\rho ^2dzd\overline{z}}{((z+i)(\overline{z}i)\nu ^2)^2}$$ (2.64) by $`SL(2,R)`$ transformation $$zGz=\frac{az+b}{cz+d},G=\left(\genfrac{}{}{0.0pt}{}{ab}{cd}\right)SL(2,R)$$ (2.65) here $$\rho =\frac{2lA}{l^22\pi A},\nu =\frac{l\sqrt{l^24\pi A}}{l^22\pi A}$$ (2.66) The semiclassical approximation to the expectation value (2.58) is then evaluated as an integral over this manifold of classical configurations, i.e. $$e^{2b\sigma \varphi }_{A,l}=Ne^{\frac{1}{b^2}S_{\mathrm{cl}}}𝑑\mu (G)\left(e^{2b\varphi _G}(z,\overline{z})\right)^\sigma $$ (2.67) where $`S_{\mathrm{cl}}`$ is the classical action (2.63) with $`\eta =0`$, $`d\mu (G)`$ stands for the $`SL(2,R)`$ invariant integration mesure and the factor $`N`$ combines the determinant of zero modes and the contributions of positive modes to the gaussian integral around given classical solution. It is important to note that while $`N`$ can very well depend on $`A/l^2`$, it carries no dependence on $`\sigma `$, i.e. all the $`\sigma `$ dependence of the one-point function in this approximation comes from the integral in (2.67). The integrand in (2.67) can be simplified by a shift of the integration variable $`GG_zG`$, where $`G_z`$ is any fixed ($`z`$-dependent) $`SL(2,R)`$ transformation which maps the point $`z`$ to the point $`i`$ in the upper half-plane; this gives for the integral in (2.67) $$\left(\frac{2i\rho }{z\overline{z}}\right)^{2\sigma }𝑑\mu (G)((a^2+b^2+c^2+d^2+2)\nu ^2(c^2+d^2))^{2\sigma }.$$ (2.68) To evaluate this integral one can introduce the following coordinates on the group manifold of $`SL(2,R)`$, $$a/d=i\frac{2y\overline{y}x(y+\overline{y})}{y\overline{y}};b/d=x;c/d=i\frac{y+\overline{y}2x}{y\overline{y}};d=\frac{1}{\sqrt{\mathrm{}\mathrm{m}y}|yx|}$$ (2.69) where $`x`$ is real and $`y`$ and $`\overline{y}`$ are complex conjugate with $`\mathrm{}\mathrm{m}y>0`$. The invariant mesure takes the form $$d\mu (G)=\frac{2idxd^2y}{(y\overline{y})|yx|^2}$$ (2.70) and the integral in (2.68) can be written as $$\left(\frac{y\overline{y}}{2i}\right)^{2\sigma 1}((y+i)(\overline{y}i)\nu ^2)^{2\sigma }\frac{dxd^2y}{|yx|^2}$$ (2.71) This integral is readily evaluated and one obtains for (2.67) $$\stackrel{~}{N}\frac{1}{2\sigma 1}\left(\frac{i}{z\overline{z}}\frac{2A}{l}\right)^{2\sigma }e^{\frac{1}{b^2}S_{\mathrm{cl}}}$$ (2.72) where the factor $`\stackrel{~}{N}=\pi l^2N/A`$ does not depend on $`\sigma `$; as is mentioned above its determination requires analysis of the fluctuations around the classical configurations which we did not perform. The $`\sigma `$-dependent part in (2.72) agrees with $`b0`$ limit of (2.58). 5. Boundary state. Once the function $`U(P)`$ is constructed, the boundary state $`B_s|`$ can be written down explicitely $$B_s|=_0^{\mathrm{}}\frac{dP}{\pi }U(P)P|$$ (2.73) where the so called Ishibashi states $$P|=v_P|\left(1+\frac{L_1\overline{L}_1}{2P^2+Q^2/2}+\mathrm{}\right)$$ (2.74) are designed in the way to match the conformal invariance of the boundary. Since the combination $`U(P)P|`$ is invariant w.r.t. the reflection $`PP`$ one can extend formally the integral (2.73) to the negative values of $`P`$ and write $$B_s|=_{\mathrm{}}^{\mathrm{}}\frac{dP}{2\pi }e^{2i\pi Ps}u(P)P|$$ (2.75) where $$u(P)=\frac{1}{iP}\left(\pi \mu \gamma (b^2)\right)^{iP/b}\mathrm{\Gamma }\left(1+2ibP\right)\mathrm{\Gamma }\left(1+2iP/b\right)$$ (2.76) It is natural to call $`u(P)`$ the boundary state wave function. Note that the state $`P|`$, although consistent with the conformal invariance, does not correspond to any conformal boundary state, i.e., to a state created by a local conformally invariant boundary condition. However, it can be constructed as a linear combination of boundary states. In view of eq.(2.75) we can write down $$\frac{u(P)}{2\pi }P|=_{\mathrm{}}^{\mathrm{}}e^{2i\pi Ps}B_s|𝑑s$$ (2.77) This equation allows to single out a conformally invariant state containing only one primary state $`v_P|`$ and its descendents. In finite dimensional situation of rational conformal field theories this trick has been friquently used by J.Cardy . ## 3 Boundary two-point function In this section the boundary two-point function $`d(\beta |\mu _1,\mu _2)`$ of (2.7) will be derived. To this purpose we apply basically the same Techner’s tric which has been used in the first section to determine the bulk structure constants. Considering the boundary operators $`B_\beta (x)`$ we find that all the operators $`B_{nb/2}(x)`$ (and also of course the dual fields $`B_{n/2b})`$ with $`n=0,1,\mathrm{}`$are degenerate, i.e., count primary states among their descendents. A complication here is that not all of these “null vectors” nessesserily vanish, contrary to what happens in the bulk situation. For example simplest non-trivial degenerate boundary operator $`B_{b/2}^{s_1s_2}`$ (form now on we shall denote the exponential boundary operators $`B_\beta ^{s_1s_2}=e_{s_1s_2}^{\beta \varphi }`$ instead of $`B_\beta ^{\mu _1\mu _2}`$ having in mind the relation (2.36)) in general does not satisfy the second order differential equation. This means that the null-vector in the corresponding Virasoro representation is some non-vanishing primary field and therefore the second order differential equation has some non-zero terms in the right hand side. This effect can be already seen at the classical level where the upper half-plane boundary Liouville problem is reduced to the classical Liouville equation for the field $`\phi =2b\varphi `$ $$\overline{}\phi =2\pi \mu b^2e^\phi $$ (3.1) in the upper half-plane with the boundary condition $$i(\overline{})\phi =4\pi \mu _Bb^2e^{\phi /2}$$ (3.2) at the real axis. The boundary value of the classical stress tensor can be easily computed $$T_{\mathrm{cl}}(x)=\frac{1}{16}\phi _x^2+\frac{1}{4}\phi _{xx}+\pi b^2(\pi \mu _B^2b^2\mu )e^\phi $$ (3.3) The boundary operator $`B_{b/2}^{s,s}`$ in the classical limit reduces to the boundary value of $`\mathrm{exp}(\phi /4)`$ for which we have $$\left(\frac{d^2}{dx^2}+T_{\mathrm{cl}}\right)e_{s,s}^{\phi /4}=\pi b^2(\pi \mu _B^2b^2\mu )e_{s,s}^{3\phi /4}$$ (3.4) In the right-hand side there is a primary Virasoro operator $`\mathrm{exp}(3\phi /4)_{s,s}`$ which has exactly the same dimension as the null-vector in the corresponding degenerate representation. It is interesting to note that there is a unique relation between the cosmological constants $`\pi \mu _B^2b^2/\mu =1`$ where the r.h.s vanishes and this operator satisfies homogeneous linear differential equation. This effect holds on the quantum level too: if the boundary and bulk cosmological constants are related as $$1=\frac{2\mu _B^2}{\mu }\mathrm{tan}\frac{\pi b^2}{2}$$ (3.5) the second order differential equation holds for the boundary operator $`B_{b/2}^{s,s}`$ (see the remark in the concluding section). Here we are interested in the general situation where this operator is of no use since it does not always satisfy the second order differential equation. It happens however that the next degenerate boundary operator $`B_b^{s,s}`$ do satisfy the third-order differential equation when placed between identical boundary conditions. Therefore it can be used in our calculations instead of $`B_{b/2}`$. As in the bulk, the differential equation predicts the following truncated OPE of this operator with any exponential boundary primary $$B_b^{s,s}B_\beta ^{s,s^{}}=c_+[B_{\beta b}]+c_0[B_\beta ]+c_{}[B_{\beta +b}]$$ (3.6) where $`c_\sigma (\beta )`$ are again the special boundary structure constants, which can be calculated as certain screening integrals. Considering again the auxilary three-point boundary function with a $`B_b`$ insertion one figures out immediately that $$\frac{d(\beta +b|\mu _1,\mu _2)}{d(\beta |\mu _1,\mu _2)}=c_{}^1(\beta )$$ (3.7) The structure constant $`c_{}(\beta )`$ can be evaluated as a combination of scrining integrals. These are of two tipes: a volume screening by the bulk Liouville interaction term $`e^{2b\varphi }`$ $`c_{}^{(\mathrm{v})}`$ $`=\mu {\displaystyle \underset{\mathrm{Im}z>0}{}}d^2ze^{2b\varphi (z)}B_\beta ^{s_1s_2}(0)B_b^{s_2s_2}(1)B_{Q\beta b}^{s_2s_1}(\mathrm{})`$ (3.8) $`=\mu {\displaystyle \underset{\mathrm{Im}z>0}{}}d^2z{\displaystyle \frac{\left|1z\right|^{4b^2}}{\left|z\right|^{4b\beta }\left|z\overline{z}\right|^{2b^2}}}`$ and two boundary screenings $`e^{b\varphi }`$ related to the boundary interaction $`c_{}^{(\mathrm{b})}`$ $`={\displaystyle \underset{i,j}{}}{\displaystyle \frac{\mu _i\mu _j}{2}}{\displaystyle _{C_i}}{\displaystyle _{C_j}}𝑑x_1𝑑x_2e^{b\varphi (x_1)}e^{b\varphi (x_2)}B_\beta ^{s_1s_2}(0)B_b^{s_2s_2}(1)B_{Q\beta b}^{s_2s_1}(\mathrm{})`$ (3.9) $`={\displaystyle \underset{i,j}{}}{\displaystyle \frac{\mu _i\mu _j}{2}}{\displaystyle _{C_i}}{\displaystyle _{C_j}}𝑑x_1𝑑x_2{\displaystyle \frac{\left|(1x_1)(1x_2)\right|^{2b^2}}{\left|x_1x_2\right|^{2b^2}\left|x_1x_2\right|^{2b\beta }}}`$ where the contours $`C_i`$ are chosen as in fig.2 while $`\mu _i`$ are the corresponding values of the boundary cosmological constant, as it is also indicated in in the same figure. Both contributions can be carried out explicitely and we have $`c_{}(\beta )`$ $`=c_{}^{(\mathrm{v})}+c_{}^{(\mathrm{b})}`$ $`=\left[\mu \mathrm{sin}^2(2\pi b\beta )+\mathrm{sin}\pi b^2\left(\mu _1^2+\mu _2^22\mu _1\mu _2\mathrm{cos}(2\pi b\beta )\right)\right]I_0(\beta )`$ (3.10) $`=4\mu \mathrm{sin}\left(\pi b{\displaystyle \frac{2\beta +i(s_1+s_2)}{2}}\right)\mathrm{sin}\left(\pi b{\displaystyle \frac{2\beta i(s_1+s_2)}{2}}\right)\times `$ $`\mathrm{sin}\left(\pi b{\displaystyle \frac{2\beta +i(s_1s_2)}{2}}\right)\mathrm{sin}\left(\pi b{\displaystyle \frac{2\beta i(s_1s_2)}{2}}\right)I_0(\beta )`$ where $$I_0(\beta )=\frac{\gamma (1+b^2)}{\pi }\mathrm{\Gamma }(12b\beta )\mathrm{\Gamma }(2b\beta 1)\mathrm{\Gamma }(1b^22b\beta )\mathrm{\Gamma }(2b\beta b^21)$$ (3.11) and $`s_1`$ and $`s_2`$ are again related to $`\mu _1`$ and $`\mu _2`$ as in eq.(2.36). To construct a solution to the functional equation (3.7) with (3.10) we need more special functions. First one is what is sometimes called the $`q`$-gamma function. Here we denote it $`𝐒(x)`$. It is self dual with repect to $`b1/b`$ and satisfies the following shift relations $`𝐒(x+b)`$ $`=2\mathrm{sin}(\pi bx)𝐒(x)`$ (3.12) $`𝐒(x+1/b)`$ $`=2\mathrm{sin}(\pi x/b)𝐒(x)`$ It has zeroes at $`x=Q+nb+m/b`$ and poles at $`x=nbm/b`$ ( $`m`$ and $`n`$ are non-negative integer numbers). In the strip $`0<\mathrm{Re}`$ $`x<Q`$ the following integral representation is allowed $$\mathrm{log}𝐒(x)=\underset{0}{\overset{\mathrm{}}{}}\frac{dt}{t}\left[\frac{\mathrm{sinh}(Q2x)t}{2\mathrm{sinh}(bt)\mathrm{sinh}(t/b)}\frac{(Q/2x)}{t}\right]$$ (3.13) With this definition it satisfies also the “unitarity” relation $$𝐒(x)𝐒(Qx)=1$$ (3.14) It is also convenient to introduce a self-dual entire function $`𝐆(x)`$ which contains only zeroes at $`x=nbm/b`$, $`m,n=0,1,2,\mathrm{}`$ and enjoes the following shift relations $`𝐆(x+b)`$ $`={\displaystyle \frac{b^{1/2bx}}{\sqrt{2\pi }}}\mathrm{\Gamma }(bx)𝐆(x)`$ (3.15) $`𝐆(x+1/b)`$ $`={\displaystyle \frac{b^{x/b1/2}}{\sqrt{2\pi }}}\mathrm{\Gamma }(x/b)𝐆(x)`$ This function is “elementary” in the sense that both $`\mathrm{{\rm Y}}(x)`$ from eq.(1.21) and $`𝐒(x)`$ are simply expressed in $`𝐆(x)`$ $`\mathrm{{\rm Y}}(x)`$ $`=𝐆(x)𝐆(Qx)`$ $`𝐒(x)`$ $`={\displaystyle \frac{𝐆(Qx)}{𝐆(x)}}`$ (3.16) The integral representation which is valid for all $`0<\mathrm{Re}`$ $`x`$ reads $$\mathrm{log}𝐆(x)=\underset{0}{\overset{\mathrm{}}{}}\frac{dt}{t}\left[\frac{e^{Qt/2}e^{xt}}{(1e^{bt})(1e^{t/b})}+\frac{(Q/2x)^2}{2}e^t+\frac{Q/2x}{t}\right]$$ (3.17) With this function one can easily construct a solution to (3.7). $`d(\beta |s_1,s_2)`$ $`={\displaystyle \frac{\left(\pi \mu \gamma (b^2)b^{22b^2}\right)^{(Q2\beta )/2b}𝐆(Q2\beta )𝐆^1(2\beta Q)}{𝐒(\beta +i(s_1+s_2)/2)𝐒(\beta i(s_1+s_2)/2)𝐒(\beta +i(s_1s_2)/2)𝐒(\beta i(s_1s_2)/2)}}`$ (3.18) $`=\left(\pi \mu \gamma (b^2)b^{22b^2}\right)^{(Q2\beta )/2b}{\displaystyle \frac{𝐆(Q2\beta )}{𝐆(2\beta Q)}}\times `$ $`\mathrm{exp}\left({\displaystyle \underset{\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{dt}{t}}\left[{\displaystyle \frac{\mathrm{sinh}(Q2\beta )t\mathrm{cos}s_1t\mathrm{cos}s_2t}{\mathrm{sinh}bt\mathrm{sinh}t/b}}{\displaystyle \frac{(Q2\beta )}{t}}\right]\right)`$ This solution satisfies also the “dual-shift” relation analogous to (3.7) so that (3.18) is the unique self-dual solution to (3.7). It is of course possible to express the ratio of two $`𝐆`$-functions in terms of $`𝐒`$-function times some ordinary $`\mathrm{\Gamma }`$-functions. We prefere to present $`d(\beta |s_1,s_2)`$ in the form (3.18) to make obvious the “unitarity” relation $$d(\beta |s_1,s_2)d(Q\beta |s_1,s_2)=1$$ (3.19) Note, that an overall independent of $`\beta `$ constant which is allowed by (3.7) and its dual is completely fixed by (3.19). ## 4 Concluding remarks * Eq.(3.5) together with the structure of singularities of the two-point function (3.18) drop a hint at the suggestion that the level 2 degenerate boundary operator $`B_{b/2}^{s_1s_2}(x)`$ has a vanishing null vector if and only if $`s_1s_2=\pm ib`$ or $`s_1+s_2=\pm ib`$ (the second condition is requred by the symmetry of boundary conditions w.r.t. $`ss`$). Let us verify this suggestion on the three-point function with one boundary field $`B_{b/2}^{s,s\pm ib}(x)`$, i.e., consider $$B_\beta ^{s_2s_1}(x_1)B_{b/2}^{s_1s_1\pm ib}(x)B_{\beta +b/2}^{s_1\pm ib,s_2}(x_2)$$ (4.1) Under our suggestion $`B_{b/2}^{s,s\pm ib}(x)`$ satisfies a second order differential equation in $`x`$ and therefore has special operator product expansion with any $`B_\beta `$ $$B_{b/2}^{s,s\pm ib}B_\beta ^{s\pm ib,s^{}}=c_+^{(\pm )}[B_{\beta b/2}^{s,s^{}}]+c_{}^{(\pm )}[B_{\beta +b/2}^{s,s^{}}]$$ (4.2) Then, exactly the same trick which led to eq.(3.7) gives the following shift relation $$\frac{d(\beta |s,s^{})}{d(\beta +b/2|s\pm ib,s^{})}=c_{}^{(\pm )}(\beta )$$ (4.3) As usual we adopt the structure constant with no screenings requred $`c_+^{(\pm )}=1`$. The structure constant $`c_{}^{(\pm )}`$ is given by the integral $$c_{}^{(\pm )}(\beta )=_{\mathrm{}}^{\mathrm{}}e^{b\varphi (x)}B_\beta ^{s_2s_1}(0)B_{b/2}^{s_1s_1\pm ib}(1)B_{Q\beta b/2}^{s_1\pm ib,s_2}(\mathrm{})𝑑x$$ (4.4) This integral is evaluated quite easily (unlike (3.8) or (3.9)) $`c_{}^{(\pm )}(\beta )`$ $`={\displaystyle \frac{\mathrm{\Gamma }(12b\beta )\mathrm{\Gamma }(2b\beta b^21)\mathrm{\Gamma }(1+b^2)}{\pi }}\times `$ (4.5) $`\left[\mu _1\mathrm{sin}\pi (b^22b\beta )+\mu _1^{(\pm )}\mathrm{sin}\pi (2b\beta )\mu _2\mathrm{sin}\pi b^2\right]`$ where $`\mu _1^{(\pm )}`$ is determined by the relation $$\mathrm{cosh}\pi b(s_1\pm ib)=\mu _1^{(\pm )}\sqrt{\mathrm{sin}\pi b^2/\mu }$$ (4.6) After some simple algebra we obtain $`c_{}^{(\pm )}(\beta )`$ $`=2({\displaystyle \frac{\mu }{\pi \gamma (b^2)}})^{1/2}\mathrm{\Gamma }(12b\beta )\mathrm{\Gamma }(2b\beta b^21)\times `$ (4.7) $`\mathrm{sin}\pi b\left(\beta ib(s_1+s_2)/2\right)\mathrm{sin}\pi b^2\left(\beta ib(s_1s_2)/2\right)`$ It is easy to see that the two-point function (3.18) satisfies both relations (4.3). After this support one may suggest further that any degenerate field $`B_{nb/2}^{s,s^{}}`$ has vanishing null-vector (and therefore has truncated operator product expansions if $`ss^{}=ibk`$ or $`s+s^{}=ibk`$ with $`k=n/2,n/2+1,n/2+2,\mathrm{},n/2`$, in close analogy with the fusion rules for degenerate bulk fields. * The boundary two-point function (3.18) is readily applied as the reflection coefficient in the reflection relations for the one-point function of an exponetial boundary operator in the boundary sin-Gordon model. The latter is defined by the following two-dimensional euclidean action $$A_{\mathrm{bsG}}=\underset{\mathrm{\Gamma }}{}d^2x\left[\frac{1}{4\pi }(_a\varphi )^22\mu \mathrm{cos}(2\beta \varphi )\right]2\mu _B\underset{\mathrm{\Gamma }}{}\mathrm{cos}(\beta (\varphi \varphi _0))$$ (4.8) where the bulk part of the action is integrated over a half-plane $`\mathrm{\Gamma }`$ so that the boundary $`\mathrm{\Gamma }`$ is a strainght line. For the moment $`\beta `$ denotes the standard sin-Gordon coupling constant. Apart from it the boundary model depends of three parameters $`\mu `$, $`\mu _B`$ and $`\varphi _0`$ . The dimensional parameters $`\mu `$ and $`\mu _B`$ can be given a precise meaning by specifying the normalisation of the composite fields they couple to. As these operators are combinations of exponentials it suffices to specify a normalisation for the exponential fields in the volume and at the boundary. Here we adopt the conventional normalisation of these fields (see e.g.) corresponding to the short distance asymptotics at $`\left|xy\right|0`$ $`e^{2ia\varphi }(x)e^{2ia\varphi }(y)`$ $`={\displaystyle \frac{1}{\left|xy\right|^{4a^2}}}+\mathrm{}\text{ for the volume fields}`$ $`e^{ia\varphi _B}(x)e^{ia\varphi _B}(y)`$ $`={\displaystyle \frac{1}{\left|xy\right|^{2a^2}}}+\mathrm{}\text{for the boundary ones}`$ (4.9) Here we present only the result for the one-point function of the boundary operator $`𝒢_B(a)=\mathrm{exp}(ia\varphi _B)`$ which reads $$𝒢_B(a)=\left(\frac{\pi \mu }{\gamma (\beta ^2)}\right)^{{\displaystyle \frac{a^2}{2(1\beta ^2)}}}g_0(a)g_S(a)g_A(a)$$ (4.10) where $`\mathrm{log}g_0(a)`$ $`={\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t}}\left[{\displaystyle \frac{2\mathrm{sinh}^2(a\beta t)\left(e^{(1\beta ^2)t/2}\mathrm{cosh}(t/2)\mathrm{cosh}(\beta ^2t/2)1\right)}{\mathrm{sinh}t\mathrm{sinh}(\beta ^2t)\mathrm{sinh}((1\beta ^2)t)}}a^2e^t\right]`$ $`\mathrm{log}g_S(a)`$ $`={\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t}}{\displaystyle \frac{\mathrm{sinh}^2(a\beta t)\left(2\mathrm{cos}(2zt)\mathrm{cos}(2z^{}t)\right)}{\mathrm{sinh}t\mathrm{sinh}(\beta ^2t)\mathrm{sinh}((1\beta ^2)t)}}`$ (4.11) $`\mathrm{log}g_A(a)`$ $`={\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t}}{\displaystyle \frac{\mathrm{sinh}(2a\beta t)\left(\mathrm{cos}(2zt)\mathrm{cos}(2z^{}t)\right)}{\mathrm{sinh}t\mathrm{sinh}(\beta ^2t)\mathrm{cosh}((1\beta ^2)t)}}`$ where the complex number $`z`$ is related to the parameters of the model (4.8) as $$\mathrm{cosh}^2\pi z=\frac{\mu _B^2e^{2i\beta \varphi _0}}{\mu }\mathrm{sin}\pi \beta ^2$$ (4.12) and $`z^{}`$ is the complex conjugate to $`z`$. The details and some applications will be published elsewhere. * In the next publication we will present an explicit expression for the bulk-boundary structure constant (2.8). Equation (2.77) then permits to resolve the system (2.11), (2.13) and (2.14) for the cross-matrix and obtain an explicit expression for a special case of symmeric cross-matrix $`K\left(\begin{array}{cc}\alpha _1& \alpha _1\\ \alpha _2& \alpha _2\end{array}P,P^{}\right)`$. * The random lattice models of 2D quantum gravity allow in many cases to find explicitely the partition functions of minimal models on fluctuating disk with some bulk and boundary operators inserted . Detailed comparison with the Liouville field thery predictions seems quite interesting. The work on that is in progress. Aknowledgements The present study started as a common project with J.Teschner. In the course of the project it turned out that the methods and results were rather complimentary, so that it was decided to present the respective points of view in seperate publications, see ref.. The work of V.F. and Al.Z. was partially supported by EU under contract ERBFMRX CT 960012. The work of A.Z. is supported in part by DOE grant #DE-FG05-90ER40559.
warning/0001/cond-mat0001232.html
ar5iv
text
# Anderson-Mott Transition in a Magnetic Field: Corrections to Scaling ## I Introduction The metal-insulator transition of interacting, disordered electrons at zero temperature ($`T=0`$), or Anderson-Mott transition (AMT), has been studied theoretically near two-dimensions ($`d=2+ϵ`$), as well as in high dimensions ($`d>6`$, and $`d6`$,). In the latter case one finds the complete universality that is characteristic of Landau theories, while in $`d=2+ϵ`$ there is a large number of different universality classes. This is because the AMT in low dimensions is driven by soft modes, and the soft-mode spectrum of electrons at $`T=0`$ is determined by the symmetries of the underlying field theory. Accordingly, systems in external magnetic fields, in the presence of magnetic impurities, in the presence of spin-orbit scattering, etc., all constitute different universality classes. A priori it is not clear whether the high-dimensional or the low-dimensional theories provide a better description of the situation in $`d=3`$. Experimentally, while different universality classes seem to exist, the situation is also far from clear. A serious problem is posed by the fact that the minimum distance from the critical point, and the ease and accuracy with which the distance from the critical point can be controlled, are much inferior to what can be routinely achieved near thermal phase transitions. Consequently, one should expect corrections to scaling to play an important role in the experimentally accessible region of parameter space, and their incorporation into any interpretation of data to be essential. This expectation notwithstanding, however, corrections to scaling near an AMT have been largely ignored by both experimentalists and theorists, with the exception of a suggestion that strong (logarithmic) corrections to scaling may be responsible for the ill-understood observations in phosphorus-doped silicon. Recently, another example of an AMT has been considered near two-dimensions, namely, the quantum phase transition from a ferromagnetic metal to a ferromagnetic insulator. Since both an intrinsic magnetization and an external magnetic field break the rotational symmetry in spin space and give a mass to the transverse spin-triplet particle-hole excitations, one expects this transition to be related to the AMT in an external field. An important question in this context relates to the Goldstone modes or spin waves that are present in a ferromagnet, but not in a system in an external field. Since the Goldstone modes contribute to the soft-mode spectrum, one might expect them to influence the critical behavior. However, it has been shown in Ref. that this is not the case. This makes it likely that the two transitions indeed belong to the same universality class. The effective model for the ferromagnetic MIT that was derived and studied in Ref. , however, is different from the one for the magnetic-field universality class considered before, and the former contains strong corrections to scaling that were absent in the latter. This suggests that the existing description of the magnetic-field universality class is incomplete. In this note we show that the two transitions belong indeed to the same universality class, as has been suggested in Ref. , and that the previously studied model for the magnetic-field AMT missed terms that contribute important corrections to scaling. As a result of these corrections to scaling, a reliable experimental determination of the critical exponents is not possible from the currently available experimental data. ## II Effective action We consider the usual microscopic model for an AMT in an external magnetic field. That is, we assume that the orbital effects of the magnetic field serve only to suppress the soft modes in the particle-particle or Cooper channel, and consider only the Zeeman term explicitly. The latter’s contribution to the action is $`S_\mathrm{B}`$ $`=`$ $`{\displaystyle \frac{1}{2}}g_\mathrm{L}\mu _\mathrm{B}B{\displaystyle }dxT{\displaystyle \underset{n}{}}{\displaystyle \underset{\alpha }{}}[\overline{\psi }_n^\alpha (𝐱)\psi _n^\alpha (𝐱)`$ (2) $`\overline{\psi }_n^\alpha (𝐱)\psi _n^\alpha (𝐱)].`$ Here $`g_\mathrm{L}`$ is the g-factor, $`\mu _\mathrm{B}`$ is the Bohr magneton, $`B`$ is the magnetic field, and $`T`$ is the temperature. $`\overline{\psi }_,`$ and $`\psi _{n,}`$ are fermionic fields for up- and down-spin electrons, with $`n`$ the index of a fermionic Matsubara frequency, $`\omega _n=2\pi T(n+1/2)`$, and $`\alpha `$ a replica index to deal with the quenched disorder. Within the formalism of Ref. , bilinear products of fermion fields are expressed in terms of a classical matrix field $`Q`$, whose elements are related to the fermion fields by means of the isomorphism $$Q_{12}\frac{i}{2}\left(\begin{array}{cccc}\psi _1\overline{\psi }_2& \psi _1\overline{\psi }_2& \psi _1\psi _2& \psi _1\psi _2\\ \psi _1\overline{\psi }_2& \psi _1\overline{\psi }_2& \psi _1\psi _2& \psi _1\psi _2\\ \overline{\psi }_1\overline{\psi }_2& \overline{\psi }_1\overline{\psi }_2& \overline{\psi }_1\psi _2& \overline{\psi }_1\psi _2\\ \overline{\psi }_1\overline{\psi }_2& \overline{\psi }_1\overline{\psi }_2& \overline{\psi }_1\psi _2& \overline{\psi }_1\psi _2\end{array}\right)$$ (3) Here all fields are understood to be taken at position $`𝐱`$, and $`1(n_1,\alpha _1)`$, etc. In terms of these matrix fields, the Zeeman term can be written $$𝒜_\mathrm{B}=ih𝑑𝐱\mathrm{tr}\left[(\tau _3s_3)Q(𝐱)\right],$$ (4) where $`h=g_\mathrm{L}\mu _\mathrm{B}B`$, $`\mathrm{tr}`$ is a trace over all discrete indices of the matrix $`Q`$, and $`\tau _3=s_3=i\sigma _z`$, with $`\sigma _z`$ a Pauli matrix. More generally, it is useful to expand $`Q`$ into a spin-quaternion basis, $$Q_{12}(𝐱)=\underset{r=0,3}{}\underset{i=0,3}{}{}_{r}{}^{i}Q_{12}^{}(𝐱)\left(\tau _rs_i\right),$$ (5) with $`\tau _0=s_0=𝟙`$ the unit $`2\times 2`$ matrix. The partition function can be written in terms of a functional integration over the field $`Q`$ and an auxiliary field $`\stackrel{~}{\mathrm{\Lambda }}`$, $$Z=D[Q]D[\stackrel{~}{\mathrm{\Lambda }}]e^{𝒜[Q,\stackrel{~}{\mathrm{\Lambda }}]},$$ (7) with an action $$𝒜[Q,\stackrel{~}{\mathrm{\Lambda }}]=𝒜_{\mathrm{B}=0}[Q,\stackrel{~}{\mathrm{\Lambda }}]+𝒜_\mathrm{B}[Q].$$ (8) The action in the absence of a magnetic field, $`𝒜_{\mathrm{B}=0}`$, is the same as in Ref. . The further formal development proceeds in exact analogy to Ref. , and we will therefore be very brief. The action allows for a saddle-point solution that can be written in terms of two Green functions, $`𝒢`$ and $``$, which are related to the saddle-point values of the matrix elements $`{}_{0}{}^{0}Q`$ and $`{}_{3}{}^{3}Q`$, respectively. They obey the equations $`\left(i\omega _n\xi _𝐤\mathrm{\Sigma }_n\right)𝒢_n(𝐤)+\mathrm{\Delta }_n_n(𝐤)=1,`$ (10) $`\left(i\omega _n\xi _𝐤\mathrm{\Sigma }_n\right)_n(𝐤)+\mathrm{\Delta }_n𝒢_n(𝐤)=0.`$ (11) Here $`\xi _𝐤=𝐤^2/2mϵ_\mathrm{F}`$ with $`m`$ the electron mass and $`ϵ_\mathrm{F}`$ the Fermi energy. The two self-energies $`\mathrm{\Sigma }`$ and $`\mathrm{\Delta }`$ are given by $`\mathrm{\Sigma }_n`$ $`=`$ $`{\displaystyle \frac{1}{\pi N_\mathrm{F}\tau }}{\displaystyle \frac{1}{V}}{\displaystyle \underset{𝐤}{}}𝒢_n(𝐤),`$ (13) $`\mathrm{\Delta }_n`$ $`=`$ $`\mathrm{\Delta }{\displaystyle \frac{1}{\pi N_\mathrm{F}\tau }}{\displaystyle \frac{1}{V}}{\displaystyle \underset{𝐤}{}}_n(𝐤),`$ (14) with $$\mathrm{\Delta }=h+2\mathrm{\Gamma }^{(t)}T\underset{n}{}\frac{1}{V}\underset{𝐤}{}_n(𝐤).$$ (15) Here $`\mathrm{\Gamma }^{(t)}`$ is the spin-triplet interaction amplitude defined in Ref. , $`\tau `$ is the transport relaxation time due to the quenched disorder, and $`N_\mathrm{F}`$ is the density of states at the Fermi level. The appearance of $`h`$ in Eq. (15) is the only difference between the saddle-point equations here and those in Ref. , and it results from the Zeeman term in the action. For later reference we note that the field configuration obtained by putting $`\mathrm{\Delta }=0`$ is not a saddle point, in constrast to the situation in the absence of an external field. The discussion of the saddle-point solution proceeds as in Ref. . In particular, $`\mathrm{\Delta }`$ obeys the equation $$\mathrm{\Delta }=hT\underset{n}{}\frac{1}{V}\underset{𝐤}{}\frac{2\mathrm{\Gamma }^{(t)}\mathrm{\Delta }}{(i\omega _n\xi _𝐤+i\mathrm{sgn}\omega _n/2\tau )^2\mathrm{\Delta }^2}.$$ (17) For $`h=0`$ there is a nonzero physical solution only if the Stoner criterion $`N_\mathrm{F}\mathrm{\Gamma }^{(t)}>1`$ is fulfilled. In paramagnetic systems, $`N_\mathrm{F}\mathrm{\Gamma }^{(t)}`$ is smaller than the critical value, and $`\mathrm{\Delta }=0`$ in the absence of a magnetic field. However, for nonzero $`h`$ we have $$\mathrm{\Delta }=h\left[1+N_\mathrm{F}\mathrm{\Gamma }^{(t)}\right]+O(b^3).$$ (18) Since $`\mathrm{\Delta }`$ is proportional to the magnetization, this reflects the fact that a magnetic field polarizes the spins. We see that, at the saddle-point level, the only difference between the current situation and the one considered in Ref. is the origin of the nonzero value of $`\mathrm{\Delta }`$. More generally, as far as an effective theory that concentrates on soft-mode effects is concerned, the only difference lies in the existence of Goldstone modes in a magnetic system. Since the latter turn out to be irrelevant for describing the metal-insulator transition, the procedure of expanding about the saddle-point, separating soft and massive modes, and deriving an effective theory for the metal-insulator transition is therefore the same in the magnetic field and the ferromagnetic cases, respectively, and we can simply take over the result for the latter. The final effective action thus reads $`𝒜`$ $`=`$ $`{\displaystyle \frac{1}{2G}}{\displaystyle 𝑑𝐱\mathrm{tr}[\stackrel{~}{Q}(𝐱)]^2}+2H{\displaystyle 𝑑𝐱\mathrm{tr}[\mathrm{\Omega }\stackrel{~}{Q}(𝐱)]}`$ (22) $`{\displaystyle \frac{1}{2G_3}}{\displaystyle 𝑑𝐱\mathrm{tr}\left((\tau _3s_3)[\stackrel{~}{Q}(𝐱)]^2\right)}`$ $`+2H_3{\displaystyle 𝑑𝐱\mathrm{tr}\left[(\tau _3s_3)\mathrm{\Omega }\stackrel{~}{Q}(𝐱)\right]}+𝒜_{\mathrm{int}}[\stackrel{~}{Q}].`$ Here $$\mathrm{\Omega }_{12}=(\tau _0s_0)\delta _{12}\mathrm{\hspace{0.17em}2}\pi Tn,$$ (23) is a bosonic Matsubara frequency matrix. The bare values of the coupling constants $`G`$, $`G_3`$, $`H`$, and $`H_3`$ are $`1/G`$ $`=`$ $`{\displaystyle \frac{\pi }{4}}m(\sigma _0^++\sigma _0^{}),`$ (24) $`1/G_3`$ $`=`$ $`{\displaystyle \frac{\pi }{4}}m(\sigma _0^+\sigma _0^{}),`$ (25) $`H`$ $`=`$ $`{\displaystyle \frac{\pi }{16}}(N_\mathrm{F}^++N_\mathrm{F}^{}),`$ (26) $`H_3`$ $`=`$ $`{\displaystyle \frac{\pi }{16}}(N_\mathrm{F}^+N_\mathrm{F}^{}),`$ (27) where $`\sigma _0^\pm `$ and $`N_\mathrm{F}^\pm `$ are the Boltzmann conductivity and the bare density of states at the Fermi level, respectively, of systems whose Fermi energy has been shifted by $`\pm \mathrm{\Delta }`$ from its level in the absence of a magnetic field. Notice that $`1/G_3`$ and $`H_3`$ vanish for a vanishing magnetic field, and for small fields are linear in the field, see Eqs. (25,27,18). The interaction piece of the action consists of three terms, $$𝒜_{\mathrm{int}}[Q]=𝒜_{\mathrm{int}}^{(s)}[Q]+𝒜_{\mathrm{int}}^{(t)}[Q]+𝒜_{\mathrm{int}}^{(3)}[Q],$$ (28) $`𝒜_{\mathrm{int}}^{(s)}[Q]`$ $`=`$ $`{\displaystyle \frac{\pi T}{4}}K_s{\displaystyle 𝑑𝐱\underset{1234}{}\delta _{\alpha _1\alpha _2}\delta _{\alpha _1\alpha _3}\delta _{12,43}}`$ (31) $`\times {\displaystyle \underset{r}{}}()^r\mathrm{tr}\left[(\tau _rs_0)Q_{12}(𝐱)\right]`$ $`\times \mathrm{tr}\left[(\tau _rs_0)Q_{34}(𝐱)\right],`$ $`𝒜_{\mathrm{int}}^{(t)}[Q]`$ $`=`$ $`{\displaystyle \frac{\pi T}{4}}K_t{\displaystyle 𝑑𝐱\underset{1234}{}\delta _{\alpha _1\alpha _2}\delta _{\alpha _1\alpha _3}\delta _{12,43}}`$ (34) $`\times {\displaystyle \underset{r}{}}()^r\mathrm{tr}\left[(\tau _rs_3)Q_{12}(𝐱)\right]`$ $`\times \mathrm{tr}\left[(\tau _rs_3)Q_{34}(𝐱)\right],`$ $`𝒜_{\mathrm{int}}^{(3)}[Q]`$ $`=`$ $`4\pi TK_3{\displaystyle 𝑑𝐱\underset{1234}{}\delta _{\alpha _1\alpha _2}\delta _{\alpha _1\alpha _3}\delta _{12,43}}`$ (36) $`{\displaystyle \underset{rs}{}}{\displaystyle \underset{ij}{}}m_{rs,ij}{}_{r}{}^{i}Q_{12}^{}(𝐱){}_{s}{}^{j}Q_{34}^{}(𝐱),`$ where $$m_{rs,ij}=\frac{1}{4}\mathrm{tr}(\tau _3\tau _r\tau _s^{})\mathrm{tr}(s_3s_is_j^{}),$$ (37) and $`K_t=2\pi \mathrm{\Gamma }^{(t)}`$ and $`K_s`$ are the usual spin-triplet and spin-singlet interaction amplitudes that are also present in the absence of a magnetic field. $`K_3`$, like $`1/G_3`$ and $`H_3`$, is magnetic field dependent, and vanishes linearly with the field for small fields. Since it is absent from the bare action, but generated by the renormalization group at one-loop order, it is also proportional to the disorder. Finally, $$\stackrel{~}{Q}_{12}=\widehat{Q}_{12}\delta _{12}(\tau _0s_0)\omega _{n_1},$$ (38) and $`\widehat{Q}`$ is subject to the constraints $$\widehat{Q}^2(𝐱)1,\widehat{Q}^{}=\widehat{Q},\mathrm{tr}\widehat{Q}(𝐱)0.$$ (39) ## III Metal-insulator transition We recognize Eqs. (II) as the generalized nonlinear $`\sigma `$ model proposed by Finkel’stein for the magnetic field universality class, except that the latter was missing the terms proportional to $`1/G_3`$, $`H_3`$, and $`K_3`$. As was shown in Ref. , these terms do not change the asymptotic critical behavior. Choosing the correlation length exponent $`\nu `$, the critical exponent for the density of states $`\beta `$, and the dynamical critical exponent $`z`$ as the three independent exponents, we thus have, to lowest order in an expansion in $`ϵ=d2`$, $`\nu `$ $`=`$ $`1/ϵ+O(1),`$ (41) $`\beta `$ $`=`$ $`1/2ϵ(1\mathrm{ln}2),`$ (42) $`z`$ $`=`$ $`d,`$ (43) and the critical exponent for the conductivity, $`s=\nu (d2)`$, is $$s=1+O(ϵ).$$ (44) While the additional coupling constants $`1/G_3`$, $`H_3`$, and $`K_3`$ do not contribute to the leading critical behavior, they lead to corrections to scaling. As was shown in Ref. , the least irrelevant operator related to these coupling constants has a scale dimension $$\lambda _3=\frac{3\mathrm{ln}22}{1\mathrm{ln}2}ϵ+O(ϵ^2).$$ (45) Consequently, in the vicinity of the metal-insulator transition any observable $`\omega `$ obeys the homogeneity law $$\omega (t,T,u,\mathrm{})=b^{x_\omega }\omega (tb^{1/\nu },Tb^z,ub^{\lambda _3},\mathrm{}),$$ (46) where $`t`$ is the dimensionless distance from the critical point. Apart from (or instead of) the temperature $`T`$, other generalized external fields may appear as arguments of $`\omega `$, e.g., a frequency, or an electric field, depending on the nature of the observable under consideration. $`u`$ denotes the least irrelevant variable (which is a linear combination of $`1/G_3`$, $`H_3`$, and $`K_3`$), $`x_\omega `$ is the scale dimension of $`\omega `$, $`b`$ is an arbitrary scale factor, and the ellipses denote the dependence of $`\omega `$ on variables that are more irrelevant than $`u`$. These corrections to scaling were absent from the original model of Ref. . ## IV Discussion To put our results in context, we give a brief discussion of corrections to scaling in general, the additional terms in Eq. (22) in particular, and why they were missed in previous treatments. Asymptotically close to a critical point, observables obey the homogeneity law, Eq. (46), and of the various coupling constants only those with positive scale dimensions need to be kept. Apart from the dimensionless distance from the critical point, whose scale dimension is by definition the inverse correlation length exponent $`1/\nu >0`$, the only coupling constants with positive scale dimensions are generalized external fields, like the temperature in our example. Technically, the homogeneity law appears as the solution of a renormalization group equation, which takes the form of a Callan-Symanzik equation or a similar partial differential equation. If one is not asymptotically close to the transition, corrections to the asymptotic scaling behavior appear from two sources: (1) The solution of the renormalization group equation becomes more complicated than a generalized homogeneous function, and needs to be expanded in a power series. (2) The “irrelevant operators”, i.e. coupling constants whose scale dimensions are negative, can no longer be ignored. Keeping them, and ordering them with respect to their irrelevancy, leads to the Wegner expansion. In our case, the leading term in the Wegner expansion, which we have kept in Eq. (47), is more important than the corrections to scaling of type (1). In fact, our value $`\nu \lambda _30.26`$ is rather small for a leading Wegner exponent, which typically is on the order of 0.5. Since these large corrections to scaling result from the coupling constants $`1/G_3`$, $`H_3`$, and $`K_3`$ in Eqs. (II), it is important to keep these terms. This brings us to our next discussion topic, viz. the origin of these additional terms in the action. As we have mentioned in Sec. II above, a field configuration that has $`\mathrm{\Delta }=0`$, and thus describes non-spin polarized electrons, is not a saddle-point solution of the field theory, Eqs. (5), underlying the effective action. Reference used the implicit assumption that the only effect of the magnetic field, as far as the metal-insulator transition is concerned, is to give a mass to certain modes that are massless in the absence of an external field. If this were true, then one could use a field configuration with $`\mathrm{\Delta }=0`$ as the starting point for deriving an effective action. The result of this procedure is Eqs. (II) with $`1/G_3=H_3=K_3=0`$. By the nature of the renormalization group flow equations, zero bare values for these coupling constants imply that they are not generated under renormalization either. The effective theory that was originally proposed for this problem is therefore incomplete. As we have seen, this omission has an influence, not on the asymptotic scaling properties, but on the corrections to scaling, within the resulting effective field theory. Since the additional terms discussed above describe a physical effect, namely the spin polarization due to the magnetic field that was neglected in the original treatment, we conclude that the leading corrections to scaling discussed in this paper are an immediate, if a priori not entirely obvious, consequence of this spin polarization. We also note that there is no obvious reason underlying our result that the additional terms do not change the asypmtotic critical behavior; this appears to be accidental. In particular, it needs to be stressed that this conclusion holds only in an $`ϵ`$-expansion near two-dimensions, and in $`d=3`$ the new terms might well influence the asymptotic scaling behavior. We finally discuss the experimental implications of our results. For definiteness, let us consider the conductivity $`\sigma `$, whose scale dimension is $`x_\sigma =s/\nu `$. A common way to analyze experiments is to extrapolate the data to $`T=0`$ and consider $`\sigma (t,T=0)`$. From Eq. (46), we have $$\sigma (t,T=0)=\sigma _0t^s[1+\mathrm{const}.\times t^{\nu \lambda _3}].$$ (47) Here $`\sigma _0`$ is a microscopic conductivity scale on the order of the Boltzmann conductivity, and the constant is nonuniversal. Assuming that the constant is of $`O(1)`$, in order to achieve, e.g., a 10% accuracy in a determination of the value of $`s`$ from a log-log plot, the $`t`$-range needs to be restricted to $`t<(0.1/\nu |\lambda _3|)^{1/\nu |\lambda _3|}`$. Extrapolating the results of our one-loop approximation to $`d=3`$, we have $`\nu |\lambda _3|0.26`$, or $`t<0.026`$. If the constant is larger, as is the case at many critical points, the asymptotic critical region is even smaller. For instance, for $`\mathrm{const}.=1/|\nu \lambda _3|4`$, we need $`t10^4`$ to measure $`s`$ with a 10% accuracy. Such small values of $`t`$ are not currently achievable for Anderson-Mott transitions. The technique that yields the best control over $`t`$, viz. stress tuning, allows to probe the region $`10^3t10^2`$, while other methods are restricted to $`t10^2`$ or larger. We conclude that it is currently not possible to determine the asymptotic critical exponents for the magnetic-field universality class of the Anderson-Mott transition with any meaningful accuracy. ###### Acknowledgements. Parts of this work were performed at the Institute for Theoretical Physics at UC Santa Barbara, and at the Aspen Center for Physics. This work was supported by the NSF under grant Nos. DMR-98-70597, DMR-99-75259, and PHY94-07194.
warning/0001/cond-mat0001111.html
ar5iv
text
# Generic Dynamic Scaling in Kinetic Roughening ## Abstract We study the dynamic scaling hypothesis in invariant surface growth. We show that the existence of power-law scaling of the correlation functions (scale invariance) does not determine a unique dynamic scaling form of the correlation functions, which leads to the different anomalous forms of scaling recently observed in growth models. We derive all the existing forms of anomalous dynamic scaling from a new generic scaling ansatz. The different scaling forms are subclasses of this generic scaling ansatz associated with bounds on the roughness exponent values. The existence of a new class of anomalous dynamic scaling is predicted and compared with simulations. The theory of kinetic roughening deals with the fate of surfaces growing in nonequilibrium conditions . In a typical situation an initially flat surface grows and roughens continuously as it is driven by some external noise. The noise term can be of thermal origin (like for instance fluctuations in the flux of particles in a deposition process), or a quenched disorder (like in the motion of driven interfaces through porous media). A rough surface may be characterized by the fluctuations of the height around its mean value. So, a basic quantity to look at is the global interface width, $`W(L,t)=[\overline{h(x,t)\overline{h}]^2}^{1/2}`$, where the overbar denotes average over all $`x`$ in a system of size $`L`$ and brackets denote average over different realizations. Rough surfaces then correspond to situations in which the stationary width $`W(L,t\mathrm{})`$ grows with the system size. Alternatively, one may calculate other quantities related to correlations over a distance $`l`$ as the height-height correlation function, $`G(l,t)=\overline{[h(x+l,t)h(x,t)]^2}`$, or the local width, $`w(l,t)=[h(x,t)h_l]^2_l^{1/2}`$, where $`\mathrm{}_l`$ denotes an average over $`x`$ in windows of size $`l`$. In absence of any characteristic length in the problem growth processes are expected to show power-law behaviour of the correlation functions in space and time and the Family-Vicsek dynamic scaling ansataz $$W(L,t)=t^{\alpha /z}f(L/\xi (t)),$$ (1) ought to hold. The scaling function $`f(u)`$ behaves as $$f(u)\{\begin{array}{ccc}u^\alpha \hfill & \mathrm{if}& u1\hfill \\ \mathrm{const}.\hfill & \mathrm{if}& u1\hfill \end{array},$$ (2) where $`\alpha `$ is the roughness exponent and characterizes the stationary regime, in which the horizontal correlation length $`\xi (t)t^{1/z}`$ ($`z`$ is the so called dynamic exponent) has reached a value larger than the system size $`L`$. The ratio $`\beta =\alpha /z`$ is called growth exponent and characterizes the short time behavior of the surface. As occurs in equilibrium critical phenomena, the corresponding critical exponents do not depend on microscopic details of the system under investigation. This has made possible to divide growth processes into universality classes according to the values of these characteristic exponents . A most intringuing feature of some growth models is that the above standard scaling of the global width differs substancially from the scaling behaviour of the local interface fluctuations (measured either by the local width or the height-height correlation). More precisely, in some growth models the local width (and the height-height correlation) scales as in Eq.(1), i.e. $`w(l,t)=t^\beta f_A(l/\xi (t))`$, but with the anomalous scaling function $$f_A(u)\{\begin{array}{ccc}u^{\alpha _{loc}}\hfill & \mathrm{if}& u1\hfill \\ \mathrm{const}\hfill & \mathrm{if}& u1\hfill \end{array},$$ (3) where the new independent exponent $`\alpha _{loc}`$ is called the local roughness exponent. This is what has been called anomalous roughening in the literature, and has been found to occur in many growth models as well as experiments . Moreover ,it has recently been shown that anomalous roughening can take two different forms. On the one hand, there are super-rough processes, i.e. $`\alpha >1`$, for which always $`\alpha _{loc}=1`$. On the other hand, there are intrinsically anomalous roughened surfaces, for which $`\alpha _{loc}<1`$ and $`\alpha `$ can actually be any $`\alpha >\alpha _{loc}`$. Anomalous scaling implies that one more independent exponent, $`\alpha _{loc}`$, may be needed in order to asses the universality class of the particular system under study. In other words, some growth models may have exactly the same $`\alpha `$ and $`z`$ values seemingly indicating that they belong to the same universality class. However, they may have different values of $`\alpha _{loc}`$ showing that they actually belong to distict classes of growth. As for the experiments, only the local roughness exponent is measurable by direct methods, since the system size remains normally fixed. Fracture experiments in systems of varying sizes have succeded in measuring both the local and global roughness exponents in good agreement with the scaling picture described above. In this Letter we introduce a new anomalous dynamics in kinetic roughening. We show that, by adopting more general forms of the scaling functions involved, a generic theory of dynamic scaling can be constructed. Our theory incorporates all the different forms that dynamic scaling can take, namely Family-Vicsek, super-rough and intrinsic, as subclasses and predicts the existence of a new class of growth models with novel anomalous scaling properties. Simulations of the Sneppen model (rule A) of self-organized depinning (and other related models) are presented as examples of the new dynamics. Firstly, let us consider the Fourier transform of the height of the surface in a system of size $`L`$, which is given by $`\widehat{h}(k,t)=L^{1/2}_x[h(x,t)\overline{h}(t)]\mathrm{exp}(ikx)`$, where the spatial average of the height has been substracted. The scaling behaviour of the surface can now be investigated by calculating the structure factor or power spectrum $$S(k,t)=\widehat{h}(k,t)\widehat{h}(k,t),$$ (4) which is related to the height-height correlation function $`G(l,t)`$ defined above by $`G(l,t)`$ $`=`$ $`{\displaystyle \frac{4}{L}}{\displaystyle \underset{2\pi /Lk\pi /a}{}}[1\mathrm{cos}(kl)]S(k,t)`$ (5) $``$ $`{\displaystyle _{2\pi /L}^{\pi /a}}{\displaystyle \frac{dk}{2\pi }}[1\mathrm{cos}(kl)]S(k,t),`$ (6) where $`a`$ is the lattice spacing and $`L`$ is the system size. In order to explore the most general form that kinetic roughening can take, we study the scaling behaviour of surfaces satisfying what we will call a generic dynamic scaling form of the correlation functions. We will consider that a growing surface satisfies a generic dynamic scaling when there exists a correlation lenght $`\xi (t)`$, i.e. the distance over which correlations have propagated up to time $`t`$, and $`\xi (t)t^{1/z}`$, being $`z`$ the dynamic exponent. If no characteristic scale exists but $`\xi `$ and the system size $`L`$, then power-law behaviour in space and time is expected and the growth saturates when $`\xi L`$ and the correlations (and from Eq.(6) also the structure factor) become time-independent. The global roughness exponent $`\alpha `$ can now be calculated in this regime from $`G(l=L,tL^z)L^{2\alpha }`$ (or $`W(L,tL^z)L^\alpha `$). In general, as we will see below, the scaling function that enters the dynamic scaling of the local width (or the height-height correlation) takes different forms depending on further restrictions and/or bounds for the roughness exponent values. These kind of restrictions are very often assumed and not valid for every growth model. For instance, only if the surface were self-affine saturation of the correlation function $`G(l,t)`$ would also occur for intermediate scales $`l`$ at times $`tl^z`$ and with the very same roughness exponent. However, the latter does not hold when anomalous roughening takes place as can be seen from the scaling of the local width in Eq.(3). Our aim here is to investigate all the possible forms that the scaling functions can exhibit when solely the existence of generic scaling is assumed. So, if the roughening process under consideration shows generic dynamic scaling (in the sense above explained), and no further assumptions (like for instance surface self-affinity or implicit bounds for the exponents values) are imposed, then we propose that the structure factor is given by $$S(k,t)=k^{(2\alpha +1)}s(kt^{1/z}),$$ (7) where the scaling function has the general form $$s(u)\{\begin{array}{ccc}u^{2(\alpha \alpha _s})\hfill & \mathrm{if}& u1\hfill \\ u^{2\alpha +1}\hfill & \mathrm{if}& u1\hfill \end{array},$$ (8) and the exponent $`\alpha _s`$ is what we will call the spectral roughness exponent. This scaling ansatz is a natural generalization of the scaling proposed for the structure factor in Ref. for anomalous scaling. In the case of the global width, one can make use of $$W^2(L,t)=\frac{1}{L}\underset{k}{}S(k,t)=\frac{dk}{2\pi }S(k,t),$$ (9) to prove easily that the global width scales as in Eqs.(1) and (2), independently of the value of the exponents $`\alpha `$ and $`\alpha _s`$. However, the scaling of the local width is much more involved. The existence of a generic scaling behaviour like (8) for the structure factor always leads to a dynamic scaling behaviour, $$w(l,t)\sqrt{G(l,t)}=t^\beta g(l/\xi )$$ (10) of the height-height correlation (and local width), but the corresponding scaling function $`g(u)`$ is not unique. When substituting Eqs.(8) and (7) into (6), one can see that the various limits involved ($`a0`$, $`\xi (t)/L\mathrm{}`$ and $`L\mathrm{}`$) do not conmute . This results in a different scaling behaviour of $`g(u)`$ depending on the value of the exponent $`\alpha _s`$. Let us now summarize how all scaling behaviours reported in the literature are obained from the generic dynamic scaling ansatz (8). We shall also show how a new roughening dynamics naturally appears in this scaling theory. Two major cases can be distinguised, namely $`\alpha _s<1`$ and $`\alpha _s>1`$. On the one hand, for $`\alpha _s<1`$ the integral in Eq.(6) has already been computed and one gets $$g_{\alpha _s<1}(u)\{\begin{array}{ccc}u^{\alpha _s}\hfill & \mathrm{if}& u1\hfill \\ \mathrm{const}\hfill & \mathrm{if}& u1\hfill \end{array}.$$ (11) So, the corresponding scaling function is $`g_{\alpha _s<1}f_A`$ and $`\alpha _s=\alpha _{loc}`$, i.e. the intrinsic anomalous scaling function in Eq.(3). Moreover, in this case the interface would satisfy a Family-Vicsek scaling (for the local as well as the global width) only if $`\alpha =\alpha _s`$ were satisfied for the particular growth model under study. Thus, the standard Family-Vicsek scaling turns out to be one of the possible scaling forms compatible with generic scaling invariant growth, but not the only one. On the other hand, a new anomalous dynamics shows up for growth models in which $`\alpha _s>1`$. In this case, one finds that, in the thermodynamic limit $`L\mathrm{}`$, the integral Eq.(6) has a divergence coming from the lower integration limit. To avoid the divergence one has to compute the integral keeping $`L`$ fixed. We then obtain the scaling function $$g_{\alpha _s>1}(u)\{\begin{array}{ccc}u\hfill & \mathrm{if}& u1\hfill \\ \mathrm{const}\hfill & \mathrm{if}& u1\hfill \end{array}.$$ (12) So that in this case one always gets $`\alpha _{loc}=1`$ for any $`\alpha _s>1`$. Thus, for growth models in which $`\alpha =\alpha _s`$ one recovers the super-rough scaling behaviour . However, it is worth noting that neither the spectral exponent $`\alpha _s`$ or the global exponent $`\alpha `$ are fixed by the scaling in Eqs(8) and (12) and, in principle, they could be different. Therefore, growth models in which $`\alpha _s>1`$ but $`\alpha \alpha _s`$ could also be possible and represent a new type of dynamics with anomalous scaling. The main feature of this new type of anomalous roughening is that it can be detected only by determining the scaling of the structure factor. Whenever such a scaling takes place in the problem under investigation the new exponent $`\alpha _s`$ will only show up when analyzing the scaling behaviour of $`S(k,t)`$ and will not be detectable in either $`W(L,t)`$, $`w(l,t)`$ or $`G(l,t)`$. In fact, as we have shown, the stationary regime of a surface exhibiting this kind of anomalous scaling will be characterized by $`W(L)L^\alpha `$ and $`w(l,L)\sqrt{G(l,L)}lL^{\alpha 1}`$, however, the structure factor scales as $`S(k,L)k^{(2\alpha _s+1)}L^{2(\alpha \alpha _s)}`$ where the spectral roughness exponent $`\alpha _s`$ is a new and independent exponent. We can summarize our analitycal results as follows $$\{\begin{array}{ccc}\mathrm{if}\hfill & \alpha _s<1\alpha _{loc}=\alpha _s\hfill & \{\begin{array}{c}\alpha _s=\alpha \mathrm{Family}\mathrm{Vicsek}\hfill \\ \alpha _s\alpha \mathrm{Intrinsic}\hfill \end{array}\hfill \\ \mathrm{if}\hfill & \alpha _s>1\alpha _{loc}=1\hfill & \{\begin{array}{c}\alpha _s=\alpha \mathrm{Super}\mathrm{rough}\hfill \\ \alpha _s\alpha \mathrm{New}\mathrm{class}\hfill \end{array}\hfill \end{array}$$ (13) In the following we present simulations of a one-dimensional growth model that is a nice example of the new dynamics. We have performed numerical simulations of the Sneppen model of self-organized depinning (model A) . We have found that this model exhibits anomalous roughening the type described by Eq.(8) for $`\alpha _s>1`$ and $`\alpha _s\alpha `$. In this model the height of the interface $`h(i,t)`$ is taken to be an integer defined on a one-dimensional discrete substrate $`i=1,\mathrm{},L`$. A random pinning force $`\eta (i,h)`$ is associated with each lattice site $`[i,h(i)]`$. The quenched disorder $`\eta (i,h)`$ is uniformly distributed in $`[0,1]`$ and uncorrelated. The growth algorithm is then as follows. At every time step $`t`$, the site $`i_0`$ with the smallest pinning force is chosen and its height $`h(i_0,t)`$ is updated $`h(i_0,t+1)=h(i_0,t)+1`$ provided that the conditions $`|h(i_0,t)h(i_0\pm 1,t)|<2`$ are satisfied. Periodic boundary conditions are assumed. We have studied the behaviour of the model in systems of different sizes from $`L=2^6`$ up to $`L=2^{13}`$. From calculations of the saturated global width $`W(L)`$ for various system sizes we find a global roughness exponent $`\alpha =1.000\pm 0.005`$ in agreement with previous simulations . We have checked that the scaling of the global width is given by Eq.(1) with a scaling function like (2). Also in agreement with previous work we find that the time exponent $`\alpha /z=0.95\pm 0.05`$. The local width $`w(l,t)`$ scales as $`w(l,t)=t^{\alpha /z}g(l/\xi )`$ where the scaling function is given by Eq.(12), and also $`\alpha _{loc}=1`$. From these simulation results one could conclude that the behaviour of the Sneppen growth model is rather trivial and that the exponents $`\alpha =\alpha _{loc}=z=1`$ describe its scaling properties. Quite the opposite, this model exhibits no trivial features that can be noticed when the structure factor is calculated. In Figure 1 we show our numerical results for the structure factor $`S(k,t)`$ in a system of size $`L=2048`$. Note that in Figure 1 the curves $`S(k,t)`$ for different times are shifted downwards reflecting that $`\alpha <\alpha _s`$. This contrasts with the case of intrinsic anomalous roughening where $`\alpha _s=\alpha _{loc}`$ and $`\alpha _{loc}1`$. The slope of the continuous line is $`3.7`$ and indicates that a new exponent $`\alpha _s=1.35`$ enters the scaling. This can be better appreciated by the data collapse shown in Figure 2, where one can observe that, instead of being constant, the scaling function $`s(u)`$ has a negative exponent $`u^{0.7}`$ for $`u1`$. The exponents used for the data collapse are $`\alpha =1`$, $`z=1`$ and the scaling function obtained is in excellent agreement with Eq.(8) and a spectral exponent $`\alpha _s=1.35\pm 0.03`$. The interface in the Sneppen model A is formed by facets with constant slope $`\pm 1`$ . The value of the exponents $`\alpha =\alpha _{loc}=1`$ and $`\alpha _s=1.35`$ is related to the faceted form of the interface at saturation. It is easy to understand how the anomalous spectral roughness exponent appears due to the faceted form of the interface. For the simpler (and trivial) case of a faceted interface formed by a finite number of identical segments, $`N`$, of constant slope, $`\pm m`$, one can show analitically that the global width $`W(L)m^2L^2/N^2`$, and the height-height correlation function $`G(l)l^2m^2Nm^2l^3/L`$, which leads to $`\alpha =\alpha _{loc}=1`$, while the spectrum $`S(k,L)k^4L^1`$ as $`k0`$. A simple comparison with the anomalous scaling form for the stationary spectrum $`S(k,L)k^{(2\alpha _s+1)}L^{2(\alpha \alpha _s)}`$ leads to $`\alpha _s=1.5`$. Actually, the facets occuring in the Sneppen model are not formed by identical segments, but rather follow a random distribution , which leads to a spectral exponent different from the trivial case. In summary, we have presented a generic theory of scaling for invariant surface growth. We have shown that the existence of power-law scaling of the correlation functions (scale invariance) does not determine a unique form of the scaling functions involved. This leads to the different dynamic scaling forms recently observed in growth models and experiments exhibiting anomalous roughening. In particular interface scale invariance does not necessarily imply Family-Vicsek dynamic scaling. We have derived all the types of scaling (Family-Vicsek, super-rough and intrinsic anomalous) from a unique scaling ansatz, which is formulated in the Fourier space. The different types of scaling are subclasses of our generic scaling ansatz associated with bounds on the values that the new spectral roughness exponent $`\alpha _s`$ may take. This generalization has allowed us to predict the existence of a new kind of anomalous scaling with interesting features. Simulations of a model for self-organized interface depinning have been shown to be in excellent agreement with the new anomalous dynamics. It has recently been shown that anomalous roughening stems from a non trivial dynamics of the mean local slopes $`\overline{(h)^2}`$. In contrast, the new anomalous dynamics can be pinned down to growth models in which the stationary state consits of faceted interfaces. The authors would like to thank R. Cuerno for an earlier collaboration that led to Ref.. This work has been supported by the DGES of the Spanish Government (project No. PB96-0378-C02-02). JJR is supported by the Ministerio de Educación y Cultura (Spain). JML is supported by a TMR Network of the European Commission (contract number FMRXCT980183).
warning/0001/cond-mat0001257.html
ar5iv
text
# String Approximation for Cooper Pair in High-Tc Superconductivity ## I Introduction In the quantum chromodynamics (QCD) there is a very fruitful analogy between superconductivity and QCD: according to the ’t Hooft and Mandelstam assumption hooft , mandel a confinement in the QCD is similar to the dual Meissner effect in a superconductor. Such confinement scenario assumes that the ground state of QCD is a condensate of magnetic monopoles <sup>1</sup><sup>1</sup>1we can name such condensate as a dual superconductor which one confines (pinches) the color electric field of color charges into flux tubes <sup>2</sup><sup>2</sup>2in ordinary superconductor this happens with magnetic field. In other words, the magnetic charges, defined as the Dirac monopoles in ground state of QCD, should condense in the vacuum in just the same way as the Cooper pair in a superconductor. This analogy works in the direction: superconductivity $``$ quantum chromodynamics. An analogy can be offered in the opposite direction: from the QCD to the High-T<sub>c</sub> superconductivity by the following manner. An interaction between gluons exists in the QCD and it is so strong that the force lines of color gauge field are stretched into a flux tube between quark and antiquark. The most important in this picture is the strong interaction between gluons. This one distinguishes the quantum electrodynamics (QED) from the quantum chromodynamics: QED does not have the photon-photon interaction. Thus, our basic assumption is that in the High-T<sub>c</sub> superconductor there is a strong phonon-phonon interaction. We assume that the jump from the ordinary superconductivity to High-T<sub>c</sub> one is analogous to the jump from QED to QCD. ordinary superconductivity $``$ High-T<sub>c</sub> superconductivity $``$ QED $``$ QCD The consequence of such conjectural analogy is that the interaction between phonons is so strong that the phonons between the Cooper pair are confined into a phonon tube (PT) dzhun . The phonon Hamiltonian $`_{ph}`$ in the continuum limit we can write as follows $$_{ph}=d^3𝐫\left[\frac{\rho }{2}\dot{s}_i^2(𝐫,t)+c^{kl}_ks_k_ls^k+V(𝐫,t)\right],$$ (1) where $`\rho `$ is the mass density, $`𝐬=\{s_i\}`$ ($`i=x,y,z`$) is the ion deviation from the equilibrium state, $`V`$ is the potential energy, $`c^{kl}`$ are some coefficients. Our assumption indicates that the potential term $`V`$ in the High-T<sub>c</sub> superconductor can not be simplified as $$V\frac{^2V}{s_is_j}s_is_j$$ (2) even though for the small deviations $`s_i`$. The same takes place in the QCD with Lagrangian $$_{QCD}F_{\mu \nu }^aF^{a\mu \nu },$$ (3) here $`\mu ,\nu =t,x,y,z`$; $`F_{\mu \nu }^a=_\mu A_\nu _\nu A_\mu +gf_{abc}A_\mu ^bA_\nu ^c`$; $`A_\mu ^a`$ is the SU(3) gauge potential; $`f_{abc}`$ are the structural constants of the SU(3) gauge group; $`a,b,c`$ are the color indexes. It is easy to see that in this QCD Lagrangian we have $`(A)^3`$ and $`(A)^4`$ terms. The most important is that coupling constant $`g`$ is not small, i.e. in contrast with QED we have a very strong interaction between gluons that probably leads to the appearance of a nonlocal object - flux tube stretched between quark and antiquark. ## II BCS-theory + string approximation. The basic idea of BCS theory is that in the presence of even a weak interaction between electrons they cooperate in the Cooper pairs that leads to decreasing the ground state energy of superconductor in comparison with the normal state. According to the above-mentioned assumption about the possible analogy with the QCD we suppose * There is a strong interaction between phonons (nonlinear potential term in (1) expression) which leads to appearing of a tube filled by phonons (like the flux tube filled by chromodynamical color fields in the QCD). We can name such tube as the PT. * In the first approximation, neglecting by cross section of the PT, such object is like a string. Analogously to the QCD in the first rough approximation the PT in the High-T<sub>c</sub> superconductor can be considered as a string. * As in the QCD we assume that the interaction between Cooper electrons is described by the nonrelativistic string potential $$V=kl,$$ (4) where $`V`$ is the potential energy of the phonon interaction between Cooper electrons, $`k`$ is a coefficient and $`l`$ is a length of the string. In this model the Cooper pair is modeled as a string with two Cooper electrons attached at its ends. In the QCD such construction is a string with quark and antiquark at the ends. The interaction quark-antiquark is so strong that we have confinement: if we increase the string length then in some moment the string is torn and again we will have the quark-antiquark pair, i.e. we can not obtain a single quark. By applying such model to the High-T<sub>c</sub> superconductivity we should be delicate: certainly a single electron exists <sup>3</sup><sup>3</sup>3this means that the Cooper pair can be destroyed but the basic idea is the same: the string interaction leads to an essential increase of the binding energy of the Cooper pair. Now we would like to repeat the calculations in the microscopic BCS theory with the potential (4) (following, for example, to Ref. tilley ). Let us assume that $`|00`$ is a quantum amplitude that the Cooper pair is unoccupied and $`|11`$ that it is occupied, then the wave function is $`\mathrm{\Phi }`$ $`=`$ $`{\displaystyle \underset{𝐤}{}}\psi _𝐤,`$ (5) $`\psi _𝐤`$ $`=`$ $`u_𝐤^{}|00+v_𝐤|11,`$ (6) where $`|u_𝐤|^2`$ is the probability that the Cooper pair $`(𝐤,)`$, $`(𝐤,)`$ is unoccupied and accordingly $`|v_𝐤|^2`$ is the probability that it is occupied; $`𝐤`$ is the wave vector. We have the normalization condition $$|v_𝐤|^2+|u_𝐤|^2=1.$$ (7) The normal ground state is given by $`u_𝐤=0`$ and $`v_𝐤=1\text{for}|𝐤|<𝐤_F,`$ (8) $`u_𝐤=1`$ and $`v_𝐤=0\text{for}|𝐤|>𝐤_F,`$ (9) where $`𝐤_F`$ is the Fermi wave vector. The energy of ground state is calculated by the variational method $$\delta \mathrm{\Phi }|\widehat{}\mu \widehat{N}|\mathrm{\Phi }=0$$ (10) with the constraint $`\mathrm{\Phi }|\widehat{N}|\mathrm{\Phi }=N`$, where $`\widehat{N}`$ is the number operator and $`N`$ is the mean number of particles in this system, $`\mu `$ is the Fermi energy. The calculation for kinetic energy term $`\widehat{K}`$ gives us $$\mathrm{\Phi }|\widehat{K}|\mathrm{\Phi }=2\underset{𝐤}{}\epsilon _𝐤|v_𝐤|^2,$$ (11) where the single particle energies $`\epsilon _𝐤`$ are measured from the Fermi surface $`\epsilon _F`$, the factor 2 appears as $`|v_𝐤|^2`$ is the occupation probability for the Cooper pair. The expectation value of the potential energy part is $$\mathrm{\Phi }|\widehat{V}|\mathrm{\Phi }=\underset{𝐤,𝐤^{}}{}V_{\mathrm{𝐤𝐤}^{}}u_𝐤v_𝐤^{}^{}u_𝐤^{}^{}v_𝐤,$$ (12) where $`u_𝐤^{}^{}v_𝐤`$ is the amplitude for the initial state in which the Cooper pair $`𝐤`$ is occupied and pair $`𝐤^{}`$ in unoccupied; $`u_𝐤v_𝐤^{}^{}`$ is the amplitude for the final state in which the reverse is true. Thus we have $$\mathrm{\Phi }|\widehat{}\mu \widehat{N}|\mathrm{\Phi }=2\underset{𝐤}{}\epsilon _𝐤|v_𝐤|^2+\underset{𝐤,𝐤^{}}{}V_{\mathrm{𝐤𝐤}^{}}u_𝐤v_𝐤^{}^{}u_𝐤^{}^{}v_𝐤.$$ (13) As usual, we minimize this functional with respect to $`v_𝐤^{}`$ that leads to the equation $$\mathrm{\Delta }_𝐤=\underset{𝐤^{}}{}V_{\mathrm{𝐤𝐤}^{}}\frac{\mathrm{\Delta }_𝐤^{}}{2E_𝐤^{}}$$ (14) where $`E_𝐤^2`$ $`=`$ $`\epsilon _𝐤^2+\mathrm{\Delta }_𝐤^2,`$ (15) $`\mathrm{\Delta }_𝐤`$ $`=`$ $`{\displaystyle \underset{𝐤}{}}V_{\mathrm{𝐤𝐤}^{}}u_𝐤^{}v_𝐤^{}`$ (16) As usual we simplify $$V_{\mathrm{𝐤𝐤}^{}}=\{\begin{array}{cc}\hfill V& \text{if}|\epsilon _𝐤|,|\epsilon _𝐤^{}|<\mathrm{}\omega _D\hfill \\ \hfill 0& \text{otherwise}\hfill \end{array}$$ (17) here $`\mathrm{}\omega _D`$ is the Debye energy of phonons. Now we can include our string assumption: we suppose that $`V`$ in the (17) expression is $$V=kl,$$ (18) where $`k`$ is some constant, $`l`$ is the length of Cooper pair (length of the string). As usual the further calculations give us $$|\mathrm{\Delta }|=\frac{\mathrm{}\omega _D}{\mathrm{sinh}\left({\displaystyle \frac{1}{N_0V}}\right)},$$ (19) where $`N_0`$ is the density of states in energy at the Fermi surface. ## III Discussion We can hope that in the High-T<sub>c</sub> superconductor the quantity $`V`$ in the (19) expression will be much more than the corresponding quantity in the ordinary superconductors $$V_{high}V_{ordinary}$$ (20) as it takes place in the QCD: the strong interaction between quarks is much more than the electromagnetic one between electrons (positrons). In the QCD the interaction between gluons is so strong that the force lines are confined into the flux tube stretched between quark and antiquark. The greatest difficulty here is the microscopical calculation of the quantity $`V`$. This difficulty is connected with the presence of the nonlinear potential term in the Hamiltonian. This means that in the presence of the strong self-interaction between phonons the expression for the exchange of one phonon between two electrons (see Fig. 1) $$V(\text{k,k}^{})=\frac{g^2\mathrm{}\omega _𝐪}{(\epsilon _{𝐤+𝐪})^2(\mathrm{}\omega _𝐪)^2}$$ (21) now is not correct. The same problem exists in the QCD: now we can not deduce the existence of the flux tube (string in the first rough approximation) from the SU(3) Lagrangian. This problem is connected with that we do not have appropriate mathematical tools in the nonperturbative QCD, i.e. without the Feynman diagram techniques. One of the first attempts of nonperturbative calculations in the QCD was a Veneziano amplitude venez (see Fig.2) $$A(s,t)=\frac{\mathrm{\Gamma }\left(\alpha (s)\right)\mathrm{\Gamma }\left(\alpha (t)\right)}{\mathrm{\Gamma }\left(\alpha (s)\alpha (t)\right)}$$ (22) where $`s=(p_1+p_2)^2`$ and $`t=(p_3+p_4)^2`$ are the Mandelstam variables; $`\mathrm{\Gamma }`$ is the gamma function; $`\alpha (s)=\alpha (0)+\alpha ^{}s`$; $`\alpha (0)`$ and $`\alpha ^{}`$ are some constants. This approach has not resulted in a great success in the QCD as this amplitude is some rough approximation for the correct 4-point Green’s function. It is possible to assume that such approximation can be made in the superconductivity theory : the scattering amplitude (21) of two electrons $`V(\text{k,k}^{})`$ is the Veneziano amplitude (22). Certainly (just as in QCD) it is a rough approximation for the correct Green’s function. We can try to obtain an expression for $`V`$ on the basis of the dimensional reasons. The result, for example, can be $`k`$ $``$ $`{\displaystyle \frac{\mathrm{}\omega _D}{\sqrt{S}}},`$ (23) $`V`$ $``$ $`\mathrm{}\omega _D{\displaystyle \frac{l}{\sqrt{S}}},`$ (24) where $`S`$ is the cross section of the PT. If we suppose that the PT is located inside of a layer of an anisotropical superconductor then $`\sqrt{S}\delta `$, where $`\delta `$ is the thickness of the layer. In this case these expressions will be $`k{\displaystyle \frac{\mathrm{}\omega _D}{\delta }},`$ (25) $`V\mathrm{}\omega _D{\displaystyle \frac{l}{\delta }}`$ (26) because of $`l\delta `$ we have $`V\mathrm{}\omega _D`$. ## IV Phonon tube What is the physical meaning of the PT ? In the QCD the flux tube is filled with the color field which is parallel to the tube axis and equal to zero outside this tube. In our case we can suppose that there can be one from the following possibilities $`𝐬`$ $`0\text{and/or}`$ $`𝐬^2`$ $`0\text{and/or}`$ $`𝐬^3`$ $`0\mathrm{}\text{inside PT and}`$ (27) $`𝐬`$ $`0\text{and}`$ $`𝐬^2`$ $`0\text{and}`$ $`𝐬^3`$ $`0\mathrm{}\text{outside PT}.`$ (28) The first possibility $`𝐬0`$ means that the nonzero static deviation of the ion lattice exists only inside PT. The second one $`𝐬^20`$ means that the nonzero mean square of the deviation exists only inside PT only and so on. ## V Heisenberg quantization model for the PT A possible strong phonon-phonon interaction can obstruct the application of Feynman diagram techniques in our case. Some time ago W. Heisenberg had conceived of the difficulties in applying an expansion in small parameters to quantum field theories having strong interactions. He had investigated the Dirac equation with nonlinear terms (Heisenberg equation) (see, for example, Ref’s heis1 \- heis2 ). In these papers he repeatedly underscored that a nonlinear theory with a large parameter requires the introduction of another quantization rule. He worked out a quantization method for strong nonlinear field unusing the expansion in a small parameter. It is possible that in the High - T<sub>c</sub> superconductivity the interaction between phonons is strong making it necessary take into account the interaction between phonons to correctly calculate the energy of the Cooper pairs. ### V.1 Heisenberg quantization of nonlinear spinor field Heisenberg’s basic idea proceeds from the fact that the n-point Green functions must be found from some infinity differential equations system derived from the field equation for the field operator. For example, we present Heisenberg quantization for nonlinear spinor field. The basic equation (Heisenberg equation) has the following form: $$\gamma ^\mu _\mu \widehat{\psi }(x)l^2\mathrm{}\left[\widehat{\psi }(\widehat{\overline{\psi }}\widehat{\psi })\right]=0,$$ (29) where $`\gamma ^\mu `$ are Dirac matrices; $`\widehat{\psi }(𝐱)`$ is the field operator; $`\widehat{\overline{\psi }}`$ is the Dirac adjoint spinor; $`\mathrm{}[\widehat{\psi }(\widehat{\overline{\psi }}\widehat{\psi })]=\widehat{\psi }(\widehat{\overline{\psi }}\widehat{\psi })`$ or $`\widehat{\psi }\gamma ^5(\widehat{\overline{\psi }}\gamma ^5\widehat{\psi })`$ or $`\widehat{\psi }\gamma ^\mu (\widehat{\overline{\psi }}\gamma _\mu \widehat{\psi })`$ or $`\widehat{\psi }\gamma ^\mu \gamma ^5(\widehat{\overline{\psi }}\gamma _\mu \gamma ^5\widehat{\psi })`$. Heisenberg emphasizes that the 2-point Green function $`G_2(𝐱_2,𝐱_1)`$ in this theory differs strongly from the propagator in linear theory. This difference lies in its behaviour on the light cone. $`G_2(𝐱_2,𝐱_1)`$ oscillates strongly on the light cone in contrast to the propagator of the linear theory which has a $`\delta `$-like singularity. Then Heisenberg introduces the $`\tau `$ functions: $$\tau (𝐱_1𝐱_2\mathrm{}|𝐲_1𝐲_2\mathrm{})=0|T\psi (𝐱_1)\psi (𝐱_2)\mathrm{}\psi ^{}(𝐲_1)\psi ^{}(𝐲_2)\mathrm{}|\mathrm{\Phi },$$ (30) where $`T`$ is the time ordering operator. $`|\mathrm{\Phi }`$ is a system state characterized by the fundamental Eq. (29). Relationship (30) allows us to establish a one-to-one correspondence between the system state $`|\mathrm{\Phi }`$ and the function set $`\tau `$. This state can be defined using the infinite function set of (30). Applying Heisenberg’s equation (29) to (30) we can obtain the following infinite equations system: $`l^2\gamma _{(r)}^\mu {\displaystyle \frac{}{x_{(r)}^\mu }}\tau (𝐱_1\mathrm{}𝐱_n|𝐲_1\mathrm{}𝐲_n)=\mathrm{}\left[\tau (𝐱_1\mathrm{}𝐱_n𝐱_r|𝐲_1\mathrm{}𝐲_n𝐲_r)\right]+`$ $`\delta (𝐱_r𝐲_1)\tau \left(𝐱_1\mathrm{}𝐱_{r1}𝐱_{r+1}\mathrm{}𝐱_n|𝐲_2\mathrm{}𝐲_{r1}𝐲_{r+1}\mathrm{}𝐲_n\right)+`$ $`\delta (𝐱_r𝐲_2)\tau \left(𝐱_1\mathrm{}𝐱_{r1}𝐱_{r+1}\mathrm{}𝐱_n|𝐲_1𝐲_2\mathrm{}𝐲_{r1}𝐲_{r+1}\mathrm{}𝐲_n\right)+\mathrm{}.`$ (31) Heisenberg then employs the Tamm - Dankoff method for getting approximate solutions to the infinite equations system of (31). The key to this method lies in the fact that the system of equation has an approximate solution derived after cutting off the infinite equation system (31) to a finite equation system. It is necessary to note that a method of solution to Eq. (31) can be various. For example, we can try to determine the Green’s functions using the numerical lattice calculations. Here the important point is the following: The technique of expansion in small parameters (Feynman diagrams) can not be employed for strong nonlinear fields. It is possible that as in quantum chromodynamics, where quarks are thought to interact strongly by means of flux tubes, so too in High-T<sub>c</sub> superconductivity phonons may strongly interact among themselves. In Ref. vds ; dzhun2 such a mechanism is applied for the QCD. In this paper we apply a variant of Heisenberg’s quantization method to solutions of the classical SU(3) Yang-Mills field equations which have bad asymptotic behavior. After quantization it has been found that the bad features (i.e. divergent fields and energy densities) of these solutions are moderated. From these results is argued that in general the n-point Green’s functions for Yang-Mills theories can have nonperturbative pieces which can not be represented as the sum of Feynman diagrams. In Heisenberg’s theory the matter and the interacting fields are identical: fundamental spinor field $`\psi (x)`$. From a more recent perspective this is not the case. An interaction is carried by some kind of boson field. In superconductivity this is the phonons, in quantum chromodynamics it is the nonabelian $`SU(3)`$ gauge field - gluons. In conclusion of this section we emphasize again Heisenberg’s statement that the perturbation theory is inapplicable to strong nonlinear fields. ### V.2 Heisenberg quantization method for High-T<sub>c</sub> superconductivity Thus, the basic assumption supposed here is the following: The energy of Cooper pair has an essential contribution coming from an interaction of phonons. This means that the corresponding sound wave is a nonlinear wave. #### V.2.1 An application of Heisenberg quantization method for the Green’s function method. In this subsection we would like to show that Heisenberg idea about quantization of strongly interacting fields in fact had been applied in the superconductivity theory for definition of Green’s function. Here we would like to show that the calculations which was made in Ref. gor is an application of the Heisenberg quantization idea. In this section we follow to Ref.abr . The Hamiltonian of the system electrons describing the properties of a metal in the superconductivity state is $$\widehat{H}=\left[\left(\widehat{\psi }_\alpha ^+\frac{^2}{2m}\widehat{\psi }_\alpha \right)+\frac{\lambda }{2}\left(\widehat{\psi }_\beta ^+\left(\widehat{\psi }_\alpha ^+\widehat{\psi }_\alpha \right)\widehat{\psi }_\beta \right)\right]𝑑V,$$ (32) where $`\widehat{\psi }_\alpha `$ is the operator of spinor field describing electrons, $`m`$ is the electron mass, $`\lambda `$ is some constant and $`\alpha ,\beta `$ are the spinor indexes. As usually (it is the same as Heisenberg equation (29) for nonlinear spinor field) the operators $`\widehat{\psi }`$ and $`\widehat{\psi }^+`$ obey the following operator equations $`\left(i{\displaystyle \frac{}{t}}+{\displaystyle \frac{^2}{2m}}\right)\widehat{\psi }_\alpha (x)\lambda \left(\widehat{\psi }_\beta ^+(x)\widehat{\psi }_\beta (x)\right)\widehat{\psi }_\alpha (x)`$ $`=`$ $`0,`$ (33) $`\left(i{\displaystyle \frac{}{t}}{\displaystyle \frac{^2}{2m}}\right)\widehat{\psi }_\alpha ^+(x)+\lambda \widehat{\psi }_\alpha ^+(x)\left(\widehat{\psi }_\beta ^+(x)\widehat{\psi }_\beta (x)\right)`$ $`=`$ $`0.`$ (34) As well as in Heisenberg method for nonlinear spinor field we have an equation for the 2-point Green’s function $`G_{\alpha \beta }(x,x^{})=iT(\widehat{\psi }_\alpha (x)\widehat{\psi }_\beta ^+(x^{}))`$ $$\left(i\frac{}{t}+\frac{^2}{2m}\right)G_{\alpha \beta }(x,x^{})+i\lambda T\left(\widehat{\psi }_\gamma ^+(x)\widehat{\psi }_\gamma (x)\widehat{\psi }_\alpha (x)\widehat{\psi }_\beta ^+(x^{})\right)=\delta (xx^{}).$$ (35) Further we have to write an equation for term $`T(\widehat{\psi }_\gamma ^+(x)\widehat{\psi }_\gamma (x)\widehat{\psi }_\alpha (x)\widehat{\psi }_\beta ^+(x^{}))`$ and so on. After this we will have an infinite set for Green’s function. The main problem here is: how can we cut off this infinite equation system ? Let us cite Ref.abr : “$`\mathrm{}`$ For interacting particles, the product of four $`\widehat{\psi }`$-operators can be expressed in terms of the vertex part, i.e. it already includes the contributions from various scattering processes. In the weak-interaction model under consideration, these scattering processes involving collisions of particles can be neglected, but at the same time, it must be borne in mind that the ground state of the system differs from the usual state with a filled Fermi sphere, because of the presence of bound pairs of electrons. $`\mathrm{}`$”. This means that in the textbook abr it was made the following approximation: the operators $`\widehat{\psi }\widehat{\psi }`$ and $`\widehat{\psi }^+\widehat{\psi }^+`$ contain terms corresponding to the annihilation and creation of bound pairs (Cooper pairs). It allows to state $$\begin{array}{c}\hfill T\left(\widehat{\psi }_\alpha (x_1)\widehat{\psi }_\beta (x_2)\widehat{\psi }_\gamma ^+(x_3)\widehat{\psi }_\delta ^+(x_4)\right)\\ \hfill T\left(\widehat{\psi }_\alpha (x_1)\widehat{\psi }_\gamma ^+(x_3)\right)T\left(\widehat{\psi }_\beta (x_2)\widehat{\psi }_\delta ^+(x_4)\right)+\\ \hfill T\left(\widehat{\psi }_\alpha (x_1)\widehat{\psi }_\delta ^+(x_4)\right)T\left(\widehat{\psi }_\beta (x_2)\widehat{\psi }_\gamma ^+(x_3)\right)+\\ \hfill N\left|T\left(\widehat{\psi }_\alpha (x_1)\widehat{\psi }_\beta (x_2)\right)\right|N+2N+2\left|T\left(\widehat{\psi }_\gamma ^+(x_3)\widehat{\psi }_\delta ^+(x_4)\right)\right|N\end{array}$$ (36) where $`|N`$ and $`|N+2`$ are ground states of system with $`N`$ and $`N+2`$ particles (Cooper pairs), respectively. We should note that this expression is approximate one and following to Heisenberg idea it allows us to cut off the above-mentioned infinite equation set for Green’s function. The key role here plays the last term in expression (36). This expression allows us to split the 4-point Green’s function with the nonperturbative way not using Feynman diagram techniques. Eq. (36) means that we have neglected all effects of scattering particles by each other and the presence of the interaction has been taken into account only as leading to the formation of bound pairs. The third term in the right-hand side of Eq. (36) has been written in complete analogy with the case of a Bose gas in the correspondence with the fact that all bound pairs (Cooper pairs) are “condensed on the lowest level”. The quantity $$N\left|T\left(\widehat{\psi }_\alpha \widehat{\psi }_\beta \right)\right|N+2N+2\left|T\left(\widehat{\psi }_\gamma ^+\widehat{\psi }_\delta ^+\right)\right|N,$$ (37) obviously has the same order of magnitude as the density of pairs. We can introduce the following functions $`e^{2i\mu t}F_{\alpha \beta }(xx^{})`$ $`=`$ $`N\left|T\left(\widehat{\psi }_\alpha (x)\widehat{\psi }_\beta (x^{})\right)\right|N+2`$ (38) $`e^{2i\mu t}F_{\alpha \beta }^+(xx^{})`$ $`=`$ $`N+2\left|T\left(\widehat{\psi }_\gamma ^+(x)\widehat{\psi }_\delta ^+(x^{})\right)\right|N,`$ (39) where $`\mu `$ is a chemical potential. We now substitute (36) into the equation (33) for the Green’s function. We can everywhere omit the first two terms in the right-hand side of (36) since they lead to an additive correction to the chemical potential in the equations for the functions $`G,F,F^+`$. As a result we obtain the following equation connecting $`G`$ and $`F^+`$: $$\left(i\frac{}{t}+\frac{^2}{2m}\right)G(xx^{})i\lambda F(0+)F^+(xx^{})=\delta (xx^{}).$$ (40) The quantity $`F(0+)`$ is defined as $$F_{\alpha \beta }(0+)=e^{2i\mu t}N\left|\widehat{\psi }_\alpha (x)\widehat{\psi }_\beta (x)\right|N+2=\underset{\begin{array}{c}xx^{},\\ tt^{}+0\end{array}}{lim}F_{\alpha \beta }(xx^{}).$$ (41) An equation for $`F^+(xx^{})`$ can be obtained in a similar way by using the equation (34) $$\left(i\frac{}{t}\frac{^2}{2m}2\mu \right)F^+(xx^{})+i\lambda F^+(0+)G(xx^{})=0.$$ (42) Analogously to Eq.(41) we have $$F_{\alpha \beta }^+(0+)=e^{2i\mu t}N+2\left|\psi _\alpha ^+(x)\psi _\beta ^+(x)\right|N$$ (43) Now we have Eq’s.(40) and (42) as the finite set of equations for the 2-point Green’s functions $`G`$ and $`F`$. Thus, above-mentioned reasonings should convince us that this Green’s function method in the superconductivity theory is a realization of Heisenberg idea developed by him for the quantization of non-linear spinor field. #### V.2.2 Phonon-phonon interaction In this subsection we would like to consider a possibility for the phonons to form a static configuration. Let us write again the Lagrangian (1) for phonons in continuous limit $$_{ph}=d^3𝐫\left[\frac{\rho }{2}\dot{s}_i^2(𝐫,t)+c^{kl}_ks_k_ls^kV(𝐫,t)\right].$$ (44) Thus, in this model it is assumed that operators of strong nonlinear fiels $`s_i`$ must satisfy the following equation (which is implied from the Lagrangian (44)): $$\frac{^2\widehat{s}_i}{t^2}+c^{kl}\frac{^2\widehat{s}_i}{x^kx^l}=\frac{dV(\widehat{s}_i)}{d\widehat{s}_i}.$$ (45) The multitime formalism of Heisenberg’s method (when in $`\tau (t_1,t_2,\mathrm{}),t_1t_2\mathrm{}`$) allows us to investigate the scattering processes in quantum theory. The simultaneous formalism (when $`\tau (t_1,t_2,\mathrm{}),t_1=t_2=\mathrm{}=t`$) allows us to calculate the mean value of the field, the energy, or any combination of field powers. Let us consider, for example, the simplest case of the scalar field when the vector $`𝐬`$ is replaced by the scalar field $`\phi `$. In this case the Lagrangian is $$=d^3𝐫\left[\left(_\mu \phi \right)^2V(\phi )\right]$$ (46) where $`\mu =t,x,y,z`$ and the potential term $$V(\phi )=\frac{\lambda }{4}\left(\phi ^2\phi _0^2\right).$$ (47) The classical field equation is $$\mathrm{}\phi (𝐫)=\lambda \phi (𝐫)\left(\phi ^\mathrm{𝟐}(𝐫)\phi _\mathrm{𝟎}^\mathrm{𝟐}\right).$$ (48) The quantization of (48) equation gives us $$\mathrm{}\widehat{\phi }(𝐫)=\widehat{\phi ^3(𝐫)}\widehat{\phi }(𝐫)\phi _0^2.$$ (49) It is easy to see that the mean value $`\phi (𝐫)=0|\widehat{\phi }(𝐫)|0`$ satisfies the following equation: $$\mathrm{}\phi (𝐫)=\phi ^3(𝐫)\phi _0^2\phi (𝐫).$$ (50) For the definition of $`\phi ^3(𝐫)`$ we turn to Eq.(49) and obtain ($`\phi ^3(𝐫)=\tau (\mathrm{𝐫𝐫𝐫}`$) in Heisenberg’s notation): $$\mathrm{}\phi ^3(𝐫)=3\lambda \left(\phi ^5(𝐫)\phi _0^2\phi ^3(𝐫)\right),$$ (51) here $`\phi ^5(𝐫)=\tau (\mathrm{𝐫𝐫𝐫𝐫𝐫})`$. Analogously it can be used to derive the infinite equation system for calculating $`\phi ^n(𝐫)`$. In the first approximation we can solve this equation system using the following assumption: $$\phi ^3(𝐫)\phi (𝐫)^3,$$ (52) and then we can derive the equation $$\mathrm{}\phi (𝐫)=\lambda \phi (𝐫)\left(\phi (𝐫)^2\phi _0^2\right).$$ (53) This equation is very interesting for us. Derrick’s Theorem derr states that Eq. (53) has a static solution with the finite energy only in two dimensional spacetime $`(t,x)`$. This solution is $$\phi (x)=\phi _1\mathrm{tanh}\left[\frac{1}{2}m\left(xx_0\right)\right]$$ (54) where $`\phi _1,m`$ and $`x_0`$ are some constants. The first assumption is that such solution can be realized on the layer of High-T<sub>c</sub> superconductor as a line (wall) separating regions with different vacua ($`\phi =\pm \phi _0`$). The existence of regions with different vacua can have very interesting physical consequences for a distribution of free electrons in the ion lattice. Following Ref.jackiw we present a model example of a fermion coupled to solution (54). Let we have 2-dimensional Dirac equation $$\left(\gamma ^\mu _\mu +g\phi (x)\right)\psi (x)=0$$ (55) where $`\psi =\left(\genfrac{}{}{0pt}{}{\psi _1}{\psi _2}\right)`$ is the two-component spinor; $`g`$ is the coupling constant; $`\phi (x)`$ is the soliton (54); $`\gamma ^\mu `$ are the Dirac matrices, which we can choose to be the Pauli matrices $$\gamma ^1=\sigma _1,\gamma ^4=\sigma _3,$$ (56) Now we search a static solution of Dirac equation in the presence of the soliton $$\left(\sigma _1_x+g\phi \right)\psi =E\sigma _3\psi =0$$ (57) here we consider case with energy $`E=0`$. In this case we have the following normalizable zero mode $$\psi =\psi _0\left\{\mathrm{cosh}\left[\frac{m}{2}\left(xx_0\right)\right]\right\}^{2m/m_\psi }\left(\genfrac{}{}{0pt}{}{1}{1}\right)$$ (58) where $`m_\psi =g\phi _0`$. We see that it is strongly localized at the position $`x_0`$ of the soliton (see, Fig.3). In the context of mechanism with different vacua this toy model can mean that the free electrons on the layer of the High-T<sub>c</sub> superconductor will be located on the boundaries between regions with different vacua. Evidently, the appearance of localized solutions is connected with the presence of two vacuums ($`\phi =\pm \phi _0`$). It allows us to assume that the mechanism offered here for the formation of Cooper pair in High-$`T_c`$ superconductivity depends on the existence of several vacuums for phonon field, i.e. on the presence of several equilibrium states (stable and unstable) for ions of the lattice. Such interpretation depends on the existence of different vacua. Thus, this mechanism takes place only in the presence of different equilibrium states for ions of the lattice. It is necessary to note that can exist other mechanism of the field localization in some region. It is possible, for example, that such mechanism (without different vacua of gauge field) occurs in QCD by forming a hypothesized flux tube between quark-antiquark. Another possibility can be connected with cut-in of an external gauge field. In the Ref. nielsen it was shown that the scalar field coupled with the electromagnetic field can forms the flux tube (Nielsen - Olesen flux tube). The detailed investigation of such possibility in the context of the Gingbug-Landau theory is given in Ref.obukh . It should be pointed out that the investigation of $`\tau (\mathrm{𝐫𝐫}\mathrm{})=\phi ^n(𝐫)`$ gives us the information about the mean value of the field $`\phi (𝐫)`$. For the investigation of questions on the scattering or interaction of phonons it is necessary to explore the functions, $`\tau (𝐫_1𝐫_2\mathrm{})=0|\phi (𝐫_1)\mathrm{}\phi (𝐫_n)|0`$. ## VI An experimental testing How can this string model for the Cooper pair in the High-T<sub>c</sub> superconductor be proved ? The best way is the direct observation of the PT between Cooper electrons. Evidently it should be connected with the detection of the ion lattice between Cooper electrons. If we can establish that $`s_i^n0`$ inside of the tube and $`s_i^n0`$ outside of the PT then probably it will indicate the existence of the PT. Certainly, such measurements of the state of ion lattice are very difficult. Another indirect way is a test for the nonlinear potential $`V(𝐫)`$. The presence of such nonlinear term can lead to: (a) the nonlinear effects by propagation of the sound wave in the High - T<sub>c</sub> superconductor along the layer; (b) the different equilibrium states of ions in the lattice. Such effects can be: (a) nonlinear scattering, propagation, absorption and so on for the sound waves; (b) appearance of regions with different vacua. Certainly, the presence of such kind of nonlinear effects is not the direct demonstration of the PT but it can indicate the important role of the strong phonon-phonon interaction in the High-T<sub>c</sub> superconductor. ## VII Conclusions The basic goal of this paper is an assumption that the quantum solid theory with the strong nonlinear interaction of ions in the lattice can be similar to the QCD in which there is the strong gluon-gluon interaction. Such similarity can lead to the appearance of nonlocal objects in the quantum solid theory like the flux tube in the QCD. It is possible that such a mechanism is realized in the High-$`T_c`$ superconductor by such a way that the phonons are confined into a nonlocal object (tube) located between two Cooper electrons. In this paper we have considered only one mechanism probably leading in emergence the PT. The problem on existence of other alternative such mechanisms was not discussed in this paper and remains while open. It is very important to note that our analysis shows that the Green’s function method in the superconductivity theory is a realization of an algorithm proposed by Heisenberg for quantization of nonlinear spinor field. We have shown that Green’s function method is such realization of discussed Heisenberg idea. We would like to emphasize that the nonperturbative Heisenberg quantization method is much more powerful than a perturbative Feynman diagram techniques. Finally, we can presuppose that in the quantum field theory nonlinearity can lead to nonlocality <sup>4</sup><sup>4</sup>4in the case of strong nonlinearity, of course: nonlinear terms $`(A)^3`$ and $`(A)^4`$ in the QCD lead to the flux tube and probably such a mechanism in the High-T<sub>c</sub> superconductivity leads to the PT. ###### Acknowledgements. This work is supported by a George Forster Research Fellowship from the Alexander von Humboldt Foundation. I would like to thank H.-J. Schmidt for the invitation to Potsdam University for research.
warning/0001/gr-qc0001009.html
ar5iv
text
# Introduction ## Introduction In classical physics it is expected that the energy-momentum tensor will satisfy the weak energy condition $`T_{\mu \nu }V^\mu V^\nu 0`$ for all timelike vectors $`V^\mu `$ (see for exceptions). This condition ensures that all observers measure a positive energy density. However, in quantum field theory there exist states for which the expectation value of the energy-momentum tensor violates the weak energy condition. Such exotic matter is of interest since it is required to maintain wormholes and to create warp drives . Over the past decade Ford and Roman have studied the properties of exotic matter extensively. In four dimensional flat spacetime they have shown that massless bosonic fields satisfy the quantum inequalities $$\widehat{\rho }=\frac{t_0}{\pi }_{\mathrm{}}^{\mathrm{}}\frac{<T_{tt}>}{(t^2+t_0^2)}𝑑t\frac{3A}{32\pi ^2t_0^4}$$ (1) where $`<T_{tt}>`$ is the expectation value of the normal ordered energy density, $`A=1`$ for massless scalar fields, and $`A=2`$ for the electromagnetic field. The quantity $`\widehat{\rho }`$ samples $`<T_{tt}>`$ over a time interval of order $`t_0`$ using the sampling function $`h(t)=t_0/[\pi (t^2+t_0^2)]`$. These quantum inequalities show that an observer can measure a negative energy density for a time $`\stackrel{<}{}|\rho |^{1/4}`$. In two dimensional flat spacetime they found that $$\widehat{\rho }\frac{1}{8\pi t_0^2}$$ (2) for a massless scalar field. The bounds found by Ford and Roman are not optimal bounds. In two dimensional flat spacetime Flanagan has shown that the optimal bound for a scalar field is given by $$\widehat{\rho }\frac{1}{24\pi }_{\mathrm{}}^{\mathrm{}}\frac{h^{^{}}(t)^2}{h(t)}𝑑t$$ (3) where $`h(t)`$ is any sampling function that satisfies $`h(t)0`$ and $`_{\mathrm{}}^{\mathrm{}}h(t)𝑑t=1`$. Non-optimal quantum inequalities have also been found for scalar fields in static Robertson-Walker, de Sitter, and Schwarzschild spacetimes . Some work has also been done on negative energy density states for the Dirac equation in four dimensional Minkowski space. In an earlier paper I examined a class of states that produced violations of the weak energy condition and showed that they satisfied a quantum inequality of the form (1). In this paper I show that (2) is also an optimal bound for the massless Dirac field in two dimensional flat spacetimes. I also find the optimal bound for massless Dirac and scalar fields in two dimensional curved spacetimes with a conformal factor dependent on $`x`$ only or on $`t`$ only. These results are then applied to two dimensional black hole and cosmological spacetimes. It is shown that the bound on the negative energies diverges to minus infinity (neglecting back reaction) as the event horizon or cosmological singularity is approached. Thus, the negative energies become unconstrained near the horizon or the initial singularity. The quantum inequalities derived in this paper can also be used to support the quantum interest conjecture . According to this conjecture any negative energy flux must be preceded or followed by a larger positive energy flux (i.e. the negative flux must be repaid with interest by the positive flux). Pretorius has shown that the quantum interest conjecture for scalar fields in Minkowski space follows from the scaling properties of the corresponding quantum inequalities. Since the Dirac equation satisfies the same quantum inequalities in two dimensional Minkowski space it will also satisfy the quantum interest conjecture. Pretorius also conjectured that in curved spacetime the quantum interest conjecture applies only to deviations from the ground state, not to the total energy. This conjecture is shown to be true for the spacetimes examined in this paper. I will take $`\mathrm{}=c=G=1`$ throughout this paper. ## The Energy-Momentum Tensor In any two dimensional spacetime the metric can be taken to have the conformally flat form $$ds^2=C(x,t)[dt^2dx^2].$$ (4) In null coordinates ($`u=tx,v=t+x`$) the metric becomes $$ds^2=C(u,v)dudv.$$ (5) The conservation laws $`_\mu T_\nu ^\mu =0`$ give $$\frac{1}{\sqrt{g}}\frac{}{x^\alpha }\left[\sqrt{g}T_\beta ^\alpha \right]=\frac{1}{2}T_\beta (\mathrm{ln}C)$$ (6) where $`T=T_\alpha ^\alpha `$ and $`T_\beta ^\alpha `$ is the energy-momentum tensor. Using $`T_u^u=T_v^v=\frac{1}{2}T`$ and $`T_{uu}=\frac{1}{2}CT_u^v,T_{vv}=\frac{1}{2}CT_v^u`$ we find that $$_vT_{uu}=\frac{1}{4}C_uT$$ (7) and $$_uT_{vv}=\frac{1}{4}C_vT.$$ (8) Thus in flat spacetime, where the conformal anomally vanishes (i.e. $`T=0`$), a conformally invariant theory, such as the massless scalar or Dirac field, will have $`T_{uu}=T_{uu}(u)`$, $`T_{vv}=T_{vv}(v)`$, and $`T_{uv}=0`$. This can be seen explicitly for the massless scalar and Dirac fields. The scalar field equation $`\mathrm{}^2\varphi =0`$ has the general solution $`\varphi (u,v)=\varphi _u(u)+\varphi _v(v)`$ and the energy-momentum tensor has the components $`T_{uu}=(_u\varphi _u)^2,T_{vv}=(_v\varphi _v)^2`$, and $`T_{uv}=0`$. The massless Dirac equation $$i\gamma ^\mu _\mu \psi =0$$ (9) with $$\gamma ^0=\left(\begin{array}{cc}0i& \\ i\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \end{array}\right)$$ (10) and $$\gamma ^1=\left(\begin{array}{cc}0i& \\ i\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \end{array}\right)$$ (11) has the general solution $$\psi =\left[\begin{array}{cc}\psi _u(u)& \\ \psi _v(v)& \end{array}\right].$$ (12) The energy-momentum tensor is given by $$T_{uu}=\frac{i}{2}[\psi _u^{}_u\psi _u(_u\psi _u^{})\psi _u],$$ (13) $$T_{vv}=\frac{i}{2}[\psi _v^{}_v\psi _v(_v\psi _v^{})\psi _v],$$ (14) and $`T_{uv}=0`$. Note that $`\psi `$ can be taken to be real since the gamma matrices are purely imaginary. For a discussion of the 1+1 dimensional Dirac equation see chapter four of Green, Schwarz, and Witten . Both of these energy-momentum tensors satisfy $`T_{uu}=T_{uu}(u)`$, $`T_{vv}=T_{vv}(v)`$, and $`T_{uv}=0`$. ## Quantum Inequalities Consider two conformally related spacetimes, one with metric $$ds^2=C(x,t)[dt^2dx^2]$$ (15) and the other one with $`C(x,t)=1`$. The energy-momentum tensor for scalar and Dirac fields satisfies $$<T_{\alpha \beta }^{(C)}>=<T_{\alpha \beta }^{(\eta )}>+\mathrm{\Theta }_{\alpha \beta }\frac{1}{48\pi }RC\eta _{\alpha \beta }$$ (16) where $`T_{\alpha \beta }^{(C)}`$ is the renormalized energy-momentum tensor with the metric given in (15), $`T_{\alpha \beta }^{(\eta )}`$ is the energy-momentum tensor with $`C(x,t)=1`$, $$\mathrm{\Theta }_{uu}=\frac{1}{12\pi }C^{\frac{1}{2}}_u^2C^{\frac{1}{2}},$$ (17) $$\mathrm{\Theta }_{vv}=\frac{1}{12\pi }C^{\frac{1}{2}}_v^2C^{\frac{1}{2}},$$ (18) $$\mathrm{\Theta }_{uv}=\mathrm{\Theta }_{vu}=0,$$ (19) and $`R=\mathrm{}^2\mathrm{ln}C`$ is the Ricci scalar. First consider the quantum inequalities in a two dimensional flat spacetime. The metric is $$ds^2=dt^2dx^2=dudv.$$ (20) Now consider the conformally related spacetime $$ds^2=f^{^{}}(v)dudv=dudV$$ (21) where $`V=f(v)`$. From (16) we find that $$<T_{vv}^{(f^{^{}})}>=<T_{vv}^{(\eta )}>+\mathrm{\Theta }_{vv},$$ (22) which gives $$(f^{^{}})^2<T_{VV}>=<T_{vv}^{(\eta )}>+\mathrm{\Theta }_{vv},$$ (23) where $`T_{VV}`$ is $`VV`$ component of the energy-momentum tensor in the (u,V) coordinate system. Now multiply by the sampling function $`h(v)`$ and integrate over $`v`$ to get $$_{\mathrm{}}^{\mathrm{}}f^{^{}}(v)h(v)<T_{VV}>f^{^{}}(v)𝑑v=_{\mathrm{}}^{\mathrm{}}h(v)<T_{vv}^{(\eta )}>𝑑v+_{\mathrm{}}^{\mathrm{}}h(v)\mathrm{\Theta }_{vv}𝑑v.$$ (24) Choose $`f^{^{}}(v)`$ such that $`f^{^{}}(v)h(v)=1`$. This gives $$_{\mathrm{}}^{\mathrm{}}<T_{VV}>𝑑V=_{\mathrm{}}^{\mathrm{}}h(v)<T_{vv}^{(\eta )}>𝑑v+\mathrm{\Delta }$$ (25) where $$\mathrm{\Delta }=\frac{1}{48\pi }_{\mathrm{}}^{\mathrm{}}\frac{h^{^{}}(v)^2}{h(v)}𝑑v.$$ (26) Note that I have taken $`h^{^{}}(v)0`$ as $`v\pm \mathrm{}`$ to obtain (26). Now the left hand side of (25) is greater than or equal to zero since it is the normal ordered Hamiltonian for the left moving sector. Thus $$_{\mathrm{}}^{\mathrm{}}h(v)<T_{vv}(v)>𝑑v\mathrm{\Delta }.$$ (27) This is an optimal bound since the equality can be reached by using the ground state of the left moving sector. A similar result holds for $`_{\mathrm{}}^{\mathrm{}}h(u)<T_{uu}(u)>𝑑u`$. Thus from $`T_{tt}=T_{uu}+T_{vv}`$ we find the optimal inequality (at fixed $`x`$) $$\widehat{\rho }=_{\mathrm{}}^{\mathrm{}}h(t)<T_{tt}(t)>𝑑t\frac{1}{24\pi }_{\mathrm{}}^{\mathrm{}}\frac{h^{^{}}(t)^2}{h(t)}𝑑t$$ (28) for scalar and Dirac fields in two dimensional flat spacetime. This is the result obtained by Flanagan for scalar fields. As discussed in the introduction, the above quantum inequality implies that the Dirac field satisfies the quantum interest conjecture in two dimensional Minkowski space. Now consider a two dimensional curved spacetime with $`C=C(x)`$. An observer with velocity $`V^\alpha `$ will measure $$\rho =<T_{\alpha \beta }>V^\alpha V^\beta .$$ (29) for the the expectation value of the energy density. From (16) we find $$\rho _{(C)}=\frac{1}{C}\rho _{(\eta )}+\frac{1}{C}\left[\mathrm{\Theta }_{\alpha \beta }V_{(\eta )}^\alpha V_{(\eta )}^\beta \frac{RC}{48\pi }\right]$$ (30) where $`\rho _{(C)}`$ is the energy density in the spacetime, $`\rho _{(\eta )}`$ is the energy density in the conformally related flat spacetime, and $`V_{(\eta )}^\alpha `$ is the four velocity of the observer in the flat spacetime (note that $`V_{(C)}^\alpha =C^{\frac{1}{2}}V_{(\eta )}^\alpha )`$. Now consider an observer at rest in the $`(x,y)`$ coordinate system. Multiply (30) by $`h_C(t)\sqrt{C}`$ and integrate to get $$_{\mathrm{}}^{\mathrm{}}\rho _{(C)}h_C\sqrt{C}𝑑t=\frac{1}{C}_{\mathrm{}}^{\mathrm{}}\rho _{(\eta )}h_\eta 𝑑t\widehat{\mathrm{\Delta }}$$ (31) where $`h_\eta =\sqrt{C}h_C`$ and $$\widehat{\mathrm{\Delta }}=\frac{1}{C}\left[\mathrm{\Theta }_{tt}\frac{RC}{48\pi }\right].$$ (32) Note that $`_{\mathrm{}}^{\mathrm{}}h_C\sqrt{C}𝑑t=_{\mathrm{}}^{\mathrm{}}h_\eta 𝑑t=1`$. Using $`R=\mathrm{}^2\mathrm{ln}C`$ and $`\mathrm{\Theta }_{tt}=\frac{1}{24\pi }C^{\frac{1}{2}}_x^2(C^{\frac{1}{2}})`$ gives $$\widehat{\mathrm{\Delta }}=\frac{1}{6\pi C}[C^{\frac{1}{4}}_x^2(C^{\frac{1}{4}})].$$ (33) Thus from (28), (31), and (33) we find that $$\widehat{\rho }_C\frac{1}{24\pi C}\left[_{\mathrm{}}^{\mathrm{}}\frac{h_\eta ^{^{}}(t)^2}{h_\eta (t)}𝑑t+4C^{\frac{1}{4}}_x^2(C^{\frac{1}{4}})\right].$$ (34) This is an optimal bound since the equality can be reached by using the state that minimizes $`_{\mathrm{}}^{\mathrm{}}\rho _{(\eta )}h_\eta 𝑑t`$. The first term on the right hand side of (34) will satisfy the scaling arguments used by Pretorius , but the second term (i.e. the vacuum polarization term) does not. Thus, the quantum interest conjecture applies to deviations from the vacuum polarization energy, not to the total energy. A similar bound can be derived for spacetimes with $`C=C(t)`$. In this case we will average over space instead of time. It is easy to show that (28) becomes $$\stackrel{~}{\rho }=_{\mathrm{}}^{\mathrm{}}h(x)<T_{tt}>𝑑x\frac{1}{24\pi }_{\mathrm{}}^{\mathrm{}}\frac{h^{^{}}(x)^2}{h(x)}𝑑x$$ (35) and that (34) becomes $$\stackrel{~}{\rho }_C\frac{1}{24\pi C}\left[_{\mathrm{}}^{\mathrm{}}\frac{h_\eta ^{^{}}(x)^2}{h_\eta (x)}𝑑x+\frac{1}{2}\left(\frac{_tC}{C}\right)^2\right].$$ (36) This is also an optimal bound since the equality can be reached by using the state that minimizes $`_{\mathrm{}}^{\mathrm{}}\rho _\eta h_\eta 𝑑x`$. From the above inequalities we see that the right hand side will generally diverge as $`C0`$. Thus, neglecting back reaction, the negative energy densities become unconstrained in regions where $`C0`$. ## Black Hole and Cosmological Spacetimes Consider the two dimensional black hole spacetime $$ds^2=\left(1\frac{2m}{\overline{r}}\right)dt^2\left(1\frac{2m}{\overline{r}}\right)^1d\overline{r}^2.$$ (37) In conformally flat coordinates $$ds^2=\left[1\frac{2m}{\overline{r}(r)}\right](dt^2dr^2)$$ (38) where $`\overline{r}(r)`$ is defined via $$r=\overline{r}+2m\mathrm{ln}\left|\frac{\overline{r}}{2m}1\right|.$$ (39) Note that the horizon is at $`\overline{r}=2m`$ and at $`r\mathrm{}`$. Now consider an observer at rest near the horizon. Close to the horizon $$C(r)e^{r/2m}$$ (40) and $$\widehat{\rho }_C\frac{1}{24\pi C}\left[_{\mathrm{}}^{\mathrm{}}\frac{h_\eta ^{^{}}(t)^2}{h_\eta (t)}𝑑t+\frac{1}{16m^2}\right].$$ (41) Thus, neglecting back reaction, as the horizon is approached the right hand side diverges and the quantum inequality does not constrain the negative energy densities. Consider a specific example using the sampling function $$h_C(t)=\frac{\tau _0}{\pi [Ct^2+\tau _0^2]},$$ (42) which has been used by Ford and Roman in the flat spacetime case with $`C=1`$. Note that $`\sqrt{C}t`$ is the proper time measured by the observer. The function has a maximum height of $`1/(\pi \tau _0)`$ and a proper width of $`\tau _0`$. The flat spacetime sampling function is $$h_\eta =\frac{\sqrt{C}\tau _0}{\pi [Ct^2+\tau _0^2]}.$$ (43) It has a maximum height of $`\sqrt{C}/(\pi \tau _0)(0`$ as $`\overline{r}2m`$) and a proper width of $`\tau _0/\sqrt{C}(\mathrm{}`$ as $`\overline{r}2m`$). The inequality (41) becomes $$\widehat{\rho }_C\frac{1}{48\pi C}\left[\frac{C}{\tau _0^2}+\frac{1}{8m^2}\right].$$ (44) Let’s see how the Boulware state near the horizon compares with the above inequality. Since the Boulware state diverges at the event horizon it will not be the quantum state outside a black hole. It is instead the quantum state exterior to a static mass distribution that is larger than its Schwarzschild radius. I will take the mass distribution to be only slightly larger than its Schwarzschild radius (of course, the matter will be under arbitrarily large stresses which is physically unrealistic). For a metric of the form $$ds^2=f(\overline{r})dt^2\frac{1}{f(\overline{r})}d\overline{r}^2$$ (45) the energy density in the Boulware state is given by $$\rho _B=\frac{1}{24\pi }\left[f^{^{\prime \prime }}\frac{(f^{^{}})^2}{4f}\right].$$ (46) Note that this will be the energy density in the $`(r,t)`$ coordinate system since $`\rho =T_t^t`$ is unaffected by changes in the radial coordinate. For $`f=(12m/\overline{r})`$ we have $$\rho _B=\frac{1}{384\pi m^2}\left(1\frac{2m}{\overline{r}}\right)^1$$ (47) near the horizon. This is just the second term on the left hand side of (41). Thus the Boulware state satisfies the quantum inequality (41), as it must. Now consider the cosmological spacetime $$ds^2=t^n\left[dt^2dx^2\right]n>0.$$ (48) The last term on the right hand side of (36) is $$\frac{1}{48\pi C}\left(\frac{_tC}{C}\right)^2=\frac{n^2}{48\pi }t^{(n+2)}$$ (49) which diverges as $`t0`$. Thus as we approach the initial singularity the negative energy densities become unconstrained. Of course, as one approaches the initial singularity quantum gravity effects are expected to become important and the above analysis will break down. ## Conclusion Optimal quantum inequalities were derived for scalar and Dirac fields in two dimensional spacetimes in which the conformal factor is a function of $`x`$ only or of $`t`$ only. For spacetimes with a metric of the form $$ds^2=C(x)(dt^2dx^2)$$ (50) it was shown that the energy density satisfies the optimal bound $$\widehat{\rho }=_{\mathrm{}}^{\mathrm{}}h(t)<T_{tt}>𝑑t\frac{1}{24\pi C}\left[_{\mathrm{}}^{\mathrm{}}\frac{h^{^{}}(t)^2}{h(t)}𝑑t+4C^{\frac{1}{4}}_2^x(C^{\frac{1}{4}})\right]$$ (51) where h(t) is a sampling function that satisfies $`h(t)0`$ and $`_{\mathrm{}}^{\mathrm{}}h(t)𝑑t=1`$. A similar quantum inequality was also found when the conformal factor is a function of time only. These inequalities were then applied to black hole and cosmological spacetimes. It was shown that the bound on the negative energies diverges to minus infinity as the horizon or initial singularity is approached. Thus, neglecting back reaction, the negative energy densities become unconstrained near the horizon or initial singularity. I also showed that the quantum interest conjecture holds for scalar and Dirac fields in two dimensional Minkowski space and that in curved spacetimes this conjecture applies to deviations from the vacuum polarization energy, not to the total energy. ## Acknowledgements I would like to thank Werner Israel, Shinji Mukohyama, Frans Pretorius, and Thomas Roman for many interesting discussions on negative energy densities.
warning/0001/physics0001068.html
ar5iv
text
# Calculation of the Electron Self Energy for Low Nuclear Charge ## Abstract We present a nonperturbative numerical evaluation of the one-photon electron self energy for hydrogenlike ions with low nuclear charge numbers $`Z=1`$ to 5. Our calculation for the $`1S`$ state has a numerical uncertainty of 0.8 Hz for hydrogen and 13 Hz for singly-ionized helium. Resummation and convergence acceleration techniques that reduce the computer time by about three orders of magnitude were employed in the calculation. The numerical results are compared to results based on known terms in the expansion of the self energy in powers of $`Z\alpha `$. Recently, there has been a dramatic increase in the accuracy of experiments that measure the transition frequencies in hydrogen and deuterium . This progress is due in part to the use of frequency chains that bridge the range between optical frequencies and the microwave cesium time standard. The most accurately measured transition is the $`1S`$-$`2S`$ frequency in hydrogen; it has been measured with a relative uncertainty of $`3.4\times 10^{13}`$ or $`840\mathrm{Hz}`$. With trapped hydrogen atoms, it should be feasible to observe the $`1S`$-$`2S`$ frequency with an experimental linewidth that approaches the $`1.3\mathrm{Hz}`$ natural width of the $`2S`$ level . Indeed, it is likely that transitions in hydrogen will eventually be measured with an uncertainty below $`1\mathrm{Hz}`$ . In order for the anticipated improvement in experimental accuracy to provide better values of the fundamental constants or better tests of QED, there must be a corresponding improvement in the accuracy of the theory of the energy levels in hydrogen and deuterium, particularly in the radiative corrections that constitute the Lamb shift. As a step toward a substantial improvement of the theory, we have carried out a numerical calculation of the one-photon self energy of the $`1S`$ state in a Coulomb field for values of the nuclear charge $`Z=1,2,3,4,5`$. This is the first complete calculation of the self energy at low $`Z`$ and provides a result that contributes an uncertainty of about 0.8 Hz in hydrogen and deuterium. This is a decrease in uncertainty of more than three orders of magnitude over previous results. Among all radiative corrections, the largest by several orders of magnitude are the one-photon self energy and vacuum polarization corrections. Of these, the larger and historically most problematic is the self energy. Analytic calculations of the electron self energy at low nuclear charge $`Z`$ have extended over 50 years. The expansion parameter in the analytic calculations is the strength of the external binding field $`Z\alpha `$. This expansion is semi-analytic \[i.e., it is an expansion in powers of $`Z\alpha `$ and $`\mathrm{ln}(Z\alpha )^2`$\]. The leading term was calculated in . It is of the order of $`\alpha (Z\alpha )^4\mathrm{ln}(Z\alpha )^2`$ in units of $`m_\mathrm{e}c^2`$, where $`m_\mathrm{e}`$ is the mass of the electron. In subsequent work higher-order coefficients were evaluated. The analytic results are relevant to low-$`Z`$ systems. For high $`Z`$, the complete one-photon self energy has been calculated without expansion in $`Z\alpha `$ by numerical methods . However, such numerical evaluations at low nuclear charge suffer from severe loss of numerical significance at intermediate stages of the calculation and slow convergence in the summation over angular momenta. As a consequence, the numerical calculations have been confined to higher $`Z`$. Despite these difficulties, the numerical calculations at higher $`Z`$ could be used together with the power-series results to extrapolate to low $`Z`$ with an assumed functional form in order to improve the accuracy of the self energy at low $`Z`$ ; up to the present, this approach has provided the most accurate theoretical prediction for the one-photon self energy of the $`1S`$ state in hydrogen . However, this method is not completely satisfactory. The extrapolation procedure gives a result with an uncertainty of 1.7 kHz, but employs a necessarily incomplete analytic approximation to the higher-order terms. It therefore contains a component of uncertainty that is difficult to reliably assess. Termination of the power series at the order of $`\alpha (Z\alpha )^6`$ leads to an error of 27 kHz. After the inclusion of a result recently obtained in for the logarithmic term of order $`\alpha (Z\alpha )^7\mathrm{ln}(Z\alpha )^2`$ the error is still 13 kHz. A detailed comparison between the analytic and numerical approaches has been inhibited by the lack of accurate numerical data for low nuclear charge. The one-photon problem is especially well suited for such a comparison because five terms in the $`Z\alpha `$ expansion have been checked in independent calculations. The known terms correspond to the coefficients $`A_{41}`$, $`A_{40}`$, $`A_{50}`$, $`A_{62}`$ and $`A_{61}`$ listed below in Eq. (4). The energy shift $`\mathrm{\Delta }E_{\mathrm{SE}}`$ due to the electron self energy is given by $$\mathrm{\Delta }E_{\mathrm{SE}}=\frac{\alpha }{\pi }\frac{(Z\alpha )^4}{n^3}m_\mathrm{e}c^2F(Z\alpha ),$$ (1) where $`n`$ is the principal quantum number. For a particular atomic state, the dimensionless function $`F`$ depends only on one argument, the coupling $`Z\alpha `$. The semi-analytic expansion of $`F(Z\alpha )`$ about $`Z\alpha =0`$ gives rise to the following terms: $`F(Z\alpha )=A_{41}\mathrm{ln}(Z\alpha )^2+A_{40}+(Z\alpha )A_{50}+(Z\alpha )^2`$ (3) $`\times [A_{62}\mathrm{ln}^2(Z\alpha )^2+A_{61}\mathrm{ln}(Z\alpha )^2+G_{\mathrm{SE}}(Z\alpha )],`$ where $`G_{\mathrm{SE}}(Z\alpha )`$ represents the nonperturbative self-energy remainder function. The first index of the $`A`$ coefficients gives the power of $`Z\alpha `$ \[including the $`(Z\alpha )^4`$ prefactor from Eq. (1)\], the second corresponds to the power of the logarithm. For the $`1S`$ ground state, which we investigate in this Letter, the terms $`A_{41}`$ and $`A_{40}`$ were obtained in . The correction term $`A_{50}`$ was found in . The higher-order corrections $`A_{62}`$ and $`A_{61}`$ were evaluated and confirmed in . The results are $`A_{41}`$ $`=`$ $`{\displaystyle \frac{4}{3}},`$ (4) $`A_{40}`$ $`=`$ $`{\displaystyle \frac{10}{9}}{\displaystyle \frac{4}{3}}\mathrm{ln}k_0,`$ (5) $`A_{50}`$ $`=`$ $`2\pi \left({\displaystyle \frac{139}{64}}\mathrm{ln}2\right),`$ (6) $`A_{62}`$ $`=`$ $`1,`$ (7) $`A_{61}`$ $`=`$ $`{\displaystyle \frac{28}{3}}\mathrm{ln}2{\displaystyle \frac{21}{20}}.`$ (8) The Bethe logarithm $`\mathrm{ln}k_0`$ has been evaluated, e.g., in as $`\mathrm{ln}k_0=2.9841285558(3)`$. For our high-accuracy, numerical calculation of $`F(Z\alpha )`$, we divide the calculation into a high- and a low-energy part (see Ref. ). Except for a further separation of the low-energy part into an infrared part and a middle-energy part, which is described in and not discussed further here, we use the same integration contour for the virtual photon energy and basic formulation as in . The numerical evaluation of the radial Green function of the bound electron \[see Eq. (A.16) in \] requires the calculation of the Whittaker function $`W_{\kappa ,\mu }(x)`$ (see , p. 296) over a very wide range of parameters $`\kappa `$, $`\mu `$ and arguments $`x`$. Because of numerical cancellations in subsequent steps of the calculation, the function $`W`$ has to be evaluated to 1 part in $`10^{24}`$. In a problematic intermediate region, which is given approximately by the range $`15<x<250`$, we found that resummation techniques applied to the divergent asymptotic series of the function $`W`$ provide a numerically stable and efficient evaluation scheme. These techniques follow ideas outlined in and are described in detail in . For the acceleration of the slowly convergent angular momentum sum in the high-energy part \[see Eq. (4.3) in \], we use the combined nonlinear-condensation transformation . This transformation consists of two steps: First, we apply the van Wijngaarden condensation transformation to the original series to transform the slowly convergent monotone input series into an alternating series . In the second step, the convergence of the alternating series is accelerated by the $`\delta `$ transformation \[see Eq. (3.14) in \]. The $`\delta `$ transformation acts on the alternating series much more effectively than on the original input series. The highest angular momentum, characterized by the Dirac quantum number $`\kappa `$, included in the present calculation is about $`3500000`$. However, even in these extreme cases, evaluation of less than $`1000`$ terms of the original series is required. As a result, the computer time for the evaluation of the slowly convergent angular momentum expansion is reduced by roughly three orders of magnitude. The convergence acceleration techniques remove the principal numerical difficulties associated with the singularity of the relativistic propagators for nearly equal radial arguments. These singularities are present in all QED effects in bound systems, irrespective of the number of photons involved. It is expected that these techniques could lead to a similar decrease in computer time in the calculation of QED corrections involving more than one photon. In the present calculation, numerical results are obtained for the scaled self-energy function $`F(Z\alpha )`$ for the nuclear charges $`Z=1,2,3,4,5`$ (see Table 1). The value of $`\alpha `$ used in the calculation is $`\alpha _0=1/137.036`$. This is close to the current value from the anomalous magnetic moment of the electron , $`1/\alpha =137.03599958(52).`$The numerical data points are plotted in Fig. 1, together with a graph of the function determined by the analytically known lower-order coefficients listed in Eq. (4). In order to allow for a variation of the fine-structure constant, we repeated the calculation with two more values of $`\alpha `$, which are $`1/\alpha _>=137.0359995\text{and}\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}1}/\alpha _<=137.0360005.`$On the assumption that the main dependence of $`F`$ on $`Z\alpha `$ is represented by the lower-order terms in (4), the change in $`F(Z\alpha )`$ due to the variation in $`\alpha `$ is $`{\displaystyle \frac{F(Z\alpha )}{\alpha }}\delta \alpha =2A_{41}{\displaystyle \frac{\delta \alpha }{\alpha }}+\left[ZA_{50}+\mathrm{O}(\alpha \mathrm{ln}^2\alpha )\right]\delta \alpha `$for a given nuclear charge $`Z`$. Based on this analytic estimate, we expect a variation $`F(Z\alpha _>)F(Z\alpha _0)F(Z\alpha _0)F(Z\alpha _<)9\times 10^9`$for the different values of $`\alpha `$. This variation is in fact observed in our calculation. E.g., for the case $`Z=2`$ we find $`F(2\alpha _<)`$ $`=`$ $`8.528325061(1),`$ (9) $`F(2\alpha _0)`$ $`=`$ $`8.528325052(1)\text{and}`$ (10) $`F(2\alpha _>)`$ $`=`$ $`8.528325043(1).`$ (11) This constitutes an important stability check on the numerics and it confirms that the main dependence of $`F`$ on its argument is indeed given by the lowest-order analytic coefficients $`A_{41}`$ and $`A_{50}`$. In addition to the results for $`F(Z\alpha _0)`$, numerical results for the nonperturbative self-energy remainder function $`G_{\mathrm{SE}}(Z\alpha _0)`$ are also given in Table 1. The results for the remainder function are obtained from the numerical data for $`F(Z\alpha _0)`$ by direct subtraction of the analytically known terms corresponding to the coefficients $`A_{41}`$, $`A_{40}`$, $`A_{50}`$, $`A_{62}`$ and $`A_{61}`$ \[see Eqs. (3,4)\]. Note that because the dependence of $`F`$ on $`Z\alpha `$ is dominated by the subtracted lower-order terms, we have at the current level of accuracy $`G_{\mathrm{SE}}(Z\alpha _<)=G_{\mathrm{SE}}(Z\alpha _0)=G_{\mathrm{SE}}(Z\alpha _>)`$. The numerical uncertainty of our calculaton is $`0.8\times Z^4\mathrm{Hz}`$ in frequency units. A sensitive comparison of numerical and analytic approaches to the self energy can be made by extrapolating the nonperturbative self-energy remainder function $`G_{\mathrm{SE}}(Z\alpha )`$ to the point $`Z\alpha =0`$. It is expected that the function $`G_{\mathrm{SE}}(Z\alpha )`$ approaches a constant in the limit $`Z\alpha 0`$. This constant is referred to as $`G_{\mathrm{SE}}(0)A_{60}`$. In the analytic approach, much attention has been devoted to the coefficient $`A_{60}`$ . The correction has proven to be difficult to evaluate, and analytic work on $`A_{60}`$ has extended over three decades. A step-by-step comparison of the analytic calculations has not been feasible, because the approaches to the problem have differed widely. An additional difficulty is the isolation of terms which contribute in a given order in $`Z\alpha `$, i.e. the isolation of only those terms which contribute to $`A_{60}`$ (and not to any higher-order coefficients). In order to address the question of the consistency of $`A_{60}`$ with our numerical results, we perform an extrapolation of our data to the point $`Z\alpha =0`$. The extrapolation procedure is adapted to the problem at hand. We fit $`G_{\mathrm{SE}}`$ to an assumed functional form which corresponds to $`A_{60}`$, $`A_{71}`$ and $`A_{70}`$ terms, with the coefficients to be determined by the fit. We find that our numerical data is consistent with the calculated value $`A_{60}=\mathrm{30.924\hspace{0.17em}15}(1)`$ . It is difficult to assess the seventh-order logarithmic term $`A_{71}`$, because the extrapolated value for $`A_{71}`$ is very sensitive to possible eighth-order triple and double logarithmic terms, which are unknown. We obtain as an approximate result $`A_{71}=5.5(1.0)`$, and we therefore cannot conclusively confirm the result $`A_{71}=\pi \left({\displaystyle \frac{139}{64}}\mathrm{ln}2\right)=4.65.`$Since our all-order numerical evaluation eliminates the uncertainty due to higher-order terms, we do not pursue this question any further. The numerical data points of the function $`G_{\mathrm{SE}}(Z\alpha )`$ are plotted in Fig. 2 together with the value $`G_{\mathrm{SE}}(0)=A_{60}=\mathrm{30.924\hspace{0.17em}15}(1)`$. For a determination of the Lamb shift, the dependence of $`G_{\mathrm{SE}}`$ on the reduced mass $`m_\mathrm{r}`$ of the system has to be restored. In general, the coefficients in the analytic expansion (3) acquire a factor $`(m_\mathrm{r}/m_\mathrm{e})^3`$, because of the scaling of the wave function. Terms associated with the anomalous magnetic moment are proportional to $`(m_\mathrm{r}/m_\mathrm{e})^2`$ . The nonperturbative remainder function $`G_{\mathrm{SE}}`$ is assumed to be approximately proportional to $`(m_\mathrm{r}/m_\mathrm{e})^3`$, but this has not been proved rigorously. Work is currently in progress to address this question . We conclude with a brief summary of the results of this Letter. (i) We have obtained accurate numerical results for the self energy at low nuclear charge. Previously, severe numerical cancellations have been a problem for these evaluations. (ii) For a particular example, we have addressed the question of how well semi-analytic expansions represent all-order results at low nuclear charge. Our numerical data is consistent with the value $`A_{60}=\mathrm{30.924\hspace{0.17em}15}(1)`$ . (iii) Numerical techniques have been developed that reduce the computer time for the problem by about three orders of magnitude. The calculation presented here is of importance for the interpretation of measurements in hydrogen, deuterium and singly-ionized helium and for the improvement of the Rydberg constant, because of recent and projected progress in accuracy. In the determination of the Rydberg constant, uncertainty due to the experimentally determined proton radius can be eliminated by comparing the frequencies of more than one transition . We have shown that an all-order calculation can provide the required accuracy if suitable numerical methods are used. The authors acknowledge helpful discussions with E.J. Weniger. U.D.J. gratefully acknowledges helpful conversations with J. Baker, J. Conlon, J. Devaney and J. Sims, and support by the Deutsche Forschungsgemeinschaft (contract no. SO333/1-2) and the Deutscher Akademischer Austauschdienst. P.J.M. is grateful to Rebecca Ghent who participated in earlier exploratory work on this calculation, and he acknowledges continued support by the Alexander-von-Humboldt Foundation. G.S. acknowledges continued support by the Gesellschaft für Schwerionenforschung and the Deutsche Forschungsgemeinschaft.
warning/0001/math0001105.html
ar5iv
text
# Lefschetz numbers of iterates of the monodromy and truncated arcs ## 1. Introduction Let $`X`$ be a smooth complex algebraic variety and let $`f:X𝐂`$ be a non constant morphism of complex algebraic varieties. We fix a smooth metric on $`X`$. Let $`x`$ be a point of $`f^1(0)`$. We set $`X_{\epsilon ,\eta }^\times :=B(x,\epsilon )f^1(D_\eta ^\times )`$, with $`B(x,\epsilon )`$ the open ball of radius $`\epsilon `$ centered at $`x`$ and $`D_\eta ^\times =D_\eta \{0\}`$, with $`D_\eta `$ the open disk of radius $`\eta `$ centered at $`0`$. For $`0<\eta \epsilon 1`$, the restriction of $`f`$ to $`X_{\epsilon ,\eta }^\times `$ is a locally trivial fibration - called the Milnor fibration - onto $`D_\eta ^\times `$ with fiber $`F_x`$, the Milnor fibre at $`x`$. The action of a characteristic homeomorphism of this fibration on cohomology gives rise to the monodromy operator $$M:H^{}(F_x,𝐐)H^{}(F_x,𝐐).$$ For any natural number $`n`$, we consider the Lefschetz number $$\mathrm{\Lambda }(M^n):=\underset{q0}{}(1)^q\mathrm{Trace}[M^n,H^q(F_x,𝐐)],$$ of the $`n`$-th iterate of $`M`$. These numbers are related to the monodromy zeta function $$Z(t):=\underset{q0}{}[\mathrm{det}(\mathrm{Id}tM,H^q(F_x,𝐐))]^{(1)^q}$$ as follows: if one writes $$\mathrm{\Lambda }(M^n)=\underset{i|n}{}s_i,$$ for $`n1`$, then $$Z(t)=\underset{i1}{}(1t^i)^{s_i/i}.$$ In the paper , A’Campo gives a formula for the Lefschetz numbers $`\mathrm{\Lambda }(M^n)`$ in terms of a resolution of the singularities of $`f=0`$. The aim of the present note is to give a formula for $`\mathrm{\Lambda }(M^n)`$ directly in terms of geometric objects associated to the function $`f`$. This formula will involve truncated arcs on $`X`$. Let us recall from , that there is a $`𝐂`$-scheme $`(X)`$, the space of formal arcs on $`X`$, whose set of $`𝐂`$-rational points $`(X)(𝐂)`$ is naturally in bijection with $`X(𝐂[[t]])`$. Similarly, for $`n0`$, we can consider the space $`_n(X)`$ of arcs modulo $`t^{n+1}`$: a $`𝐂`$-rational point of $`_n(X)`$ corresponds to a $`𝐂[t]/t^{n+1}𝐂[t]`$-rational point on $`X`$. The space $`_n(X)`$ is canonically endowed with the structure of a complex algebraic variety. For instance when $`X`$ is the affine space $`𝐀_𝐂^m`$, a $`𝐂`$-rational point of $`(X)`$ is just an $`m`$-tuple of power series in the variable $`t`$ with coefficients in $`𝐂`$, while $`_n(X)(𝐂)`$ is the set of $`m`$-tuples of complex polynomials of degree $`n`$ in the variable $`t`$. Furthermore, there is a natural morphism $$\pi _n:(X)_n(X)$$ which corresponds to truncation on $`𝐂`$-rational points. Since $`X`$ is assumed to be smooth, the morphism $`\pi _n`$ is surjective (also for $`𝐂`$-rational points). In what follows, we identify $`(X)`$ with its set of $`𝐂`$-rational points, and similarly for $`_n(X)`$. With a “complex algebraic variety”, we mean the set of $`𝐂`$-rational points of a separated reduced scheme of finite type over $`𝐂`$, not necessarly irreducible. Now, we consider the set $`𝒳_n`$ of points $`\phi `$ in $`_n(X)`$ such that $`\pi _0(\phi )=x`$ and such that the $`t`$-valuation of $`f(\phi )`$ is exactly $`n`$. This subset $`𝒳_n`$ is a locally closed subvariety of $`_n(X)`$ and we may consider the morphism $$\overline{f}_n:𝒳_n𝐂^\times ,$$ sending a point $`\phi `$ in $`𝒳_n`$ to the coefficient of $`t^n`$ in $`f(\phi )`$. There is a natural action of $`𝐂^\times `$ on $`𝒳_n`$ given by $`a\phi (t)=\phi (at)`$. Since $`\overline{f}_n(a\phi )=a^n\overline{f}_n(\phi )`$, it follows that $`\overline{f}_n`$ is a locally trivial fibration and has a geometric monodromy of order $`n`$. We denote by $`𝒳_{n,1}`$ the fiber $`\overline{f}_n^1(1)`$. The group $`\mu _n`$ of $`n`$-th roots of unity in $`𝐂`$ acts naturally on $`𝒳_{n,1}`$, by restriction of the $`𝐂^\times `$-action on $`𝒳_n`$. We can now state the main result of this note. ###### 1.1 Theorem. For every integer $`n1`$, the Lefschetz number $`\mathrm{\Lambda }(M^n)`$ is equal to $`\chi (𝒳_{n,1})`$, the Euler characteristic of $`𝒳_{n,1}`$. In fact, we shall deduce Theorem 1.1 from a more general result, Theorem 2.4, where we give a formula for the class of $`𝒳_{n,1}`$ in the ring $`_{\mathrm{loc}}`$ obtained by localisation of the class of the affine line in the Grothendieck ring of complex algebraic varieties, whose definition is recalled at the beginning of section 2. Actually, it would be also possible to prove Theorem 1.1 as an easy consequence of Theorem 2.2.1 in , by a reasonning involving motives and Hodge polynomials. In fact, Theorem 2.4 may be further generalized to take in account the monodromy action. This is done in Theorem 2.10, once we introduced the “monodromic” Grothendieck ring $`_{\mathrm{loc}}^{\mathrm{mon}}`$. We denote by $`T_n`$ the monodromy operator $$T_n:H^{}(𝒳_{n,1},𝐐)H^{}(𝒳_{n,1},𝐐)$$ of the locally trivial fibration $`\overline{f}_n:𝒳_n𝐂^\times `$. Note that $`T_n`$ is induced by the automorphism $`\phi (t)\phi (e^{2\pi i/n}t)`$ of $`𝒳_{n,1}`$. For any $`d`$ in $`𝐍`$, we denote the Lefschetz number of $`T_n^d`$ by $$\mathrm{\Lambda }(T_n^d):=\underset{q0}{}(1)^q\mathrm{Trace}[T_n^d,H^q(𝒳_{n,1},𝐐)].$$ Since $`T_n^n`$ is the identity, we have $`\mathrm{\Lambda }(T_n^d)=\mathrm{\Lambda }(T_n^{\mathrm{gcd}(d,n)})`$. Hence to know the $`\mathrm{\Lambda }(T_n^d)`$’s for every $`d`$ in $`𝐍`$, it is enough to know them for every $`d`$ dividing $`n`$. This information is provided by the following result: ###### 1.2 Theorem. If $`n1`$ and $`d`$ divides $`n`$, then $`\mathrm{\Lambda }(T_n^d)=\mathrm{\Lambda }(M^d)`$. Finally, in section 3, we extend Theorem 2.10 to the case of quasi-projective varieties over a field of characteristic zero. This enables us to construct what we believe to be the “virtual motivic incarnation” of the the Milnor fibre at $`x`$ in $`_{k,\mathrm{loc}}^{\mathrm{mon}}`$, the analogue over $`k`$ of the ring $`_{\mathrm{loc}}^{\mathrm{mon}}`$. Taking Euler characteristic with values into virtual Chow motives, we apply this to settle an issue that remained open in . The first author is endebted to Eduard Looijenga for a conversation in November 1998 which influenced the present paper. The second author would like to thank Paul Seidel for interesting discussions. ## 2. Calculation of a motivic volume and proof of the main results To state Theorem 2.4, we shall start with some reminders from motivic integration as developped in and . ### 2.1. We denote by $``$ the abelian group generated by symbols $`[S]`$, for $`S`$ a complex algebraic variety, with the relations $`[S]=[S^{}]`$ if $`S`$ and $`S^{}`$ are isomorphic and $`[S]=[S^{}]+[SS^{}]`$ if $`S^{}`$ is Zariski closed in $`S`$. There is a natural ring structure on $``$, the product being induced by the cartesian product of varieties, and to any constructible set $`S`$ in some complex algebraic variety one naturally associates a class $`[S]`$ in $``$. We denote by $`𝐋`$ the class of the affine line in $``$ and we denote by $`_{\mathrm{loc}}`$ the localisation $`_{\mathrm{loc}}:=[𝐋^1]`$. Let $`X`$ be a smooth complex algebraic variety of pure dimension $`m`$. We call a subset $`A`$ of $`(X)`$ cylindrical at level $`n`$ if $`A=\pi _n^1(C)`$, with $`C`$ a constructible subset of $`_n(X)`$. We say that $`A`$ is cylindrical if it is cylindrical at some level $`n`$. Let $`X`$, $`Y`$ and $`F`$ be complex algebraic varieties, and let $`A`$, resp. $`B`$, be a constructible subset of $`X`$, resp. $`Y`$. We say that a map $`\pi :AB`$ is a piecewise morphism if there exists a finite partition of the domain of $`\pi `$ into locally closed subvarieties of $`X`$ such that the restriction of $`\pi `$ to any of these subvarieties is a morphism of varieties. We say that a map $`\pi :AB`$ is a piecewise trivial fibration with fiber $`F`$, if there exists a finite partition of $`B`$ in subsets $`S`$ which are locally closed subvarieties of $`Y`$ such that $`\pi ^1(S)`$ is a locally closed subvariety of $`X`$ and isomorphic, as a complex algebraic variety, to $`S\times F`$, with $`\pi `$ corresponding under the isomorphism to the projection $`S\times FS`$. We say that the map $`\pi `$ is a piecewise trivial fibration over some constructible subset $`C`$ of $`B`$, if the restriction of $`\pi `$ to $`\pi ^1(C)`$ is a piecewise trivial fibration onto $`C`$. We call a cylindrical subset $`A`$ of $`(X)`$ stable at level $`n𝐍`$ if $`A`$ is cylindrical at level $`n`$ and $`\pi _{k+1}((X))\pi _k((X))`$ is a piecewise trivial fibration over $`\pi _k(A)`$ with fiber $`𝐀_𝐂^m`$, for all $`kn`$. We call $`A`$ stable if it is stable at some level $`n`$. Since $`X`$ is smooth, any cylindrical subset $`A`$ of $`(X)`$ is stable (at the same level), by Lemma 4.1 of . Denote by $`𝐂_0`$ the family of stable cylindrical subsets of $`(X)`$. Clearly there exists a unique additive measure $$\stackrel{~}{\mu }:𝐂_0_{\mathrm{loc}}$$ satisfying $$\stackrel{~}{\mu }(A)=[\pi _n(A)]𝐋^{(n+1)m}$$ when $`A𝐂_0`$ is stable at level $`n`$. In particular, the relation $$[𝒳_{n,1}]=\stackrel{~}{\mu }(𝒵_{n,1})𝐋^{(n+1)m}$$ holds in $`_{\mathrm{loc}}`$, where $`𝒵_{n,1}`$ is the set of points $`\phi `$ in $`(X)`$ such that $`\pi _0(\phi )=x`$, such that the $`t`$-valuation of $`f(\phi )`$ is exactly $`n`$, and such that the coefficient of $`t^n`$ in $`f(\phi )`$ is equal to 1. The following geometric lemma, which is a special case of Lemma 3.4 in , will play a crucial role in the proof of Theorem 2.4. ###### 2.2 Lemma. Let $`X`$ and $`Y`$ be connected smooth complex algebraic varieties and let $`h:YX`$ be a birational morphism. For $`e`$ in $`𝐍`$, let $`\mathrm{\Delta }_e`$ be the subset of $`(Y)`$ defined by $$\mathrm{\Delta }_e:=\{\phi Y(𝐂[[t]])|\mathrm{ord}_t\mathrm{det}\mathrm{Jac}_h(\phi )=e\},$$ where $`\mathrm{Jac}_h(\phi )`$ is the jacobian of $`h`$ at $`\phi `$. For $`k`$ in $`𝐍`$, let $`h_k:_k(Y)_k(X)`$ be the morphism induced by $`h`$, and let $`\mathrm{\Delta }_{e,k}`$ be the image of $`\mathrm{\Delta }_e`$ in $`_k(Y)`$. If $`k2e`$, the following holds. 1. The constructible subset $`\mathrm{\Delta }_{e,k}`$ of $`_k(Y)`$ is a union of fibers of $`h_k`$. 2. The restriction of $`h_k`$ to $`\mathrm{\Delta }_{e,k}`$ is a piecewise trivial fibration with fiber $`𝐀_𝐂^e`$ onto its image. ### 2.3. Now we shall use Lemma 2.2 to compute $`\stackrel{~}{\mu }(𝒵_{n,1})`$ on a resolution of $`f`$. Let $`D`$ be the divisor defined by $`f=0`$ in $`X`$. Let $`(Y,h)`$ be a resolution of $`f`$. By this, we mean that $`Y`$ is a smooth and connected complex algebraic variety, $`h:YX`$ is proper, that the restriction $`h:Yh^1(D)XD`$ is an isomorphism, and that $`(h^1(D))_{\mathrm{red}}`$ has only normal crossings as a subvariety of $`Y`$. Furthermore, we choose $`h`$ in such a way that $`(h^1(x))_{\mathrm{red}}`$ is a union of irreducible (smooth) components of $`(h^1(D))_{\mathrm{red}}`$, which we shall denote by $`E_i`$, $`iJ`$. For each $`iJ`$, denote by $`N_i`$ the multiplicity of $`E_i`$ in the divisor of $`fh`$ on $`Y`$, and by $`\nu _i1`$ the multiplicity of $`E_i`$ in the divisor of $`h^{}dx`$, where $`dx`$ is a local non vanishing volume form at $`x`$, i.e. a local generator of the sheaf of differential forms of maximal degree at $`x`$. For $`iJ`$ and $`IJ`$, we consider the varieties $`E_i^{}:=E_i_{ji}E_j`$, $`E_I:=_{iI}E_i`$, and $`E_I^{}:=E_I_{jJI}E_j`$. We shall also set $`m_I=\mathrm{gcd}(N_i)_{iI}`$. We introduce an unramified Galois cover $`\stackrel{~}{E}_I^{}`$ of $`E_I^{}`$, with Galois group $`\mu _{m_I}`$, as follows. Let $`U`$ be an affine Zariski open subset of $`Y`$, such that, on $`U`$, $`fh=uv^{m_I}`$, with $`u`$ a unit on $`U`$ and $`v`$ a morphism form $`U`$ to $`𝐀_𝐂^1`$. Then the restriction of $`\stackrel{~}{E}_I^{}`$ above $`E_I^{}U`$, denoted by $`\stackrel{~}{E}_I^{}U`$, is defined as $$\left\{(z,y)𝐀_𝐂^1\times (E_I^{}U)\right|z^{m_I}=u^1\}.$$ Note that $`E_I^{}`$ can be covered by such affine open subsets $`U`$ of $`Y`$. Gluing together the covers $`\stackrel{~}{E}_I^{}U`$, in the obvious way (cf. the proof of Lemma 3.2.2 in ), we obtain the cover $`\stackrel{~}{E}_I^{}`$ of $`E_I^{}`$ which has a natural $`\mu _{m_I}`$-action (obtained by multiplying the $`z`$-coordinate with the elements of $`\mu _{m_I}`$). ###### 2.4 Theorem. With the previous notations, the following relation holds in $`_{\mathrm{loc}}`$: (2.4.1) $$[𝒳_{n,1}]=𝐋^{nm}\underset{\genfrac{}{}{0pt}{}{IJ}{I\mathrm{}}}{}(𝐋1)^{|I|1}[\stackrel{~}{E}_I^{}]\left(\underset{\genfrac{}{}{0pt}{}{k_i1,iI}{{\scriptscriptstyle k_iN_i}=n}}{}𝐋^{_{iI}k_i\nu _i}\right).$$ ###### Proof. Let $`\stackrel{~}{𝒵}_{n,1}`$ be the preimage of $`𝒵_{n,1}`$ in $`(Y)`$. Remark that, the morphism $`h`$ being proper, the induced function $`h_{}:\stackrel{~}{𝒵}_{n,1}𝒵_{n,1}`$ is bijective. Now, for $`e0`$, define $`\stackrel{~}{𝒵}_{n,1,e}`$ as the set of points $`\phi `$ in $`\stackrel{~}{𝒵}_{n,1}`$ such that $`\mathrm{ord}_t\mathrm{det}\mathrm{Jac}_h(\phi )=e`$. ###### 2.5 Lemma. Let $`I`$ be a non empty subset of $`J`$. Let $`U`$ be an affine Zariski open subset of $`Y`$, such that, on $`U`$, $`fh=u_{iI}y_i^{N_i}`$ and $`\mathrm{det}\mathrm{Jac}_h=v_{iI}y_i^{\nu _i1}`$, with $`u`$ and $`v`$ units on $`U`$ and $`y_i`$ a regular function on $`U`$ with divisor $`E_iU`$. Let $`k_i`$, $`iI`$, be natural numbers with $`_{iI}k_iN_i=n`$, $`k_i1`$. Let $`𝒰_{(k_i)}`$ be the set of points $`\phi `$ in $`_n(U)`$ such that $`\mathrm{ord}_ty_i(\phi )=k_i`$, for $`iI`$, and $`\overline{f}_n(h_n(\phi ))=1`$, where $`h_n:_n(Y)_n(X)`$ is the morphism induced by $`h`$. Then $$[𝒰_{(k_i)}]=(𝐋1)^{|I|1}[\stackrel{~}{E}_I^{}U]𝐋^{mn_{iI}k_i}$$ in $`_{\mathrm{loc}}`$. ###### Proof. Note that the projection $`_n(U)U`$ maps $`𝒰_{(k_i)}`$ into $`E_I^{}U`$. Thus we may assume that $`E_I^{}U`$ is not empty, and - using additivity - that there exist $`m|I|`$ functions on $`U`$ which, together with the functions $`(y_i)_{iI}`$, induce an étale map $`U𝐀_𝐂^m`$. By Lemma 4.2 of (with $`n=e=0`$), this map induces an isomorphism $`_n(U)U\times _{𝐀_𝐂^m}_n(𝐀_𝐂^m)`$. It follows now from the very definitions that $$[𝒰_{(k_i)}]=[W_I]𝐋^{mn_{iI}k_i},$$ with $$W_I:=\left\{(z_i,y)(𝐂^\times )^{|I|}\times (E_I^{}U)\right|\underset{iI}{}z_i^{N_i}u=1\},$$ and we have $$[W_I]=(𝐋1)^{|I|1}[\stackrel{~}{E}_I^{}U].$$ To verify the last equality, consider an automorphism $`z(z^{a_i})_{iI}`$ of $`(𝐂^\times )^{|I|}`$, with $`(a_i)_{iI}`$ a basis for $`𝐙^{|I|}`$ and $`a_1=(N_i/m_I)_{iI}`$. ∎ By covering $`Y`$ with affine Zariski open subsets $`U`$ verifying the assumptions in Lemma 2.5, one sees that all the subsets $`\stackrel{~}{𝒵}_{n,1,e}`$ of $`(Y)`$ are cylindrical and that there exists $`e_0`$ such that $`\stackrel{~}{𝒵}_{n,1,e}`$ is empty for $`e>e_0`$. Now set $`𝒵_{n,1,e}=h_{}(\stackrel{~}{𝒵}_{n,1,e})`$. Since $`\pi _kh_{}=h_k\pi _k`$, with the notation of Lemma 2.2, it follows from assertion a) of Lemma 2.2 that the subsets $`𝒵_{n,1,e}`$ of $`(X)`$ are cylindrical. Since $`𝒵_{n,1}`$ is equal to the disjoint union of the subsets $`𝒵_{n,1,e}`$ for $`ee_0`$, we have $$\stackrel{~}{\mu }(𝒵_{n,1})=\underset{ee_0}{}\stackrel{~}{\mu }(𝒵_{n,1,e}).$$ Now it follows from Lemma 2.2 that $$\stackrel{~}{\mu }(𝒵_{n,1,e})=𝐋^e\stackrel{~}{\mu }(\stackrel{~}{𝒵}_{n,1,e}).$$ By using Lemma 2.5, one gets, for every $`U`$ as in 2.5, $$\stackrel{~}{\mu }(\stackrel{~}{𝒵}_{n,1,e}(U))=𝐋^m\underset{\genfrac{}{}{0pt}{}{IJ}{I\mathrm{}}}{}(𝐋1)^{|I|1}[\stackrel{~}{E}_I^{}U]\underset{\genfrac{}{}{0pt}{}{k_i1,iI}{{\scriptscriptstyle k_iN_i}=n,{\scriptscriptstyle k_i(\nu _i1)}=e}}{}𝐋^{_{iI}k_i},$$ and the result follows by additivity of $`\stackrel{~}{\mu }`$. ∎ ### 2.6. Proof of Theorem 1.1 We use a resolution $`(Y,h)`$ of $`f`$ satisfying the conditions in 2.3. We shall view the Euler characteristic of complex constructible sets as a ring morphism $`\chi :𝐙`$. Since $`\chi (𝐋)=1`$, this morphism extends uniquely to a ring morphism $`\chi :_{\mathrm{loc}}𝐙`$. Since $`\chi ((𝐋1)^{|I|1})=0`$ when $`|I|>1`$, it follows from Theorem 2.4 that $$\chi (𝒳_{n,1})=\underset{N_i|n}{}N_i\chi (E_i^{}).$$ The result follows since $`\mathrm{\Lambda }(M^n)`$ is equal to the right hand side of the previous formula by A’Campo’s formula for $`\mathrm{\Lambda }(M^n)`$ given in . ∎ ###### 2.7 Remark. As observed by Paul Seidel, Theorem 1.1 bears some similarity with properties of Floer homology for a symplectic automorphism (see, e.g., ). ###### 2.8 Remark. Of course, it is also possible to prove Theorem 1.1 directly, without considering the ring $`_{\mathrm{loc}}`$, by using Lemma 2.2 and working with Euler characteristics all the way. ### 2.9. To take in account the monodromy action we introduce a ring $`^{\mathrm{mon}}`$. As an abelian group $`^{\mathrm{mon}}`$ is generated by symbols $`[S,\tau ]`$, with $`S`$ a complex algebraic variety and $`\tau `$ an automorphism of $`S`$. The relations are $`[S,\tau ]=[S^{},\tau ^{}]`$ if the pairs $`(S,\tau )`$ and $`(S^{},\tau ^{})`$ are isomorphic, $`[S,\tau ]=[S^{},\tau _{|S^{}}]+[SS^{},\tau _{|SS^{}}]`$ when $`S^{}`$ is Zariski closed in $`S`$ and stable under $`\tau `$, and $`[S\times 𝐀_𝐂^n,\sigma ]=[S\times 𝐀_𝐂^n,\sigma ^{}]`$ whenever $`\sigma `$ and $`\sigma ^{}`$ are liftings<sup>1</sup><sup>1</sup>1meaning that $`\tau p=p\sigma =p\sigma ^{}`$, where $`p`$ is the projection of $`S\times 𝐀_𝐂^n`$ onto $`S`$. of a same automorphism $`\tau `$ of $`S`$. There is a natural ring structure on $`^{\mathrm{mon}}`$, induced by the cartesian product of varieties. We denote by $`𝐋`$ the class of $`(𝐀_𝐂^1,\mathrm{id})`$ in $`^{\mathrm{mon}}`$ and we denote by $`_{\mathrm{loc}}^{\mathrm{mon}}`$ the localisation $`_{\mathrm{loc}}^{\mathrm{mon}}:=^{\mathrm{mon}}[𝐋^1]`$. We will often write $`[S]`$ instead of $`[S,\tau ]`$ when it is clear from the the context what $`\tau `$ is. For example we write $`[𝒳_{n,1}]`$ and $`[\stackrel{~}{E}_I^{}]`$, the automorphism being the obvious one induced by multiplication by $`e^{2\pi i/n}`$, resp. $`e^{2\pi i/m_I}`$. Let $`T`$ be an automorphism of the scheme $`(X)`$ which permutes the fibers of $`\pi _n`$ for every $`n`$. Denote by $`𝐂_{0,T}`$ the family of stable cylindrical subsets $`A`$ of $`(X)`$ with $`T(A)=A`$. Clearly there exists a unique additive measure $$\stackrel{~}{\mu }^{\mathrm{mon}}:𝐂_{0,T}_{\mathrm{loc}}^{\mathrm{mon}}$$ satisfying $$\stackrel{~}{\mu }^{\mathrm{mon}}(A)=[\pi _n(A)]𝐋^{(n+1)m}$$ when $`A𝐂_{0,T}`$ is stable at level $`n`$. ###### 2.10 Theorem. Relation (2.4.1) holds in $`_{\mathrm{loc}}^{\mathrm{mon}}`$. ###### Proof. The proof of Theorem 2.4 carries over literally to the monodromic situation. ∎ ### 2.11. For any complex algebraic variety with an action of a finite abelian group $`G`$, and for any character $`\alpha `$ of $`G`$, we denote by $`H^{}(X,𝐂)_\alpha `$ the part of $`H^{}(X,𝐂)`$ on which $`G`$ acts by multiplication by $`\alpha `$, and we set $$\chi (X,\alpha ):=\underset{q0}{}(1)^q\mathrm{dim}H^q(X,𝐂)_\alpha .$$ ### 2.12. Proof of Theorem 1.2 The monodromy zeta function $`Z(t)`$ of the Milnor fibration being equal to $`_{iJ}(1t^{N_i})^{\chi (E_i^{})}`$ by , Theorem 1.2 is equivalent to the assertion that the monodromy zeta function of the fibration $`\overline{f}_n`$ is equal to $$\underset{\genfrac{}{}{0pt}{}{iJ}{N_i|n}}{}(1t^{N_i})^{\chi (E_i^{})}.$$ But this assertion is equivalent to the validity of the equality $$\chi (𝒳_{n,1},\alpha )=\underset{\genfrac{}{}{0pt}{}{iJ}{\mathrm{ord}(\alpha )|N_i|n}}{}\chi (E_i^{}),$$ for every character $`\alpha `$ of $`\mu _n`$, which is a direct consequence of Theorem 2.10. ∎ ###### 2.13 Remark. In fact, Theorem 2.10 still remains valid if one changes the definition of $`_{\mathrm{loc}}^{\mathrm{mon}}`$ in 2.9 by imposing the relation $`[S\times 𝐀_𝐂^n,\sigma ]=[S\times 𝐀_𝐂^n,\sigma ^{}]`$ only when $`\sigma `$ and $`\sigma ^{}`$ are cartesian products which coincide on $`S`$. To verify this claim one needs a straighforward refinement of Lemma 4.1 of and of Lemma 2.2. ## 3. Some further results ### 3.1. There is an algebraic analogue of Theorem 2.10 over an arbitrary field $`k`$ of characteristic zero. To state it, we first define the ring $`_{k,\mathrm{loc}}^{\mathrm{mon}}`$, which is the algebraic analogue of the ring $`_{\mathrm{loc}}^{\mathrm{mon}}`$. For $`n1`$ an integer, we denote by $`\mu _n`$ the group *scheme* over $`k`$ of $`n`$-th roots of unity. Note that it is not assumed that all geometric points of $`\mu _n`$ are rational over $`k`$. By an action of $`\mu _n`$ on a quasi-projective scheme over $`k`$, we mean an action in the sense of group schemes and schemes over $`k`$. Set $`\widehat{\mu }:=\begin{array}{c}\text{lim}\hfill \\ \hfill \end{array}_n\mu _n`$. By an action of $`\widehat{\mu }`$ on a quasi-projective scheme over $`k`$, we mean an action which factors through a suitable $`\mu _n`$. We define $`_{k,\mathrm{loc}}^{\mathrm{mon}}`$ in the same way as $`_{\mathrm{loc}}^{\mathrm{mon}}`$, working now with pairs consisting of a quasi-projective scheme over $`k`$ and a $`\widehat{\mu }`$-action on it. Let $`X`$ be a smooth algebraic variety over $`k`$ and let $`f:X𝐀_k^1`$ be a non constant morphism. Similarly as in the complex case one defines the arc space $`(X)`$ and the quasi-projective variety $`𝒳_{n,1}`$ with a natural $`\widehat{\mu }`$-action. One defines also similarly as before resolutions $`(Y,h)`$ of $`f`$ and the varieties $`\stackrel{~}{E}_I^{}`$ with $`\widehat{\mu }`$-action. The proof of Theorem 2.10 carries over to the algebraic case to give the following: ###### 3.2 Theorem. Let $`X`$ be a smooth algebraic variety over $`k`$ and let $`f:X𝐀_k^1`$ be a non constant morphism. Let $`(Y,h)`$ be a resolution of $`f`$. With the previous notations, relation (2.4.1) holds in $`_{k,\mathrm{loc}}^{\mathrm{mon}}`$. ### 3.3. Consider the power series $$P(T):=\underset{n1}{}[𝒳_{n,1}]𝐋^{nm}T^n$$ in the variable $`T`$ over the ring $`_{k,\mathrm{loc}}^{\mathrm{mon}}`$. It follows directly from Theorem 3.2 that $`P(T)`$ is a rational power series and that its “limit for $`T\mathrm{}`$” (cf. section 4 of ) is equal to $`𝒮`$, where (3.3.1) $$𝒮:=\underset{\mathrm{}IJ}{}(1𝐋)^{|I|1}[\stackrel{~}{E}_I^{}].$$ In particular it follows that the right hand side of (3.3.1) is independent of the resolution $`(Y,h)`$, as an element of $`_{k,\mathrm{loc}}^{\mathrm{mon}}`$. As we shall explain in 3.5, we believe that $`𝒮`$ is the the “virtual motivic incarnation” of the Milnor fibre at $`x`$. ### 3.4. We denote by $`E`$ the smallest subfield of $`𝐂`$ containing all roots of unity. Assume $`E`$ is contained in $`k`$. To any quasi-projective variety $`X`$ over $`k`$ with a $`\widehat{\mu }`$-action, and to any character $`\alpha `$ of $`\widehat{\mu }`$ of finite order, we associate, as in Theorem 1.3.1 of (see also ), an element $`\chi _{\mathrm{mot},c}(X,\alpha )`$ of the Grothendieck group $`K_0(\mathrm{Mot}_{k,E})`$ of the category $`\mathrm{Mot}_{k,E}`$ of Chow motives over $`k`$ with coefficients in $`E`$. One can check that $`\chi _{\mathrm{mot},c}(\mathrm{\_},\alpha )`$ factorizes through a ring morphism $`_{k,\mathrm{loc}}^{\mathrm{mon}}K_0(\mathrm{Mot}_{k,E})`$. One verifies that $`\chi _{\mathrm{mot},c}(\mathrm{\_},\alpha )`$ respects the last relation in the definition of $`_{k,\mathrm{loc}}^{\mathrm{mon}}`$ by an argument similar to the proof of Proposition 2.6 in . We still write $`𝐋`$ for $`\chi _{\mathrm{mot},c}(𝐀_k^1,1)`$. When $`k=𝐂`$, the topological Euler characteristic of $`\chi _{\mathrm{mot},c}(X,\alpha )`$ is equal to $`\chi (X,\alpha )`$, with the notation of 2.11. ### 3.5. Let $`\alpha `$ be a character of $`\widehat{\mu }`$ of order $`d`$ and set $`S_\alpha :=\chi _{\mathrm{mot},c}(𝒮,\alpha ^1)`$. It follows from the definition that $$S_\alpha =\underset{\genfrac{}{}{0pt}{}{\mathrm{}IJ}{d|m_I}}{}(1𝐋)^{|I|1}\chi _{\mathrm{mot},c}(\stackrel{~}{E}_I^{},\alpha ^1).$$ This element $`S_\alpha `$ plays a key role in section 4 of , where it was proved that modulo $`(𝐋1)`$-torsion $`S_\alpha `$ does not depend on the chosen resolution $`(Y,h)`$ of $`f`$. However it follows now from the above considerations that $`S_\alpha `$, as an element of $`K_0(\mathrm{Mot}_{k,E})`$, does not depend on the chosen resolution. As explained in , we believe that $`S_\alpha `$ is the “virtual motivic incarnation” of the $`\alpha `$-isotypic part of the Milnor fibre. It was shown in that this is indeed true at the level of $`𝐂`$-Hodge realizations.
warning/0001/gr-qc0001060.html
ar5iv
text
# Locally Anisotropic Kinetic Processes and Thermodynamics in Curved Spaces ## 1 Introduction The experimental data on anisotropies in the microwave background radiation (see, for instance, Ref. and ) and modern physical theories support the idea that in the very beginning the Universe has to be described as an anisotropic and higher dimension spacetime. It is of interest the investigation of higher dimension generalized Kaluza–Klein and string anisotropic cosmologies. Therefore, in order to understand the initial dynamical behavior of an anisotropic universe, in particular, to study possible mechanisms of anisotropic inflation connected with higher dimensions we have to know how we can compute the parameters (transport coefficients, damping terms and viscosity coefficients) that characterize the cosmological fluids in spacetimes with generic anisotropy. The first relativistic macroscopic thermodynamic theories have been proposed in Refs. and . The further developments and applications in gravitational physics, astrophysics and cosmology are connected with papers . The Israel’s approach to a microscopic Boltzmann like kinetic theory for relativistic gases makes possible to express the transport coefficients via the differential cross sections of the fluid’s particles. Here we also note the Chernikov’s results on Boltzmann equations with collision integral on (pseudo) Riemannian spaces . A complete formulation of relativistic kinetics is contained in the monograph by de Groot, van Leeuwen and van Weert. We also emphasize the Vlasov’s monograph were an attempt to statistical motivation of kinetic and thermodynamic theory on phase spaces enabled with Finsler like metrics and connection structures was proposed. Recent developments and applications of kinetics and nonequilibrium thermodynamics could be found in . The extension of the four dimensional considerations to higher dimensions is due to and . The generalization of kinetic and thermodynamic equations and formulas to curved spaces is not a trivial task. In order to consider flows of particles with noninteger spins, interactions with gauge fields, and various anisotropic processes of microscopic or macroscopic nature it is necessary a reformulation of the kinetic theory in curved spacetimes by using the Cartan’s moving frame method . A general approach with respect to anholonomic frames contains the possibility to take into account generic spacetime anisotropies which play an important role in the vicinity of cosmological or astrophysical singularities and for non–trivial reductions of higher dimension theories to lower dimensional ones. In our works we proposed to describe such locally anisotropic spacetimes and interactions by applying the concept of nonlinear connection (in brief, N–connection) field which models the local splitting of spacetime into horizontal (isotropic) and vertical (anisotropic) subspaces. For some values of components the N–connection could parametrize, for instance, toroidal compactifications of higher dimensions but, in general, its dynamics is to be determined, in different approaches, by field equations or constraints in some generalized Kaluza–Klein, gauge gravity or (super)string theories . The concept of N–connection was firstly applied in the framework of Finsler and Lagrange geometry and gravity and their higher order extensions (see Refs. on details). Here it is to be emphasized that every generalized Finsler like geometry could be modelled equivalently on higher dimensions (pseudo)Riemannian spacetimes (for some models by introducing additional torsion and/or nonmetricity structures) with correspondingly adapted anholonomic frame structures. If we restrict our considerations only to the higher dimensional Einstein gravity, the induced N–connection structure becomes a ”pure” anholonomic frame effect which points to the fact that the set of dynamical gravitational field variables given by metric’s components was redefined by introducing local frame variables. The components of a fixed basis define a system (equivalently, frame) of reference (in four dimensions one uses, for instance, the terms of vierbien, or tetradic field) with respect to which, in its turn, one states the coefficients of curved spacetime’s metric and of fundamental physical values and field equations. It should be noted that the procedure of choosing (establishing) of a system of reference must be also physically motivated and that in general relativity this task is not considered as a dynamical one following from the field equations. For trivial models of physical interactions on curved spacetimes one can restrict our considerations only with holonomic frames which locally are linearly equivalent to some coordinate basis. Extending the class of physical fields and interactions (even in the framework of Einstein’s gravity), for example, by introducing spinor fields, statistical and fluid models with spinning particles we have to apply frame bundle methods and deal with general anholonomic frames and phase spaces provided with Cartan’s Finsler like connections (induced by the Levi Civita connection on (pseudo)Riemannian spacetime) . The modelling of kinetic processes with respect to anholonomic frames (which induces corresponding N–connection structures) is very useful with the aim to elucidate flows of fluids of particles not being in local equilibrium. More exactly, such generic anisotropic fluids are considered to be like in a thermodynamic equilibrium with respect to some frames which are locally adapted both to the spacetime metric and N–connection structures but in general (for instance, with respect to local coordinate frames) the conditions of local equilibrium are not satisfied. We developed a proper concept of locally anisotropic spacetime (in brief, la–spacetime) in order to provide a unified (super) geometric background for generalized Kaluza–Klein (super) gravities and low energy (super) string models when the higher dimension spacetime is characterized by generic local anisotropies and compatible metric, N–connection and correspondingly adapted linear connection structures. It should be noted that the term la–spacetime can be used even for Einstein spaces if they are provided with anholonomic frame structures. Our treatment of local anisotropy is more general than that from Refs. which is used for a subclass of Finsler like metrics being conformally equivalent (with conformal factors depending both on spacetime coordinates and tangent vectors) to the flat (pseudo) Euclidean, equivalently, Minkovschi metric. If in the Bogoslovsky, Goenner and Asanov works on Finsler like spacetimes there are investigated possible effects of violation of local Lorentz and/or Poincaré symmetries, our approach to locally anisotropic strings, field interactions and stochastics is backgrounded on the fact that such models could be constructed as to be locally Lorentz invariant with respect to frames locally adapted to N–connection structures. The purpose of the present article is to formulate a Boltzmann type kinetic theory and non–equilibrium thermodynamics on spacetimes of higher dimension with respect to anholonomic frames which models local anisotropies in both type Einstein or generalized Finsler–Kaluza–Klein theories. We shall also discuss possible applications in modern cosmology and astrophysics. On kinetic and thermodynamic part our notations and derivations are often inspired by the Refs. and . A number of formulas will be similar for both locally isotropic and anistropic spacetimes if in the last case we shall consider the equations and vector, tensor, spinor and connection objects with respect to anholonomic rest frames locally adapted to the N–connection structures. In consequence, some tedious proofs and intermediary formulas will be omitted by referring the reader to the corresponding works. To keep the article self–consistent we start up in Section 2 with an overview of the so–called locally anisotropic spacetime geometry and gravity. In Sections 3 and 4 we present the basic definitions for locally anisotropic distribution functions, particle flow and energy momentum tensors and generalize the kinetic equations for la–spacetimes. The equilibrium state and derivation of expressions for the particle density and the entropy density for systems that are closed to a locally anisotropic state of chemical equilibrium are considered in Section 5. Section 6 is devoted to the linearized locally anisotropic transport theory. There we discuss the problem of solution of kinetic equations, prove the linear lows for locally anisotropic non–equilibrium thermodynamics and obtain the explicit formulas for transport coefficients. As an example, in Section 7, we examine the transport theory in curved spaces with rotation ellipsoidal horizons. Concluding remarks are contained in Section 8. ## 2 Spacetimes with Local Anisotropy We outline the necessary background on anholonomic frames and nonlinear connections (N–connection) modelling local anisotropies (la) in curved spaces and on locally anisotropic gravity (see Refs. and for details on spacetime differential geometry and N–connections structures). We shall prove that the Cartan’s moving frame method allows a geometric treatment of both type of locally isotropic (for simplicity, we shall consider (pseudo) Riemannian spaces) and anisotropic (the so–called generalized Finsler–Kaluza–Klein spaces). ### 2.1 Anholonomic frames and Einstein equations In this paper spacetimes are modelled as smooth (i.e class $`C^{\mathrm{}})`$ manifolds $`V_{(d)}`$ of finite integer dimension $`d3,4,\mathrm{},`$ being Hausdorff, paracompact and connected. We denote the local coordinates on $`V_{(d)}`$ by variables $`u^\alpha ,`$ where Greek indices $`\alpha ,\beta ,\mathrm{}=3,4,\mathrm{}`$ could be both type coordinate or abstract (Penrose’s) ones. A spacetime is provided with corresponding geometric structures of symmetric metric $`g_{\alpha \beta }`$ and of linear, in general nonsymmetric, connection $`\mathrm{\Gamma }_{\beta \gamma }^\alpha `$ defining the covariant derivation $`_\alpha `$ satisfying the metricity conditions $`_\alpha g_{\beta \gamma }=0.`$ We shall underline indices, $`\underset{¯}{\alpha },\underset{¯}{\beta },\mathrm{},`$ if one would like to emphasize them as abstract ones. Let a set of basis (frame) vectors $`e^{\underset{¯}{\alpha }}=\{e_\alpha ^{\underset{¯}{\alpha }}=g_{\alpha \beta }e^{\underset{¯}{\alpha }\beta }\}`$ on $`V_{(d)}`$ be numbered by an underlined index. We shall consider only frames associated to symmetric metric structures via relations of type $$e_{\underset{¯}{\alpha }}^\alpha e_{\underset{¯}{\beta }\alpha }=\eta _{\underset{¯}{\alpha },\underset{¯}{\beta }}\text{ and }e_\alpha ^{\underset{¯}{\alpha }}e_\beta ^{\underset{¯}{\beta }}\eta _{\underset{¯}{\alpha },\underset{¯}{\beta }}=g_{\alpha \beta },$$ where the Einstein summation rule is accepted and $`\eta _{\underset{¯}{\alpha },\underset{¯}{\beta }}`$ is a given constant symmetric matrix, for simplicity a pseudo–Euclidean metric of signature $`\left(,+\mathrm{}.+\right)`$ (the sign minus is used in this work for the time like coordinate of spacetime). Operations with underlined and non–underlined indices are correspondingly performed by using the matrix $`\eta _{\underset{¯}{\alpha },\underset{¯}{\beta }},`$ its inverse $`\eta ^{\underset{¯}{\alpha }\underset{¯}{\beta }},`$ and the metric $`g_{\alpha \beta }`$ and its inverse $`g^{\alpha \beta }.`$ A frame (local basis) structure $`e_\alpha `$ on $`V_{(d)}`$ is characterized by its anholonomy coefficients $`w_{\beta \gamma }^\alpha `$ defined from relations $$e_\alpha e_\beta e_\beta e_\alpha =w_{\alpha \beta }^\gamma e_\gamma .$$ (1) With respect to a fixed basis $`e_\alpha `$ and its dual $`e^\beta `$ we can decompose tensors and write down their components, for instance, $$T=T_{\alpha \beta }^\gamma e_\gamma e^\alpha e^\beta $$ where by $``$ it is denoted the tensor product. A spacetime $`V_{(d)}`$ is holonomic (locally integrable) if it admits a frame structure for which the anholonomy coefficients from (1) vanishes, i.e. $`w_{\alpha \beta }^\gamma =0.`$ In this case we can introduce local coordinate bases, $$_\alpha =/u^\alpha $$ (2) and their duals $$d^\alpha =du^\alpha $$ (3) and consider components of geometrical objects with respect to such frames. We note that the general relativity theory was formally defined on holonomic pseudo-Riemannian manifolds. Even on holonomic spacetimes, for various (geometrical, computational and physically motivated) purposes, it is convenient to use anholonomic frames $`e_{\underset{¯}{\alpha }},`$ but we emphasize that for such spacetimes one can always define some linear transforms of frames to a coordinate basis, $`e_{\underset{¯}{\alpha }}=a_{\underset{¯}{\alpha }}^{\underset{¯}{\alpha ^{}}}_{\underset{¯}{\alpha }^{}}`$. By applying both holonomic and anholonomic frames and theirs mutual transforms on holonomic pseudo-Riemannian spaces there were developed different variants of tetradic and spinor gravity and extensions to linear, affine and de Sitter gauge group gravity models . A spacetime is generically anholonomic (locally non-integrable) if it does not admit a frame structure for which the anholonomy coefficients from (1) vanishes, i.e. $`w_{\alpha \beta }^\gamma 0.`$ In this case the anholonomy becomes a proper spacetime characteristics. For instance, a generic anholonomy could be obtained if we consider nontrivial reductions from higher dimension spaces to lower dimension ones. It induces nonvanishing additional terms into the torsion, $$T(\delta _\gamma ,\delta _\beta )=T_{\beta \gamma }^\alpha \delta _\alpha ,$$ and curvature, $$R(\delta _\tau ,\delta _\gamma )\delta _\beta =R_{\beta \gamma \tau }^\alpha \delta _\alpha ,$$ tensors of a linear connection $`\mathrm{\Gamma }_{\beta \gamma }^\alpha ,`$ with coefficients defined respectively as $$T_{\beta \gamma }^\alpha =\mathrm{\Gamma }_{\beta \gamma }^\alpha \mathrm{\Gamma }_{\gamma \beta }^\alpha +w_{\beta \gamma }^\alpha $$ (4) and $$R_{\beta \gamma \tau }^\alpha =\delta _\tau \mathrm{\Gamma }_{\beta \gamma }^\alpha \delta _\gamma \mathrm{\Gamma }_{\beta \delta }^\alpha +\mathrm{\Gamma }_{\beta \gamma }^\phi \mathrm{\Gamma }_{\phi \tau }^\alpha \mathrm{\Gamma }_{\beta \tau }^\phi \mathrm{\Gamma }_{\phi \gamma }^\alpha +\mathrm{\Gamma }_{\beta \phi }^\alpha w_{\gamma \tau }^\phi .$$ (5) The Ricci tensor is defined $$R_{\beta \gamma }=R_{\beta \gamma \alpha }^\alpha $$ (6) and the scalar curvature is $$R=g^{\beta \gamma }R_{\beta \gamma }.$$ (7) The Einstein equations on an anholonomic spacetime are introduced in a standard manner, $$R_{\beta \gamma }\frac{1}{2}g_{\beta \gamma }R=k\mathrm{{\rm Y}}_{\beta \gamma },$$ (8) where the energy–momentum d–tensor $`\mathrm{{\rm Y}}_{\beta \gamma }`$ includes the cosmological constant terms and possible contributions of torsion (4) and matter and $`k`$ is the coupling constant. For a symmetric linear connection the torsion field can be considered as induced by some anholonomy (or equivalently, by some imposed constraints) conditions. For dynamical torsions there are necessary additional field equations, see, for instance, the case of locally anisotropic gauge like theories . The usual locally isotropic Einstein gravity is obtained on the supposition that for every anholonomic frame could be defined corresponding linear transforms to a coordinate frame basis. It is a topic of further theoretical and experimental investigations to establish if the present day experimental data on anisotropic structure of Universe is a consequence of matter and quantum fluctuation induced anisotropies and for some scales the anisotropy is a consequence of anholonomy of observer’s frame. The spacetime anisotropy could be also a generic property following, for instance, from string theory, and from a more general self–consistent gravitational theory when both the left (geometric) and right (matter energy–momentum tensor) parts of Einstein equations depend on anisotropic parameters. ### 2.2 The local anisotropy and nonlinear connection A subclass of anholonomic spacetimes consists from those with local anisotropy modelled by a nonlinear connection structure. In this subsection we briefly outline the geometry of anholonomic frames with induced nonlinear connection structure. The la–spacetime dimension is split locally into two components, $`n`$ for isotropic coordinates and $`m`$ for anisotropic coordinates, when $`n_{(a)}=n+m`$ with $`n2`$ and $`m1.`$ We shall use local coordinates $`u^\alpha =(x^i,y^a),`$ where Greek indices $`\alpha ,\beta ,\mathrm{}`$ take values $`1,2,\mathrm{},n+m`$ and Latin indices $`i`$ and $`a`$ are correspondingly $`n`$ and $`m`$ dimensional, i.e. $`i,j,k\mathrm{}=1,2,`$ $`\mathrm{},n`$ and $`a,b,c,\mathrm{}=1,2,\mathrm{},m.`$ Now, we consider an invariant geometric definition of spacetime’s splitting into isotropic and anisotropic components. For modelling la–spacetimes one uses a vector bundle $`=(E_{n+m},p,M_{(n)},F_{(m)},Gr)`$ provided with nonlinear connection (in brief, N–connection) structure $`N=\{N_j^a\left(u^\alpha \right)\},`$ where $`N_j^a\left(u^\alpha \right)`$ are its coefficients . We use denotations: $`E_{m+n}`$ is a $`(n+m)`$–dimensional total space of a vector bundle, $`M_{(n)}`$ is the $`n`$–dimensional base manifold, $`F_{(m)}`$ is the typical fiber being a $`m`$–dimensional real vector space, $`Gr`$ is the group of automorphisms of $`F_{(m)}`$ and $`p`$ is a surjective map. For simplicity, we shall consider only local constructions on vector bundles. The N–connection is a new geometric object which generalizes that of linear connection. This concept came from Finsler geometry (see the Cartan’s monograph ), the global formulation of it is due to W. Barthel , and it is studied in details in Miron and Anastasiei works . We have extended the geometric constructions for spinor bundles and superbundles with further applications in locally anisotropic field theory and strings and modern cosmology and astrophysics . We have also illustrated that the N–connection could be introduced on (pseudo) Riemnannian spacetimes and in Einstein gravity if we consider anholonomic frames consisting from sets of basic vectors some of them being holonomic and the rest anholonomic. In this case the N–connection coefficients are associated to the frame structure and transforms into some metric components if the considerations are transferred with respect to a coordinate basis. The rigorous mathematical definition of N–connection is based on the formalism of horizontal and vertical subbundles and on exact sequences of vector bundles. Here, for simplicity, we define a N–connection as a distribution which for every point $`u=(x,y)`$ defines a local decomposition of the tangent space of our vector bundle, $`T_uE,`$ into horizontal, $`H_uE,`$ and vertical (anisotropy), $`V_uE,`$ subspaces, i.e. $$T_uE=H_uEV_uE.$$ If a N–connection with coefficients $`N_j^a\left(u^\alpha \right)`$ is introduced on the vector bundle $``$ the modelled spacetime posses a generic local anisotropy and in this case we can not apply in a usual manner the operators of partial derivatives and their duals, differentials. Instead of coordinate bases (2) and (3) we must consider some bases adapted to the N–connection structure: $$\delta _\alpha =(\delta _i,_a)=\frac{\delta }{u^\alpha }=(\delta _i=\frac{\delta }{x^i}=\frac{}{x^i}N_i^b(x^j,y^c)\frac{}{y^b},_a=\frac{}{y^a})$$ (9) and $$\delta ^\beta =(d^i,\delta ^a)=\delta u^\beta =\left(d^i=dx^i,\delta ^a=\delta y^a=dy^a+N_k^a(x^j,y^b)dx^k\right).$$ (10) A nonlinear connection (N–connection) is characterized by its curvature $$\mathrm{\Omega }_{ij}^a=\frac{N_i^a}{x^j}\frac{N_j^a}{x^i}+N_i^b\frac{N_j^a}{y^b}N_j^b\frac{N_i^a}{y^b}.$$ (11) Here we note that the class of usual linear connections can be considered as a particular case when $$N_j^a(x,y)=\mathrm{\Gamma }_{bj}^a(x)y^b.$$ The elongation (by N–connection) of partial derivatives in the adapted to the N–connection partial derivatives (9), or the locally adapted basis (la–basis) $`\delta _\beta `$, reflects the fact that the spacetime $``$ is locally anisotropic and generically anholonomic because there are satisfied anholonomy relations (1), $$\delta _\alpha \delta _\beta \delta _\beta \delta _\alpha =w_{\alpha \beta }^\gamma \delta _\gamma ,$$ where anolonomy coefficients are as follows $`w_{ij}^k`$ $`=`$ $`0,w_{aj}^k=0,w_{ia}^k=0,w_{ab}^k=0,w_{ab}^c=0,`$ $`w_{ij}^a`$ $`=`$ $`\mathrm{\Omega }_{ij}^a,w_{aj}^b=_aN_i^b,w_{ia}^b=_aN_i^b.`$ On a la–spacetime the geometrical objects have a distinguished (by N–connection), into horizontal and vertical components, character. They are briefly called d–tensors, d–metrics and/or d–connections. Their components are defined with respect to a la–basis of type (9), its dual (10), or their tensor products (d–linear or d–affine transforms of such frames could also be considered). For instance a covariant and contravariant d–tensor $`Z,`$ is expressed as $$Z=Z_\beta ^\alpha \delta _\alpha \delta ^\beta =Z_j^i\delta _id^j+Z_a^i\delta _i\delta ^a+Z_j^b_bd^j+Z_a^b_b\delta ^a.$$ (12) A symmetric d–metric on la–space $``$ is written as $$\delta s^2=g_{\alpha \beta }\left(u\right)\delta ^\alpha \delta ^\beta =g_{ij}(x,y)dx^idx^j+h_{ab}(x,y)\delta y^a\delta y^b.$$ A linear d–connection $`D`$ on la–space $`,`$ $$D_{\delta _\gamma }\delta _\beta =\mathrm{\Gamma }_{\beta \gamma }^\alpha (x^k,y)\delta _\alpha ,$$ is parametrized by non–trivial h–v–components, $$\mathrm{\Gamma }_{\beta \gamma }^\alpha =(L_{jk}^i,L_{bk}^a,C_{jc}^i,C_{bc}^a).$$ (13) Some d–connection and d–metric structures are compatible if there are satisfied the conditions $$D_\alpha g_{\beta \gamma }=0.$$ For instance, a canonical compatible d–connection $${}_{}{}^{c}\mathrm{\Gamma }_{\beta \gamma }^{\alpha }=\left({}_{}{}^{c}L_{jk}^{i},^cL_{bk}^a,^cC_{jc}^i,^cC_{bc}^a\right)$$ is defined by the coefficients of d–metric (2.2), $`g_{ij}(x,y)`$ and $`h_{ab}(x,y),`$ and by the coefficients of N–connection, $`{}_{}{}^{c}L_{jk}^{i}`$ $`=`$ $`{\displaystyle \frac{1}{2}}g^{in}\left(\delta _kg_{nj}+\delta _jg_{nk}\delta _ng_{jk}\right),`$ (14) $`{}_{}{}^{c}L_{bk}^{a}`$ $`=`$ $`_bN_k^a+{\displaystyle \frac{1}{2}}h^{ac}\left(\delta _kh_{bc}h_{dc}_bN_i^dh_{db}_cN_i^d\right),`$ $`{}_{}{}^{c}C_{jc}^{i}`$ $`=`$ $`{\displaystyle \frac{1}{2}}g^{ik}_cg_{jk},`$ $`{}_{}{}^{c}C_{bc}^{a}`$ $`=`$ $`{\displaystyle \frac{1}{2}}h^{ad}\left(_ch_{db}+_bh_{dc}_dh_{bc}\right)`$ The coefficients of the canonical d–connection generalize for la–spacetimes the well known Cristoffel symbols. For a d–connection (13) we can compute the components of, in our case d–torsion, (4) $`T_{.jk}^i`$ $`=`$ $`T_{jk}^i=L_{jk}^iL_{kj}^i,T_{ja}^i=C_{.ja}^i,T_{aj}^i=C_{ja}^i,`$ $`T_{.ja}^i`$ $`=`$ $`0,T_{.bc}^a=S_{.bc}^a=C_{bc}^aC_{cb}^a,`$ $`T_{.ij}^a`$ $`=`$ $`\mathrm{\Omega }_{ij}^a,T_{.bi}^a=_bN_i^aL_{.bj}^a,T_{.ib}^a=T_{.bi}^a.`$ In a similar manner, putting non–vanishing coefficients (13) into the formula for curvature (5), we can compute the non–trivial components of a d–curvature $`R_{h.jk}^{.i}`$ $`=`$ $`\delta _kL_{.hj}^i\delta _jL_{.hk}^i+L_{.hj}^mL_{mk}^iL_{.hk}^mL_{mj}^iC_{.ha}^i\mathrm{\Omega }_{.jk}^a,`$ $`R_{b.jk}^{.a}`$ $`=`$ $`\delta _kL_{.bj}^a\delta _jL_{.bk}^a+L_{.bj}^cL_{.ck}^aL_{.bk}^cL_{.cj}^aC_{.bc}^a\mathrm{\Omega }_{.jk}^c,`$ $`P_{j.ka}^{.i}`$ $`=`$ $`_kL_{.jk}^i+C_{.jb}^iT_{.ka}^b(_kC_{.ja}^i+L_{.lk}^iC_{.ja}^lL_{.jk}^lC_{.la}^iL_{.ak}^cC_{.jc}^i),`$ $`P_{b.ka}^{.c}`$ $`=`$ $`_aL_{.bk}^c+C_{.bd}^cT_{.ka}^d(_kC_{.ba}^c+L_{.dk}^cC_{.ba}^dL_{.bk}^dC_{.da}^cL_{.ak}^dC_{.bd}^c)`$ $`S_{j.bc}^{.i}`$ $`=`$ $`_cC_{.jb}^i_bC_{.jc}^i+C_{.jb}^hC_{.hc}^iC_{.jc}^hC_{hb}^i,`$ $`S_{b.cd}^{.a}`$ $`=`$ $`_dC_{.bc}^a_cC_{.bd}^a+C_{.bc}^eC_{.ed}^aC_{.bd}^eC_{.ec}^a.`$ The components of the Ricci tensor (15) with respect to locally adapted frames (9) and (10) (in this case, d–tensor) are as follows: $`R_{ij}`$ $`=`$ $`R_{i.jk}^{.k},R_{ia}=^2P_{ia}=P_{i.ka}^{.k},`$ (15) $`R_{ai}`$ $`=`$ $`{}_{}{}^{1}P_{ai}^{}=P_{a.ib}^{.b},R_{ab}=S_{a.bc}^{.c}.`$ We point out that because, in general, $`{}_{}{}^{1}P_{ai}^{}{}_{}{}^{2}P_{ia}^{}`$ the Ricci d-tensor is non symmetric. Having defined a d-metric of type (2.2) in $``$ we can compute the scalar curvature (7) of a d-connection $`D,`$ $$\stackrel{}{R}=G^{\alpha \beta }R_{\alpha \beta }=\widehat{R}+S,$$ (16) where $`\widehat{R}=g^{ij}R_{ij}`$ and $`S=h^{ab}S_{ab}.`$ Now, by introducing the values (15) and (16) into anholonomic gravity field equations (8) we can write down the system of Einstein equations for la–gravity with prescribed N–connection structure : $`R_{ij}{\displaystyle \frac{1}{2}}\left(\widehat{R}+S\right)g_{ij}`$ $`=`$ $`k\mathrm{{\rm Y}}_{ij},`$ (17) $`S_{ab}{\displaystyle \frac{1}{2}}\left(\widehat{R}+S\right)h_{ab}`$ $`=`$ $`k\mathrm{{\rm Y}}_{ab},`$ $`{}_{}{}^{1}P_{ai}^{}`$ $`=`$ $`k\mathrm{{\rm Y}}_{ai},`$ $`{}_{}{}^{2}P_{ia}^{}`$ $`=`$ $`k\mathrm{{\rm Y}}_{ia},`$ where $`\mathrm{{\rm Y}}_{ij},\mathrm{{\rm Y}}_{ab},\mathrm{{\rm Y}}_{ai}`$ and $`\mathrm{{\rm Y}}_{ia}`$ are the components of the energy–momentum d–tensor field. We note that such decompositions into h– and v–components of gravitational field equations have to be considered even in general relativity if physical interactions are examined with respect to an anholonomic frame of reference with associated N–connection structure. There are variants of la–gravitational field equations derived in the low–energy limits of the theory of locally anisotropic (super)strings or in the framework of gauge like la–gravity when the N–connection and torsions are dynamical fields and satisfy some additional field equations. ### 2.3 Modelling of generalized Finsler geometries in <br>(pseudo) Riemannian spaces The present day trend is to consider the Finsler like geometries and their generalizations as to be quite sophisticate for straightforward applications in quantum and classical field theory. The aim of this subsection is to proof that, a matter of principle, such geometries could be equivalently modelled on corresponding (pseudo) Riemannian manifolds (tangent or vector bundles) by using the Cartan’s moving frame method and from this viewpoint a wide class of Finsler like metrics could be treated as some solutions of usual Einstein field equations. #### 2.3.1 Almost Hermitian Models of Lagrange and Finsler Spaces This topic was originally investigated by Miron and Anastasiei ; here we outline some basic results. Let us model a la–spacetime not on a vector bundle $``$ but on a manifold $`\stackrel{~}{TM}=TM\backslash \{0\}`$ associated to a tangent bundle $`TM`$ of a $`n`$–dimensional base space $`M`$ (when the dimensions of the typical fiber and base are equal, $`n=m`$ and $`\backslash \{0\}`$ means that there is eliminated the null cross–section of the bundle projection $`\tau :TMM)`$ and consider d–metrics of type $$\delta s^2=g_{\alpha \beta }\left(u\right)\delta ^\alpha \delta ^\beta =g_{ij}(x,y)dx^idx^j+g_{ij}(x,y)\delta y^i\delta y^i.$$ On $`TM`$ we can define a natural almost complex structure $`C_{(a)}`$ as follows $$C_{(a)}\left(\delta _i\right)=/y^i\text{ and }C_{(a)}\left(/y^i\right)=\delta _i$$ where the la–derivative $`\delta _i=/x^iN_i^k/y^k`$ (9) and la–differential $`\delta ^i=dy^i+N_k^idx^k`$ (10) act on $`\stackrel{~}{TM}`$ being adapted to a nontrivial N–connection structure $`N=\{N_j^k(x,y)\}`$ in $`TM`$. It is obvious that $`C_{(a)}^2=I.`$ The pair $`(\delta s^2,C_{(a)})`$ defines an almost Hermitian structure on $`\stackrel{~}{TM}`$ with an associate 2–form $$\theta =g_{ij}(x,y)\delta ^i\mathrm{\Lambda }dx^j$$ and the triad $`K^{2n}=(\stackrel{~}{TM},\delta s^2,C_{(a)})`$ is an almost Kählerian space. By straightforward calculations we can verify that the canonical d–connection (13) satisfies the conditions $${}_{}{}^{c}D_{X}^{}\left(\delta s^2\right)=0,{}_{}{}^{c}D_{X}^{}\left(C_{(a)}\right)=0$$ for any d–vector $`X`$ on $`TM`$ and has zero $`hhh`$– and $`vvv`$–torsions. The notion of Lagrange space was introduced as a generalization of Finsler geometry in order to geometrize the fundamental concepts in mechanics. A regular Lagrangian $`L(x^i,y^i)`$ on $`\stackrel{~}{TM}`$ is introduced as a continuity class $`C^{\mathrm{}}`$ function $`L:TM\text{IR}`$ for which the matrix $$g_{ij}(x,y)=\frac{1}{2}\frac{^2L}{y^iy^j}$$ (18) has rank $`n`$ and is of constant signature on $`\stackrel{~}{TM}.`$ A d–metric (2.3.1) with coefficients of form (18), a corresponding canonical d–connection (13) and almost complex structure $`C_{(a)}`$ defines an almost Hermitian model of Lagrange geometry. For arbitrary metrics $`g_{ij}(x,y)`$ of rank $`n`$ and constant signature on $`\stackrel{~}{TM}`$, which can not be determined as a second derivative of a Lagrangian, one defines the so–called generalized Lagrange geometry on $`TM`$ (see details in ). A particular subclass of metrics of type (18) consists from those where instead of a regular Lagrangian one considers a Finsler metric function $`F`$ on $`M`$ defined as $`F:TM\text{IR}`$ having the properties that it is of class $`C^{\mathrm{}}`$ on $`\stackrel{~}{TM}`$ and only continuous on the image of the null cross–section in $`TM,`$ the restriction of $`F`$ on $`\stackrel{~}{TM}`$ is a positive function homogeneous of degree 1 with respect to the variables $`y^i,`$ i. e. $$F(x,\lambda y)=\lambda F(x,y),\lambda \text{IR}^n,$$ and the quadratic form on $`\text{IR}^n`$ with coefficients $$g_{ij}(x,y)=\frac{1}{2}\frac{^2F^2}{y^iy^j},$$ (19) defined on $`\stackrel{~}{TM,}`$ is positive definite. Different approaches to Finsler geometry, its generalizations and applications are examined in a number of monographs and as a rule they are based on the assertion that in this type of geometries the usual (pseudo)Riemannian metric interval $$ds=\sqrt{g_{ij}\left(x\right)dx^idx^j}$$ on a manifold $`M`$ is changed into a nonlinear one defined by the Finsler metric $`F`$ (fundamental function) on $`\stackrel{~}{TM}`$ (we note an ambiguity in terminology used in monographs on Finsler geometry and on gravity theories with respect to such terms as Minkowschi space, metric function and so on) $$ds=F(x^i,dx^j).$$ (20) Geometric spaces with a ’combersome’ variational calculus and a number of curvatures, torsions and invariants connected with nonlinear metric intervals of type (18) are considered as less suitable for purposes of modern field and particle physics. In our investigations of generalized Finsler geometries in (super) string, gravity and gauge theories we advocated the idea that instead of usual geometric constructions based on straightforward applications of derivatives of (19) following from a nonlinear interval (20) one should consider d–metrics (2.2) and/or (2.3.1) with the coefficients of necessity determined via an almost Hermitian model of a Lagrange (18), Finsler geometry (19) and/or their extended variants. This way, by a synthesis of the moving frame method with the geometry of N–connection, we can investigate a various class of higher and lower dimension gravitational models with generic or induced anisotropies in a unified manner on some anholonomic and/or Kaluza–Klein spacetimes. As a matter of principle, having a physical model with a d–metric and geometrical objects associated to la–frames, we can redefine the physical values with respect to a local coordinate base on a (pseudo) Riemannian space. The coefficients $`g_{ij}(x,y)`$ and $`h_{ab}(x,y)`$ of d–metric written for a la–basis $`\delta u^\alpha =(dx^i,\delta y^a)`$ transforms into a usual (pseudo) Riemannian metric if we rearrange the components with respect to a local coordinate basis $`du^{\widehat{\alpha }}=(dx^{\widehat{i}},dy^{\widehat{a}})`$, $$g_{\widehat{\alpha }\widehat{\beta }}=\left(\begin{array}{cc}g_{\widehat{i}\widehat{j}}+N_{\widehat{i}}^{\widehat{a}}N_{\widehat{j}}^{\widehat{b}}h_{\widehat{a}\widehat{b}}& N_{\widehat{j}}^{\widehat{e}}h_{\widehat{a}\widehat{e}}\\ N_{\widehat{i}}^{\widehat{e}}h_{\widehat{b}\widehat{e}}& h_{\widehat{a}\widehat{b}}\end{array}\right),$$ (21) where ’hats’ on indices emphasize that coefficients of the metric are given with respect to a coordinate (holonomic) basis on a spacetime $`V^{n+m}.`$ Parametrizations (ansatzs) of metrics of type (21) are largely applied in Kaluza–Klein gravity and its generalizations . In our works , following the geometric constructions from , we proved that the physical model of interactions is substantially simplified, as well we can correctly elucidate anisotropic effects, if we work with diagonal blocks of d–metrics (2.2) with respect to anholonomic frames determined by a N–connection structure. #### 2.3.2 Finsler like metrics in Einstein’s gravity There are obtained (see Appendix) some classes of locally anisotropic cosmological and black hole like solutions (in three, four and higher dimensions) which can be treated as generalized Finsler metrics being of nonspheric symmetry (with rotation ellipsoid, torus and cylindrical event horizons, or with elliptical oscillations of horizons). Under corresponding conditions such metrics could be solutions of field equations in general relativity or its lower or higher dimension variants. Here we shall formulate the general criteria when a Finsler like metric could be a solution of gravitational field equations in Einstein gravity. Let consider on $`\stackrel{~}{TM}`$ an ansatz of type (21) when $`g_{ij}=h_{ij}=\frac{1}{2}^2F^2/y^iy^j`$ (for simplicity, we omit ’hats’ on indices) i.e. $$g_{\widehat{\alpha }\widehat{\beta }}=\frac{1}{2}\left(\begin{array}{cc}\frac{^2F^2}{y^iy^j}+N_i^kN_j^l\frac{^2F^2}{y^ky^l}& N_j^l\frac{^2F^2}{y^ky^l}\\ N_i^k\frac{^2F^2}{y^ky^l}& \frac{^2F^2}{y^iy^j}\end{array}\right).$$ (22) A metric (22), induced by a Finsler quadratic form (19) could be treated in a framework of a Kaluza–Klein model if for some values of Finsler metric $`F(x,y)`$ and N–connection coefficients $`N_i^k(x,y)`$ this metric is a solution of the Einstein equations (8) written with respect to a holonomic frame. For the dimension $`n=2,`$ when the values $`F`$ and $`N`$ are chosen to induce locally a (pseudo) Riemannian metric $`g_{\widehat{\alpha }\widehat{\beta }}`$ of signature $`(,+,+,+)`$ and with coefficients satisfying the four dimensional Einstein equations we define a subclass of Finsler metrics in the framework of general relativity. Here we note that, in general, a N–connection, on a Finsler space, subjected to the condition that the induced (pseudo) Riemannian metric is a solution of usual Einstein equations does not coincide with the well known Cartan’s N–connection . We have to examine possible compatible deformations of N–connection structures . Instead of Finsler like quadratic forms we can consider ansatzs of type (21) with $`g_{ij}`$ and $`h_{ij}`$ induced by a Lagrange quadratic form (18). A general approach to the geometry of spacetimes with generic local anisotropy can be developed on embeddings into corresponding Kaluza–Klein theories and adequate modelling of la–interactions with respect to anholonomic or holonomic frames and associated N–connection structures. ## 3 Collisionless relativistic kinetic equation As argued in Section 2, the spacetimes could be of generic local anisotropy after nontrivial reductions from some higher dimension theories or posses a local anisotropy induced by anholonomic frame structures even we restrict our considerations to the general relativity theory. In this line of particular interest is the formulation of relativistic kinetic theory with respect to general anholonomic frames and elucidation of locally anisotropic kinetic and thermodynamic processes. ### 3.1 The distribution function and its moments We use the relativistic approach to kinetic theory (we refer readers to monographs for hystory and complete treatment). Let us consider a simple system consisting from $`\varpi `$ point particles of mass $`\stackrel{~}{m}`$ in a la–spacetime with a d–metric $`g_{\alpha \beta }.`$ Every particle is characterized by its coordinates $`u_{(l)}^\alpha =(x_{(l)}^i,y_{(l)}^a),(u^{\overline{1}}=x^1=ct`$ is considered the time like coordinate, for simplicity we put hereafter the light velocity $`c=1);`$ $`x_{(l)}^2,x_{(l)}^3,\mathrm{},x_{(l)}^n`$ and $`y_{(l)}^1,y_{(l)}^2,\mathrm{},y_{(l)}^m`$ are respectively space like and anisotropy coordinates, where the index $`(l)`$ enumerates the particles in the system. The particles momenta are denoted by $`p_{\alpha \left(l\right)}=g_{\alpha \beta }p_{(l)}^\beta ,`$ . We shall use the distribution function $`\mathrm{\Phi }(u^\alpha ,p_\beta ),`$ given on the space of supporting elements $`(u^\alpha ,p_\beta ),`$ as a general characteristic of particle system. The system (of particles) is to be defined by using the random function $$\varphi (u^\alpha ,p_\beta )=\underset{l=1}{\overset{\varpi }{}}\delta s\delta ^{(n_{(a)})}\left(u^\alpha u_{(l)}^\alpha (s)\right)\delta ^{(n_{(a)})}\left(p^\beta p_{(l)}^\beta (s)\right),$$ where the sum is taken on all system’s particles, $$\delta s=\sqrt{g_{\alpha \beta }\delta u^\alpha \delta u^\beta }$$ is the interval element along the particle trajectory $`u_{(l)}^\alpha (s)`$ parametrized by a natural parameter $`s`$ and $`\delta ^{n_{(a)}}\left(u^\alpha \right)`$ is the $`n_{(a)}`$–dimensional delta function. The functions $`u_{(l)}^\alpha (s)`$ and $`p_{(l)}^\beta (s)`$ describing the propagation of the $`l`$–particle are found from the motion equations on la–spacetime $`\stackrel{~}{m}`$ $`{\displaystyle \frac{\delta u_{(l)}^\alpha }{ds}}=p_{(l)}^\alpha `$ $`\stackrel{~}{m}`$ $`D{\displaystyle \frac{\delta p_{\alpha (l)}}{ds}}=\stackrel{~}{m}{\displaystyle \frac{\delta p_{\alpha (l)}}{ds}}{}_{}{}^{c}\mathrm{\Gamma }_{\alpha \beta }^{\tau }\left(u_{(l)}^\gamma (s)\right)p_{\tau (l)}p_{(l)}^\beta =F_{\alpha (l)},`$ where $`{}_{}{}^{c}\mathrm{\Gamma }_{\beta \gamma }^{\alpha }`$ is the canonical d–connection with coefficients (14) and by $`F_{\alpha (l)}`$ is denoted an exterior force (electromagnetic or another type) acting on the $`l`$th particle. The distribution function $`\mathrm{\Phi }(u,p)`$ is defined by averaging on paths of random function $$\mathrm{\Phi }(u,p)=<<\varphi (u,p)>>$$ where brackets $`<<\mathrm{}>>`$ denote path averaging. Let consider a space like hypersurface $`(u^\alpha )=const`$ with elements $`\delta \mathrm{\Sigma }_\alpha =n_\alpha \delta \mathrm{\Sigma },`$ where $`n_\alpha =\frac{D_\alpha 𝒯}{\left|D𝒯\right|}`$ and $`d\mathrm{\Sigma }=\frac{\sqrt{|g|}\delta ^{n_{(a)}}u}{n_\alpha \delta u^\alpha },`$ $`\left|D𝒯\right|=\sqrt{g^{\alpha \beta }D_\alpha D_\beta 𝒯}`$ and $`\sqrt{|g|}\delta ^{n_{(a)}}u`$ is the invariant space–time volume. A local system of reference in a point $`u_{(0)}^\alpha ,`$ with the metric $`g_{ij}^{(0)}=\eta _{ij}=diag(1,1,1,\mathrm{},1),`$ is obtained if $`(u_{(0)}^\alpha )=x_{(0)}^1`$ . In this case $`\delta \mathrm{\Sigma }_\alpha =\delta _\alpha ^1\delta ^{n_{(a)}1}u,`$ $`\delta ^{n_{(a)}1}u=dx^2dx^3\mathrm{}dx^n\delta y^1\delta y^2\mathrm{}\delta y^m`$ and $`\delta _\beta ^\alpha `$ is the Kronecker symbol. The value $$\mathrm{\Phi }(u,p)v^\alpha d\mathrm{\Sigma }_\alpha \frac{\delta ^{n_{(a)}}p}{\sqrt{|g|}},$$ where $`v^\alpha =p^\alpha /\stackrel{~}{m}`$ and $`\delta P=\delta ^{n_{(a)}}p/\sqrt{|g|}=\delta p_1\delta p_2\mathrm{}\delta p_{n_{(a)}}/\sqrt{|g|}`$ is the invariant volume in the momentum space and defines the quantity of particles intersecting the hypersurface element $`\delta \mathrm{\Sigma }_\alpha `$ with momenta $`p_\alpha `$ included into the element $`\delta P`$ in the vicinity of the point $`u^\alpha .`$ The first $`<p^\alpha >`$ and second moments $`<p^\alpha p^\beta >`$ of distribution function $`\mathrm{\Phi }(x,p)`$ give respectively the flux of particles $`n^\alpha `$ and the energy–momentum $`T^{\alpha \beta },`$ $$<p^\alpha >=\mathrm{\Phi }(u,p)p^\alpha \delta P=\stackrel{~}{m}n^\alpha $$ (24) and $$<p^\alpha p^\beta >=\mathrm{\Phi }(u,p)p^\alpha p^\beta \delta P=\stackrel{~}{m}\mathrm{{\rm Y}}^{\alpha \beta }.$$ (25) We emphasize that the motion equations (3.1) have the first integral, $`g^{\alpha \beta }p_\alpha ^{(l)}p_\beta ^{(l)}=\stackrel{~}{m}^2=const`$ for every $`l`$–particle, so the functions (3.1) and, in consequence, $`\mathrm{\Phi }(u,p)`$ are non–zero only on the mass hypersurface $$g^{\alpha \beta }p_\alpha p_\beta =\stackrel{~}{m}^2,$$ (26) which in distinguished by a N–connection form (see the dual to d–metric (2.2)) is written $$g^{ij}p_ip_j+g^{ab}p_ap_b=\stackrel{~}{m}^2.$$ For computations it is convenient to use a new distribution function $`f(u,p)`$ on a $`n_{(a)}1`$ dimensional la–hyperspace (26) $$\mathrm{\Phi }(u^\alpha ,p_\beta )=f(u^\alpha ,p_{\widehat{\beta }})\delta \left(\sqrt{g^{\alpha \beta }p_\alpha p_\beta }\stackrel{~}{m}\right)\theta \left(p_{\overline{1}}\right)$$ where $$\theta \left(p_{\overline{1}}\right)=\left\{\begin{array}{c}1,p_{\overline{1}}0\\ 0,p_{\overline{1}}<0\end{array}\right\}$$ and the index $`\widehat{\beta }`$ runs values in $`n_{(a)}1`$ dimensional la–space. The flux of particles (24) and of energy–momentum (25) are computed by using $`f(u,p)`$ as $$n^\alpha =\delta \varsigma p^\alpha f(u^\gamma ,p_{\widehat{\beta }}),$$ with respective base $`n^i`$ and fiber $`n^a`$ components, $$n^i=\delta \varsigma p^if(u^\gamma ,p_{\widehat{\beta }})\text{ and }n^a=\delta \varsigma p^af(u^\gamma ,p_{\widehat{\beta }}),$$ when $$\mathrm{{\rm Y}}^{\alpha \beta }=\delta \varsigma p^\alpha p^\beta f(u^\gamma ,p_{\widehat{\beta }})$$ where $`p^{\overline{1}}`$ is expressed via $`p^2,p^3,\mathrm{},p^{n_{(a)}}`$ by using the equation (26) and it is used the abbreviation $`\delta \varsigma =\delta ^{n_{(a)}1}p/(\sqrt{|g|}p^{\overline{1}}).`$ The energy–momentum can be also split into base–fiber (horizontal–vertical, in brief, h– and v–, or hv–components) by a corresponding distinguishing of momenta, $`\mathrm{{\rm Y}}^{\alpha \beta }=\{\mathrm{{\rm Y}}^{ij},\mathrm{{\rm Y}}^{aj},\mathrm{{\rm Y}}^{ib},\mathrm{{\rm Y}}^{ab}\}.`$ The one–particle distribution function $`f(u,p)`$ characterizes the number of particles (with mass $`\stackrel{~}{m},`$ or massless if $`\stackrel{~}{m}=0)`$ at a point $`u^\alpha `$ of a la–spacetime of dimension $`n_{(a)}`$ having distinguished by N–connection momentum vector (in brief, momentum d–vector) $`p^\alpha =(p^i,p^a)=(p^{\overline{1}},\stackrel{}{p}).`$ One states that $$f(u,p)p^\beta \delta \sigma _\beta \delta \varsigma =f(u,p)p^i\delta \sigma _i\delta \varsigma +f(u,p)p^a\delta \sigma _a\delta \varsigma ,$$ gives the number of world–lines of particles with momentum d–vector $`p^\alpha `$ in an interval $`\delta p^\alpha `$ around $`p,`$ crossing a space like hypersurface $`\delta \sigma _\beta `$ at a point $`u.`$ When $`\delta \sigma _\beta `$ is taken to be time like one has $`\delta \sigma _\beta =(\delta ^{n_{(a)}}u,0,\mathrm{},0).`$ The velocity d–field of a fluid $`U^\mu `$ in la–spacetime can be defined in some way like in relativistic kinetic theory . To obtain a particularly simple form for the energy–momentum d–tensor one should follow the Landau–Lifshitz approach when the locally anisotropic fluid velocity $`U^\mu \left(u\right)=(U^i\left(u\right),U^a\left(u\right))`$ and the energy density $`\epsilon \left(u\right)`$ are defined respectively as the eigenvector and eigenvalue of the eigenvalue equation $$\mathrm{{\rm Y}}^{\alpha \beta }U_\beta =\epsilon U^\alpha .$$ (27) A unique value of $`U^\alpha `$ is to be found from the conditions to be a time like and normalized to the unity d–vector, $`U^\alpha U_\alpha =1.`$ Having so fixed the fluid d–velocity we can define correspondingly the particle density $$\stackrel{~}{n}\left(u\right)=n^\alpha U_\alpha ,$$ (28) the energy density $$\stackrel{~}{\epsilon }\left(u\right)=U_\alpha \mathrm{{\rm Y}}^{\alpha \beta }U_\beta $$ (29) and the average energy per particle $$e=\stackrel{~}{\epsilon }/\stackrel{~}{n}.$$ (30) We can introduce also the pressure d–tensor $$P^{\alpha \beta }\left(u\right)=\mathrm{\Delta }^{\alpha \tau }\mathrm{{\rm Y}}_{\tau \nu }\mathrm{\Delta }^{\nu \beta }$$ (31) by applying the projector $$\mathrm{\Delta }^{\alpha \beta }=g^{\alpha \beta }U^\alpha U^\beta $$ (32) with properties $$\mathrm{\Delta }^{\alpha \beta }\mathrm{\Delta }_{\beta \mu }=\mathrm{\Delta }_\mu ^\alpha ,\mathrm{\Delta }^{\alpha \beta }U_\beta =0\text{ and }\mathrm{\Delta }_\mu ^\mu =n_{(a)}1.$$ With respect to a la–frame (9) we can introduce a particular Lorentz system, called the locally anisotropic rest frame of the fluid when $`U^\mu =(1,0,0,\mathrm{},0)`$ and $`\mathrm{\Delta }_\mu ^\alpha =diag(0,1,1,\mathrm{},1).`$ So the pressure d–tensor was defined as to coincide with the la–space part of the energy–momentum d–tensor with respect to a locally anisotropic rest frame. ### 3.2 Collisionless kinetic equation We start our proof by applying the identity $$𝑑s\frac{d}{ds}\delta ^{n_{(a)}}\left(u^\alpha u_{(l)}^\alpha (s)\right)\delta ^{n_{(a)}}\left(p^\beta p_{(l)}^\beta (s)\right)=0.$$ Differentiating under this integral and taking into account the equation (26) we get the relation $$\frac{\delta \left(p^\alpha \varphi \right)}{u^\alpha }+\frac{}{p_\beta }\left({}_{}{}^{c}\mathrm{\Gamma }_{\beta \epsilon }^{\mu }p_\mu p^\epsilon \varphi \right)=0.$$ (33) In our further considerations we neglect the interactions between the particles and suppose that the metric of the background gravitational field $`g_{\alpha \beta }`$ does not depend on motion of particles. After averaging (33) on paths we define the equation for the one–particle distribution function $$\frac{\delta \left(p^\alpha \mathrm{\Phi }\right)}{u^\alpha }+\frac{}{p_\beta }\left({}_{}{}^{c}\mathrm{\Gamma }_{\beta \epsilon }^{\mu }p_\mu p^\epsilon \mathrm{\Phi }\right)=0.$$ (34) As a matter of principle we can generalize the problem when particles are considered to be also sources of gravitational field which is self–consistently defined from the Einstein equations with the energy–momentum tensor defined by the formula (25). Taking into account the identity $$\frac{\delta p^\alpha }{u^\alpha }+\frac{}{p_\beta }({}_{}{}^{c}\mathrm{\Gamma }_{\beta \epsilon }^{\mu }p_\mu p^\epsilon )=0$$ we get from (34) the collisionless kinetic equation for the distribution function $`\mathrm{\Phi }(u^\alpha ,p_\beta )`$ $$p^\beta \widehat{D}_\beta \mathrm{\Phi }=0,$$ (35) where $$\widehat{D}_\beta =\frac{\delta }{u^\beta }{}_{}{}^{c}\mathrm{\Gamma }_{\beta \alpha }^{\epsilon }p^\alpha \frac{}{p^\epsilon }$$ (36) generalizes the Cartan’s covariant derivation for a space with higher order anisotropy (locally parametrized by supporting elements $`(u^\alpha ,p_\beta )),`$ provided with a (extended to momenta coordinates) higher order nonlinear connection $$\widehat{N}_\beta ^\epsilon =\delta _a^\epsilon \delta _\beta ^iN_i^a{}_{}{}^{c}\mathrm{\Gamma }_{\beta \alpha }^{\epsilon }p^\alpha $$ (37) (on the geometry of higher order anisotropic spaces and superspaces and possible applications in physics see ). The equation (35) can be written in equivalent form for the distribution function $`f(u^\alpha ,p_{\widehat{\beta }}),`$ $$p^\lambda \widehat{D}_\lambda f=0,$$ (38) with the Cartan’s operator defined on $`n_{(a)}1`$ dimensional momentum space. The kinetic equations (35) and, equivalently, (38) reflect the conversation low of the quantity of particles in every volume of the space of supporting elements which holds in absence of collisions. Finally, in this subsection, we note that the kinetic equation (35) is formulated on the space of supporting elements $`(u^\alpha ,p_\beta ),`$ which is characterized by coordinate transforms $$x^i^{}=x^i^{}\left(x^k\right),y^a^{}=y^aK_a^a^{}\left(x^i\right)$$ (39) and $$p_j^{}=p_i\frac{x^i}{x^j^{}},p_a^{}=p_aK_a^{}^a(x^i),$$ where $`K_a^a^{}\left(x^i\right)`$ and $`K_a^{}^a(x^i)`$ take values correspondingly in the set of matrices parametrizing the group of linear transforms $`GL(m,\text{IR}^m),`$ where IR denotes the set of real numbers. In a particular case, for dimensions $`n=m`$ we can parametrize $$K_i^i^{}\left(x^i\right)=\frac{x^i^{}}{x^i}\text{ and }K_i^{}^i(x^i)=\frac{x^i}{x^i^{}}_{x^i=x^i(x^i^{})}.$$ A distinguished by higher order nonlinear connection (37) tensor $`Q_{\beta _1\beta _2\mathrm{}\beta _q}^{\alpha _1\alpha _2\mathrm{}\alpha _r}(u^\epsilon ,p_\tau )`$ being contravariant of rang $`r`$ and covariant of rang $`q`$ satisfy the next transformation laws under chaingings of coordinates of type (39): $$Q_{\beta _1^{}\beta _2^{}\mathrm{}\beta _q^{}}^{\alpha _1^{}\alpha _2^{}\mathrm{}\alpha _r^{}}(u^\epsilon ^{},p_\tau ^{})=\frac{u^{\alpha _1^{}}}{u^{\alpha _1}}\mathrm{}\frac{u^{\alpha _r^{}}}{u^{\alpha _r}}\frac{u^\beta }{u^{\beta _1^{}}}\mathrm{}\frac{u^{\beta _q}}{u^{\beta _q^{}}}Q_{\beta _1\beta _2\mathrm{}\beta _q}^{\alpha _1\alpha _2\mathrm{}\alpha _r}(u^\epsilon ,p_\tau ).$$ (40) Operators of type (37) and transformations (39) and (40), on first order of anisotropy, on tangent bundle $`TM`$ on a $`n`$ dimensional manifold $`M`$ were considered by E. Cartan in his approach to Finsler geometry and by A. A. Vlasov in order to formulate the statistical kinetic theory on a Finsler geometry background. It should be emphasized that the Cartan–Vlasov approach is to be applied even in general relativity because the kinetic processes are to be examined in a phase space provided with local coordinates $`(u^\epsilon ,p_\tau )`$. Our recent generalizations to higher order anisotropy (including spinor and supersymmetric spaces) are to applied in the case of models with nontrivial reductions (modelled by N–connections) from higher dimensions to lower dimensional ones. ## 4 Kinetic equation with pair collisions In this section we shall prove the relativistic kinetic equations for one–particle distribution function $`f(u^\alpha ,p_\beta )`$ with pair collisions . We summarize the related results and generalize the constructions for Minkowski spaces . ### 4.1 Integral of collisions, differential cross–section and velocity of transitions If collisions of particles are taken into account (for simplicity we shall consider only pair collisions), the quantity of particles from a volume in the space of supporting elements is not constant. We have to introduce a source of particles $`C(x,p),`$ for instance, in the kinetic equations (38), $$p^\alpha \widehat{D}_\alpha f=C\left(f\right)=C(u,p).$$ (41) The scalar function $`C(u,p)`$ is called the integral of collisions (in the space of supporting elements). The value $$\mathrm{}^{n_{(a)}}u\frac{\mathrm{}^{n_{(a)}1}p}{p^1}C(u,p)$$ is the changing of the quantity of particles under pair collisions in a region $`\mathrm{}^{n_{(a)}}u\mathrm{}^{n_{(a)}1}p.`$ Let us denote by $$\frac{W(p,p_{[1]}|p^{},p_{[1]}^{})}{p^1p_{[1]}^1p^1p_{[1]}^1}$$ (42) the probability of transition for two particles which before scattering have the momenta $`p_{\widehat{i}}`$ and $`p_{1\widehat{i}}`$ and after scattering the momenta $`p_{\widehat{i}}^{}`$ and $`p_{1\widehat{i}}^{}`$ with respective inaccuracies $`\mathrm{}^{n_{(a)}1}p^{}`$ and $`\mathrm{}^{n_{(a)}1}p_1^{}.`$ The function $`W(p,p_{[1]}|p^{},p_{[1]}^{})`$ (the so–called collision rate) is symmetric on arguments $`p,p_{[1]}`$ and $`p^{},p_{[1]}^{}`$ and describes the velocity of transitions with conservation of momenta of type $$p^\alpha +p_{[1]}^\alpha =p^\alpha +p_{[1]}^\alpha $$ when the conditions (26) hold. The number of binary collisions within a la–spacetime interval $`\delta u`$ around a point $`u`$ between particles with initial momenta in the ranges $$(\stackrel{}{p},\stackrel{}{p}+\delta \stackrel{}{p})\text{ and }(\stackrel{}{p}_{[1]},\stackrel{}{p}_{[1]}+\delta \stackrel{}{p}_{[1]})$$ and final momenta in the ranges $$(\stackrel{}{p}^{},\stackrel{}{p}^{}+\delta \stackrel{}{p}^{})\text{ and }(\stackrel{}{p}_{[1]}^{},\stackrel{}{p}_{[1]}^{}+\delta \stackrel{}{p}_{[1]}^{})$$ is given by $$f(u,p)f(u,p_{[1]})W(p,p_{[1]};p^{},p_{[1]}^{})\delta \varsigma _p\delta \varsigma _{p_{[1]}}\delta \varsigma _p^{}\delta \varsigma _{p_{[1]}^{}}\delta u$$ (43) were, for instance, we abbreviated $`\delta \varsigma _p=\delta ^{n_{(a)}}p/(\sqrt{|g|}p^{\overline{1}}).`$ The collision integral of the Boltzmann equation in la–spacetime (41) is expressed in terms of the collision rate (42) $$C\left(f\right)=f(u,p)f(u,p_{[1]})W(p,p_{[1]};p^{},p_{[1]}^{})\delta \varsigma _{p_{[1]}}\delta \varsigma _p^{}\delta \varsigma _{p_{[1]}^{}}.$$ The collision integral is related to the differential cross section (see below). ### 4.2 The Cross Section in La–spacetime The number $`n_{(bin)}`$ of binary collisions per unit time and unit volume when the initial momenta of the colliding particles lie in the ranges $$(\stackrel{}{p},\stackrel{}{p}+\delta \stackrel{}{p})\text{ and }(\stackrel{}{p}_{[1]},\stackrel{}{p}_{[1]}+\delta \stackrel{}{p}_{[1]})$$ and the final momenta are in some interval $`\varsigma `$ in the space of variables $`\stackrel{}{p}^{}`$ and $`\stackrel{}{p}_{[1]}^{}`$ is to be obtained from (43) by dividing on $`\delta u=\delta t\delta x^2\mathrm{}\delta x^ndy^1\mathrm{}dy^m`$ and integrating with respect to the primed variables, $$n_{(bin)}=f(u,p)f(u,p_{[1]})\times \delta \varsigma _p\delta \varsigma _{p_{[1]}}_\varsigma W(p,p_{[1]};p^{},p_{[1]}^{})\delta \varsigma _p^{}\delta \varsigma _{p_{[1]}^{}}.$$ (44) Let us introduce an auxiliary velocity $$v=F/(p^{\overline{1}}p_{[1]}^{\overline{1}})$$ (45) were the so–called M$`\ddot{o}`$ller flux factor is defined by $$F=\left[\left(p^\alpha p_{[1]\alpha }\right)^2\stackrel{~}{m}^4\right]^{1/2}.$$ (46) It may be verified that the speed $`v`$ reduces to the relative speed of particles in a frame in which one of the particles initially is at rest. We also consider the product of the number density of target particles with momenta in the range $`(\stackrel{}{p}_{[1]},\stackrel{}{p}_{[1]}+\delta \stackrel{}{p}_{[1]})`$ (given by $`f(u,p_{[1]})\delta ^{n_{(a)}1}p_{[1]})`$ and the flux of incoming particles with momenta in the range $`(\stackrel{}{p},\stackrel{}{p}+\delta \stackrel{}{p})`$ (given by $`f(u,p)\delta ^{n_{(a)}1}pv,`$ where $`v`$ is the speed (45). This product can be written $$Ff(u,p)f(u,p_{[1]})\delta \varsigma _p\delta \varsigma _{p_{[1]}}.$$ (47) By definition (like in usual the relativistic Boltzmann theory ) the cross section $`\stackrel{~}{\sigma }`$ is the division of the number (44) to the number (47) $$\stackrel{~}{\sigma }=\frac{1}{F}_\varsigma W(p,p_{[1]};p^{},p_{[1]}^{})\delta \varsigma _p^{}\delta \varsigma _{p_{[1]}^{}}.$$ (48) The condition that collisions are local implies that the collision rate must contain $`n_{(a)}`$ delta–functions $$W(p,p_{[1]};p^{},p_{[1]}^{})=w(p,p_{[1]};p^{},p_{[1]}^{})\times \delta ^{n_{(a)}}\left(p+p_{[1]}p^{}p_{[1]}^{}\right).$$ (49) Let $`\text{IR}^{n_{(a)}1}`$ be a $`(n_{(a)}1)`$–dimensional Euclidean space enabled with an orthonormal basis $`(e_1,e_2,\mathrm{},e_{n_{(a)}1})`$ and denote by $`(u^1,u^2,\mathrm{},u^{n_{(a)}1})`$ the Cartesian coordinates with respect to this basis. For calculations on scattering of particles it is useful to apply the system of spherical coordinates $`(r_{n_{(a)}}=r,\theta _{n_{(a)}1},\theta _{n_{(a)}2},\mathrm{},\theta _1)`$ associated with the Cartesian coordinates $`(u_1,u_2,\mathrm{},).`$ The volume element can be expressed as $$d^{n_{(a)}1}u=(r)^{n_{(a)}2}drd^{n_{(a)}2}\mathrm{\Omega }_\theta ,$$ where the element of solid angle, the $`(n_{(a)}1)`$–spherical element, is given by $`d^{n+m2}`$ $`\mathrm{\Omega }_\theta =\mathrm{sin}^{n+m3}\theta _{n+m2}\mathrm{sin}^{n+m4}\theta _{n+m3}\mathrm{}`$ (50) $`\mathrm{sin}^2\theta _3\mathrm{sin}\theta _2\times d\theta _{n+m2}\mathrm{}d\theta _2d\theta _1.`$ In our further considerations we shall consider that the Cartesian and spherical $`(n_{(a)}1)`$–dimensional coordinates are given with respect to a la–frame of type (9). In order to eliminate the delta functions from (49) put into (48) we fix as a reference frame the center of mass frame for the collision between two particles with initial momenta $`\stackrel{}{p}`$ and $`\stackrel{}{p}_{[1]}.`$ The quantities defined with respect to the center of mass frame will be enabled with the subindex $`CM.`$ One denotes by $`\stackrel{}{p}_{CM}`$ the polar axis and characterizes the directions $`\stackrel{}{p}_{CM}^{}`$ of the outgoing particles with respect to polar axis by means of generalized spherical coordinates. In this case $$\delta ^{n_{(a)}1}p_{CM}^{}=\left|\stackrel{}{p}_{CM}^{}\right|^{n_{(a)}2}d\left|\stackrel{}{p}_{CM}^{}\right|d\mathrm{\Omega }_{CM}$$ (51) where $`d\mathrm{\Omega }_{CM}`$ is given by the formula (50). The total $`\left(n_{(a)}+1\right)`$–momenta before and after collision in la–spacetime are given by d–vectors $$P^\alpha =p^\alpha +p_{[1]}^\alpha ,P^\alpha =p^\alpha +p_{[1]}^\alpha .$$ (52) Following from $$P_{CM}^1=2p_{CM}^1=2\sqrt{\left|\stackrel{}{p}_{CM}^{}\right|^2+\stackrel{~}{m}^2}$$ we have $$\left|\stackrel{}{p}_{CM}^{}\right|^2d\left|\stackrel{}{p}_{CM}^{}\right|=\frac{1}{4}P_{CM}^1dP_{CM}^1.$$ In result the formula for the volume element (51) transforms into $$\delta ^{n_{(a)}1}p_{CM}^{}=\frac{1}{4}\left|\stackrel{}{p}_{CM}^{}\right|^{n_{(a)}2}P_{CM}^1dP_{CM}^1d\mathrm{\Omega }_{CM}.$$ Inserting this collision rate (49) into (48) we obtain an integral which can be rewritten (see the Appendices 1 and 7 to the paper for details on transition to the center of mass variables; in la–spacetimes one holds good similar considerations with that difference that me must work with respect to la–frames (9) and (10) and d–metric (2.2) ) $$\sigma =\frac{F^{n_{(a)}4}}{E^{n_{(a)}2}}_\varsigma w𝑑\mathrm{\Omega }_{CM}$$ (53) where $`E=\sqrt{P^\alpha P_\alpha },,`$ is the total energy (devided by $`c=1).`$ The differential section in the center of mass system is found from (53) $$\frac{\delta \sigma }{d\mathrm{\Omega }}_{_{CM}}=\frac{F^{n_{(a)}4}}{E^{n_{(a)}2}}w$$ which allow to express the collision rate (49) as $$W(p,p_{[1]};p^{},p_{[1]}^{})=\frac{E^{n_{(a)}2}}{F^{n_{(a)}4}}\times \frac{\delta \sigma }{d\mathrm{\Omega }}_{_{CM}}\delta ^{n_{(a)}}\left(p+p_{[1]}p^{}p_{[1]}^{}\right).$$ (54) We shall use the formula (54) in order to compute the transport coefficients. ## 5 Equilibrium States in La–Spacetimes When the system is in equilibrium we can derive an expression for the particle distribution function $`f(u,p)=f_{(eq)}(u,p)`$ in a similar way as for locally isotropic spaces and write $$f_{(eq)}(u,p)=\frac{1}{(2\pi \mathrm{})^{n_{(a)}1}}\mathrm{exp}\left(\frac{\mu p^iU_ip^aU_a}{k_BT}\right),$$ (55) where $`k_B`$ is Boltzmann’s constant and $`\mathrm{}`$ is Planck’s constant divided by $`2\pi `$ and $`\mu =\mu \left(u\right)`$ and $`T=T\left(u\right)`$ are respectively the thermodynamic potential and temperature. For simplicity, in this section we shall omit explicit dependencies of $`\mu `$ and $`T`$ on la–spacetime coordinates $`u`$. Our thermodynamic systems will be considered in local equilibrium in a vicinity of a point $`u_{[0]}`$ with respect to a rest frame locally adapted to N–connection structure, like (9) and (10). In the simplest case the $`n+m`$ splitting is trivially given by a N–connection with vanishing curvature (11). For such conditions of trivial la–spacetime $`\mu `$ and $`T`$ can be considered as constant values. ### 5.1 Particle Density The formula relating the particle density $`\stackrel{~}{n},`$ temperature $`T,`$ and thermodynamic potential $`\mu `$ is obtained by inserting (55) into (28), with (24). Choosing the locally adapted to the N–connection (9) to be the rest frame, when $`U_\mu =(1,0,\mathrm{},0)`$ the calculus is to be performed as for isotropic $`n_{(a)}`$ dimensional spaces . The integral for the number of particles per unit volume in the rest la–frame is $$\stackrel{~}{n}(\mu ,T)=\frac{1}{\left(2\pi \mathrm{}\right)^{n_{(a)}1}}\mathrm{exp}\left(\frac{\mu }{k_BT}\right)\times p^{\overline{1}}\mathrm{exp}\left(\frac{p^{\overline{1}}}{k_BT}\right)\frac{d^{n_{(a)}1}p}{p^{\overline{1}}},$$ (56) with $$p^{\overline{1}}=\left(\left|\stackrel{}{p}\right|^2+\stackrel{~}{m}^2\right)^{1/2}$$ (57) and depends only on length of $`n_{(a)}1`$ dimensional d–vector $`\stackrel{}{p}.`$ By applying spherical coordinates, when $$d^{n_{(a)}1}p=\left|\stackrel{}{p}\right|^{n_{(a)}2}d\left|\stackrel{}{p}\right|d\mathrm{\Omega }$$ and $`d\mathrm{\Omega }`$ is given by the expression (50), and differentiating on radius $`\rho `$ the well known formula for the volume $`V_{n_{(a)}1}`$ in a $`n_{(a)}1`$ dimensional Euclidean space $$V_{n_{(a)}1}=\frac{\left(n_{(a)}1\right)\pi ^{\left(n_{(a)}1\right)/2}}{\mathrm{\Gamma }[(n_{(a)}+1)/2]}\rho ,$$ (58) where $`\mathrm{\Gamma }`$ is the Euler gamma function, then putting $`\rho =1`$ we get $$𝑑\mathrm{\Omega }=\frac{\left(n_{(a)}1\right)\pi ^{\left(n_{(a)}1\right)/2}}{\mathrm{\Gamma }[(n_{(a)}+1)/2]}.$$ (59) We note that from (57) one follows $$\left|\stackrel{}{p}\right|d\left|\stackrel{}{p}\right|=p^{\overline{1}}dp^{\overline{1}}$$ and we can transform (56) into an integral with respect to the angular variables through the replacement $$\frac{d^{n_{(a)}1}p}{p^{\overline{1}}}\frac{\left(n_{(a)}1\right)\pi ^{\left(n_{(a)}1\right)/2}}{\mathrm{\Gamma }[(n_{(a)}+1)/2]}\left|\stackrel{}{p}\right|^{n_{(a)}3}dp^{\overline{1}}.$$ Introducing the dimensionless quantities $$\xi =p^{\overline{1}}/k_BT,\text{ and }\vartheta =\stackrel{~}{m}/k_BT,$$ (60) for which, respectively, $$dp^{\overline{1}}=k_BTd\xi ,\text{ and }\left|\stackrel{}{p}\right|=k_BT\sqrt{\xi ^2\vartheta ^2},$$ (61) the integral (56) is computed $`\stackrel{~}{n}(\mu ,T)`$ $`=`$ $`2^{n_{(a)}/2}\left({\displaystyle \frac{\pi \stackrel{~}{m}}{2\pi \mathrm{}}}\right)^{n_{(a)}1}\left({\displaystyle \frac{k_BT}{\stackrel{~}{m}}}\right)^{\left(n_{(a)}2\right)/2}`$ $`\times `$ $`\mathrm{exp}\left({\displaystyle \frac{\mu }{k_BT}}\right)K_{(n_{(a)}+1)/2}\left({\displaystyle \frac{\stackrel{~}{m}}{k_BT}}\right),`$ where the modified Bessel function of the second kind of order $`n_{(a)}/2`$ has the integral representation $$K_s\left(\eta \right)=\frac{\sqrt{\pi }}{\left(2\eta \right)^s\mathrm{\Gamma }\left(s+\frac{1}{2}\right)}_\eta ^{\mathrm{}}e^q\left(q^2\eta ^2\right)^{s1/2}𝑑q.$$ (63) We note the dependence of (5.1) (but not of (61)) on $`m`$ anisotropic parameters (coordinates). The formulas proved in this subsection transforms into locally anisotropic ones if the N–connection is fixed to be trivial (with vanishing N–connection curvature (11)) and the d–metric (2.2) transforms into a usual (pseudo) Riemannian one. ### 5.2 Average energy and pressure The average energy per particle, for a system in equilibrium, can be calculated by introducing the distribution function (55) into (30) and applying the formulas (25), (29) and (5.1). In terms of dimensionless variables (60) we have $`\epsilon (\mu ,T)`$ $`=`$ $`{\displaystyle \frac{\left(k_BT\right)^{n_{(a)}}}{\left(2\pi \mathrm{}\right)^{n_{(a)}1}}}{\displaystyle \frac{\left(n_{(a)}1\right)\pi ^{\left(n_{(a)}1\right)/2}}{\mathrm{\Gamma }[(n_{(a)}+1)/2]}}`$ $`\times `$ $`\mathrm{exp}\left({\displaystyle \frac{\mu }{k_BT}}\right){\displaystyle _\vartheta ^{\mathrm{}}}\left(\xi ^2\vartheta ^2\right)^{(n_{(a)}3)/2}\xi ^2𝑑\xi .`$ After carrying out partial integrations together with applications of the formula $$\left(q^2\zeta ^2\right)^{\left(s2\right)/2}=\frac{1}{s}\frac{d}{dq}\left[\left(q^2\zeta ^2\right)^{s/2}\right]$$ we obtain $`\epsilon \left(\vartheta \right)`$ $`=`$ $`\pi ^{\left(n_{(a)}2\right)/2}{\displaystyle \frac{\stackrel{~}{m}^{n_{(a)}}}{\left(2\pi \mathrm{}\right)^{n_{(a)}1}}}\left({\displaystyle \frac{2}{\vartheta }}\right)^{n_{(a)}/2}`$ $`\times `$ $`\mathrm{exp}\left({\displaystyle \frac{\mu }{k_BT}}\right)\left[\vartheta K_{(n_{(a)}+2)/2}\left(\vartheta \right)K_{n_{(a)}/2}\left(\vartheta \right)\right].`$ Dividing the energy density (5.2) to the particle density (56) and substituting the inverse dimensionless temperature $`\vartheta =\vartheta \left(T\right)`$ (60) we get the energy per particle (30) $$e\left(T\right)=k_BT\left(\frac{\stackrel{~}{m}}{k_BT}\frac{K_{(n_{(a)}+2)/2}\left(\frac{\stackrel{~}{m}}{k_BT}\right)}{K_{n_{(a)}/2}\left(\frac{\stackrel{~}{m}}{k_BT}\right)}1\right).$$ (65) This formula is the thermal equation of state of a equilibrium system having $`m`$ anisotropy parameters (we remember that $`n_{(a)}=n+m)`$ with respect to a la–frame (9). The pressure d–tensor is to be computed by substituting (55) into (31), applying also the integral (25). We get $$P^{\alpha \beta }=k_BT\stackrel{~}{n}(\mu ,T)\mathrm{\Delta }^{\alpha \beta },$$ (66) were, by definition, the coefficients before $`\mathrm{\Delta }^{\alpha \beta }`$ determine the pressure $$\stackrel{~}{P}=k_BT\stackrel{~}{n}(\mu ,T).$$ (67) So, for a system of $`n_{(a)}1=n+m1`$ dimensions the expression (67) defines the equation of state of ideal gas of particles with respect to a la–frame, having $`m`$ anisotropic parameters. ### 5.3 Enthalpy, specific heats and entropy The average enthalpy per particle $$h\left(T\right)=e\left(T\right)+\stackrel{~}{P}(\mu ,T)/\stackrel{~}{n}(\mu ,T)$$ (68) is computed directly by substituting in this formula the values (65),(67) and (5.1). The result is $$h\left(T\right)=\stackrel{~}{m}\frac{K_{(n_{(a)}+2)/2}\left(\frac{\stackrel{~}{m}}{k_BT}\right)}{K_{n_{(a)}/2}\left(\frac{\stackrel{~}{m}}{k_BT}\right)}.$$ (69) Using (65) and (69) we can compute respectively the specific heats at constant pressure and at constant volume $$c_p=\left(\frac{h}{T}\right)_p\text{ and }c_v=\left(\frac{e}{T}\right)_v.$$ Subtracting of $`h`$ and $`e,`$ after carrying out the differentiation with respect to $`T,`$ one finds Mayer’s relation, $$c_pc_v=k_B$$ and, introducing the adiabatic constant $`\gamma =c_p/c_v,`$ $$\frac{\gamma }{\gamma 1}=\vartheta ^2+\left(n_{(a)}+1\right)\frac{h}{k_BT}\left(\frac{h}{k_BT}\right)^2.$$ (70) The relation (70) can be proven by straightforward calculations by using the properties of Bessel’s function (63) (see a similar proof for isotropic spacetimes in ). The entropy per particle is introduced as in the isotropic case $$s=\frac{h\mu }{T}.$$ (71) We have to insert the values of $`\mu `$ (found from (5.1) $$\mu (\stackrel{~}{n},T)=k_BT\times \mathrm{ln}\frac{\left(2\pi \mathrm{}\right)^{n_{(a)}1}\stackrel{~}{n}}{(2\stackrel{~}{m})^{n_{(a)}/2}(\pi k_BT)^{(n_{(a)}2)/2}K_{n_{(a)}/2}\left(\frac{\stackrel{~}{m}}{k_BT}\right)}$$ (72) and of $`h`$ (see (69) in order to obtain the dependence $`s=s\left(T\right)`$ (for simplicity, we omit this cumbersome formula). ### 5.4 Low and high energy limits Let us investigate the non–relativistic locally anisotropic limit by applying the asymptotic formula $$K_s\left(\vartheta \right)\sqrt{\frac{\pi }{2\vartheta }}e^\vartheta $$ valid for large $`\vartheta .`$ The formula (65) gives $$e\stackrel{~}{m}+\frac{n+m1}{2}k_BT.$$ Thus, each isotropic and anisotropic dimension contributes with $`k_BT/2`$ to the average energy per particle. Now, we consider high temperatures. For $`\vartheta 0`$ one holds the asymptotic formula $$K_s\left(\vartheta \right)2^{s1}\mathrm{\Gamma }\left(s\right)\vartheta ^s$$ and the particle number density (5.1), the energy (65), enthalpy (69) and entropy (71) can be approximated respectively (we state explicitly the dimensions $`n+m)`$ by $`\stackrel{~}{n}(\mu ,T)`$ $`=`$ $`2^{n+m1}\pi ^{n+m2}\mathrm{\Gamma }\left({\displaystyle \frac{n+m}{2}}\right)\times \left({\displaystyle \frac{k_BT}{2\pi \mathrm{}}}\right)^{n+m1}\mathrm{exp}\left({\displaystyle \frac{\mu }{k_BT}}\right),`$ (73) $`e`$ $`=`$ $`\left(n+m1\right)k_BT,h=\left(n+m\right)k_BT,`$ $`s`$ $`=`$ $`\left(n+m\right)k_B.`$ In consequence, the corresponding specific heats and adiabatic constant are $$c_p=\left(n+m\right)k_B,c_v=\left(n+m1\right)k_B\text{ and }\gamma =\frac{n+m}{n+m1}.$$ These formulas imply that for large $`T`$ the spacetime anisotropy (very possible at the beginning of our Universe) could modify substantially the thermodynamic parameters. Finally we emphasize that in the ultrarelativistic limit $$s=2^{n+m1}\pi ^{n+m2}\left(n+m\right)\times \mathrm{\Gamma }\left(\frac{n+m}{2}\right)\left(\frac{k_BT}{2\pi \mathrm{}}\right)^{n+m1}\frac{k_B}{\stackrel{~}{n}}.$$ (74) We also note that our treatment was based on Maxwell–Boltzmann instead of Bose–Einstein statistics. ## 6 Linearized Locally Anisotropic <br>Transport Theory For systems with local anisotropy outside equilibrium the distribution function $`f_{(eq)}(u,p)`$ (55) must be generalized to another one, $`f(u,p),`$ solving the kinetic la–equation (41). We follow a standard procedure of linearization and construction of solutions of kinetic equations by generalizing to la–spacetimes the Chapman–Enskog approach (we shall extend to locally anisotropic backgrounds the results presented in ). We write $$f(u,p)=f_{[0]}(u,p)\left[1+\phi (u,p)\right]$$ (75) with the lowest order of approximation to $`f`$ taken similarly to (55) $$f_{[0]}(u,p)=\frac{1}{(2\pi \mathrm{})^{n_{(a)}1}}\times \mathrm{exp}\left(\frac{\mu (u)p^iU_i(u)p^aU_a(u)}{k_BT(u)}\right)$$ (76) where the constant variables of a trivial equilibrium $`\mu ,T`$ and $`U_\alpha `$ are changed by some local conterparts $`\mu \left(u\right),T\left(u\right)`$ and $`U_\alpha \left(u\right).`$ Outside equilibrium the first two dependencies $`\mu \left(u\right)`$ and $`T\left(u\right)`$ are defined respectively from relations (72) and (65). ### 6.1 Linearized Transport Equations The so–called deviation function $`\phi (u,p)`$ from (75) describes the deformation by nonequilibrium flows of $`f_{[0]}(u,p)`$ into $`f(u,p).`$ It is considered that in equilibrium $`\phi `$ vanishes and not too far from equilibrium states it must be small. The Chapman–Enskog method states that after substituting $`f_{[0]}`$ into the left side and $`f_{[0]}\left(1+\phi \right)`$ into the right hand of (41) we shall neglect the quadratic and higher terms in $`\phi .`$ In result we obtain a linearized equation for the deviation function (for simplicity, hereafter we shall not point to the explicit dependence of functions, kinetic and thermodynamic values on spacetime coordinates) $$p^\alpha \widehat{D}_\alpha f_{[0]}=f_{[0]}L[\phi ]$$ where, having introduced (76) into (4.1), we write for the right side $$L[\phi ]=p^\alpha \widehat{D}_\alpha \left(\frac{\mu p^iU_ip^aU_a}{k_BT}\right).$$ The left side is to be computed by applying the generalized Cartan derivative (36). We introduce a locally anisotropic generalization of the gradient operator by considering the operator (in brief, la–gradient) $$_\alpha =\mathrm{}_\alpha ^\beta D_\beta $$ (77) with $`\mathrm{}_\alpha ^\beta `$ given by (32) (we use the a d–covariant derivative defined by a d–connection (13) instead of partial and/or isotropic derivatives for isotropic spaces ). In a local rest frame of the isotropic fluid $`_\alpha (0,_1,_2,\mathrm{}),`$ i.e. in the flat isotropic space the la–gradient reduces to the ordinary space gradient. By applying $`_\alpha `$ we can eliminate the time la–derivatives $`k_BTL[\phi ]`$ $`=`$ $`p^\alpha p^\beta _\alpha U_\beta Tp^\alpha _\alpha \left({\displaystyle \frac{\mu }{T}}\right)p^\alpha p^\beta U_\alpha \left({\displaystyle \frac{_\beta T}{T}}{\displaystyle \frac{_\beta p}{h\stackrel{~}{n}}}\right)`$ $`+[\left(\gamma 1\right)p^\alpha U_\alpha +T^2\left(\gamma 1\right){\displaystyle \frac{}{T}}\left({\displaystyle \frac{\mu }{T}}\right)+\stackrel{~}{n}\left({\displaystyle \frac{\mu }{n}}\right)]p^\epsilon U_\epsilon ^\upsilon U_\upsilon .`$ We can verify, using the expression $`\mu =\mu (\stackrel{~}{n},T)`$ from (72) that one holds the equalities $$\frac{}{T}\left(\frac{\mu }{T}\right)=\frac{e}{T^2},\frac{\mu }{\stackrel{~}{n}}=\frac{k_BT}{\stackrel{~}{n}}$$ and $$\frac{1}{\stackrel{~}{n}}_\alpha p=h\frac{_\alpha T}{T}+T_\alpha \left(\frac{\mu }{T}\right).$$ Introducing these expressions into (6.1) we get $`k_BT`$ $`L[\phi ]=\mathrm{\Xi }X+\left(hp^\alpha U_\alpha \right)\mathrm{}_\epsilon ^\beta p_\beta X^\beta `$ $`+\left(\mathrm{}_\epsilon ^\beta \mathrm{}_\gamma ^\alpha {\displaystyle \frac{1}{n+m1}}\mathrm{}_{\epsilon \gamma }\mathrm{}^{\beta \alpha }\right)p_\beta p_\alpha {}_{}{}^{(0)}X_{}^{\epsilon \gamma },`$ where $$\mathrm{\Xi }=\left(\frac{n+m}{n+m1}\gamma \right)\left(p^\alpha U_\alpha \right)^2+\left[\left(\gamma 1\right)h\gamma k_BT\right]p^\alpha U_\alpha \frac{\stackrel{~}{m}^2}{n+m1}$$ (79) and there were considered forces deriving the system towards equilibrium $$X=^\mu U_\mu ,X^\alpha =\frac{^\alpha T}{T}\frac{^\alpha p}{\stackrel{~}{n}h}$$ and $${}_{}{}^{(0)}X_{}^{\epsilon \gamma }=\frac{1}{2}\left(^\epsilon U^\gamma +^\gamma U^\epsilon \right)\frac{1}{n+m1}\mathrm{}^{\epsilon \gamma }^\mu U_\mu .$$ We note that the values $`k_BTL[\phi ],\mathrm{\Xi }`$ and $`{}_{}{}^{(0)}X_{}^{\epsilon \gamma }`$ depend explicitly on the dimensions of the base subspace, $`n,`$ and of the fiber subspace, $`m.`$ The anisotropy also modify both the thermodynamic and kinetic values via operators $`^\mu `$ and $`\mathrm{}_\epsilon ^\beta `$ which depend explicitly on d–metric and d–connection coefficients. ### 6.2 On the solution of locally anisotropic transport equations Let us suppose that is known the solution of these three equations: $`k_BTL[A\left(p\right)]`$ $`=`$ $`\mathrm{\Xi },`$ (80) $`k_BTL[B\left(p\right)\mathrm{}_\alpha ^\beta p_\beta ]`$ $`=`$ $`\left(hp^\alpha U_\alpha \right)\mathrm{}_\alpha ^\beta p_\beta `$ and $$k_BTL\left[C\left(p\right)\left(\mathrm{}_\alpha ^\beta \mathrm{}_\epsilon ^\sigma \frac{1}{n+m1}\mathrm{}_{\alpha \epsilon }\mathrm{}^{\beta \sigma }\right)p_\beta p_\sigma \right]$$ $$=\left(\mathrm{}_\alpha ^\beta \mathrm{}_\epsilon ^\sigma \frac{1}{n+m1}\mathrm{}_{\alpha \epsilon }\mathrm{}^{\beta \sigma }\right)p_\beta p_\sigma $$ for some functions $`A\left(p\right),B\left(p\right)`$ and $`C\left(p\right).`$ Because the integral operator $`L`$ is linear the linear combination of thermodynamic forces $$\phi =AX+B_\alpha X^\alpha +C_{\alpha \beta }{}_{}{}^{(0)}X_{}^{\alpha \beta },$$ where $$B_\alpha =B\left(p\right)\mathrm{}_\alpha ^\beta p_\beta $$ and $$C_{\alpha \beta }=B\left(p\right)\left(\mathrm{}_\alpha ^\epsilon \mathrm{}_\beta ^\sigma \frac{1}{n+m1}\mathrm{}_{\alpha \beta }\mathrm{}^{\epsilon \sigma }\right)p_\epsilon p_\sigma ,$$ is a solution for the deviation function $`\phi .`$ It can also be verified that every function of the form $`\phi =a+b_\alpha p^\alpha `$ with some parameters $`a`$ and $`b_\alpha `$ not depending on $`p^\alpha `$ is a solution of the homogeneous equation $`L\left[\phi \right]=0.`$ In consequence, it was proved (see and ) that the scalar parts of $`A\left(p\right)`$ and $`B\left(p\right)`$ are determined respectively up to functions of the forms $`\phi =a+b_\alpha p^\alpha `$ and $`b^\alpha p_\alpha .`$ We emphasize that solutions of the transport equations (80) (see next subsections) will be expressed in terms of $`A,B,`$ and $`C.`$ ### 6.3 Linear laws for locally anisotropic non–equilibrium <br>thermodynamics Let $`U_\beta \left(u\right)`$ be a velocity field of Landau–Lifshitz type characterizing a locally anisotropic fluid flow. The heat flow and the viscous pressure d–tensor with respect to a such la–field are correspondingly defined similarly to but in terms of objects on la–spacetime (see formulas (69), (66), (32), (31), (25), (24), (13), (9), (10), (2.2) $`I_{(heat)}^\alpha `$ $`=`$ $`\left(U_\alpha \mathrm{{\rm Y}}^{\alpha \beta }hn^\beta \right)\mathrm{}_\beta ^\epsilon ,`$ (81) $`\mathrm{\Pi }^{\alpha \beta }`$ $`=`$ $`P^{\alpha \beta }+p\mathrm{}^{\alpha \beta }.`$ It should be noted that if the Landau–Lifshitz condition (27) is satisfied the first term in $`I_{(heat)}^\nu `$ vanishes and the heat flow is the enthalpy carried away by the particles. The pressure $`P\left(u\right)`$ is defined by $`\stackrel{~}{n}\left(u\right)k_BT\left(u\right),`$ see (67). Inserting (75) and (76) into (81) we prove (see locally isotropic cases in ) the linear laws of locally anisotropic non–equilibrium thermodynamics $$I_{(heat)}^\nu =\lambda TX^\nu $$ and $$\mathrm{\Pi }^{\alpha \beta }=2\eta {}_{}{}^{(0)}X_{}^{\alpha \beta }\eta _{(v)}X\mathrm{}^{\epsilon \sigma }$$ with the transport coefficients (the heat conductivity $`\lambda ,`$ the shear viscosity $`\eta ,`$ and the volume viscosity coefficient $`\eta _{(v)};`$ in order to compare with formulas from we shall introduce explicitly the light velocity constant $`c)`$ defined as $`\lambda `$ $`=`$ $`{\displaystyle \frac{c}{(n+m1)T}}{\displaystyle p^\sigma B_\epsilon \mathrm{}_\sigma ^\epsilon \left(hp^\nu U_\nu \right)f_{[0]}𝑑\varsigma _p},`$ $`\eta _{(v)}`$ $`=`$ $`{\displaystyle \frac{c}{n+m1}}{\displaystyle p_\epsilon p_\sigma \mathrm{}^{\epsilon \sigma }Af_{[0]}𝑑\varsigma _p},`$ (82) $`\eta `$ $`=`$ $`{\displaystyle \frac{c}{(n+m)\left(n+m4\right)}}{\displaystyle p^\sigma p^\epsilon C_{\nu \mu }\mathrm{}_{\sigma \epsilon }^{\nu \mu }f_{[0]}𝑑\varsigma _p},`$ where one uses the d–tensor $$\mathrm{}_{\sigma \epsilon }^{\nu \mu }=\frac{1}{2}\left(\mathrm{}_\sigma ^v\mathrm{}_\epsilon ^\mu \mathrm{}_\epsilon ^v\mathrm{}_\sigma ^\mu \right)\frac{1}{n+m1}\mathrm{}^{\nu \mu }\mathrm{}_{\sigma \epsilon }$$ with the properties $$\mathrm{}_{\rho \sigma }^{\nu \mu }\mathrm{}_{\varphi \epsilon }^{\rho \sigma }=\mathrm{}_{\varphi \epsilon }^{\nu \mu }$$ and $$\mathrm{}_{\rho \sigma }^{\rho \sigma }=1+\frac{(n+m)(n+m1)}{2}.$$ The main purpose of non–equilibrium thermodynamics is the calculation of transport coefficients. With respect to la–frames the formulas are quite similar with those for isotropic spaces with that difference that we have to consider the values as d–tensors and take into account the number $`m`$ of anisotropic variables. On both type of locally isotropic and anisotropic spacetimes one holds the so–called conditions of fit (see, for instance, ) $$p_\epsilon U^\epsilon Af_{[0]}𝑑\varsigma _p=0\text{ and }\left(p_\epsilon U^\epsilon \right)^2Af_{[0]}𝑑\varsigma _p=0$$ which allow us to write the volume viscosity (see (79)) $$\eta _{(v)}=c\mathrm{\Xi }Af_{[0]}𝑑\varsigma _p.$$ For some d–tensors $`H^{\alpha _1,\mathrm{}\alpha _q}\left(p\right)`$ and $`S_{\alpha _1,\mathrm{}\alpha _q}(p)`$ we define the symmetric bracket $$\{H,S\}=\frac{1}{\stackrel{~}{n}^2}H^{\alpha _1,\mathrm{}\alpha _q}\left(p\right)L\left[S_{\alpha _1,\mathrm{}\alpha _q}(p)\right]f_{[0]}𝑑\varsigma _p$$ where $`\stackrel{~}{n}`$ is the particle density and $`L`$ is a linearized operator. In terms of such brackets the coefficients (82) can be rewritten in an equivalent form $`\lambda `$ $`=`$ $`{\displaystyle \frac{ck_B\stackrel{~}{n}^2}{n+m1}}\{B^\alpha ,B_\alpha \},`$ (83) $`\eta `$ $`=`$ $`{\displaystyle \frac{ck_BT\stackrel{~}{n}^2}{(n+m1)^23}}\{C^{\alpha \beta },C_{\alpha \beta }\},`$ $`\eta _{(v)}`$ $`=`$ $`ck_BT\stackrel{~}{n}^2\{A,A\}.`$ The addition of invariants of type $`\phi =a+b_\alpha p^\alpha `$ to some solutions for $`A\left(p\right),`$ $`B_\alpha \left(p\right),`$ or $`C_{\alpha \beta }\left(p\right)`$ does not change the values of the transport coefficients. ### 6.4 Integral and algebraic equations The solutions of transport equations (80) are approximated by considering power series on $`\zeta =p^\alpha U_\alpha /k_BT`$ for functions $$A\left(\zeta \right)=\underset{\widehat{a}=2}{\overset{\mathrm{}}{}}A_{\widehat{a}}\zeta ^{\widehat{a}},B\left(\zeta \right)=\underset{\widehat{b}=1}{\overset{\mathrm{}}{}}B_{\widehat{b}}\zeta ^{\widehat{b}},C\left(\zeta \right)=\underset{\widehat{c}=0}{\overset{\mathrm{}}{}}C_{\widehat{c}}\zeta ^{\widehat{c}}.$$ The starting values $`\widehat{a}=2`$ and $`\widehat{b}=1`$ have been introduced with the aim to determine the scalar functions $`A\left(\zeta \right)`$ and $`B\left(\zeta \right),`$ respectively, up to contributions of the forms $`a+b_\alpha p^\alpha `$ and $`b^\alpha p_\alpha .`$ Inserting these power series into the integral equations (80), multiplying respectively on $`\zeta ^{\widehat{s}}f_{[0]},\zeta ^{\widehat{s}}p^\alpha f_{[0]},`$ and $`\zeta ^{\widehat{s}}p^\alpha p^\beta f_{[0]},`$ after integrating on $`d\varsigma _p`$ we get $$\underset{\widehat{a}=2}{\overset{\mathrm{}}{}}a_{\widehat{a}_1\widehat{a}}A_{\widehat{a}}=\alpha _{\widehat{a}_1},\underset{\widehat{b}=1}{\overset{\mathrm{}}{}}b_{\widehat{b}_1\widehat{b}}B_{\widehat{b}}=\beta _{\widehat{b}_1},\underset{\widehat{c}=0}{\overset{\mathrm{}}{}}c_{\widehat{c}_1\widehat{c}}C_{\widehat{c}}=\gamma _{\widehat{c}_1},$$ where $`\widehat{a}_1=2,3,\mathrm{},\widehat{b}_1=1,2,\mathrm{},\widehat{c}_1=0,1,\mathrm{},`$ and there are symmetric brackets $`a_{\widehat{a}_1\widehat{a}}`$ $`=`$ $`\{\zeta ^{\widehat{a}_1},\zeta ^{\widehat{a}}\},`$ (84) $`b_{\widehat{b}_1\widehat{b}}`$ $`=`$ $`\{\zeta ^{\widehat{b}_1}p^\alpha ,\zeta ^{\widehat{b}}\mathrm{}_\alpha ^vp_\nu \},`$ $`c_{\widehat{c}_1\widehat{c}}`$ $`=`$ $`\{\zeta ^{\widehat{c}_1}p^\alpha p^\beta ,\zeta ^{\widehat{c}}\mathrm{}_{\alpha \beta }^{v\mu }p_\nu p_\mu \}`$ and integrals $`\alpha _{\widehat{a}_1}`$ $`=`$ $`{\displaystyle \frac{\zeta ^{\widehat{a}_1}\mathrm{\Xi }f_{[0]}}{k_BT\stackrel{~}{n}^2}𝑑\varsigma _p},`$ (85) $`\beta _{\widehat{b}_1}`$ $`=`$ $`{\displaystyle \frac{\zeta ^{\widehat{b}_1}\left(hp^\alpha U_\alpha \right)\mathrm{}_{\mu \nu }p^\mu p^\nu f_{[0]}}{k_BT\stackrel{~}{n}^2}𝑑\varsigma _p},`$ $`\gamma _{\widehat{c}_1}`$ $`=`$ $`{\displaystyle \frac{\zeta ^{\widehat{c}_1}\mathrm{}_{\alpha \beta }^{v\mu }p_\nu p_\mu p^\alpha p^\beta f_{[0]}}{k_BT\stackrel{~}{n}^2}𝑑\varsigma _p}.`$ The lowest approximation is given by the coefficients $$A_{\widehat{2}}=\alpha _{\widehat{2}}/a_{\widehat{2}\widehat{2}},B_{\widehat{1}}=\beta _{\widehat{1}}/b_{\widehat{1}\widehat{1}},C_{\widehat{0}}=\gamma _{\widehat{0}}/c_{\widehat{0}\widehat{0}}.$$ Introducing these values into (83) we obtain the first–order approximations to the transport coefficients $`\lambda `$ $`=`$ $`{\displaystyle \frac{ck_B\stackrel{~}{n}^2}{n+m1}}{\displaystyle \frac{\beta _{\widehat{1}}^2}{a_{\widehat{2}\widehat{2}}}},`$ (86) $`\eta `$ $`=`$ $`{\displaystyle \frac{ck_BT\stackrel{~}{n}^2}{\left(n+m\right)\left(n+m1\right)2}}{\displaystyle \frac{\gamma _{\widehat{0}}^2}{c_{\widehat{0}\widehat{0}}}},`$ $`\eta _{(v)}`$ $`=`$ $`ck_BT\stackrel{~}{n}^2{\displaystyle \frac{\alpha _{\widehat{2}}^2}{a_{\widehat{2}\widehat{2}}}}.`$ The values $`\alpha _{\widehat{2}},\beta _{\widehat{1}}`$ and $`\gamma _{\widehat{0}}`$ are $`\left(n+m1\right)`$–fold integrals which can be expressed in terms of enthalpy $`h`$ and temperature $`T,`$ $`\alpha _{\widehat{2}}`$ $`=`$ $`{\displaystyle \frac{n+m+1\left(n+m1\right)\gamma }{ck_BT\stackrel{~}{n}}}{\displaystyle \frac{\left(n+m1\right)\gamma }{c\stackrel{~}{n}}}\underset{T\mathrm{}}{\underset{}{}}0,`$ (87) $`\beta _{\widehat{1}}`$ $`=`$ $`{\displaystyle \frac{k_BT}{c\stackrel{~}{n}}}{\displaystyle \frac{(n+m1)\gamma }{\gamma 1}}\underset{T\mathrm{}}{\underset{}{}}{\displaystyle \frac{k_BT}{c\stackrel{~}{n}}}\left(n+m1\right)\left(n+m\right),`$ $`\gamma _{\widehat{0}}`$ $`=`$ $`{\displaystyle \frac{k_BT}{c^2\stackrel{~}{n}}}\left(n+m2\right)\left(n+m+1\right)h\underset{T\mathrm{}}{\underset{}{}}{\displaystyle \frac{1}{\stackrel{~}{n}}}\left({\displaystyle \frac{k_BT}{c}}\right)^2\left[(n+m)^21\right]\left(n+m\right).`$ In order to complete the calculus of the transport coefficients it is necessary to compute the brackets from the denominators of (86). ### 6.5 Brackets from flat to la–spacetimes The calculation of brackets is a quite tedious task (see ) which should involve the curved spacetime metric and connection. For simplicity, we consider a background flat spacetime with trivial local anisotropy. In this case we can apply directly the formulas proved for the Boltzmann theory in $`n_{(a)}`$ dimension but with respect to la–frames and by introducing the locally anisotropic d–connections and d–metrics instead of their isotropic analogous. Let us consider a locally adapted to N–connection rest frame, $`\stackrel{}{e}`$ being a unit vector and denote by $`(\mathrm{\Theta }_{n_{(a)}1},\mathrm{\Theta }_{n_{(a)}2},\mathrm{},\mathrm{\Theta }_1)`$ the angles of spherical coordinates with respect to a Cartesian one $`(X^1,X^2,\mathrm{},X^{n_{(a)}})`$ chosen as the $`X^{n_{(a)}}`$–axis to be along $`\stackrel{}{p}_{CM}`$ (where $`|\stackrel{}{p}_{CM}|=F/\stackrel{~}{P},\stackrel{~}{P}=|P^\alpha |,`$ see (46) (52), and $`\stackrel{}{p}_{CM}`$ $`\stackrel{}{e}`$ in the plane of $`X^{n_{(a)}}`$ and $`X^{n_{(a)}1}`$ axes. We suppose that the scattering angle is $`\mathrm{\Theta }_{n_{(a)}1}`$ and the differential cross section $`\left(d\sigma /d\mathrm{\Omega }\right)_{CM}`$ is a function only on the module of total energy and on scattering angle $`\mathrm{\Theta }=\mathrm{\Theta }_{n_{(a)}1},`$ $$\left(\frac{d\sigma }{d\mathrm{\Omega }}\right)_{CM}=\left(\frac{d\sigma }{d\mathrm{\Omega }}\right)_{CM}(\stackrel{~}{P},\mathrm{\Theta }).$$ The results of calculations of symmetric brackets will be expressed into terms of twofold integrals, which for arbitrary integers $`J_{\widehat{r}\widehat{s}}^{(\widehat{i},\widehat{j},\widehat{k})}`$ $`(\zeta ,\widehat{t},\widehat{u},\widehat{v},\widehat{w})=\nu \left(\zeta \right)\left(\begin{array}{c}\widehat{r}\\ \widehat{t}\end{array}\right)\left(\begin{array}{c}\widehat{s}\\ \widehat{u}\end{array}\right)\left(\begin{array}{c}\widehat{d}/2\\ \widehat{v}\end{array}\right){\displaystyle \frac{2^{\widehat{b}+\widehat{v}\widehat{d}}}{\sqrt{\pi }}}`$ $`\times `$ $`\mathrm{\Gamma }\left[\widehat{b}+\widehat{v}+{\displaystyle \frac{1}{2}}\right]F(\widehat{t},\widehat{u},\widehat{v}){\displaystyle \underset{2\zeta }{\overset{\mathrm{}}{}}}𝑑\chi \chi ^{(n+m1+\widehat{d}\widehat{u}\widehat{b}\widehat{v}+\widehat{i})}`$ $`\times `$ $`\left({\displaystyle \frac{\chi ^2}{4}}\zeta ^2\right)^{\left(n+m2+\widehat{t}+\widehat{u}+\widehat{j}\right)/2}K_{\widehat{b}+\widehat{v}\widehat{h}}\left(\zeta \right)\left[\delta _{0\widehat{w}}\left(\begin{array}{c}\widehat{u}\\ \widehat{w}\end{array}\right)\mathrm{sin}^{\widehat{w}}\mathrm{\Theta }\mathrm{cos}^{\widehat{u}\widehat{w}}\mathrm{\Theta }\right]`$ $`\times `$ $`{\displaystyle \underset{2\zeta }{\overset{\pi }{}}}𝑑\mathrm{\Theta }\mathrm{sin}^{n+m1}\mathrm{\Theta }\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_{CM}(\stackrel{~}{P},\mathrm{\Theta }),`$ where there are introduced the integers $$\widehat{d}=\widehat{r}+\widehat{s}\widehat{t}\widehat{u},\widehat{b}=\left(\widehat{t}+\widehat{u}+n+m2\right)/2,\widehat{h}=[1\left(1\right)^{\widehat{d}}]$$ and variables $`\chi =c/\left(k_BT\right)`$ and $`\zeta =\stackrel{~}{m}c^2/\left(k_BT\right).`$ The factors in (6.5) are defined $$\nu \left(\zeta \right)=\frac{\pi ^{\left(n+m3\right)/2}}{2\mathrm{\Gamma }\left((n+m1)/2\right)\mathrm{\Gamma }\left(n+m3\right)\zeta ^2\left[K_{\frac{n+m}{2}}\left(\zeta \right)\right]^2},$$ and $`F(\widehat{t},\widehat{u},\widehat{w})`$ $`={\displaystyle \frac{1}{4}}\left[1+\left(1\right)^{\widehat{w}}\right]\left[1+\left(1\right)^{\widehat{t}+\widehat{u}\widehat{w}}\right]`$ $`\times B({\displaystyle \frac{n+m3}{2}},{\displaystyle \frac{\widehat{w}+1}{2}})\times B({\displaystyle \frac{n+m2+\widehat{w}}{2}},{\displaystyle \frac{\widehat{t}+\widehat{u}\widehat{w}+1}{2}}),`$ where the beta function is given by gamma functions, $$B(x,y)=\frac{\mathrm{\Gamma }\left(x\right)\mathrm{\Gamma }\left(y\right)}{\mathrm{\Gamma }\left(x+y\right)},$$ and by $`\left(\begin{array}{c}\widehat{r}\\ \widehat{t}\end{array}\right)`$ it is denoted the Newton’s binomium. #### 6.5.1 Scalar type brackets The first type of brackets necessary for calculation of transport coefficients (83) (see the series approximation (84) are the so–called the scalar brackets, decomposed into a sum of two integrals $$a_{\widehat{a}_1\widehat{a}}=a_{\widehat{a}_1\widehat{a}}^{}+a_{\widehat{a}_1\widehat{a}}^{\prime \prime },$$ (98) where $`\widehat{a}_1,\widehat{a}=2,3,\mathrm{},`$ $`a_{\widehat{a}_1\widehat{a}}^{}`$ $`=`$ $`\stackrel{~}{n}^2{\displaystyle f_{[0]}\left(p\right)f_{[0]}\left(p_{[1]}\right)\left(\zeta _p\right)^{\widehat{a}_1}\times [\left(\zeta _{p_{[1]}}\right)^{\widehat{a}}\left(\zeta _{p_{[1]}^{}}\right)^{\widehat{a}}]W𝑑\varsigma _p𝑑\varsigma _{p_{[1]}}𝑑\varsigma _p^{}𝑑\varsigma _{p_{[1]}^{}}},`$ $`a_{\widehat{a}_1\widehat{a}}^{\prime \prime }`$ $`=`$ $`\stackrel{~}{n}^2{\displaystyle f_{[0]}\left(p\right)f_{[0]}\left(p_{[1]}\right)\left(\zeta _p\right)^{\widehat{a}_1}\times [\left(\zeta _p\right)^{\widehat{a}}\left(\zeta _p^{}\right)^{\widehat{a}}]W𝑑\varsigma _p𝑑\varsigma _{p_{[1]}}𝑑\varsigma _p^{}𝑑\varsigma _{p_{[1]}^{}}},`$ with the integration variables $`\zeta _p`$ $`=`$ $`p^\alpha U_\alpha /k_BT,\zeta _{p_{[1]}}=p_{[1]}^\alpha U_\alpha /k_BT,`$ $`\zeta _p^{}`$ $`=`$ $`p^\alpha U_\alpha /k_BT,\zeta _{p_{[1]}^{}}=p_{[1]}^\alpha U_\alpha /k_BT.`$ The explicit calculations of integrals from (98) (see the method and basic intermediar formulas in ; in this paper we deal with la–values) give $`a_{\widehat{a}_1\widehat{a}}^{}`$ $`=`$ $`{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{a}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{a}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{a}_1\widehat{a}}^{(0,0,0)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}),`$ $`a_{\widehat{a}_1\widehat{a}}^{\prime \prime }`$ $`=`$ $`{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{a}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{a}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}J_{\widehat{a}_1\widehat{a}}^{(0,0,0)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}).`$ In the lowest approximation, for $`\widehat{a}_1=\widehat{a}=2,`$ we obtain $$a_{\widehat{2}\widehat{2}}=2\left[J_{\widehat{2}\widehat{2}}^{(0,0,0)}(\widehat{2},\widehat{2},\widehat{0},\widehat{0})+J_{\widehat{2}\widehat{2}}^{(0,0,0)}(\widehat{2},\widehat{2},\widehat{0},\widehat{2})\right].$$ So, we have reduced the scalar brackets to twofold integrals which are expressed in terms o spherical coordinates with respect to a locally anisotropic rest frame. #### 6.5.2 Vector type brackets The vector type brackets from (83) and (84) are also split into two types of integrals $$b_{\widehat{b}_1\widehat{b}}=b_{\widehat{b}_1\widehat{b}}^{}+b_{\widehat{b}_1\widehat{b}}^{\prime \prime }$$ (99) where $`\widehat{b}_1,\widehat{b}=1,2,\mathrm{},`$ $`b_{\widehat{b}_1\widehat{b}}^{}`$ $`=`$ $`\stackrel{~}{n}^2{\displaystyle f_{[0]}\left(p\right)f_{[0]}\left(p_{[1]}\right)\mathrm{}_{\sigma \epsilon }\left(\zeta _p\right)^{\widehat{b}_1}p^\sigma }`$ $`\times \left[p_{[1]}^\epsilon \left(\zeta _{p_{[1]}}\right)^{\widehat{b}}p_{[1]}^\epsilon \left(\zeta _{p_{[1]}^{}}\right)^{\widehat{b}}\right]Wd\varsigma _pd\varsigma _{p_{[1]}}d\varsigma _p^{}d\varsigma _{p_{[1]}^{}},`$ $`b_{\widehat{b}_1\widehat{b}}^{\prime \prime }`$ $`=`$ $`\stackrel{~}{n}^2{\displaystyle f_{[0]}\left(p\right)f_{[0]}\left(p_{[1]}\right)\mathrm{}_{\sigma \epsilon }\left(\zeta _p\right)^{\widehat{b}_1}p^\sigma }`$ $`\times \left[p^\epsilon \left(\zeta _p\right)^{\widehat{b}}p^\epsilon \left(\zeta _p^{}\right)^{\widehat{b}}\right]Wd\varsigma _pd\varsigma _{p_{[1]}}d\varsigma _p^{}d\varsigma _{p_{[1]}^{}}.`$ A tedious calculus similar to that presented in implies further decompositions of coefficients and their representation as $$b_{\widehat{b}_1\widehat{b}}^{}=\underset{\left(i\right)=1}{\overset{3}{}}b_{(i)\widehat{b}_1\widehat{b}}^{}\text{ and }b_{\widehat{b}_1\widehat{b}}^{\prime \prime }=\underset{\left(i\right)=1}{\overset{3}{}}b_{(i)\widehat{b}_1\widehat{b}}^{\prime \prime },$$ with corresponding sums $`b_{(1)\widehat{b}_1\widehat{b}}^{}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left({\displaystyle \frac{k_BT}{c}}\right)^2\times {\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{b}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{b}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{a}_1\widehat{a}}^{(2,0,0)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}),`$ $`b_{(2)\widehat{b}_1\widehat{b}}^{}`$ $`=`$ $`\left({\displaystyle \frac{k_BT}{c}}\right)^2{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{b}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{b}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{a}_1\widehat{a}}^{(0,2,1)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}),`$ $`b_{(2)\widehat{b}_1\widehat{b}}^{}`$ $`=`$ $`\left({\displaystyle \frac{k_BT}{c}}\right)^2{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{b}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{b}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{a}_1+1\widehat{a}+1}^{(0,2,1)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w});`$ with corresponding sum expressions for $`b_{(i)\widehat{b}_1\widehat{b}}^{\prime \prime }`$ by omitting the factor $`\left(1\right)^{\widehat{u}}.`$ In the lowest approximation (for $`\widehat{b}_1=\widehat{b}=1)`$ we have $`b_{\widehat{1}\widehat{1}}`$ $`=`$ $`2\left({\displaystyle \frac{k_BT}{c}}\right)^2\times [J_{\widehat{1}\widehat{1}}^{(0,2,1)}(\widehat{1},\widehat{1},\widehat{0},\widehat{0})`$ $`+J_{\widehat{2}\widehat{2}}^{(0,0,0)}(\widehat{2},\widehat{2},\widehat{0},\widehat{0})+J_{\widehat{2}\widehat{2}}^{(0,0,0)}(\widehat{2},\widehat{2},\widehat{0},\widehat{2})].`$ Here should be noted that in general $`T=T\left(u\right)`$ is a function on la–spacetime coordinates. #### 6.5.3 Tensor type brackets In a similar fashion as for scalar and vector type symmetric brackets from (83) and (84) one holds the decomposition $$c_{\widehat{c}_1\widehat{c}}=c_{\widehat{c}_1\widehat{c}}^{}+c_{\widehat{c}_1\widehat{c}}^{\prime \prime }$$ (100) where $`\widehat{c}_1,\widehat{c}=1,2,\mathrm{},`$ $`c_{\widehat{c}_1\widehat{c}}^{}`$ $`=\stackrel{~}{n}^2{\displaystyle f_{[0]}\left(p\right)f_{[0]}\left(p_{[1]}\right)\mathrm{}_{\sigma \epsilon }^{\nu \mu }\left(\zeta _p\right)^{\widehat{c}_1}p_\nu p_\mu }`$ $`\times [p_{[1]}^\sigma p_{[1]}^\epsilon \left(\zeta _{p_{[1]}}\right)^{\widehat{c}}p_{[1]}^\sigma p_{[1]}^\epsilon \left(\zeta _{p_{[1]}^{}}\right)^{\widehat{c}}]Wd\varsigma _pd\varsigma _{p_{[1]}}d\varsigma _p^{}d\varsigma _{p_{[1]}^{}},`$ $`c_{\widehat{c}_1\widehat{c}}^{\prime \prime }`$ $`=\stackrel{~}{n}^2{\displaystyle f_{[0]}\left(p\right)f_{[0]}\left(p_{[1]}\right)\mathrm{}_{\sigma \epsilon }^{\nu \mu }\left(\zeta _p\right)^{\widehat{c}_1}p_\nu p_\mu }`$ $`\times [p^\sigma p^\epsilon \left(\zeta _{p_{[1]}}\right)^{\widehat{c}}p^\sigma p^\epsilon \left(\zeta _{p_{[1]}^{}}\right)^{\widehat{c}}]Wd\varsigma _pd\varsigma _{p_{[1]}}d\varsigma _p^{}d\varsigma _{p_{[1]}^{}}.`$ For tensor like brackets we have to consider sums on nine terms $$c_{\widehat{c}_1\widehat{c}}^{}=\underset{(s)=1}{\overset{9}{}}c_{(s)\widehat{c}_1\widehat{c}}^{}\text{ and }c_{\widehat{c}_1\widehat{c}}^{\prime \prime }=\underset{(s)=1}{\overset{9}{}}c_{(s)\widehat{c}_1\widehat{c}}^{\prime \prime }.$$ These terms are (with that exception that we have dependencies on the number of anisotropic variables and $`T=T\left(u\right)`$ is a function on la–spacetime coordinates) $`c_{(1)\widehat{c}_1\widehat{c}}^{}`$ $`=`$ $`{\displaystyle \frac{1}{16}}\left({\displaystyle \frac{k_BT}{c}}\right)^4{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{c}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{c}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{c}_1\widehat{c}}^{(4,0,0)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}),`$ $`c_{(2)\widehat{c}_1\widehat{c}}^{}`$ $`=`$ $`\left({\displaystyle \frac{k_BT}{c}}\right)^4{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{c}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{c}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{c}_1\widehat{c}}^{(0,4,2)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}),`$ $`c_{(3)\widehat{c}_1\widehat{c}}^{}`$ $`=`$ $`{\displaystyle \frac{n+m2}{n+m1}}\left({\displaystyle \frac{k_BT}{c}}\right)^4{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{c}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{c}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{c}_1+2\widehat{c}+2}^{(0,0,0)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}),`$ $`c_{(4)\widehat{c}_1\widehat{c}}^{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{k_BT}{c}}\right)^4{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{c}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{c}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{c}_1\widehat{c}}^{(2,2,1)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}),`$ $`c_{(5)\widehat{c}_1\widehat{c}}^{}`$ $`=`$ $`2\left({\displaystyle \frac{k_BT}{c}}\right)^4{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{c}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{c}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{c}_1+1\widehat{c}+1}^{(0,2,1)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}),`$ $`c_{(6)\widehat{c}_1\widehat{c}}^{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{k_BT}{c}}\right)^4{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{c}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{c}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{c}_1+1\widehat{c}+1}^{(2,0,0)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}),`$ $`c_{(7)\widehat{c}_1\widehat{c}}^{}`$ $`=`$ $`{\displaystyle \frac{\zeta ^4}{n+m1}}\left({\displaystyle \frac{k_BT}{c}}\right)^4{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{c}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{c}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{c}_1\widehat{c}}^{(0,0,0)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}),`$ $`c_{(8)\widehat{c}_1\widehat{c}}^{}`$ $`=`$ $`{\displaystyle \frac{\zeta ^4}{n+m1}}\left({\displaystyle \frac{k_BT}{c}}\right)^4{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{c}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{c}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{c}_1+2\widehat{c}}^{(0,0,0)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}),`$ $`c_{(9)\widehat{c}_1\widehat{c}}^{}`$ $`=`$ $`{\displaystyle \frac{\zeta ^4}{n+m1}}\left({\displaystyle \frac{k_BT}{c}}\right)^4{\displaystyle \underset{\widehat{t}=0}{\overset{\widehat{c}_1}{}}}{\displaystyle \underset{\widehat{u}=0}{\overset{\widehat{c}}{}}}{\displaystyle \underset{\widehat{v}=0}{\overset{\widehat{d}/2}{}}}{\displaystyle \underset{\widehat{w}=0}{\overset{\widehat{u}}{}}}\left(1\right)^{\widehat{u}}J_{\widehat{c}_1\widehat{c}+2}^{(0,0,0)}(\widehat{t},\widehat{u},\widehat{v},\widehat{w}).`$ The twice primed values $`c_{(s)\widehat{c}_1\widehat{c}}^{\prime \prime }`$ are given by similar sums by omitting the factor $`\left(1\right)^{\widehat{u}}`$ and by changing into $`c_{(4)\widehat{c}_1\widehat{c}}^{\prime \prime }`$ and $`c_{(5)\widehat{c}_1\widehat{c}}^{\prime \prime }`$ the overall signs. In the lower approximation, for $`\widehat{c}_1=\widehat{c}=0,`$ one holds $`c_{\widehat{0}\widehat{0}}`$ $`=2\left({\displaystyle \frac{k_BT}{c}}\right)^4\{J_{\widehat{0}\widehat{0}}^{(0,4,2)}(\widehat{0},\widehat{0},\widehat{0},\widehat{0})+2J_{\widehat{1}\widehat{1}}^{(0,2,1)}(\widehat{1},\widehat{1},\widehat{0},\widehat{0})`$ $`+{\displaystyle \frac{n+m2}{n+m1}}[J_{\widehat{2}\widehat{2}}^{(0,0,0)}(\widehat{2},\widehat{2},\widehat{0},\widehat{0})+J_{\widehat{2}\widehat{2}}^{(0,0,0)}(\widehat{2},\widehat{2},\widehat{0},\widehat{2})]\}.`$ ### 6.6 Locally anisotropic transport coefficients for a class of cross sections For astrophysical applications we can substitute $$\left(\frac{d\sigma }{d\mathrm{\Omega }}\right)_{CM}=\xi P^r,$$ (101) with some scalar factor $`\xi `$ (with or without dimension) and $`r`$ being a positive or negative number into (6.5). In this subsection we put $`\zeta =\stackrel{~}{m}c^2/\left(k_BT\right)`$ for high values of $`T.`$ The chosen type of differential cross section (101) is used, for instance, for calculations of neutrino–neutrino scattering (when $`r=2`$ and $`\xi `$ is connected with the weak coupling constant). In the first approximation the symmetric brackets (84) are $`a_{\widehat{2}\widehat{2}}`$ $`=\xi \left({\displaystyle \frac{2k_BT}{c}}\right)^r{\displaystyle \frac{\pi ^{(n+m1)/2}\mathrm{\Gamma }\left[n+m+1+r/2\right]\mathrm{\Gamma }\left[\left(n+m+r\right)/2\right]}{\mathrm{\Gamma }\left[n+m2\right]\mathrm{\Gamma }\left[\left(n+m\right)/2\right]\mathrm{\Gamma }\left[\left(n+m+1\right)/2\right]}},`$ $`b_{\widehat{1}\widehat{1}}`$ $`=\left({\displaystyle \frac{k_BT}{c}}\right)^2{\displaystyle \frac{n+m+2+r}{2}}a_{\widehat{2}\widehat{2}},`$ $`c_{\widehat{0}\widehat{0}}`$ $`=\left({\displaystyle \frac{k_BT}{c}}\right)^2\left[{\displaystyle \frac{\left(n+m+r\right)\left(n+m+4+r\right)}{2}}+{\displaystyle \frac{n+m2}{n+m1}}\right]a_{\widehat{2}\widehat{2}}.`$ Putting these values into (87) we obtain the locally anisotropic variant of transport coefficients in the first approximation, when $`\eta _{(v)}0`$ but with nonzero $`\lambda `$ $`=`$ $`{\displaystyle \frac{2k_Bc}{\xi }}\left({\displaystyle \frac{c}{2k_BT}}\right)^r{\displaystyle \frac{\left(n+m\right)^2\left(n+m1\right)}{n+m+2+r}}`$ $`\times {\displaystyle \frac{\mathrm{\Gamma }\left[n+m2\right]\mathrm{\Gamma }\left[\left(n+m\right)/2\right]\mathrm{\Gamma }\left[\left(n+m+1\right)/2\right]}{\mathrm{\Gamma }\left[n+m+1+r/2\right]\mathrm{\Gamma }\left[\left(n+m+r\right)/2\right]}},`$ and $`\eta `$ $`=`$ $`{\displaystyle \frac{1}{\xi }}\left({\displaystyle \frac{c}{2k_BT}}\right)^{r1}{\displaystyle \frac{1}{\pi ^{(n+m1)/2}\left[\left(n+m\right)^2nm2\right]}}`$ $`\times `$ $`{\displaystyle \frac{\left[(n+m1)^21\right]\left(n+m+1\right)^2}{\left[\left(n+m+r\right)\left(n+m+4+r\right)+2\frac{\left(n+m2\right)}{\left(n+m1\right)}\right]}}`$ $`\times `$ $`{\displaystyle \frac{\mathrm{\Gamma }\left[n+m2\right]\mathrm{\Gamma }\left[\left(n+m\right)/2\right]\mathrm{\Gamma }\left[\left(n+m+1\right)/2\right]}{\mathrm{\Gamma }\left[n+m+1+r/2\right]\mathrm{\Gamma }\left[\left(n+m+r\right)/2\right]}}.`$ These formulas present a locally anisotropic generalization of the Boisseau and van Leeuwen results. ## 7 Transport Theory in Curved Spaces with Rotation Ellipsoidal Horizons After having established in a general way the scheme for calculation of transport coefficients on la–spacetimes, we now specify an example for a four dimensional static metric (being a solution of Einstein equations in both general relativity and la–gravity) with the event horizon described by a hypersurface of rotation ellipsoid (see the Appendix). We note that in this case $`n=3,m=1`$ and $`n_{(a)}=4.`$ The main formulas for kinetic and thermodynamic variables from the previous Sections 3–6 were based on spherical symmetry of $`n_{(a)}1`$ dimensional volume (58) and spherical integral (59). For the rotation ellipsoid we have to modify the volume’s formula by introducing the ellipsoidal dependence $$V_{n_{(a)}1}=\frac{\left(n_{(a)}1\right)\pi ^{(n_{(a)}1)/2}}{\mathrm{\Gamma }\left[(n_{(a)}1)/2\right]}\rho \left(\stackrel{~}{\theta }\right)\underset{n_{(a)}=4}{\underset{}{=}}\frac{3\pi ^{3/2}}{\mathrm{\Gamma }\left[3/2\right]}\rho \left(\stackrel{~}{\theta }\right)$$ were $$\rho \left(\stackrel{~}{\theta }\right)=\frac{\rho _{(0)}}{1\epsilon \mathrm{cos}\stackrel{~}{\theta }}$$ is the parametric formula of an ellipse with constant parameter $`\rho _{(0)},`$ angle variable $`\stackrel{~}{\theta }`$ and eccentricity $`\epsilon =1/\sigma <1`$ is defined by the axes of rotation ellipsoid (see the formula (123) from Appendix). If in the case of spherical symmetry $`4\underset{0}{\overset{\pi /2}{}}𝑑\stackrel{~}{\theta }=2\pi ,`$ the ellipse deformation gives the result $$4\underset{0}{\overset{\pi /2}{}}\frac{d\stackrel{~}{\theta }}{1\epsilon \mathrm{cos}\stackrel{~}{\theta }}=\frac{8}{\sqrt{1\epsilon ^2}}\mathrm{arctan}\sqrt{\frac{1\epsilon }{1+\epsilon }}.$$ So, performing integrations on solid angles in spaces with rotational ellipsoid symmetry we can use the same formulas as for spherical symmetry but multiplied on $$q_{(\epsilon )}=\frac{4}{\pi \sqrt{1\epsilon ^2}}\mathrm{arctan}\sqrt{\frac{1\epsilon }{1+\epsilon }}.$$ For instance, the integral (59) transforms $$𝑑\mathrm{\Omega }𝑑\mathrm{\Omega }_{(\epsilon )}=q_{(\epsilon )}\frac{\left(n_{(a)}1\right)\pi ^{(n_{(a)}1)/2}}{\mathrm{\Gamma }\left[(n_{(a)}1)/2\right]}\underset{n_{(a)}=4}{\underset{}{=}}q_{(\epsilon )}\frac{3\pi ^{3/2}}{\mathrm{\Gamma }\left[3/2\right]}.$$ The formulas for the particle density (56) and energy density (5.2) of point particles must be multiplied on $`q_{(\epsilon )},`$ $$\stackrel{~}{n}\stackrel{~}{n}_{(\epsilon )}=q_{(\epsilon )}\stackrel{~}{n}\text{ and }\stackrel{~}{\epsilon }\stackrel{~}{\epsilon }_{(\epsilon )}=q_{(\epsilon )}\stackrel{~}{\epsilon },$$ but the energy per particle will remain constant. We also have to modify the formula for pressure (67), been proportional to the particle density, but consider unchanged the averaged enthalpy (68). The entropy per particle (71) and chemical potential (72) depends explicitly on $`q_{(\epsilon )}`$–factor because theirs formulas were derived by using the particle density $`\stackrel{~}{n}_{(\epsilon )}.`$ Here we note that all proved formulas depends on $`m(=1`$, in this Section) anisotropic parameters and on volume element determined by d–metric. In the first approximation of transport coefficients we could chose a locally isotropic background but introducing the factor $`q_{(\epsilon )}`$ and taking into account the dependence on anisotropic dimension. Putting $`\epsilon `$–corrections into (86) we obtain the first–order approximations to the transport coefficients in a la–spacetime with the symmetry of rotation ellipsoid $`\lambda `$ $`=`$ $`{\displaystyle \frac{ck_B\stackrel{~}{n}_{(\epsilon )}^2}{3}}{\displaystyle \frac{\beta _{\widehat{1}}^2}{a_{\widehat{2}\widehat{2}}}},`$ (105) $`\eta `$ $`=`$ $`{\displaystyle \frac{ck_BT\stackrel{~}{n}_{(\epsilon )}^2}{10}}{\displaystyle \frac{\gamma _{\widehat{0}}^2}{c_{\widehat{0}\widehat{0}}}},`$ $`\eta _{(v)}`$ $`=`$ $`ck_BT\stackrel{~}{n}_{(\epsilon )}^2{\displaystyle \frac{\alpha _{\widehat{2}}^2}{a_{\widehat{2}\widehat{2}}}}.`$ The values $`\alpha _{\widehat{2}},\beta _{\widehat{1}}`$ and $`\gamma _{\widehat{0}}`$ are $`3`$–fold integrals expressed in terms of enthalpy $`h`$ and temperature $`T`$ and have the limits $`\alpha _{\widehat{2}}`$ $`=`$ $`{\displaystyle \frac{5\gamma }{ck_BT\stackrel{~}{n}_{(\epsilon )}}}{\displaystyle \frac{3\gamma }{c\stackrel{~}{n}_{(\epsilon )}}}\underset{T\mathrm{}}{\underset{}{}}0,`$ (106) $`\beta _{\widehat{1}}`$ $`=`$ $`{\displaystyle \frac{k_BT}{c\stackrel{~}{n}_{(\epsilon )}}}{\displaystyle \frac{3\gamma }{\gamma 1}}\underset{T\mathrm{}}{\underset{}{}}12{\displaystyle \frac{k_BT}{c\stackrel{~}{n}_{(\epsilon )}}},`$ $`\gamma _{\widehat{0}}`$ $`=`$ $`{\displaystyle \frac{k_BT}{c^2\stackrel{~}{n}_{(\epsilon )}}}10h\underset{T\mathrm{}}{\underset{}{}}60{\displaystyle \frac{1}{\stackrel{~}{n}_{(\epsilon )}}}\left({\displaystyle \frac{k_BT}{c}}\right)^2.`$ In the next step we compute the $`q_{(\epsilon )}`$–deformations of brackets from the denominators of (105). Because the squares of $`\alpha ,\beta `$ and $`\gamma `$ coefficients from (85) are proportional to deformations of $`\stackrel{~}{n}^2`$ (this conclusion follows from the formulas (6.6) in the $`T\mathrm{}`$ limit, see (106)) and the scalar (98), vector (99) and tensor (100) type brackets do not change under $`q_{(\epsilon )}`$–deformations (see (6.6)) we conclude that the transport coefficients (86) do not contain the factor $`q_{(\epsilon )}`$ but depends only on the number $`m`$ of anisotropic dimensions. This conclusion is true only in the first approximation and for locally isotropic backgrounds. In consequence, the final formulas for the transport coefficient, see (6.6) and (6.6), in a (3+1) locally anisotropic spacetime with rotation ellipsoid symmetry are $`\eta _{(v)}`$ $``$ $`0`$ $`\lambda `$ $`=`$ $`{\displaystyle \frac{2k_Bc}{\xi }}\left({\displaystyle \frac{c}{2k_BT}}\right)^r{\displaystyle \frac{48}{6+r}}{\displaystyle \frac{\mathrm{\Gamma }\left[2\right]^2\mathrm{\Gamma }\left[5/2\right]}{\mathrm{\Gamma }\left[5+r/2\right]\mathrm{\Gamma }\left[\left(4+r\right)/2\right]}},`$ $`\eta `$ $`=`$ $`{\displaystyle \frac{1}{\xi }}\left({\displaystyle \frac{c}{2k_BT}}\right)^{r1}{\displaystyle \frac{20}{\pi ^{3/2}[(4+r)(8+r)+4/3]}}{\displaystyle \frac{\mathrm{\Gamma }[2]^2\mathrm{\Gamma }\left[5/2\right]}{\mathrm{\Gamma }\left[5+r/2\right]\mathrm{\Gamma }\left[\left(4+r\right)/2\right]}}.`$ This is consistent with the fact that we chose as an example a static metric with a local anisotropy that does not cause drastic changings in the structure of transport coefficients. Nevertheless, there are $`q_{(\epsilon )}`$–deformations of such values as the density of particles, kinetic potential and entropy which reflects modifications of kinetic and thermodynamic processes even by static spacetime local anisotropies. ## 8 Concluding Remarks The formulation of Einstein’s theory of relativity with respect to anholonomic frames raises a number of questions concerning locally anisotropic field interactions and kinetic and thermodynamic effects. We argue that spacetime local anisotropy (la) can be modeled by applying the Cartan’s moving frame method with associated nonlinear connection (N–connection) structures. A remarkable fact is that this approach allows a unified treatment of various type of theories with generic local anisotropy like generalized Finsler like gravities, of standard Kaluza–Klein models with nontrivial compactifications (modelled by N–connection structures), of standard general relativity with anholonomic frames and even of low dimensional models with distinguished anisotropic parameters. We have shown a relationship between a subclass of Finsler like metrics with (pseudo) Riemannian ones being solutions of Einstein equations. This paper has provided a generalization of relativistic kinetics and nonequilibrium thermodynamics in order to be included possible spacetime local anisotropies. It should not be considered as a work on definitions of some sophisticated theories on Finsler like spaces but developing an approach to kinetics and thermodynamics in (pseudo) Riemannian spacetimes of arbitrary dimension provided with anholonomic frame structures. The item of introducing into consideration of generalized Finsler and Kaluza–Klein spaces was imposed by the facts that one has recently constructed (see Appendix and solutions of the Einstein equations with generic anisotropy, of Finsler and another type, like ellipsoidal static black holes, black tora, anisotropic solitonic backgrounds and so on) and that in the low energy limits of string theories various classes of generalized Finsler–Kaluza–Klein metrics could be obtained alternatively to the well known (pseudo) Riemannian ones . By applying the moving frame method it is possible to elaborate a general schema for defining of physical values, basic equations and approximated calculations of kinetic and thermodynamic values with respect to la–frames in all type of the mentioned theories. The crucial ingredient in definition of collisionless relativistic locally anisotropic kinetic equation was the extension of the moving frame method to the space of supporting elements $`(u^\alpha ,p^\beta )`$ provided with an induced higher order anisotropic structure. The former Cartan–Vlasov approach , proposing a variant of statistical kinetic theory on curved phase spaces provided with Finsler like and Cartan N–connection structures, was self–consistently modified for both type of locally isotropic (the Einstein theory) and anisotropic (generalized Finsler–Kaluza–Klein) spacetimes with N–connection structures induced by local anholonomic frame or via reductions from higher dimensions in (super) string or (super) gravity theories. The physics of pair collisions in la–spacetimes was examined by introducing on the space of supporting elements of (correspondingly adapted to the N–connection structure) integral of collisions, differential cross–sections and velocity of transitions. Despite all the complexities of definition of equilibrium states with generic anisotropy it is possible a rigorous definition of local equilibrium particle distribution functions by fixing some anholonomic frames of reference adapted to the N–connection structure. The basic kinetic and thermodynamic values such as particle density, average energy and pressure, enthalpy, specific heats and entropy are derived via integrations on volume elements determined by metric components given with respect to la–bases. In the low and high energy limits the formulas reflect explicit dependencies on the number of anisotropic dimensions as well on anisotropic deformations of spacetime metric and linear connection. One can linearize the transport equations and prove the linear laws for locally anisotropic non–equilibrium thermodynamics. It has also established a general scheme for calculation of transport coefficients (the heat conductivity, the shear viscosity and the volume viscosity) in la–spacetimes. An explicit computation of such values was performed for a metric with rotation ellipsoidal event horizon (an example of spacetime with static local anisotropy), recently found as a new solution of the Einstein equations. Our overall conclusion is that in order to obtain a self–consistent formulation of the locally anisotropic kinetic and thermodynamic theory in curved spacetimes and calculation of basic physical values we must consider moving frames with correspondingly adapted nonlinear connection structures. Aknowledgments The author want to thank Professors G. Neugebauer and H. Dehnen for hospitality and support of his participation at Journees Relativistes 99 at Weimar and visit at Konstantz University, Germany, where the bulk of results were communicated. ## Appendix: A Locally Anisotropic Solution of Einstein’s Equations Before presenting an explicit construction of a four dimensional solution with local anisotropy of the Einstein equations (8) we briefly review the properties of four dimensional metrics which transforms into (2+2) or (3+1) anisotropic d–metrics (we note that this is not a (space + time) but a (isotropic+anisotropic) decomposition of coordinates) by transitions to correspondingly defined anholonomic bases of tetrads (vierbeins). Let us consider a four dimensional (in brief, 4D) spacetime $`V^{(2+2)}`$ (with two isotropic plus two anisotropic local coordinates) provided with a metric of signature (+,+,–,+) parametrized by a symmetric matrix of type $$\left[\begin{array}{cccc}g_1+q_1{}_{}{}^{2}h_{3}^{}+n_1{}_{}{}^{2}h_{4}^{}& 0& q_1h_3& n_1h_4\\ 0& g_2+q_2{}_{}{}^{2}h_{3}^{}+n_2{}_{}{}^{2}h_{4}^{}& q_2h_3& n_2h_4\\ q_1h_3& q_2h_3& h_3& 0\\ n_1h_4& n_2h_4& 0& h_4\end{array}\right]$$ (107) with coefficients being some functions $$g_i=g_i(x^j),q_i=q_i(x^j,t),n_i=n_i(x^j,t),h_a=h_a(x^j,t)$$ of necessary smooth class. With respect to a la–basis (10) the ansatz (107) results in diagonal $`2\times 2`$ h– and v–metrics for a d–metric (2.2) (for simplicity, we shall consider only diagonal 2D nondegenerated metrics because for such dimensions every symmetric matrix can be diagonalized). An equivalent diagonal d–metric of type (2.2) is obtained for the associated N–connection with the coefficients being functions on three coordinates $`(x^i,z),`$ $`N_1^3`$ $`=`$ $`q_1(x^i,z),N_2^3=q_2(x^i,z),`$ (108) $`N_1^4`$ $`=`$ $`n_1(x^i,z),N_2^4=n_2(x^i,z).`$ For simplicity, we shall use brief denotations of partial derivatives, like $`\dot{a}`$ $`=a/x^1,`$ $`a^{}=a/x^2,`$ $`a^{}=a/z`$ $`\dot{a}^{}`$ $`=^2a/x^1x^2,`$ $`a^{}=^2a/zz.`$ The non–trivial components of the Ricci d–tensor (15) (for the ansatz 107) when $`R_1^1=R_2^2`$ and $`S_3^3=S_4^4,`$ are computed $`R_1^1`$ $`=R_2^2={\displaystyle \frac{1}{2g_1g_2}}[(g_1^{^{\prime \prime }}+\ddot{g}_2)+{\displaystyle \frac{1}{2g_2}}\left(\dot{g}_2^2+g_1^{}g_2^{}\right)+{\displaystyle \frac{1}{2g_1}}\left(g_1^{2}+\dot{g}_1\dot{g}_2\right)];`$ (109) $`S_3^3`$ $`=S_4^4={\displaystyle \frac{1}{h_3h_4}}[h_4^{}+{\displaystyle \frac{1}{2h_4}}(h_4^{})^2+{\displaystyle \frac{1}{2h_3}}h_3^{}h_4^{}];`$ (110) $`P_{31}`$ $`={\displaystyle \frac{q_1}{2}}[\left({\displaystyle \frac{h_3^{}}{h_3}}\right)^2{\displaystyle \frac{h_3^{}}{h_3}}+{\displaystyle \frac{h_4^{}}{2h_4^{2}}}{\displaystyle \frac{h_3^{}h_4^{}}{2h_3h_4}}]+{\displaystyle \frac{1}{2h_4}}[{\displaystyle \frac{\dot{h}_4}{2h_4}}h_4^{}\dot{h}_4^{}+{\displaystyle \frac{\dot{h}_3}{2h_3}}h_4^{}],`$ $`P_{32}`$ $`={\displaystyle \frac{q_2}{2}}[\left({\displaystyle \frac{h_3^{}}{h_3}}\right)^2{\displaystyle \frac{h_3^{}}{h_3}}+{\displaystyle \frac{h_4^{}}{2h_4^{2}}}{\displaystyle \frac{h_3^{}h_4^{}}{2h_3h_4}}]+{\displaystyle \frac{1}{2h_4}}[{\displaystyle \frac{h_4^{}}{2h_4}}h_4^{}h_4^{}+{\displaystyle \frac{h_3^{}}{2h_3}}h_4^{}];`$ $`P_{41}`$ $`={\displaystyle \frac{h_4}{2h_3}}n_1^{}+{\displaystyle \frac{1}{4h_3}}({\displaystyle \frac{h_4}{h_3}}h_3^{}3h_4^{})n_1^{},`$ $`P_{42}`$ $`={\displaystyle \frac{h_4}{2h_3}}n_2^{}+{\displaystyle \frac{1}{4h_3}}({\displaystyle \frac{h_4}{h_3}}h_3^{}3h_4^{})n_2^{}.`$ The curvature scalar $`\stackrel{}{R}`$ (16) is defined by the sum of two non-trivial components $`\widehat{R}=2R_1^1`$ and $`S=2S_3^3.`$ The system of Einstein equations (17) transforms into $`R_1^1`$ $`=`$ $`\kappa \mathrm{{\rm Y}}_3^3=\kappa \mathrm{{\rm Y}}_4^4,`$ (113) $`S_3^3`$ $`=`$ $`\kappa \mathrm{{\rm Y}}_1^1=\kappa \mathrm{{\rm Y}}_2^2,`$ (114) $`P_{3i}`$ $`=`$ $`\kappa \mathrm{{\rm Y}}_{3i},`$ (115) $`P_{4i}`$ $`=`$ $`\kappa \mathrm{{\rm Y}}_{4i},`$ (116) where the values of $`R_1^1,S_3^3,P_{ai},`$ are taken respectively from (109), (110), (Appendix: A Locally Anisotropic Solution of Einstein’s Equations), (Appendix: A Locally Anisotropic Solution of Einstein’s Equations). We note that we can define the N–coefficients (108), $`q_i(x^k,z)`$ and $`n_i(x^k,z),`$ by solving the equations (115) and (116) if the functions $`h_i(x^k,z)`$ are known as solutions of the equations (114). An elongated rotation ellipsoid hypersurface is given by the formula $$\frac{\stackrel{~}{x}^2+\stackrel{~}{y}^2}{\sigma ^21}+\frac{\stackrel{~}{z}^2}{\sigma ^2}=\stackrel{~}{\rho }^2,$$ (117) where $`\sigma 1`$ and $`\stackrel{~}{\rho }`$ is similar to the radial coordinate in the spherical symmetric case. The space 3D coordinate system is defined $`\stackrel{~}{x}`$ $`=`$ $`\stackrel{~}{\rho }\mathrm{sinh}u\mathrm{sin}v\mathrm{cos}\phi ,\stackrel{~}{y}=\stackrel{~}{\rho }\mathrm{sinh}u\mathrm{sin}v\mathrm{sin}\phi ,`$ $`\stackrel{~}{z}`$ $`=`$ $`\stackrel{~}{\rho }\mathrm{cosh}u\mathrm{cos}v,`$ where $`\sigma =\mathrm{cosh}u,(0u<\mathrm{},0v\pi ,0\phi <2\pi ).`$ The hypersurface metric is $`g_{uu}`$ $`=`$ $`g_{vv}=\stackrel{~}{\rho }^2\left(\mathrm{sinh}^2u+\mathrm{sin}^2v\right),`$ (118) $`g_{\phi \phi }`$ $`=`$ $`\stackrel{~}{\rho }^2\mathrm{sinh}^2u\mathrm{sin}^2v.`$ Let us introduce a d–metric $$\delta s^2=g_1(u,v)du^2+dv^2+h_3(u,v,\phi )\left(\delta t\right)^2+h_4(u,v,\phi )\left(\delta \phi \right)^2,$$ (119) where $`\delta t`$ and $`\delta \phi `$ are N–elongated differentials. As a particular solution (113) for the h–metric, considering $`\mathrm{{\rm Y}}_3^3=\mathrm{{\rm Y}}_4^4,`$ we choose the coefficient $$g_1(u,v)=\mathrm{cos}^2v.$$ (120) The $`h_3(u,v,\phi )=h_3(u,v,\stackrel{~}{\rho }(u,v,\phi ))`$ is considered as $$h_3(u,v,\stackrel{~}{\rho })=\frac{1}{\mathrm{sinh}^2u+\mathrm{sin}^2v}\frac{\left[1\frac{r_g}{4\stackrel{~}{\rho }}\right]^2}{\left[1+\frac{r_g}{4\stackrel{~}{\rho }}\right]^6}.$$ (121) In order to define the $`h_4`$ coefficient solving the Einstein equations, for simplicity, with a diagonal energy–momentum d–tensor for vanishing pressure, we must solve the equation (114) which transforms into a linear equation if $`\mathrm{{\rm Y}}_1=0.`$ In our case $`s(u,v,\phi )=\beta ^1(u,v,\phi ),`$ where $`\beta =\left(h_4/\phi \right)/h_4,`$ must be a solution of $$\frac{s}{\phi }+\frac{\mathrm{ln}\sqrt{\left|h_3\right|}}{\phi }s=\frac{1}{2}.$$ After two integrations (see ) the general solution for $`h_4(u,v,\phi ),`$ is $$h_4(u,v,\phi )=a_4(u,v)\mathrm{exp}\left[\underset{0}{\overset{\phi }{}}F(u,v,z)𝑑z\right],$$ (122) where $$F(u,v,z)=(\sqrt{|h_3(u,v,z)|}[s_{1(0)}(u,v)+\frac{1}{2}\underset{z_0(u,v)}{\overset{z}{}}\sqrt{|h_3(u,v,z)|}𝑑z])^1,$$ $`s_{1(0)}(u,v)`$ and $`z_0(u,v)`$ are some functions of necessary smooth class. We note that if we put $`h_4=a_4(u,v)`$ the equations (114) are satisfied for every $`h_3=h_3(u,v,\phi ).`$ Every d–metric (119) with coefficients of type (120), (121) and (122) solves the Einstein equations (113)–(116) with the diagonal momentum d–tensor $$\mathrm{{\rm Y}}_\beta ^\alpha =diag[0,0,\epsilon =m_0,0],$$ when $`r_g=2\kappa m_0;`$ we set the light constant $`c=1.`$ If we choose $$a_4(u,v)=\frac{\mathrm{sinh}^2u\mathrm{sin}^2v}{\mathrm{sinh}^2u+\mathrm{sin}^2v}$$ our solution is conformally equivalent (if not considering the time–time component) to the hypersurface metric (118). The condition of vanishing of the coefficient (121) parametrizes the rotation ellipsoid for the horizon $$\frac{\stackrel{~}{x}^2+\stackrel{~}{y}^2}{\sigma ^21}+\frac{\stackrel{~}{z}^2}{\sigma ^2}=\left(\frac{r_g}{4}\right)^2,$$ (123) where the radial coordinate is redefined via relation $`\stackrel{~}{r}=\stackrel{~}{\rho }\left(1+\frac{r_g}{4\stackrel{~}{\rho }}\right)^2.`$ After multiplication on the conformal factor $$\left(\mathrm{sinh}^2u+\mathrm{sin}^2v\right)\left[1+\frac{r_g}{4\stackrel{~}{\rho }}\right]^4,$$ approximating $`g_1(u,v)=\mathrm{sin}^2v0,`$ in the limit of locally isotropic spherical symmetry, $$\stackrel{~}{x}^2+\stackrel{~}{y}^2+\stackrel{~}{z}^2=r_g^2,$$ the d–metric (119) reduces to $$ds^2=\left[1+\frac{r_g}{4\stackrel{~}{\rho }}\right]^4\left(d\stackrel{~}{x}^2+d\stackrel{~}{y}^2+d\stackrel{~}{z}^2\right)\frac{\left[1\frac{r_g}{4\stackrel{~}{\rho }}\right]^2}{\left[1+\frac{r_g}{4\stackrel{~}{\rho }}\right]^2}dt^2$$ which is just the Schwazschild solution with the redefined radial coordinate when the space component becomes conformally Euclidean. So, the d–metric (119), the coefficients of N–connection being solutions of (115) and (116), describe a static 4D solution of the Einstein equations when instead of a spherical symmetric horizon one considers a locally anisotropic deformation to the hypersurface of rotation elongated ellipsoid.
warning/0001/hep-ph0001030.html
ar5iv
text
# Probing Intermediate Mass Higgs Interactions at the CERN Large Hadron Collider ## I Introduction The search for the Higgs boson and the study of the $`SU(2)_LU(1)_Y`$ symmetry breaking mechanism are the main goals of the present and future high energy experiments . At present, the best available limit on the Higgs mass arises from searches at the CERN LEP collider. The ALEPH Collaboration analysis of the 1999 data with integrated luminosities of 29 pb<sup>-1</sup> at $`\sqrt{s}=191.6`$ GeV and 69.5 pb<sup>-1</sup> at $`\sqrt{s}=195.6`$ GeV yields $`M_H>98.8`$ GeV at 95% CL. A next step will be given by the CERN Large Hadron Collider (LHC) that will be able to detect the standard Higgs boson in the decay channel $`H\gamma \gamma `$ for masses in the range of 100–150 GeV . If a Higgs boson is observed then it is imperative to verify whether its couplings are in agreement with the ones predicted by the Standard Model (SM) with a single scalar doublet. In the SM, the precise form of the Higgs couplings to the gauge bosons and its self–couplings are completely determined in terms of one free parameter which can be chosen to be the Higgs mass $`M_H`$. However, in this simple realization, the theory presents the problem that the quantum corrections to the Higgs mass are quadratically divergent with the cut–off. This implies the necessity of a large fine–tuning in order to keep the theory perturbative up to very high energies, or, conversely, the existence of new physics which manifest itself above a certain scale $`\mathrm{\Lambda }`$. If the energy scale of new physics is large compared to the electroweak scale and there is no new light resonances, we can represent its impact on Higgs boson properties via the introduction of effective operators . In this approach, we parametrize the Higgs anomalous interactions by the most general dimension–six effective lagrangian with the linear realization of the $`SU(2)_LU(1)_Y`$ symmetry. In the linear realization of the SM there are eleven $`C`$ and $`P`$ conserving dimension–six operators with some of them contributing at tree level to well measured observables, and consequently being severely constrained . In Ref. it was argued that it is unnatural to expect a large hierarchy between the coefficients of the various dimension–six operators independently on whether they do or do not contribute at tree level to the low energy observables. As a consequence the existing limits coming from the LEP I physics would preclude the direct observation of new effects in the processes accessible at LEP II and Tevatron. In this work, we study the potentiality of the CERN LHC to probe the interactions of an intermediate mass Higgs boson and consequently its sensitivity to new physics scales beyond the constraints stemming from the tree level contribution of the effective operators to low energy observables. For intermediate Higgs masses, the decay channel $`H\gamma \gamma `$ is the most important one for the Higgs search at the LHC. Moreover the effect of the additional Higgs interactions on the properties of an intermediate mass Higgs can be more easily seen in processes that are suppressed in the SM, such as the Higgs decay into two photons. In the SM, this decay occurs only at one–loop level, and it can be enhanced (or suppressed) by the anomalous interactions. We analyze the Higgs production and subsequent decay into two photons through gluon–gluon fusion $$ppggH(\gamma \gamma ),$$ (1) as well as through vector–boson fusion mechanism $$ppqq^{}VVj+j+H(\gamma \gamma ).$$ (2) with $`V=W^\pm `$ or $`Z^0`$. We show that the LHC will be able to expand considerably the present sensitivity on the dimension–six Higgs couplings, being able to probe new physics scales as large as 2.2 TeV, provided that the Higgs is observed. In this case, our results show that the LHC will be able to improve the limits arising from the tree level contribution to the LEP I observables, which are presently the most severe constraint. Furthermore, there is also a distinct possibility, i.e. the existence of anomalous Higgs interactions reduces its decay into two photons. In this case, no signal will be observed in the above reactions, and, consequently, if the Higgs is observed in other decay channel, the existence of non–vanishing anomalous couplings could also be established. ## II Effective Higgs Interactions and Present Bounds In the linear representation of the $`SU(2)_LU(1)_Y`$ symmetry breaking mechanism, the SM model is the lowest order approximation while the first corrections, which are of dimension six, can be written as $$_{\text{eff}}=\underset{n}{}\frac{f_n}{\mathrm{\Lambda }^2}𝒪_n,$$ (3) where the operators $`𝒪_n`$ involve vector–boson and/or Higgs–boson fields with couplings $`f_n`$ . This effective Lagrangian describes the phenomenology of models that are somehow close to the SM since a light Higgs scalar doublet is still present at low energies. Of the eleven possible operators $`𝒪_n`$ that are $`P`$ and $`C`$ even, only seven of them modify the Higgs–boson couplings to vector bosons , $`𝒪_{BW}=\mathrm{\Phi }^{}\widehat{B}_{\mu \nu }\widehat{W}^{\mu \nu }\mathrm{\Phi },`$ (4) $`𝒪_{WW}=\mathrm{\Phi }^{}\widehat{W}_{\mu \nu }\widehat{W}^{\mu \nu }\mathrm{\Phi },`$ (5) $`𝒪_{BB}=\mathrm{\Phi }^{}\widehat{B}_{\mu \nu }\widehat{B}^{\mu \nu }\mathrm{\Phi },`$ (6) $`𝒪_W=(D_\mu \mathrm{\Phi })^{}\widehat{W}^{\mu \nu }(D_\nu \mathrm{\Phi }),`$ (7) $`𝒪_B=(D_\mu \mathrm{\Phi })^{}\widehat{B}^{\mu \nu }(D_\nu \mathrm{\Phi }),`$ (8) $`𝒪_{\mathrm{\Phi },1}=\left(D_\mu \mathrm{\Phi }\right)^{}\mathrm{\Phi }^{}\mathrm{\Phi }\left(D^\mu \mathrm{\Phi }\right),`$ (9) $`𝒪_{\mathrm{\Phi },2}={\displaystyle \frac{1}{2}}^\mu \left(\mathrm{\Phi }^{}\mathrm{\Phi }\right)_\mu \left(\mathrm{\Phi }^{}\mathrm{\Phi }\right),`$ (10) where $`\mathrm{\Phi }`$ is the Higgs doublet, $`D_\mu `$ the covariant derivative, $`\widehat{B}_{\mu \nu }=i(g^{}/2)B_{\mu \nu }`$, and $`\widehat{W}_{\mu \nu }=i(g/2)\sigma ^aW_{\mu \nu }^a`$, with $`B_{\mu \nu }`$ and $`W_{\mu \nu }^a`$ being respectively the $`U(1)_Y`$ and $`SU(2)_L`$ field strength tensors. It is interesting to notice that the operators $`𝒪_{\mathrm{\Phi },1}`$ and $`𝒪_{\mathrm{\Phi },2}`$ contribute to the weak boson mass and Higgs wave function, which in turn leads to new Higgs couplings to the gauge bosons. The effective operators in Eq. (7) give rise to anomalous $`H\gamma \gamma `$, $`HZ\gamma `$, $`HZZ`$, and $`HW^+W^{}`$ couplings, which, in the unitary gauge, are given by $`_{\text{eff}}^{\text{HVV}}`$ $`=`$ $`g_{H\gamma \gamma }HA_{\mu \nu }A^{\mu \nu }+g_{HZ\gamma }^{(1)}A_{\mu \nu }Z^\mu ^\nu H+g_{HZ\gamma }^{(2)}HA_{\mu \nu }Z^{\mu \nu }`$ (11) $`+`$ $`g_{HZZ}^{(1)}Z_{\mu \nu }Z^\mu ^\nu H+g_{HZZ}^{(2)}HZ_{\mu \nu }Z^{\mu \nu }+h_{HZZ}^{(3)}HZ_\mu Z^\mu `$ (12) $`+`$ $`g_{HWW}^{(1)}\left(W_{\mu \nu }^+W^\mu ^\nu H+\text{h.c.}\right)+g_{HWW}^{(2)}HW_{\mu \nu }^+W^{\mu \nu }+g_{HWW}^{(3)}HW_\mu ^+W^\mu ,`$ (13) where $`A(Z)_{\mu \nu }=_\mu A(Z)_\nu _\nu A(Z)_\mu `$. The effective couplings $`g_{H\gamma \gamma }`$, $`g_{HZ\gamma }^{(1,2)}`$, and $`g_{HZZ}^{(1,2,3)}`$ are related to the coefficients of the operators appearing in (3) through, $`g_{H\gamma \gamma }`$ $`=`$ $`\left({\displaystyle \frac{gM_W}{\mathrm{\Lambda }^2}}\right){\displaystyle \frac{s^2(f_{BB}+f_{WW}f_{BW})}{2}},`$ (14) $`g_{HZ\gamma }^{(1)}`$ $`=`$ $`\left({\displaystyle \frac{gM_W}{\mathrm{\Lambda }^2}}\right){\displaystyle \frac{s(f_Wf_B)}{2c}},`$ (15) $`g_{HZ\gamma }^{(2)}`$ $`=`$ $`\left({\displaystyle \frac{gM_W}{\mathrm{\Lambda }^2}}\right){\displaystyle \frac{s[2s^2f_{BB}2c^2f_{WW}+(c^2s^2)f_{BW}]}{2c}},`$ (16) $`g_{HZZ}^{(1)}`$ $`=`$ $`\left({\displaystyle \frac{gM_W}{\mathrm{\Lambda }^2}}\right){\displaystyle \frac{c^2f_W+s^2f_B}{2c^2}},`$ (17) $`g_{HZZ}^{(2)}`$ $`=`$ $`\left({\displaystyle \frac{gM_W}{\mathrm{\Lambda }^2}}\right){\displaystyle \frac{s^4f_{BB}+c^4f_{WW}+c^2s^2f_{BW}}{2c^2}},`$ (18) $`g_{HZZ}^{(3)}`$ $`=`$ $`\left({\displaystyle \frac{gM_Wv^2}{\mathrm{\Lambda }^2}}\right){\displaystyle \frac{f_{\mathrm{\Phi },1}f_{\mathrm{\Phi },2}}{4c^2}},`$ (19) $`g_{HWW}^{(1)}`$ $`=`$ $`\left({\displaystyle \frac{gM_W}{\mathrm{\Lambda }^2}}\right){\displaystyle \frac{f_W}{2}},`$ (20) $`g_{HWW}^{(2)}`$ $`=`$ $`\left({\displaystyle \frac{gM_W}{\mathrm{\Lambda }^2}}\right)f_{WW},`$ (21) $`g_{HWW}^{(3)}`$ $`=`$ $`\left({\displaystyle \frac{gM_Wv^2}{\mathrm{\Lambda }^2}}\right){\displaystyle \frac{f_{\mathrm{\Phi },1}+2f_{\mathrm{\Phi },2}}{4}},`$ (22) with $`g`$ being the electroweak coupling constant and $`s(c)\mathrm{sin}(\mathrm{cos})\theta _W`$. In the couplings $`g_{HZZ}^{(3)}`$ and $`g_{HWW}^{(3)}`$ we have also included the effects arising from the contribution of the operators $`𝒪_{\mathrm{\Phi },1}`$ and $`𝒪_{\mathrm{\Phi },2}`$ to the renormalization of the weak boson masses and the Higgs field wave function. The operators $`𝒪_{\mathrm{\Phi },1}`$ and $`𝒪_{BW}`$ contribute at tree level to the vector–boson two–point functions, and consequently are severely constrained by low–energy data . The present 95% CL limits on these operators for 90 GeV $`M_H`$ 800 GeV and $`m_{\text{top}}=175`$ GeV read , $`1.2`$ $`{\displaystyle \frac{f_{\mathrm{\Phi },1}}{\mathrm{\Lambda }^2}}0.56\text{ TeV}^2,`$ (23) $`1.0`$ $`{\displaystyle \frac{f_{BW}}{\mathrm{\Lambda }^2}}8.6\text{ TeV}^2.`$ (24) On the order hand, the remaining operators can be indirectly constrained via their one–loop contributions to low–energy observables, which are suppressed by factors $`1/(16\pi ^2)`$. Using the “naturalness” assumption that large cancellations do not occur among their contributions, we can consider only the effect of one operator at a time. In this case, the following constraints at 95% CL (in units of TeV<sup>-2</sup>) arise $`12.{\displaystyle \frac{f_W}{\mathrm{\Lambda }^2}}2.5,`$ (25) $`7.6{\displaystyle \frac{f_B}{\mathrm{\Lambda }^2}}22,`$ (26) $`24{\displaystyle \frac{f_{WW}}{\mathrm{\Lambda }^2}}14,`$ (27) $`79{\displaystyle \frac{f_{BB}}{\mathrm{\Lambda }^2}}47.`$ (28) These limits depend in a complex way on the Higgs mass. The values quoted above for the sake of illustration were obtained for $`M_H=200`$ GeV. Some of the anomalous Higgs interaction can also be constrained by their effect on the triple gauge–boson vertices. Recently, the LEP and Tevatron Collaborations have studied the production of pairs of gauge bosons and derived bounds on the anomalous interactions that modify the $`WW\gamma `$ and $`WWZ`$ vertices. Combining the published results from DØ and the four LEP experiments, the 95% CL bounds on the anomalous Higgs interactions are (in TeV<sup>-2</sup>) : $`31{\displaystyle \frac{(f_W+f_B)}{\mathrm{\Lambda }^2}}68,\text{ for }f_{WWW}=0.`$ (29) Notice that, since neither $`f_{WW}`$ nor $`f_{BB}`$ contribute to the triple gauge–boson vertices, no direct constraint on these couplings can be derived from this analysis. Notwithstanding, these couplings can be constrained by data on Higgs searches at LEP II and Tevatron colliders. The combined analysis of these signatures yields the following 95% CL bounds on the anomalous Higgs interactions (in TeV<sup>-2</sup>): $`7.5{\displaystyle \frac{f_{WW(BB)}}{\mathrm{\Lambda }^2}}18`$for $`M_H150`$ GeV. These limits can be improved by a factor 2–3 in the upgraded Tevatron runs. ## III Intermediate Mass Higgs Production and New Interactions ### A Gluon–gluon fusion At the LHC, light Higgs bosons are copiously produced through gluon-gluon fusion and it was established that the decay mode $`H\gamma \gamma `$ provides the best signature for Higgs masses in the range $`90<M_H<150`$ GeV . However, this channel also possesses a very large background, requiring a good energy resolution from the detectors in order to be observed. In our analyses we computed the SM and anomalous Higgs production cross sections using the cuts and efficiencies that the ATLAS Collaboration applied in their studies. We imposed the acceptance cuts $`p_T^{\gamma _{1(2)}}>25(40)\text{GeV},|\eta _{\gamma _{(1,2)}}|<2.5,`$where $`p_T^{\gamma _{1(2)}}`$ stands for the largest (second largest) transverse momentum of the photons in the event and $`\eta _{\gamma _{(1,2)}}`$ are the rapidities of these photons. We also required that $`{\displaystyle \frac{p_T^{\gamma _1}}{(p_T^{\gamma _1}+p_T^{\gamma _2})}}<0.7`$to reduce the quark bremsstrahlung background. The efficiency for reconstruction and identification of one photon was taken to be 80%. We collected the photon pair events whose invariant masses fall in bins of size $`2\mathrm{\Delta }M_H`$ around the Higgs mass with $`\left({\displaystyle \frac{\mathrm{\Delta }M_H}{M_H}}\right)={\displaystyle \frac{5\%}{\sqrt{M_H}}}0.5\%.`$ In order to take advantage of the careful estimates of backgrounds and detector efficiencies made by the ATLAS collaboration, we normalized our predictions for the SM Higgs cross section to their value for each value of the Higgs mass and then we rescaled the anomalous production cross section by the same factor. This factor varies from 0.65 to 0.80, depending on the Higgs mass. We present in Table I the expected number of Higgs and background events in the SM for a center of mass energy $`\sqrt{s}=14`$ TeV and an integrated luminosity of 100 fb<sup>-1</sup> after cuts and efficiencies. The SM irreducible background is by far the most important one, however, there is still rather large jet–jet and $`\gamma `$–jet reducible backgrounds. It is worth mentioning that the SM Higgs can be observed in this channel with an statistical significance of more than 3.9 standard deviations. For an intermediate mass Higgs, $`𝒪_{BB}`$, $`𝒪_{WW}`$, and $`𝒪_{BW}`$ are the only effective operators (7) that contribute significantly at tree level to the process $`ggH\gamma \gamma `$. For the sake of illustration, we exhibit in Fig. 1 the expected number of events after cuts for a Higgs boson with $`M_H=130`$ GeV as a function of the anomalous coupling $$f\frac{(f_{BB}+f_{WW}f_{BW})}{\mathrm{\Lambda }^2}.$$ (30) Notice that the anomalous contribution to $`H\gamma \gamma `$ depends only on this combination of the effective interactions \[see Eq. (18)\]. Let us initially assume that the Higgs boson has indeed been observed in the $`\gamma \gamma `$ channel and that its production cross section is in agreement with the SM prediction. In this case we can extract the sensitivity of this process to the anomalous interactions using as background the processes listed in Table I together with the SM Higgs signal. We present in Table II the 95% CL allowed range of $`f`$ for an integrated luminosity of 100 fb<sup>-1</sup>. In Fig. 1, we also exhibit the 95% CL interval around the SM Higgs signal (upper shadowed region) for $`M_H=130`$ GeV. The two allowed parameter ranges given in Table II correspond to the two intersections of the signal curve with the 95% CL region. One should notice that the existence of a sizable interference between the SM and anomalous contributions to the Higgs into two photons width allows the existence of a range of anomalous couplings not compatible with zero even if its production rate is in accordance with the SM predictions. This range would correspond to $`2.5<f/\mathrm{\Lambda }^2<3.0`$ TeV<sup>-2</sup>, for $`M_H=130`$ GeV (see Fig. 1). The bounds in Table II, derived from the SM Higgs observation, are more restrictive than the ones emanating from the contribution of these operators to the low–energy and LEP physics either at tree level (24) or at the one–loop level (26). As seen in this table for any Higgs mass, this process is sensitive to new physics which characteristic scale up to $`\mathrm{\Lambda }2.2`$ TeV for $`f_{BB}=f_{WW}=f_{BW}=1`$. Another possible scenario is the one where no SM Higgs signal is observed in the $`\gamma \gamma `$ channel. In this case, we obtain different constraints on the anomalous couplings for a given value of $`M_H`$. In order to extract the sensitivity region for the anomalous couplings we have to consider only the SM Higgs backgrounds listed in Table I without including in that the SM Higgs signal. In Table III we present sensitivity range for the anomalous coupling $`f`$ at 95% CL, assuming that only the SM Higgs background is observed for an integrated luminosity of 100 fb<sup>-1</sup>. The interpretation of these results depend on the observation of the Higgs in other channels. Should the Higgs be observed in another decay channel insensitive to these anomalous couplings, such as in $`H\tau ^+\tau ^{}`$ , the non observation of the corresponding signal in $`\gamma \gamma `$ should be interpreted as the inevitable existence of new physics in the Higgs couplings to photons with characteristic strength as given in Table III. If, on the other hand, no signal is observed in any other decay channel, the existence of the Higgs of a given mass can be ruled out and no limit can be extracted from the non observation of the $`\gamma \gamma `$ signal, becoming meaningless the constraints in Table III. ### B Vector boson fusion Higgs bosons can also be produced in $`pp`$ collisions via the vector boson fusion (VBF) process (2). A nice feature of this production mechanism is that the jets in the VBF signal events tend to populate the forward direction and can be used to tag these events and reduce the background. Moreover, the background can be further suppressed by vetoing additional jet activity in the central region . In our analysis of this process we followed closely the study of Ref. . We considered all backgrounds from QCD and electroweak processes, as well as the corresponding interferences, which can lead to events with two photons and two jets. We generated the scattering amplitudes for both the SM Higgs signal and the background with the package Madgraph which makes use of the helicity amplitudes contained in the package Helas . In our calculations we have used the MRS(G) proton structure function. In order to reduce the background, we required the photons to satisfy the acceptance cuts $`p_T^{\gamma _{1(2)}}>25(50)\text{GeV},|\eta _{\gamma _{(1,2)}}|<2.5,`$and we tagged the jets in the forward region and vetoed their presence in the central detector $`p_T^{j_{(1,2)}}>20(40)\text{GeV},`$ $`|\eta _{j_{(1,2)}}|<5.0,`$ $`|\eta _{j_1}\eta _{j_2}|>4.4,\eta _{j_1}.\eta _{j_2}<0.`$ We also imposed that the photons are isolated from the jets ($`\mathrm{\Delta }R_{\gamma j}>0.7`$) and, as before, used the photon detection efficiency of 80%. We display in Table IV the expected number of SM Higgs signal and background events using the above cuts and efficiencies for an integrated luminosity of 100 fb<sup>-1</sup>. Our results show good quantitative agreement with the analysis in Ref. , taking into account the different choice of structure functions and the inclusion of the photon detection efficiency. Notice that the SM irreducible background and the SM Higgs signal are of the same order of magnitude. The dimension–six operators in Eq. (7) contribute to the $`\gamma \gamma `$ signal of Higgs production via vector boson fusion by modifying both the production cross section as well as the Higgs decay width. On one hand, the operators $`𝒪_{BB}`$, $`𝒪_{WW}`$, and $`𝒪_{BW}`$ contribute to the Higgs production and decay while $`𝒪_W`$, $`𝒪_B`$, $`𝒪_{\mathrm{\Phi },1}`$, and $`𝒪_{\varphi ,2}`$ change only the production vertices. In our calculation of the anomalous contribution we have included all amplitudes generated by these operators in the spirit of Refs. . As in the gluon–gluon fusion analysis, we first assume that the Higgs boson has indeed been observed in the $`\gamma \gamma jj`$ channel and that its production cross section is in agreement with the SM prediction. In this case, the SM backgrounds for the anomalous interactions are the processes listed in Table IV together with the SM Higgs signal. We present in Table V the 95% CL sensitivity to anomalous coupling combination (30) assuming that the other anomalous couplings vanish, for an integrated luminosity of 100 fb<sup>-1</sup>. This Table also contains the 95% CL allowed range for the “super–blind” operator $`𝒪_{\mathrm{\Phi },2}`$ alone as well as the limits attainable assuming all couplings to be equal $`(f_{all}=f_{BB}=f_{WW}=f_{BW}=f_W=f_B=f_{\mathrm{\Phi }_1}=f_{\mathrm{\Phi }_2})`$. Here we also find two allowed ranges of anomalous parameters due to the sizeable interference between the SM and the anomalous contributions to the Higgs production and decay. Comparing the results in Tables II and V we see that for those operators contributing to the Higgs decay into two photons, the VBF process leads to slightly better sensitivity to new interactions. On the other hand, the LHC capability to probe the operators contributing only to the Higgs production in VBF is much smaller, being $`f_{\mathrm{\Phi },2}`$ the only coupling that can be meaningfully constrained. For $`f_B`$, $`f_W`$, and $`f_{\mathrm{\Phi },1}`$ we verified that the expected sensitivity from this analysis is much worse than present limits in Eqs. (23), (24), and (29). Different sensitivity bounds are obtained if no Higgs boson signal is observed in the VBF channel for a given value of $`M_H`$. In this scenario, we assumed that only the SM background for the Higgs search were observed (see Table IV). In Table VI we show sensitivity range of anomalous couplings at the 95% CL for an integrated luminosity of 100 fb<sup>-1</sup> assuming that no Higgs signal is observed in the corresponding mass bin. As discussed above the interpretation of these results depend whether the Higgs has been observed in other channels. The effective interactions $`𝒪_{BB}`$, $`𝒪_{WW}`$, and $`𝒪_{BW}`$ can diminish the Higgs decay width into $`\gamma \gamma `$ and consequently this Higgs signature will not be observed neither in gluon–gluon fusion nor in VBF. In this case, the VBF Higgs production can be seen in the $`\tau ^+\tau ^{}`$ channel. If the Higgs is really discovered in this channel, the analysis of the $`\gamma \gamma `$ channel is a strong sign of the existence of new Higgs interactions. Of course, the constraints on the anomalous couplings are meaningless if the Higgs is not seen in any channel. For the operators $`𝒪_W`$, $`𝒪_B`$, $`𝒪_{\mathrm{\Phi },1}`$, and $`𝒪_{\varphi ,2}`$ the main effect is to modify the production cross section, affecting equally the Higgs signal in all decay modes. Therefore, the bounds on these operators make sense only if the Higgs production via gluon–gluon fusion is observed. ## IV Conclusions In this work we have studied the sensitivity of the LHC collider to new physics in the Higgs sector of the SM. In particular we have concentrated on new signals associated to the decay of an intermediate mass Higgs boson in two photons both in gluon–gluon fusion (1) as well as in gauge–boson fusion (2). We have shown that the LHC will be able to expand considerably the present sensitivity on the dimension–six Higgs couplings that modify the $`H\gamma \gamma `$ vertex, being able to probe new physics scales as large as 2.2 TeV, provided that the Higgs is observed. In both channels our result show that of the LHC is sensitive to new physics scales beyond the present constraints originating from their contribution at tree level to the LEP I and low energy observables. For the effective operators that do not change the Higgs coupling to photons, the LHC possible bounds are weaker than the ones presently available, with the exception of the “super–blind” operator $`𝒪_{\mathrm{\Phi },2}`$. We have also found that due to the presence of a sizeable interference between the anomalous and the SM contributions there is the distinct possibility that the anomalous Higgs interactions dilute its decay into two photons and the Higgs may be not observable in the above reactions. Thus the observation of the Higgs in other decay channel, namely $`\tau ^+\tau ^{}`$, would imply the existence of new physics in the Higgs couplings to gauge bosons at a characteristic scale of 0.7–1.4 TeV. ## Acknowledgments M. C. G.-G. is thankful to the IFT for their kind hospitality during her visit. This work was supported in part by the Director, Office of Science, Office of High Energy and Nuclear Physics, Division of High Energy Physics of the U.S. Department of Energy under Contract DE-AC03-76SF00098 and in part by Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), by Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP), and by Programa de Apoio a Núcleos de Excelência (PRONEX). It was also supported by Spanish DGICYT under grants PB95-1077 and PB98-0693, and by the European Union TMR network ERBFMRXCT960090.
warning/0001/astro-ph0001009.html
ar5iv
text
# THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES ## 1 Introduction X-ray radiation from a spiral galaxy consists of emission from discrete X-ray sources, hot gaseous media, and in some cases the active galactic nucleus (AGN); for a review, see Fabbiano (1988b). Majority of the discrete X-ray sources are presumably accreting collapsed objects, in particular low-mass X-ray binaries (LMXBs; a close binary consisting of a weakly-magnetized neutron star and a low-mass star). In fact, the X-ray emission from M31, with the total 2–10 keV luminosity of $`5\times 10^{39}`$ ergs s<sup>-1</sup>, is dominated by some 100 LMXBs (Van Speybroeck et al. 1979; Fabbiano, Trinchieri & van Speybroeck 1987; Makishima et al. 1989). The most luminous ones in M31 have X-ray luminosities close to the Eddington limit for a 1.4 $`M_{}`$ neutron star, $`L_\mathrm{E}^{\mathrm{NS}}=2\times 10^{38}`$ ergs s<sup>-1</sup> (Supper et al. 1997). Similarly, discrete X-ray sources in M33 have X-ray luminosities below $`L_\mathrm{E}^{\mathrm{NS}}`$ (Long et al. 1996), except the one (called X-8) to be discussed later. In many other nearby spiral galaxies, however, there reside point-like off-center X-ray sources of which the X-ray luminosity significantly exceeds $`L_\mathrm{E}^{\mathrm{NS}}`$ (e.g., Long 1982; Fabbiano 1988b; Marston et al. 1995; Colbert & Mushotzky 1999). Actually, the X-ray luminosity function of point-like non-AGN sources in nearby spiral galaxies derived with ROSAT extends well beyond $`L_\mathrm{E}^{\mathrm{NS}}`$ and reaches even $`2\times 10^{40}`$ ergs s<sup>-1</sup> (Read et al. 1997). Although some of these luminous X-ray objects are identified with young supernova remnants (e.g., Fabbiano & Trinchieri 1987; Schlegel 1994), others have not been identified securely in other wavelengths. These luminous non-AGN and non-supernova-remnant X-ray objects may be called “ultra-luminous compact X-ray sources” (hereafter abbreviated as ULXs). The ULXs cannot generally be explained as collections of multiple sources with luminosity below $`L_\mathrm{E}^{\mathrm{NS}}`$ each, since many of the ULXs exhibit significant time variability. Statistical calculations show that these objects are not likely to be less luminous foreground objects or background AGNs, either (Fabbiano 1988b). Assuming that ULXs are single compact objects powered by mass accretion, there are basically two alternative ways of explaining them. One is to regard a ULX as a sub-Eddington binary involving a relatively massive black hole (BH) with mass up to $`100M_{}`$, because the Eddington limit for a body of mass $`M`$ is given, when electron scattering dominates the opacity, as $$L_\mathrm{E}=1.5\times 10^{38}(M/M_{})\mathrm{ergs}\mathrm{s}^1.$$ (1) Here we assume a spherical symmetry, and the hydrogen to helium composition ratio of 0.76:0.23 by weight. The BH interpretation may be supported by a signature of ULXs as a young population; they are more abundant in spirals with later morphological types, and they often locate in spiral arms or HII regions (Fabbiano 1988b). However, there is not necessarily a consensus on the presence of $`100M_{}`$ BHs. The other is to regard a ULX as a binary where the X-ray emission is highly collimated toward us, or a super-Eddington radiation is maintained. Then, the object can be accreting BHs in an ordinary mass range, or even LMXBs. This interpretation has some empirical support, because LMC X-2, an LMXB in the Large Magellanic Cloud, exhibits a marginally super-Eddington luminosity of $`2\times 10^{38}`$ ergs s<sup>-1</sup> (Tanaka 1997). However, the difficulty of this alternative is that there is no known mechanism of producing such an efficient X-ray beaming, or a steady super-Eddington radiation. Thus, the nature of ULXs has remained a big mystery in the modern X-ray astrophysics. It affects our basic understanding of X-ray properties of normal spiral galaxies, because a single ULX could overwhelm the summed contribution from the entire ordinary LMXBs in the galaxy. Furthermore, a ULX located near the galaxy center would (and actually does; Mizuno et al. 1999; Colbert & Mushotzky 1999) mimic a low-luminosity AGN. Clearly, a key to the issue is provided by detailed X-ray spectroscopy, which has been enabled by ASCA (Tanaka, Inoue, & Holt 1994). In § 2, we briefly summarize the “multi-color disk” (MCD) modeling of emission from optically-thick accretion disks around BHs. We show in § 3 that this particular spectral model gives adequate description of the ASCA spectra of a fair number of ULXs, on condition that the disk temperature is made rather high up to $`2`$ keV. We show in § 4 that the high disk temperature is also observed from Galactic BHB binaries (BHBs) with superluminal jets. In § 5, we point out a particular problem associated with the BH interpretation of ULXs; that they exhibit disk temperatures which are too high for the large BH masses required by their high X-ray luminosities. We present various attempts to solve the problem, finally arriving at a hypothesis of spinning stellar-mass BHs. Throughout the paper, errors and uncertainties refer to 90% confidence limits unless otherwise stated. ## 2 The Multicolor-Disk (MCD) Formalism of Accretion Disk Emission ### 2.1 Assumed spectral model The X-ray emission from a BHB, in so called soft (or high) state, consists of a bright soft component (White & Marshall 1984) with a characteristic temperature of $`1`$ keV, and a hard tail. The soft component is thought to originate from an optically thick accretion disk around the BH, of which the standard accretion disk model by Shakura & Sunyaev (1973) provides a prototypical formalism. There are in fact many recent progresses in the theory of accretion disks, including the concept of slim accretion disk (Abramowicz et al. 1988) and advection-dominated accretion flow (ADAF; Narayan & Yi 1995). A unified picture is presented by by Esin, McClintock, & Narayan (1997). Nevertheless, as long as the high (soft) state of BHBs are concerned, there is no strong observational evidences for deviations from the standard Shakura & Sunyaev type accretion disk. Therefore, we take the standard accretion disk as a start point, and employ its simple mathematical approximation called multi-color disk blackbody (MCD) model (Mitsuda et al. 1984). The model is known to give a successful physical description of the X-ray spectra of many BHBs (Makishima et al. 1986; Ebisawa et al. 1993; Mineshige et al. 1994; Tanaka & Lewin 1995; Tanaka & Shibazaki 1996; Zhang, Cui & Chen 1997; Dotani et al. 1997; Kubota et al. 1998). The MCD model is a superposition of many blackbody elements, up to a certain maximum color temperature (sometimes simply called dsik temperature) $`T_{\mathrm{in}}`$ that is expected to occur near the innermost disk boundary. The local disk temperature $`T(R)`$ is considered to scale as $$T(R)R^{3/4},$$ (2) where $`R`$ is radial distance from the BH. Although an exact treatment of the disk emission must take into account the inner disk boundary condition and relativistic effects (e.g. Ebisawa, Mitsuda & Hanawa 1991), the MCD formalism is known to give a reasonable approximation (e.g., Ebisawa 1991; Dotani et al. 1997; Kubota et al. 1998), providing a simple physical insight and a straightforward comparison among different BHs. In particular, the validity of equation (2) has been confirmed observationally in the case of Nova Muscae (Mineshige et al. 1994). ### 2.2 Assumed geometry Assuming a flat disk geometry with inclination $`i`$ and distance $`D`$, the bolometric luminosity of an optically thick accretion disk is given as $$L_{\mathrm{bol}}=2\pi D^2f_{\mathrm{bol}}(\mathrm{cos}i)^1,$$ (3) where $`f_{\mathrm{bol}}`$ is the bolometric flux calculated from the observed band-limited flux, via bolometric correction using the MCD model. This $`L_{\mathrm{bol}}`$ is related to the maximum disk color temperature $`T_{\mathrm{in}}`$ and the innermost disk radius $`R_{\mathrm{in}}`$ as $$L_{\mathrm{bol}}=4\pi (R_{\mathrm{in}}/\xi )^2\sigma (T_{\mathrm{in}}/\kappa )^4.$$ (4) Here $`\sigma `$ is the Stefan-Boltzmann constant, $`\kappa 1.7`$ (e.g. Shimura & Takahara 1995) is ratio of the color temperature to the effective temperature, or “spectral hardening factor”, and $`\xi `$ is a correction factor reflecting the fact that $`T_{\mathrm{in}}`$ occurs at a radius somewhat larger than $`R_{\mathrm{in}}`$. Kubota et al. (1998) give $`\xi =(3/7)^{1/2}(6/7)^3=0.412`$. Because $`T_{\mathrm{in}}`$ is directly observable and $`L_{\mathrm{bol}}`$ can be estimated through equation (3), we can solve equation (4) as $$R_{\mathrm{in}}=\xi \kappa ^2\sqrt{\frac{L_{\mathrm{bol}}}{4\pi \sigma T_{\mathrm{in}}^4}}=\xi \kappa ^2\frac{D}{\sqrt{\mathrm{cos}i}}\sqrt{\frac{f_{\mathrm{bol}}}{2\sigma T_{\mathrm{in}}^4}},$$ (5) which provides us with the basic tool to estimate the innermost disk radius $`R_{\mathrm{in}}`$. When $`R_{\mathrm{in}}`$ is thus estimated, we may identify it with the radius of the last stable Keplerian orbit (e.g. Makishima et al. 1986; Tanaka & Shibazaki 1996). For a non-spinning BH of mass $`M`$, this condition takes place at $`3R_\mathrm{S}`$, where $$R_\mathrm{S}=\frac{2GM}{c^2}=2.95\left(\frac{M}{M_{}}\right)\mathrm{km}$$ (6) is the Schwarzschild radius, $`G`$ is the gravitational constant, and $`c`$ is the light speed. Thus, we may in general write as $$R_{\mathrm{in}}=3\alpha R_\mathrm{S}=8.86\alpha \left(\frac{M}{M_{}}\right)\mathrm{km}$$ (7) using a positive parameter $`\alpha `$, with $`\alpha =1`$ corresponding to the Schwarzschild BH. Conversely, we may obtain an “X-ray estimated” BH mass as $$MM_{\mathrm{XR}}=R_{\mathrm{in}}/8.86\alpha M_{}.$$ (8) ### 2.3 Temperature-luminosity relations By substituting equation (7) into equation (4), we obtain a temperature-luminosity relation for an accretion disk around a BH of mass $`M`$, as $$L_{\mathrm{bol}}=7.2\times 10^{38}\left(\frac{\xi }{0.41}\right)^2\left(\frac{\kappa }{1.7}\right)^4\alpha ^2\left(\frac{M}{10M_{}}\right)^2\left(\frac{T_{\mathrm{in}}}{\mathrm{keV}}\right)^4\mathrm{erg}\mathrm{s}^1.$$ (9) Note that the mass accretion rate is unspecified in this relation. It may be convenient here to write $`L_{\mathrm{bol}}`$ as $$L_{\mathrm{bol}}=\eta L_\mathrm{E}\eta M,$$ (10) where a non-dimensional parameter $`\eta `$ means the disk bolometric luminosity normalized to the Eddington luminosity of equation (1). Using this and equation (1), we can then eliminate $`M`$ from equation (9), to obtain another form of temperature-luminosity relation as $$L_{\mathrm{bol}}=3.1\times 10^{39}\left(\frac{\xi }{0.41}\right)^2\left(\frac{\kappa }{1.7}\right)^4\alpha ^2\eta ^2\left(\frac{T_{\mathrm{in}}}{\mathrm{keV}}\right)^4\mathrm{erg}\mathrm{s}^1.$$ (11) Note that the BH mass is unspecified in this relation. By equating equations (9) and (11), and solving it for $`T_{\mathrm{in}}`$, we obtain an important scaling of $$T_{\mathrm{in}}=1.2\left(\frac{\xi }{0.41}\right)^{1/2}\left(\frac{\kappa }{1.7}\right)\alpha ^{1/2}\eta ^{1/4}\left(\frac{M}{10M_{}}\right)^{1/4}\mathrm{keV}.$$ (12) Thus, a heavier BH tends to show a lower disk temperature. ## 3 Analysis of the ULX Spectra Although ASCA has a rather limited angular resolution, we can utilize its wide-band (0.5–10 keV) spectroscopic capability for those ULXs which are relatively isolated from other X-ray sources in the target galaxies, particularly when helped with the ROSAT and/or Einstein images. The ULXs so far studied with ASCA include; the central source (X-8) in M33 (Takano et al. 1994), so called Source 1 in IC 342 (Okada et al. 1998), source X-6 in M81 (Kohmura 1994; Uno 1997), two sources (A and B) in NGC 1313 (Petre et al. 1994), one (called X-1) in the newly discovered spiral galaxy Dwingeloo 1 (Reynolds et al. 1997), and two ULXs in NGC 4565 (Mizuno et al. 1999). Below we review these results, and in some cases, reanalyze the published spectra or newly process the ASCA data. In Table 1, we give a brief summary of these results. ### 3.1 M33 X-8 The source X-8 (Long et al. 1981; Gottwald, Pietsch & Hasinger 1987; Trinchieri, Fabbiano & Peres 1988) at the center of M33 has a 2–10 keV luminosity reaching $`1\times 10^{39}`$ ergs s<sup>-1</sup>, which makes it the most luminous X-ray source in the Local Group. In spite of the positional coincidence with the optical nucleus, X-8 is unlikely to be a low-lumiosity AGN, because of the 106 day periodicity found in its ROSAT lightcurve (Dubus et al. 1997), the lack of nuclear activity in M33 in other wavelengths, and the stellar kinematic evidence for the absence of a massive BH (Lauer et al. 1998). Then, considering its luminosity exceeding $`L_\mathrm{E}^{\mathrm{NS}}`$, X-8 may be interpreted alternatively as a ULX. Takano et al. (1994), and subsequently Mizuno (2000), analyzed the spectra of M33 X-8, taken with the Gas Imaging Spectrometer (GIS; Ohashi et al. 1996; Makishima et al. 1996) and the Solid-State Imaging Spectrometer (SIS; Burke et al. 1994; Yamashita et al. 1997) onboard ASCA. The spectra have been described successfully with an MCD model of $`T_{\mathrm{in}}1.15`$ keV, as quoted in Table 1, on condition that an additional power-law component is included to describe a hard tail above $`5`$ keV. Although the MCD emission is associated with either LMXBs or BHBs, X-8 is too luminous for an ordinary LMXB, and its power-law hard tail is characteristic of BHBs. Takano et al. (1994) and Mizuno (2000) therefore conclude that X-8 is a mass-accreting stellar-mass ($`10M_{}`$) BH. These results have been reconfirmed by Colbert & Mushotzky (1999). Quantitatively, if we adopt $`i=0^{}`$, the measured values of $`T_{\mathrm{in}}=1.15`$ keV and $`L_{\mathrm{bol}}=3.8\times 10^{38}`$ ergs s<sup>-1</sup> of X-8 (Table 1) can be consistently reproduced by substituting $`M=5.5M_{}`$ and $`\eta =0.46`$ into equations (9) and (11), together with canonical values of $`\alpha =1`$, $`\xi =0.41`$, and $`\kappa =1.7`$. If instead, e.g., $`i=60^{}`$ is adopted, $`L_{\mathrm{bol}}`$ is doubled to $`7.6\times 10^{38}`$ ergs s<sup>-1</sup>, and the solution becomes $`M7.8M_{}`$ and $`\eta 0.65`$. These estimates reconfirm the inference made by Takano et al. (1994). We are then inspired to regard M33 X-8 as the nearest and prototypical ULX. ### 3.2 IC 342 Source 1 Another case of significant interest is “Source 1” (Fabbiano & Trinchieri 1987; Bregman et al. 1993; Makishima 1994) located in a spiral arm of the galaxy IC 342. As reported by Okada et al. (1998), its spectrum obtained with ASCA can be expressed by an MCD model of maximum color temperature $`T_{\mathrm{in}}1.8`$ keV, whereas the power-law and Bremsstrahlung models disagree with the data. We here reanalyze the same ASCA data in the same way as Okada et al. (1998). Detailed study of the time variation is reported elsewhere (Mizuno 2000). The derived SIS and GIS spectra of Source 1 are presented in Figures THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES. Here and hereafter, all the ASCA spectra are presented after background subtraction, but without removing the instrumental responses. We fitted the SIS/GIS spectra jointly, using three different spectral models; power-law, thermal Bremsstrahlung, or MCD, all modified at low energies with photoelectric absorption. Then, as shown in Table 2, only the MCD model among the three has given an acceptable joint fit, in agreement with Okada et al. (1998). Predictions of the best-fit power-law and MCD models are shown respectively in Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIESa and Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIESb as solid histograms. The successful MCD fit encourages us to interpret this source as an accreting BH, like M33 X-8. However, the BH must be fairly massive (Table 1) for the high source luminosity to stay within the Eddington limit. More importantly, the measured disk temperature, $`T_{\mathrm{in}}1.8`$ keV, much exceeds those (0.4–1.0 keV) typically found among established BHBs in the Milky Way and Magellanic clouds, in contradiction to the prediction of equation (12) that more massive BHs should have lower disk temperatures. These problems, pointed out by Okada et al. (1998), Mizuno et al. (1999) and Mizuno (2000), are found commonly with other ULXs, as revealed hereafter. We further tried two other spectral models of convex shapes. One is unsaturated Comptonization (UC) model by Sunyaev & Titarchuk (1980), describing thermal electrons of temperature $`T_\mathrm{e}`$ Compton up-scattering soft seed photons into X-rays. It has given a fairly good fit to the spectra of IC 342 Source 1, although not as good as the MCD model (Table 2). This is of no surprise, because the MCD and UC models can take nearly identical shapes (Ebisawa 1991), when $`T_\mathrm{e}`$ is made close to $`T_{\mathrm{in}}`$ and the optical depth for scattering, $`\tau _{\mathrm{es}}`$, is adjusted. The other is a broken power-law, often used to approximate synchrotron radiation spectra emitted from jet-dominated sources. This model, modified with photoelectric absorption, gave a value of chi-square in between those from the MCD and Bremsstrahlung fits (Table 2). Physical meanings of these alternative modelings are examined in § 5.1. ### 3.3 M81 X-6 Because of the explosion of the supernova SN1993J, the region around the spiral galaxy M81 has been observed repeatedly with ASCA. These observations have provided valuable information on SN1993J (Kohmura et al. 1994; Kohmura 1994; Uno 1997), the low-luminosity AGN in M81 (Ishisaki et al. 1996; Iyomoto 1999), and the previously known ULX called M81 X-6 (Fabbiano 1988a) which lies close to SN1993J. A merit of X-6 is the accurate knowledge of the distance to the host galaxy, $`3.6\pm 0.3`$ Mpc (Freedman et al. 1994). Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIESc presents the SIS spectrum of X-6 (plus SN1993J), produced by Uno (1997) who screened the data through standard procedure, accumulated photons falling within $`1.^{}5`$ of X-6, and then subtracted background as well as contamination from the M81 nucleus. The spectrum co-adds two observations, conducted on 1994 April 1 through 2 for 28 ks and on 1994 October 21 through 22 for 37 ks. The supernova had faded away significantly by these observations, and its contribution to the present spectrum can be expressed by a steep power law having a photon index of 3.0 and a photon flux density of $`1.0\times 10^4`$ ergs s<sup>-1</sup> cm<sup>-2</sup> keV<sup>-1</sup> at 1 keV (Uno 1997). Therefore, the contamination is limited to energies below $`2`$ keV. In the present study of M81 X-6 we utilize this SIS spectrum only, since the poorer spatial resolution of the GIS increases the supernova contamination. Further data analysis of X-6, including the intensity-sorted spectroscopy, is reported elsewhere (Mizuno 2000). We fitted the SIS spectrum with five alternative spectral models; power-law, thermal Bremsstrahlung, MCD, UC, or broken power-law, all modified at low energies with photoelectric absorption. The SN1993J contribution was taken into account as a fixed model component mentioned above. As summarized in Table 2, the results have been qualitatively similar to those from IC 342 Source 1: the power-law model failed, the Bremsstrahlung fit is much better but still unacceptable, while the remaining three models have been generally successful. In any case, reservation should be put on the obtained $`N_\mathrm{H}`$, because of the supernova contamination below $`2`$ keV. Based on the MCD fit and equation (3), the disk bolometric luminosity becomes as given in Table 1; clearly, X-8 is a ULX, and thanks to the accurate distance, its luminosity is considerably more reliable than that of IC 342 Source 1. The BH mass required by the Eddington limit (Table 1) may be reasonable as long as $`i`$ is rather small. However, in reference to equation (12), the measured value of $`T_{\mathrm{in}}`$ (Table 2) is too high for the rather high BH mass. The high value of $`T_{\mathrm{in}}`$ suggests the presence of a separate harder spectral component, e.g., a power-law hard tail normally seen from soft-state BHBs just as is the case with M33 X-8 (§ 3.1). Accordingly, we refitted the X-6 spectrum by a sum of an MCD continuum, and a hard power-law of which the photon index is fixed at 2.3, typical of high-state BHBs. However the data did not require the hard component, with its 0.5–10 keV flux being at most $`18\%`$ of the total source flux in the same range. Furthermore, inclusion of the hard tail component at the allowed upper limit increased the MCD temperature slightly, from 1.48 keV to 1.52 keV. Alternatively, a hard blackbody (BB) component of temperature $`2`$ keV might be present, like in X-ray luminous LMXBs (Mitsuda et al. 1984; Tanaka 1997). Accordingly, we replaced the power-law hard tail with a blackbody of temperature fixed at 2.0 keV. However, the data did not require the BB component either, and its contribution to the overall 0.5–10 keV flux turned out to be at most 14%. ### 3.4 Dwingeloo 1 X-1 Through ASCA observations of the recently discovered nearby spiral galaxy Dwingeloo 1 located behind the Milky Way, Reynolds et al. (1997) detected a ULX, and named it X-1. Its spectrum, presented in Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIESd, was described well with several alternative models because of rather poor data statistics, but the MCD fit was not attempted by Reynolds et al. (1997). We hence reanalyze the same GIS and SIS spectra, in terms of the MCD model as well as power-law and Bremsstrahlung models. As shown in Table 2, all these models have been acceptable, with the power-law and Bremsstrahlung results being consistent with those reported by Reynolds et al. (1997). The MCD model is somewhat preferred among the three, because the absorption associated with it agrees with the line-of-sight Galactic value of $`N_\mathrm{H}=7\times 10^{21}`$ cm<sup>-1</sup> (see Reynolds et al. 1997); the other two models require excess absorption which could be artificial. The source again exhibits a quite high disk temperature, $`1.8`$ keV. ### 3.5 NGC 1313 Source B The spiral galaxy NGC 1313 is known to contain three bright X-ray objects; one identified with SN1978k, another called Source A located near the nucleus, and the other called Source B associated with a spiral arm (Fabbiano & Trinchieri 1987). Here we focus on Source B which is regarded as a ULX (Petre et al. 1994; Colbert et al. 1995); Source A could be composite. The NGC 1313 region was observed twice with ASCA, on 1993 July 12 and 1995 November 29. The first ASCA observation was reported by Petre et al. (1994), who found that the Source B spectra can be described well with a Bremsstrahlung or a Raymond-Smith model (Raymond & Smith 1977), whereas the power-law model is rejected. However, they did not try an MCD fit. We therefore reanalyzed the same SIS and GIS spectra of NGC 1313 Source B, which are reproduced in our Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIESe. As summarized in Table 2, the power-law fit is unacceptable, and the Bremsstrahlung fit is marginally acceptable; the derived parameters agree with those of Petre et al. (1994). The MCD fit is fully acceptable, and the obtained disk temperature is again rather high. We have also analyzed the ASCA data of Source B from the second ASCA observation, by screening the data in the standard way, and accumulating the SIS and GIS photons separately over a common region of radius $`3^{}`$ centered on the source. We subtracted the GIS background using blank-sky data, and the SIS background using source-free regions of the same on-source data. As shown in Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIESf and Table 2, the MCD fit is again most successful. As listed in Table 2, the source flux decreased by a factor of 2.5 from 1993 to 1995, meanwhile the disk temperature decreased by a factor of $`1.4`$. There is marginal evidence (Table 1) for an increase in the disk radius as the source became fainter. This phenomenon is reported in further detail by Mizuno (2000). ### 3.6 Two ULXs in NGC 4565 Finally, we refer to the recent results by Mizuno et al. (1999), who studied X-ray emission from the edge-on spiral galaxy NGC 4565 using ASCA and ROSAT. The galaxy hosts two bright point-like X-ray sources, the fainter one coincident with the nucleus (hereafter called center source) while the brighter one displaced $`2`$ kpc above the galaxy disk (hereafter off-center source), both showing high bolometric luminosities (Table 1). Their ASCA spectra have been described successfully by the MCD model with rather high disk temperatures (Table 1), and marginally well by the Bremsstrahlung model, but the power-law fit was unacceptable for the off-center source. Based on the absence of significant spectral absorption, Mizuno et al. (1999) conclude that the center source is in fact a ULX in NGC 4565 rather than an AGN. ## 4 Galactic Transients with Super-luminal Jets Although no ULXs are known in our Galaxy or M31, the issue of rather high disk temperature is also seen among some Galactic transient sources (Zhang, Cui & Chen 1997). The best examples are GRO J1655$``$40 and GRS 1915+105, both exhibiting super-luminal radio jets which have allowed accurate determinations of the system inclination and distances. We here analyze the ASCA data of these two sources, because their resemblance to ULXs is considered to be meaningful. As they are too bright for the SIS, we utilize the GIS data only. ### 4.1 GRO J1655$``$40 The transient source GRO J1655$``$40 has $`i=69.^{}50\pm 0.^{}08`$ (Orosz et al. 1997) and $`D=3.2`$ kpc (Hjellming & Rupen 1995). It has a secure BH secondary, with the optically determined BH mass $`M_{\mathrm{opt}}`$ of $`7.0\pm 0.2M_{}`$ (Orosz et al. 1997) or $`5.57.9M_{}`$ (Shahbaz et al. 1999). The RXTE observations are reported by Mendez et al. (1998). We reanalyzed the deadtime-corrected ASCA GIS spectrum of this source, acquired on 1995 August 15–16 and published by Ueda et al. (1998) as their Figure 2a; they fitted it by a power-law model with a high-energy cutoff, to focus upon the iron absorption feature. We fit it with the MCD model, plus a power-law of which the photon index is fixed at 2.36 after Zhang et al. (1997). The iron absorption feature was represented with a spectral notch at 6.95 keV according to Ueda et al. (1998). We have obtained an acceptable fit with $`\chi ^2/\nu =91.6/91`$, in spite of very high signal statistics. This reconfirms the correctness of the spectral modeling, and the accuracy of the instrumental calibration (Makishima et al. 1996). The derived MCD parameters are given in Table 1; the observed disk bolometric luminosity is $`1/8`$ of the Eddington limit for $`M_{\mathrm{opt}}`$. The absorption turned out to be $`(9\pm 1)\times 10^{21}`$ cm<sup>-2</sup>. Through equation (5), we obtain $`R_{\mathrm{in}}\sqrt{\mathrm{cos}i}=11.4`$ km, which in turn yields an X-ray mass estimate of $`M_{\mathrm{XR}}2.2M_{}`$ via equation (8). This is in fact much lower than the $`M_{\mathrm{opt}}`$ quoted above. This $`M_{\mathrm{XR}}`$ does not change by more than 10% even when forcing the power-law normalization to be zero (then the fit becomes unacceptable), or leaving the power-law slope free. Thus, the disk temperature is apparently too high, and the disk radius is too small, just like in the case of the ULXs described in § 3. We also attempted the spectral fitting using general relativistic accretion disk (GRAD) model (Hanawa 1989; Ebisawa, Mitsuda & Hanawa 1991), which improves over the MCD formalism by taking into account the inner boundary condition and relativistic effects. It has the BH mass $`M`$ and the accretion rate $`\dot{M}`$ as free parameters, with $`i`$ and $`\kappa `$ adjustable auxiliary parameters. We fixed $`i`$ at $`69^{}.5`$ and $`\kappa `$ at 1.7, and again fixed the power-law photon index at 2.36. As shown in Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIESa, an acceptable ($`\chi ^2/\nu =70.4/91`$) fit has been obtained, with $`M=2.91\pm 0.10M_{}`$ and $`\dot{M}=(2.64_{0.07}^{+0.03})\times 10^{18}`$ g s<sup>-1</sup>. The value of $`N_\mathrm{H}`$ was not much different from that of the MCD fit (Table 1). Thus, the mass estimate somewhat increases, but it still remains significantly below $`M_{\mathrm{opt}}`$. These results reconfirm the report by Zhang, Cui & Chen (1997), who used the ASCA data acquired on the same occasion to derive $`R_{\mathrm{in}}=22`$ km, employing also $`\kappa =1.7`$ and relativistic corrections to the MCD results. This has led Zhang, Cui & Chen (1997) to propose that the BH in GRO J1655$``$40 is rapidly spinning, and the accretion disk is prograde to the BH rotation. We come back to the issue in § 5.6. ### 4.2 GRS 1915+105 Similarly, the transient GRS 1915+105 is a promising BH candidate, and has been studied by a number of authors including Belloni et al. (1997) and Zhang, Cui & Chen (1997). It has $`i=70^{}`$ and $`D=12.5`$ kpc (Mirabel & Rodriguez 1994), although the optical estimate of the BH mass is not yet available. Belloni et al. (1997) report $`L_{\mathrm{bol}}1.4\times 10^{39}`$ ergs s<sup>-1</sup>, which indicates $`M_\mathrm{E}9M_{}`$. Then, the disk temperature should be $`1.2`$ keV, from equation (12). However, in reality, a much higher value of $`T_{\mathrm{in}}=2.27\pm 0.03`$ keV was obtained with RXTE during the outburst peak (Belloni et al. 1997). We analyzed the ASCA GIS data of GRS 1915+105, acquired on 1995 April 20 for 18 ks and described partially by Ebisawa (1997) and Kotani et al. (2000). The GIS data were screened in the standard way, and the spectra from GIS2 and GIS3 were added together after appropriate dead-time corrections (Makishima et al. 1996) as presented in Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIESb. Background is completely negligible. It has been fitted successfully ($`\chi ^2/\nu =82.5/81`$) with the MCD continuum, on condition that it is multiplied by the following six factors; (i) a photoelectric absorption by $`N_\mathrm{H}=(4.3\pm 0.1)\times 10^{22}`$ cm<sup>-2</sup>; (ii) a smeared Fe-K edge with the center energy $`7.12\pm 0.08`$ keV, optical depth $`1.6\pm 0.2`$, and a fixed width of 10 keV; (iii) a narrow K<sub>α</sub> line of He-like and/or H-like iron, centered at $`6.98\pm 0.03`$ keV with an equivalent width (EW) of $`55\pm 6`$ eV, (iv) a narrow Fe K<sub>β</sub> line at $`8.15\pm 0.09`$ keV with an EW of $`20\pm 6`$ eV, (v) a third narrow absorption line at $`1.90\pm 0.02`$ keV with an EW of $`19\pm 2`$ eV, attributable to K<sub>α</sub> line of H-like and/or He-like Si; and (vi) a broad absorption line at $`2.52\pm 0.04`$ keV, having width of 0.8 keV and an EW of $`22\pm 3`$ eV, presumably a blend of Si-K<sub>β</sub> and S-K<sub>α</sub> lines. No power-law tail was required. The derived MCD parameters are given in Table 1. The bolometric luminosity becomes $`1.2\times 10^{39}`$ ergs s<sup>-1</sup>, which is similar to that measured with RXTE. The value of $`M_{\mathrm{XR}}=3.5\pm 0.4M_{}`$ (Table 1) falls significantly below $`M_\mathrm{E}`$. Thus, as reported by Belloni et al. (1997), GRS 1915+105 exhibit a very high $`T_{\mathrm{in}}`$ and too small $`R_{\mathrm{in}}`$, like GRO J1655$``$40 and the ULXs. ## 5 Discussion ### 5.1 Spectral properties of the sample ULXs We have studied the ASCA spectra of seven ULXs, partially referring to published results. The Bremsstrahlung model failed for least in two spectra (IC 342 Source 1, and M81 X-6), and the power-law fit is generally unacceptable except in the faintest objects. In contrast, all the spectra have been described successfully by the MCD model (plus a hard power-law in M33 X-8) with a relatively high disk temperature in the range 1.0–1.8 keV. Thus, the seven ULXs form a rather homogeneous sample, and hence are thought to be the same type of objects. With a possible exception of Dwingeloo 1 X-1 with rather unconstrained spectral modeling, none of them are background AGNs, since no AGN is known to exhibit the MCD-type spectra in the ASCA range. We found that the UC model and the broken power-law model can also describe fairly well the two ULX spectra of high statistics, those of IC 342 Source 1 (§ 3.2) and M81 X-6 (§ 3.3). These models will give acceptable fits to the remaining ULX spectra that have poorer statistics. However if the UC interpretation were correct, the spectrum should bear a prominent Fe-K edge absorption feature at $`7`$ keV, because the large optical depth for electron scattering required by the two ULX spectra, $`\tau _{\mathrm{es}}=2030`$, implies a huge hydrogen column density of $`5\times 10^{25}`$ cm<sup>-2</sup>, and because the heavy ions would not be completely ionized at the suggested electron temperature of $`T_\mathrm{e}1`$ keV. Evidently, such a feature is absent in the observed ULX spectra, making the UC interpretation of the ULX spectra physically unrealistic. Similarly, the broken power-law model (appropriate for synchrotron emission) is physically unrealistic, since the break in the photon index of 1.1–1.5 required by the two sources is too abrupt to be interpreted in terms of synchrotron cooling. These considerations strongly suggest that the emission from these ULXs actually originates in optically thick accretion disks as approximated by the MCD modeling, and hence the objects are either BHBs or LMXBs. The disk temperatures of the ULXs are closer to those of luminous LMXBs (Mitsuda et al. 1984; Tanaka 1997), rather than to those of BHBs. However, in order for LMXBs to appear as ULXs, the emission must be highly anisotropic, and/or super-Eddington by nearly two orders of magnitude. Furthermore, the hard BB component characterizing the LMXB spectra is absent, at least in the M81 X-6 spectrum, with an upper limit of $`14\%`$ (§ 3.3). This is considerably lower than the fractional BB contribution of $`50\%`$ observed from the most luminous LMXBs (Mitsuda et al. 1984). We therefore conclude that the ULXs are accreting BHBs rather than a version of LMXB. This makes a very important step towards understanding these enigmatic objects. ### 5.2 Problems with the accretion-disk interpretation In Table 1, we compile the disk bolometric luminosity $`L_{\mathrm{bol}}`$ via equation (3), the minimum BH mass $`M_\mathrm{E}`$ necessary to make $`L_{\mathrm{bol}}`$ stay within the Eddington limit, and the disk inner radius $`R_{\mathrm{in}}`$ given by equation (5). The assumed distance $`D`$ to each host galaxy is also given there. For all ULXs, we have assumed the face-on inclination, i.e. $`i0`$, that makes $`M_\mathrm{E}`$ lowest. This assumption is not necessarily too arbitrary, because face-on systems must have higher fluxes for a given bolometric luminosity according to equation (3), and hence more selectively detectable, than those with larger inclinations; we hence retain the assumption of $`i0`$ for the ULXs throughout the paper. When $`D`$ and $`i`$ are changed, these quantities scale as $$M_\mathrm{E}D^2(\mathrm{cos}i)^1,M_{\mathrm{XR}}R_{\mathrm{in}}D(\mathrm{cos}i)^{1/2}.$$ (13) In Table 1, $`M_\mathrm{E}`$ takes rather high values, reaching $`7080M_{}`$ for IC 342 Source 1 and NGC 4565 off-center source. Although there is not yet a consensus on the formation of BHs in such a mass range, some (e.g. Fryer 1999) argue that a star heavier than $`40M_{}`$ can directly form a BH without supernova explosion (and hence without losing much of the progenitor mass), and there are a fair number of observations of stars in the mass range of $`100150M_{}`$ (e.g. Krabbe et al. 1995). Therefore, the high BH mass may not be a big difficulty. A more profound problem is that, except M33 X-8 (and NGC 1313 Source B in 1995), the observed $`T_{\mathrm{in}}`$ is too high for the required high BH mass (Okada et al. 1998; Mizuno et al. 1999; Mizuno 2000), as mentioned in § 3. The high values are unlikely to be an artifact of the data analysis (§ 3), and substitution of these values into equation (12) would make the BH masses $`<5M_{}`$ even at the extreme of $`\eta =1`$. This makes self-inconsistent the BH interpretation of ULXs, where a rather high BH mass is inevitable. This problem is best visualized by Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES, to be called an X-ray “H-R diagram” for BHs, which summarizes various BHBs and ULXs on the plane of $`T_{\mathrm{in}}`$ vs. $`L_{\mathrm{bol}}`$. There, a family of grid lines with a positive slope $`(L_{\mathrm{bol}}T_{\mathrm{in}}^4)`$ represent the loci of constant $`M`$, expressed by equation (9). The other family of grid lines with a negative slope $`(L_{\mathrm{bol}}T_{\mathrm{in}}^4)`$ indicate the loci of constant $`\eta `$, expressed by equation (11). In order to consistently interpret an X-ray source as an accreting non-spinning BH, the data point must fall on the region of $`\eta <1`$ (non-violation of the Eddington limit) and $`M>3M_{}`$ (heavier than a neutron star) in Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES. The reality, however, is that six out of the seven ULXs (except M33 X-8) fall on the super-Eddington regime in this plot. As to NGC 1313 Source B, the 1993 data point is super-Eddington, whereas the 1995 point is marginally sub-Eddington. The problem may equivalently be stated in the following way. The observed high values of $`T_{\mathrm{in}}`$, together with equation (5), yield rather small values of $`R_{\mathrm{in}}`$ (Table 1). Then, the BH mass $`M_{\mathrm{XR}}`$, determined by equation (8), becomes uncomfortably low. Quantitatively, the derived $`M_{\mathrm{XR}}`$ falls by a factor of 1.6–6.0 short of the minimum BH mass $`M_\mathrm{E}`$ to satisfy the Eddington limit. The issue may be illustrated in Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES, to be called a “mass-radius diagram” for BHs, where we compare the values of $`R_{\mathrm{in}}`$ against $`M_\mathrm{E}`$ or $`M_{\mathrm{opt}}`$. It is important to note that the same problem is found also with the two Galactic jet sources studied in § 4. We accordingly plot these two objects as well in Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES (more than once for each object) and Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES. Indeed, in the H-R diagram of Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES, GRS 1915+105 falls on the super-Eddington regime, and GRO J1655$``$40 is located near the lower-mass boundary of a BH. Similarly in the mass-radius diagram of Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES, the two jet sources (both referring to our own data) deviate from the relation of ordinary BHBs (see § 5.4), and line up better with the ULXs. ### 5.3 Simple solutions As first-cut attempts to solve the above problem, let us consider trivial possibilities, namely revisions of the inclination $`i`$, the Eddington limit, or the distance. So far, we have assumed $`i=0`$. However, as is apparent from equation (13), the ratio of $`M_\mathrm{E}/M_{\mathrm{XR}}`$ increases with $`i`$ as $`(\mathrm{cos}i)^{1/2}`$. Therefore the problem of high $`M_\mathrm{E}/M_{\mathrm{XR}}`$ ratio gets worse if $`i`$ is increased from $`0`$. Thus, $`i`$ has already been selected to make the problem least severe. Larger values of $`i`$ would also make the BH mass still higher. Alternatively, the problems may result from our wrong application of the spherically-symmetric Eddington limit of equation (1) to the systems considered here, which are basically axi-symmetric. However, the critical upper-limit luminosity of an axi-symmetric BH system does not differ very much from the spherical one (e.g., Abramowicz et al. 1988; Beloborodov 1998). This justifies our use of equation (1). Yet another simple possibility is that the source distances are systematically overestimated: note $`M_\mathrm{E}/M_{\mathrm{XR}}D`$ from equation (13). However, in order to bring $`M_\mathrm{E}/M_{\mathrm{XR}}`$ down to unity, we have to reduce the distances to IC 342, M81, NGC 1313, and NGC 4565 by a factor of 6.0, 2.2, 3.1, and 3.7, respectively. Such large distance revisions would be unwarranted, especially for M81 of which the distance is known to within $`\pm 10`$% (Freedman et al. 1994). Therefore, this attempt is unsuccessful, as already noticed by Mizuno et al. (1999) regarding the ULXs in NGC 4565. ### 5.4 Possible errors in $`\xi \kappa ^2`$ As is obvious from equations (5) and (8), the $`M_\mathrm{E}`$ vs. $`M_{\mathrm{XR}}`$ discrepancy may alternatively be caused because our choice of $`\xi =0.41`$ (Kubota et al. 1998) and $`\kappa =1.7`$ (Shimura & Takahara 1995), or $`\xi \kappa ^2=1.18`$, is wrong. In this subsection, we therefore attempt to “calibrate” the values of $`\xi `$ and $`\kappa `$ mainly using established BHBs. The factor $`\xi `$ corrects the MCD formalism for the inner boundary condition of the accretion disk, in such a way that the disk temperature reaches maximum at $`R_{\mathrm{in}}/\xi `$ instead of $`R_{\mathrm{in}}`$. We may then examine our choice of $`\xi =0.41`$ (Kubota et al. 1998) in reference to the GRAD model mentioned in § 4.1 that properly takes into account this effect. We accordingly simulated a number of spectra using the GRAD model, by changing $`M`$, $`\dot{M}`$ and $`i`$, but fixing $`\kappa `$ at 1.7. The spectra were analyzed with the MCD formalism, employing $`\xi =0.41`$ and $`\kappa =1.7`$. Then, over the inclination range of $`0^{}`$ to $`75^{}`$ and over $`T_{\mathrm{in}}=1.22.0`$ keV, the X-ray mass $`M_{\mathrm{XR}}`$ from the MCD analysis agreed within $`\pm 25\%`$ with the BH mass $`M`$ employed as initial inputs to the GRAD model simulations. We can also utilize Cygnus X-1, the prime BHB, for this purpose. Applying the GRAD model to the ASCA data of Cygnus X-1 acquired in the 1997 May soft state, and assuming $`D=2.5`$ kpc and $`i=30^{}`$, Dotani et al. (1997) derived $`M_{\mathrm{XR}}=12_1^{+3}M_{}`$. If we apply our MCD formalism with $`\xi \kappa ^2=1.18`$ to the same dataset for Cygnus X-1, we obtain $`R_{\mathrm{in}}=90\pm 18`$ km (Table 1), and hence $`M_{\mathrm{XR}}=10\pm 2M_{}`$, which is close to the GRAD result by Dotani et al. (1997). Furthermore, our value of $`M_{\mathrm{XR}}`$ is consistent with the latest optical estimate of the BH mass in Cygnus X-1, $`M_{\mathrm{opt}}=10.1_{5.3}^{+4.6}M_{}`$ (Herreo et al. 1995). Thus, our overall knowledge of Cygnus X-1 supports the value of $`\xi \kappa ^21.2`$. We plot Cygnus X-1 in Figures THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES and THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES. We may further utilize the two Magellanic BHBs, LMC X-1 and LMC X-3, which have well established distances of 50–55 kpc; here we adopt 55 kpc. We hence analyzed the ASCA GIS data of LMC X-1 and LMC X-3, obtained on 1995 April 2 for 22 ks and 1995 April 14 for 23 ks, respectively. Their GIS spectra have both been fitted successfully by a model consisting of an MCD component and a power-law tail, whenall model parameters are left free to vary. The derived MCD parameters are given in Table 1, while the power-law index turned out to be $`2.36_{0.18}^{+0.14}`$ for LMC X-1 and $`2.52_{0.25}^{+0.22}`$ for LMC X-3. The photoelectric absorption is $`N_\mathrm{H}=(5.9\pm 0.3)\times 10^{21}`$ cm<sup>-2</sup> in LMC X-1 , and $`N_\mathrm{H}<2.9\times 10^{21}`$ cm<sup>-2</sup> in LMC X-3. The measured disk temperatures, given in Table 1, agree with the Ginga measurements (Ebisawa 1991; Ebisawa et al. 1993). Using $`i60^{}`$ for LMC X-1 (Cowley 1992; Cowley et al. 1995) and $`i=66^{}\pm 2^{}`$ for LMC X-3 (Kuiper et al. 1988), we obtain the disk radii as given in Table 1. In Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES, these radii are in good consistency with the optical mass estimates of $`M_{\mathrm{opt}}=410M_{}`$ for LMC X-1 (Cowley 1992; Cowley et al. 1995) and $`M_{\mathrm{opt}}=5.07.2M_{}`$ for LMC X-3 (Kuiper et al. 1988, $`D=55`$ kpc). Thus, the value of $`\xi \kappa ^2`$ we have employed is not particularly wrong, and may not be changed by more than a factor of 1.5 or so. This tolerance is not large enough to solve the $`M_\mathrm{E}`$ vs. $`M_\mathrm{X}`$ discrepancy for many of our sources. ### 5.5 Luminosity dependences Even though we have argued in § 5.4 for the justification of our estimates on $`R_{\mathrm{in}}`$, the disk structure may depend on $`L_{\mathrm{bol}}`$ or $`\dot{M}`$, particularly when $`L_{\mathrm{bol}}`$ approaches the Eddington limit. The electron scattering effect may get severer, which will increase $`\kappa `$, e.g., from 1.7 to 1.8–2.0 (Shimura & Takahara 1995). Furthermore, the true disk temperature may become higher than is described by the standard accretion disk model, because entropy of the accreting matter may start being “advected” into the BH as described by the slim disk model (Abramowicz et al. 1988). Recently, Watarai et al. (1999) have shown via calculation that significant amount of X-rays are radiated even from the region inside $`3R_\mathrm{s}`$ in a slim accretion disk accreting close to the Eddington limit. (We may also refer to Reynolds & Begelman (1997) who discuss the effect of accretion flow inside $`3R_\mathrm{s}`$.) While this mechanism can potentially explain what we have observed, its effects will vanish when the accretion rate becomes sufficiently low. Therefore, we are urged to search the available data for luminosity-dependent effects, such as apparent changes in $`R_{\mathrm{in}}`$. As mentioned in § 5.4.3, LMC X-3 was observed extensively with Ginga. In Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES, we therefore plot all the Ginga and ASCA data points for LMC X-3. Thus, the luminosity of LMC X-3 varied by a factor of $`7`$, reaching $`2/3`$ of the Eddington limit. However, the data points are aligned very well with a grid line for a constant mass as already reported previously (Ebisawa 1991; Ebisawa et al. 1993), and the implied mass is consistent with $`M_{\mathrm{opt}}`$. Therefore, we do not see any luminosity-dependent changes in $`R_{\mathrm{in}}`$. The bright BH transient GS 2000+25, discovered with the Ginga ASM (Tsunemi et al. 1989), is still more suited for the study of the luminosity dependence of accretion disks, because of its large luminosity swing. After the discovery, GS 2000+25 was observed with the Ginga LAC eight times spanning 240 days, covering almost the outburst peak (Ebisawa 1991). The obtained 2–30 keV spectra were all expressed well with an MCD component plus a hard power-law tail. In Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES, we plot seven out of the eight Ginga data points (discarding the last one), assuming a source distance of 3 kpc and $`i=65^{}`$ after Callanan et al. (1996). Thus, over a very wide range of $`\dot{M}`$ up to $`\eta 0.7`$, the data points generally distribute along a constant-mass grid line, as already noticed by Ebisawa (1991). The value of $`M_\mathrm{X}7M_{}`$ indicated by Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES agrees with $`M_{\mathrm{opt}}`$ measured by various authors (Filippenko, Matheson & Barth 1995; Beekman et al. 1996; Callanan et al. 1996), which cluster around $`58M_{}`$. Thus, the standard-disk interpretation is fully self consistent as to this objects, with little evidence of luminosity-dependent effects. The superluminal source GRO J1655-40 itself exhibited a large luminosity decline as well. In Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES, the ASCA and RXTE data points generally align along a constant-mass locus, even though the absolute value of $`M_{\mathrm{XR}}`$ falls considerably below $`M_{\mathrm{opt}}`$ as argued repeatedly. In short, the small value of $`M_{\mathrm{XR}}`$ (or the too high $`T_{\mathrm{in}}`$) of this jet source is not specific to its high-luminosity states, but appears intrinsic to it. We may further argue against luminosity-dependent effects by comparing GRO J1655$``$40 and LMC X-3. These two objects are very similar in terms of the BH mass, inclination, and $`L_{\mathrm{bol}}`$ at the time of the ASCA observation (Table 1). Nevertheless, their disk temperatures differ systematically by 40%, which makes their disk radii significantly different in Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES. We therefore conclude that luminosity-dependent changes in the accretion disk, if any, cannot fully explain the high disk temperature seen among a certain class of BHs. We regard it as an intrinsic property of such BHs. ### 5.6 The spinning black hole scenario We are finally urged to examine the remaining degree of freedom in our MCD scenario, i.e., the assumption of $`\alpha =1`$ in equation (7), which reflects the fact that the stable circular orbit can exist only outside $`3R_\mathrm{s}`$ in the case of a Schwarzschild (i.e., a non-rotating) BH. If instead the BH is rotating significantly (i.e., a Kerr BH), and the accretion disk is prograde to it, the last stable Keplerian orbit gets closer to the BH, down to $`R_{\mathrm{in}}=0.5R_\mathrm{s}`$ for a maximally rotating BH. As a result, we expect a Kerr BH to have a smaller value of $`\alpha `$, down to the extreme of $`\alpha =1/6`$, than a Schwarzschild BH of the same mass and the same accretion rate. Then, a Kerr BH can attain a higher value of $`T_{\mathrm{in}}`$, neglecting for the moment stronger relativistic effects. In Figure THE NATURE OF ULTRA-LUMINOUS COMPACT X-RAY SOURCES IN NEARBY SPIRAL GALAXIES, the dotted line shows the prediction for the maximally rotating BHs, obtained by substituting $`\alpha =1/6`$ into equation (7). Thus, except for IC 342 Source 1 for which the distance is rather uncertain, all the objects now fall on the physically reasonable region. Therefore we propose that the ULXs contain Kerr BHs. This idea was first proposed by Zhang, Cui & Chen (1997), to explain the high disk temperatures of GRS 1915+105 and GRO J1655$``$40, and subsequently applied by Mizuno et al. (1999) to the two ULXs in NGC 4565. Of course, investigation of the emission from rotating BHs must fully take into account relativistic effects. Such studies have been conduced by several authors, including, e.g., Asaoka (1987), Zhang, Cui & Chen (1997), and Beloborodov (1998). Asaoka (1987) showed that the spectrum from an optically-thick accretion disk around a Kerr BH is similar in shape to that around a Schwarzschild BH, except a higher temperature. Zhang, Cui & Chen (1997) calculated how the maximum color temperature of the accretion disk, $`T_{\mathrm{in}}`$, depends on the BH rotation. They show that, when the disk is prograde to a maximally rotating BH, $`T_{\mathrm{in}}`$ can be higher by a factor of $`3`$ than that around a Schwarzschild BH with the same BH mass and the same mass accretion rate. This is large enough to account for the high disk temperature of our sample objects. (This temperature enhancement factor becomes $`6^{0.75}=3.8`$ in terms of the MCD formalism which neglects the general relativistic effects.) There is another merit in appealing to the BH rotation, that the radiation efficiency, defined as the ratio $`L_{\mathrm{bol}}/\dot{M}c^2`$, gets higher as the BH angular momentum increases. The efficiency is 0.057 for a Schwarzschild BH, while it increases to $`0.4`$ for a maximally rotating BH fed with a prograde disk (see Zhang, Cui & Chen 1997). Therefore, in a maximally rotating BH, the same mass accretion rate can produces 7 times higher luminosity than in a non-rotating BH, although the same Eddington limit should still apply. How to make rather massive (several tens $`M_{}`$) Kerr BHs is an intriguing issue. There seems at least three different scenarios. One is the simplest scenario, to assume that a massive star with a rapid rotation collapses into a spinning BH. Another is to invoke a merger of two ordinary stellar-mass BHs, where the angular momentum is brought in from their orbital motion, and the BH mass is doubled. The other is to assume that a slowly spinning BH is spun up due to angular momentum supplied by the accreting matter, just as binary X-ray pulsars are spun up. Among the three scenarios, the last one is particularly attractive because it can potentially explain the large BH mass and the high angular momentum at the same time. Although Chen, Cui & Zhang (1997) argued that this mechanism is inefficient, the scenario may change significantly if the accretion is allowed to become considerably advective and super critical (Beloborodov 1998). More quantitative account of this scenario, however, is beyond the scope of the present paper. ## 6 Summary and Conclusion We have shown that the multi-color disk blackbody (MCD) emission model can successfully reproduce the 0.5–10 keV ASCA spectra of the seven ULXs in nearby spiral galaxies. These objects therefore form a homogeneous sample, and are considered to be mass-accreting BHs of which the X-ray emission originates from optically-thick accretion disks. Their bolometric luminosities, reaching $`10^{3940}`$ ergs s<sup>-1</sup> even allowing for possible distance errors, require the BHs to be relatively massive, up to several tens solar masses. The innermost disk temperatures of the ULXs have been found in the range $`T_{\mathrm{in}}=1.11.8`$ keV. These values significantly exceed those typically found in Galactic and Magellanic BHBs, and are inconsistent with the high BH mass indicated by the high ULX luminosities. Similarly high disk temperatures are observed from the two Galactic superluminal jet sources, GRO J1655-40 and GRS 1915+105. This problem cannot be explained away by changing the disk inclination, the source distance, or the value of $`\xi \kappa ^2`$, within tolerances. Although an optically-thick advective disk (a slim disk) may account qualitatively for these observations, we do not find evidence of luminosity-dependent deviations from the predictions of the standard disk scinario. We hence suggest that the BHs in these objects are spinning rapidly (i.e. Kerr BHs), so that the accretion disks can get closer to the BHs and get hotter. In short, a ULX may be a Kerr BH with a mass of several tens $`M_{}`$, in which an optically-thick accretion disk is radiating at a near-Eddington luminosity. This provides the first concrete working hypothesis with which the ULX phenomenon can be investigated. We would like to thank Drs. T. Hanawa, S. Mineshige, J. Fukue, and Mr. K. Watarai for helpful discussions. We also thank members of the ASCA team. The present work is supported in part by the Grant-in-Aid for Center-of-Excellence, No. 07CE2002, from Ministry of Education, Science, Sports and Culture of Japan.
warning/0001/hep-ph0001041.html
ar5iv
text
# Seesaw mechanism and the neutrino mass matrix ## 1 INTRODUCTION In this talk, I discuss how phenomenologically viable neutrino masses and mixings can be generated within the framework of the the seesaw mechanism of neutrino mass generation, constrained by the following set of assumptions: (i) Three generation $`SU(2)_L\times U(1)`$ model with the addition of three right-handed neutrino fields, which are singlets under $`SU(2)_L\times U(1)`$. No Higgs triplets are introduced and thus the effective mass matrix for the left-handed Majorana neutrinos is entirely generated by the seesaw mechanism, being given by $$m_L=m_DM_R^1m_D^T,$$ (1) where $`m_D`$ and $`M_R`$ denote the neutrino Dirac mass matrix and the Majorana mass matrix of right-handed neutrinos. (ii) The neutrino Dirac mass matrix $`m_D`$ has a hierarchical eigenvalue structure, analogous to the one for the up-type quarks. (iii) Charged lepton and neutrino Dirac mass matrices, $`m_l`$ and $`m_D`$, are “aligned” in the sense that in the absence of the right-handed mass $`M_R`$, the leptonic mixing would be small, as it is in the quark sector. In other words, we assume that the left-handed rotations that diagonalize $`m_l`$ and $`m_D`$ are the same or nearly the same. We therefore consider that the large lepton mixing results from the fact that neutrinos acquire their mass through the seesaw mechanism. (iv) The Dirac and Majorana neutrino mass matrices are unrelated. ## 2 GENERAL FRAMEWORK Under our assumptions, in the basis where the mass matrix of charged leptons has been diagonalized, the effective mass matrix of light neutrinos $`m_L`$ can be written as $$\left(\begin{array}{ccc}m_u^2M_{11}^1& m_um_cM_{12}^1& m_um_tM_{13}^1\\ m_um_cM_{12}^1& m_c^2M_{22}^1& m_cm_tM_{23}^1\\ m_um_tM_{13}^1& m_cm_tM_{23}^1& m_t^2M_{33}^1\end{array}\right)$$ (2) Here $`M_{ij}^1(M_R^{})_{ij}^1`$, where $`M_R^{}`$ is the mass matrix of the right handed neutrinos in the basis where the neutrino Dirac mass matrix $`m_D`$ is diagonal, and $`m_u`$, $`m_c`$ and $`m_t`$ are the eigenvalues of $`m_D`$. For our numerical estimates we take them to be equal to the masses of the corresponding up-type quarks, but for our general arguments their precise values are unimportant. The mass matrix (2) has to be compared with the phenomenologically allowed neutrino mass matrices. Consider first the direct neutrino mass hierarchy $`m_1,m_2m_3`$. Assuming that $`\theta _{23}45^{}`$ which is the best fit value of the Super-Kamiokande data , and taking into account that the CHOOZ experiment indicates that $`\theta _{13}1`$ , one can show that $`m_L`$ must have the approximate form $$m_L=m_0\left(\begin{array}{ccc}\kappa & \epsilon & \epsilon ^{}\\ \epsilon & 1+\delta \delta ^{}& 1\delta \\ \epsilon ^{}& 1\delta & 1+\delta +\delta ^{}\end{array}\right),$$ (3) where $`\kappa `$, $`\epsilon `$, $`\epsilon ^{}`$, $`\delta `$ and $`\delta ^{}`$ are small dimensionless parameters. Comparing (2) and (3), one concludes that the following relations should hold, in leading order: $$m_c^2M_{22}^1=m_t^2M_{33}^1=m_cm_tM_{23}^1.$$ (4) These relations seem to indicate that in order to obtain the form of Eq. (3), strong correlations are required between the eigenvalues of $`m_D`$ and the entries of $`(M_R^{})^1`$, in apparent contradiction with our assumption (iv). However, in fact there is no conflict. Obviously, the form of $`(M_R^{})^1`$ depends on the $`\nu _R`$ basis one chooses. We have defined $`M_R^{}`$ in the basis where $`m_D`$ is diagonal, i.e. have included into its definition the right-handed rotation arising from the diagonalization of $`m_D`$. Therefore $`(M_R^{})^1`$ contains information about the Dirac mass sector, and Eq. (4) is not necessarily in conflict with our assumption (iv). This assumption has to be formulated in terms of weak-basis invariants. What should be required is that the ratios of the eigenvalues of $`(M_R^{})^1`$ should not be related to the ratios of the eigenvalues of $`m_D`$. In order to see how the phenomenologically favoured form of $`m_L`$ can be achieved without contrived fine tuning between the parameters of the Dirac and Majorana sectors, let us first consider the two-dimensional sector of $`m_L`$ in the 2-3 subspace, which is responsible for a large $`\theta _{23}`$. We shall write the diagonalized Dirac mass matrix $`m_D^{diag}`$ as $$m_D^{diag}=m_tdiag(p^2q,p,\mathrm{\hspace{0.33em}1}),p=m_c/m_t10^2,$$ $$q=m_um_t/m_c^20.4.$$ (5) The 2-3 sector of $`(M_R^{})^1`$, in order to lead to the 2-3 structure of Eq. (3) (with all elements approximately equal to unity up to a common factor), should have the following form: $$M_R^1\left(\begin{array}{cc}1& p\\ p& p^2\end{array}\right).$$ (6) The eigenvalues of the matrix in Eq. (6) are 0 and $`1+p^2`$, and thus by choosing the pre-factor to be $`const/(1+p^2)`$ one arrives at the matrix $`M_R^1`$ of the desired form with $`p`$\- and $`q`$-independent eigenvalues. It turns out to be possible to find a $`3\times 3`$ matrix $`(M_R^{})^1`$ with $`p`$\- and $`q`$-independent eigenvalues, whose 2-3 sector generalizes (6) so as to obtain the realistic form of $`m_L`$ in (3) with $`\delta ,\delta ^{}0`$ : $$(M_R^{})^1=S_R^T(M_R^0)^1S_R$$ (7) with $$S_R=\left(\begin{array}{ccc}1& 0& 0\\ 0& c_\varphi & s_\varphi \\ 0& s_\varphi & c_\varphi \end{array}\right),$$ (8) where $`c_\varphi =\mathrm{cos}\varphi `$, $`s_\varphi =\mathrm{sin}\varphi `$, $`\varphi =\mathrm{arctan}p`$, and $$(M_R^0)^1=\frac{1}{2M}\left(\begin{array}{ccc}\gamma & \beta & \alpha \\ \beta & 1& 0\\ \alpha & 0& r\end{array}\right).$$ (9) The dimensionless parameters $`\alpha `$, $`\beta `$, $`\gamma `$ and $`r`$ in (9) do not depend on $`p`$ and $`q`$. Eqs. (7) - (9) yields the following mass matrix for the light neutrinos: $$m_L\frac{m_t^2}{2M}\frac{p^2}{1+p^2}\times $$ $$\left(\begin{array}{ccc}q_{}^{}{}_{}{}^{2}p^2\gamma & q^{}p(\beta \alpha p)& q^{}(\alpha +\beta p)\\ q^{}p(\beta \alpha p)& 1r/4p^2& 1r/4p^2\\ q^{}(\alpha +\beta p)& 1r/4p^2& 1+3r/4p^2\end{array}\right)$$ (10) Here $`q^{}q\sqrt{1+p^2}q`$. Thus, we have obtained $`m_L`$ of the desired form, while abiding by our assumptions. The particular case of the neutrino mass matrix of the form (10) with $`\beta =\gamma =r=0`$ (which allows only the vacuum oscillations solution of the solar neutrino problem) was obtained in . Comparison of Eqs. (3) and (10) allows one to express the parameters of the phenomenological mass matrix of light neutrinos $`\kappa `$, $`\epsilon `$, $`\epsilon ^{}`$ $`\delta `$ and $`\delta ^{}`$ in terms of the parameters of the mass matrix of the right handed neutrinos $`\alpha `$, $`\beta `$, $`\gamma `$ and $`r`$ and the parameters of the Dirac mass matrix $`m_t`$, $`p`$ and $`q`$. The largest eigenvalue of the matrix $`m_L`$ in (10) is $$m_3\frac{m_t^2}{M}\frac{p^2}{1+p^2}\frac{m_c^2}{M}.$$ (11) It scales as $`m_c^2`$ rather than as usually expected $`m_t^2`$. It has to be identified with $`\mathrm{\Delta }m_{atm}^2(26)\times 10^3`$ eV<sup>2</sup>, which gives $`M10^{10}10^{11}`$ GeV, i.e. an intermediate mass scale rather than the GUT scale. ## 3 INVERTED MASS HIERARCHY AND QUASIDEGENERACY Direct inspection of the neutrino mass textures that lead to the inverted mass hierarchy $`m_3m_1m_2`$ and the quasi-degenerate case with $`m_1m_2m_3`$ show that all of them except one do not satisfy our assumptions (i)-(iv) . In particular, for these textures it is impossible to have both traces ($`\mathrm{\Lambda }_1+\mathrm{\Lambda }_2+\mathrm{\Lambda }_3`$) and second invariants ($`\mathrm{\Lambda }_1\mathrm{\Lambda }_2+\mathrm{\Lambda }_1\mathrm{\Lambda }_3+\mathrm{\Lambda }_2\mathrm{\Lambda }_3`$) of the corresponding matrices $`(M_R^{})^1`$ to be $`p`$\- and $`q`$-independent (here $`\mathrm{\Lambda }_i`$ ($`i=1,2,3`$) are the eigenvalues of $`(M_R^{})^1`$). The remaining texture corresponds to the mass matrix of the heavy singlet neutrinos $`M_R^{}`$ with the following eigenvalues: two singlet neutrinos are almost degenerate with $`M_1M_210^8`$ GeV; the third mass eigenvalue turns out to be well above the Planck scale: $`M_310^{22}`$ GeV, clearly not a physical value. Thus, this case of the inverted mass hierarchy is ruled out as well. ## 4 DISCUSSION We have shown that the seesaw mechanism, supplemented by the set of assumptions listed in the Introduction, leads to phenomenologically viable mass matrices of light active neutrinos. The mixing angle $`\theta _{23}`$ responsible for the dominant channel of the atmospheric neutrino oscillations can be naturally large without any fine tuning. We have found that all three main neutrino oscillations solutions to the solar neutrino problem – small mixing angle MSW, large mixing angle MSW and vacuum oscillations – are possible within the constrained seesaw. The numerical examples of the requisite values of the parameters of the Majorana mass matrix $`M_R^{}`$ of singlet neutrinos are given in . Although the constrained seesaw mechanism allows to obtain a large mixing angle $`\theta _{23}`$ in a very natural way, it does not explain why $`\theta _{23}`$ is large: the largeness of this mixing angle is merely related to the choice of the inverse mass matrix of heavy singlet neutrinos, Eq. (7) - (9). However, once this choice has been made, the smallness of the mixing angle $`\theta _{13}`$ which determines the element $`U_{e3}`$ of the lepton mixing matrix can be readily understood. For the case of the normal mass hierarchy $`m_1,m_2m_3`$ the value of $`\theta _{13}`$ can be expressed in terms of the entries of the effective mass matrix $`m_L`$ in Eq. (3) as $`\mathrm{sin}\theta _{13}(\epsilon +\epsilon ^{})/2\sqrt{2}`$ . One then finds $`\mathrm{sin}\theta _{13}q(\alpha +2\beta p)/2\sqrt{2}q\alpha /2\sqrt{2}0.14\alpha `$, assuming $`\beta \alpha p^1100\alpha `$. Since all the solutions of the solar neutrino problem require $`|\alpha |<1`$ in order to have small enough $`\mathrm{\Delta }m_{}^2`$ , the smallness of $`\theta _{13}`$ follows. The seesaw mechanism we have studied naturally leads to the normal neutrino mass hierarchy while disfavouring the inverted mass hierarchy and quasi-degenerate neutrinos. For LMA and SMA solutions of the solar neutrino problem, the masses of the heavy singlet neutrinos are of the order $`10^{10}10^{11}`$ GeV. For the VO solution, the lightest of the singlet neutrinos has the mass of the same order of magnitude, whereas the masses of the other two are $`10^{12}10^{13}`$ GeV.
warning/0001/hep-th0001001.html
ar5iv
text
# Compactification, Geometry and Duality: 𝑁=2. ## 1 Introduction One of the most basic properties one may study about a class of string compactifications is its moduli space of vacua. If the class is suitably chosen one may find this a challenging subject which probes deeply into our understanding of string theory. In four dimensions it is the $`N=2`$ cases which provide the “Goldilocks” theories to study. As we will see, $`N=4`$ supersymmetry is too constraining and determines the moduli space exactly, leaving no room for interesting corrections from quantum effects. $`N=1`$ supersymmetry is highly unconstrained leaving the possibility that our supposed moduli acquire mass ruining the moduli space completely. $`N=2`$ however is just right — quantum effects are not potent enough to kill the moduli but they can affect the structure of the moduli space. (It is therefore a pity that the real world does not have $`N=2`$ supersymmetry — such a theory is necessarily non-chiral.) The subject of $`N=2`$ compactifications is enormous and we will present here only a rather biased set of highlights. These lectures will be sometimes closely-related to a set of lectures I gave at TASI 3 years ago . Having said that, the focus of these lectures differs from the former and the set of topics covered is not identical. I will however often refer to for details of certain subjects. Related to the problem of finding the moduli space of a class of theories is the following problem in string duality. Consider these four possibilities for obtaining an $`N=2`$ theory in four dimensions: 1. A type IIA string compactified on a Calabi–Yau threefold $`X`$. 2. A type IIB string compactified on a Calabi–Yau threefold $`Y`$. 3. An $`E_8\times E_8`$ heterotic string compactified on K3$`\times T^2`$. 4. A $`\mathrm{Spin}(32)/_2`$ heterotic string compactified on K3$`\times T^2`$. Can we find cases where the resulting 4 dimensional physics is identical for two or more of these possibilities and, if so, how do we match the moduli of these theories to each other? This is a story that began with some time ago but many details are poorly-understood to this day. One might suppose that knowing the moduli space of each theory listed above is a prerequisite for solving this problem but actually it is often useful, as we will see, to consider this duality problem at the same time as the moduli space problem. Note that there are other possibilities for producing $`N=2`$ theories in four dimensions such as the type I open string on K3$`\times T^2`$. We will stick with the four listed above in these lectures as they are quite sufficient for our purposes. As will be discussed shortly this problem breaks up into two pieces. One factor of the moduli space consists of the vector multiplet moduli space and the other factor consists of the hypermultiplet moduli space. By most criteria the moduli space of vector multiplets is well-understood today. This complex Kähler space can be modeled exactly in terms of the deformation space of a Calabi–Yau threefold. We will therefore be able to review this subject fairly extensively. In contrast the hypermultiplet moduli space remains a subject of research very much “in progress”. We will only be able to discuss in detail the classical boundaries of these moduli spaces. The interior of these spaces may offer considerable insight into string theory but we will only be able to cover some tantalizing hints of such possibilities. These lectures divide naturally into three sections. In section 2 we discuss generalities about moduli spaces of various numbers of supersymmetries in various numbers of dimensions. Although these lectures are intended to focus on the case of $`N=2`$ in four dimensions, there are highly relevant observations that can be made by considering other possibilities. Of particular note in this section is the rigid structure which emerges with more supersymmetry than the case in question. The heart of these lectures then consists of a discussion of the vector multiplet moduli space in section 3 and then the hypermultiplet moduli space in section 4. It is perhaps worth mentioning again that these lectures do not do justice to this vast subject and should be viewed as a biased account. Topic such as open strings, D-branes and M-theory have been neglected only because of the author’s groundless prejudices. The paragraphs starting with a “✄” are technical and can be skipped if the reader does not wish to be embroiled in subtleties. ## 2 General Structure ### 2.1 Holonomy We begin this section with a well-known derivation of key properties of moduli spaces based on $`R`$-symmetries and holonomy arguments. We should warn the more mathematically-inclined reader that we shall not endeavour to make completely watertight rigorous statements in the following. There may be a few pathological special cases which circumvent some of our (possibly implicit) assumptions. Suppose we are given a vector bundle, $`V`$, with a connection. We may define the “holonomy”, $`\mathrm{Hol}(V)`$, of this bundle as the group generated by parallel transport around loops in the base with respect to this connection. (A choice of basepoint is unimportant.) We may also define the restricted holonomy group $`\mathrm{Hol}^0(V)`$ to be generated by contractable loops. This notion can be very useful when applied to supersymmetric field theories as noted in . First let us consider the moduli space of a given class of theories. We will consider the moduli space as the base space of a bundle. Note that the moduli space of theories, $`\mathrm{M}`$, is equipped with a natural metric — that of the sigma-model. The tangent directions in the moduli space are given by the massless scalar fields with completely flat potentials. These massless fields may thus be given vacuum expectation values leading to a deformation of the theory. Let us denote these moduli fields $`\varphi ^i`$, $`i=1,\mathrm{},dim(\mathrm{M})`$. The low-energy effective action in the uncompactified space-time is then given by $$d^dx\sqrt{g}G_{ij}^\mu \varphi ^i_\mu \varphi ^j+\mathrm{}$$ (1) where $`G_{ij}`$ is our desired metric on $`\mathrm{M}`$. We therefore have a natural torsion-free connection on the tangent bundle of $`\mathrm{M}`$ given by the Levi-Civita connection with respect to this metric. Now consider the supersymmetry generators given by spinors $`Q_A`$, $`A=1,\mathrm{},N`$, where as usual $`N`$ denotes the number of supersymmetries (we suppress the spinor index). These objects are representations of $`\mathrm{Spin}(1,d1)`$ and are * Real if $`d=1,2,3mod8`$ * Complex if $`d=0,4mod8`$ * Quaternionic (or symplectic Majorana) if $`d=5,6,7mod8`$. The bundle of supersymmetry generators over $`\mathrm{M}`$ will also have a natural connection related to that on the tangent bundle. The key relation in supersymmetry is the equation $$\gamma _{\alpha \beta }^\mu \{\overline{Q}_{\overline{A}}^\alpha ,Q_B^\beta \}=\delta _{\overline{A}B}P^\mu ,$$ (2) where $`\gamma `$ are the usual gamma matrices, $`P`$ is translation and the bars in this equation are to be interpreted according to whether the spinors are real, complex or quaternionic.<sup>1</sup><sup>1</sup>1We ignore central charges which are irrelevant for this argument. Because parallel transport must preserve $`\delta _{\overline{A}B}`$ in (2) we see immediately that under holonomy $`Q_A`$ must transform as a fundamental representation of * $`\mathrm{SO}(N)`$ if $`d=1,2,3mod8`$ * $`\mathrm{U}(N)`$ if $`d=0,4mod8`$ * $`\mathrm{Sp}(N)`$ if $`d=5,6,7mod8`$, if the loop around which we transport is contractable. These groups are the “$`R`$-symmetries” of the supersymmetric field theories and give $`\mathrm{Hol}^0`$ of this bundle. We also note that in $`4M+2`$ dimensions, for integer $`M`$, the supersymmetries are chiral. This means that we consider left and right supersymmetries separately as we will illustrate in some examples below. The massless scalar fields live in supermultiplets. Within each supermultiplet the set of scalar fields will form a particular (possibly trivial) representation of the $`R`$-symmetry. We refer to for a detailed account of this. Occasionally the supermultiplet contains only one scalar component and this then transforms trivially under $`R`$. So long as this is not the case the holonomy of our tangent bundle is related to the $`R`$-symmetry. We may be more precise than this. As we go around a loop in $`\mathrm{M}`$ the scalars within every given supermultiplet will be mixed simultaneously by the $`R`$-symmetry. The supermultiplets themselves may also be mixed as a whole into each other by holonomy. This implies that, so long as the scalars transform nontrivially under $`R`$, the holonomy of the tangent bundle is factorized with the $`R`$-symmetry forming one factor. It is important to note however that we may not mix a scalar from one supermultiplet freely with any scalar from another supermultiplet in a way that violates this factorization. This is incompatible with the detailed supersymmetry transformation laws (as the reader might verify if they are unconvinced). Note in particular that the scalars within two different types of supermultiplets can never mix under holonomy. This is a useful observation given the following due to De Rham (see, for example, ) ###### Theorem 1 If a Riemannian manifold is complete, simply connected and if the holonomy of its tangent bundle with respect to the Levi-Civita connection is reducible, then this manifold is a product metrically. Thus if $`\mathrm{M}`$ is simply-connected we see that $`\mathrm{M}`$ factorizes exactly into parts labelled by the type of supermultiplet containing the massless scalars. If $`\mathrm{M}`$ is not simply-connected we may pass to the universal cover and use this theorem again. The general statement is therefore that the moduli space factorizes up to the quotient of a discrete group acting on the product. * Actually we should treat the word “complete” in the above theorem with a little more care. There are nasty points at finite distance in the moduli space where the manifold structure breaks down. These points also lead to a breakdown in the factorization of the moduli space. These extremal transitions will be studied more in section 4.2. We should only say that the moduli space factorizes locally away from such points. We may now analyze the structure of each factor of $`\mathrm{M}`$ from the Berger-Simons theorem (for an account of this we refer again to ) which states that the manifold must appear as a row in the following list:<sup>2</sup><sup>2</sup>2The notation $`\mathrm{Sp}(1).\mathrm{Sp}(n)`$ means $`\mathrm{Sp}(1)\times \mathrm{Sp}(n)`$ divided by the diagonal central $`_2`$. $$\begin{array}{cc}\mathrm{Hol}^0& dim(\mathrm{M})\\ & \\ \mathrm{SO}(n)& n\\ \mathrm{U}(n)& 2n\\ \mathrm{SU}(n)& 2n\\ \mathrm{Sp}(1).\mathrm{Sp}(n)& 4n\\ \mathrm{Sp}(n)& 4n\\ \mathrm{Spin}(7)& 8\\ G_2& 7\end{array}$$ (3) or be a “symmetric space” (which we will define shortly). Note that the following names are given to some of these holonomies: * Kähler if $`\mathrm{Hol}\mathrm{U}(n)`$, * Ricci-flat Kähler if $`\mathrm{Hol}\mathrm{SU}(n)`$, * Quaternionic Kähler if $`\mathrm{Hol}\mathrm{Sp}(1).\mathrm{Sp}(n)`$, * Hyperkähler if $`\mathrm{Hol}\mathrm{Sp}(n)`$. A symmetric space is a Riemannian manifold which admits a “parity” $`_2`$-symmetry about every point. This parity symmetry acts as $`1`$ in every direction on the tangent space. All symmetric spaces are of the form $`G/H`$ for groups $`G`$ and $`H`$, where the holonomy is given by $`H`$. They have been classified by E. Cartan and we list all the noncompact forms in table 1.<sup>3</sup><sup>3</sup>3We have been sloppy about the precise global form of the group $`H`$. As listed one often needs to quotient by a finite group to get the correct answer. For example in entry “E V”, $`\mathrm{SU}(8)`$ is not a subgroup of $`E_{7(7)}`$ — the correct form of $`H`$ should be $`\mathrm{SU}(8)/_2`$. The notation $`\mathrm{SO}_0(p,q)`$ refers to the part of the Lie group connected to the identity. The noncompact forms are the ones relevant to moduli spaces. A key point to note here is that the symmetric spaces are rigid — they have no deformations of the metric which would preserve the holonomy. The same is not true for the non-symmetric spaces listed in (3). Thus if the holonomy is of a type which forces a symmetric space as the only possibility we will refer to this as a rigid case. Let us consider a few examples. * $`N=(1,1)`$ in 6 dimensions (i.e., one left-moving supersymmetry and one right-moving supersymmetry). This implies that the $`R`$-symmetry is $`\mathrm{Sp}(1)\times \mathrm{Sp}(1)=\mathrm{SO}(4)`$ (up to irrelevant discrete groups). Analysis of the supermultiplets shows that matter supermultiplets have 4 scalars transforming as a 4 of $`\mathrm{SO}(4)`$. If we only had one such supermultiplet we could say nothing about the moduli space as a generic Riemannian manifold of 4 dimensions has holonomy $`\mathrm{SO}(4)`$. If we have a generic number, $`n`$, of supermultiplets and assuming the moduli space doesn’t factorize unnaturally then a quick look at the list above shows that the only possibility is the symmetric space. $`\mathrm{SO}_0(4,n)/(\mathrm{SO}(4)\times \mathrm{SO}(n))`$. Thus this case is rigid. The gravity supermultiplet has a single scalar giving an additional factor of $``$ to the moduli space. * $`N=(2,0)`$ in 6 dimensions. This implies that the $`R`$-symmetry is $`\mathrm{Sp}(2)=\mathrm{SO}(5)`$ (up to irrelevant discrete groups). Analysis of the supermultiplets shows that matter supermultiplets have 5 scalars transforming as a 5 of $`\mathrm{SO}(5)`$. If we have a generic number, $`n`$, of supermultiplets and assuming the moduli space doesn’t factorize unnaturally then the list above shows that the only possibility is the symmetric space $`\mathrm{SO}_0(5,n)/(\mathrm{SO}(5)\times \mathrm{SO}(n))`$. Thus this case is rigid. There are no further moduli. * $`N=4`$ in 4 dimensions. This implies that the $`R`$-symmetry is $`\mathrm{U}(4)=\mathrm{SO}(6)\times \mathrm{U}(1)`$ (up to irrelevant discrete groups). Analysis of the supermultiplets shows that matter supermultiplets have 6 scalars transforming as a 6 of $`\mathrm{SO}(6)`$. If we have a generic number, $`n`$, of supermultiplets and assuming the moduli space doesn’t factorize unnaturally then the only possibility for this factor is the symmetric space $`\mathrm{SO}_0(6,n)/(\mathrm{SO}(6)\times \mathrm{SO}(n))`$. Thus this case is rigid. The gravity multiplet contains a complex scalar transforming under the $`\mathrm{U}(1)`$ factor of the holonomy. By holonomy arguments this contributes a complex Kähler factor to the moduli space. Closer analysis of this supergravity shows that this factor is actually $`\mathrm{SL}(2,)/\mathrm{U}(1)`$. This last example demonstrates an important point. Analysis of the $`R`$-symmetry may be sufficient to imply that we have a rigid moduli space but sometimes the moduli space is rigid even when the holonomy may imply otherwise. A more detailed analysis of the supergravity action is required in some cases to show that we indeed have a symmetric space. The rule of thumb is as follows: If we have maximal (32 supercharges, e.g. $`N=8`$ in four dimensions) or half-maximal (16 supercharges, e.g. $`N=4`$ in four dimensions) supersymmetry then, and usually only then, is the moduli space rigid. Note that there are a few strange examples such as where the moduli space is rigid even when there are fewer than 16 supercharges. ### 2.2 U-Duality In this section we will focus on global properties of the rigid moduli spaces. The analysis of the moduli spaces so far is not quite complete. The problem is that the moduli space need not be a manifold. There may be singular points corresponding to the theories with special properties. In the rigid case however the fact that the moduli space is symmetric wherever it is not singular is a very powerful constraint. Let us suppose first that we have an orbifold point. That is a region in the moduli space which looks locally like a manifold divided by a discrete group fixing some point $`x`$. Away from the fixed point set the moduli space is symmetric and thus “homogeneous”. That is, there exist a transitive set of translation symmetries. Assuming geodesic completeness of the moduli space, these translations may be used to extend the local orbifold property to a global one. That is, the moduli space is globally of the form of a manifold divided by a discrete group . This homogeneous structure of the moduli space may also be used to rule out other possibilities of singularities which occur at finite distance. Consider beginning at a smooth point in moduli space and approaching a singularity. The homogeneous structure implies that nothing about the local structure of the moduli space may change as you approach the singularity — everything happens suddenly as you hit the singularity. This rules out every other type of “reasonable” singularity that one may try to put in the moduli space. To be completely rigorous would require us to make precise technical definitions about the allowed geometry of the moduli space. Instead we shall just assert here that any type of singularity at finite distance that one might think of (such as a conifold) would ruin the homogeneous nature of the moduli space and so is not allowed in the rigid case. We therefore arrive at the conclusion that the only allowed global form of a rigid moduli space is of a symmetric space divided by a discrete group. This implies that any analysis of the moduli space of string theories in the case of maximal or half-maximal supersymmetry comes down to question of this discrete group. This group is precisely the group known as S-duality, T-duality or U-duality depending on the context. Many examples of such dualities were discussed in and we refer the reader there for details as well as references. For example the general rule is that a space locally of the form $`\mathrm{SO}_0(p,q)/(\mathrm{SO}(p)\times \mathrm{SO}(q))`$ becomes $$\mathrm{O}(\mathrm{{\rm Y}}_{p,q})\backslash \mathrm{O}(p,q)/(\mathrm{O}(p)\times \mathrm{O}(q)),$$ (4) where $`\mathrm{{\rm Y}}_{p,q}`$ is some lattice (often even and unimodular) of signature $`(p,q)`$ and $`\mathrm{O}(\mathrm{{\rm Y}}_{p,q})`$ is its discrete group of isometries. Indeed the only interesting question one may ask about the moduli space in the rigid case is what exactly this discrete group is! Any quantum corrections to the local structure are not allowed due to rigidity. It is not therefore surprising that S, T and U-duality are so ubiquitous when studying theories with a good deal of supersymmetry. As we will we see however, the picture becomes quite different when the supersymmetry is less that half-maximal. One final word of warning here. We have not been clear about what we mean by a “class” of string theories. If we determine the moduli space of some kind of string compactified on some kind of space, up to topology, then our moduli space may have numerous disconnected components. In this case the above results apply to each component separately. This reducibility often happens when we have half-maximal supersymmetry. ### 2.3 Eight supercharges Now we turn our attention to theories with quarter-maximal supersymmetry, or a total of 8 supercharges. Here we will also specify how one might obtain such a theory from string theory. If we compactify a ten-dimensional supersymmetric theory on $`^{1,d1}\times M`$, where $`M`$ is some compact manifold, then holonomy arguments may be again used to determine the number of unbroken supersymmetries in $`^{1,d1}`$. This time it is the holonomy of the compact space $`M`$ rather than the moduli space which we analyze. The basic idea is roughly that a symmetry in ten dimensions will be broken by the holonomy of (a suitable bundle on) $`M`$ to the centralizer of this holonomy group. That is, a symmetry in uncompactified space is broken if it can be transformed by parallel transport around a loop in the internal compactified dimensions.<sup>4</sup><sup>4</sup>4While this seems a very reasonable statement it is probably not rigorous. Breaking the gauge group of the heterotic string in this way does not always lead to the correct global form. We begin with $`N=(1,0)`$ in six dimensions. We may obtain this by compactifying a heterotic string theory on a four-dimensional manifold with holonomy $`\mathrm{SU}(2)`$. The only such manifold is a K3 surface. We refer to for an explanation of these points. The $`R`$-symmetry in this case is $`\mathrm{Sp}(1)`$. An analysis of supermultiplets show that scalars may occur in either of two types: 1. The Hypermultiplet contains 4 scalars which we may view as a quaternion. The holonomy $`\mathrm{Sp}(1)`$ may then be viewed as multiplication on the left by another quaternion of unit norm. 2. The Tensor multiplet contains a single scalar. Thus holonomy tells us nothing interesting. Note that the vector supermultiplet contains no scalars. The moduli space of such theories will locally factorize into a moduli space of hypermultiplets, which will be quaternionic Kähler and a moduli space of tensor multiplets. Now let us consider $`N=2`$ theories in four dimensions. We may obtain this by compactifying a heterotic string on a six-dimensional manifold with holonomy $`\mathrm{SU}(2)`$. The only such manifold is a product of a K3 surface and a 2-torus (or a finite quotient thereof). Alternatively we may compactify a type IIA or IIB superstring (which has twice as much supersymmetry than the heterotic string) on a manifold of holonomy $`\mathrm{SU}(3)`$. As usual, we will refer to such a Ricci-flat Kähler manifold as a “Calabi–Yau threefold”. Again we refer to for extensive details and references on these points. The $`R`$-symmetry in this case is $`\mathrm{U}(2)=\mathrm{Sp}(1)\times \mathrm{U}(1)`$. An analysis of supermultiplets show that scalars may occur in either of two types: 1. The Hypermultiplet contains 4 scalars which we may view as a quaternion. The holonomy $`\mathrm{Sp}(1)`$ may then be viewed as multiplication on the left by another quaternion of unit norm. 2. The Vector multiplet contains 2 scalars transforming as a complex scalar under the $`\mathrm{U}(1)`$ factor of the holonomy. Because of this the moduli space of such theories locally factorizes into a moduli space of hypermultiplets, which is quaternionic Kähler and we denote $`\mathrm{M}_H`$, and a moduli space of vector multiplets which is Kähler and we denote $`\mathrm{M}_V`$. Although we will see below that $`\mathrm{M}_V`$ is not any old complex Kähler manifold and $`\mathrm{M}_H`$ is not any old quaternionic Kähler manifold, it is true that they are not completely determined by supersymmetry and consequently have deformations. Thus in contrast to the rigid cases with a lot of supersymmetry, $`N=2`$ theories in 4 dimensions (and $`N=(1,0)`$ theories in six dimensions, etc.) can have interesting quantum corrections which warp the moduli space away from that which would be expected classically. Further analysis of $`\mathrm{M}_H`$ by Bagger and Witten yielded a property which is worth noting. They showed that the scalar curvature of $`\mathrm{M}_H`$ is negative and proportional to the gravitational coupling constant. Thus $`\mathrm{M}_H`$ is not hyperkähler unless the gravitational coupling is taken to zero. What is particularly nice about $`N=2`$ theories is that their moduli cannot gain mass through quantum effects. This is to be contrasted with the $`N=1`$ case in four dimensions where the moduli can become massive. This is discussed in M. Dine’s lectures in this volume. ### 2.4 Type II compactification Let us now turn to the coarse structure of the moduli space of type IIA and type IIB compactifications on Calabi–Yau threefolds. Begin with a type IIA string compactified on a Calabi–Yau threefold $`X`$. To leading order we demand that the metric be Ricci-flat. Actually this statement is not exact and receives quantum corrections. At higher loop in the non-linear $`\sigma `$-model we discuss below, the metric is warped away from the Ricci-flat solution and when one takes nonperturbative effects into account it is unlikely that one can faithfully represent $`X`$ in terms of a Riemannian metric at all. We will see this breakdown of Riemannian geometry later in section 3.3.6. What is true however is that if the Calabi–Yau is large then the Ricci-flat metric is a good approximation. Thus we may at least get the dimension of the moduli space correct using this metric. Thanks to Yau’s proof of the Calabi conjecture we do not have to undertake the unpleasant (and as yet unsolved) problem of explicitly constructing the Ricci-flat metric. We may instead assert its unique existence given a complex structure on $`X`$ together with the cohomology class of its Kähler form, $`[J]H^2(X,)`$. Deformation of the complex structure of $`X`$ yields $`h^{2,1}(X)`$ complex moduli whereas the Kähler form degree of freedom yields $`h^{1,1}(X)`$ real degrees of freedom. (We refer to section 2 of for a discussion of the classical geometry we use here.) The deformation of the Ricci-flat metric on $`X`$ thus produces $`h^{1,1}(X)+2h^{2,1}(X)`$ real moduli. There is also the ten-dimensional dilaton, $`\mathrm{\Phi }`$, which controls the string coupling constant. This contributes one real modulus. All the remaining moduli arise from objects which naïvely appear as $`p`$-forms in the ten-dimensional type II theory. The basic idea is that both type II strings (and indeed all closed string theories) have a “$`B`$-field” 2-form arising in the NS-NS spectrum while the type IIA string also contains a 1-form and a 3-form from the R-R sector and the type IIB string contains a 0-form, a 2-form and a self-dual 4-form from the R-R sector. We refer to for this basic property of string theory. This description of these ten-dimensional fields in terms of de Rham cohomology is rather vague and unfortunately does not really tell us the full truth about these objects and the resulting degrees of freedom they yield as moduli. The aspects which are poorly described concern both what happens when $`X`$ is singular and the discrete degrees of freedom (arising from torsion in the cohomology for example). We cannot pretend to understand the basic nature of string theory until we have a better description of the geometry of these fields. At present there are two leading contenders namely “gerbes” and “K-theory”. Unfortunately at the time of writing, neither of these theories together with its application to string theory is completely understood although the subject is maturing rapidly. The idea of a gerbe is best understood by first considering the R-R 1-forms of the type IIA string. These 1-forms are believed to describe a $`\mathrm{U}(1)`$ gauge theory in the ten-dimensional spacetime given by the type IIA string theory. Thus this R-R 1-form is actually a connection on a $`\mathrm{U}(1)`$-bundle and is the vector potential of some ten-dimensional “photon”. Consider now what happens if we compactify this $`\mathrm{U}(1)`$ gauge theory on a manifold $`X`$. This requires a choice of $`\mathrm{U}(1)`$-bundle $`VX`$ satisfying the Yang-Mills equations of motion. Such a bundle may have a first Chern class $`c_1(V)`$ corresponding to “magnetic monopoles” in $`X`$. If we demand that there are no such monopoles then our bundle must be flat. A flat bundle is over $`X`$ is described purely by the monodromy of the bundle around the various non-contractable loops in $`X`$. That is, by a homomorphism from $`\pi _1(X)`$ to $`\mathrm{U}(1)`$ — an element of $`\mathrm{Hom}(\pi _1(X),\mathrm{U}(1))`$. This is equal to $`\mathrm{Hom}(H_1(X),\mathrm{U}(1))`$ as $`\mathrm{U}(1)`$ is an abelian group. Using the universal coefficients theorem this in turn is equal to $`H^1(X,\mathrm{U}(1))`$. We arrive at the conclusion that the moduli space of flat $`\mathrm{U}(1)`$ bundles over $`X`$ is given by $`H^1(X,\mathrm{U}(1))`$. This then would be the contribution to the moduli space of the R-R 1-form of the type IIA string. The idea of gerbes is to extend the notion of a $`\mathrm{U}(1)`$-bundle with a connection to an object whose connection is a form of degree greater than one. Thus the $`B`$-field of string theory is treated as some connection on a gerbe where the string itself carries unit electric charge with respect to this generalized gauge theory. The theory of gerbes is described clearly in terms of Čech cohomology by Hitchin in . (See also , for example, for further discussion.) The basic property which we require is ###### Proposition 1 The moduli space of flat gerbes over $`X`$ whose connection corresponds to a $`p`$-form is given by $`H^p(X,\mathrm{U}(1))`$. Thus assuming no solitons in the background corresponding to a gerbe curvature (such as an “$`H`$-monopole”) this yields the desired moduli space. The exact sequence $$0\mathrm{U}(1)0$$ (5) yields the exact sequence $$H^p(X,)H^p(X,)H^p(X,\mathrm{U}(1))H^{p+1}(X,)H^{p+1}(X,).$$ (6) Thus if the cohomology of $`X`$ is torsion-free, we have $$H^p(X,\mathrm{U}(1))\frac{H^p(X,)}{H^p(X,)},$$ (7) which is a torus whose dimension is given by the Betti number $`b_p`$. Torsion in $`H^{p+1}(X,)`$ will extend this moduli space although it will not change the dimension. It is worth mentioning that discrete degrees of freedom associated to torsion are very poorly understood at present even though they will appear whenever a type IIA string is compactified on a non-simply-connected Calabi–Yau threefold. In the work of (see also for a fuller description of the degrees of freedom) these discrete choices were not treated as choices at all and were fixed by a process known as “black hole level matching”. Clearly more work needs to be done to understand this better. Alternatively one associates the R-R $`p`$-forms to the associated electrically-charged D-branes of dimension $`p1`$. The R-R field then measures the phase one associates to a D-brane instanton. This gives a nice physical picture of the meaning of the R-R moduli. It is believed however following the work of Minasian and Moore and Witten that the charges of D-branes are classified by K-theory and not cohomology (see also the lectures by J. Schwarz ). One might therefore suppose that the R-R moduli spaces may be given by something more in the spirit of K-theory like, for example, $`K^1(X,\mathrm{U}(1))`$ for the type IIA string and $`K^0(X,\mathrm{U}(1))`$ for the type IIB string.<sup>5</sup><sup>5</sup>5K-theory may be regarded as a generalized cohomology theory based on the Eilenberg-Steenrod axioms for cohomology. To define $`K^p(X,\mathrm{U}(1))`$ we may assert that $`K^{\mathrm{even}}`$ for a point is $`\mathrm{U}(1)`$ whereas $`K^{\mathrm{odd}}`$ for a point is 0 and all cohomology axioms are satisfied. The Chern character gives an isomorphism over the rational numbers between $`K^0`$ and $`H^{\text{even}}`$ and between $`K^1`$ and $`H^{\text{odd}}`$. This means that as far as simple dimension counting is concerned the moduli space of R-R fields is the same whether we use the gerbe picture or whether we use the K-theory picture. These pictures may not be equivalent globally over the entire moduli space however. Again more work is needed here. Either way, for the type II string compactified on a Calabi–Yau threefold we have the dimension of the moduli space given by certain Betti numbers. The $`B`$-field gives us $`b_2(X)=h^{1,1}(X)`$ real degrees of freedom. These can be paired up with the Kähler form degrees of freedom to produce $`h^{1,1}(X)`$ complex degrees of freedom. This complexification of the Kähler form is seen clearly from mirror symmetry as we will see in section 3.2.3. The uncompactified components of the $`B`$-field give an antisymmetric tensor field in four dimensions. Such a field may be dualized in the usual way to produce a real scalar. This is usually called the “axion” and is paired up with the dilaton to form a complex degree of freedom. Analysis of the R-R fields is as follows. For the type IIA string on $`X`$ we have $`b_1(X)+b_3(X)`$ real moduli. A manifold with precisely $`\mathrm{SU}(3)`$ holonomy has $`b_1(X)=0`$. (One way of seeing this is that a nonzero number would imply a continuous isometry leading to a torus factor.) We also have $`h^{3,0}=1`$ from the holomorphic 3-form which is nonzero and unique up to isometry. Thus in total we have $`2+2h^{2,1}`$ degrees of freedom from the R-R sector. All told we have produced $`2h^{1,1}+4(h^{2,1}+1)`$ real moduli. Since we expect our moduli space to factorize (up to a discrete quotienting) as $`\mathrm{M}_H\times \mathrm{M}_V`$, we need to label these moduli as to whether they form scalar fields in hypermultiplets or vector multiplets. A careful analysis of this was performed in but we may obtain the same result by a simple crude argument as follows. Clearly the type of field determines which kind of supermultiplet it lives in. For example, all the R-R fields must be in hypermultiplets or they must all be in vector multiplets. We also do not expect the labelling to depend on the specific values of $`h^{1,1}(X)`$ and $`h^{2,1}(X)`$. These facts together with the fact that the dimension of $`\mathrm{M}_H`$ is a multiple of four immediately tells is that ###### Proposition 2 For the type IIA string compactified on a Calabi–Yau threefold $`X`$ we have * $`\mathrm{M}_V`$ is spanned by the deformation of the complexified Kähler form and has complex dimension $`h^{1,1}(X)`$. * $`\mathrm{M}_H`$ is spanned by the deformations of complex structure of $`X`$, the dilaton-axion, and the R-R fields. It has quaternionic dimension $`h^{2,1}(X)+1`$. A similar analysis for the type IIB string differs only in the fact that the R-R fields consist of even forms rather than odd forms. This results in ###### Proposition 3 For the type IIB string compactified on a Calabi–Yau threefold $`Y`$ we have * $`\mathrm{M}_V`$ is spanned by the deformation of the complex structure of $`Y`$ and has complex dimension $`h^{2,1}(Y)`$. * $`\mathrm{M}_H`$ is spanned by the deformations of the complexified Kähler form, the dilaton-axion, and the R-R fields. It has quaternionic dimension $`h^{1,1}(Y)+1`$. We emphasize that these results are subject to quantum corrections. That is we may find the dimensions of these moduli spaces and the forms of these moduli spaces around some limit point using the above results, but the precise geometry of the moduli space may vary. We will discuss this in detail shortly. ### 2.5 Heterotic compactification Now we deal with the heterotic string compactified on a product of a K3 surface, which we denote $`S_H`$, and a 2-torus (or “elliptic curve”), which we denote $`E_H`$. (We may also take a quotient of this product by a finite group preserving the $`\mathrm{SU}(2)`$ holonomy. This makes a little difference to the analysis below but we ignore this possibility for the sake of exposition.) Again, to leading order, one of the things we are required to specify is a Ricci-flat metric on $`S_H\times E_H`$. In the case of $`E_H`$ this is easy as a Ricci-flat metric is a flat metric. We simply give one complex parameter specifying the complex structure of $`E_H`$ and a real number specifying the area of $`E_H`$. The moduli space of Ricci-flat metrics on $`S_H`$ is well-understood but a full explanation is rather lengthy. It is described in detail in . One of the most important points is that, unlike the threefold case, it does not factorize into a product of deformations of the complex structure and deformations of the Kähler form. This can be traced to the fact that a K3 surface has a hyperkähler structure which allows for a choice (parametrized by an $`S^2`$) of complex structures for a fixed Ricci-flat metric. Indeed, this choice allows for a deformation of complex structure to be reinterpreted as a deformation of the Kähler form. The result which we quote here is as follows. Let $`\mathrm{\Gamma }_{3,19}`$ be an even self-dual lattice of signature $`(3,19)`$ representing $`H^2(S_H,)`$ together with its cup product. The moduli space of Ricci-flat metrics on a K3 surface is then $$_+\times \mathrm{O}(\mathrm{\Gamma }_{3,19})\backslash \mathrm{O}(3,19)/(\mathrm{O}(3)\times \mathrm{O}(19)).$$ (8) where the $`_+`$ factor represents the total volume of $`S_H`$. As in the type II strings, the heterotic string has a dilaton which is complexified by adding the axion originating in the uncompactified parts of the $`B`$-field. Next we come to one of the awkward and interesting parts of the heterotic string — the “vector bundle”. Naïvely stated we take a smooth principal $`\mathrm{G}_0`$-bundle, $`V`$, on $`S_H\times E_H`$. The group $`\mathrm{G}_0`$ should be $`(E_8\times E_8)_2`$ or $`\mathrm{Spin}(32)/_2`$ according to which heterotic string we use.<sup>6</sup><sup>6</sup>6We explain this mouthful in the former case at the end of section 2.5.1. This bundle is used to “compactify” the gauge degrees of freedom of the ten-dimensional heterotic string. The vector bundle $`V`$ is equipped with a connection and this must satisfy certain conditions for the equations of motion of the string theory to be satisfied. Such a connection should be considered analogous to the Levi-Civita connection on the tangent bundle derived from the metric. Indeed, a simply ansatz frequently used is to embed the holonomy of the tangent bundle into $`\mathrm{G}_0`$ and obtain an effective choice for $`V`$. This process is often referred to as “embedding the spin connection in the gauge group” and was used in the earliest models of the heterotic string . The equations of motion imply a certain topological constraint on $`V`$. This topological condition can also be interpreted as that required for the cancellation of gravitational and Yang-Mills anomalies. We explain this shortly. In abstract terms, the homotopy class of our $`\mathrm{G}_0`$-bundle determines a characteristic class in $`H^4(S_H\times E_H,\pi _3(\mathrm{G}_0))`$ which may be thought of as the generalization of the second Chern class or the first Pontryagin class of $`V`$. This must be equal to the second Chern class of the tangent bundle of the base space $`S_H\times E_H`$. To leading order, the condition that $`V`$ must satisfy is simply that it obeys the Yang-Mills equations of motion. Fortunately the moduli space of solutions to these equations over a compact Kähler manifold is a well-studied problem in the mathematics literature. The trick is to complexify the problem giving a holomorphic bundle whose structure group lies in the complexification of $`\mathrm{G}_0`$. In most of what follows we will assume this complexification process implicitly and still refer to $`V`$ as a $`\mathrm{G}_0`$-bundle. The situation is easiest to explain in the case that $`V`$ is a $`\mathrm{U}(n)`$-vector bundle. In this case the complexification is a generic holomorphic vector bundle whose structure group is $`\mathrm{GL}(n,)`$. The Hermitian-Yang-Mills equations of interest then impose that the connection satisfies $`F_{\overline{ı}\overline{ȷ}}=F_{ij}=0`$ and $`g_{i\overline{ȷ}}F^{i\overline{ȷ}}=0`$, where $`g_{i\overline{ȷ}}`$ is the Kähler metric on the base manifold. We may integrate the equation $`g_{i\overline{ȷ}}F^{i\overline{ȷ}}=0`$ to obtain the necessary condition on the “degree” of $`V`$: $$\mathrm{deg}(V)=c_1(V)(J)=0,$$ (9) where $`J`$ is the Kähler form. We also need to explain what is meant by a “stable” vector bundle. To a given bundle $`E`$ we associate its “slope” $$\mu (E)=\frac{\mathrm{deg}(E)}{\mathrm{rank}(E)}.$$ (10) A bundle $`E`$ is said to be stable if every coherent subsheaf<sup>7</sup><sup>7</sup>7We could almost say “subbundle” here. $`\mathrm{F}`$ of lower rank satisfies $`\mu (\mathrm{F})<\mu (E)`$. A “semistable” bundle is allowed to satisfy $`\mu (\mathrm{F})\mu (E)`$. We then have the following theorem due to Donaldson, Uhlenbeck and Yau <sup>8</sup><sup>8</sup>8The original form of this theorem does not restrict attention to curvatures satisfying $`g_{i\overline{ȷ}}F^{i\overline{ȷ}}=0`$. Instead the case of constant $`g_{i\overline{ȷ}}F^{i\overline{ȷ}}`$ is considered. This is often called “Hermitian-Einstein” and is analogous to the case of an Einstein metric as opposed to a Ricci-flat metric. ###### Theorem 2 A bundle is stable and satisfies (9) if and only if it admits an irreducible Hermitian-Yang-Mills connection. This connection is unique. This reduces the difficult problem of finding the moduli space of bundles in terms of solution sets of differential equations to a more algebraic problem of finding the moduli space of holomorphic vector bundles. This is exactly analogous to replacing the problem of finding the moduli space of Ricci-flat metrics for a Calabi–Yau manifold to that of finding the moduli space of complex structures. Note that the theorem above imposes that the connection be irreducible. In many cases this is a little strong a we need to consider semistable bundles. This is discussed in . Continuing the analogy of solving the Yang-Mills equations for $`V`$ to the finding of a Ricci-flat metric we might suppose that looking for higher-order corrections to the equations of motion may require corrections to be made to the connection. These corrections will affect our moduli space problem. In addition we should expect that worldsheet instantons might ruin the very interpretation of these degrees of freedom of the heterotic string as coming from a vector bundle. The act of replacing the differential geometry problem of finding vector bundles satisfying the Yang-Mills equations by the algebraic geometry question of looking at the moduli space of stable holomorphic bundles might actually be seen as moving a step closer to the truth in string theory. As we move around the moduli space we will often encounter degenerations of the bundle data which can be interpreted easily in the algebraic picture by using the language of “sheaves”. See for example for more on this. Anyway, to return to our problem, we require a (semi)stable holomorphic bundle over the product of a K3 surface and an elliptic curve. The first simplification is to assume that this bundle factorizes nicely. That is, we have two bundles $$\begin{array}{cc}\hfill V_S& S_H\hfill \\ \hfill V_E& E_H.\hfill \end{array}$$ (11) Let the structure group of $`V_S`$ be $`\mathrm{G}_S`$ and let the structure group of $`V_E`$ be $`\mathrm{G}_E`$. This is then a special case of a $`\mathrm{G}_0`$-bundle over $`S_H\times E_H`$ if $`\mathrm{G}_0\mathrm{G}_S\times \mathrm{G}_E`$. Our problem nicely factorizes into finding the moduli space of $`V_S`$ and the moduli space of $`V_E`$. Finally we come to the other interesting part of the heterotic string — the $`B`$-field. In the case of the heterotic string a deep understanding of this object is even more troublesome than the $`B`$-field of the type II string. This was analyzed recently by Witten . Let us assume the heterotic string is compactified on a generic Calabi–Yau space $`Z`$. We can make a simple statement — the number of real degrees of freedom of the $`B`$-field is given by $`dimH^2(Z)`$ as it was for the type II strings. Beyond this simple dimension counting we have to work harder. The general idea is that anomaly cancellation in the heterotic string requires an equation in differential forms as follows. $$H=dB+\frac{\alpha ^{}}{4\pi }(\omega _Y\omega _T),$$ (12) where $`H`$ is the physically significant, and thus gauge-invariant, field strength associated to the heterotic string. The terms $`\omega _Y`$ and $`\omega _L`$ are Chern–Simons 3-forms associated to the connections of the Yang–Mills gauge bundle and the tangent bundle respectively. We refer to for a general review of these facts. Note that the exterior derivative of this formula gives $$dH=\frac{\alpha ^{}}{4\pi }\left(\mathrm{tr}RR\mathrm{tr}FF\right),$$ (13) where $`R`$ and $`F`$ are the curvatures of the tangent bundle and $`V`$ respectively. Taking cohomology classes this gives the topological constraint on $`V`$ discussed above.<sup>9</sup><sup>9</sup>9This argument using De Rham cohomology misses the torsion part. The fact that $`\omega _Y`$ and $`\omega _L`$ are not gauge invariant objects implies that $`B`$ will have some nontrivial transformation properties. An effect of $`B`$, as in the type II string, is to weight instantons as will be explained briefly in section 3.2.3. Namely, if a 2-sphere $`S`$ in the target space represents a worldsheet instanton then the action is weighted by a factor given by $$c=\mathrm{exp}\left(2\pi i_SB\right).$$ (14) In the simpler case of the type IIA string this phase is determined by the homology class of $`S`$. That is, $`B\mathrm{Hom}(H_2(Z),\mathrm{U}(1))H^2(Z,\mathrm{U}(1))`$. Witten noted the following awkward property of this phase when dealing with the heterotic string. Suppose we have a family of rational curves in the target space. For simplicity we assume the space contains $`^1\times C`$ for some complex curve $`C`$. Fix a particular $`^1`$ in this family. Let $`c_0`$ be the phase associated to this curve given by (14). Now move this curve in a contractable loop $`\gamma `$ within $`C`$. Let $`WC`$ be a disc in $`C`$ with boundary $`\gamma `$. When we return back to our original $`^1`$ one finds the phase induced by the $`B`$-field equal to $$c_1=\mathrm{exp}\left(2\pi i_{W\times ^1}𝑑H\right)c_0,$$ (15) where $`dH`$ is given by (13). Thus unless $`dH=0`$ the contribution of the $`B`$-field to the phase factor in the instanton is not single-valued. Physically the theory is OK because there is another contribution to the phase of the instanton action given by a Pfaffian associated to the worldsheet fermions in the heterotic string. This exactly cancels the above holonomy . Instead of taking a single Calabi–Yau target space with a family of 2-spheres we may take a family of Calabi–Yau target spaces containing a given 2-sphere. The above analysis holds with little modification and shows that going around a contractable loop in the moduli space of Calabi–Yau spaces can introduce an ambiguity in the associated $`B`$-field phase. In other words the $`B`$-field does not live in the flat bundle $`H^2(Z,\mathrm{U}(1))`$ over the moduli space! All we can say in general is that the $`B`$-field lives in a bundle over the moduli space whose generic fibre is a torus of dimension $`dimH^2(Z)`$. One way of avoiding this nastiness is by “embedding the spin connection in the gauge group”. In this very special case, the above holonomies disappear. One may also get the holonomies to vanish by taking the sizes to infinity by taking $`\alpha ^{}0`$. In both of these cases $`B`$ really does live in $`H^2(Z,\mathrm{U}(1))`$. #### 2.5.1 $`E_H`$ and its bundle Let us deal first with the bundle $`V_E`$ over the fixed torus $`E_H`$. This case is rather easy to analyze as the only bundles over $`E_H`$ which solve the equations of motion are ones which are flat. Because the tangent bundle and gauge bundle are flat we have $`dH=0`$ and avoid any curvature of the bundle in which the $`B`$-field lives. Thus $`BH^2(E_H,\mathrm{U}(1))`$. We are required to find the moduli space of flat $`\mathrm{G}_E`$-bundles over $`E_H`$. Let us assume that $`\mathrm{G}_E`$ is simply-connected. This problem was extensively analyzed in . A flat $`\mathrm{G}_E`$-bundle over $`E_H`$ is specified by its “Wilson lines”. That is, we specify a homomorphism $`\pi _1(E_H)\mathrm{G}_E`$ up to conjugation by $`\mathrm{G}_E`$. Since $`\pi _1(E_H)`$ is the abelian group $``$, we need to specify two commuting elements of $`\mathrm{G}_E`$. A useful result of Borel states that any two commuting elements of $`\mathrm{G}_0`$ may be conjugated simultaneously into the maximal Cartan subgroup $`T\mathrm{G}_0`$. This implies that our desired moduli space is $`T\times T`$ divided by any remnants of the conjugation equivalence. The latter is given precisely by the Weyl group $`W(\mathrm{G}_0)`$. The desired moduli space of bundles over a fixed $`E_H`$ is therefore $$\frac{T\times T}{W(\mathrm{G}_0)}.$$ (16) Now consider supplementing this data by the moduli of $`E_H`$ to get the full moduli space related to $`E_H`$. The Kähler form and $`B`$-field classically live in $`_+\times /`$ which may be exponentiated to give $`^{}`$. The moduli space of complex structures is given by the upper half plane, $`𝖧`$, divided by $`\mathrm{SL}(2,)`$ as is well-known. Note that this $`\mathrm{SL}(2,)`$ acts on the generators of $`\pi _1(E_H)`$ and thus on (16) by mixing the two $`T`$’s. We thus have ###### Proposition 4 The classical moduli space of $`\mathrm{G}_E`$-bundles on $`E_H`$ together with the moduli space of Ricci-flat metrics and $`B`$-fields on $`E_H`$ is given by $$\mathrm{SL}(2,)\backslash \left(𝖧\times \frac{T\times T}{W(\mathrm{G}_0)}\right)\times ^{}.$$ (17) This rather ugly-looking result becomes more pleasant when stringy considerations are taken into account. For example, let us divert our attention briefly to the case of a heterotic string compactified only on $`E_H`$.<sup>10</sup><sup>10</sup>10To be precise, we consider the component of the moduli space containing the trivial bundle. This implies $`\mathrm{G}_E=\mathrm{G}_0`$. This case was studied by Narain (see also ). The exact result is that the moduli space is given by $$\mathrm{O}(\mathrm{\Gamma }_{2,18})\backslash \mathrm{O}(2,18)/(\mathrm{O}(2)\times \mathrm{O}(18)),$$ (18) (times a real line for the dilaton). The lattice $`\mathrm{\Gamma }_{2,18}`$ is the even self-dual lattice of signature $`(2,18)`$ which is given by the root lattice of $`E_8\times E_8`$, or $`\mathrm{SO}(32)`$, supplemented by two orthogonal copies of $`U`$. We use the standard notation $`U`$ for the even self-dual lattice of signature $`(1,1)`$. Note that the heterotic string compactified on a single 2-torus has half-maximal supersymmetry and indeed the moduli space (18) is of a form promised in section 2.2. The only way that (18) differs from the classical statement (17) is that there are extra discrete identifications. See, for example, section 3.5 of for details of how these moduli spaces are mapped to each other. These extra identifications in the exact case are called “T-Dualities”. These T-Dualities include the familiar $`R1/R`$ dualities of the torus as well as dualities which mix moduli corresponding to the bundle with moduli corresponding to the base. When compactifying the heterotic string on $`S_H\times E_H`$ we will have fewer supersymmetries and so we have every reason to expect that quantum effects will have a more serious effect on the classical moduli space of vector bundles. We will see that this is so. * Let us return again for a moment to the eight dimensional case of the heterotic string only compactified on $`E_H`$. It is known that the moduli space of flat $`\mathrm{G}_0`$-bundles on a 2-torus is not connected. In the case of the $`E_8\times E_8`$ heterotic string it is believed to be a valid string model if the two $`E_8`$ factors are exchanged under holonomy around a non-contractable loop in the torus. These models were explored in . Such a bundle is not really an $`E_8\times E_8`$-bundle but is more accurately described as an $`(E_8\times E_8)_2`$-bundle where this latter $`_2`$ acts to exchange the two $`E_8`$ factors. Pedants who like to say “$`\mathrm{Spin}(32)/_2`$ heterotic string” rather than $`\mathrm{SO}(32)`$ heterotic string” should by all rights be expected to say “$`(E_8\times E_8)_2`$ heterotic string” rather than “$`E_8\times E_8`$ heterotic string”! Similarly a $`\mathrm{Spin}(32)/_2`$-bundle may have a nontrivial second Stiefel-Whitney class over the torus. Such a bundle is not homotopic to the trivial bundle and so lies in a different component of the moduli space. These classes of bundles have been studied in . In particular, a connection between these two classes was discussed in . See also for a nice mathematical treatment of these issues. We should expect the same kind of effects for various possibilities of $`\mathrm{G}_E`$ when we now compactify down to four dimensions. Monodromy can be expected to play a rôle around the cycles in $`E_H`$ whenever $`\mathrm{G}_E`$ admits an outer automorphism (possibly even if this outer automorphism was not induced by an endomorphism of $`\mathrm{G}_0`$). We may also obtain second Stiefel-Whitney classes whenever $`\mathrm{G}_E`$ is not simply-connected. It is probably fair to say that we do not have a full understanding of these disconnected components of the moduli space in the context of string duality at the present time. We will ignore this problem in these notes and implicitly assume that the flat bundles on $`E_H`$ are always homotopic to the trivial bundle. See also where another issue to do with the global form of the gauge group is raised. #### 2.5.2 $`S_H`$ and its bundle We now need to consider the bundle $`V_SS_H`$ subject to the anomaly cancellation condition. In the case that $`V_S`$ is an $`\mathrm{SU}(n)`$ bundle this would amount to $`c_2(V_S)=24`$ for example. In general this is a much harder problem to solve than the preceding case. Having said that, the bundle part of the problem is not too bad so long as we ignore quantum corrections. Work by Mukai (see also for a nice account of this work) tells us that we may put the hyperkähler structure of the K3 surface $`S_H`$ to good use. The basic result we will use is that the moduli space of stable vector bundles over $`S_H`$ will also have a hyperkähler structure. In fact, Mukai has shown that in many cases one may obtain a moduli space which is itself another K3 surface! The relationship between $`S_H`$ and this latter K3 surface may be viewed as a kind of mirror symmetry in some cases . We will have more to say about the bundle $`V_S`$ and its moduli space in the case that $`S_H`$ is an elliptic fibration in section 4.3.1 but for now we will just content ourselves with the knowledge that the moduli space has a hyperkähler structure. The moduli space of Ricci-flat metrics and $`B`$-fields on $`S_H`$ is given by $$\mathrm{O}(\mathrm{\Gamma }_{4,20})\backslash \mathrm{O}(4,20)/(\mathrm{O}(4)\times \mathrm{O}(20)).$$ (19) See for more details. The fact that it is a symmetric space may be deduced from its appearance in the moduli space of a type IIA string compactified on a K3 surface — which has half-maximal supersymmetry. Our complete moduli space of deformations of $`S_H`$ together with its bundle $`V_S`$ may therefore itself be viewed as a fibration. The base space of this fibration is given by (19) (or perhaps only some subspace of it) while the fibre is given by the hyperkähler moduli space of the bundle $`V_S`$. Note that (19) may be viewed as a quaternionic Kähler manifold (well, orbifold to be precise) from the fact that $`\mathrm{Sp}(1).\mathrm{Sp}(20)\mathrm{SO}(4)\times \mathrm{SO}(20)`$ (up to finite groups). Assuming the moduli space of $`S_V`$ varies over this space in a way compatible with this quaternionic structure we see that the total moduli space will also have a quaternionic Kähler structure. Our crude counting argument tells us immediately that this total moduli space of $`V_SS_H`$ should be identified with $`\mathrm{M}_H`$ leaving the remaining moduli in $`\mathrm{M}_V`$. Again one may be more careful along the lines of if one wishes. Anyway, to recap we have ###### Proposition 5 For the heterotic string compactified on $`(V_SS_H)\times (V_EE_H)`$ we have * $`\mathrm{M}_V`$ is spanned by the deformations of $`V_EE_H`$ (i.e., deformations of $`V_E`$ and deformations of the complex structure and complexified Kähler form on $`E_H`$) and by the dilaton-axion. It has complex dimension $`\mathrm{rank}(V_E)+3`$. * $`\mathrm{M}_H`$ is spanned by the deformations of $`V_SS_H`$. The dimension of the space $`\mathrm{M}_H`$ depends on several considerations and we do not compute it here. Note in particular that certain bundles put constraints on the K3 they live on and the complete form of (19) may not be seen. ### 2.6 Who gets corrected? So far we have listed the degrees of freedom present in a given string theory and then determined the classical picture of the resulting moduli space. This is not expected to be exact however — there will be corrections from various sources. To specify exactly how these corrections arise will again strongly test our knowledge about what string theory is exactly. Even though we don’t really know what string theory is, we do know enough to make statements about where we might expect quantum corrections to arise. An irrefutable statement about string theory is that it contains at least two limits in which we expect quantum field theory to provide a good picture (at least most of the time). The first quantum field theory consist of the two-dimensional worldsheet conformal field theory, i.e., the “pre-duality” picture of string theory. Indeed this picture gives us the “stringiness” in string theory! Secondly we have the effective quantum field theory which lives in the target spacetime dimensions. Consider first the worldsheet quantum field theory. This has an action $$\frac{1}{4\pi \alpha ^{}}_\mathrm{\Sigma }d^2\sigma \sqrt{\gamma }\left(\gamma ^{ab}g_{\mu \nu }(x)+iϵ^{ab}B_{\mu \nu }(x)\right)_ax^\mu _bx^\nu +\frac{1}{4\pi }_\mathrm{\Sigma }d^2\sigma \sqrt{\gamma }\mathrm{R}\mathrm{\Phi }_0+\mathrm{},$$ (20) where $`x`$ maps the worldsheet $`\mathrm{\Sigma }`$ into ten-dimensional spacetime. We have a worldsheet metric $`\gamma _{ab}`$, and target space metric and $`B`$-field given by $`g_{\mu \nu }`$ and $`B_{\mu \nu }`$. In addition $`\mathrm{R}`$ represents the worldsheet scalar curvature and $`\mathrm{\Phi }_0`$ is the dilaton which we assume to be independent of $`x`$. The difficulty in analyzing this model is that the metric and $`B`$-field vary as a function of the position in target space, $`x`$. The important point to note is that $`\alpha ^{}`$ (which sets the “string scale” in units of area) acts as a coupling constant. If $`\sqrt{\alpha ^{}}`$ is much less than a characteristic distance scale, $`R`$, of variations in the metric and $`B`$-field then $`x`$ represents almost “free” fields. We can then use a perturbation theory expanding in powers of $`\alpha ^{}/R^2`$. We may also have nonperturbative effects due to worldsheet instantons which contribute towards correlation functions as $`\mathrm{exp}(R^2/\alpha ^{})`$. These instantons are the maps $`x`$ which solve the equations of motion of (20) and are given in our context as holomorphic maps . To compute any correlation function using this worldsheet field theory version of string theory it is necessary to integrate over all worldsheets. This includes a sum over all genera with genus zero corresponding to tree-level, genus one giving one loop, etc. Such summands will be weighted by a relative factor of $`\mathrm{exp}(g\mathrm{\Phi }_0)`$, where $`g`$ is the genus of $`\mathrm{\Sigma }`$, thanks to the last term in (20). This picture of string theory induces an effective spacetime action proportional to<sup>11</sup><sup>11</sup>11We are being thoroughly negligent with factors of 2 etc., and we have omitted an overall factor. See section 3.7 of for a better discussion. $$d^{10}x\sqrt{g}e^{2\mathrm{\Phi }_0}\left(R_g+|\mathrm{\Phi }_0|^2+|dB|^2\right)+\mathrm{}$$ (21) We may use this as the basis of a spacetime quantum field theory. The important thing to note here is that $`\lambda =\mathrm{exp}(\mathrm{\Phi }_0)`$ appears as a coupling constant in this quantum field theory. This is hardly surprising given that the number of loops in this field theory corresponds to the genus of the worldsheet in the previous field theory. $`\lambda `$ is often called the “string coupling”. At the heart of the subtlety of string theory is that each of these field theories above contains the seeds for the other field theory’s downfall! As we have already mentioned, there are good reasons for believing that worldsheet instanton effects in the worldsheet conformal field theory make a complete understanding of spacetime in terms of Riemannian geometry unlikely. Thus the spacetime quantum field theory cannot really be considered in the form of the action (21). Equally, nonperturbative effects, such as instantons, coming from the spacetime field theory cannot be understood in terms of the genus expansion of the worldsheet theory. The best we can do is to assume that true string theory knows about both of these field theories and includes the nonperturbative effects from both simultaneously. This idea will become very important in section 4.4.2. The worldsheet picture of string theory can only really be considered to be an accurate picture of string theory when $`\lambda 0`$ and equally the spacetime effective action point of view can only be relied upon safely when $`\alpha ^{}0`$. These are the two limits of string theory where we really understand what is going on. We need to look at the moduli spaces of the previous section and ask how they may be warped by corrections coming from quantum effects of either of our two field theories. Fortunately it is not the case that all of the moduli spaces are affected by both corrections. We can see this from the holonomy argument in section 2.1 that the moduli space factorizes as $`\mathrm{M}_H\times \mathrm{M}_V`$ exactly. Let us consider $`\lambda `$-corrections first from the spacetime field theory. These must vanish as $`\lambda 0`$. Because of this they cannot affect the factor of the moduli space which does not contain the dilaton. Similarly the $`\alpha ^{}`$-corrections must disappear in the large radius limit of the compactification and so cannot affect a factor of the moduli space which does not know about sizes. One may try to argue that the moduli space of complex structures of a Calabi–Yau threefold does not know about size. Algebraic geometers can compute the moduli space of an algebraic variety without knowing about feet and inches! On the other hand it is the Kähler form which determines the volume of the threefold and so we might expect its moduli space to be subject to $`\alpha ^{}`$-corrections. One should be a little careful with this argument as varying the complex structure can vary volumes of object such as minimal 3-cycles in the threefold. That being said, this argument can be shown to be rigorously correct. For example, one may use topological field theory methods to show that the moduli space of complex structures is unaffected by quantum corrections from the worldsheet field theory . The results for which parts of the moduli spaces are affected by quantum corrections are given in table 2. We should note that some entries in this table may only be valid if only one of the coupling constants $`\alpha ^{}`$ or $`\lambda `$ is nonzero. For example if $`\lambda =0`$ then the moduli space $`\mathrm{M}_V`$ for the heterotic string is not prone to $`\alpha ^{}`$-corrections but this may not be true when $`\lambda `$ is nonzero. Upon compactification on a space $`X`$ to flat $`d`$-dimensional spacetime we obtain the spacetime effective action $$d^dx\sqrt{g}e^{2\mathrm{\Phi }}\left(R_g+|\mathrm{\Phi }|^2+|dB|^2\right)+d^dx\sqrt{g}g^{\mu \nu }G_{ij}_\mu \varphi ^i_\nu \varphi ^j+\mathrm{},$$ (22) from (21) where $`\varphi ^i`$ are coordinates on the moduli space as in (1). The quantity $`\mathrm{\Phi }`$ represents the effective $`d`$-dimensional dilaton and is given basically by $`\mathrm{\Phi }_0\frac{1}{2}\mathrm{log}\mathrm{Vol}(X)`$. In the compactification scenario this field theory is declared to be accurate. Because this part of spacetime is flat Minkowski space (or very nearly) we assert that worldsheet instantons are not allowed to spoil this field theory. This assumption is implicit in all of these lectures. Of course, this means that we are not allowed to ask questions about the $`d`$-dimensional physics which might probe effects such as quantum gravity. Then the compactification model would be invalid. ## 3 The Moduli Space of Vector Multiplets ### 3.1 The special Kähler geometry of $`\mathrm{M}_V`$ In order to discuss quantum corrections we need to establish limits on how much we are allowed to warp the moduli spaces consistent with the supersymmetry. We have said that $`\mathrm{M}_V`$ is Kähler and we can now put further limits on the structure of this moduli space. We wish to exploit the fact that the moduli space factor $`\mathrm{M}_V`$ for the type IIB string compactified on a Calabi–Yau space $`Y`$ is not warped at all by quantum corrections. The fact that $`\mathrm{M}_V`$ is given exactly in the form of a moduli space of complex structures on a Calabi–Yau threefold will allow us to ask more detailed questions about the differential geometry of $`\mathrm{M}_V`$. The deformations of complex structure of $`Y`$ are best thought of as variations of Hodge structure as we now explain. Any Calabi–Yau threefold has Hodge numbers $`h^{p,q}`$ in the form of a Hodge diamond $$\begin{array}{cccccccc}& & & \multicolumn{2}{c}{1}& & & \\ & & \multicolumn{2}{c}{0}& \multicolumn{2}{c}{0}& & \\ & \multicolumn{2}{c}{0}& \multicolumn{2}{c}{h^{1,1}}& \multicolumn{2}{c}{0}& \\ \multicolumn{2}{c}{1}& \multicolumn{2}{c}{h^{2,1}}& \multicolumn{2}{c}{h^{2,1}}& \multicolumn{2}{c}{1}\\ & \multicolumn{2}{c}{0}& \multicolumn{2}{c}{h^{1,1}}& \multicolumn{2}{c}{0}& \\ & & \multicolumn{2}{c}{0}& \multicolumn{2}{c}{0}& & \\ & & & \multicolumn{2}{c}{1}& & & \end{array}$$ (23) Of interest to us is the middle row of this diamond which relates to $`H^3(Y)`$. In particular we have a relationship between the Dolbeault cohomology groups and the integral cohomology: $$H^3(Y,)=H^{3,0}(Y)H^{2,1}(Y)H^{1,2}(Y)H^{0,3}(Y)=H^3(Y,)_{}.$$ (24) As we vary the complex structure the way in which the lattice $`H^3(Y,)`$ embeds itself into the space $`H^3(Y,)`$ “rotates” with respect to the decomposition of $`H^3(Y,)`$ into the Dolbeault cohomology groups. Consider a holomorphic 3-form $`\mathrm{\Omega }H^{3,0}(Y)`$. This is never zero anywhere on $`Y`$ and is uniquely defined up to a constant multiple thanks to the Calabi–Yau condition. Now consider a symplectic basis for $`H_3(Y)`$ given by $`A^a`$ and $`B_a`$ for $`a=1,\mathrm{},h^{2,1}+1`$ with intersections $`A^aB_b=\delta _b^a`$. Define the periods $$t^a=_{A^a}\mathrm{\Omega }\text{and}\mathrm{F}_a=_{B_a}\mathrm{\Omega }.$$ (25) These periods “measure” the complex structure of $`Y`$. Since $`Y`$ has only $`h^{2,1}`$ deformations of complex structure it is clear that not all of these periods may be independent. Firstly we have noted that $`\mathrm{\Omega }`$ is defined only up to a constant multiple so the periods can at best only be homogeneous coordinates in a projective space. Secondly it was shown by Bryant and Griffiths that, given all the $`t^a`$’s, all the $`\mathrm{F}_a`$’s are determined. That is, we may express the $`\mathrm{F}_a`$’s as functions of the $`t^a`$’s. Thus we are locally modeling the moduli space by $`^{h^{2,1}}`$. This gives us the correct dimension for the moduli space. (Note that the topology of the moduli space is unlikely to be that of a projective space as we have ignored the subtleties of degenerations so far. Also, the metric on the moduli space will not be the Fubini-Study metric. One way of seeing this is that some degenerations will be an infinite distance away from generic points in the moduli space.) It is then not hard to show, see for example section 3 of , that we may define a function $`\mathrm{F}`$ locally on the moduli space such that $$\begin{array}{cc}\hfill \mathrm{F}& =\frac{1}{2}\underset{c}{}t^c\mathrm{F}_c\hfill \\ \hfill \mathrm{F}(\lambda t^0,\lambda t^1,\mathrm{})& =\lambda ^2\mathrm{F}(t^0,t^1,\mathrm{})\hfill \\ \hfill \mathrm{F}_a& =\frac{\mathrm{F}}{t^a}.\hfill \end{array}$$ (26) We may rephrase this more globally in terms of bundle language following Strominger . The moduli space $`\mathrm{M}_V`$ has an “ample” line bundle $`L`$ such that $`c_1(L)`$ is given by the cohomology class of the Kähler form on $`\mathrm{M}_V`$. We also have an $`\mathrm{Sp}(h^{2,1}+1)`$-bundle $`\mathrm{H}`$ over $`\mathrm{M}_V`$ whose fibre is given by $`H^3(Y,)`$ in the fundamental representation. We then have sections $$\begin{array}{cc}\hfill \mathrm{\Omega }& \mathrm{\Gamma }(\mathrm{H}L)\hfill \\ \hfill \mathrm{F}& \mathrm{\Gamma }(L^2).\hfill \end{array}$$ (27) The important point is that the function $`\mathrm{F}`$, which is called the “prepotential” contains all the useful information we will need. The geometry of $`\mathrm{M}_V`$ is completely determined by it. This fact shows that $`\mathrm{M}_V`$ cannot be any old Kähler manifold. It is conventional to denote the special property that we have a prepotential by saying that $`\mathrm{M}_V`$ is “special Kähler”. Our discussion above takes the point of view that special Kähler geometry appears from the moduli space of complex structure on Calabi–Yau threefolds. This is not the original definition however. Special Kähler was first used to denote the geometry of the moduli space of scalar fields in vector multiplets of arbitrary $`N=2`$ supersymmetric field theories coupled to gravity in four dimensions as in . In this context, the projective coordinates $`t^a`$ are known as “special” or “flat” coordinates. The link between these points of view is that the Kähler metric on $`\mathrm{M}_V`$ given in the effective action (22) is given by $$\begin{array}{cc}\hfill G_{a\overline{b}}& =\frac{^2K}{t^a\overline{}t^b}\hfill \\ \hfill K& =\mathrm{log}\mathrm{Im}\left(\overline{t}^a\frac{\mathrm{F}}{t^a}\right).\hfill \end{array}$$ (28) The remarkable fact, as proved by Strominger in (see also a discussion of this in ), is that these points of view are equivalent. That is to say the local conditions arising from differential geometry for deformations of Hodge structure of a Calabi–Yau threefold are identical to the conditions on the moduli space of vector moduli in an $`N=2`$ theory of supergravity in four dimensions. It is well worth pausing to reflect on the implications of this statement. Since we have approached the question of supergravity in lower dimensions from the point of view of string theorists this statement may not seem particularly stunning — it is just a confirmation that things are working out nicely. Our moduli space of compact Calabi–Yau manifolds ties in nicely with the geometry of the moduli space of vacuum expectation values of the massless scalar particles in the uncompactified dimensions. This statement of equivalence does not depend on string theory however. What would we have made of it if we had not yet discovered string theory? It is as if the $`N=2`$ theories of supergravity in four dimensions “knew” that they were related in some way the Calabi–Yau threefolds. String theory, or at least ten-dimensional supergravity, provides this link nicely via compactification. Even if string theory turns out to be wrong for some reason, this link between $`N=2`$ theories and Calabi–Yau threefolds is irrefutable. We should provide a word of caution about the strength of the statements above. Just because a moduli space is consistent with these conditions that it be a deformation of Hodge structure of a Calabi–Yau manifold, it does not imply that such a Calabi–Yau manifold must exist. It is perhaps worthwhile to mention the following structure about special Kähler geometry which gives a hint as to why the $`N=2`$ theory “knows” about the Calabi–Yau 3-fold. Consider the following Hermitian form on $`H^3(Y,)`$: $$H_Y(\omega _1,\omega _2)=2i_Y\omega _1\overline{\omega }_2.$$ (29) It is easy to show (see for example) that the imaginary part of this form coincides with the usual cup product structure when restricted to $`H^3(Y,)`$. One may also show that * on $`H^{3,0}(Y)`$ the form $`H_Y`$ is negative definite, and * on $`H^{2,1}(Y)`$ the form $`H_Y`$ is positive definite. One may show that this is reflected in the fact that the the matrix $$\mathrm{Im}\left(\frac{^2\mathrm{F}}{t^at^b}\right),$$ (30) has signature $`(1,h^{2,1})`$. This signature nicely “separates” the $`H^{3,0}`$ part of the cohomology from the $`H^{2,1}`$ part. We can now argue (very) roughly as follows. If we had an even number of dimensions to our compact space then we wouldn’t have the right symplectic structure (e.g. the $`\mathrm{Sp}(h^{2,1}+1)`$ group above) on the middle dimension cohomology to see the correct variation of Hodge structure. If the compact space were complex dimension one, we wouldn’t have “enough room” in the Hodge diamond for an indefinite Hermitian form. If we had five or more complex compact directions we would expect something more complicated. Thus three dimensions is the most natural. We also obtain $`h^{3,0}=1`$ from the signature telling us that we must have a Calabi–Yau!<sup>12</sup><sup>12</sup>12Of course, we could do something like take a fivefold with Hodge numbers $`h^{5,0}=0`$ and $`h^{4,1}=1`$ which might give us the right structure. We consider this less natural than the Calabi–Yau threefold. We would like to emphasize again the fact that this discussion of special Kähler geometry depends on the exactness of the effective action (22). If true quantum gravity effects in four dimensions are considered we may expect much of this discussion to break down. Indeed, the statement that we have a moduli space in the form of a Riemannian manifold (or orbifold etc.) would then be suspect. ### 3.2 $`\mathrm{M}_V`$ in the type IIA string Now we wish to look at the way that $`\mathrm{M}_V`$ is seen in the type IIA string on the Calabi–Yau threefold $`X`$. This involves the old work of mirror symmetry. Since there have been numerous reviews of mirror symmetry we will be fairly brief here and focus only the warping of the special geometry of $`\mathrm{M}_V`$. #### 3.2.1 Before corrections and five dimensions As noted in section 2.6 $`\mathrm{M}_V`$ consists of the moduli space of the Kähler form on $`X`$ but is subject to corrections coming from worldsheet instantons. Let us first establish what it would look like if there were no quantum corrections. One may approach this directly as in or one may proceed in a slightly different way via M-theory. We will do the latter (as most string theory students these days are perhaps even better acquainted with M-theory than with string theory!). The first thing to note is that an $`N=2`$ theory in four dimensions may be obtained by compactifying an $`N=1`$ theory in five dimensions on a circle. Then if we reinterpret the IIA string theory as M-theory on a circle we see that this five-dimensional theory may be obtained by compactifying M-theory on the Calabi–Yau threefold $`X`$. To put it another way we may consider the limit of a type IIA string on $`X`$ as the string coupling becomes very strong. In the same way that the ten-dimensional type IIA string becomes eleven-dimensional M-theory in this limit, the four dimensional $`N=2`$ theory will turn into the five-dimensional $`N=1`$ theory. The useful thing about this limit for our purposes is that the effective string scale given by $`\alpha ^{}`$ tends to zero in this limit. Thus stringy effects such as worldsheet instantons are completely suppressed. This is explained nicely by Witten . The general idea is that the metric in the uncompactified directions needs to be rescaled as we change dimension (just as it is going from ten dimensions to eleven dimensions ). This rescaling is infinite taking the string scale to zero size. A vector multiplet in five dimensions contains only one real scalar as opposed to the two scalars coming from the four dimensional vector multiplet. The rescaling between the type IIA theory and M-theory also causes a slight reshuffling of moduli as explained in . The result is that we have a moduli space $`\mathrm{M}_V^5`$ of vector multiplets which is a real space of dimension $`h^{1,1}1`$. This is the classical moduli space of cohomology classes Kähler forms on $`X`$ of fixed volume. The deformation corresponding to the volume defects to the hypermultiplets replacing the lost dilaton leaving the moduli space $`\mathrm{M}_H`$ unchanged between four and five dimensions. The compactification of M-theory on a smooth Calabi–Yau threefold yields $`h^{1,1}`$ vector fields from the M-theory 3-form in eleven dimensions. This yields a supersymmetric $`\mathrm{U}(1)^{h^{1,1}}`$ gauge theory (with gravity) in five dimensions. The action for such a field theory contains the interesting “Chern-Simons”-like term $$d^5x\kappa _{abc}F^aF^bA^c$$ (31) where $`\kappa _{abc}`$ is symmetric in its indices $`a,b,c=1,\mathrm{},h^{1,1}(X)`$. As usual with these topological types of terms in field theory one may compute $`\kappa _{abc}`$ from the intersection theory of $`X`$. In this case one discovers that $$\kappa _{abc}=_Xe_ae_be_c,$$ (32) where the $`e_a`$’s are the generators of (the free part of) $`H^2(X,)`$. Equivalently we may use 4-cycles $`D_a`$ in $`H_4(X,)`$ dual to $`e_a`$ and obtain intersection numbers: $$\kappa _{abc}=D_aD_bD_c.$$ (33) Furthermore, as explained in , we may put homogeneous coordinates $`\xi ^a`$ on $`\mathrm{M}_V^5`$ such that the metric is given by $$G_{ab}=\frac{^2}{\xi ^a\xi ^b}\mathrm{log}\left(\kappa _{cde}\xi ^c\xi ^d\xi ^e\right).$$ (34) This should be regarded as the “special” real geometry of $`\mathrm{M}_V^5`$ where the “prepotential” is given by a pure cubic $`\mathrm{F}_5=\kappa _{cde}\xi ^c\xi ^d\xi ^e`$. Relationships of this to Jordan algebras are discussed in . We may instead regard $`\xi ^a`$ as the affine coordinates in $`H^2(X,)=^{h^{1,1}}`$. The Kähler form is then given by $$J=\xi ^ae_a,$$ (35) and the condition that we fix the volume to, say one, for $`\mathrm{M}_V^5`$, is given by $$JJJ=\mathrm{F}_5=1.$$ (36) This latter condition can also be seen directly from supergravity without reference to the geometrical interpretation of $`X`$ as in . Thus again we see strong hints that five dimensional $`N=1`$ supergravity “knows” that it has something to do with Kähler threefolds even without direct reference to M-theory. Note that the moduli space $`\mathrm{M}_V^5`$ is not the complete hypersurface $`\mathrm{F}_5=1`$ in $`H^2(X,)`$. It turns out that phase transitions occur precisely on the walls of the Kähler cone to truncate this hypersurface to lie completely within the Kähler cone. This is discussed in for example but we will not pursue it here. We may now perform crude dimensional reduction of this five-dimensional field theory to render it a four-dimensional theory. Recall that dimensional reduction simply asserts that the fields have no dependence on motion in the directions we wish to lose and we decompose the vectors, tensors, etc accordingly into lower-dimensional objects. Performing this operation is a lengthy operation but the result for the moduli space is straight-forward. As required, we obtain special Kähler geometry for $`\mathrm{M}_V`$ in four dimensions. Now we have complex homogeneous coordinates $`t^0,t^a`$, where $`a`$ still runs $`1,\mathrm{},h^{1,1}`$ and a prepotential $$\mathrm{F}_0=\frac{\kappa _{abc}t^at^bt^c}{t^0}.$$ (37) The very important point to realize however is that dimensional reduction is not necessarily the same thing as compactification on a circle (as emphasized in for example). The problem is that solitons present in the five-dimensional theory can become instantons in the four dimensional theory and add quantum corrections to the picture. We may only regard (37) as the classical contribution to the prepotential. We may expect quantum corrections to appear with respect to $`\alpha ^{}`$ as noted earlier. Note that (37) may be computed as the classical contribution directly without a foray into five dimensional physics . The five dimensional picture is probably worth being aware of however as it offers many insights. #### 3.2.2 Mirror Pairs The easiest way of computing the quantum corrections to the prepotential of the type IIA string compactified on $`X`$ is to use a duality argument in the form of mirror symmetry. That is, can we find a Calabi–Yau threefold $`Y`$ such that the type IIB string compactified on $`Y`$ yields the same physics in four dimensions as the type IIA string compactified on $`X`$? If this is the case $`X`$ and $`Y`$ are said to be “mirror” Calabi–Yau threefolds. Given the current state of our knowledge of string theory it is probably not possible to rigorously prove that any such pairs $`X`$ and $`Y`$ satisfy this condition. We can come fairly close however. The reason is that because the dilaton of each of the two type II strings appears in the moduli space of hypermultiplets in a similar way, we may choose both strings to be simultaneously very weakly coupled over the whole moduli space $`\mathrm{M}_V`$. This allows us to reliably use the worldsheet approach to analyze mirror pairs. Thus the construction of mirror pairs of Calabi–Yau can be reduced to a conformal field theory question in two dimensions. We will then assume that if two theories are mirror at this conformal field theory level then they will be mirror pairs in the full nonperturbative string theory picture. The canonical example of mirror pairs of conformal field theories is provided by the Greene–Plesser construction . An explanation of this would require a major diversion into the subtleties of conformal field theories which would take us way beyond the scope of these lectures. We will then content ourselves to quote their result. We refer to for more details. Consider the weighted projective space $`_{\{w_0,w_1,w_2,w_3,w_4\}}^4`$ with homogeneous coordinates $$[x_0,x_1,\mathrm{},x_4][\lambda ^{w_0}x_0,\lambda ^{w_1}x_1,\mathrm{},\lambda ^{w_4}x_4],$$ (38) for $`\lambda ^{}`$. We may now consider the hypersurface $`X`$ given by $$x_0^{\frac{d}{w_0}}+x_1^{\frac{d}{w_1}}+\mathrm{}+x_4^{\frac{d}{w_4}}=0,$$ (39) where $`d=w_i`$ and we impose the condition $$\frac{d}{w_i},\text{for all }i.$$ (40) The projective space will generically have orbifold singularities along subspaces. These orbifold loci may be blown-up to smooth the space and we assume that $`X`$ is transformed suitably along with this blowing up process to render it smooth. The Greene–Plesser statement is then ###### Proposition 6 $`X`$ is mirror to $`Y`$, where $`Y`$ is the (blown-up) orbifold $`X/G`$ and $`G`$ is the group with elements $$g:[x_0,x_1,\mathrm{},x_4][\alpha _0x_0,\alpha _1x_1,\mathrm{},\alpha _4x_4],$$ (41) where we impose $`\alpha _i^{\frac{d}{w_i}}=1`$ for all $`i`$ and $`\alpha _i=1`$. The essence of this statement can be generalized considerably to hypersurfaces in toric varieties and to complete intersections in toric varieties as was done by Batyrev and Borisov . This Batyrev–Borisov statement is not yet understood at the level of conformal field theory but the evidence that it does produce mirror pairs is very compelling. Thus there are a very large number of candidate mirror pairs of Calabi–Yau threefolds. #### 3.2.3 The mirror map Knowing the mirror partner $`Y`$ of $`X`$ is a good start to knowing how to compute the quantum corrections to the prepotential of the type IIA compactification but we need a little more information. Namely, we need to know exactly how to map the coordinates on our special Kähler moduli spaces between the type IIA and the type IIB picture. This is known as the “mirror map”. The most direct way of finding the mirror map is to take a little peek into the moduli space of hypermultiplets even though we are only concerned with vector multiplets in this section. The fact we need to borrow from hypermultiplets is that the Ramond-Ramond moduli must be mapped into each other under mirror symmetry. For the hypermultiplet moduli spaces we wrote down in section 2.4 this shows that $`H^{\text{odd}}(X,\mathrm{U}(1))`$ is mapped to $`H^{\text{even}}(Y,\mathrm{U}(1))`$. The next statement we need concerns the symmetry of mirror pairs themselves. We may state this as ###### Proposition 7 If $`X`$ and $`Y`$ are mirror pairs then so are $`Y`$ and $`X`$. That is, the type IIA string on $`Y`$ is physically equivalent to the type IIB string on $`X`$. This statement is completely trivial in terms of the definition of mirror pairs at the level of conformal field theory. See for example for more details. Here, since we are trying to be careful about not specifying our definition of string theory, we will just have to assume that this proposition is true. We therefore may assume that $`H^{\text{even}}(X,\mathrm{U}(1))`$ is mapped to $`H^{\text{odd}}(Y,\mathrm{U}(1))`$ under the mirror map. This implies some map between $`H^{\text{even}}(X,)`$ and $`H^{\text{odd}}(Y,)`$. This map between the integral structures of the even cohomology of $`X`$ and the odd cohomology of $`Y`$ is very interesting and forms one of the most powerful ideas in mirror symmetry. Clearly it cannot be an exact statement at the level of classical geometry. This is because as we wander about the moduli space of complex structures on $`Y`$ we may induce monodromy on $`H^{\text{odd}}(Y,)=H^3(Y,)`$. That is, if we pick a certain basis for integral 3-cycles in $`Y`$ we may go around a closed loop in moduli space which maps this basis nontrivially back into itself. If this statement were then mapped into a statement about the type IIA string on $`X`$ we would conclude that some even-dimensional cycle, such as a 0-cycle, could magically transmute into a 2-cycle as we move about the moduli space of complexified Kähler forms. Clearly this does not happen! * To explain this effect in geometric terms Kontsevich has a very interesting proposal based on some ideas by Mukai . Rather than thinking in terms of $`H^{\text{even}}(X,)`$ directly one may consider $`𝐃(X)`$, the derived category of coherent sheaves on $`X`$. Objects in $`𝐃(X)`$ are basically complexes of sheaves of the form $`\mathrm{}\mathrm{F}_1\mathrm{F}_2\mathrm{F}_3\mathrm{}`$. The automorphisms of $`H^3(Y,)`$ induced by monodromy can then be understood in many cases in terms of automorphisms of $`𝐃(X)`$ . Objects in $`𝐃(X)`$ can then be mapped into $`H^{\text{even}}(X,)`$ essentially by using the Chern character. This is a fascinating subject somewhat in its infancy that promises much insight into mirror symmetry and stringy geometry. Instead this statement must only be true classically at the large radius limit of $`X`$ and thus the corresponding “large complex structure” limit of $`Y`$. Somehow near this limit point, and in particular monodromy about this point, these two integral structures must align classically. This was first suggested in and then explained more clearly in . Let us suppose we fix a point in the moduli space of complex structures, $`\mathrm{M}_V`$, on $`Y`$ which will be our candidate limit point. As this is a limit point it is natural to expect that $`Y`$ will be singular here. Actually one expects to find singular $`Y`$’s along complex co-dimension one subspaces of $`\mathrm{M}_V`$. This special limit point turns out to be a particularly nasty singularity as it lies on an intersection of many such divisors. We will assume there are in fact $`h^{2,1}(Y)=dim\mathrm{M}_V`$ such divisors intersecting transversely at this limit point. If this is not the case then one may blowup using standard methods in algebraic geometry to reduce back to this case. We may now consider the monodromy matrices $`M_k`$ which act on $`H^3(Y,)`$ as we go around each of these divisors. Mapping back to $`X`$ this limit point should be the large radius limit as every component of the Kähler form tends to infinity. The monodromy about this limit is $`BB+v`$, where $`vH^2(X,)`$. This fixes the mirror map as follows. First we need to switch back to the dual language of periods defined in (25). We will find one period which we denote $`t_0`$ which is completely invariant under the monodromies $`M_k`$. We also find periods $`t_k`$ such that $$\begin{array}{cc}\hfill M_k:t_k& t_k+t_0\hfill \\ \hfill M_j:t_k& t_k,\text{for }kj\hfill \end{array}$$ (42) where $`k=1,\mathrm{},h^{2,1}(Y)`$. The fact we may do this is a special property of the limit point we have chosen and defines the property that it can represent the mirror of a large radius limit point. This is explained in more detail in . We now give the mirror map: $$(B+iJ)_k=\frac{t_k}{t_0},$$ (43) where $`B+iJ=(B+iJ)_ke_k`$ is the complexified Kähler form on $`X`$ expanded over a basis $`e_kH^2(X,)`$. This is the only map possible which gives the correct monodromy and reflects the projective symmetry of the homogeneous coordinates $`t_a`$. The canonical example is that of the quintic threefold as computed in . In this case $`X`$ is the quintic hypersurface in $`^4`$ and thus $`Y`$ is $`X/(_5)^3`$ according to proposition 6. This case is fairly straight-forward as $`\mathrm{M}_V`$ is only one dimensional since $`h^{1,1}(X)=h^{2,1}(Y)=1`$. In this case one can compute the periods and use (26) to compute $$\mathrm{F}=(t_0)^2\left(5t^3+\frac{33}{2}t^2\frac{25}{72}t+\frac{25i}{12\pi ^3}\zeta (3)+\frac{2875i}{72\pi ^3}e^{2\pi it}+O(e^{4\pi it})\right),$$ (44) where $`t=t_1/t_0`$ can be viewed as the single component of the complexified Kähler form on $`X`$. Note we indeed get the correct leading term $`5t^3`$ from the intersection theory but there are an infinite number of quantum corrections. The quadratic and linear terms in $`t`$ are physically meaningless whereas the constant term proportional to $`\zeta (3)`$ is a loop term correcting the metric. The power series in $`q=e^{2\pi it}`$ corresponds to the worldsheet instanton corrections. An instanton in the worldsheet quantum field theory (20) corresponds to a holomorphic map from $`\mathrm{\Sigma }`$ into the target space $`X`$ . At tree-level in string theory such objects are therefore “rational curves” (i.e. holomorphic complex curves of genus zero). This is an important subject and any respectable review of $`N=2`$ theories in string theory should go into some detail about these rational curves. We will not do this however as there are already a number of reviews of this subject. As is explained in numerous places elsewhere, the quantum corrections may be used to count the numbers of rational curves in $`X`$. Indeed the 2875 appearing in (44) corresponds to the number of lines on a quintic surface. The interested reader should consult , for example, for much more information about this vast subject. One rough and ready way to appreciate why rational curves should make an appearance in mirror symmetry is as follows. We have already argued that the truly stringy geometry of $`X`$ must somehow mix up the notion of 0-cycles, 2-cycles, 4-cycles, etc as we move away from the large radius limit. These worldsheet instanton corrections near the large radius limit can be thought of as the way that 2-cycles (i.e., rational curves) start to mix into our notion of 0-cycles (i.e., points). This stringy geometry which can mix the notion of points and rational curves has yet to be understood properly. ### 3.3 $`\mathrm{M}_V`$ in the heterotic string Now we will consider the moduli space $`\mathrm{M}_V`$ in terms of the heterotic string compactified on $`S_H\times E_H`$. For a field theorist with a bias towards gauge theories this is actually the most useful way of viewing the resulting $`N=2`$ theories in four dimensions as we now explain. #### 3.3.1 Supersymmetric abelian gauge theories An abelian gauge theory of $`\mathrm{U}(1)^{n+2}`$ in flat space is based on the action $$d^4x\left(\frac{1}{\lambda ^2}\underset{a=1}{\overset{n+2}{}}F^a^2+\mathrm{}\right),$$ (45) where $`\lambda `$ is the gauge coupling constant. If this is an $`N=2`$ supersymmetric theory then $`n+1`$ of the $`\mathrm{U}(1)`$’s are associated to vector multiplets and the extra $`\mathrm{U}(1)`$ is the “graviphoton” coming from the supergravity multiplet. If the gauge particles are actually fundamental strings then the coupling constant should be given by the string dilaton as in equation (21) and the discussion following this equation. In the heterotic string, the dilaton lives in one of the $`n+1`$ vector multiplets and pairs up with the axion to form the complex field $$s=a+\frac{i}{\lambda ^2},$$ (46) where $`a`$ is the axion. We should note the difficulty of trying to find such a theory by compactifying a type II string. Here the dilaton lives in a hypermultiplet which cannot couple to the vector bosons in the desired way. Thus the gauge coupling cannot be interpreted as a type II string coupling — gauge bosons cannot be fundamental strings. The fact that such a term is expected in the action immediately tells us the form of the prepotential governing $`\mathrm{M}_V`$. This is perhaps easiest to see if we compactify the heterotic string first on $`S_H`$ times a circle to get an $`N=1`$ theory in five dimensions, and then compactify on a circle to get our desired four-dimensional theory. The theory in five dimensions will have generic gauge symmetry $`\mathrm{U}(1)^{n+1}`$ as compactifying on a circle gives a $`\mathrm{U}(1)`$ via the Kaluza–Klein mechanism. The interesting term in the five-dimensional theory is the Chern-Simons-like term (31). What will this reduce to when we compactify on the circle down to four dimensions? The answer is that we will replace the vector field $`A`$’s by four-dimensional real scalar $`a`$’s to form a term $$d^4x\kappa _{ebc}a^eF^bF^c.$$ (47) But this is the famous CP-violating term of a gauge theory and the scalar field is playing the rôle of an axion. If we want the kinetic term in the standard form (45) then the only scalar which is allowed to play the rôle of an axion in a theory with $`N=2`$ supersymmetry is the axion partner of the dilaton, namely the $`a`$ in (46). In addition this axion is not allowed to appear as a coefficient in front of field strengths associated with the $`\mathrm{U}(1)`$ gauge boson in the same multiplet as the axion and dilaton. This would lead to rather unorthodox terms proportional to $`\lambda ^4`$ in the action under supersymmetrization. This implies immediately that the superpotential in five dimensions is of the form $`\mathrm{F}_5=st^it^j\gamma _{ij}`$, for some matrix $`\gamma _{ij}`$, for $`i,j=1,\mathrm{},n`$ and where the $`t^i`$’s are the five-dimensional moduli fields associated to the vector supermultiplets other than the dilaton. Of course we expect this cubic superpotential to be corrected when we are in four dimensions just as the cubic potential was corrected for the type IIA compactifications in section 3.2. This time however the corrections will not be $`\alpha ^{}`$-corrections due to worldsheet instantons but they will be $`\lambda `$-corrections due to gauge instantons in spacetime. The cubic superpotential $`\mathrm{F}=st^it^j\gamma _{ij}`$ which is exact in five dimensions and correct to leading order in four dimensions is rather constraining. We may also note that in order for the kinetic term for the photons to be positive-definite it is required that $`\gamma _{ij}`$ have signature $`(+,,,,\mathrm{})`$ . Running through a lengthy process using the definitions of special Kähler geometry this leads to a moduli space for our four-dimensional theory locally of the form $$\mathrm{M}_{V,0}^{\text{Het}}=\frac{\mathrm{SL}(2,)}{\mathrm{U}(1)}\times \frac{\mathrm{SO}_0(2,n)}{\mathrm{SO}(2)\times \mathrm{SO}(n)},$$ (48) before quantum corrections. The first thing to note about this space is that it is a product of two symmetric spaces — just the kind of thing we would expect if we had more supersymmetry. The second thing to note is that the second term looks a lot like Narain’s moduli spaces as we discussed in section 2.5.1. This term represents just what we would expect if we look at the stringy moduli space of vector bundles of rank $`n2`$ over a 2-torus, together with deformations of the torus itself. This is excellent news. It means that if we identify the first term with the dilaton and axion then we have a very natural interpretation of this moduli space in terms of the data discussed in section 2.5.1. To get the moduli space perfectly correct we need to worry about the global form. If the second term really is the Narain moduli space of the bundle $`V_EE_H`$ discussed in section 2.5.1 then we should really expect it to be of the form $$\mathrm{O}(\mathrm{{\rm Y}})\backslash \mathrm{O}(2,n)/(\mathrm{O}(2)\times \mathrm{O}(n)),$$ (49) where $`\mathrm{{\rm Y}}`$ is the lattice of signature $`(2,n)`$ given by $`\mathrm{\Gamma }_{2,2}L`$ and where $`L`$ is the Cartan lattice of the structure group of $`V_E`$ with negative definite signature. One might also be tempted to replace the first term of (48) by the expression $$\mathrm{SL}(2,)\backslash \mathrm{SL}(2,)/\mathrm{U}(1).$$ (50) This $`\mathrm{SL}(2,)`$ would certainly respect the axion shift symmetry $`aa+2\pi `$ which we expect to be correct but it would also imply some strong-weak coupling duality for $`N=2`$ theories in four dimensions. This does not exist in general. The problem is that moduli space (48) ignores quantum corrections and it therefore only correct as the dilaton tends to $`\mathrm{}`$. The only part of $`\mathrm{SL}(2,)`$ which preserves this limit is the axion shift symmetry. If we could find a type IIB string compactification dual to this heterotic model then we could compute the prepotential exactly, just as we did by mirror symmetry for the type IIA string. This would allow us to compute the nonperturbative corrections to the moduli space arising from quantum corrections due to $`\lambda `$. Oddly enough it is much more natural to ask first for a type IIA string dual to the heterotic model we desire. #### 3.3.2 Heterotic/Type IIA duality If the type IIA string compactified on $`X`$ is dual to the heterotic string model giving the gauge theory of section 3.3.1 then we know a surprising amount of the geometry of $`X`$ with very little effort. The fact that the prepotential is $$\mathrm{F}_0=st^it^j\gamma _{ij}$$ (51) to leading order tells us about the cup product structure of $`H^2(X,)`$ or equivalently, the intersection form on $`H_4(X,)`$. In particular from (33) it tells us that the 4-cycle $`S`$ representing the complexified dilaton $`s`$ satisfies $$SSD=0,$$ (52) for any $`D`$ (whether it be associated to $`s`$ or a $`t^i`$). This implies that $`SS`$ is empty. One may now proceed to show that 1. $`S`$ can be represented by an algebraic surface embedded in $`X`$. 2. $`S`$ is a K3 surface. 3. Moving $`S`$ parallel to itself (as suggested by $`SS=0`$) sweeps out all of $`X`$. That is, $`X`$ is a K3-fibration. As this is reviewed at length in we will not repeat the proof here. It is not hard to show that in order for $`X`$ to be a Calabi–Yau manifold with $`\mathrm{SU}(3)`$ holonomy it must have finite (or trivial) fundamental group $`\pi _1(X)`$. For a K3 fibration $`XW`$, this implies that the base $`W`$ also has finite $`\pi _1`$. Thus if $`W`$ is a smooth space of complex dimension one, it must be isomorphic to $`^1`$. Anyway, not only do we now know that $`X`$ is a K3-fibration, we also know exactly which modulus of the complexified Kähler form corresponds to the dilaton-axion. We know that the element of $`H_4(X)`$ corresponding to $`S`$ is the homology class of a generic K3 fibre. We need the component of the Kähler form which controls the size of a 2-cycle which is dual (via intersection theory) to this K3 fibre. For simplicity we could assume that $`X`$ as a K3-fibration has a global section.<sup>13</sup><sup>13</sup>13This section need not be unique and in the example in section 3.3.4 it will not be. Its homology class and hence its area is unique however. That is, we have an embedding $`WX`$ which is an “inverse” of the fibration projection. This section acts as a 2-cycle dual to the K3 fibre. We have thus shown ###### Proposition 8 If a type IIA string on $`X`$ is dual to a heterotic string on a K3 surface times a torus, then $`X`$ must be a K3 fibration. Assuming this fibration has a section then the area of this section (and the corresponding component of the $`B`$-field) maps to the dilaton (and axion) on the heterotic side. We refer to for a careful statement of the assumptions which go into this proposition. People often loosely refer to the area of the section as the “area of the base”. If $`X`$ does not have a section then this duality can still work — we just have to work a little harder to determine the dilaton. We will always assume there is a section. At this point it is worthwhile to consider a sketchy picture of instanton corrections in this dual pair. On the heterotic side we have spacetime instanton effects<sup>14</sup><sup>14</sup>14The observant reader will note that we had assumed that we had an abelian gauge theory. Therefore we don’t really have any instantons in the gauge theory. We will see in section 3.3.3 that, if we want, there really is a nonabelian gauge theory lurking here. which produce effects of the order $`\mathrm{exp}(ns)`$ in correlation functions. In the type IIA picture one gets exactly the same effects thanks to the above mapping by wrapping worldsheet instantons around the section of the fibration. Thus spacetime instanton effects in the heterotic string are exchanged with worldsheet instanton effects in the type IIA string. One can consider this statement to be rather profound. It shows that neither the worldsheet picture nor the spacetime picture of the quantum field which “models” string theory can be more fundamental than the other. At least in the sense of instanton corrections, the two pictures may be interchanged. Now we have discussed one of the moduli in $`\mathrm{M}_V`$, let us find the others. Note first that although $`X`$ is a K3 fibration, not all of its fibres need be K3 surfaces. We only demand that the generic fibre is a K3 surface. We refer to fig. 1 for a picture of the K3 fibration. There are three ways to obtain contributions towards $`H^2(X)`$ in terms of $`X`$ as a K3 fibration. Let us list the different generators of $`H^2(X)`$ in the language of deformations of the Kähler form: 1. Deform the area of the section $`WX`$. 2. Deform the areas of curves within the generic K3 fibres. 3. Deform the independent volumes of the irreducible components of a reducible bad fibre. In terms of elements of $`H_4(X)`$, the contributions of type II are obtained by taking a 2-cycle $`C_i`$ in a generic fibre and then sweeping it out by moving over $`W`$ to produce a 4-cycle which we denote $`D_i`$. In this way, the intersection pairing $`C_iC_j`$ between 2-cycles in the K3 fibre is copied into the intersection numbers $`SD_iD_j`$ for $`X`$. The $`C_i`$’s in the K3 fibre are not just any old 2-cycle. They have to be algebraic curves, i.e., holomorphically embedded. It can be shown that the intersection form with a K3 surface of algebraic curves is an indefinite quadratic form of signature $`(+,,,,\mathrm{})`$. This shows that the moduli coming from contributions of type I and II to $`H^2(X)`$ from the K3 fibrations form the special Kähler geometry with prepotential $`\mathrm{F}=st^it^j\gamma _{ij}`$ to leading order as required in section 3.3.1. Indeed, one may prove the following (as is done in section 3.4 of for example): ###### Proposition 9 The moduli space of the Kähler form and $`B`$-field for a type IIA string on an algebraic K3 surface $`S`$ is given by $$\mathrm{O}(\mathrm{{\rm Y}})\backslash \mathrm{O}(2,n)/(\mathrm{O}(2)\times \mathrm{O}(n)),$$ (53) where $`\mathrm{{\rm Y}}=\mathrm{Pic}(S)\mathrm{\Gamma }_{1,1}`$ and $`\mathrm{Pic}(X)`$ is the “Picard lattice” given by the algebraic curves in $`S`$ together with their intersection form. The integer $`n`$ is given by the dimension of the Picard lattice. This gives a precise isomorphism between the moduli of type II above and the Narain factor of the moduli space of the heterotic string. There is one more result we may state here which will be useful later on. Given the Narain moduli space of the heterotic string on $`T^2`$ as given in (49) we can see that $`\mathrm{{\rm Y}}`$ must contain $`\mathrm{\Gamma }_{2,2}`$ and so $`\mathrm{Pic}(S)`$ will contain $`\mathrm{\Gamma }_{1,1}`$. This is actually a necessary and sufficient condition for $`S`$ to be an elliptic K3 surface with a least one section . The fibration structure of each K3 fibre extends to the following statement about the whole of $`X`$: ###### Proposition 10 If a type IIA string on $`X`$ is dual to a heterotic string on a K3 surface times a torus, and we see the full moduli space of the heterotic torus, then $`X`$ is an elliptic fibration over some complex surface with at least one section. So finally what about contributions of type III? These 4-cycles can be associated with components of reducible fibers which do not intersect the section. This lack of intersection with $`S`$ violates the expected special Kähler geometry from section 3.3.1. It turns out that these moduli will be something to do with the full nonperturbative physics of the heterotic string — more than is described by the effective action discussed in section 3.3.1. We will have more to say about these type III divisors later. #### 3.3.3 Enhanced gauge symmetry We now want to deal with the important subject of enhanced nonabelian gauge symmetries in the effective four-dimensional uncompactified dimensions. To simplify our discussion we will tackle only the subject of simply-laced Lie algebras, i.e., the “ADE” series of Lie algebras whose roots are all the same length. We will also ignore the subject of the global topology of the gauge group. That is we will not concern ourselves too much with whether or not a gauge group is really $`\mathrm{SU}(2)`$ or $`\mathrm{SO}(3)`$ for example. We refer the reader to for details about these subtleties which we are ignoring. To leading order we have a factor looking like (49) in the moduli space which we recognize as the Narain moduli space for some vector bundle, $`V_E`$, on a 2-torus $`E_H`$. The moduli here may be regarded as deformations of the flat metric on the torus itself together with deformations of the flat bundle. As discussed in section 2.5.1, the parameters controlling the bundle are known as “Wilson lines”. They measure the holonomy of the bundle as we go around non-contractable loops within $`E_H`$. As observed in section 2.3, the observed gauge group in the uncompactified dimensions which remains unbroken by the compactification process can be regarded as the centralizer of the holonomy acting on the ten-dimensional primordial gauge group. For generic values of the Wilson lines the holonomy of $`V_E`$ is $`\mathrm{U}(1)^{n2}`$, where $`n2`$ is the rank of the structure group $`\mathrm{G}_E`$ (which we assume to be simply-laced) of $`V_E`$. This holonomy is simply the Cartan subgroup of $`\mathrm{G}_E`$ and so the unbroken part of the gauge symmetry is $`\mathrm{U}(1)^{n2}`$. (Note that the compactification process on $`E_H`$ adds four more $`\mathrm{U}(1)`$’s to bring the total to $`n+2`$ as in section 3.3.1.) The interesting question arises as to what happens when the holonomy of the bundle $`V_E`$ is not generic. If we switch off some of the Wilson lines, we might expect the structure group of $`V_E`$ to decrease allowing for a larger centralizer. That is, the observed gauge symmetry in four dimensions should become larger. The idea is that the moduli space $`\mathrm{O}(2,n)/(\mathrm{O}(2)\times \mathrm{O}(n))`$ is viewed as the Grassmannian of space-like (positive) 2-planes $`\mathrm{}^{2,n}`$. One may also embed the lattice $`\mathrm{{\rm Y}}`$ into this same $`^{2,n}`$. The desired moduli space (49) is then this Grassmannian divided out by the automorphisms of the lattice $`\mathrm{{\rm Y}}`$. The rule is then as follows: ###### Proposition 11 The observed gauge group in uncompactified space has rank $`n+2`$. The roots of the semi-simple part of this gauge group correspond to elements of $`\mathrm{{\rm Y}}`$ which have length squared $`2`$ and which are orthogonal to $`\mathrm{}`$. A few points are worth noting: 1. At a generic point in the moduli space $`\mathrm{}`$ is orthogonal to no such elements of $`\mathrm{{\rm Y}}`$ and so the gauge group is $`\mathrm{U}(1)^{2+n}`$ as expected. 2. This rule is completely derivable from classical geometry for the case that the roots are in $`L\mathrm{{\rm Y}}`$, where $`L`$ is the root lattice of $`\mathrm{G}_E`$. Picking up roots in the rest of $`\mathrm{{\rm Y}}`$ is a stringy effect — the analogue of the $`\mathrm{SU}(2)`$ gauge symmetry one sees on a circle of self-dual radius (see for example). 3. The maximal rank of the semi-simple part of the observed gauge group is $`n`$. There are always at least two $`\mathrm{U}(1)`$ factors which are not enhanced to nonabelian groups. This is because the GSO projection of the supersymmetric half of the heterotic string projects out the would-be vector bosons which would like to enhance these gauge group factors. 4. This Grassmannian picture for the moduli space is only true to leading order. We can expect quantum corrections to break anything — including the nonabelian enhanced gauge symmetry. Now we would like to map this picture of gauge symmetry enhancement back into the language of the type IIA string compactified on $`X`$. What do we need to do to $`X`$ to get an enhanced gauge symmetry? This is explained in great detail in . First of all note that the factor (49) of the moduli space corresponds exactly to the Kähler form parameters of “type II” above. We know this because of the special Kähler geometry discussed in section 3.3.1 and the intersection numbers discussed in section 3.3.2. This means that moving around in this Narain component of $`\mathrm{M}_V`$ corresponds to changing the size (and $`B`$-field) of the algebraic curves in the generic K3 fibres of $`X`$. The result is ###### Proposition 12 Let a set of algebraic genus zero curves collapse to zero area in every K3 fiber in $`X`$. Thus $`X`$ acquires a curve of singularities. In addition set the corresponding components of the $`B`$-field to zero. Then one obtains a nonabelian enhanced gauge symmetry. The ADE classification of curves one may collapse in a K3 surface corresponds to the ADE classification of the resulting Lie gauge groups. Again we need to note a few points: 1. We are assuming that there is no monodromy in these curves in the K3 fibres as we move around the base $`W`$. If there is monodromy one can obtain non-simply-laced gauge symmetries which we do not wish to discuss here. 2. We also assume that the overall volume of each K3 fiber is generic. By tuning the volume to the right values one may enhance the gauge symmetry further. One usual way of picturing the appearance of a nonabelian gauge symmetry is as follows. The type IIA string theory contains 2-branes in its spectrum (as discussed in many other lectures at this school). These 2-branes may be “wrapped” around the 2-spheres living in the K3 fibres. The mass of the resulting solitons in the four-dimensional theory is given by the area (and $`B`$-field) of these 2-spheres. In the limit that these spheres shrink to zero size we obtain new massless states in the theory. These massless states may lie in either hypermultiplets or vector multiplets. Which type is determined by the moduli space of the 2-cycle that shrank down to zero size. Witten showed that isolated curves give rise to hypermultiplets and curves that live in families parametrized by other curves give vectors. Thus, in our case where we are shrinking down whole families of curves in order to obtain a Calabi–Yau threefold with a singular curve we expect extra vectors. These vectors are the “W-bosons” which enhance the gauge group to a nonabelian group. The case we have considered here is actually a special case of acquiring a singular curve in $`X`$ and so must be considered to be a special case of acquiring nonabelian gauge symmetry. Consider the projection given by the K3-fibration $`\pi :X_1W`$ when $`X_1`$ is a singular space made by shrinking down a particular curve (or set of curves) within every K3 fibre. Let $`C_{\text{sing}}X_1`$ be the resulting singular curve within $`X_1`$. The restriction of the fibration $$\pi |_{C_{\text{sing}}}:C_{\text{sing}}W$$ (54) is an isomorphism. Suppose that we can find another family of curves within $`X`$ which can be shrunk down to form another singular space $`X_2`$ with a singular set $`C_{\text{sing}}^{}X_2`$. The projection under $`\pi `$ of a general singular set may or may not be surjective onto $`W`$. In particular we may have that the image under $`\pi `$ is a point (or a set of points) in $`W`$. It is not hard to see that the fibre over such a point in $`W`$ is peculiar and could not possibly be a smooth K3 fibre. Indeed we are talking about contributions of “type III” to the moduli space of vector multiplets when we shrink such 2-cycles down. We therefore claim that a singular curve lying over a point in $`W`$ must correspond to a nonperturbative enhanced gauge group. We will see examples of nonperturbative gauge groups in section 4.3. #### 3.3.4 An example Now that we have spoken rather abstractly about duality let us give an example which illustrates most of what we have discussed above. This example first appeared in . We begin by describing the Calabi–Yau threefold $`X`$ on which we will compactify the type IIA string. Let $`X`$ be the hypersurface $$x_0^2+x_1^3+x_2^{12}+x_3^{24}+x_4^{24}=0$$ (55) in the weighted projective space $`_{\{12,8,2,1,1\}}^4`$ with homogeneous coordinates $$[x_0,x_1,x_2,x_3,x_4][\lambda ^{12}x_0,\lambda ^8x_1,\lambda ^2x_2,\lambda x_3,\lambda x_4].$$ (56) Note that this satisfies the Calabi–Yau condition (40). We also need to note that this Calabi–Yau threefold is not smooth. In particular, putting $`\lambda =i`$ we obtain $$[x_0,x_1,x_2,x_3,x_4][x_0,x_1,x_2,ix_3,ix_4].$$ (57) which produces a $`_4`$ singularity at $`[x_0,x_1,0,0,0]`$ which is a single point in $`X`$. Similarly putting $`\lambda =1`$ puts a $`_2`$ singularity along $`[x_0,x_1,x_2,0,0]`$, which is a curve in $`X`$ (containing the previous $`_4`$ fixed point). These quotient singularities need to be blown up if we want a nice smooth Calabi–Yau threefold for $`X`$. For the singular curve in $`X`$ fixed by $`_2`$ we may replace each point in this curve by a $`^1`$. The homogeneous coordinates of this $`^1`$ may be considered to be $`[x_3,x_4]`$ (which are not now allowed to vanish simultaneously — we have removed the singularity after all!). Actually we may view $`[x_3,x_4]`$ as the coordinates of $`W^1`$ and project in the obvious way $$\pi :XW.$$ (58) Let us denote a given point on $`W`$ by $`\mu `$. That is, let $`x_4=\mu x_3`$. Then the inverse image of a point in $`W`$ under $`\pi `$ is $$x_0^2+x_1^3+x_2^{12}+x_3^{24}(1+\mu ^{24})=0$$ (59) in the weighted projective space $`_{\{12,8,2,1,\}}^3`$. This is a K3 surface as required. This is most easily seen by putting $`x_3^{}=x_3^2`$ giving us an equation in $`_{\{6,4,1,1\}}^3`$. Thus we have written $`X`$ as a K3-fibration. We may now play the same trick again on each K3 fibre. Each K3 fibre has a $`_2`$ singularity in it (as a side effect of the $`_4`$ singularity in the original threefold). This may be resolved by replacing it with a $`^1`$ which we denote $`C`$. Thus the fibre itself may be written as a bundle over $`C^1`$ with fibre given by a cubic equation in $`_{\{3,2,1\}}^2`$ — namely an elliptic curve. Thus our final smooth $`X`$ consists of a K3-fibration over $`W^1`$ where each K3 fibre is itself an elliptic fibration over another $`C^1`$. All these fibrations have sections allowing us to identify $`W`$ and $`C`$ as the bases of fibrations with subspaces of $`X`$. Now we may describe $`H^2(X)`$, or equivalently $`H_4(X)`$, in terms of this K3 fibration in the language of section 3.3.2. 1. We have the size of the section $`W`$. This gives one vector multiplet. 2. We may vary the sizes of the section $`C`$ of each K3 fibre and we may vary the size of each elliptic fibre of these K3’s. This gives two more vector multiplets. 3. The only bad K3 fibres occur where $`\mu ^{24}=1`$ in (59). The resulting polynomial does not factorize and so this bad fibre is still irreducible. Therefore we obtain no more vector multiplets associated with bad fibres. So we have a theory with three vector multiplets (indeed, $`h^{1,1}(X)=3`$). We may now write down the form of the moduli space to leading order using proposition 9. First we need the Picard lattice of the generic K3 fibre. There are two generators: the elliptic fibre, $`e`$ and the $`^1`$ section $`f`$. It is not hard to show that $`ee=0`$, $`ef=1`$, and $`ff=2`$. This intersection matrix is isomorphic to $`\mathrm{\Gamma }_{1,1}`$. Thus $`\mathrm{{\rm Y}}\mathrm{\Gamma }_{2,2}`$ and $`n=2`$ in (53). Let us try to find a heterotic string interpretation of this moduli space. Going back to the discussion around equation (49) we see that we have the simplest case where $`L`$, the Cartan lattice of $`\mathrm{G}_E`$, is empty and indeed the rank of $`\mathrm{G}_E`$ is $`n2=0`$. The vector moduli space is purely described in terms of deformations of the dilaton-axion and the Narain moduli space of the 2-torus $`E_H`$ with no bundle degrees of freedom. This accounts for all three vector moduli. In other words, all the the primordial gauge group in ten dimensions must have been sucked up with the bundle on the K3 surface $`S_H`$ leaving nothing left for $`E_H`$ to play with. To describe exactly what this bundle on $`S_H`$ is requires a knowledge of the hypermultiplet moduli space and so we won’t be able to discover this until section 4.3.1. We get enhanced gauge symmetries in the following ways. We may shrink down the section $`f`$ in every K3 fibre. The undoes the second blow-up we did when resolving at the start of this section. It produces a single curve of “$`A_1`$” singularities within $`X`$. It corresponds to putting the space-like 2-plane $`\mathrm{}`$ perpendicular to the single vector $`s\mathrm{{\rm Y}}`$. Either way, we get an $`\mathrm{SU}(2)`$ gauge symmetry. We may also squeeze out a rank 2 gauge symmetry — either $`\mathrm{SU}(2)\times \mathrm{SU}(2)`$ or $`\mathrm{SU}(3)`$ by tuning the vector moduli further. This can be seen by noting that $`A_1A_1`$ and $`A_2`$ can both be embedded in $`\mathrm{\Gamma }_{2,2}`$ and we may arrange $`\mathrm{}`$ to be orthogonal to either. This corresponds to shrinking the elliptic fiber, $`e`$, down to an area of order 1 as well as tuning the size of $`f`$. The precise details are given in . Now let us turn our attention to the type IIB picture. Using proposition 6 we see that $`Y`$ is given by $`X/(_6\times _{12})`$ where the generators of the quotienting group are given by $$\begin{array}{cc}\hfill g_1:[x_0,x_1,x_2,x_3,x_4]& [x_0,x_1,x_2,e^{\frac{2\pi i}{24}}x_3,e^{\frac{2\pi i}{24}}x_4]\hfill \\ \hfill g_2:[x_0,x_1,x_2,x_3,x_4]& [x_0,x_1,e^{\frac{2\pi i}{12}}x_2,x_3,e^{\frac{2\pi i}{12}}x_4].\hfill \end{array}$$ (60) The general form of $`Y`$ may be written as a quotient of $`X`$ with defining equation $$x_0^2+x_1^3+x_2^{12}+x_3^{24}+x_4^{24}+\alpha x_0x_1x_2x_3x_4+\beta x_2^6x_3^6x_4^6+\gamma x_3^{12}x_4^{12}=0.$$ (61) The three parameters $`\alpha `$, $`\beta `$ and $`\gamma `$ then give the three deformations of complex structure of $`Y`$ (as $`h^{2,1}=3`$). Knowing the details of the mirror map allows us to map these parameters to the complexified Kähler form of the type IIA description. One may determine this using the “monomial-divisor” mirror map of when one has a hypersurface in a weighted projective space. This particular model was also studied in . The upshot is that if $`X`$ is in the “Calabi–Yau phase” where the areas of all possible algebraic curves are large then essentially * Letting $`x=\beta /\alpha ^60`$ will take the size of the elliptic fibre off to infinity. * Letting $`y=4/\gamma ^20`$ sends the size of the section $`W`$ off to infinity. * Letting $`z=4\gamma /\beta ^20`$ sends the area of the rational curve $`f`$ within each K3 fibre off to infinity. The parameters $`(x,y,z)`$ are chosen so that the interior of the Kähler cone of $`X`$ is described asymptotically by $`x1`$, $`y1`$ and $`z1`$. Away from this limit these parameters can get mixed up and everything is less clear although well-understood. #### 3.3.5 Quantum corrections to $`N=2`$ gauge theories So far we have discussed purely the classical limit of the heterotic string theory where we assume the dilaton is such that the coupling is very weak and that the prepotential is purely cubic. Thanks to the duality of the heterotic string to the type IIB string we may try to continue our analysis of the heterotic string away from this classical limit. This is an enormous subject but we will be very brief here. Our intention is to give only a flavour of the subject. Let us explain what happens in terms of the example of the previous section. In particular let us study what happens to the would-be $`\mathrm{SU}(2)`$ gauge theory which appears when every K3 fibre of $`X`$ contains an $`A_1`$ singularity. First of all we mentioned that we could actually get the gauge group to be $`\mathrm{SU}(2)\times \mathrm{SU}(2)`$ or $`\mathrm{SU}(3)`$ if we tuned the size of the K3 fibre suitably. Let’s not concern ourselves with this fact here and let us instead assume that the parameter $`\alpha `$ (or equivalently $`x`$) in the last section is at any generic value. Now we can ask ourselves if anything interesting happens to $`Y`$ as we vary $`y`$ and $`z`$. In particular, the most obvious question to ask is whether $`Y`$ is ever singular. $`Y`$ is singular whenever $`f=f/x_0=\mathrm{}f/x_4=0`$ has a solution for (61). With a little algebra we find that this has a fairly simple solution for $`y=1`$. In this case we have 12 singular points in $`Y`$ lying in the subspace $`x_0=x_1=x_2=0`$. We know that varying $`y`$ has something to do with varying the dilaton in the heterotic string so this suggests that something curious happens in our model when the heterotic string coupling is of order 1. While this sounds interesting it is a bit too exotic for our purposes here! We would rather discover something interesting which happens near weak coupling. The next simplest solution one finds is when we have singular points in the larger subspace $`x_0=x_1=0`$. This demands that $$(1z)^2yz^2=0.$$ (62) If our heterotic string has zero coupling we set $`y=0`$ and so this has a solution when $`z=1`$. One may show that $`z=1`$ is exactly the value required to make the little curves $`f`$ in each K3 fibre of $`X`$ acquire zero size . So this must be exactly where we expect to see enhanced $`\mathrm{SU}(2)`$ gauge symmetry. To summarize we expect to see an $`\mathrm{SU}(2)`$ gauge symmetry whenever $`y=0`$ and $`z=1`$. Now we may probe into nonzero coupling by letting $`y`$ acquire a small nonzero value. The odd thing to note is that (62) then has two solutions for $`z`$ near 1. Somehow our single $`\mathrm{SU}(2)`$ theory has split into two interesting things for nonzero heterotic dilaton. At this point we could easily go off and explore the wonders of these quantum corrections. This subject is generally called “Seiberg–Witten” theory . These lecture notes would be dwarfed by a full treatment of this subject so instead we will refer to , for example, for a review. Here we will just review some basic properties. In its basic form Seiberg–Witten theory is not a theory which includes gravity. It is a very interesting question as to how one can remove gravity from the four-dimensional theory we have constructed. One might regard the removal of gravity as a rather regressive thing to do — after all it was precisely because string theory contains gravity that string theory became so popular in the first place. Nevertheless going to a limit where gravity can be ignored provides a very useful way of making contact between what is known about string theory and quantum field theory. Indeed this process has often dominated work in string theory in recent years. In order to switch off gravity we need to take the string coupling to zero. As we discussed in the type IIA language this corresponds to taking the area of the section $`W`$ to infinity. In type IIB language we are taking $`y0`$. If we were to do this process alone then everything would become rather trivial. Instead let us “zoom in” on the splitting effect that we saw above. In particular let us rescale $`z1`$ as we take $`y0`$ so that we fix the location of the two solutions of (62) at some fixed scale determined by a constant traditionally called $`\mathrm{\Lambda }^2`$. This leaves us with one complex parameter $`u`$, where the $`u=\pm \mathrm{\Lambda }^2`$ at the discriminant. We show this in figure 2. This scale $`\mathrm{\Lambda }`$ encodes the effective coupling constant of the gravity-free Yang–Mills theory which remains. This process is explained in detail in . In a way this limit is one of the cleanest ways of viewing the process of “dynamical scale generation” in quantum field theory. We desire to zoom in on the part of the moduli space where gravity is weakly coupled but the structure of the $`\mathrm{SU}(2)`$ gauge theory of interest forces us to fix a scale. This is the same scale which appears when computing the running of a coupling constant in an asymptotically free theory! The two main statements of Seiberg–Witten theory for $`\mathrm{SU}(2)`$ are 1. The gauge symmetry $`\mathrm{SU}(2)`$ never appears. It is broken by quantum effects (assuming $`\mathrm{\Lambda }`$ is nonzero). 2. At $`u=\pm \mathrm{\Lambda }^2`$ massless solitons appear. These are the remnants of the “W-bosons” which appeared classically to enhance the gauge symmetry. There is one aspect of this “zooming in” process which is of great interest when discussing the geometry of $`N=2`$ theories. Namely, the structure of special Kähler geometry changes. If one considers the geometry of the moduli space with no gravity then (28) becomes $$K=\mathrm{Im}\left(\overline{t}^i\frac{\mathrm{F}}{t^i}\right).$$ (63) where $`t^i`$ are affine coordinates. This form of special Kähler geometry is often referred to as “rigid” special Kähler geometry while that of section 3.1 is called “local” special Kähler geometry. The key point, as discussed in for example, is that while local special Kähler geometry is associated to the moduli space of complex Calabi–Yau threefolds, rigid special Kähler geometry is associated to the moduli space of complex curves. Thus we should expect the theory of $`N=2`$ supersymmetric field theories without gravity to be associated to Riemann surfaces in much the same way that these theories with gravity were associated to Calabi–Yau threefolds. This is pretty much exactly what Seiberg–Witten theory does. An $`\mathrm{SU}(2)`$ gauge theory is associated with an elliptic curve for example. The exact way in which this curve appears in the limit of the Calabi–Yau threefold as we decouple gravity is not at all clear. A fairly systematic way of doing this construction was explained in in the case that $`Y`$ is constructed using toric geometry. See also for an earlier analysis of this problem and , for example, for further discussion. The geometry of the Calabi–Yau threefold makes an explicit appearance for $`N=2`$ theories with gravity — it is the Calabi–Yau threefold $`Y`$ on which the type IIB string is compactified. The manifest geometry of the Riemann surface in the case of Seiberg–Witten theory is a little more obscure. Possibly the best suggestion for a direct picture in which this curve appears was given by Witten in terms of M-theory and world-volume theories of D-branes. #### 3.3.6 Breaking T-Duality Our discussion of the moduli space of the type IIA picture and the heterotic picture for $`\mathrm{M}_V`$ were in excellent agreement so long as we ignored quantum effects. In both cases we had a “Narain” factor in the form of the symmetric space given in (49). In the language of the heterotic string this consisted of the moduli of the 2-torus $`E_H`$ together with the degrees of freedom of the Wilson lines of the flat bundle $`V_E`$. The group $`\mathrm{O}(\mathrm{{\rm Y}})`$ gave the T-dualities of the heterotic string on a torus. In particular if we consider the example of section 3.3.4 then we have a Narain factor of the form $$\mathrm{O}(\mathrm{\Gamma }_{2,2})\backslash \mathrm{O}(2,2)/(\mathrm{O}(2)\times \mathrm{O}(2))(C_m\times C_c)\backslash \left(\frac{𝖧_\sigma }{\mathrm{SL}(2,)}\times \frac{𝖧_\tau }{\mathrm{SL}(2,)}\right).$$ (64) Here we have used the standard decomposition of the Grassmannian into a form which makes it more recognizable for our purposes. We have two copies of the upper half-plane $`𝖧\mathrm{SL}(2,)/\mathrm{U}(1)`$ which we parameterize by complex numbers $`\sigma `$ and $`\tau `$ respectively. The groups $`C_m`$ and $`C_c`$ are both isomorphic to $`_2`$ and are generated by $$\begin{array}{cc}\hfill g_m:(\tau ,\sigma )& (\sigma ,\tau )\hfill \\ \hfill g_c:(\tau ,\sigma )& (\overline{\tau },\overline{\sigma }),\hfill \end{array}$$ (65) respectively. We refer to , for example, for details of this isomorphism. We depict this moduli space in figure 3. The interpretation of this moduli space in terms of $`E_H`$ is straight-forward. We let $`\tau `$ denote the complex structure in the standard way and we let $`\sigma `$ denote the single component of $`B+iJ`$. Thus the $`\mathrm{SL}(2,)`$ action on $`\tau `$ is the standard modular invariance of a 2-torus. The $`\mathrm{SL}(2,)`$ acting on $`\sigma `$ is composed of the familiar $`BB+1`$ symmetry as well as a $`J1/J`$ T-duality. Note that $`C_m`$ is “mirror symmetry” for a 2-torus as was first seen in . $`C_c`$ can be thought of as a complex conjugation symmetry of the theory. This is all very well but we have noticed in the previous section that this picture of the moduli space is subject to quantum corrections. That is, this Narain picture of the moduli space of $`E_H`$ is not exact. We will now argue that the effect of these quantum corrections is to completely ruin the description of the moduli space as a quotient and so any notion of T-duality for $`E_H`$ is lost. To argue this let us discuss what can go wrong with T-duality arguments in another example. We will consider the classical quintic hypersurface in $`^4`$ as was analyzed in . The Calabi–Yau manifold has $`h^{1,1}=1`$. When we flatten out the single complex coordinate describing $`B+iJ`$ we obtain the moduli space depicted in figure 4. Now although this moduli space looks similar to the fundamental region of $`𝖧/\mathrm{SL}(2,)`$ there is a big difference. There is no action of any discrete group on $`𝖧`$ for which the region in figure 4 is a fundamental region. One may see this as follows. Note that there is an angle of $`2\pi /5`$ formed at the lowest point in this region. One should therefore need 5 fundamental regions touching at this point. Indeed one may find such regions and they are pictured in figure 5.2 of . One can also see that the $`BB+1`$ symmetry should allow us to translate these fundamental regions one unit to the left or right to form new fundamental regions. The problem is that doing this translation gives a region which overlaps in an open set with one the regions we built earlier by rotating by $`2\pi /5`$. Thus these supposed fundamental regions do not tessellate in $`𝖧`$ and therefore cannot be derived in terms of a group action on $`𝖧`$. Indeed we argued in section 2 that $`N=2`$ theories in four dimensions do not generically have locally symmetric moduli space. It was in the context of symmetric spaces that we saw the natural appearance of T-duality. It should not therefore be a surprise that we do not find the true analogue of a modular group for the quintic threefold. We should therefore expect that the quintic threefold represents the generic case of a Calabi–Yau moduli space. In particular once we turn on the heterotic string coupling, i.e., give finite size to the section $`W`$ of the example in section 3.3.4 the Narain description of the moduli space is lost. This is argued in . So if the heterotic string on a torus does not respect T-duality how should we really describe the moduli space? The principle should be the same as that for the quintic threefold. One should begin with a weakly coupled heterotic string on a large circle or torus. Here one unambiguously sees the geometry of the compactification. Now move about the moduli space of compactifications. In this we can label every point in the moduli space by a set of moduli (such as radii) for the torus. One problem we have to be careful about is that we may follow loops in the moduli space which allow us to identify more than one torus with a given point in the moduli space of theories. We must avoid this by putting cuts in the moduli space. If we do not put in such cuts then generically one would expect to be able to identify every possible torus with each point in the moduli space! (Note that since the classical $`\mathrm{SL}(d,)`$ symmetry of a $`d`$-torus is lost one must describe the torus directly in terms of data which chooses a fundamental region of the classical moduli space of flat metrics.) Once we have completed this labelling process (the details of which depend on a choice of cuts) we have defined every possible torus to be considered. Tori excluded by the process, such a circle of radius less that $`\sqrt{\alpha ^{}}`$, do not exist and should not be considered. It is only the accidental T-duality of the weakly-coupled string that led us to believe that we could make real sense of small tori. * This example consisted of a moduli space $`\mathrm{M}_V`$ which became locally a symmetric space on its boundary at infinite distance corresponding to some classical limit. It is interesting to note that there are other known examples where a subspace of $`\mathrm{M}_V`$ can be locally symmetric. For example consider the so-called Z-orbifold $`T^6/_3`$ with 27 fixed points. The rational curves in this space (after blowing up) conspire to only give certain quantum corrections to $`\mathrm{M}_V`$. The effect of this is to make the prepotential $`\mathrm{F}`$ exactly cubic if none of the 27 blow-up modes are switched on . The result is that we get a slice of the moduli space (at finite distance) of the form $$\mathrm{M}_{V,\mathrm{orb}}\mathrm{U}(3,3;)\backslash \mathrm{U}(3,3)/\left(\mathrm{U}\left(3\right)\times \mathrm{U}\left(3\right)\right).$$ (66) Moving away from this subspace there are instanton corrections and the symmetric space structure is lost. Note that in general we lose the classical $`\mathrm{SL}(2,)`$ symmetry of the complex structure of the torus in addition to any T-duality. How can this be? The moduli space of a 2-torus of volume one is determined by considering the ways of making a a lattice of area one, dividing out by rotations, and then dividing out by the modular group $`\mathrm{SL}(2,)`$. This gives us the familiar form $`\mathrm{SL}(2,)\backslash \mathrm{SL}(2,)/\mathrm{SO}(2)`$. If we declare that quantum effects break this structure then quantum effects must be having a drastic effect on this construction of the torus. As well as breaking the $`\mathrm{SL}(2,)`$ invariance, we are also modifying the $`\mathrm{SL}(2,)/\mathrm{SO}(2)`$ part. It is as if we are breaking the picture of the 2-torus as a Riemannian manifold. Hopefully once stringy geometry is better-understood it will be more clear what is happening here. It is worth mentioning that there are two distinct types of U-dualities discussed in the literature. One is an “internal duality” statement where one says that a string theory of type $`\mathrm{S}_1`$ (e.g., type IIA, $`E_8\times E_8`$ heterotic etc.) compactified on $`X_1`$ with coupling $`\lambda _1`$ is dual to a string theory of the same type $`\mathrm{S}_1`$ compactified on $`X_2`$ with coupling $`\lambda _2`$. Alternatively one has an “external duality” where one says that a string theory of type $`\mathrm{S}_1`$ compactified on $`X_1`$ with coupling $`\lambda _1`$ is dual to a string theory of a different type $`\mathrm{S}_2`$ compactified on $`X_2`$ with coupling $`\lambda _2`$. Our discussion of the breaking of T-dualities (and by implication U-dualities) was in the context of internal dualities. In particular we were fixing our string as an $`E_8\times E_8`$ heterotic string. What happens when the external duality relating an $`E_8\times E_8`$ heterotic string on a given torus and a given choice of Wilson lines to a $`\mathrm{Spin}(32)/_2`$ heterotic string on another torus and set of Wilson lines? Our discussion of mapping out the moduli space should apply again. Map out the moduli space of tori and Wilson lines as above using the $`E_8\times E_8`$ heterotic string. Now do the same thing with the $`\mathrm{Spin}(32)/_2`$ heterotic string. Note that the starting point for the large torus will not be the same limit point in moduli space as the former case. This means that every point in the moduli space will now have two labels — one for each heterotic string. One should not obtain small radii for either heterotic string interpretation. Thus strictly external U-dualities need not be broken by quantum effects. The precise mapping between $`(X_1,\lambda _1)`$ and $`(X_2,\lambda _2)`$ can be expected to be modified however. ## 4 The Moduli Space of Hypermultiplets Now we come to the considerably more tricky subject of trying to map out the moduli space of hypermultiplets for our $`N=2`$ theories in four dimensions. In the case of the vector multiplet moduli space, the type IIB string compactified on $`Y`$ gave an exact model. For the hypermultiplets there is no exact model. This makes the subject much more difficult and potentially much more interesting! ### 4.1 Related Dimensions The purpose of these lectures is to discuss some special properties of $`N=2`$ theories in four dimensions. It turns out to be very useful to be aware of some other closely-related theories in both higher and lower dimensions than four to help gain insight into the hypermultiplet moduli space. #### 4.1.1 $`N=(1,0)`$ in six dimensions Imagine compactifying the heterotic string on a K3 surface $`S_H`$. This would yield a theory with $`N=(1,0)`$ supersymmetry in six dimensions. We refer the reader to for a good discussion of many aspects of such theories. This has an $`R`$-symmetry of $`\mathrm{Sp}(1)`$. We discussed the supermultiplets of such theories in section 2.3. Such a theory may then be compactified on a 2-torus to yield our familiar $`N=2`$ theory in four dimensions. Upon dimensional reduction, the $`N=(1,0)`$ supermultiplets in six dimensions become $`N=2`$ supermultiplets in four dimensions as follows: * A six-dimensional hypermultiplet becomes a hypermultiplet in four dimensions. * A six-dimensional vector multiplet becomes a vector multiplet in four dimensions. * A six-dimensional tensor multiplet becomes a vector multiplet in four dimensions. In particular the hypermultiplet moduli space of a heterotic string compactified on a K3 surface $`S_H`$ is exactly the same as the hypermultiplet moduli space of a heterotic string compactified on $`S_H\times E_H`$. This is consistent with our earlier comment that all the hypermultiplet information comes from the K3 surface $`S_H`$. It is therefore quite common to analyze the hypermultiplet moduli space in terms of six-dimensional physics rather than four-dimensional physics. Having said that, our duality statements might now sound a bit peculiar. We want to say something to the effect that we can model the hypermultiplet moduli space of a heterotic string on a K3 surface in terms of a type IIA string on a Calabi–Yau threefold $`X`$ but the former is six-dimensional while the latter is four-dimensional. It is important to note that we cannot necessarily completely ignore the 2-torus $`E_H`$ in the product $`S_H\times E_H`$. In effect we can think of arriving at our six-dimensional theory by beginning in four dimensions and decompactifying $`E_H`$. To do this we certainly need the full moduli space of $`E_H`$ and from proposition 10 this in turn implies that the Calabi–Yau threefold $`X`$ is an elliptic fibration with a section. Assuming this is the case, we may model the six-dimensional physics of the heterotic string on $`S_H`$ in terms of the type IIA string on $`X`$ by implicitly decompactifying $`E_H`$. This mechanism of using type IIA strings on $`X`$ to model six-dimension physics is known as “F-theory”. The reader should be warned that there are at least two other ways of defining F-theory common in the literature. One is to treat F-theory as twelve dimensional (although whether it lives in $`^{2,10}`$ or $`^{1,11}`$ is unclear). Another way is to view it as a type IIB string compactification with a varying dilaton. We refer to for more details. The type IIA definition of F-theory is well-suited for our purposes of linking the subject to four dimensions. Let us denote the elliptic fibration as $`p:X\mathrm{\Theta }`$, where $`\mathrm{\Theta }`$ is a complex surface. We also know we have a K3-fibration $`\pi :XW`$, where $`W^1`$, and a fibration $`\mathrm{\Theta }W`$ with generic fibre given by $`^1`$. That is, $`\mathrm{\Theta }`$ is a “ruled surface”. If $`\mathrm{\Theta }`$ is a smooth $`^1`$-bundle over $`W`$, it is the “Hirzebruch surface” $`𝐅_n`$. Here the section $`W\mathrm{\Theta }`$ has self-intersection $`n`$ within $`\mathrm{\Theta }`$. Blowing up $`𝐅_n`$ at a few points replaces some of the smooth $`^1`$-fibres by chains of $`^1`$’s. It is common to then draw $`X`$ (representing a complex dimension as a real dimension) in the following form. We may use the plane of the paper to represent $`\mathrm{\Theta }`$ by letting the horizontal direction represent the section and the vertical direction represent the $`^1`$-fibre. That is, the “ruling” of the ruled surface $`\mathrm{\Theta }`$ is given by vertical lines. Now over a (complex) codimension one subspace of $`\mathrm{\Theta }`$ the elliptic fibration $`p:X\mathrm{\Theta }`$ will degenerate. We may draw this “discriminant” locus as a set of curves and lines in the plane of the paper. Kodaira has classified the possibilities for how an elliptic fibre may degenerate in the case of one parameter family . We show the possibilities in figure 5. With the exception of $`\mathrm{I}_0`$ which is the smooth elliptic case, and $`\mathrm{II}`$ which is an elliptic curve with a cusp, each line in the figure represents a rational curve. This curve may appear with a multiplicity given by the small numbers in the figure. This classification can be used to label the generic points on the irreducible components of the discriminant locus. The result is that one obtains a picture somewhat typically like figure 6 for $`X`$ in the form of an elliptic fibration. In this figure the dotted lines represent lines ($`^1`$’s) within $`\mathrm{\Theta }`$. $`C_0`$ is a section and $`f`$ is a generic $`^1`$ fibre. Note that this notation is consistent with the $`f`$ which appeared in section 3.3.4. At one point over $`W`$ we have put a fibre as a chain of three $`^1`$’s. The solid lines represent the discriminant locus. Each irreducible component is labelled by its Kodaira type. When these components collide, the elliptic fibration will degenerate further and the resulting fibre need not lie in Kodaira’s classification. Since we wish to study the moduli space of hypermultiplets we are particularly interested in the deformations of complex structure of $`X`$. When we draw $`X`$ as an elliptic fibration, the complex structure is encoded in the discriminant locus. Thus, deformations of $`X`$ are given simply by the deformations of the discriminant locus. It will also be worthwhile to note how the Kähler form data appears in the elliptic fibration. Deforming the Kähler form may either affect areas in the fibre direction (i.e., the area of the generic fibre as well as areas within the chains of special Kodaira fibres) or affect areas within $`\mathrm{\Theta }`$. As we decompactify $`E_H`$ to go from 4 dimensions to 6 dimensions one can show that the areas in the fibre direction become meaningless . Our discussion of the types of supermultiplets in four and six dimensions given at the start of this section leads one to conclude: * Using the Kähler form to vary areas in the fibre direction corresponds to moduli in a six-dimensional vector supermultiplet. * Using the Kähler form to vary areas in $`\mathrm{\Theta }`$ corresponds to moduli in a six-dimensional tensor supermultiplet. #### 4.1.2 $`N=4`$ in three dimensions Imagine taking our $`N=2`$ theory in four dimension and compactifying further on a circle. This leads to a theory in three dimensions with $`N=4`$ supersymmetry. This has an $`R`$-symmetry of $`\mathrm{SO}(4)\mathrm{Sp}(1)\times \mathrm{Sp}(1)`$ (up to irrelevant discrete factors) which implies that the moduli space should factorize into a product of two quaternionic Kähler spaces. These three dimensional theories have two different types of “hypermultiplets” whose moduli spaces cannot mix. In the literature one often refers to one of these types of hypermultiplets as “vector multiplets” to reflect their four-dimensional origin. However, one should be aware that, within the context of the three-dimensional physics, such a distinction is arbitrary. Note that the hypermultiplet moduli space $`\mathrm{M}_H`$ from four dimensions comes through unscathed into the three dimensional picture whereas our vector multiplet moduli space becomes “quaternionified” in the compactification. It is remarkable how resilient $`\mathrm{M}_H`$ is! It is unchanged as we compactify on circles a theory in six dimensions with $`N=(1,0)`$ supersymmetry down to three dimensions. Compare this with the capricious vector multiplet moduli space which is non-existent in 6 dimensions, real in 5 dimensions, complex in 4 dimensions and quaternionic in 3 dimensions! Because the vector multiplet moduli space becomes a hypermultiplet moduli space upon compactification to three dimensions, this picture provides a potentially useful way of using our knowledge of special Kähler manifolds to uncover some of the mysteries of quaternionic Kähler manifolds. Suppose we wish to study $`\mathrm{M}_H(X)`$ for a type IIA string compactified on the Calabi–Yau threefold $`X`$. Consider instead the moduli space $`\mathrm{M}_V(Y)`$ of the type IIA string compactified on $`Y`$, the mirror of $`X`$. Compactifying further on $`S_R^1`$, a circle of radius $`R`$, the special complex Kähler space $`\mathrm{M}_V(Y)`$ becomes a quaternionic Kähler space which we will denote $`\mathrm{M}_V(Y)_{}`$. Since the type IIA string on a circle of radius $`R`$ is supposedly T-dual to the type IIB string on a circle of radius $`1/R`$, the type IIA string on $`Y\times S_R^1`$ should be dual to the type IIB string on $`Y\times S_{1/R}^1`$. Using mirror symmetry this is then dual to the type IIA string on $`X\times S_{1/R}^1`$. The space $`\mathrm{M}_V(Y)_{}`$ must now represent the factor of the moduli space containing deformations of complex structures of $`X`$. That is, it descended from $`\mathrm{M}_H(X)`$ upon compactification on the circle. Having said that, $`\mathrm{M}_H(X)`$ is unchanged by this circle compactification and so $$\mathrm{M}_H(X)\mathrm{M}_V(Y)_{}.$$ (67) * We already questioned the validity of T-duality for the heterotic string in section 3.3.6. It is natural to question whether T-duality is valid for the type II strings when we have only modestly extended supersymmetry. The crude statement that the type IIA string compactified on $`Y\times S_R^1`$ is dual to the type IIB string compactified on $`Y\times S_{1/R}^1`$ is almost certainly incorrect. It is true however that one should expect this to be exact when the strings are very weakly coupled. Most analyses of strings using this statement of T-duality such as do use only weakly-coupled strings. We will not try to elucidate the exact meaning of T-duality in type II strings in these lectures. Determining $`\mathrm{M}_V(Y)_{}`$ from the complex space $`\mathrm{M}_V(Y)`$ is not easy. An interesting attempt at this problem was made some time ago by Cecotti et al. in .<sup>15</sup><sup>15</sup>15see also for further analysis along these lines. This paper assumed that the moduli space $`\mathrm{M}_V`$ was determined by a prepotential that was exactly cubic. Particular attention was paid to the cases where $`\mathrm{M}_V`$ is a symmetric space. If one then ignored quantum corrections upon compactification on a circle, this symmetric space was mapped via the so-called “$`c`$-map” to another symmetric space. For example one might have something like<sup>16</sup><sup>16</sup>16The map $`c`$ is not intended to be viewed as a map of topological spaces! We are replacing one space by another. $$c:\frac{\mathrm{SU}(3,3)}{\mathrm{S}(\mathrm{U}(3)\times \mathrm{U}(3))}\frac{E_{6(+2)}}{\mathrm{SU}(2)\times \mathrm{SU}(6)}.$$ (68) A notable case of the $`c`$-map is $$c:\frac{\mathrm{SL}(2,)}{\mathrm{U}(1)}\times \frac{\mathrm{SO}_0(2,n2)}{\mathrm{SO}(2)\times \mathrm{SO}(n2)}\frac{\mathrm{SO}_0(4,n)}{\mathrm{SO}(4)\times \mathrm{SO}(n)}.$$ (69) We will revisit this briefly in section 4.4.3. Of course, unless we pick a very special model to examine<sup>17</sup><sup>17</sup>17See for such an example., there will be quantum corrections and the analysis of will not be directly applicable. However, this method may provide a good starting point for the analysis of the quaternionic Kähler moduli spaces as it does give the asymptotic behaviour where quantum effects can be neglected. An exact version of the $`c`$-map was elucidated by Seiberg and Witten in the case of rigid special Kähler geometry. As discussed in section 3.3.5, $`\mathrm{M}_V(Y)`$ is described in this limit by the deformation of a complex curve $`C_{\mathrm{SW}}`$. Seiberg and Witten’s remarkably simple result is then ###### Proposition 13 In the case that $`\mathrm{M}_Y(V)`$ is a rigid special Kähler space, $`\mathrm{M}_V(Y)_{}`$ is simply the hyperkähler space given by an abelian (i.e., complex algebraic torus) fibration over $`\mathrm{M}_V(Y)`$ where the fibre is given by the Jacobian $`H^1(C_{\mathrm{SW}},\mathrm{U}(1))`$. In addition the volume of the fibre is determined by $`R`$, the radius of the circle on which one compactifies. ### 4.2 Extremal Transitions Since direct analysis of the hypermultiplet moduli space is so formidable the most prudent course of action is to try to squeeze as much information out of our knowledge of the vector multiplet moduli space as we possible can. This is facilitated by the occurrence of phase transitions or “extremal transitions”. We go to a funny point in moduli space where vector moduli disappear and new hypermultiplet moduli appear. We may then pretend that we actually did this process in reverse and claim that we know something about what happens when we move around in the moduli space of hypermultiplets! #### 4.2.1 Conifolds Let us consider the simplest type of extremal transition first — the “conifold” of . We may understand this both from the point of view of geometry and from the point of view of field theory as explained in . We begin with the geometrical picture. Consider the type IIB string compactified on the Calabi–Yau manifold $`Y`$. We move about the moduli space of vector multiplets by deforming the complex structure of $`Y`$. Let us consider a one-dimensional family of such $`Y`$’s and denote an element of this family by $`Y_t`$ where $`t`$ parameterizes the family. At a special point in this part of the moduli space, say $`t=0`$, $`Y`$ may become singular. The simplest thing that can happen as $`t0`$ is that an $`S^3`$ can contract to a point. Locally such a singularity would look like the hypersurface $$w^2+x^2+y^2+z^2=0,$$ (70) in $`^4`$. This is called a “conifold singularity”. Locally such a conifold point can be resolved by replacing the point by a $`^1`$ (see, for example, for a nice explanation of this). Since the Kähler form controls the areas of $`^1`$’s such a resolution might be pictured as a deformation of Kähler form. In other words we have turned a degree of freedom from a deformation of complex structure into a deformation of Kähler form. Globally this picture does not work quite this simply. We need to consider the case of $`P`$ disjoint $`S^3`$’s, each shrinking to a point at $`t=0`$. If $`Y_t`$ represents the smooth $`Y`$ for a generic value of $`t`$ then a simple application of the Mayer-Vietoris sequence gives a relationship between the homology of $`Y_t`$ and the homology of $`Y_0`$. See for a full description of this process. Now resolve the resulting $`P`$ conifold points by adding $`^1`$’s and call the resulting smooth manifold $`Y^{}`$. Another application of the Mayer-Vietoris sequence gives a relationship between the homology of $`Y_0`$ and the homology of $`Y^{}`$. Combining these results we obtain $$\begin{array}{cc}\hfill 0H_4(Y_t)H_4(Y^{})\stackrel{f_1}{}& ^PH_3(Y_t)H_3(Y_0)0\hfill \\ \hfill 0H_3(Y^{})H_3(Y_0)& ^P\stackrel{f_2}{}H_2(Y^{})H_2(Y_t)0.\hfill \end{array}$$ (71) Let us denote by $`Q`$ the rank of the map labelled $`f_1`$. By Poincaré duality the rank of $`f_2`$ must also then be $`Q`$. Note that $`Q`$ represents the dimension of the kernel of the map $`^PH_3(Y_t)`$, i.e., the number of homology relations between the $`P`$ 3-spheres in the smooth $`Y`$. The above exact sequences give $$\begin{array}{cc}\hfill b_2(Y^{})& =b_2(Y_t)+Q\hfill \\ \hfill b_3(Y^{})& =b_3(Y_t)2(PQ).\hfill \end{array}$$ (72) That is, as we go through the conifold transition, we lose $`PQ`$ vector multiplets and gain $`Q`$ hypermultiplets. Note that $`P>1`$ is required for this transition to make sense and so a single conifold point is not sufficient. From the point of view of field theory this process is a supersymmetric variant of the Higgs mechanism. As we wander about the moduli space of vector multiplets it is possible that some hypermultiplets suddenly become massless. Indeed, Strominger noted that the singularities in the moduli space metric associated to a conifold are exactly the same as seen by Seiberg and Witten when a hypermultiplet becomes massless. Suppose $`P`$ hypermultiplets become massless and that these hypermultiplets are charged under $`PQ`$ of the $`\mathrm{U}(1)`$ gauge symmetries in our original theory. We may try to give these new hypermultiplets vacuum expectation values which would then spontaneously break this $`\mathrm{U}(1)^{PQ}`$ gauge symmetry. Our $`N=2`$ gauge theory in four dimensions has the standard gauge theory couplings and so these broken gauge symmetries must “eat up” some Goldstone bosons in order to become massive. What’s more they must do this in a way consistent with $`N=2`$ supersymmetry. The only way this can happen is for us to lose $`PQ`$ of our $`P`$ new massless hypermultiplets leaving us with $`Q`$ new hypermultiplets. This is the field theory picture for losing $`PQ`$ vector multiplets and gaining $`Q`$ hypermultiplets. Since we obtain $`Y^{}`$ via the Higgs mechanism, this is often referred to as the “Higgs phase”. Since $`Y_t`$ has more $`\mathrm{U}(1)`$’s (massless photons) it is referred to as the “Coulomb phase”. That is, the Higgs phase is the one with more hypermultiplets and the Coulomb phase is the one with more vector multiplets. The conifold transition is just the simplest example of all kinds of extremal transitions which may occur. #### 4.2.2 Enhanced gauge symmetry The Higgs phase transition of the preceding section was a little boring because there was no nonabelian gauge symmetry at the phase transition point. We know how to get enhanced gauge symmetry (at least in some limit) from section 3.3.3. We need to consider the type IIA string compactified on $`X`$, where $`X`$ has a curve of ADE singularities. This is easy to arrange using the elliptic fibration language of section 4.1.1. We can describe the situation using the “Weierstrass form” of the elliptic fibration which is standard when discussing F-theory. Let $`s`$ and $`t`$ be affine complex coordinates on some patch of the base $`\mathrm{\Theta }`$. We may then write the elliptic fibration as $$y^2=x^3+a(s,t)x+b(s,t).$$ (73) The discriminant is then given by $`\mathrm{\Delta }=4a^3+27b^2`$. The geometry of such fibrations was discussed in detail in and so we will be brief here. Let us assume $`a`$ and $`b`$ are independent of $`t`$ for the time being. We wish to put a line of interesting fibres along $`s=0`$. Table 3 lists the resulting fibres where $`as^L`$, $`bs^K`$ and $`\mathrm{\Delta }s^N`$ near $`s=0`$. The final column denotes the resulting singularity if all the components of the fibre not intersecting the section are shrunk down to zero area. Note that the fibres $`\mathrm{I}_0`$, $`\mathrm{I}_1`$ and II only have one component and thus cannot produce a singularity. This results in an explicit description of an extremal transition involving nonabelian gauge symmetry. Begin with a type IIA string on a smooth Calabi–Yau threefold $`X`$ where all the components of all the fibres have nonzero area. Now shrink down all the components of the fibres which do not hit the section. This will result in curves of ADE singularities producing some gauge group $`\mathrm{G}`$. We may then be free to deform the discriminant by a deformation of complex structure to smooth the threefold. Let us recast this transition in terms of the language of a heterotic string compactified on $`S_H\times E_H`$. The process begins by a deformation of the Kähler form of $`X`$ which is thus a deformation of the bundle over $`E_H`$ (or $`E_H`$ itself). That is, we vary Wilson lines over $`E_H`$. We then obtain the gauge group $`\mathrm{G}`$ by switching these lines “off”. The deformation of complex structure of $`X`$ then corresponds to deforming the bundle over the K3 surface $`S_H`$ to reabsorb the enhanced gauge symmetry $`\mathrm{G}`$ into a bundle. This extremal transition therefore appears as reducing the structure group of the bundle $`V_EE_H`$ and increasing the structure group of $`V_SS_H`$. We begin in the “Coulomb” branch where $`\mathrm{G}`$ is broken to its Cartan subgroup $`\mathrm{U}(1)^{\mathrm{rank}(\mathrm{G})}`$. We end up in the “Higgs” branch where $`\mathrm{G}`$ may be completely broken. This process therefore decreases the number of vector multiplets as one would expect. An interesting point to bear in mind is that the gauge group $`\mathrm{G}`$ can be broken by quantum effects, i.e., effects due to $`\lambda `$-corrections in the heterotic string and $`\alpha ^{}`$-corrections (specifically worldsheet instantons wrapped around the base $`W`$) in the type IIA string. Even though $`\mathrm{G}`$ is broken however it does not mean that the phase transition cannot happen. Quantum effects cannot obstruct motion in the moduli space and these extremal transitions most certainly exist in terms of Calabi–Yau threefolds. What tends to happen, as explained in , is that the phase transition point does not happen at a point of enhanced gauge symmetry (which need not exist) but rather at a point where some solitons become massless. Only if quantum effects are ignored would these solitons actually produce the enhanced gauge symmetry. In a particularly interesting class of examples the extremal transition can become more complicated. One may have more than one Higgs phase joining on to the Coulomb branch. This is actually understood both in terms of field theory and in terms of the geometry of Calabi–Yau threefolds. An example of a field theory with two Higgs branches was discussed in . The geometry was explained in based on an earlier observation by Gross . As mentioned above, when we go to the six-dimensional picture of this field theory, the degrees of freedom associated to the areas of the elliptic fibration $`p:X\mathrm{\Theta }`$ become frozen. That is, the vector supermultiplets associated to the above gauge groups lose their moduli. Because of this we lose the Coulomb branch of the theory. In other words there are special points in $`\mathrm{M}_H`$ where we may acquire enhanced gauge symmetry but there is never any phase transition associated with such events. #### 4.2.3 Massless Tensors Having said that we lose the standard Higgs-Coulomb phase transitions associated to enhanced gauge symmetry when we look at six dimensional $`N=(1,0)`$ theories, one may ask if we have any transitions at all. There are indeed still interesting phase transitions in six dimensions as was explained in . Going to the six dimensional decompactification limit of the four dimensional theories may freeze out the Kähler form degrees of freedom associated to the fibres of $`p:X\mathrm{\Theta }`$, but there are still Kähler degrees of freedom remaining within $`\mathrm{\Theta }`$ itself. Since these degrees of freedom are present as moduli in six dimensions and descend to vector multiplet moduli in four dimensions, they must be associated to scalars living in the six-dimensional tensor multiplets . Note that the scalar fields in tensor multiplets have only one real degree of freedom. There is no modulus associated to varying the $`B`$-field on $`\mathrm{\Theta }`$. Effectively the periodicity of $`B`$ tends to zero as we decompactify the four dimensional theory to six dimensions. The geometry of the tensor moduli space is given by the real special Kähler geometry of section 3.2.1. The six-dimensional phase transitions are then between a phase spanned by hypermultiplets, which we still call the Higgs phase, and a phase spanned by tensor multiplets, which is called the Coulomb phase for consistency with the four-dimensional picture. In terms of F-theory on a Calabi–Yau threefold $`X`$ this phase transition is really nothing more than the conifold transition we discussed in section 4.2.1. We will give an example here to explicitly give the geometry of the elliptic fibration. For more details on the geometry we refer to . Consider an elliptic fibration whose local Weierstrass form is $$y^2=x^3+s^4x+s^5t.$$ (74) This has a type $`\mathrm{II}^{}`$ fibre running along $`s=0`$ and so one would associate this to an $`E_8`$ gauge group. At $`t=0`$ something special happens. The elliptic fibration degenerates so badly that the only fibre that would smooth the space out would actually be complex dimension two rather than some algebraic curve. To avoid this one may blow up the point $`s=t=0`$ in the base to introduce a new rational curve into $`\mathrm{\Theta }`$. One is certainly not always free to do this! Blowing up any old point in $`\mathrm{\Theta }`$ would usually result in breaking the Calabi–Yau condition. It is only because (74) is so singular that one can do this. The form (74) is therefore precisely at the phase transition point. We may go off into the Higgs phase by deforming the equation, and thus the complex structure of $`X`$. We may go off into the Coulomb phase by blowing up the base $`\mathrm{\Theta }`$ at $`s=t=0`$. ### 4.3 The classical limit The preceding section on extremal transitions gives us invaluable information about specific points in $`\mathrm{M}_H`$ — those which allow phase transitions into new dimensions in $`\mathrm{M}_V`$. We now explore the other part of $`\mathrm{M}_H`$ which is accessible. We will look at the boundary where all quantum effects may be ignored. We have asserted that the heterotic string compactified on $`(V_SS_H)\times (V_EE_H)`$ is dual to the type IIA string compactified on a Calabi–Yau threefold $`X`$. If we could go to a limit in the moduli space where the $`\alpha ^{}`$-corrections to the heterotic string and the $`\lambda `$-corrections to the type IIA string were simultaneously switched off then we should be able to map the two respective moduli spaces of hypermultiplets exactly onto each other. In order to completely ignore $`E_H`$ and its bundle we will assert that we are in the F-theory situation where $`X`$ is a K3 fibration and an elliptic fibration with a section. We will also demand that $`S_H`$ is itself an elliptic surface with a section. This latter demand kills many moduli and one might ask whether one really needed to impose such a drastic constraint. As we will see, it appears to be necessary to get a simple description of the classical moduli spaces. In proposition 8 of section 3.3.2 we showed that the dilaton of the heterotic string is mapped to the area of the $`^1`$ base of $`X`$ as a K3-fibration. While we tried to be quite rigorous in showing proposition 8, there is a quicker (but dirtier) way showing the same thing. Suppose that $`X`$ were not a K3-fibration over $`^1`$ but simply a product of a K3 surface times an elliptic curve. This would yield an $`N=4`$ theory in four dimensions. It is also dual to a heterotic string on $`T^6`$. One may then use a simple dimensional reduction argument to show that the coupling of the heterotic string is given by the area of the elliptic curve on which the type IIA string was compactified. The same argument shows that the coupling of the type IIA string is given by the area of one of the $`T^2`$’s in the heterotic 6-torus. If we assume that $`T^2\times Q`$ (for any space $`Q`$) is equivalent to a $`Q`$-fibration over $`^1`$ as far as areas are concerned then this simple dimensional reduction argument reproduces proposition 8. It also implies that the coupling of the type IIA string is determined by the area of the section of the K3 surface $`S_H`$, as an elliptic fibration, on which the heterotic string is compactified. We will assume this statement is true even though this argument considerably lacks rigour. See for a more thorough treatment of this question. In order to make the type IIA string very weakly coupled we are therefore required to make the section of $`S_H`$ very large on the heterotic side. This will eliminate $`\lambda `$-corrections on the type IIA side. Now in order to remove the $`\alpha ^{}`$-corrections on the heterotic side we are required to make the K3 surface $`S_H`$ very large. Since we have made the section of $`S_H`$ large we have already fulfilled this requirement partially. If we assume that $`S_H`$ is a completely generic elliptic K3 surface with a section, then the only other area we need care about is that of the generic elliptic fibre. If both the section and the fibre have large area then every minimal 2-cycle in $`S_H`$ will be large, unless we have chosen to be close to a special point in the moduli space of Ricci-flat metrics where a 2-cycle shrinks down to zero size. How exactly we take the area of the generic fibre of $`S_H`$ to be infinite was first explained in following an observation in . It was then explored more fully in . We refer the reader to for details of the following argument. We will approach this problem as an algebraic geometer would. For a discussion of the link of this approach with a more metric-minded picture see . The basic idea is that taking the areas of the fibres of $`S_H`$ to be large corresponds to a deformation of complex structure of $`X`$. There is therefore some limiting complex structure of $`X`$ which represents $`S_H`$ at infinite size. We may construct this by considering a one-dimensional family of $`X`$’s. Let $`w:\mathrm{X}D`$ be a fibration of some 4-dimensional complex manifold $`\mathrm{X}`$ over some complex disc $`D`$. Let $`u`$ be a complex parameter for $`D`$. If $`u0`$ then the the fibre $`w^1(u)`$ will be a smooth Calabi–Yau threefold in the class $`X`$. When $`u=0`$, our fibre $`X_0`$ will be singular. It is $`X_0`$ which will correspond to $`S_H`$ with a generic elliptic fibre of infinite area. #### 4.3.1 The $`E_8\times E_8`$ heterotic string In order to proceed further we need to specify whether we are talking about the $`E_8\times E_8`$ heterotic string or the $`\mathrm{Spin}(32)/_2`$ heterotic string. We will deal with the $`E_8\times E_8`$ case first. A picture of what happens as $`X`$ turns into $`X_0`$ is depicted in figure 7 for the case of the $`E_8\times E_8`$ heterotic string. What happens is that the Calabi–Yau threefold $`X`$ “breaks in two” to give a reducible space $`X_1X_2`$ intersecting along a complex surface $`S_{}`$. This surface is an elliptic fibration over a $`^1`$ which we denote $`C_{}`$ in figure 7. The surface $`S_{}`$ is in fact a K3 surface and is isomorphic to $`S_H`$! The way one shows this is via an adiabatic argument where one thinks of $`S_H`$ as a slowly-varying elliptic fibration. One then focuses attention on one elliptic fibre and pretends that the heterotic string compactified on this single fibre is dual to F-theory on a K3 surface. Such an adiabatic argument might be considered a little dangerous when trying to obtain exact results. The fact that we indeed recover a K3 surface $`S_{}`$ in the stable degeneration shows that the result is in fact exact. The only way of mapping the moduli space of $`S_H`$ onto the moduli space of $`S_{}`$ is to identify them with each other! Having determined the heterotic K3 surface $`S_H`$ from the degeneration $`XX_1X_2`$, we should now like to determine the bundle data $`V_S`$. This may be done by a very direct but rather technical process. Whereas $`XW`$ was a K3-fibration, each of $`X_1W`$ and $`X_2W`$ is a fibration with fibre given by a “rational elliptic surface” (sometimes called an “$`E_9`$ Del Pezzo Surface”). Each rational elliptic surface is itself an elliptic fibration over a $`^1`$ (the vertical dotted lines in figure 7). We now need to introduce the notion of the “Mordell–Weil” group $`\mathrm{\Phi }`$ of an elliptic fibration. If we have an elliptic fibration with a given section $`\sigma _0`$ we may associate $`\sigma _0`$ with the identity element of $`\mathrm{\Phi }`$. Any further sections give further elements of $`\mathrm{\Phi }`$. $`\mathrm{\Phi }`$ has a group structure given by the obvious $`S^1\times S^1`$ structure of the elliptic curve. Nontrivial elements of the Mordell–Weil group of the rational elliptic surfaces intersect $`S_{}`$ at points. The locus of all these points generates curves $`C_1`$ and $`C_2`$ within $`S_{}`$ associated to $`X_1`$ and $`X_2`$ respectively. These curves $`C_i`$ each specify an $`E_8`$ bundle over $`S_{}`$. The way that these “spectral curves” (or “cameral curves” to be more precise) determine the bundles is beyond the scope of these lectures. We refer to for details. See also for a discussion of this problem from a toric point of view. Note also that we have the R-R degrees of freedom in the type IIA string from 3-cycles which are invariant under monodromy in $`D`$ around $`u=0`$. Some of these R-R degrees of freedom are essential in determining the $`E_8`$ bundle structure. They translate into specifying a line bundle over the spectral curve. The remaining R-R degrees of freedom describe much of the $`B`$-field degree of freedom of the heterotic string on $`S_H`$ . The “lost” R-R degrees of freedom which are not invariant around the stable degeneration $`u=0`$ can be matched up with the deformations of $`S_H`$ which kill the elliptic fibration and/or the section . While we will not explain here how to determine the $`E_8`$ bundles exactly we will list some of the interesting results we discover in this classical limit. There are a plethora of possibilities! As is common we will refer to the characteristic class of the bundle in $`H^4(S_H,)`$ as “$`c_2`$” even when this bundle is not a $`\mathrm{U}(n)`$-bundle. 1. We may deform a smooth vector bundle so that all of its curvature is concentrated at points. The fundamental such point has $`c_2=1`$ and is known as a “point-like instanton” . It was shown in that such objects can naturally be thought of as an ideal sheaf of a point. These point-like instantons produce a phase transition as described in section 4.2.3 . That is, once we deform a bundle to obtain such an instanton, we obtain a new massless tensor which we may use to move down into the Coulomb phase. 2. We may acquire ADE singularities in $`S_H`$. If the bundle is suitably generic in this case nothing interesting happens. 3. We may acquire ADE singularities in $`S_H`$ and let point-like instantons collide with these singularities. All possible cases were determined in . For example, a collection of $`k`$ point-like $`E_8`$ instantons on a $`^2/_m`$ (that is, type $`A_{m1}`$) quotient singularity, where $`k2m`$, yields $`k`$ new tensor directions in the Coulomb branch and a local contribution to the gauge symmetry of<sup>18</sup><sup>18</sup>18It is possible that the actual group is a discrete quotient of this. This comment also applies to later examples of this nature. $$\mathrm{G}\mathrm{SU}(2)\times \mathrm{SU}(3)\times \mathrm{}\times \mathrm{SU}(m1)\times \mathrm{SU}(m)^{k2m+1}\times \mathrm{SU}(m1)\times \mathrm{}\times \mathrm{SU}(2).$$ (75) One may show that the case $`k<2m`$ reduces to the case obtained by replacing $`m`$ with the integer part of $`k/2`$. 4. One may put fractional point-like instantons on orbifold points. That is, one may concentrate all the curvature of a vector bundle at an orbifold point such that the remaining holonomy is a discrete group which embeds into group associated to the orbifold singularity. Note that for such a bundle we need not have a local integral contribution to $`c_2`$. Many possibilities were discussed in . The interesting feature here is that the finite part of the Mordell–Weil group of the fibration $`p:X\mathrm{\Theta }`$ plays an important rôle. Also in this case, the specific embedding of the holonomy in $`E_8\times E_8`$ must be specified. For example, suppose we take $`S_H`$ to have a singularity of the form $`^2/_2`$ (or $`A_1`$) and take the $`B`$-field associated to this to be zero. Then we build the simplest point-like instanton on this which has monodromy $`_2`$ and breaks $`E_8`$ to $`(E_7\times \mathrm{SU}(2))/_2`$. Such an instanton then has $`c_2=\frac{1}{2}`$ and produces no Coulomb branch or new gauge group enhancement. 5. If however we take the same $`^2/_2`$ singularity but now break $`E_8`$ to $`\mathrm{Spin}(16)/_2`$ then the resulting instantons have $`c_2=1`$ and each produces a nonperturbative contribution of $`\mathrm{SU}(2)`$ to the gauge group. 6. One may “embed the spin connection in the gauge group” to break $`E_8\times E_8`$ to $`E_8\times E_7`$ and then take the limit where one again acquires a $`^2/_2`$ singularity with zero $`B`$-field. This was analyzed in . In this case one obtains a point-like instanton with $`c_2=\frac{3}{2}`$ and a nonperturbative contribution of $`\mathrm{SU}(2)`$ to the gauge group. No new massless tensors appear. By counting point-like instantons one may also arrive at the following (see also for more details) ###### Proposition 14 A type IIA string compactified on an elliptic fibration (with section) over the Hirzebruch surface $`𝐅_n`$ is dual to an $`E_8\times E_8`$ heterotic string compactified on $`(V_SS_H)\times (V_EE_H)`$ where $`V_S=V_S^{(1)}V_S^{(2)}`$. The bundles $`V_S^{(1)}`$ and $`V_S^{(2)}`$ are then $`E_8`$ bundles where $`c_2(V_S^{(1)})=12n`$ and $`c_2(V_S^{(2)})=12+n`$ . The example of section 3.3.4 corresponds to an elliptic fibration over $`𝐅_2`$. Thus it corresponds to an $`E_8\times E_8`$ bundle on $`S_H`$ whose $`c_2`$ is split $`(10,14)`$. Some of the above results may also be approached using toric methods. We refer to for some examples. See also for an interesting conjecture concerning mirror symmetry and these results. Note that in addition to nonperturbatively enhanced gauge symmetry and new massless tensor multiplets, one may also acquire new massless hypermultiplets nonperturbatively. Although these hypermultiplets are massless, they need not provide new directions in the hypermultiplet moduli space. In order to do so they must give massless fields which remain massless when we try to use the fields to move in the moduli space. The usual Higgs mechanism as described above dictates which hypermultiplets remain massless even when one tries to move off into a Higgs branch. Although we have only specified the F-theory rules for analyzing enhanced gauge symmetry and extra massless tensors, there is an assortment of rules for determining the hypermultiplet spectrum and its transformation rules under the gauge symmetry. This is a fascinating subject which links the theory of Lie algebras to the geometry of elliptically fibred Calabi–Yau threefolds. We will not discuss this subject here as it is still a little incomplete. We refer the reader to for more details. A quantum field theory with $`N=(1,0)`$ supersymmetry in six dimensions coupled to gravity may have chiral anomalies coming from both gravity and Yang–Mills. One of the remarkable facts about the F-theory description of these six-dimensional theories is that a massless spectrum is always generated such that all these anomalies cancel. See for an example of this. Why the geometry of Calabi–Yau threefolds should know about these anomalies is currently a mystery. #### 4.3.2 The $`\mathrm{Spin}(32)/_2`$ heterotic string Since we have discussed many of the peculiar properties of the classical limit of an $`E_8\times E_8`$ heterotic string on various bundles on a K3 surface, we should now be able to have just as much fun with the $`\mathrm{Spin}(32)/_2`$ heterotic string. Unfortunately at the present point in time there has been less attention paid to this string, at least in the context of F-theory. Having said that the $`\mathrm{Spin}(32)/_2`$ string is more amenable to analysis in terms of open strings — the $`\mathrm{Spin}(32)/_2`$ is believed to be dual to the type I open string. This allows $`D`$-brane technology to be used as was done in . There is a stable degeneration in the $`\mathrm{Spin}(32)/_2`$ case but it is quite different to the $`E_8\times E_8`$ case . This time, the elliptic fibres break in half as opposed to the base. A generic elliptic fibre becomes two rational curves intersecting at two points (i.e., an $`\mathrm{I}_2`$ fibre in Kodaira’s notation) as depicted in figure 8. At some of the fibres these two rational curves only intersect at a single point. Thus $`X`$ becomes a reducible space $`X_0=X_aX_b`$ where $`X_aX_b`$ is a double cover of the base branched over some subspace. This intersection is again a K3 surface which we take to be equivalent to $`S_H`$. Another difference between the $`E_8\times E_8`$ heterotic string and the $`\mathrm{Spin}(32)/_2`$ heterotic string is that in the latter case the bundle data has yet to be elucidated. Determining the exact way the $`\mathrm{Spin}(32)/_2`$ vector bundle data is encoded in $`X_a`$ and $`X_b`$ may not be particularly difficult and is a problem which should be investigated. Here is a collection of some known results: 1. We may deform a smooth vector bundle so that all of its curvature is concentrated at points. The fundamental such point has $`c_2=1`$ and is known as a “point-like instanton” . $`k`$ such instantons coincident at a smooth point in $`S_H`$ will yield an enhanced gauge symmetry of $`\mathrm{Sp}(k)`$. 2. We may acquire ADE singularities in $`S_H`$. If the bundle is suitably generic in this case nothing interesting happens. 3. We may acquire ADE singularities in $`S_H`$ and let point-like instantons collide with these singularities. All possible cases were determined in . For example, consider a collection of $`k`$ point-like $`\mathrm{Spin}(32)/_2`$ instantons on a $`^2/_m`$ (that is, type $`A_{m1}`$) quotient singularity. If $`m`$ is even and $`k2m`$, then we have $`\frac{1}{2}m`$ new tensor directions in the Coulomb branch and a local contribution to the gauge symmetry of $$\mathrm{Sp}(k)\times \mathrm{SU}(2k8)\times \mathrm{SU}(2k16)\times \mathrm{}\mathrm{}\times \mathrm{SU}(2k4m+8)\times \mathrm{Sp}(k2m).$$ (76) If $`m`$ is odd and $`k2m2`$, then we have $`\frac{1}{2}(m1)`$ new tensor directions in the Coulomb branch and a local contribution to the gauge symmetry of $$\mathrm{Sp}(k)\times \mathrm{SU}(2k8)\times \mathrm{SU}(2k16)\times \mathrm{}\mathrm{}\times \mathrm{SU}(2k4m+4).$$ (77) For smaller values of $`k`$ we refer to . 4. Suppose we put a point-like instanton with $`_2`$ monodromy on an $`A_1`$ singularity such that $`\mathrm{Spin}(32)/_2`$ in broken to $`\mathrm{U}(16)/_2`$. A minimal such instanton has $`c_2=1`$ and gives no new gauge symmetry or tensors . 5. One may produce a peculiar point-like instanton called a “hidden obstructer” which may live anywhere in $`S_H`$, has $`c_2=4`$, and produces a massless tensor leading to a Coulomb phase . Again by counting point-like instantons one may arrive at the following ###### Proposition 15 A type IIA string compactified on an elliptic fibration (with section) over the Hirzebruch surface $`𝐅_n`$ is dual to a $`\mathrm{Spin}(32)/_2`$ heterotic string compactified on $`(V_SS_H)\times (V_EE_H)`$ with $`4n`$ hidden obstructers and where $`c_2(V_S)=8+4n`$. ### 4.4 Into the interior So far we have danced around the edges of the moduli space $`\mathrm{M}_H`$ where we may ignore both the $`\alpha ^{}`$-corrections to the heterotic moduli space and the $`\lambda `$-corrections of the type II moduli space. Surprisingly little is known about what happens if one ventures into the interior of the moduli space. We collect here briefly the few known results. #### 4.4.1 The hyperkähler limit We already mentioned this in section 4.1.2. In effect we may look at the “first order” behaviour as we move away from the classical limit. In any of the examples where we had a perturbative gauge symmetry we may ask what happens if we allow this theory to interact (i.e., allow some coupling or some effective scale $`\mathrm{\Lambda }`$ to be nonzero) while keeping the effective gravitational coupling zero. This would lead to a field theory limit which is described by a hyperkähler moduli space. This is the “rigid limit” of the quaternionic kähler manifold in the same sense as we had a rigid special Kähler limit of a special Kähler manifold. Proposition 13 by Seiberg and Witten gives a powerful tool in this respect. In terms of the heterotic compactification picture we go to the hyperkähler limit by rescaling the overall size of $`S_H`$ to infinity. In order to get something interesting we simultaneously scale down some minimal 2-spheres to keep their areas finite. The result is that we end up describing a heterotic string on an ALE space. The analysis of such systems is perhaps best done by using various dualities involving D-branes along the lines of . Because of this we will regard this subject as beyond the scope of these lectures. We will give one interesting result however. Suppose one were to consider perhaps the simplest case of $`k`$ point-like instantons moving around an ALE space of type $`A_{m1}`$. One can then show that the resulting hyperkähler moduli space with $`k+m1`$ quaternionic dimensions is the same as you would get from the $`c`$-map of section 4.1.2 applied to the rigid limit of $`\mathrm{M}_V`$ for a theory with gauge symmetry $`\mathrm{SU}(m)\times \mathrm{U}(1)^k`$. In other words, suppose our desired moduli space is the hyperkähler limit of $`\mathrm{M}_H`$ which is given by the type IIA string on $`X`$. Then the type IIA string compactified on $`Y`$, the mirror of $`X`$, would yield a gauge symmetry of $`\mathrm{SU}(m)\times \mathrm{U}(1)^k`$. We know from section 4.3.1 that when we go to the classical limit of this theory we will get a gauge group of the form (75). That is, we are in the Higgs branch of a field theory associated to the gauge group (75). From section 4.1.2 this implies that in the three-dimensional picture, mirror symmetry exchanges a field theory with gauge group $`\mathrm{SU}(m)\times \mathrm{U}(1)^k`$ with a field theory with gauge group given by (75). This is a statement of “Intriligator–Seiberg mirror symmetry”. See for many examples of such mirror pairs and for further discussion of this example. Clearly analysis of this hyperkähler limit is much easier than a discussion of the quaternionic Kähler $`\mathrm{M}_H`$ in its full glory. This is essentially because one ends up studying field theory (without gravity) rather than full string theory. #### 4.4.2 Mixed instantons Both the type IIA and type IIB strings suffer from $`\lambda `$-corrections when studying $`\mathrm{M}_H`$. In it was argued that one could study the associated instantons by considering maps of certain cycles into the Calabi–Yau space. These cycles represent the world-volume of D-brane solitons. In a way therefore these $`\lambda `$-corrections could be modeled by something that looks like a generalization of worldsheet instantons. In the case of a the type IIA string on a Calabi–Yau space $`X`$, one needs to consider “supersymmetric” or “special Lagrangian” minimal 3-cycles embedded in $`X`$. (On a related note, such 3-cycles have also achieved prominence from the mirror conjecture of .) Because counting these 3-cycles is very difficult, this approach to computing the quantum corrections has not to date been very useful. Indeed, it will probably be easier to compute the quantum corrections in some other way and then use this to predict the number of 3-cycles — just as was done for rational curves. For the type IIB string, the instanton $`\lambda `$-corrections come from even-dimensional cycles in $`Y`$, including rational curves. Remember that we also have worldsheet instanton corrections coming from rational curves in $`Y`$. Thus it would appear at first that in order to compute the quantum corrections to $`\mathrm{M}_H`$ we should count the rational curves in terms of worldsheet instantons and then add to this the contribution of rational curves from D-1-brane worldsheets. It was shown in that this is not the full story. The subtleties of our discussion of quantum corrections in section 2.6 turn out to have real significance. We only really understand worldsheet instantons when $`\lambda =0`$ and we only understand the D-brane instantons when $`\alpha ^{}=0`$. We have no right to trust either of these pictures when we set both $`\lambda `$ and $`\alpha ^{}`$ to be nonzero. By analyzing a heterotic string on $`S_H\times E_H`$ which is dual to the type IIB string on $`Y`$, one may show that there are many quantum corrections which correspond to instantons which depend on many different combinations of $`\alpha ^{}`$ and $`\lambda `$ . It is as if we had instantons which are both worldsheet and spacetime simultaneously. One very rough way of saying what happens is that the type IIB string in ten dimensions has an $`\mathrm{SL}(2,)`$ symmetry which permutes the fundamental string with “$`(p,q)`$-strings” for any relatively prime $`(p,q)`$. One then needs to add up the contribution from instantons from all of these $`(p,q)`$-strings. On closer inspection this description as it stands is flawed. Firstly, S-duality, like any U-duality, is broken when we have only modestly extended supersymmetry. This was shown explicitly for the type IIB string on $`Y`$ in . Secondly we do not really have a formulation of $`(p,q)`$-strings which allows one to make much sense of a computation of instanton corrections. Understanding these mixed instanton corrections may be one of the most challenging problems for our current definitions of string theory. It may be that we need to replace our basic formulation of string theory to be able to make sense of this problem. #### 4.4.3 Hunting the universal hypermultiplet We will close our discussion of the hypermultiplet moduli space by further demonstrating how troublesome analysis of $`\mathrm{M}_H`$ can be. We want to analyze the question of whether the dilaton belongs to some special hypermultiplet which may have some universal properties for any $`\mathrm{M}_H`$. We will begin by a quick review of some general facts about quaternionic geometry. It is well-known that we may put patches of complex coordinates on a complex manifold $`M_{}`$. That is, we may take some open neighbourhood in $`M_{}`$ with a homeomorphism to some open subset of $`^n`$. Then do this for a collection of patches covering $`M_{}`$ such that the coordinates are related by elements of $`\mathrm{GL}(n,)`$ between patches. We may also consider complex submanifolds of $`M_{}`$. The patches on such submanifolds map holomorphically to the patches of $`M_{}`$. Unfortunately this does not work at all as nicely for quaternionic Kähler manifolds $`M_{}`$. We refer to section 14.F of for more details and references. One might suppose that one could consider patches homeomorphic to an open subset of $`^n`$ such that these coordinates were related by elements of $`\mathrm{Sp}(1).\mathrm{GL}(n,)\mathrm{GL}(4n,)`$. We multiply by $`\mathrm{Sp}(1)`$ on the left and by $`\mathrm{GL}(n,)`$ on the right to try to match the holonomy structure discussed in section 2.1. These would be patches of “quaternionic coordinates”. Unfortunately the only spaces which can admit such a structure are necessarily locally projectively equivalent to quaternionic projection space $`^n`$ . The hypermultiplet moduli spaces one encounters in string theory are not expected to be of this specific form. In other words we would not expect the quaternionic structure of $`\mathrm{M}_H`$ to be “integrable”. For a typical $`\mathrm{M}_H`$ one cannot think in terms of quaternionic coordinates. While it is true that the scalars in a hypermultiplet give a quaternion, these scalars only give tangent directions in the moduli space. There is no way to integrate such a quaternionic structure a nonzero distance along such directions. In other words if one tries to start at a generic point in space and then integrate along the tangent directions given by the 4 massless scalars of a chosen hypermultiplet then one will lose the hypermultiplet structure. The four scalars one ends up with will not be mapped purely into each other by the $`\mathrm{Sp}(1)`$ $`R`$-symmetry. There is also generically a lack of existence of quaternionic submanifolds in a generic quaternionic Kähler manifold, by which we mean the following. If one considers the tangent bundle at a given point $`\mathrm{M}_H`$ one can certainly see a quaternionic structure. One may pick a quaternionic subspace of this and try to integrate along these quaternionic directions to map out a submanifold. After integrating a nonzero distance one will generically discover that one has rotated out of the desired quaternionic structure. In other words, the $`\mathrm{Sp}(1)`$ part of the holonomy will no longer have a closed action within the new tangent directions. Having said this, if one chooses the starting point and tangent directions carefully one can sometimes integrate to find closed manifolds which are compatible with the quaternionic structure. We may call such rare objects quaternionic submanifolds. We emphasize that finding quaternionic submanifolds of a quaternionic manifold is a much harder problem than finding complex submanifolds of a complex manifold. In the notion of a “universal hypermultiplet” was introduced. If one ignores $`\lambda `$-corrections to a type II compactification one might argue from the conformal field theory that the hypermultiplet in which the dilaton lives somehow decouples from the rest of the theory. If this were the case then one could find this universal hypermultiplet by studying any particularly simple example. Consider compactifying the type II string on a 6-torus to obtain a theory in four dimensions with $`N=8`$ supersymmetry. Now imagine what would happen to the moduli space if one embedded the $`\mathrm{U}(2)`$ $`R`$-symmetry of $`N=2`$ into the $`\mathrm{U}(8)`$ $`R`$-symmetry of $`N=8`$. It was argued in that this leads to a natural embedding $$\frac{E_{7(+7)}}{\mathrm{SU}(8)}\frac{\mathrm{SL}(2,)}{\mathrm{U}(1)}\times \frac{\mathrm{SU}(2,1)}{\mathrm{S}(\mathrm{U}(2)\times \mathrm{U}(1))}.$$ (78) The right-hand-side is therefore a possible moduli space for an $`N=2`$ system (embedded in an $`N=8`$ system). Clearly the first factor would be $`\mathrm{M}_V`$ and the second factor would be $`\mathrm{M}_H`$. This would suggest that if a universal hypermultiplet exists it must be of the form $`\mathrm{SU}(2,1)/\mathrm{S}(\mathrm{U}(2)\times \mathrm{U}(1))`$. Even this simplest of examples shows that one cannot expect the universal hypermultiplet to appear as a factor in the moduli space. Equation (78) represents an embedding of the universal hypermultiplet into the moduli space which does not factorize. One should therefore immediately question the validity of saying that the dilaton can be decoupled in a special way from the other fields (even when quantum effects are ignored). One might argue that the failure of the universal hypermultiplet to appear as a factor might be due to an excess of supersymmetry in the above example. This is not so as we see shortly. The best we might hope for then is that the dilaton lives in a hypermultiplet which can be integrated at least at some special points in $`\mathrm{M}_H`$ to give a quaternionic submanifold of dimension one. Let us consider a class of genuine $`N=2`$ examples. We know from the heterotic string that there are many cases where $`\mathrm{M}_H`$ can be described asymptotically (as the K3 surface gets large) by the moduli space of K3 surfaces with bundles. In many of these cases we may freeze the bundle moduli as well as some of the deformations of the K3 itself by pushing point-like instantons into singularities and moving off in the corresponding Coulomb branch. An example of this was studied in . This implies that many examples of $`\mathrm{M}_H`$ look asymptotically like $$\mathrm{M}_H\mathrm{O}(\mathrm{\Lambda }_{4,n})\backslash \mathrm{O}(4,n)/(\mathrm{O}(4)\times \mathrm{O}(n)),$$ (79) for some $`n`$ and some lattice $`\mathrm{\Lambda }_{4,n}`$. Indeed in a few special examples such as there are no quantum corrections and this moduli space is exact (see for the classification of this type of example). Now it is known that any quaternionic submanifold of $`\mathrm{M}_H`$ must be totally geodesic. From an old result of E. Cartan, the totally geodesic submanifolds of a symmetric space are always determined by Lie triples which have been classified (see for example). This will actually allow for an embedding of the universal hypermultiplet (assuming $`n>1`$): $$\frac{\mathrm{SO}_0(4,n)}{\mathrm{SO}(4)\times \mathrm{SO}(n)}\frac{\mathrm{SO}_0(4,2)}{\mathrm{SO}(4)\times \mathrm{SO}(2)}\frac{\mathrm{SU}(2,2)}{\mathrm{S}(\mathrm{U}(2)\times \mathrm{U}(2))}\frac{\mathrm{SU}(2,1)}{\mathrm{S}(\mathrm{U}(2)\times \mathrm{U}(1))}.$$ (80) Note however that (79) does not factorize in any way. This embedding relies very much on the special properties of symmetric spaces. The question we should address however is whether this delicate embedding can be expected to remain when $`\lambda `$-corrections are taken into account. If the deformation of $`\mathrm{M}_H`$ produced by these quantum corrections is sufficiently generic then this embedding will be destroyed even if we were to allow for deformations of the universal hypermultiplet itself. Until we know more about $`\lambda `$-corrections this is impossible to address but for now it would seem to be most prudent to assume that any notion of a universal hypermultiplet, even if only as a quaternionic submanifold of $`\mathrm{M}_H`$ rather than a factor, should be doubted. Since it was the quaternionic structure that caused problems above one might consider an alternative approach to finding the dilaton without trying to keep it cooped up in a special hypermultiplet. It is tempting to conjecture that (79) is the universal behaviour of $`\mathrm{M}_H`$ in the weakly-coupled limit. We can then try something like a decomposition of this symmetric space along the lines of into a warped product such as $$\frac{\mathrm{SO}_0(4,n)}{\mathrm{SO}(4)\times \mathrm{SO}(n)}\frac{\mathrm{SL}(2,)}{\mathrm{U}(1)}\times \frac{\mathrm{SO}_0(2,n2)}{\mathrm{SO}(n2)\times \mathrm{SO}(2)}\times (_+\times )\times ^{2n},$$ (81) where we have pulled the dilaton out as the $`_+`$ factor. Actually this decomposition is well-suited to understanding the stable degenerations of section 4.3. We leave it as an interesting exercise for the reader to interpret each factor (although see for hints!). Of course, this symmetric space is only the asymptotic form of the moduli space $`\mathrm{M}_H`$. The quantum corrections will make everything much more difficult to analyze. Clearly we have much about $`\mathrm{M}_H`$ to learn! ## Acknowledgements It is a pleasure to thank R. Bryant, D. Morrison, R. Plesser and E. Sharpe for numerous conversations and collaborations on topics covered in these lectures. I would also like to thank S. Kachru, J. Harvey, K. T. Mahanthappa and E. Silverstein for organizing TASI99. The author is supported in part by a research fellowship from the Alfred P. Sloan Foundation.
warning/0001/hep-ph0001133.html
ar5iv
text
# Reexamination of the Constraint on Topcolor-Assisted Technicolor Models from 𝑅_𝑏 ## Abstract Recent study on the charged top-pion correction to $`R_b`$ shows that it is negative and large, so that the precision experimental value of $`R_b`$ gives rise to a severe constraint on the topcolor-assisted technicolor models such that the top-pion mass should be of the order of 1 TeV. In this paper, we restudy this constraint by further taking account of the extended technicolor gauge boson correction which is positive. With this positive contribution to $`R_b`$, the constraint on the topcolor-assisted technicolor models from $`R_b`$ changes significantly. The top-pion mass is allowed to be in the region of a few hundred GeV depending on the models. PACS number:12.60.CN,12.60.NZ,13.38.DG The mechanism of electroweak symmetry breaking remains an open question in current particle physics despite the success of the standard model (SM) tested by the CERN $`e^+e^{}`$ collider LEP and SLAC Large Detector (SLD) precision measurement data. In the SM, an elementary Higgs field is assumed to be responsible to break the electroweak symmetry. So far the Higgs bosons has not been found. Recent investigation shows that the LEP-SLD precision measurement data do not really require the existence of a light Higgs boson . Furthermore, theories with elementary scalar fields suffer from the problems of triviality and unnaturalness. To completely avoid these problems arising from the elementary Higgs field, various kinds of dynamical electroweak symmetry breaking mechanisms have been proposed, and among which the topcolor-assisted technicolor theory is an attractive idea. In the topcolor-assisted technicolor theory, there are two kinds of new heavy gauge bosons: (a) the extended technicolor (ETC) gauge bosons including the sideways and diagonal gauge bosons, (b) the topcolor gauge bosons including the color-octet colorons $`C_\mu ^a`$ and an extra $`U(1)`$ gauge boson $`Z^{}`$. The technicolor interactions play the major role in breaking the electroweak gauge symmetry and, in addition, give rise to the masses of the ordinary leptons and quarks including a very small portion of the top-quark mass, namely $`\epsilon m_t`$ with a model-dependent parameter $`\epsilon 1`$. The topcolor interactions also make small contributions to the breaking of the electroweak symmetry, and give rise to the main part of the top-quark mass $`(1\epsilon )m_t`$ similar to the constituent masses of the light quarks in QCD. So that the heaviness of the top quark emerges naturally in the topcolor-assisted technicolor theory. Furthermore, this kind of theory predicts a number of pseudo Goldstone bosons (PGBs) including the technipions in the technicolor sector and the top-pion in the topcolor sector. All the new particles in this theory can give corrections to the $`Z`$-pole observables at LEP and SLC, and thus the LEP-SLD precision data may give constraints on the parameters in the topcolor-assisted technicolor theory. These constraints have recently been studied in Refs.. Due to the strong coupling between the top-pion and the top and bottom quarks, the top-pion gives rise to a large negative correction to the $`Zb\overline{b}`$ branching ratio $`R_b`$. Together with the positive contributions from the colorons and $`Z^{}`$, the total topcolor correction to $`R_b`$ is shown to be quite negative which is of the wrong sign when comparing the SM value of $`R_b`$ to the LEP-SLD data. Since the negative top-pion corrections become smaller when the top-pion is heavier, the LEP-SLD data of $`R_b`$ give rise to certain lower bound on the top-pion mass. It is shown in Ref. that the top-pion mass $`m_{\pi _t}`$ should not be lighter than the order of 1 TeV to make the theory consistent with the LEP-SLD data. This implies that the scale of topcolor should be much higher than what the original model expected . However, in those analyses, the ETC contributions to $`R_b`$ are not taken into account. The main ETC corrections to $`R_b`$ are from the ETC gauge boson contributions. It has been shown in Ref. that the positive diagonal ETC gauge boson contribution is larger than the negative sideways gauge boson contribution, and thus the total ETC correction to $`R_b`$ is positive. It is the purpose of this short paper to investigate how much the constraint on the topcolor-assisted technicolor theory from $`R_b`$ changes when this positive ETC correction is included. Since the corrections to $`R_b`$ in the topcolor-assisted technicolor models depends on the values of the parameters in the models, we shall consider the original topcolor-assisted technicolor model (it will be referred to as Model-I in this paper) and the topcolor-assisted multiscale technicolor model (it will be referred to as Model-II in this paper) as two typical examples in the investigation. These two models are different only in their ETC parts. For the model-dependent parameter $`\epsilon `$, it has been shown that the $`bs\gamma `$ rate is sensitive to the value of $`\epsilon `$, and the CLEO data on the $`bs\gamma `$ rate gives constraint on the value of $`\epsilon `$, namely $`\epsilon 0.1`$ . We shall take three values $`\epsilon =0.05,0.08,`$ and 0.1 in our calculation to see its effect. The left-handed and right-handed $`Zb\overline{b}`$ and $`Zt\overline{t}`$ coupling constants $`g_L^b,g_R^b,g_L^t`$ and $`g_R^t`$ in the SM are, respectively, $`g_L^b=\frac{1}{2}+\frac{1}{3}\mathrm{sin}^2\theta _W,g_R^b=\frac{1}{3}\mathrm{sin}^2\theta _W,g_L^t=\frac{1}{2}\frac{2}{3}\mathrm{sin}^2\theta _W,`$ and $`g_R^t=\frac{2}{3}\mathrm{sin}^2\theta _W`$ Here we have ignored the coupling constant $`\frac{e}{\mathrm{sin}\theta _W\mathrm{cos}\theta _W}`$ which is irrelevant to $`R_b`$.. Let $`\delta g_L^b`$ and $`\delta g_R^b`$ denote, respectively, the corrections to $`g_L^b`$ and $`g_R^b`$ from the topcolor-assisted technicolor theory. Then the correction to $`R_b`$ can be expressed as $`{\displaystyle \frac{\delta R_b}{R_b^{SM}}}{\displaystyle \frac{R_bR_b^{SM}}{R_b^{SM}}}=(1R_b^{SM}){\displaystyle \frac{2[g_L^b\delta g_L^b+g_R^b\delta g_R^b]}{(g_L^b)^2+(g_R^b)^2}},`$ (1) where $`R_b^{SM}=0.2158\pm 0.0002`$ is the SM prediction for $`R_b`$. We shall calculate $`\delta g_L^b`$ and $`\delta g_R^b`$ from various sectors in Model-I and Model-II. We first consider the topcolor sector which is the same in Model-I and Model-II. The Feynman diagrams for the one-loop charged top-pion corrections to the $`Zb\overline{b}`$ vertex and the dependence of $`\delta R_b/R_b^{SM}`$ on $`m_{\pi _t}`$ have been shown in Figs. 1-2 in Ref.. Compared with the charged top-pion contributions, the neutral top-pion contributions are suppressed by a factor of $`m_b^2/m_t^2`$ and thus can be ignored. In Ref., the effect of the technicolor contribution to the top-quark mass $`\epsilon m_t`$ is not taken into account (the result in Ref. corresponds to taking $`\epsilon =0`$). Taking account of the $`\epsilon `$ effect, the total one-loop top-pion correction to $`R_b`$ in the on-shell renormalization scheme reads $`\delta g_L^{b(\pi _t)}=\left({\displaystyle \frac{v_\pi }{v_w}}\right)^2{\displaystyle \frac{[(1\epsilon )m_t]^2V_{tb}^2}{16\pi ^2F_{\pi _t}^2}}\{g_L^b\overline{B}_1(p_b,m_t,m_{\pi _t})+g_R^t[2\overline{C}_{24}^{}+\overline{B}_0(k,m_t,m_t)m_{\pi _t}^2C_0^{}(p_b,k,m_{\pi _t},m_t,m_t)]`$ (2) $`+g_L^tm_t^2C_0^{}(p_b,k,m_{\pi _t},m_t,m_t)+(12\mathrm{sin}^2\theta _W)\overline{C}_{24}(p_b,k,m_t,m_{\pi _t},m_{\pi _t})\},`$ (3) $`\delta g_R^{b(\pi _t)}=0,`$ (4) where $`v_\pi /v_w=(167\mathrm{GeV})/(174\mathrm{GeV})`$ reflects the effect of the mixing between the top-pion and the would-be Goldstone boson , $`F_{\pi _t}=50`$ GeV is the top-pion decay constant, $`p_b`$ and $`k`$ are, respectively, the momenta of the external $`b`$ quark and $`Z`$ boson, $`B_i`$ and $`C_{ij}`$ are the two-point and three-point scalar integral functions. The factor $`[(1\epsilon )m_t]^2`$ in (3) comes from the $`\pi _tt\overline{b}`$ coupling when the technicolor contribution to the top-quark mass is considered. This factor causes the $`\epsilon `$-dependence of $`\delta g_L^{b(\pi _t)}`$. The negative correction to $`R_b`$ from the top-pion decreases with $`\epsilon `$. In Fig. 1, we plot the top-pion contributed $`\delta R_b/R_b^{SM}`$ versus $`m_{\pi _t}`$ with $`\epsilon =0,0.05,0.08,0.1`$. The $`\epsilon =0`$ curve is just the result given in Ref.. The contributions to $`\delta g_L^b`$ and $`g_R^b`$ from the topcolor gauge bosons $`C_\mu ^a`$ and $`Z_\mu ^{}`$ have been calculated in Ref., which are $`\delta g_L^{b(C^a)}=g_L^b{\displaystyle \frac{\kappa _3}{6\pi }}C_2(R)\left[{\displaystyle \frac{m_Z^2}{M_C^2}}\mathrm{ln}{\displaystyle \frac{M_C^2}{m_Z^2}}\right],\delta g_R^{b(C^a)}=g_R^b{\displaystyle \frac{\kappa _3}{6\pi }}C_2(R)\left[{\displaystyle \frac{m_Z^2}{M_C^2}}\mathrm{ln}{\displaystyle \frac{M_C^2}{m_Z^2}}\right],`$ (5) $`\delta g_L^{b(Z^{})}=g_L^b{\displaystyle \frac{\kappa _1}{6\pi }}(Y_L^b)^2\left[{\displaystyle \frac{m_Z^2}{M_Z^{}^2}}\mathrm{ln}{\displaystyle \frac{M_Z^{}^2}{m_Z^2}}\right],\delta g_R^{b(Z^{})}=g_R^b{\displaystyle \frac{\kappa _1}{6\pi }}(Y_R^b)^2\left[{\displaystyle \frac{m_Z^2}{M_Z^{}^2}}\mathrm{ln}{\displaystyle \frac{M_Z^{}^2}{m_Z^2}}\right],`$ (6) where $`\kappa _3`$ and $`\kappa _1`$ are, respectively, the coloron and the $`Z^{}`$ couplings , $`M_C`$ and $`M_Z^{}`$ are, respectively, the masses of $`C_\mu ^a`$ and $`Z^{}`$, $`C_2(R)=\frac{4}{3}`$, $`Y_L^b=\frac{1}{3}`$, and $`Y_R^b=\frac{2}{3}`$. We shall take $`M_B=M_Z^{}=1`$ TeV in the calculation. To obtain proper vacuum tilting (the topcolor interactions only condense the top quark but not the bottom quark), the couplings $`\kappa _3`$ and $`\kappa _1`$ should satisfy certain constraint. There is a region of $`\kappa _3`$ and $`\kappa _1`$, namely $`\kappa _3=2,\kappa _11`$, satisfying the requirement of vacuum tilting and the constraints from Z-pole physics and $`U(1)`$ triviality shown in Refs.. We shall take $`\kappa _3=2`$ and $`\kappa _1=1`$ in the following calculation. Next, we consider the ETC sector corrections to $`R_b`$. In the topcolor-assisted technicolor theory, the technipion-top-bottom coupling is proportional to $`\frac{\epsilon m_t}{F_\pi }`$, and the technipion corrections to $`\delta g_L^b`$ and $`\delta g_R^b`$ are proportional to $`(\frac{\epsilon m_t}{F_\pi })^2`$ which is very small, so that the technipion correction to $`R_b`$ is negligible. Therefore, the main contribution is from the ETC gauge bosons. This has been calculated in Ref. which reads $`\delta g_L^{b(ETC)}={\displaystyle \frac{1}{A}}{\displaystyle \frac{\epsilon m_t}{16\pi F_\pi }}\sqrt{{\displaystyle \frac{N_{TC}}{N_C}}}[{\displaystyle \frac{2N_C}{N_{TC}+1}}\xi _t(\xi _t^1+\xi _b)\xi _t^2],`$ (7) where $`N_{TC}`$ and $`N_C`$ are, respectively, the number of technicolors and the numbers of ordinary colors, $`\xi _t`$ and $`\xi _b`$ are coupling coefficients with $`\xi _b=(m_s/m_c)\xi _t^1`$ , and $`\xi _t`$ is ETC gauge-group dependent. Following Refs., we take $`\xi _t=1/\sqrt{2}`$. The factor $`\frac{1}{A}`$ reflects the walking effect in the ETC sector which is taken to be $`A=1.7`$ in Refs.. The decay constant $`F_\pi `$ is different in Model-I and Model-II. In Model-I, the ETC sector is the one-family ETC model. Considering the mixing between the top-pion and the would-be Goldstone boson, we have $`N_d(F_\pi /\sqrt{2})^2+F_{\mathrm{\Pi }_t}^2=v_w^2`$ ($`N_d=4`$ is the number of $`SU(2)`$ doublets in the one-family technifermion sector), and thus $`F_\pi =118`$ GeV. In Model-II, the ETC sector is the multiscale technicolor model in which $`F_\pi =40`$ GeV . This difference makes the ETC corrections to $`R_b`$ very different in Model-I and Model-II. This positive ETC correction to $`R_b`$ is larger in Model-II than in Model-I. Finally, we add all the corrections together and obtain the total corrections $`\delta g_L^b`$ $`=`$ $`\delta g_L^{b(\pi _t)}+\delta g_L^{b(C^a)}+\delta g_L^{b(Z^{})}+\delta g_L^{b(ETC)},`$ (8) $`\delta g_R^b`$ $`=`$ $`\delta g_R^{b(C^a)}+\delta g_R^{b(Z^{})},`$ (9) in which $`\delta g_L^{b(\pi _t)}`$, $`\delta g_L^{b(C^a)}`$, $`\delta g_R^{b(C^a)}`$, $`\delta g_L^{b(Z^{})}`$, $`\delta g_R^{b(Z^{})}`$, and $`\delta g_L^{b(ETC)}`$ are given in eqs.(3),(5),(6), and (7), respectively. Plugging (8) and (9) into (1), we obtain the total correction $`\delta R_b/R_b^{SM}`$ and the predicted $`R_b=R_b^{SM}+\delta R_b`$ in Model-I and Model-II. Before presenting the numerical results, let us make an examination of the parameter range $`\epsilon =0.050.1`$ which we take in the calculation. It has been noticed that the ETC sector not only gives rise to a positive contribution to $`R_b`$ but also contributes positive correction to the oblique correction parameter $`T`$ (or equivalently $`\mathrm{\Delta }\rho `$, $`\mathrm{\Delta }\rho =\alpha T`$) due to the violation of the custodial symmetry $`SU(2)_c`$ in the ETC sector . In the original ETC model, the top quark mass is completely generated by the ETC dynamics, so that the violation of $`SU(2)_c`$ in the ETC sector is very serious and the positive contribution to $`T`$ (or $`\mathrm{\Delta }\rho `$) is so large that it can barely be consistent with the experiment . Now we examine this problem in the topcolor-assisted technicolor models. In the topcolor-assisted technicolor models, the ETC sector only gives rise to a very small portion of the top quark mass, $`\epsilon m_t`$, therefore the violation of $`SU(2)_c`$ in the ETC sector is significantly smaller depending on the values of $`\epsilon `$. It has been shown that the most dangerous positive contrbution to $`T`$ in the ETC sector is from the exchange of the diagonal ETC gauge boson whose couplings to the up-type and down-type techniquarks are different, and this has been studied in Ref.. Since the four-fermion operators contributing the positive correction to $`T`$ is also related to the ETC generation of the top and bottom quark masses, the formula in Ref. can be further expressed in terms of the parameters $`\epsilon m_t,\xi _t`$ and $`\xi _b`$ as follows $`T^{ETC}={\displaystyle \frac{1}{A}}{\displaystyle \frac{1}{16\mathrm{sin}^2\theta _W\mathrm{cos}^2\theta _W}}{\displaystyle \frac{N_C^2}{N_{TC}(N_{TC}+1)}}{\displaystyle \frac{\epsilon m_tF_\pi }{m_Z^2}}\sqrt{{\displaystyle \frac{N_{TC}}{N_C}}}[\xi _t^1\xi _b]^2.`$ (10) For $`\epsilon =0.05,0.08`$ and $`0.1`$, the values of $`T^{ETC}`$ are 0.006, 0.009 and 0.01, respectively. These are to be compared with the experimental value $`T=0.00\pm 0.15`$ . We see that, for the parameter range $`\epsilon =0.050.1`$ which we take in this paper, the ETC contributed positive correction to $`T`$ is small enough to make the theory consistent with the experiment The corrections to $`T`$ from the exchange of topcolor gauge boson has been studied in Ref... Now we compare our predicted $`R_b`$ with the experimental value $`R_b^{expt}=0.21642\pm 0.00073`$ to get the new constraints on the two typical topcolor-assisted technicolor models. The results of the predicted $`R_b`$ in Model-I with $`\epsilon =0.05,0.08,`$ and $`0.1`$ are plotted in Fig. 2 together with the experimental value $`R_b^{expt}`$. The horizontal solid line denotes the central value $`R_b^{expt}`$, and the horizontal dotted lines indicate the $`1\sigma `$ and $`2\sigma `$ deviations. We see from Fig. 1 that the $`2\sigma `$ constraints on Model-I are $`\epsilon =0.05:400\mathrm{GeV}m_{\pi _t},`$ (11) $`\epsilon =0.08:340\mathrm{GeV}m_{\pi _t}900\mathrm{GeV},`$ (12) $`\epsilon =0.1:280\mathrm{GeV}m_{\pi _t}770\mathrm{GeV}.`$ (13) The results of the predicted $`R_b`$ in Model-II with $`\epsilon =0.05,0.08`$ and $`0.1`$ are plotted in Fig. 3 together with the experimental value $`R_b^{expt}`$ and the $`1\sigma `$ and $`2\sigma `$ deviations. Fig. 3 shows that the $`2\sigma `$ constraints on Model-II are $`\epsilon =0.05:350\mathrm{GeV}m_{\pi _t}900\mathrm{GeV},`$ (14) $`\epsilon =0.08:250\mathrm{GeV}m_{\pi _t}560\mathrm{GeV},`$ (15) $`\epsilon =0.1:220\mathrm{GeV}m_{\pi _t}430\mathrm{GeV}.`$ (16) We see that the constraints on Model-I and Model-II are different due to the different values of $`F_\pi `$ in the two models. Since $`F_\pi `$ takes a smaller value (causing a larger positive ETC correction to $`R_b`$) in Model-II, the allowed top-pion mass is lower in Model-II than in Model-I. From Fig. 2 and Fig. 3, we see that, when the positive ETC gauge boson correction to $`R_b`$ is taken into account, the constraints on the two typical topcolor-assisted technicolor models are significantly different from that shown in Refs.. As is mentioned in Ref., this kind of constraint should only be regarded as a rough estimate since the $`\pi _tt\overline{b}`$ coupling is so strong that higher order corrections from the top-pion are expected to be important. Anyway, the conclusion of the present investigation is that, to be consistent with the experimental value $`R_b^{expt}`$, the top-pion mass is roughly in a region of a few hundred GeV, and thus the scale of topcolor is likely to be around a couple of TeV which is not much higher than what is expected in the original topcolor-assisted technicolor theory . Acknowledgment This work is supported by the National Natural Science Foundation of China, the Fundamental Research Foundation of Tsinghua University, a special grant from the Ministry of Education of China, and the Natural Science Foundation of the Henan Scientific Committee.
warning/0001/hep-ph0001325.html
ar5iv
text
# Strangeness Form Factors of the Proton in the Chiral Quark Model ## I Introduction The HAPPEX experiment shows the combination $`G_E^s+0.39G_M^s`$ of the strange charge and magnetic form factors of the proton at $`Q^2=0.48`$ (GeV/$`c)^2`$ to be consistent with $`0(0.023\pm 0.034\pm 0.022\pm 0.026)`$. Similarly the SAMPLE experiment shows that $`G_M^s`$ at $`Q^2=0.1`$ (GeV/$`c)^2`$ is consistent with 0, modulo uncertainties in the calculated value of the weak axial form factor of the nucleon . The experimental result that the strangeness form factors of the proton are small may be used to constrain or test theoretical models for nucleon structure, as the theoretical predictions for these observables have covered a fairly wide range . A calculation of the strangeness form factors of the proton based on the chiral quark model is reported here. The approach considers the strangeness component of the proton as loop fluctuations, with intermediate strange mesons and $`s`$-quarks, of the constituent quarks that form the proton. The constituent quark model approach represents an alternative to the hadronic approach, in which the strangeness components are considered as fluctuations of the nucleon into intermediate strange mesons and hyperons. The chiral quark model approach brings the advantages of much smaller coupling constants and consequently the possibility of a converging loop expansion that takes all baryonic intermediate states into account. Moreover, as the amplitudes of the loops mainly scale with the inverse squared mass of the intermediate meson, heavy meson contributions are suppressed. Examples of this are the desirably small meson loop contributions to the neutron charge radius and the recent demonstration that the strange meson loop contributions to the proton give rise to but a very small strangeness magnetic moment . It is shown here that the kaon and $`K^{}`$ loop fluctuations of the light quarks, that are illustrated in Fig. 1, and which lead to but a small value for the strangeness magnetic moment of the proton, also lead to strange form factors of the proton with very small magnitudes. The results are found to be fairly insensitive to the value of the cut-off scale for the loop integrals, provided that this is taken to be about 1 GeV or larger, ie. values commensurate with the chiral symmetry restoration scale $`4\pi f\pi 1.2`$ GeV, at which scale the pseudoscalar mesons are expected to decouple from the constituent quarks. The magnitude of the strange loop contributions to the strangeness form factors of the proton is of the order $`(g^2/4\pi ^2)(m_q^2/(m_q^2+m_M^2))`$, where $`g`$ is the meson-quark coupling, and $`m_q`$ and $`m_M`$ are the masses of the light constituent quarks and strange mesons respectively. As $`g^2/4\pi 0.7`$ for $`K`$ and $`K^{}`$ mesons, it follows with $`m_q300`$ MeV that the loop amplitude in the case of kaons is expected to only be about $`0.06`$, and smaller still in the case of $`K^{}`$ mesons. In comparison the expected magnitude of a typical loop amplitude in the case of the hadronic approach, where $`g^2/4\pi 10`$ and $`m_q`$ is replaced by the proton mass, is more than an order of magnitude greater. This is also revealed by a comparison of the calculated values for the strangeness magnetic moment of the proton in Refs. and . It suggests that a small net loop contribution in the hadronic approach only can result as a consequence of strong cancellations between several large amplitudes unless strong cut-offs are invoked. This paper falls into 5 sections. In section 2 the contribution of the strange loop amplitudes to the proton form factors, that contain intermediate kaons are derived. The corresponding results for the loop amplitudes that involve intermediate $`K^{}`$ mesons are derived in section 3. The contribution from loop fluctuations with $`K^{}K\gamma `$ vertices is derived in section 4. The numerical results for the strangeness form factors of the proton are calculated in section 5. ## II Kaon loop contributions The strangeness form factors of the nucleon are defined as the invariant coefficients of the matrix elements of the operators $`\overline{s}\gamma _\mu s`$ in the proton. In standard notation the strangeness current is $$p^{}|j_\mu ^s(0)|p=i\overline{u}(p^{})\left[F_1^s(Q^2)\gamma _\mu F_2^s(Q^2)\frac{\sigma _{\mu \nu }q_\nu }{2m_N}\right]u(p).$$ (1) Here $`q=p^{}p`$, and $`Q^2=q^2=𝒒^2q_0^2>0`$, and $`m_N`$ is the proton mass. The two form factors are calculated here in the constituent quark model from the strangeness loop fluctuations illustrated by the Feynman diagrams in Fig. 1, where the meson lines represents $`K`$ and $`K^{}`$ mesons. Consider first the kaon loop diagrams. To calculate these, we consider the kaon-quark pseudovector coupling: $$=i\frac{f_{Kqs}}{m_K}\overline{\psi }\gamma _5\gamma _\mu \underset{a=4}{\overset{7}{}}\lambda ^a_\mu K^a\psi .$$ (2) The pseudovector kaon-quark coupling constant is obtained as $$f_{Kqs}=\frac{g_A^a}{2}\frac{m_K}{f_K},$$ (3) where $`g_A^q=0.87`$ for quarks and $`f_K`$ is the kaon decay constant $`(f_K=113`$ MeV). The numerical value for $`f_{Kqs}`$ is then $`f_{Kqs}=1.9`$. The kaon and strange quark current density operators have the form $`j_\mu =`$ $`ie\{_\mu K^{}K+\text{h.c.}\},`$ () $`j_\mu =`$ $`i{\displaystyle \frac{e}{3}}\overline{\psi }_s\gamma _\mu \psi _s.`$ () The standard convention on the strangeness form factors assigns the $`s`$-quark a strangeness charge of $`+1`$ and the kaon, which contains an $`\overline{s}`$-quark, a strangeness charge of $`1`$. In the calculation of the loop amplitudes the $`s`$-quark current (2.4b) should therefore be multiplied by $`3`$ and the kaon current (2.4a) by $`1`$. The pseudovector coupling term (2.2) requires introduction of a contact coupling term for current conservation. This contact current term gives rise to two seagull diagrams, which exactly cancel the corresponding seagull diagrams, which arise in the evaluation of the amplitudes of the loop diagrams in Fig. 1, upon application of the Dirac equation for the external quarks. The remaining loop amplitudes are equivalent to those, which arise if the loop amplitudes are calculated using the pseudoscalar coupling $$_{Kqs}=ig_{Kqs}\overline{\psi }\gamma _5\underset{a=4}{\overset{7}{}}\lambda ^aK^a\psi ,$$ (4) where the pseudoscalar coupling constant $`g_{Kqs}`$ is defined as $$g_{Kqs}=\frac{m_q+m_s}{m_K}f_{Kqs}.$$ (5) In these expressions $`m_q`$ represents the constituent mass of the light flavor quarks $`(u,d)`$ and $`m_s`$ represents the strange quark mass. For these masses we shall use the values $`m_q=340`$ MeV and $`m_s=460`$ MeV respectively . With these mass values we obtain the value $`g_{Kqs}^2/4\pi =0.75`$ for the (squared) kaon-quark pseudoscalar coupling constant. The kaon loop contributions to the Dirac form factors $`F_1^s`$ of the quarks are logarithmically divergent. We regularize these loop amplitudes by cutting off the loop momentum integrals smoothly at the chiral restoration scale $`\mathrm{\Lambda }_\chi =4\pi f_\pi =1.2`$ GeV. The cut-off is implemented by replacing the meson propagator $`v(k^2)=1/(k^2+m_K^2)`$ in the loop diagram, that contains the $`s`$-quark current coupling (Fig. 1a), by the propagator multiplied by a dipole form factor $$v(k^2)\frac{1}{m_K^2+k^2}\left[\frac{\mathrm{\Lambda }^2m_K^2}{\mathrm{\Lambda }^2+k^2}\right]^2.$$ (6) Current conservation then demands that the product of the two meson propagators in the loop amplitude that corresponds to the kaon current loop (Fig. 1b) be modified as $$\frac{1}{m_K^2+k_1^2}\frac{1}{m_K^2+k_2^2}\frac{v(k_2^2)v(k_1^2)}{k_1^2k_2^2}.$$ (7) The calculation then proceeds by first calculating the strangeness form factors $`F_1^s`$ and $`F_2^s`$ for the constituent quarks. These form factors are the same for the $`u`$\- and $`d`$-quarks. The relation of these form factors to the corresponding strangeness form factors of the proton is simple only under the assumption that $`m_q=m_p/3`$, which implies an equipartition of the total proton momentum between the quarks: $$F_1^s(Q^2)=3F_{1q}^s(Q^2),F_2^s(Q^2)=\frac{m_p}{m_q}F_{2q}^s(Q^2).$$ (8) From these relations we obtain the charge and magnetic form factors of the proton as $$G_E^s(Q^2)=F_1^s(Q^2)\frac{Q^2}{4m_p^2}F_2^s(Q^2),$$ (9) $$G_M^s(Q^2)=F_1^s(Q^2)+F_2^s(Q^2).$$ (10) The magnetic form factor at $`Q^2=0`$ then yields the strangeness magnetic moment in units of nuclear magnetons. Charge conservation requires that $`F_1^s(0)=0`$. This requirement is satisfied by subtracting the value of $`F_1^s(0)`$ from the expression below. The contribution to the strangeness form factor $`F_1^s`$ of the proton from the two loop diagrams in Fig. 1 are obtained as $`F_{1q}^s`$ $`(Q^2)\{a,K\}={\displaystyle \frac{g_{Kqs}^2}{8\pi ^2}}{\displaystyle _0^1}dx(1x){\displaystyle _0^1}dy\{[(m_sm_q)^2+2m_q(m_sm_q)(1x)`$ () $`+m_q^2(1x)^2Q^2(1x)^2(1y)y]K_1(Q^2)+\mathrm{ln}{\displaystyle \frac{H_1(\mathrm{\Lambda }_\chi ^2)}{H_1(m_K^2)}}x{\displaystyle \frac{\mathrm{\Lambda }_\chi ^2m_K^2}{H_1(\mathrm{\Lambda }^2)}}\},`$ $`F_{1q}^s`$ $`(Q^2)\{b,K\}={\displaystyle \frac{g_{Kqs}^2}{8\pi ^2}}{\displaystyle _0^1}dxx{\displaystyle _0^1}dy\{2m_q(m_sm_qx)(1x)K_2(Q^2)`$ () $`\mathrm{ln}{\displaystyle \frac{H_2(\mathrm{\Lambda }_\chi ^2)}{H_2(m_k^2)}}+x{\displaystyle \frac{\mathrm{\Lambda }^2m_k^2}{H_2(\mathrm{\Lambda }_\chi ^2)}}\}.`$ Here the functions $`K_1(Q^2)`$ and $`K_2(Q^2)`$ have been defined as $`K_1(Q^2)=`$ $`{\displaystyle \frac{1}{H_1(m_K^2)}}{\displaystyle \frac{1}{H_1(\mathrm{\Lambda }_\chi ^2)}}x{\displaystyle \frac{\mathrm{\Lambda }_\chi ^2m_K^2}{H_1^2(\mathrm{\Lambda }_\chi ^2)}},`$ () $`K_2(Q^2)=`$ $`{\displaystyle \frac{1}{H_2(m_K^2)}}{\displaystyle \frac{1}{H_2(\mathrm{\Lambda }_\chi ^2)}}x{\displaystyle \frac{\mathrm{\Lambda }_\chi ^2m_K^2}{H_2^2(\mathrm{\Lambda }_\chi ^2)}}.`$ () The denominator functions $`H_1(m^2)`$ and $`H_2(m^2)`$ are defined as $`H_1(m^2)=`$ $`m_s^2(1x)m_q^2x(1x)+m^2x+Q^2(1x)^2y(1y),`$ () $`H_2(m^2)=`$ $`m_s^2(1x)m_q^2x(1x)+m^2x+Q^2x^2y(1y).`$ () After subtraction of the corresponding values at $`Q^2=0(F_1^s(0))`$ from the form factors $`F_1^s(Q^2)`$ the integrals remain finite even in the limit $`\mathrm{\Lambda }_\chi ^2\mathrm{}`$. The corresponding contributions to the strangeness Pauli form factors of the quarks is $`F_{2q}^s(Q^2)\{a,K\}=`$ $`{\displaystyle \frac{g^2}{4\pi ^2}}{\displaystyle _0^1}𝑑x(1x)^2{\displaystyle _0^1}𝑑ym_q(m_sm_qx)K_1(Q^2),`$ () $`F_{2q}^s(Q^2)\{b,K\}=`$ $`{\displaystyle \frac{g^2}{4\pi ^2}}{\displaystyle _0^1}𝑑xx(1x){\displaystyle _0^1}𝑑ym_q(m_sm_qx)K_2(Q^2).`$ () These expressions reduce to those given in Ref. in the limit $`Q^20`$, if multiplied by the factor $`m_p/m_q`$ to give strangeness magnetic moments in units of nuclear magnetons. In Fig. 2 the kaon loop contributions to the proton strangeness Dirac form factor $`F_1^s`$ are shown as functions of momentum transfer, after subtraction of the irrelevant constant $`F_1^s(0)`$. This loop contribution to the proton strangeness form factor is very small and negative, and for $`Q^21`$ (GeV/$`c)^2`$ it decreases slowly from 0 to $`0.01`$. As shown below, the magnitude of this contribution is smaller than that of the strange vector meson loops. The contributions from the kaon loop amplitudes to the strangeness Pauli form factor $`F_2^s`$ are shown in Fig. 3. These contributions, while small, are notably larger than the corresponding vector meson loop contributions that are derived in section 3 below. The momentum dependence of the kaon loop contribution to $`F_2^s(Q^2)`$ is fairly weak for $`Q^2`$ values below $`1`$ (GeV/c)<sup>2</sup>. ## III Strange vector meson loop fluctuations The coupling of $`K^{}`$ mesons to constituent quarks is described by the Lagrangian $$_{K^{}qs}=ig_{K^{}qs}\overline{\psi }_s\left(\gamma _\mu +\frac{m_sm_q}{m_K^{}^2}_\mu +i\frac{\kappa _{K^{}qs}}{m_s+m_q}\sigma _{\mu \nu }_\nu \right)\underset{a=4}{\overset{7}{}}\lambda ^aK_\mu ^a\psi +\text{h.c.}.$$ (13) This coupling is a generalization to fermions of unequal mass of the conventional transverse Proca coupling for vector mesons. The coupling constants $`g_{K^{}qs}`$ and $`\kappa _{K^{}qs}`$ may be determined from the corresponding couplings of $`K^{}`$ mesons to the baryon octet by the quark model relations : $`g_{K^{}qs}=`$ $`g_{K^{}\overline{B}B},`$ (14) $`g_{K^{}qs}(1+\kappa _{K^{}qs})=`$ $`{\displaystyle \frac{3}{5}}{\displaystyle \frac{m_s+m_q}{\overline{M}}}g_{K^{}\overline{B}B}(1+\kappa _{K^{}\overline{B}B}).`$ () Here $`\overline{M}`$ represents the average of the nucleon and $`S=1`$ hyperon $`(\mathrm{\Lambda },\mathrm{\Sigma })`$ masses. A recent comprehensive boson exchange potential model fit to nucleon-nucleon scattering data gives $`g_{K^{}\overline{B}B}=2.97`$ and $`\kappa _{K^{}\overline{B}B}=4.22`$, with a liberal uncertainty margin . These values yield $`g_{K^{}qs}^2/4\pi 0.7`$ and $`\kappa _{K^{}qs}0.21`$. The small value of the tensor coupling $`\kappa _{K^{}qs}`$ and its large uncertainty range suggests that it is consistent with 0. At this stage it is therefore justified to neglect the Pauli term in (3.1) altogether. The current density operator for the $`K^{}`$ mesons takes the form $$j_\mu =\pm ie\{K_\nu ^{}_\mu K_\nu ^{}K_\nu ^{}_\nu K_\mu ^{}\}+\mathrm{h}.\mathrm{c}.$$ (15) The contribution to the strangeness from factors $`F_1^s`$ of the $`u`$\- and $`d`$-quarks from the $`K^{}`$ meson loops described by the Feynman diagrams in Figs. 1a and b (when the meson line represents a $`K^{}`$ meson) are $`F_{1q}^s`$ $`(Q^2)\{a,K^{}\}={\displaystyle \frac{g_{K^{}qs}^2}{4\pi ^2}}{\displaystyle _0^1}dx(1x){\displaystyle _0^1}dy\{[m_s^24m_qm_sx+m_q^2x^2`$ () $`Q^2(x+(1x)^2y(1y))]\overline{K}_1(Q^2)+\mathrm{ln}{\displaystyle \frac{H_1(\mathrm{\Lambda }_\chi ^2)}{H_1(m_K^{}^2)}}x{\displaystyle \frac{\mathrm{\Lambda }_\chi ^2m_K^{}^2}{H_1(\mathrm{\Lambda }_\chi ^2)}}\}+𝒪({\displaystyle \frac{1}{m_K^{}^2}}).`$ $`F_{1q}^s`$ $`(Q^2)\{b,K^{}\}={\displaystyle \frac{g_{K^{}qs}^2}{8\pi ^2}}{\displaystyle _0^1}dxx{\displaystyle _0^1}dy\{[6m_q^2x(1x)6m_qm_s(1x)`$ () $`+Q^2x(12xy(1y))]\overline{K}_2(Q^2)+6[\mathrm{ln}{\displaystyle \frac{H_2(\mathrm{\Lambda }_\chi ^2)}{H_2(m_K^{}^2)}}x{\displaystyle \frac{\mathrm{\Lambda }_\chi ^2m_K^{}^2}{H_2(\mathrm{\Lambda }_\chi ^2)}}]\}+𝒪({\displaystyle \frac{1}{m_K^{}^2}}).`$ Here the auxiliary functions $`\overline{K}_1(Q^2)`$ and $`\overline{K}_2(Q^2)`$ have been defined as the functions $`K_1(Q^2)`$ and $`K_2(Q^2)`$ in Eqs. (2.12), with the replacement of $`m_K^2`$ by $`m_K^{}^2`$. The terms of order $`m_K^{}^2`$ and higher powers of $`m_K^{}^2`$ in (3.3) arise from the terms proportional to $`m_K^{}^2`$ in the vector meson propagator $`(\delta _{\mu \nu }+k_\mu k_\nu /m_K^{}^2)/(m_K^{}^2+k^2)`$ and the coupling (3.1). These terms are small in comparison to the terms of Eqs. (3.3) at low values of $`Q^2`$. The explicit expressions for the contributions of order $`m_K^{}^2`$ to $`F_{1q}^s`$ that are indicated in Eqs. (3.4a) and (3.4b) from the two loop diagram amplitudes illustrated in Figs. 1a and b are $`F_{1q}^s`$ $`(Q^2)\{a,K^{},𝒪(m_K^{}^2)\}={\displaystyle \frac{g_{K^{}qs}^2}{8\pi ^2}}{\displaystyle \frac{m_q^2}{m_K^{}^2}}{\displaystyle _0^1}𝑑x(1x){\displaystyle _0^1}𝑑y`$ (18) $`\left\{\right[`$ $`m_s^2(1x)^2{\displaystyle \frac{m_s^2}{m_q^2}}Q^2(1x)^2y(1y)`$ () $`+m_sm_q[2x(1x)+{\displaystyle \frac{Q^2}{m_q^2}}(1x)^2(1+2xy(1y))]+m_q^2x^2(1x)^2`$ $`+Q^2(1x)^2(2x^2y(1y)y(1y^2)+x(13y+4y^2y^3))`$ $`+{\displaystyle \frac{Q^4}{m_q^2}}(1x)^2(1y)y(x+(1x)^2(1y)y)]\overline{K}_1(Q^2)+[{\displaystyle \frac{m_s^2}{m_q^2}}+2{\displaystyle \frac{m_s}{m_q}}(13x)`$ $`+1+6x(1x)+{\displaystyle \frac{Q^2}{m_q^2}}(23x6(1x)^2y(1y))\left]\right[\mathrm{ln}{\displaystyle \frac{H_1(\mathrm{\Lambda }_\chi ^2)}{H_1(m_K^{}^2)}}x{\displaystyle \frac{\mathrm{\Lambda }_\chi ^2m_K^{}^2}{H_1(\mathrm{\Lambda }_\chi ^2)}}]`$ $`+{\displaystyle \frac{6}{m_q^2}}[H_1(m_K^{}^2)\mathrm{ln}{\displaystyle \frac{H_1(m_K^{}^2)}{H_1(\mathrm{\Lambda }_\chi ^2)}}H_1(m_K^{}^2)+H_1(\mathrm{\Lambda }_\chi ^2)]\},`$ $`F_{1q}^s`$ $`(Q^2)\{b,K^{},𝒪(m_K^{}^2)\}={\displaystyle \frac{g_{K^{}qs}^2}{8\pi ^2}}{\displaystyle \frac{m_q^2}{m_K^{}^2}}{\displaystyle 𝑑xx𝑑y}`$ () $`\{[m_sm_q[(1x)^3+{\displaystyle \frac{Q^2}{m_q^2}}(1x)].+Q^2x^2(1x)[12y(1y)(1+x)]`$ $`+{\displaystyle \frac{Q^4}{m_q^2}}x^2y(1y)[1x+2x^2y(1y)]]\overline{K}_2(Q^2)`$ $`+[6{\displaystyle \frac{m_s}{m_q}}(1x)+6(1x)^2(1x)(35x).`$ $`+{\displaystyle \frac{Q^2}{m_q^2}}(2+3x11x^2y(1y))\left]\right[\mathrm{ln}{\displaystyle \frac{H_2(\mathrm{\Lambda }_\chi ^2)}{H_2(m_K^{}^2)}}x{\displaystyle \frac{\mathrm{\Lambda }_\chi ^2m_K^{}^2}{H_2(\mathrm{\Lambda }_\chi ^2)}}]`$ $`+{\displaystyle \frac{9}{m_q^2}}[H_2(m_K^{}^2)\mathrm{ln}{\displaystyle \frac{H_2(m_K^{}^2)}{H_2(\mathrm{\Lambda }_\chi ^2)}}H_2(m_K^{}^2)+H_2(\mathrm{\Lambda }_\chi ^2)]\}`$ The contributions from the $`K^{}`$ loop diagrams in Fig. 1 to the strangeness factors $`F_2^s`$ of the $`u`$\- and $`d`$-quarks are obtained as $`F_{2q}^2`$ $`(Q^2)\{a,K^{}\}={\displaystyle \frac{g_{K^{}qs}^2}{4\pi ^2}}{\displaystyle _0^1}𝑑x(1x){\displaystyle _0^1}𝑑y\mathrm{\hspace{0.17em}2}m_qx[2(m_sm_q)+m_q(1x)]\overline{K}_1(Q^2)`$ () $`+𝒪\left({\displaystyle \frac{1}{m_K^{}^2}}\right),`$ $`F_{2q}^s`$ $`(Q^2)\{b,K^{}\}={\displaystyle \frac{g_{K^{}qs}^2}{4\pi ^2}}{\displaystyle _0^1}𝑑xx{\displaystyle _0^1}𝑑ym_q[m_s(3x2)m_qx(2x1)]\overline{K}_2(Q^2)`$ () $`+𝒪({\displaystyle \frac{1}{m_K^{}^2}}).`$ Finally the corresponding terms that are proportional to $`1/m_K^2`$ have the expressions: $`F_{2q}^2`$ $`(Q^2)\{a,K^{},𝒪(m_K^{}^2)\}={\displaystyle \frac{g_{K^{}qs}^2}{4\pi ^2}}{\displaystyle \frac{m_q^2}{m_K^{}^2}}{\displaystyle _0^1}dx(1x){\displaystyle _0^1}dy\left\{\right[(1x)^2(m_sm_q)(m_s+m_qx)`$ () $`+{\displaystyle \frac{Q^2}{m_q}}(m_sm_q)(1x^2)(1x)(1y)y]\overline{K}_1(Q^2)`$ $`2(1{\displaystyle \frac{m_s}{m_q}})(1{\displaystyle \frac{3}{2}}x)(\mathrm{ln}{\displaystyle \frac{H_1(\mathrm{\Lambda }_\chi ^2)}{H_1(m_K^{}^2)}}x{\displaystyle \frac{\mathrm{\Lambda }_\chi ^2m_K^{}^2}{H_1(\mathrm{\Lambda }_\chi ^2)}})\}`$ $`F_{2q}^s`$ $`(Q^2)\{b,K^{},𝒪(m_K^{}^2)\}={\displaystyle \frac{g_{K^{}qs}^2}{4\pi ^2}}{\displaystyle \frac{m_q^2}{m_K^{}^2}}{\displaystyle _0^1}dxx{\displaystyle _0^1}dy\{[m_q(1x)^2(m_sm_qx^2)`$ () $`+Q^2x^2y(1y)(22x+x^2{\displaystyle \frac{m_s}{m_q}})]\overline{K}_2(Q^2)`$ $`2[{\displaystyle \frac{m_s}{m_q}}2(1x)^2x][\mathrm{ln}{\displaystyle \frac{H_2(\mathrm{\Lambda }_\chi ^2)}{H_2(m_K^{}^2)}}x{\displaystyle \frac{\mathrm{\Lambda }_\chi ^2m_K^{}^2}{H_2(\mathrm{\Lambda }_\chi ^2)}}]\}`$ These expressions reduce to those derived in Ref. in the limit $`Q^20`$, once multiplied by $`m_p/m_q`$ in order to obtain the results in units of nuclear magnetons. The strange vector meson loop contributions to the strangeness Dirac form factor $`F_1^s(Q^2)`$ of the proton are obtained after subtraction of the irrelevant values $`F_{1q}^s(0)`$ from the expressions (3.4) and (3.5) and multiplication of the sum of the remainders by a factor 3 (2.9). This contribution is shown in Fig. 2 along with the corresponding kaon loop contribution. In this case the vector meson loop contribution is larger in magnitude than the kaon loop contribution. Even so the sum of the $`K`$ and $`K^{}`$ loop contributions remains very small in magnitude, reaching only the value -0.086 around $`Q^2=1(`$GeV/c$`)^2`$. The contribution of the strange vector meson loops to the strangeness Pauli form factor is very small because of a near cancellation between the two involved loop diagrams . This contribution is shown in Fig. 3, with the much larger contribution from the kaon loop diagrams. ## IV The $`K^{}K`$ loop contribution We finally consider the strangeness loop fluctuation, for which the e.m. coupling is to the $`K^{}K`$ transition vertex (Fig. 4). The amplitude of this loop fluctuation may be calculated from the empirically known radiative widths of the $`K^{}`$ mesons. The $`K^{}K`$ transition current vertex has the form $$K^a(k^{})|J_\mu |K_\sigma ^b(k)=i\frac{g_{K^{}K\gamma }}{m_K^{}}ϵ_{\mu \lambda \nu \sigma }k_\lambda k_\nu ^{}\delta ^{ab}$$ (35) The coupling constant $`g_{K^{}K\gamma }`$ depends on the charge state of the strange mesons. It may be determined from the radiative decay widths using the expression $$\mathrm{\Gamma }(K^{}K\gamma )=\alpha \frac{g_{K^{}K\gamma }^2}{24\pi }m_K^{}\left[1\left(\frac{m_K}{m_K^{}}\right)^2\right]^3.$$ (36) Here $`\alpha `$ is the fine structure constant. Given the empirical radiative widths $`\mathrm{\Gamma }(K_{}^{}{}_{}{}^{+}K^+\gamma )=50`$ keV and $`\mathrm{\Gamma }(K_{}^{}{}_{}{}^{0}K^0\gamma )=116`$ keV , the corresponding coupling constant values are obtained as $`g_{K_{}^{}{}_{}{}^{\pm }K^\pm \gamma }=0.75`$ and $`g_{K_{}^{}{}_{}{}^{0}K^0\gamma }=1.14`$ (with the sign convention of ). The $`K^{}K`$ loop diagrams (Fig. 4) only contribute to the strange Pauli form factors of the quarks. With the convention of assigning the $`K`$ and $`K^{}`$ mesons a “strangeness charge” of $`1`$, the contribution to the strangeness Pauli form factor $`F_{2q}^s(Q^2)`$ from these loop diagrams is found to be $`F_{2q}^s(Q^2)=`$ $`{\displaystyle \frac{g_{Kqs}g_{K^{}qs}g_{K^{}K\gamma }}{2\pi ^2}}{\displaystyle \frac{m_q}{m_K^{}}}{\displaystyle _0^1}dxx{\displaystyle _0^1}dy\{m_q(1x)(m_sm_qx)`$ () $`\times ({\displaystyle \frac{1}{G_1}}{\displaystyle \frac{1}{G_2}}{\displaystyle \frac{1}{G_3}}+{\displaystyle \frac{1}{G_4}})\mathrm{ln}\left({\displaystyle \frac{G_2G_3}{G_1G_4}}\right)\}.`$ The quantities $`G_i`$ here have been defined as $$\begin{array}{ccc}G_1=G(m_K,m_K^{}),\hfill & G_2=G(\mathrm{\Lambda }_\chi ,m_K^{}),\hfill & \\ G_3=G(m_K,\mathrm{\Lambda }_\chi ),\hfill & G_4=G(\mathrm{\Lambda }_\chi ,\mathrm{\Lambda }_\chi ).\hfill & \end{array}$$ (38) The auxiliary function $`G(m,m^{})`$ is defined as $$G(m,m^{})=m_s^2(1x)m_q^2x(1x)+m^2x(1y)+m^2xy+Q^2x^2y(1y).$$ (39) These expressions represent direct generalizations of the corresponding expressions defined in Ref. for the case $`Q^2=0`$. To account for the different coupling constants $`g_{K^{}K\gamma }`$ in the case of $`u`$\- and $`d`$-quarks, (2.9) is generalized to $$F_2^s(Q^2)=\frac{m_p}{m_s}\left[\frac{4}{3}F_{2u}^s(Q^2)\frac{1}{3}F_{2d}^s(Q^2)\right].$$ (40) The $`K^{}K`$ loop contribution to $`F_2^s(Q^2)`$ is shown in Fig. 3. This loop contribution has the opposite sign to that of the diagonal strangeness loop fluctuations. ## V SAMPLE and HAPPEX The SAMPLE experiment measures the strangeness magnetic form factor of the proton $`G_M^s`$ at $`Q^2=0.1`$ (GeV/$`c)^2`$. The HAPPEX experiment measures the combination $`G_E^s+0.39G_M^s`$ at $`Q^2=0.48`$ (GeV/$`c)^2`$. The contributions to these observables from the kaon and $`K^{}`$ loops in Figs. 1 and 4 are shown in Figs. 5 and 6 as a function of momentum transfer. The cut-off dependence is small , and for $`\mathrm{\Lambda }_\chi =1.2`$ GeV we obtain $`G_M^s(0.1)=0.06`$ and $`G_E^s(0.48)+0.39G_M^s(0.48)=0.08`$. The latter value is slightly below the uncertainty range of the result of the HAPPEX experiment. Similarly the former value is within the uncertainty range of the result of the SAMPLE experiment, provided that $`G_A^Z`$ is small and positive rather than large and negative as originally suggested . The smallness of the calculated strangeness observables of the proton is an inherent feature of the chiral quark model. The smallness of the measured strangeness observables suggests that this model may provide a useful framework for describing those observables. The calculated momentum dependence of $`G_E^s`$ agrees fairly well with that obtained by heavy baryon chiral perturbation theory . The strangeness radius is $`0.02`$ fm<sup>2</sup>, which value is also obtained in the baryon loop calculations in Ref. if the cut-off scale is set to the chiral symmetry restoration scale. Because of the large negative vector meson contribution to the Dirac form factor $`F_1^s`$ the calculated $`G_M^s`$ form factor grows more negative with increasing $`Q^2`$, whereas the third order chiral perturbation theory result, which does not consider vector mesons, for $`G_M^s`$ is that it increases slowly with momentum transfer. Hitherto QCD lattice calculations have been made only for the strangeness magnetic moment $`G_M^s(0)`$, but not for the form factors. The calculated values are negative, with substantial uncertainty limits (-0.36 $`\pm `$ 0.20 , -0.16 $`\pm `$ 0.18 ). The chiral quark model value for $`G_M^s(0)0.06`$ falls within the uncertainty range of the latter value. ###### Acknowledgements. We are grateful for the hospitality of Professor R. D. McKeown at the W. K. Kellogg Radiation Laboratory of the California Institute of Technology where this work was completed. L. H. thanks the Waldemar von Frenckell foundation for a stipend. This work was supported in part by the Academy of Finland under contract 43982.
warning/0001/hep-ex0001018.html
ar5iv
text
# Simulation of UHE muons propagation for GEANT3 ## 1 Introduction Underground and underwater detectors for neutrino astronomy require simulation tools capable to correctly handle the propagation of high energy muons up to $`PeV`$ energy region and above. The new object-oriented (C++) version of the detector simulation tool GEANT (GEANT4 ), will probably be suitable for this goal. However many experiments still make use of GEANT3 to simulate the detector response. Even if such package can in principle perform the simulation of particle propagation above 10 $`TeV`$, it has been mainly designed for accelerator experiments, whose tipical energy range doesn’t exceed the $`TeV`$ region. The simulation of muonic interactions is actually guaranteed by the authors for muon energy below 10 $`TeV`$ . This is due to the parametrizations of cross sections for radiative processes contained in GEANT3, which are reliable only for energy up to 10 $`TeV`$. Moreover the description of photonuclear interaction is realized in a frame which significantly disagrees with theoretical calculations, for each muon energy. Our new simulation code (GMU) has been carried out to make the GEANT3 standard library reliable in reproducing UHE muons propagation through matter. GMU replaces the simulation of the radiative muonic interactions performed by GEANT3, for each energy and for each material, keeping the same structure of the original program. The procedure to apply the new code is completely transparent in such a way that no adjustments must be implemented in the programs that make use of the standard GEANT3 library. ## 2 Radiative Cross sections in GEANT3 and GMU High-energy muons propagating through matter mainly interact by quasi-continuous (ionization) and discrete (bremsstrahlung, direct electron-positron pair production, photonuclear interaction) energy losses mechanisms. Ionization dominates at energy lower than few hundreds of $`GeV`$ while above the $`TeV`$ region energy losses are mainly due to radiative processes. The radiative processes are simulated stocastically by GEANT3 above a fixed (user supplied) transferred energy threshold; below this threshold they are treated as continuous. The models which describe the interaction processes are the physical input for the code. In more detail, simulating a given process requires: * To evaluate the probability of occurrence of the process by sampling the total cross section of the process * To generate the final state after interaction by sampling the differential cross section of the process The reliability of the simulation is then affected both by the formulas (and/or parametrizations) chosen for cross sections and by the alghoritms used for sampling and for numerical integration. Standard reference formulas for UHE calculations can be found in the paper by Lohmann and Voss which also tabulate the average muon energy losses for many materials and compounds up to $`10`$ $`TeV`$. Direct electron pair production differential cross section has been first calculated by Kelner and Kotov in the framework of QED theory . We have used the well-known parametrization performed by Kokoulin and Petrukhin which considers the corrections for atomic and nuclear form factors. For bremsstrahlung differential cross section we have used the formula derived by Andreev and Bugaev which takes into account the structure of nuclear target (elastic and inelastic form factors) and the exact contributions due to atomic electrons (screening effect and bremsstrahlung of the muons on electrons). The formula used by Lohmann and Voss had been carried out just from this one by Petrukhin and Shestakov by neglecting bremsstrahlung on electrons and nuclear effects in light materials ($`Z<10`$). The discrepancy between the two approaches reaches few percents, in terms of bremsstrahlung energy losses, in the worse case (for materials with $`Z10`$) and it is almost always negligible in the calculation of the total muon energy losses. For the photonuclear interaction we have used the differential cross section calculated by Bezrukov and Bugaev within the vector meson dominance hypothesis. We also have considered the recent accelerator data coming from ZEUS and H1 according to the ref. . For this process we have not considered, at present, the angle between the ingoing and outgoing muon, taking the reasonable approximation of completely forward scattering. As a consequence, the code is not suitable for studies dedicated to muon-nucleus scattering at large angles. Total cross section and average energy loss for each radiative process $`k`$ can be calculated, starting from differential cross sections, as follows: $$\sigma _{Tot}^k=_{vmin}^{vmax}\frac{d\sigma ^k}{dv}(v,E)𝑑v$$ (1) $$\frac{dE}{dx}_k=\frac{N_A}{A}E_{vmin}^{vmax}v\frac{d\sigma ^k}{dv}(v,E)𝑑v$$ (2) * $`N_A`$ and $`A`$ are Avogadro’s number and the mass number respectively * $`v`$ is the fraction of initial energy $`E`$ lost by muon at the occurrence of the process $`k`$ * $`vmin`$ and $`vmax`$ are the kinematical limits for the allowed values of $`v`$ (,) * $`x`$ is the thickness of throughgone matter, expressed in $`g/cm^2`$ * $$\frac{d\sigma ^k}{dv}(v,E)$$ is the differential cross section for the process $`k`$ (explicit formulas are given in the Appendix) The calculation of energy losses performed with the formulas described above agree with the Lohmann and Voss results at percent level in the energy range they considered ($`1GeV`$-$`10TeV`$.) In fig. 1 we report the total cross sections for the radiative processes as they are tabulated at the initialization procedures by standard GEANT321 and by GEANT321 plus GMU. In the first case the points are calculated with approximated analytical parametrizations whose accuracy is guaranteed (at least for bremsstrahlung and pair production) within $`5\%`$ in the declared energy range ($`E<10`$ $`TeV`$). In the second case the points are calculated by performing numerical integration of the formulas given in the Appendix. (This operation requires a small amount of CPU time at the initialization, depending also on the number of materials considered.) The photonuclear cross sections differ by more than one order of magnitude. This discrepancy has already been observed in previous works and it is expected to affect the total energy loss especially in the case of light materials. The total cross sections for bremsstrahlung and pair production below 100 $`TeV`$ seem to be in excellent agreement but for energy above 100 $`TeV`$ the GEANT321 cross sections strongly deviate from expectation. ## 3 Simulation of muon energy loss In order to evaluate the correctness of the simulation code we have performed some checks for different materials. For each test run we have activated the full stochastical regime and we have evaluated the average energy loss resulting from the simulation. The values achieved with standard GEANT321 and with GEANT321 plus GMU have been compared with numerical calculations of the same quantities \[see eq.(2)\]. In fig. 2 and 3 we present the results for water and for standard rock respectively. The simulation performed by GEANT321 plus GMU reproduces very well the numerical calculation at any energy, both for total energy losses and for the individual processes. For such materials, the GEANT321 simulation is in good agreement up to 100 $`TeV`$, but it produces much greater average energy loss for energy exceeding 100 $`TeV`$. In table 1 we report the total and the photonuclear energy losses in hydrogen as they result from simulation performed by the two codes ($``$ GEANT321; $``$ GEANT321+GMU). The discrepancy is not negligible, even at low energy, being around $`20\%`$ at 10 $`TeV`$. The main source of this disagreement is clearly due to a different estimation of the photonuclear process, whose relative importance grows for light materials. In order to further verify the reliability of the simulation code we performed some specific runs by recording the fraction $`v`$ of energy lost by muon at the occurrence of each process. (Indeed the average energy loss is not especially sensitive to very big or very small energy transfer). The distributions of such variable have been compared with analytical expressions for differential cross sections. The results are shown in fig. 4, 5 and 6 for a fixed muon energy ($`E=100`$ $`TeV`$): they confirm the correctness of the sampling performed by GMU. ## 4 Conclusions The simulation of muon propagation carried out by standard GEANT3 seems to be out of control for energy above $`10^5`$ $`GeV`$. Moreover GEANT3 slightly underestimates the energy losses for very light materials (at lower energies too), due to a strong discrepancy between expected and simulated values of the muon photonuclear cross section. GEANT321 in addition to the new code GMU performs complete agreement with expectation in the tested energy range (up to $`10^8`$ $`GeV`$). Further improvements will consist in a more accurate simulation of the direction of outgoing particle produced in photonuclear interaction and in taking into account the influence of the medium (LPM effect) wich becomes important in the process of muon bremsstrahlung at extreme high energy. In this region direct muon pair production by muons should be taken into account too. The GMU code can be requested by conctacting the authors. E-mail: * Bottai`@`fi.infn.it * Lorenzo`@`le.infn.it ## 5 Acknowledgments We would like to thank the members of the Lecce MACRO group and prof. V.A.Naumov for useful discussions and concrete help. APPENDIX Radiative cross sections formulae * $`N_A`$ is Avogadro’s number * $`Z`$ and $`A`$ are the atomic number and the atomic weight of the material * $`v`$ is the fraction of initial energy $`E`$ lost in the interaction ($`E^{}=E(1v)`$) * $`\alpha `$ is the fine structure constant * $`r_e`$ is the classical electron radius * $`\lambda _e=r_e\alpha ^1`$ is the electron Compton wavelength * $`m_e`$, $`m_\mu `$, $`m_\pi `$ and $`M_p`$ are the electron, muon, pion and proton rest masses ($`c=1`$) respectively The formulae for radiative differential cross sections are taken from ref. for $`e^+e^{}`$ pair-production, from ref. for bremsstrahlung and from ref. and for photonuclear interaction. $``$ PAIR PRODUCTION $$\frac{d\sigma _p}{dvd\rho }=\alpha ^4\frac{2}{3\pi }(Z\lambda _e)^2\frac{1v}{v}[\mathrm{\Phi }_e+\frac{m_e^2}{m_\mu ^2}\mathrm{\Phi }_\mu ]$$ $$\begin{array}{cc}\hfill \mathrm{\Phi }_e=& \{[(2+\rho ^2)(1+\beta )+\xi (3+\rho ^2)]ln(1+\frac{1}{\xi })\hfill \\ & +\frac{1\rho ^2\beta }{1+\xi }(3+\rho ^2)\}S_e\hfill \end{array}$$ $$\begin{array}{cc}\hfill \mathrm{\Phi }_\mu =& \{[(1+\rho ^2)(1+\frac{3\beta }{2})\frac{1}{\xi }(1+2\beta )(1\rho ^2)]ln(1+\xi )\hfill \\ & +\frac{\xi (1\rho ^2\beta )}{1+\xi }+(1+2\beta )(1\rho ^2)\}S_\mu \hfill \end{array}$$ $$\begin{array}{cc}\hfill S_e& =ln\frac{RZ^{1/3}\sqrt{(1+\xi )(1+Y_e)}}{1+\frac{2m_e\sqrt{e}RZ^{1/3}(1+\xi )(1+Y_e)}{Ev(1\rho ^2)}}\hfill \\ & \frac{1}{2}ln[1+(\frac{3}{2}\frac{m_e}{m_\mu }Z^{1/3})^2(1+\xi )(1+Y_e)]\hfill \end{array}$$ $$S_\mu =ln\frac{(2/3)(m_\mu /m_e)RZ^{2/3}}{1+\frac{2m_e\sqrt{e}RZ^{1/3}(1+\xi )(1+Y_\mu )}{Ev(1\rho ^2)}}$$ $$Y_e=\frac{5\rho ^2+4\beta (1+\rho ^2)}{2(1+3\beta )ln(3+\frac{1}{\xi })\rho ^22\beta (2\rho ^2)}$$ $$Y_\mu =\frac{4+\rho ^2+3\beta (1+\rho ^2)}{(1+\rho ^2)(\frac{3}{2}+2\beta )ln(3+\xi )+1\frac{3}{2}\rho ^2}$$ $$\xi =\frac{m_\mu ^2v^2}{4m_e^2}\frac{(1\rho ^2)}{(1v)};\beta =\frac{v^2}{2(1v)};$$ $$\rho =\frac{E^+E}{E^++E}$$ The integration limits are: $$\frac{4m_e}{E}v1\frac{3\sqrt{e}}{4E}m_\mu Z^{1/3}$$ $$0|\rho |\left(1\frac{6m_\mu ^2}{E^2(1v)}\right)\sqrt{1\frac{4m_e}{Ev}}$$ where: * $`E^+`$ and $`E^{}`$ are the energies of the positron and electron * $`R`$=189 is the radiation logarithm value * $`e=exp(1)`$ We have taken into account the influence of the atomic electrons replacing $`Z^2`$ by $`Z(Z+1)`$. $``$ BREMSSTRAHLUNG $$\frac{d\sigma _b}{dv}=\alpha (\frac{1}{v})\left(2r_eZ\frac{m_e}{m_\mu }\right)^2\left\{\left(1+\frac{E^{}_{}{}^{}2}{E^2}\right)\mathrm{\Phi }_1(q_{\mathrm{min}},Z)\frac{2}{3}\frac{E^{^{}}}{E}\mathrm{\Phi }_2(q_{\mathrm{min}},Z)\right\}$$ $$\mathrm{\Phi }_{1,2}(q_{\mathrm{min}},Z)=\mathrm{\Phi }_{1,2}^0(q_{\mathrm{min}},Z)$$ $$\begin{array}{cc}\hfill \mathrm{\Phi }_1^0(q_{\mathrm{min}},Z)=& \frac{1}{2}\left(1+\mathrm{ln}\frac{m_\mu ^2a_1^2}{1+x_1^2}\right)x_1\mathrm{arctan}\frac{1}{x_1}\hfill \\ & +\frac{1}{Z}\left[\frac{1}{2}\left(1+\mathrm{ln}\frac{m_\mu ^2a_2^2}{1+x_2^2}\right)x_2\mathrm{arctan}\frac{1}{x_2}\right]\hfill \end{array}$$ $$\begin{array}{cc}\hfill \mathrm{\Phi }_2^0(q_{\mathrm{min}},Z)=& \frac{1}{2}\left(\frac{2}{3}+\mathrm{ln}\frac{m_\mu ^2a_1^2}{1+x_1^2}\right)+2x_1^2\left(1x_1\mathrm{arctan}\frac{1}{x_1}+\frac{3}{4}\mathrm{ln}\frac{x_1^2}{1+x_1^2}\right)\hfill \\ & +\frac{1}{Z}\left[\frac{1}{2}\left(\frac{2}{3}+\mathrm{ln}\frac{m_\mu ^2a_2^2}{1+x_2^2}\right)+2x_2^2\left(1x_2\mathrm{arctan}\frac{1}{x_2}+\frac{3}{4}\mathrm{ln}\frac{x_2^2}{1+x_2^2}\right)\right],\hfill \end{array}$$ $$\mathrm{\Delta }_1(q_{\mathrm{min}},Z)=\mathrm{ln}\frac{m_\mu }{q_c}+\frac{a}{2}\mathrm{ln}\frac{a+1}{a1}$$ $$\mathrm{\Delta }_2(q_{\mathrm{min}},Z)=\mathrm{ln}\frac{m_\mu }{q_e}+\frac{a}{4}(3a^2)\mathrm{ln}\frac{a+1}{a1}+\frac{2m_\mu ^2}{q_c^2}$$ $$q_{\mathrm{min}}\frac{m_\mu ^2v}{2E(1v)},x_i=a_iq_{\mathrm{min}},$$ $$a=\sqrt{1+\frac{4m_\mu ^2}{q_c^2}},q_c=1.9m_\mu Z^{1/3},$$ $$a_1=\frac{184.15}{\sqrt{e}Z^{1/3}m_e},a_2=\frac{1194.0}{\sqrt{e}Z^{2/3}m_e},$$ $``$ where $`e=exp(1)`$ The integration limits are: $$0v1\frac{3\sqrt{e}}{4E}m_\mu Z^{1/3}$$ $``$ PHOTONUCLEAR INTERACTION $$\begin{array}{cc}\hfill \frac{d\sigma _n}{dv}& =\frac{\alpha }{8\pi }A\sigma _{\gamma p}\left(\nu \right)v\{H\left(v\right)\mathrm{ln}(1+\frac{m_2^2}{t})\frac{2m_\mu ^2}{t}\hfill \\ \\ & +G\left(z\right)\left[H\left(v\right)\mathrm{ln}\left(1+\frac{m_1^2}{t}\right)H\left(v\right)\frac{2m_1^2}{m_1^2+t}\frac{2m_\mu ^2}{t}\right]\hfill \\ \\ & +\frac{2\xi m_\mu ^2}{t}[G\left(z\right)\frac{2m_1^2}{m_1^2+t}+\frac{m_2^2}{t}\mathrm{ln}(1+\frac{t}{m_2^2})]\}\hfill \end{array}$$ $`\nu =vE`$ is the energy of virtual photon $$\begin{array}{c}H(v)=1\frac{2}{v}=\frac{2}{v^2},G(z)=\frac{9}{z}\left\{\frac{1}{2}+\frac{1}{z^2}[(1+z)e^z1]\right\},\\ \\ z=0.00282A^{\frac{1}{3}}\sigma _{\gamma p}(\nu ),t=\frac{m_\mu ^2v^2}{1v},\\ \\ m_1^2=0.54\mathrm{GeV}^2,\mathrm{m}_2^2=1.80\mathrm{GeV}^2,\xi =0.25.\end{array}$$ The differential cross section is proportional to the total cross section $`\sigma _{\gamma N}`$, for absorption of a real photon of energy $`\nu =s/2m_N=vE`$ by a nucleon. In this calculation we have used the Regge-type parametrization for $`\sigma _{\gamma N}`$, according ref. This model performs the best fit to accelerator data (,). $$\sigma _{\gamma N}=\left[67.7s^{0.0808}+129s^{0.4525}\right]\mu b$$ The integration limits are: $$\frac{1}{E}(m_\pi +\frac{m_\pi ^2}{2M_p})v1(1+\frac{m_\mu ^2}{M_p^2})\frac{M_p}{2E}$$ Figure 1: Total Cross Sections for radiative interaction processes vs muon energy (PP $``$ $`e^+e^{}`$ pair-production, BS $``$ bremsstrahlung, PH $``$ photonuclear interaction). The plot shows the values tabulated by GEANT321 (dashed lines) and by GEANT321 plus GMU (full lines). The material is Standard Rock ($`Z=11`$, $`A=22`$). Figure 2: Average total muon energy loss in Water ($`Z/A=0.555`$) vs the $`Log_{10}E`$ ($`E`$ is the muon energy). The plot shows the results from GEANT321 (empty triangles) and from GEANT321 plus GMU (full stars); the contributions from individual radiative processes (performed by GEANT321 plus GMU) and results from numerical calculations (lines) are also shown. Statistical error bars are comparable with dots size in case of photonuclear interactions (PH) and bremsstrahlung (BR). Figure 3: Average total muon energy loss in Standard Rock ($`Z=11`$, $`A=22`$) vs the $`Log_{10}E`$ ($`E`$ is the muon energy). The plot shows the results from GEANT321 (empty triangles) and from GEANT321 plus GMU (full stars); the contributions from individual radiative processes (performed by GEANT321 plus GMU) and results from numerical calculations (lines) are also shown. Statistical error bars are comparable with dots size in case of photonuclear interactions (PH) and bremsstrahlung (BR). Figure 4: Differential cross section for muon bremsstrahlung vs $`Log_{10}v`$ ($`v`$ is the fraction of energy $`E`$ lost by muon in the interaction). The numerical calculation (dashed line) is compared with the result obtained from the simulation performed with the code GEANT321 plus GMU. The material is Standard Rock ($`E=100`$ $`TeV`$). Figure 5: Differential cross section for pair production vs $`Log_{10}v`$ ($`v`$ is the fraction of energy $`E`$ lost by muon in the interaction). The numerical calculation (dashed line) is compared with the result obtained from the simulation performed with the code GEANT321 plus GMU. The material is Standard Rock ($`E=100`$ $`TeV`$). Figure 6: Differential cross section for photonuclear interaction vs $`Log_{10}v`$ ($`v`$ is the fraction of energy $`E`$ lost by muon in the interaction). The numerical calculation (dashed line) is compared with the result obtained from the simulation performed with the code GEANT321 plus GMU. The material is Standard Rock ($`E=100`$ $`TeV`$).
warning/0001/cond-mat0001452.html
ar5iv
text
# Decoherence of Schrődinger cat states in a Luttinger liquid ## I Introduction and motivations Decoherence is a very important issue for mesoscopic systems since it governs the crossover between quantum and quasi-classical transport regimes. Coherence of mesoscopic conductors gives rise to various Aharonov-Bohm interference effects such as permanent currents in a mesoscopic ring, and conductance oscillations as a function of the external magnetic field. The problem of electron decoherence in metals at zero temperature is an active area of research both from the theoretical and experimental point of view. Recent experiments claim to observe a saturation of the dephasing time $`\tau _\varphi `$ at very low temperatures Mohanty:1997-1 . Strong discussions among theorists arose from these observationsAleiner:1998-1 ; Golubev:1998-4 ; Aleiner:1999-1 ; Golubev:1999-1 ; Golubev:1999-2 . The heart of the debate, summarized for example in Mohanty’s recent letter Mohanty:1999-1 , is to determine whether the conventional theory of dephasing in Fermi liquids Altshuler:1982-1 ; Altshuler:1985-1 ; Chakravarty:1986-1 could explain the saturation of $`\tau _\varphi `$ (for example as an effect of an external microwave radiation Aleiner:1998-1 ; Cohen:1999-1 ) or whereas one should reconsider the theory completely Golubev:1998-1 ; Golubev:1998-2 ; Golubev:1999-3 . Although such an excitement is a strong motivation for working on decoherence in disordered mesoscopic conductors, we shall present a model for studying the decay of Schrődinger cat states in 1D ballistic conductors. From a methodological point of view, our line of thought is very close to the one used in atomic physics, for example in theoretical works Davidovich:1996-1 ; Raimond:1997-1 on decoherence experiments in cavity QED Brune:1996-1 . It relies heavily on the use of simple exactly solvable models of decoherence, such as the Caldeira-Leggett model CL:1983-2 . We think that this point of view brings a different light on the question of decoherence by taking into account the electron fluid as a whole and studying the coupling of this many-body system to the external reservoirs. It makes a bridge between atomic physics situations, where decoherence is extremely well controlled and for which simplified decoherence models apply almost directly and mesoscopic conductors which are complex interacting systems. In particular, this point of view is well suited to the 1D case because of interactions. In a non-Fermi liquid, one cannot keep track of an individual electron because of orthogonality catastrophe effects. Moreover, electromagnetic or acoustic radiation emitted by one part of the system and absorbed by another part may have drastic effect on the strongly correlated electron state. These remarks motivated our method, based on the study of the coupling of an external environment to the low energy excitations of the 1D electron system. Following general ideas of Stern, Aharonov and Imry Stern:1990-1 , we have computed the decoherence rate of Schrődinger cat states built from localized excitations (e.g Luttinger fermions) introduced at different places in the system or at the same place but moving in different directions (left and right moving components). As expected, we have found that such linear superpositions decay into statistical mixtures even at zero temperatures and computed various decoherence scenarii. This is the main result of this paper and it means that a 1D pure ballistic conductor exhibits decoherence at absolute zero in the sense of Schrődinger cat states decay. One dimensional conductors in the ballistic regime are appropriately described by an effective interacting theory: the Luttinger liquid Haldane:1981-1 . In this effective theory, interactions between electrons are put by means of an electrostatic short-ranged potential. Obviously, within the approximation of a linear dispersion relation for particle-hole excitations, the Luttinger effective theory contains no source for decoherence. A coupling to an external quantum environment is necessary to introduce decoherence in the Luttinger liquid. The quantized electromagnetic field and a 2D or 3D bath of longitudinal phonons will be considered in the present paper. The combined use of bosonization and non-equilibrium techniques make it possible to solve this problem. Within the bosonization framework, the pioneering work Loss:1993-1 by Martin and Loss investigates the question of equilibrium permanent currents induced by fluctuations of the quantized electromagnetic field. They have shown that coupling the Luttinger liquid to QED leads to a renormalization of the Luttinger liquid parameters. Our discussion of dynamical and therefore non-equilibrium aspects of the coupled Luttinger & QED system will show how this renormalization appears dynamically. Let us mention that coupling a Luttinger liquid to one dimensional phonons also renormalizes the Luttinger liquid parameters and drives a 1D Fermi liquid to a non Fermi liquid fixed pointLoss:1994-1 ; Martin:1995-1 ; Martin:1995-2 ; Eggert:1996-1 . But a 1D phonon bath does not introduce any intrinsic decoherence in the Luttinger liquid since, roughly speaking, it does not have enough modes. That’s why 2D and 3D phonon baths are considered in this paper and we have shown that these baths have enough modes to kill Schrődinger cat states. To be more precise on the results presented in this paper, we have shown that the electromagnetic decoherence time is much larger, although not infinite, than the natural time associated with the Luttinger system. This is mainly due to the weakness and the transversality of the coupling between photons and the Luttinger liquid. Therefore the coherent Luttinger liquid paradigm is not ruled out by its coupling to QED. We have also shown that the acoustic decoherence is much stronger than the electromagnetic one. This difference comes from the fact that many bosonic modes of the Luttinger liquid have an efficient acoustic radiation rate whereas only very few modes dominate the decoherence process in the electromagnetic case. In the acoustic case, decoherence takes place over a much shorter time scale than dissipation contrarily to the electromagnetic case. This paper is organized as follows: the model is presented and the Feynman-Vernon and Keldysh basic tools are briefly recalled in section II. Section III makes contact with the Quantum Brownian Motion problem (QBM). The central problem of mutual decoherence of Schrődinger cat states is addressed in section IV. Results are summarized and possible extensions are discussed in the conclusion. All along the paper, the electromagnetic and acoustic cases are discussed separately. Technical details are gathered in appendices. ## II Electron systems coupled to external reservoirs (QED or phonons) ### II.1 The Feynman-Vernon-Keldysh method Studying an out of equilibrium quantum system boils down to the computation of its density matrix as a function of time. A functional integral approach to this problem has been given long time ago in the context of perturbative field theory by Keldysh Keldysh:1965-1 . Let $`\rho (t)`$ denote the density operator for a closed system at time $`t`$ (in Schrődinger’s picture) and $`U[t_i,t_f]`$ denote the evolution operator between $`t_i`$ and $`t_f`$, then $`\rho (t_f)=U[t_i,t_f]\rho (t_i)U[t_i,t_f]^1`$. The first of these operators takes into account what will be called the “forward time branch” and the other one the “backward time branch” of the evolution. Evaluating matrix elements of $`\rho (t_f)`$ can be done with an appropriate propagator which can be represented as a double functional integral (one for each evolution operator). Correlation functions are generated by introducing external sources coupled to the system’s degrees of freedom which have different values on the forward and backward time branches Cugliandolo:1998-1 . Then, one usually takes a trace at time $`t_f`$. The generating functional obtained this way is called the Keldysh generating functional and can be represented as a path integral over a special contour $`K`$ that goes from $`t_i`$ to $`t_f`$ and then back to $`t_i`$. Denoting by $`\phi `$ (respectively $`\xi `$) fields (respectively external sources) describing the dynamics of the system, and by $`\phi _+`$ and $`\phi _{}`$ their restrictions to the upper and lower branch of Keldysh’s contour (respectively $`\xi _+`$, $`\xi _{}`$), we have: $$Z[\xi _+,\xi _{}]=𝒟[\phi _+,\phi _{}]e^{i(S[\phi _+,\xi _+]S[\phi _{},\xi _{}])}\rho [\phi _+|_{t_i},\phi _{}|_{t_i}]$$ where $`\rho (t_i)[\phi _+|_{t_i},\phi _{}|_{t_i}]`$ denotes the kernel of the initial density operator. This is the basis for Keldysh’s non equilibrium perturbation theory. Because of the doubling of degrees of freedom, one obtains a 2 by 2 matrix Green function which can be related to well known Green functions by: $$𝑮(x,y)=\left(G_{ϵϵ^{}}(x,y)\right)_{(ϵ,ϵ^{})\{1,1\}}=\left(\begin{array}{cc}G_T(x,y)& G_<(x,y)\\ G_>(x,y)& G_{\stackrel{~}{T}}(x,y)\end{array}\right)$$ (1) Here $`T`$ denotes the usual time ordering, $`\stackrel{~}{T}`$ the anti-chronological time ordering, $`<`$ and $`>`$ the lesser and upper time orderings Mahan . For bosonic oscillators initially in thermal equilibrium, Green’s functions are explicitly known (see appendix A). For a matter system coupled to QED, we are interested in the evolution of the reduced density matrix for the matter system. Bosonization techniques enable us to treat the electron fluid through an effective bosonic theory. Our strategy will then be to integrate over the electromagnetic field degrees of freedom and to deal with matter degrees of freedom in a second time. It is convenient to choose the Coulomb gauge, in which dynamical degrees of freedom of the quantum electromagnetic field (transverse photons) are decoupled from the Coulomb interaction. The latter is taken into account through the effective action for the matter system which uses only instantaneous interactions such as the Luttinger liquid description. Integrating out transverse photons gives a non-local functional integral of the current density, called a Feynman-Vernon influence functional Feynman:1963-1 . For the sake of simplicity, we shall assume that these two systems are independent at initial time (the initial density operator factorizes $`\rho (t_i)=\rho _{\mathrm{Matter}}\rho _{\mathrm{QED}}`$) and that the electromagnetic field is initially at equilibrium at temperature $`T=1/k_B\beta `$. In the present case, using bold symbols for vectors and matrices, the influence functional is nothing but the influence functional for the current density: $$[𝒋_+,𝒋_{}]=𝒟[𝑨_+,𝑨_{}]e^{i(S[𝑨_+]S[𝑨_{}])}\times e^{i{\scriptscriptstyle (𝑨_+𝒋_+𝑨_{}𝒋_{})}}\times \rho _\beta [𝑨_+(t_i),𝑨_{}(t_i)]$$ (2) where $`\rho _\beta `$ is the appropriate functional kernel representing the density operator for the quantum electromagnetic field at given temperature. The functional integral 2 can be evaluated in terms of Keldysh’s Green function for the electromagnetic field at finite temperature $`D_{ϵϵ^{}}^{\alpha \beta }(x,y)=iT_KA_ϵ^\alpha (x)A_ϵ^{}^\beta (y)`$ and is equal to: $$[𝒋_+,𝒋_{}]=\mathrm{exp}\left(\frac{i}{2}\underset{(ϵ,ϵ^{})\{1,1\}}{}ϵϵ^{}d^4xd^4y𝒋_ϵ^\alpha (x)D_{ϵϵ^{}}^{\alpha \beta }(x,y)𝒋_ϵ^{}^\beta (y)\right)$$ (3) Noise and dissipation kernels are defined by: $`\nu ^{\alpha \beta }(x,y)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{}(D_T^{\alpha \beta }(x,y))`$ (4) $`\eta ^{\alpha \beta }(x,y)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(D_R^{\alpha \beta }(x,y)D_A^{\alpha \beta }(x,y))`$ (5) and we have: $`[𝒋_+,𝒋_{}]`$ $`=`$ $`\mathrm{exp}({\displaystyle }d^4xd^4y{}_{}{}^{t}(𝒋_+(x)𝒋_{}(x)).𝝂(x,y).(𝒋_+(y)𝒋_{}(y)))`$ (6) $`\times `$ $`\mathrm{exp}(i{\displaystyle }d^4xd^4y{}_{}{}^{t}(𝒋_+(x)𝒋_{}(x)).𝜼(x,y).(𝒋_+(y)𝒋_{}(y)))`$ (7) Correlation functions can be obtained from a generating functional which takes the environment into account: $$𝒟[\phi _+,\phi _{}]e^{i(S[\phi _+,\xi _+]S[\phi _{},\xi _{}])}[𝒋[\phi _+],𝒋[\phi _{}]]\rho (t_i)[\phi _+(t_i),\phi _{}(t_i)]$$ Within the elastic approximation, coupling to acoustic phonons can be treated along the same line. Longitudinal phonons create a potential $`V(𝒙,t)`$ such that $`eV(𝒙,t)=D\rho (𝒙,t)\mathrm{div}(𝒖(𝒙,t))`$ where $`𝒖(𝒙,t)`$ is the elastic deformation field, $`\rho (𝒙,t)`$ is the electric charge density and $`D`$ denotes the electron-phonon coupling energy. Longitudinal phonon dynamics is described by a quadratic action: $$S[𝒖(𝒙,t)]=\frac{\rho _M}{2}\left((_t𝒖)^2c_S^2(\mathrm{div}(𝒖))^2\right)d^3𝒙𝑑t$$ (8) where $`c_S`$ denotes the sound velocity and $`\rho _M`$ the volumic mass. The resulting phonon influence functional is a Gaussian in terms of $`\rho _\pm (x)`$ with kernel $`𝒟_{ϵ,ϵ^{}}^{\alpha ,\beta }(x,y)=i𝒖_ϵ^\alpha (x)𝒖_ϵ^{}^\beta (y)_\beta `$: $$[\rho _+,\rho _{}]=\mathrm{exp}\left(\frac{iD^2}{2}d^4xd^4y\underset{(ϵ,ϵ^{})\{+,\}^2}{}\rho _ϵ(x)\rho _ϵ^{}(y)(_{x,\alpha }_{y,\beta }𝒟_{ϵ,ϵ^{}}^{\alpha ,\beta })(x,y)\right)$$ (9) In a generic electron system, this influence functional a priori contains quartic fermion terms. The one dimensional interacting electron gas Luttinger:1963-1 is quite interesting since it can be described by an effective free bosonic theoryHaldane:1981-1 . Neglecting environment-induced umklapp processes, the influence functional is Gaussian (although non local) in terms of the bosonic field. Such terms arise from the introduction of $`2k_F`$ components of the charge or current densities in the environment’s influence functional. There is a factor $`\mathrm{exp}(2ik_F(\sigma +\sigma ^{}))`$ in front of the term involving two $`\psi _R^{}\psi _L`$s. At incommensurate filling, the usual averaging argument can be invoked to rule out these Umklapp contributions. Other terms contain both $`\psi _R^{}\psi _L`$ and $`\psi _L^{}\psi _R`$ and would imply momentum transfer of order $`2k_F`$ (except for special values of filling). In QED’s case, $`2k_F`$ photons would have a very high energy compared to the Luttinger typical energy $`\mathrm{}v_S/L`$. In the acoustic case, since $`k_FL>>v_S/c_S`$, it would also be the case. That’s why we will not take into account $`2k_F`$ components of charge and current densities. Thanks to a Fourier mode analysis, this problem boils down to a set of independent harmonic oscillators, each of them being linearly coupled to a bath of quantum oscillators. This problem is known under the name of Quantum Brownian Motion (QBM) and has widely been studied CL:1983-1 ; CL:1983-2 . Necessary results will be recalled in section III. Elementary excitations of the Luttinger liquid can be created using vertex operators. States created this way are nothing but coherent states. Since the evolution of coherent states in QBM can be exactly computed CL:1985-1 , so can the evolution of elementary excitations in the Luttinger liquid. ### II.2 Mode decomposition for the Luttinger liquid Let us consider a Luttinger liquid on a circle of length $`L`$, with both left and right chiralities. As in Degiovanni:1998-1 , $`\sigma `$ denotes the coordinate along the circle. Low energy excitations of the Luttinger liquid are described by the theory of a free compactified boson, whose action is given by $$S[\phi ]=\frac{g}{2\pi }𝑑\sigma 𝑑t(v_S^1(_t\phi )^2v_S(_\sigma \phi )^2)$$ (10) The field $`\phi `$ is compactified on a circle of radius $`R_c`$, and the Euclidean path integral representing the finite temperature equilibrium partition function contains a topological term taking into account zero mode quantization. The space of states of this free bosonic theory is a representation of a $`\widehat{U(1)}_R\times \widehat{U(1)}_L`$ algebra (two commuting copies of an oscillator algebra): $`[J_k,J_l]=k\delta _{k,l}\mathbf{\hspace{0.17em}1}`$ and $`[\overline{J}_k,\overline{J}_l]=k\delta _{k,l}\mathbf{\hspace{0.17em}1}`$. In Wen’s work Wen:1990-1 , $`J_l`$ and $`\overline{J}_l`$ for $`l0`$ are called hydrodynamic modes, whereas the $`J_0`$ and $`\overline{J}_0`$ are called zero modes. In term of these modes, the Hamiltonian of the Luttinger system is given by: $$H_{\mathrm{tot}}=\frac{\pi v_S}{L}\underset{l}{}(J_lJ_l+\overline{J}_l\overline{J}_l).$$ (11) The finite temperature equilibrium partition function with a magnetic flux $`\chi e/h`$ is given by: $$Z_{\mathrm{Lutt}}(\chi )=\frac{1}{|\eta (\tau )|^2}\underset{\genfrac{}{}{0pt}{}{(n,m)(/2)\times }{2nm(mod2)}}{}q^{\frac{1}{2}p_{n,m+2\chi }^2}\overline{q}^{\frac{1}{2}\overline{p}_{n,m+2\chi }^2},$$ (12) where $`\alpha =gR_c^2`$ encodes the interaction strength ($`\alpha =1`$ for free electrons) and $$p_{n,m}=n\sqrt{\alpha }+\frac{m}{2\sqrt{\alpha }}\overline{p}_{n,m}=n\sqrt{\alpha }\frac{m}{2\sqrt{\alpha }}.$$ (13) States of interest will be $`\widehat{U(1)}\times \widehat{U(1)}`$ highest weight states: with $`J_0`$ (resp. $`\overline{J}_0`$) eigenvalues $`p_{n,m}`$ (resp. $`\overline{p}_{n,m}`$). The charge and current densities can be expressed in terms of the $`\widehat{U(1)}`$ modes as follows: $`\rho (\sigma ,t)`$ $`=`$ $`{\displaystyle \frac{e}{L\sqrt{\alpha }}}{\displaystyle \underset{n}{}}\left(J_n(t)e^{2\pi i\frac{n}{L}\sigma }+\overline{J}_n(t)e^{2\pi i\frac{n}{L}\sigma }\right)`$ (14) $`j(\sigma ,t)`$ $`=`$ $`{\displaystyle \frac{ev_S}{L\sqrt{\alpha }}}{\displaystyle \underset{n}{}}\left(J_n(t)e^{2\pi i\frac{n}{L}\sigma }\overline{J}_n(t)e^{2\pi i\frac{n}{L}\sigma }\right).`$ (15) The $`|n,m_\chi =|p_{n,m+2\chi },\overline{p}_{n,m+2\chi }`$ state carries a charge $`2n`$ and a current $`(m+2\chi )v_S/L\alpha `$ Loss:1992-1 ; Degiovanni:1998-1 . If we assume that our Luttinger system is realized using the edge states of a FQH fluid, then the two edges of the sample are polarized with charges ($`\alpha =1/\nu `$ is an odd integer): $$q_R=e\left(n+\frac{m}{2\alpha }\right)\mathrm{and}q_L=e\left(n\frac{m}{2\alpha }\right).$$ (16) Obviously, $`\widehat{U(1)}`$ descendants carry the same global charges and current than the highest weight states. Remember that $`\widehat{U(1)}`$ primaries are created using vertex operators $`V_{n,m}(\sigma ,t)`$ which are normal ordered exponentials of bosonic modes Kac ; FLMbis . To complete this brief description, let us recall that the original fermionic operators get renormalized (orthogonality catastrophe historically introduced by Anderson Anderson:67-1 ), and that the corresponding renormalized fields are the so-called Luttinger “fermions” which correspond to $`n=\pm 1/2`$ and $`m=\pm 1`$. Within the context of fractional quantum Hall effect on a cylinder – or an annulus – edge fermions carrying a charge localized on one of the two edges, can be created or destroyed using vertex operators Stone:94-1 with $`n=\pm 1/2`$ and $`m=\pm \nu ^1`$. Typical experiments which may be performed on edge excitations of an annular 2DEG in an AsGaAl heterostructure lead to the following numerical values which we shall use in the rest of this paper : $`v_S3.10^5\mathrm{m}\mathrm{s}^1`$, $`L30\mu \mathrm{m}`$, $`\alpha =3`$ and $`T010K`$. ### II.3 Coupling of modes to the environments In order to respect cylindrical geometry, the environment field is supposed to live in a cylindrical cavity with the same revolution axis than the Luttinger circle. The cavity is of radius $`R`$ and of height $`h`$ and $`𝒱=\pi R^2h`$ will denote its volume. Periodic boundary conditions are assumed in this vertical direction. Electromagnetic modes are quantized using their momentum along $`z`$ axis $`k_z`$, their angular momentum moment around $`z`$ axis $`l`$ and the number $`n`$ of radial zeroes of the electric field. The boundary condition on the cavity plays the role of polarization and accounts for the helicity degeneracy. Transverse electric ($`TE`$) modes have their (transverse) electric field orthogonal to the cavity’s edge, and transverse magnetic ($`TM`$) modes have their magnetic field orthogonal to the cavity’s edge. Normalized expressions for the modes can be found in any book on optical fibers and are recalled in appendix B. Of course, the box is considered as large meaning that electromagnetic modes form a continuum compared to the Luttinger modes. Phonons are quantized according to the vanishing of displacement field $`𝒖`$ on the cavity’s boundary. Longitudinal modes can be written as a gradient $`𝒖=\mathbf{}(\phi )`$. As shown in appendix B, longitudinal modes are classified by $`(k_z,l)`$ and $`n`$ where $`n`$ is the number of zeroes of the $`\phi `$ potential. It is convenient to introduce some kind of “phase space coordinates” to describe the $`l1`$ modes of the Luttinger Liquid : $`q_l^{(+)}`$ $`={\displaystyle \frac{1}{2\sqrt{l}}}\left(J_l+J_l\overline{J}_l\overline{J}_l\right)\mathrm{and}ip_l^{(+)}={\displaystyle \frac{1}{2\sqrt{l}}}\left(J_lJ_l\overline{J}_l+\overline{J}_l\right)`$ (17) $`iq_l^{()}`$ $`={\displaystyle \frac{1}{2\sqrt{l}}}\left(J_lJ_l+\overline{J}_l\overline{J}_l\right)\mathrm{and}p_l^{()}={\displaystyle \frac{1}{2\sqrt{l}}}\left(J_l+J_l+\overline{J}_l+\overline{J}_l\right)`$ (18) Formulating the problem in this way shows that the interaction Hamiltonian is nothing but a linear coupling between harmonic oscillators ($`qx`$ coupling). More precisely, for each $`l1`$, two harmonic oscillators are coupled to the acoustic or electromagnetic modes of angular momentum $`l`$. The two Luttinger modes of fixed $`l`$ are coupled to different sets of environment modes. The same conclusion is true for zero modes: the global current couples to $`l=0`$ electromagnetic modes whereas the global charge decouples from the field’s propagating modes (Coulomb gauge effect). In the acoustic case, only the total electric charge couples to $`l=0`$ acoustic modes. The influence functional for each of these environments can explicitly be obtained in a nice form. First of all, the charge and current densities can be expressed as linear combinations of the above phase space coordinates. Then, performing integration over the vector potential and spatial coordinates leads to the QED functional (assuming that $`q_0^{(+)}=q_0`$ and $`q_0^{()}=0`$): $$_{FV}=\mathrm{exp}\left(\frac{ig}{2}\frac{L^3}{𝒱}\underset{\genfrac{}{}{0pt}{}{I}{l_I0}}{}\underset{(ϵ,ϵ^{},\alpha )\{+,\}^3}{}ϵϵ^{}\frac{c^2}{L^2}𝒟_I^2l_I𝑑t𝑑sq_{l_I}^{(\alpha )ϵ}(t)G_{ϵϵ^{}}^{(\omega _I)}(ts)q_{l_I}^{(\alpha )ϵ^{}}(s)\right)$$ (19) where the dimensionless coupling constant only depends on the fine structure constant $`\alpha _{QED}`$, the Luttinger liquid interaction parameter $`\alpha `$ and the ratio of velocities $`v_S/c`$: $$g=4\pi \frac{\alpha _{QED}}{\alpha }.\left(\frac{v_S}{c}\right)^2$$ (20) This mode expansion is related to a multipolar expansion. Luttinger modes of momentum $`l`$ contribute to the electric (resp. magnetic) multipolar expansion starting at order $`2l`$ (resp. $`2l+1`$) In particular, $`l=0`$ is mainly a magnetic dipole and an electric monopole (corresponding to the global charge and current around the Luttinger ring). The $`l=1`$ modes contribute first to the electric dipole and the magnetic quadrupole. Longitudinal phonons can be treated in the same way. In this case, one needs to perform a canonical transformation on the Luttinger phase space coordinates, exchanging $`p_l^{(\pm )}`$ and $`q_l^{(\pm )}`$ in order to obtain a Gaussian expression similar to 19. The dimensionless coupling constant corresponding to longitudinal phonons in dimension $`d=2,3`$ is given by: $$g_{ph}(L)=\frac{D^2}{\alpha \rho _ML^{d1}\mathrm{}c_S^3}$$ (21) where $`c_S`$ is the sound velocity, $`\rho _M`$ the volumic ($`d=3`$) or surfacic ($`d=2`$) mass and $`D`$ the typical electron/phonon coupling energy. With typical values $`D7eV`$, $`c_S3.10^3ms^1`$ and $`\rho _M3.10^3kgm^3`$, we get, in dimension three $`g_{ph}(L)4.10^6`$. ## III Relation to the Quantum Brownian motion ### III.1 The Quantum Brownian Motion model This basic model consists in a single quantum harmonic oscillator (called “the system”) coupled to a bath of oscillators (called the “reservoir”): $$L=\frac{m}{2}\left(\dot{q}^2\mathrm{\Omega }^2q^2\right)+\underset{I}{}C_Iqq_I+\underset{I}{}\frac{M_I}{2}\left(\dot{q_I}^2\omega _I^2q_I^2\right)$$ (22) We assume that these two systems are initially independent and that the oscillator bath is at equilibrium with inverse temperature $`\beta `$. One could also assume that the whole system is initially at equilibrium but it makes things more involved without illuminating the discussion. Interested readers are refered to Grabert’s review Grabert:1988-1 , to Hakim and Ambegaokar’s paperHakim:1985-1 or to Romero and Paz Romero:1996-1 . The question addressed here is to compute the time evolution of the system. The physics depends on the reservoir’s influence at each frequency, encoded in appropriate spectral densities. Therefore, several natural time scales will appear in the Luttinger & environment problem: * The cut-off time : above a certain UV cutoff, the effective description by a Luttinger liquid is no longer valid. Another natural cutoff could also be provided by the spectral distribution itself (Debye frequency for phonons). For this reason, an UV cutoff is present in the model. The associated time will be denoted by $`1/\mathrm{\Lambda }`$. Below this time scale, our simple model cannot be considered as valid. * The environment time scale $`\tau _E`$ is the time needed by the light ($`\tau _{EM}=L/c`$) or by phonons ($`\tau _S=L/c_S`$) to circle around the system. Spectral densities and influence functionals can be normalized with respect to this characteristic time (they only depend on $`\omega \tau _E`$). The low frequency regime is defined by $`\omega \tau _E<<1`$ and the high frequency regime by $`\omega \tau _E>>1`$. * The Luttinger time scale $`\tau _L=L/v_S`$ is the time needed by one excitation of the Luttinger liquid to circle around the system. This is the natural time scale from the Luttinger liquid point of view. * The thermal time scale $`\tau _{Th}=\mathrm{}/k_BT`$ is the inverse frequency associated with temperature $`T`$. The corresponding thermal length is $`l(\beta )=c\tau _{Th}`$ for photons and $`l(\beta )=c_S\tau _{Th}`$ for phonons. Let us notice that at $`T1\mathrm{K}`$, $`l(\beta )=1\mathrm{mm}`$ for photons whereas it is of order 10 $`\mu \mathrm{m}`$ for phonons. The general solution to the Quantum Brownian Motion problem is due to Hu, Paz and Zhang (HPZ) HPZ:1992-1 who computed the evolution kernel for the system’s reduced density matrix. Caldeira and Leggett’s work CL:1983-2 is concerned with a special case of reservoir, corresponding to the so-called Ohmic spectral density $`𝒥(\omega )\omega `$. As we have already noticed, the Luttinger liquid can be seen as a collection of harmonic oscillators and some zero modes. As we shall see, each harmonic mode coupled to the environment is a QBM problem. Each reservoir is characterized by a set of spectral functions (one for each $`l0`$), which are made dimensionless for simplicity. In the electromagnetic case, using expressions 94 and 95 in appendix B, we obtain: $$𝒥_l(\omega )=g\frac{L^3}{𝒱}\underset{I/l_I=l}{}\frac{l_I𝒟_I^2}{2\omega _I\tau _{EM}}\delta (\tau _{EM}(\omega \omega _I))$$ (23) The $`l=0`$ case needs a slight modification: $$𝒥_0(\omega )=g\frac{L^3}{𝒱}\underset{\genfrac{}{}{0pt}{}{ITE}{l_I=0}}{}\frac{𝒟_I^2}{2\omega _I\tau _{EM}}\delta (\tau _{EM}(\omega \omega _I))$$ (24) The electromagnetic influence functional can be rewritten as: $$\begin{array}{c}_{FV}[𝒋_+,𝒋_{}]=\mathrm{exp}\left(i\frac{c^2}{L^2}\underset{l=0}{\overset{\mathrm{}}{}}\underset{\alpha =+,}{}_{t_i}^{t_f}𝑑t_{t_i}^t𝑑s(q_l^{(\alpha )+}q_l^{(\alpha )})(t)\eta _l(ts)(q_l^{(\alpha )+}+q_l^{(\alpha )})(s)\right)\hfill \\ \hfill \times \mathrm{exp}\left(\frac{c^2}{L^2}\underset{l=0}{\overset{\mathrm{}}{}}\underset{\alpha =+,}{}_{t_i}^{t_f}𝑑t_{t_i}^t𝑑s(q_l^{(\alpha )+}q_l^{(\alpha )})(t)\nu _l(ts)(q_l^{(\alpha )+}q_l^{(\alpha )})(s)\right)\end{array}$$ (25) Dimensionless dissipation and noise kernels for each mode are expressed in terms of the spectral density by: $`\eta _l(s)`$ $`={\displaystyle \frac{L}{c}}{\displaystyle _0^{\mathrm{}}}𝑑\omega 𝒥_l(\omega )\mathrm{sin}(\omega s)`$ (26) $`\nu _l(s)`$ $`={\displaystyle \frac{L}{c}}{\displaystyle _0^{\mathrm{}}}𝑑\omega 𝒥_l(\omega )\mathrm{coth}\left({\displaystyle \frac{\beta \omega }{2}}\right)\mathrm{cos}(\omega s)`$ (27) An acoustic reservoir leads to similar expressions for spectral functions (using expressions 96 and 97): $$𝒥_l(\omega )=g_{Ph}(L)\frac{L^3}{𝒱}\underset{I/l_I=l}{}l_I𝒩_I^2(\tau _S\omega _I)^3\delta (\tau _S(\omega \omega _I))$$ (28) The $`l=0`$ case needs a slight modification: $$𝒥_0(\omega )=g_{Ph}(L)\frac{L^3}{𝒱}\underset{I/l_I=0}{}𝒩_I^2(\tau _S\omega _I)^3\delta (\tau _S(\omega \omega _I))$$ (29) Numerical computations of spectral densities as well as analytic estimates of their asymptotics are available (see appendix D). The electromagnetic case is illustrated on figure 3. In this case, all $`l1`$ modes are supraohmic ($`𝒥_l(\omega )`$ decreases faster than $`\omega `$) at low frequency. The $`l=1`$ modes show an ohmic behavior ($`𝒥_1(\omega )\omega `$). In this case, the dissipation kernel is local in time and, as we shall see in section IV.6, an effective Caldeira-Leggett model can be used to perform analytic computations. In the acoustic case $`𝒥_l(\omega )`$ goes as $`(\omega \tau _S)^{2l+d}`$ in the low frequency regime and as $`(\omega \tau _S)^{d1}`$ at higher frequencies. The main difference between the electromagnetic and acoustic reservoirs is that the natural Luttinger frequency $`2\pi /\tau _L`$ falls in the low frequency domain for QED, whereas it does not for phonons since $`\tau _S/\tau _L>>1`$. ### III.2 Phase space evolution using Wigner functions The time evolution can be computed using the Wigner function associated with the system’s density operator. This form is especially adapted to the study of decoherence of Gaussian wave packets. Moreover, it provides a nice quasi-classical insight on the evolution of the system since in the Luttinger liquid, charge and current density fluctuations play the role of “phase space coordinates” for the hydrodynamic modes and within the FQH effect framework, encode the shape of the incompressible quantum Hall fluid droplet Wen:1990-1 . The Wigner function associated with an operator $`B`$ is defined by: $$W_B(p,q)=𝑑ye^{ipy/\mathrm{}}q\frac{y}{2}|B|q+\frac{y}{2}$$ (30) We use the following notation for phase space: $`\varphi =\left(\begin{array}{c}p\\ q\end{array}\right)`$. The evolution kernel for the Wigner function can be computed from the density operator evolution kernel and is given by: $$J_W(p,q,t|p_0,q_0,0)=𝒩(t)\mathrm{exp}(\frac{1}{4}{}_{}{}^{t}(N.\varphi +N_0.\varphi _0).A(t).(N.\varphi +N_0\varphi _0))$$ (31) where the $`N_0`$ and $`N`$ time dependent matrices are given by: $$N=\left(\begin{array}{cc}1& \dot{u}_2(t)\\ 0& \dot{u}_2(0)\end{array}\right)\mathrm{and}N_0=\left(\begin{array}{cc}0& \dot{u}_1(t)\\ 1& \dot{u}_1(0)\end{array}\right)$$ (32) The $`u_i(t)`$ ($`i\{1,2\}`$) functions are defined in Hu, Paz and Zhang’s paper HPZ:1992-1 as solutions to the classical equations of motion with dissipation and boundary conditions $`(u_1(0),u_1(t))=(1,0)`$ and $`(u_2(0),u_2(t))=(0,1)`$. The $`A(t)=(a_{i,j}(t))_{(i,j)\{1,2\}^2}`$ matrix is defined by $$a_{i,j}(t)=\frac{1}{2}_0^t𝑑s_1_0^t𝑑s_2u_i(s_1)\nu (s_1s_2)u_j(s_2).$$ The reader can check that, when the reservoir decouples, the evolution kernel reduces to a delta function, giving back the classical evolution in phase space. Henceforth, we shall call $`N^1N_0`$ the “Wigner evolution operator” and denote it by $`U_t`$. When turning on the coupling to the environment, the classical delta distribution in phase space spreads into a Gaussian, the center of which moves according to $`U_t`$. As we shall now see, the evolution kernel encodes all effects of dissipation and decoherence. The evolution of a density operator built from Gaussian wave packets can be computed exactly. The generic form of a Gaussian Wigner function is, up to some normalization: $$𝒲[\overline{\varphi },K,Q](\varphi )=\mathrm{exp}(\frac{1}{2}{}_{}{}^{t}(\varphi \overline{\varphi }).Q.(\varphi \overline{\varphi })+i{}_{}{}^{t}K.\varphi )$$ (33) The 2 by 2 matrix $`Q`$ encodes the spreading of the packet, $`\overline{\varphi }`$ is the center of the packet and $`K`$ a phase modulation. Gaussian wave packets and in particular coherent states lead to Gaussian Wigner functions. A Gaussian Wigner function remains Gaussian. If $`W(t=0)=𝒲[\varphi _0,K_0,Q_0]`$ then, after time $`t`$, $$W(t)=e^{d[K_0,Q_0](t)}\times 𝒲[U_t.\varphi _0,K(t),Q(t)]$$ (34) where the basic parameters at time $`t`$ are given by $`d[K_0,Q_0](t)={}_{}{}^{t}K_{0}^{}.D(t).K_0`$ and: $`D(t)`$ $`=`$ $`\left(2Q_0+{}_{}{}^{t}N_{0}^{}A(t)^1N_0\right)^1`$ (35) $`K(t)`$ $`=`$ $`{}_{}{}^{t}U_{t}^{1}.(\mathrm{𝟏}2Q_0D(t)).K_0`$ (36) $`Q(t)`$ $`=`$ $`{}_{}{}^{t}U_{t}^{1}.\left(Q_02Q_0D(t)Q_0\right).U_t^1`$ (37) The $`D(t)`$ matrix contains the decoherence effect. To understand this, let us start with a coherent superposition of two Gaussian wave packets $`(\psi _1+\psi _2)/\sqrt{2}`$. The initial Wigner function is given by: $$W(0)=\frac{1}{2}\left(W_{11}+W_{12}+W_{21}+W_{22}\right)\mathrm{where}W_{\alpha \beta }(0)=W[\overline{\varphi }_\alpha ,\overline{\varphi }_\beta ,Q_0]=𝒲[\frac{\overline{\varphi }_\alpha +\overline{\varphi }_\beta }{2},i\sigma ^y.(\overline{\varphi }_\alpha \overline{\varphi }_\beta ),Q_0]$$ Under time evolution, the form of the wave packet is preserved up to a global factor and a phase modulation: $$W_{\alpha \beta }(t)=e^{d_{\alpha \beta }(t)+i\mathrm{\Theta }_{\alpha \beta }(t,\varphi )}W[U_t.\overline{\varphi }_\alpha ,U_t.\overline{\varphi }_\beta ,Q(t)]$$ (38) where: $`d_{\alpha \beta }(t)`$ $`=`$ $`d[i\sigma ^y.(\varphi _\alpha \varphi _\beta ),Q_0]`$ (39) $`\mathrm{\Theta }_{\alpha \beta }(t,\varphi )`$ $`=`$ $`{}_{}{}^{t}K_{q}^{}(t).(\varphi \overline{\varphi }_{12}(t))`$ (40) $`K_q(t)`$ $`=`$ $`K(t){\displaystyle \frac{\dot{u}_1(t)}{\dot{u}_2(0)}}{}_{}{}^{t}U_{t}^{1}.K_0`$ (41) Each density operator $`|\psi _\alpha \psi _\alpha |`$ has its own evolution, described by $`\overline{\varphi }_\alpha (t)=U_t.\overline{\varphi }_\alpha `$ and $`Q(t)`$. Note that $`d_{11}(t)=d_{22}(t)=0`$. The coherence part is contained in $`W_{12}`$, and evolves according to $`U_t.\overline{\varphi }_{12}`$ and $`Q(t)`$ plus an exponential factor $`e^{d_{12}(t)+i\mathrm{\Theta }_{12}(t,\varphi )}`$. The $`d_{12}(t)`$ factor gives the attenuation of off-diagonal correlations and should therefore be interpreted as the decoherence factor between the two wave packets. Unfortunately, as noticed before, explicit and closed expressions for $`U_t`$ and $`D(t)`$ are not known for a general supra-ohmic environment. Therefore, we shall perform a perturbative expansion in the coupling constant. In the case of the decoherence factor $`d_{12}(t)`$, it is enough to start from standard harmonic oscillators expressions for the $`u_1`$ and $`u_2`$ functions since $`A(t)`$ is of first order in the coupling constant. The decoherence matrix can then be expressed as an integral over the spectral density: $$D(t)=_0^+\mathrm{}𝒥(\omega )D(\mathrm{\Omega },\omega ,t)\mathrm{coth}\left(\frac{\beta \mathrm{}\omega }{2}\right)𝑑\omega .$$ (42) where $`D(\mathrm{\Omega },\omega ,t)`$ has an explicit but involved expression. Let us discuss its asymptotics in various physically relevant limits: * At short times $`\mathrm{\Omega }t<<1`$, $`\mathrm{\Omega }`$ can be neglected provided the spectral density of the bath contains modes with frequencies much higher than $`\mathrm{\Omega }`$. Then, this regime is dominated by the high frequencies of the bath and: $$D(\mathrm{\Omega },\omega ,t)\left(\begin{array}{cc}\frac{t^2}{2}& \frac{\omega t^3}{4}\\ \frac{\omega t^3}{4}& \frac{t^4}{8}\end{array}\right)$$ (43) The last expression is only valid at very short times and contains non-Markovian effects. Even in the Caldeira-Leggett model, at very short times, all frequencies of the bath take part in the evolution of the system, leading to these non-Markovian effects. * When the condition $`\mathrm{\Omega }t<<1`$ is no longer valid, dissipation effects with exponential relaxation will appear through linear terms in $`t`$. We must of course assume that $`t`$ is much smaller than typical relaxation times. At growing times, linear terms will dominate oscillating resonant ones and provide a “Golden rule” estimate for the relaxation and decoherence times. This approximation will be used in the next section in order to evaluate the typical decoherence time of Schrődinger cat states in the Luttinger model. When $`t`$ reaches the dissipation time scale, this perturbative treatment breaks down. The long time regime of decoherence can however be computed in some cases using the Caldeira-Leggett model. ## IV Mutual decoherence of elementary excitations in a Luttinger liquid ### IV.1 Statement of the problem As recalled in section II.2, elementary excitations of the Luttinger liquid are created by vertex operators $`V_{n,m}(\sigma )`$. Such a state is not an eigenstate of the Hamiltonian but rather corresponds to the introduction of a “localized” excitation at point $`\sigma `$ around the circle. Let us consider a Schrődinger cat state build as a superposition of a right moving Luttinger fermion ($`n=1/2`$ and $`m=1`$) at different places around the circle: $$|\psi _{RR}(0)=\frac{1}{\sqrt{2}}\left(\psi _R^{}(\sigma _1)|0+\psi _R^{}(\sigma _2)|0\right)$$ (44) In an isolated system, such a state will evolve according to: $$|\psi _{RR}(t)=\frac{1}{\sqrt{2}}\left(\psi _R^{}(\sigma _1,t)|0+\psi _R^{}(\sigma _2,t)|0\right)$$ (45) Therefore, this coherent superposition will remain coherent, as it should in any isolated quantum system. Switching on the coupling to the quantum electromagnetic field or phonons changes the situation: according to general works on decoherence Unruh:1989-1 ; Zurek:1982-1 ; Unruh:1989-1 ; Zurek:1991-1 , we expect this Schrődinger cat to decohere into a statistical mixing of two states: one excitation at one position, or the excitation at the other position. Here, two questions will be addressed: what is the strength of the decoherence process and on which time scale does it take place? In the following, two cases will be considered: the $`R/R`$ Schrődinger cat, already presented in equation 44 and the $`R/L`$ Schrődinger cat, defined as: $$|\psi _{RL}(0)=\frac{1}{\sqrt{2}}\left(\psi _R^{}(\sigma _1)|0+\psi _L^{}(\sigma _2)|0\right)$$ (46) Practically, the case of zero modes is simpler and will be considered in the next paragraph. We shall then turn to the $`l0`$ modes in section IV.3. Decoherence time estimates for electromagnetic (resp. acoustic) reservoirs are given in IV.4 (resp. IV.5). An effective Caldeira-Leggett model will be used to deal with the long time behavior in sections IV.6 to IV.8. ### IV.2 Evolution of zero modes For the zero modes, the zero coupling evolution corresponds to a free particle (and not to an harmonic oscillator). The Ambegaokar and Hakim Hakim:1985-1 method (exact diagonalization for the coupled system) can be used to compute the evolution of the density matrix for the zero modes. The corresponding explicit formula in the Luttinger liquid case is, at finite temperature: $$n,m|\rho (t)|n^{},m^{}=n,m|\rho (0)|n^{},m^{}\times \mathrm{exp}\left(i(\omega _{n,m}(t)\omega _{n,m}(t))t\right)\times \mathrm{exp}\left(\frac{d(t)}{\alpha }(mm^{})^2\right)$$ (47) where $`\mathrm{}\omega _{n,m}(0)`$ denotes the energy of $`|n,m`$ in the isolated Luttinger system and: $`\omega _{n,m}(t)`$ $`=`$ $`\omega _{m,n}(0)+{\displaystyle \frac{\pi v_S}{2L\alpha }}m^2{\displaystyle _0^+\mathrm{}}{\displaystyle \frac{c}{\pi v_S}}𝒥_0(\omega )\left({\displaystyle \frac{\mathrm{sin}(\omega t)}{\omega t}}1\right){\displaystyle \frac{d\omega }{\omega }}`$ (48) $`d(t)`$ $`=`$ $`{\displaystyle \frac{t^2}{\tau _{EM}}}{\displaystyle _0^+\mathrm{}}𝑑\omega 𝒥_0(\omega )\mathrm{coth}\left({\displaystyle \frac{\beta \mathrm{}\omega }{2}}\right){\displaystyle \frac{1\mathrm{cos}(\omega t)}{(\omega t)^2}}`$ (49) The first term in 47 can be interpreted as a dynamical renormalization of $`v_S/\alpha `$. Only $`v_S/\alpha `$ is renormalized since the charge density does not couple to the transverse degrees of freedom of the electromagnetic field. To be precise, at $`t+\mathrm{}`$, the velocity and interacting parameters of the Luttinger liquid get renormalized as $`v_S^{}=v_S.\zeta `$ and $`\alpha ^{}=\alpha /\zeta `$ where the renormalization constant $`\zeta `$ is equal to the $`t+\mathrm{}`$ limit of: $$\zeta (t)^2=1+\frac{c}{\pi v_S}_0^+\mathrm{}𝒥_0(\omega )\left(\frac{\mathrm{sin}(\omega t)}{\omega t}1\right)\frac{d\omega }{\omega }$$ (50) The renormalization effect is of course strongly cut-off dependent. The dimensionless coupling constant appearing here is $`\frac{\alpha _{QED}}{\alpha }.\frac{v_S}{c}10^5`$. The second term is the decoherence coefficient between two different highest weight states $`|n,m`$ and $`|n^{},m^{}`$ of the LL. Decoherence takes place in a time of the order of the cutoff time $`\mathrm{\Lambda }^1`$ and then reaches saturation. Figure 4 summarizes $`d(t)`$ and $`\zeta (t)`$’s behavior. The typical value of the $`t+\mathrm{}`$ value of the decoherence exponent is of typical order $`g`$: $$d(+\mathrm{})=_0^+\mathrm{}\frac{c}{L\omega ^2}𝒥_0(\omega )\mathrm{coth}\left(\frac{\beta \mathrm{}\omega }{2}\right)𝑑\omega $$ (51) In the acoustic case, computations can be performed in the same way. Since acoustic zero modes couple to the total charge of the Luttinger system, the decoherence factor between states $`|n,m`$ and $`|n^{},m^{}`$ is found to be $`\mathrm{exp}(2\alpha d(t)(n^2(n^{})^2))`$ where $`d(t)`$ is obtained from 49 by using the acoustic spectral density and $`\tau _S`$ insted of their electromagnetic counterparts. The Luttinger parameters $`\alpha `$ and $`v_S`$ also get renormalized. In the acoustic case, only $`v_S\alpha `$ is renormalized. Then, $`\alpha ^{}=\alpha \zeta _{ph}`$ and $`v_S^{}=v_S\zeta _{ph}`$ where $`\zeta _{ph}`$ is the $`t+\mathrm{}`$ limit of $`\zeta _{ph}(t)`$, obtained by using the acoustic spectral density and the speed of sound in formula 50. In both cases, the final decoherence exponent is proportional to the square of the difference between the total current (resp. charge), quantities which measure the “distance” between the two quantum states. Such a result is expected since, as explained in C. Cohen Tannoudji’s lectures Cohen:CDF:1989 , such a dependence is common in the case of a linear coupling with a conserved quantity. In particular, a Schrődinger cat obtained by superposing the same elementary excitation of the Luttinger liquid at two different positions along the ring has all its decoherence due to hydrodynamic modes ! We also notive that zero mode decoherence has a weak dependence in the cut off and temperature. ### IV.3 Spatial dependence of decoherence Using the explicit form of vertex operators, one easily finds the relevant parameters to be used for the decoherence of each mode. Of course, these parameters depend on positions of each of vertex operator. Here, we shall only present the results for $`R/R`$ and $`R/L`$ Schrődinger cats ($`\sigma _{12}=\sigma _1\sigma _2`$): $`d_{RR}(t,\sigma _1,\sigma _2)`$ $`={\displaystyle \frac{4}{\alpha }}{\displaystyle \underset{l=1}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{1}{l}}(m^2D_{11}^{(l)}(t)+4n^2\alpha ^2D_{22}^{(l)}(t))\mathrm{sin}^2\left({\displaystyle \frac{\pi l\sigma _{12}}{L}}\right)`$ (52) $`d_{RL}(t,\sigma _1,\sigma _2)`$ $`={\displaystyle \frac{4}{\alpha }}{\displaystyle \underset{l=1}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{1}{l}}\{m^2D_{11}^{(l)}(t)(1\mathrm{sin}^2\left({\displaystyle \frac{\pi l\sigma _{12}}{L}}\right))`$ (53) $`+4n^2\alpha ^2D_{22}^{(l)}(t)\mathrm{sin}^2\left({\displaystyle \frac{\pi l\sigma _{12}}{L}}\right)+2nm\alpha D_{12}^{(l)}(t)\mathrm{sin}\left({\displaystyle \frac{2\pi l\sigma _{12}}{L}}\right)\}`$ (54) Here $`D^{(l)}(t)`$ denotes the decoherence matrix for the $`q_l^{(\pm )}`$ modes computed along the lines of section III.2. The main change from the HPZ computations arises from our normalization choice for the $`q_l`$s. The effective spectral density to be used in the HPZ formulas is given by: $$2\pi l\frac{v_S}{c}.\frac{𝒥_l(\omega )}{\tau _E^2}.$$ This rescaling takes into account the ratio of the Luttinger mode $`l`$ time scale (i.e. $`\tau _L/l`$) and of the environment time scale $`\tau _E=L/c`$. The appearance of an odd dependence – in term of $`\sigma _{12}`$ – in the $`d_{RL}(t,\sigma _1,\sigma _2)`$ coefficient is understood by noticing that an appropriate parity operation transforms the $`R/L`$ Schrődinger cat into an $`L/R`$ one. Therefore $`d_{RL}(t,\sigma _1,\sigma _2)`$ is invariant into simultaneous changes $`\sigma _1\sigma _2`$ and $`nmnm`$. A first estimate is obtained using a perturbative approach. We perform a secular approximation and retain only terms linear in time. The corresponding decoherence rates are given by: $`{\displaystyle \frac{\tau _L}{\tau _l^{(R/R)}(\sigma _1,\sigma _2)}}`$ $`=`$ $`8\pi ^2\mathrm{\Delta }_{n,m}.\mathrm{sin}^2\left({\displaystyle \frac{l\pi \sigma _{12}}{L}}\right){\displaystyle \frac{𝒥_l(\omega _l)}{\omega _l\tau _E}}`$ (55) $`{\displaystyle \frac{\tau _L}{\tau _l^{(R/L)}(\sigma _1,\sigma _2)}}=`$ $`=`$ $`8\pi ^2\mathrm{\Delta }_{n,m}.\left(1+{\displaystyle \frac{m^24n^2\alpha ^2}{m^2+4n^2\alpha ^2}}\mathrm{cos}\left({\displaystyle \frac{2\pi l\sigma _{12}}{L}}\right)\right){\displaystyle \frac{𝒥_l(\omega _l)}{\omega _l\tau _E}}`$ (56) Here $`\mathrm{\Delta }_{n,m}`$ is the conformal dimension of the vertex operator $`V_{n,m}(\sigma )`$. Not surprisingly, the decoherence time of a $`R/R`$ Schrődinger cat diverges when $`\sigma _{12}0`$. This result is obvious since in this limit, the initial state is a pure state. For $`L/R`$ cats, the decoherence time shows a slow variation in term of the differences of positions. ### IV.4 Decoherence time estimations: QED’s case Using asymptotics of spectral densities (see appendix D), one obtains the decoherence time of the $`l`$th modes in the $`R/R`$ case: $$\frac{\tau _L}{\tau _l^{(R/R)}(\sigma _1,\sigma _2)}=4g\mathrm{\Delta }_{n,m}\left(\frac{2\pi v_S}{c}\right)^{2(l1)}.\frac{l^2(l+1)}{(2l+1)!}.\mathrm{sin}^2\left(\frac{\pi l\sigma _{12}}{L}\right)$$ (57) Similarly, for the $`R/L`$ case: $$\frac{\tau _L}{\tau _l^{(R/L)}(\sigma _1,\sigma _2)}=4g\mathrm{\Delta }_{n,m}\left(\frac{2\pi v_S}{c}\right)^{2(l1)}.\frac{l^2(l+1)}{(2l+1)!}.\left(1+\frac{m^24\alpha ^2n^2}{m^2+4\alpha ^2n^2}\mathrm{cos}\left(\frac{2\pi l\sigma _{12}}{L}\right)\right)$$ (58) Since $`v_S/c10^3`$, decoherence times for the $`l`$ and $`l+1`$ modes are related by a typical factor of $`10^6`$. This argument shows that the $`l=1`$ modes dominate the decoherence process. Physically, higher Luttinger modes contribute to higher electric and magnetic multipoles, for which radiative dissipation is known to be weaker. Since dissipation governs decoherence, this is the physical reason for the predominance of the $`l=1`$ Luttinger mode in the decoherence process. The decoherence time of the $`l=1`$ mode is nothing but the electromagnetic relaxation time: $$\tau ^1\frac{16\pi }{3}\frac{\alpha _{\mathrm{QED}}}{\alpha }\left(\frac{v_S}{c}\right)^2.\tau _L10^8\tau _L^1$$ Numerical results for the decoherence times are shown on figure 5 for Luttinger fermions. The temperature dependence can be found easily since $`\mathrm{coth}(\beta \mathrm{}\omega /2)`$ varies slowly around $`\omega _l`$ in a scale $`g\omega _l`$. Therefore: $$\frac{\tau _l^{(R/R)}(\sigma _1,\sigma _2,T)}{\tau _l^{(R/R)}(\sigma _1,\sigma _2,T=0)}=\mathrm{tanh}\left(\frac{\mathrm{}\omega _l}{2k_BT}\right)$$ (59) ### IV.5 Decoherence time estimations: acoustic case Since the sound velocity $`c_S`$ is much smaller than $`v_S`$, the condition $`\omega _l\tau _E<<1`$ does not hold for the coupling to phonons. Explicit computations show that the behavior of spectral densities for the coupling to longitudinal phonons differs from the electronic one. Luttinger modes frequencies fall into a range of frequencies where the acoustic spectral densities are proportional to $`(\omega \tau _S)^{d1}`$. Remember also that the natural cut-off frequency for the phonon bath is given by $`\omega _D=c_S/a`$ where $`a`$ is a typical microscopic length<sup>1</sup><sup>1</sup>1One would naturally think of $`\omega _D`$ as the Debye frequency but a linear dispersion relation for phonons has been assumed, an assumption which certainly fails near the Debye frequency. The $`a`$ length scale should rather be considered as small compared to the mesoscopic size of the system but large compared to the atomic length scale.. The zero temperature acoustic decoherence rate of Luttinger modes is given by formulas 55 where the $`\mathrm{\Gamma }_l^{(d)}=𝒥_l(\omega _l)/\omega _l\tau _S`$ factor is given by: $`\mathrm{\Gamma }_l^{(3)}`$ $`=`$ $`{\displaystyle \frac{g_{ph}(L)l^2}{2}}{\displaystyle \frac{v_S}{c_S}}`$ (60) $`\mathrm{\Gamma }_l^{(2)}`$ $`=`$ $`{\displaystyle \frac{g_{ph}(L)l}{\pi }}`$ (61) In opposition to QED’s case, these damping rates do not decrease with increasing $`l`$. In the QED case, higher $`l`$ modes are bad antennas for the microwave radiation emitted by the system. In the case of phonons the situation goes the other way because of the longitudinal coupling. Although in some cases the coupling constant $`g_{ph}(L)`$ is very small, “decoherence repartition” effects between modes plays a much more important role here than in QED’s case since one has to sum up over many mode contributions to decoherence. For Luttinger modes of index $`l`$, the perturbative expansion is governed by the relative damping rate $`\gamma _l/\omega _l`$, an upper value of which is given by $$\frac{g_{ph}(a)}{8\pi ^2}\frac{\omega _l\tau _S}{(\omega _D\tau _S)^2}\mathrm{for}d=3\mathrm{and}\frac{g_{ph}(a)}{\pi (\omega _D\tau _S)}\mathrm{for}d=2$$ (62) where $`g_{ph}(a)`$ is the rescaled coupling constant for the length $`a`$ (typically of order $`1`$). Assuming that $`\omega _D\tau _S=L/a`$ is much greater than one, we see that all $`l0`$ modes can be considered as weakly damped. The total decoherence exponent in the linear regime is obtained by summing over all the modes up to the Debye frequency. For $`R/R`$ Schrődinger cat states, one finds: $$\tau _L.\gamma ^{(R/R)}(\sigma _1,\sigma _2,T)=8\pi ^2\mathrm{\Delta }_{n,m}\underset{l=1}{\overset{l_{max}}{}}\mathrm{\Gamma }_l^{(d)}\mathrm{coth}\left(\frac{\beta \mathrm{}\omega _l}{2}\right)\mathrm{sin}^2\left(\frac{\pi l\sigma _{12}}{L}\right)$$ (63) Since we sum over a large number of modes, the decoherence time rapidly decreases when $`\sigma _{12}>>av_S/c_S`$, a spectacular effect due to the ratio $`c_S/v_S<<1`$. Roughly speaking, the Luttinger fermion has the time to circle many times around the loop before emitted phonons escape whereas it barely has the time to move in the electromagnetic case. This “averaging effect” explains why the dependence in the initial relative position is much weaker for acoustic than for electromagnetic decoherence. The maximal inverse decoherence rate can be expressed as an integral in the limit $`L>>av_S/c_S`$ ($`k_B\mathrm{\Theta }_D=\mathrm{}\omega _D)`$: $$\tau _L.\gamma ^{(R/R)}(T)=\mathrm{\Delta }_{n,m}\frac{g_{ph}^{(d)}(a)}{4^{d2}\pi }\left(\frac{c_S}{v_S}\right)^2\frac{L}{a}_0^1x^{d1}\mathrm{coth}\left(\frac{x\mathrm{\Theta }_D}{2T}\right)𝑑x$$ (64) The temperature dependence is very weak (remember we are typically working in situations where $`\omega _D\tau _L>10^5`$ and $`k_BT\mathrm{}\omega _L`$). To be precise, it goes like: $$\frac{\gamma ^{(R/R)}(T)\gamma ^{(R/R)}(0)}{\gamma ^{(R/R)}(0)}\left(\frac{T}{\mathrm{\Theta }_D}\right)^d$$ (65) In opposition with the photon bath case, the total acoustic decoherence time scales as $`L^1`$ in units of $`\tau _L`$. ### IV.6 Caldeira-Leggett computations The Caldeira-Leggett model corresponds to the Ohmic spectral density: at low frequencies, $`𝒥(\omega )=\frac{M\gamma \omega }{\pi }`$. In this case, for time scales large compared to the cut-off time, noise and dissipation kernels are local in time. The solution of the model is then much simpler. In the HPZ approach, the equation of motion defining the $`(u_i)_{i=1,2}`$ functions can be solved exactly, taking into account non perturbatively all effects of dissipation. For general spectral densities, solutions of the equation of motions are given in full generality by a Laplace transform of the form: $$\stackrel{~}{u}(p)=\frac{\dot{u}(0)+pu(0)}{1+p^2+2\stackrel{~}{\eta }(p)}\mathrm{and}\stackrel{~}{\eta }(p)=_0^{\mathrm{}}𝑑\omega \frac{\omega 𝒥_l(\omega )}{\omega ^2+p^2}.$$ This expression has clearly a cut along the $`p`$-imaginary axis for $`|p|<\omega _c`$ ($`\omega _c`$ is an UV cutoff such that $`𝒥(\omega )=0`$ for $`\omega \omega _c`$). It has no poles in the physical sheet. The Bromwich contour that encircles the cut is used to find the inverse Laplace transform Alamoudi:1998-1 : $`u(t)`$ $`=`$ $`2{\displaystyle _0^{\omega _c}}S(\omega )\mathrm{sin}(\omega t)𝑑\omega `$ $`S(w)`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Sigma }_I(\omega )}{(\omega ^2\omega _R^2\mathrm{\Sigma }_R(\omega ))^2+\mathrm{\Sigma }_I^2(\omega )}}`$ The self-energy due to the bath is given by: $$\mathrm{\Sigma }_R(\omega )+i\mathrm{\Sigma }_I(\omega )=2\mathrm{P}\mathrm{P}\frac{\omega ^{}𝒥(\omega ^{})}{\omega ^2\omega _{}^{}{}_{}{}^{2}}𝑑\omega ^{}+i\pi \text{sgn}(\omega )𝒥(|\omega |).$$ (66) In the weak coupling regime ($`\gamma \tau _L<<1`$), a Breit-Wigner approximation can be performed since $`\mathrm{\Sigma }_I(\omega _R)<<\omega _R^2`$. Within this approximation, the $`u_i`$ functions can be approximated by standard damped oscillator solutions. In fact, one could then imagine to use an effective Caldeira-Leggett model in order to estimate the decoherence properties of all Luttinger modes. In the electromagnetic case, the decoherence process for the $`l>1`$ modes takes place on a typical time scale of order $`\tau _l\tau _1(c/v_S)^{2l2}`$ which is much longer than for the Ohmic $`l=1`$ modes. In our particular problem, using asymptotics 104, electromagnetic dissipation for the dominant $`l=1`$ modes is given by $`\gamma =g\tau _L^1/3`$. An effective Caldeira-Leggett model for this mode can then be used. In the acoustic case, for a two dimensional phonon bath and within the relevant range of frequencies $`\omega _L<\omega <\omega _D`$, the spectral density $`𝒥_l(\omega )`$ is ohmic for any $`l`$, the Caldeira-Leggett model can be used to describe the long time dependence of decoherence. Although the HPZ method can be used to perform explicit computations, an approximate master equation can often be used to obtain the time evolution of the density matrix of a small subsystem $`𝒮`$ coupled to a “reservoir” $``$. The usual approximation consists first in assuming that the state of the reservoir is unaltered by the coupling to the system, and to forget correlations between $`𝒮`$ at time $`t^{}t`$ and $``$ at time $`t`$. The second approximation usually done consists in neglecting non-Markovian terms which can occur for example in the noise kernel (the dissipation kernel is always local in the CL model). The condition for these two approximations to be valid is Cohen:CDF:1989 : $$\tau _c^2.\mathrm{Tr}(H_𝒮^2\rho _𝒮(0))<<\mathrm{}^2$$ (67) where $`H_𝒮`$ denotes the coupling between the system and the reservoir, $`\rho _𝒮(0)`$ the initial total density operator and $`\tau _c`$ the correlation time of the reservoir. In the present context, this condition can be expressed as: $$\xi _l_0^+\mathrm{}𝒥_l(x/\tau _E)\mathrm{coth}\left(\frac{l(\beta )}{2L}x\right)𝑑x<<\left(\frac{\tau _E}{\tau _c}\right)^2$$ (68) where $`\xi _l`$ is a dimensionless coefficient that characterizes the spreading of the $`l`$-th mode initial state : $`\xi _l=1`$ for any coherent state and $`\xi _l=\mathrm{coth}(\pi \mathrm{}lv_S\beta /L)`$ for thermal equilibrium. Within the temperature range used here $`T/T_L010`$ ($`T_L=2\pi \mathrm{}v_S/k_BL`$), we have $`l(\beta )>>L`$ for the electromagnetic bath. Therefore, an estimate of the correlation time for the QEM field is given by the thermal time. High frequency asymptotics 106 show that temperature dependence has negligible effect on the l.h.s. of 68 (the effect could be important only for the $`l=1`$ mode). Then, within our temperature range, the l.h.s is approximately bounded by $`gL/a`$. The r.h.s is bounded by $`L/a`$ at high temperatures and $`(v_S/c)(T/T_L)`$ at lower temperatures. Therefore, since $`g10^8`$, the validity condition is valid for the photon bath at temperatures above $`g(L/a)(c/v_S)\times T_L`$. In the acoustic case, the l.h.s. temperature dependance is only relevant for values of $`x`$ below $`(c_S/v_S)\times (T/T_L)`$. But the main contribution to the l.h.s comes from higher frequencies is of order $`g_{ph}(L)(L/a)^dg_{ph}(a)(L/a)`$. At low temperatures, the r.h.s can be bounded by $`(c_S/v_S)(T/T_L)`$ using $`\mathrm{}\beta c_S`$ as an upper value for $`\tau _c`$. Condition 68 can therefore be rewritten as $`(Lv_S/c_Sa)g_{ph}(a)<<T/T_L`$. Subsequent computations will assume that the acoustic UV cutoff is much higher than the typical Luttinger liquid frequency $`2\pi v_S/L`$, i.e that $`Lv_S/c_Sa>>1`$. The master equation can therefore only be used for quite high temperatures. In the case of an underdamped oscillator, as pointed out by C. Cohen-Tannoudji Cohen:CDF:1989 , the approximate Markovian description can be used provided the noise kernel is local compared to the observation time scale. In the present case, this means $`\gamma <<\omega _0`$ and $`k_BT>>\mathrm{}\gamma `$. In the limit of very low temperatures $`k_BT<<\mathrm{}\gamma `$, this description is still valid but one could expect very long time algebraic tails which precisely take into account non Markovian effects induced by the divergence of the bath’s correlation time $`\mathrm{}\beta `$. This possibility is discussed in appendix E where these corrections are computed in a weakly damped regime and shown to be extremely weak. ### IV.7 Long time decoherence (Markovian master equation treatment) Within the secular approximation, the master equation can be translated in the following evolution equation for the characteristic function. In the case of a single oscillator of renormalized eigenfrequency $`\mathrm{\Omega }`$ coupled to an Ohmic bath with dissipation rate $`\gamma `$: $`Z_t(\lambda ,\overline{\lambda })=\mathrm{Tr}(\rho _𝒮(t)e^{\lambda a^{}}e^{\overline{\lambda }a})`$ is given by: $$\left[\frac{}{_t}\left(i\mathrm{\Omega }\frac{\gamma }{2}\right)\lambda \frac{}{\lambda }+\left(i\mathrm{\Omega }+\frac{\gamma }{2}\right)\overline{\lambda }\frac{}{\overline{\lambda }}\right]Z_t(\lambda ,\overline{\lambda })=\frac{\gamma }{e^{\beta \mathrm{}\mathrm{\Omega }}1}Z_t(\lambda ,\overline{\lambda })$$ (69) The general solution of this equation is given by: $$Z_t(\lambda ,\overline{\lambda })=Z_0(\lambda e^{\frac{\gamma t}{2}+i\mathrm{\Omega }t},\overline{\lambda }e^{\frac{\gamma t}{2}i\mathrm{\Omega }t}).\mathrm{exp}\left(\frac{1e^{\gamma t}}{e^{\beta \mathrm{}\mathrm{\Omega }}1}\lambda \overline{\lambda }\right)$$ (70) Such a formula immediately shows that, at zero temperature, coherent states remain coherent but their parameters evolve according to $`\alpha (t)=\alpha (0)e^{\frac{\gamma t}{2}i\mathrm{\Omega }t}`$. A coherent superposition of two coherent states decohere as follows: $`|\alpha \beta |`$ $``$ $`|\alpha (t)\beta (t)|\times e^{i\theta (t)}e^{d(t)}`$ (71) $`\theta (t)`$ $`=`$ $`(1e^{\gamma t})\mathrm{}(\alpha \overline{\beta })`$ (72) $`d(t)`$ $`=`$ $`{\displaystyle \frac{|\alpha \beta |^2}{2}}(1e^{\gamma t})`$ (73) The decoherence time is therefore given by a very simple formula: $$\tau _{\mathrm{Dec}}^1=\gamma .\frac{|\alpha \beta |^2}{2}$$ (74) Using formula 35 in the Caldeira-Leggett case provides the same result and the temperature dependence is given, in the limit $`\gamma /2\mathrm{\Omega }_10`$ ($`\mathrm{\Omega }_1=\sqrt{\mathrm{\Omega }^2\gamma ^2/4}`$) by: $$D_T(t)=\frac{\mathrm{\Theta }(T)(1e^{\gamma t})}{1+(\mathrm{\Theta }(T)1).(1e^{\gamma t})}\times \mathrm{𝟏}\mathrm{with}\mathrm{\Theta }(T)=\mathrm{coth}\left(\frac{\beta \mathrm{}\mathrm{\Omega }_1}{2}\right)$$ (75) Therefore, the $`t+\mathrm{}`$ limit of decoherence is independent of temperature but the decoherence time will scale with temperature according to the $`\mathrm{\Theta }(T)`$ factor: $$\frac{\tau _{Dec}(T)}{\tau _{Dec}(0)}\mathrm{tanh}\left(\frac{\mathrm{}\mathrm{\Omega }_1}{2k_BT}\right)$$ (76) ### IV.8 Decoherence at long time Applying previous results to the electromagnetic decoherence ($`l=1`$ modes), we obtain: $`d_{\mathrm{}}^{(R/R)}(\sigma _1,\sigma _2)`$ $`=`$ $`4\mathrm{\Delta }_{n,m}.\mathrm{sin}^2\left({\displaystyle \frac{\pi \sigma _{12}}{L}}\right)`$ (77) $`d_{\mathrm{}}^{(L/R)}(\sigma _1,\sigma _2)`$ $`=`$ $`2\mathrm{\Delta }_{n,m}.\left(1+{\displaystyle \frac{m^24n^2\alpha ^2}{m^2+4n^2\alpha ^2}}\mathrm{cos}\left({\displaystyle \frac{2\pi \sigma _{12}}{L}}\right)\right)`$ (78) As expected, the $`R/R`$ limiting decoherence vanishes for $`\sigma _1=\sigma _2`$ whereas the $`R/L`$ one does not. The typical asymptotic value is proportional to $`\mathrm{\Delta }_{n,m}`$. In fact, this number can be viewed as measuring the “distance” between the two quantum states which built our Schrődinger cat. Using vertex operator with small values of$`m`$ and $`n`$ in Schrődinger cats 44 and 46 produces mesoscopically separated coherent states in each mode. Figure 6 summarizes the electromagnetic decoherence exponent as a function of time for $`\gamma t05`$ and for various temperatures. In the acoustic case (two dimensional phonon bath), contributions of all relevant modes should be summed. As before, the $`\sigma _{12}`$ dependence disappears as soon as $`\sigma _{12}>>av_S/c_S`$. Introducing $`T=2\pi ^2\tau _Lv_S/(c_Sg_{ac}(a))`$, the sum over all Luttinger modes up to the cut-off frequency can be evaluated: $$\frac{d^{(R/R)}(t)}{d^{(R/R)}(\mathrm{})}=\frac{d^{(R/L)}(t)}{d^{(R/L)}(\mathrm{})}=1+\frac{e^{t/T}1}{(t/T)}$$ (79) where $`d^{(R/R)}(\mathrm{})=d^{(R/L)}(\mathrm{})=\mathrm{\Delta }_{n,m}c_SL/(v_Sa)`$. For the continuum approximation to be valid, we have assumed that $`L/a>>v_S/c_S`$ and therefore $`d^{(R/R)}(\mathrm{})>>1`$. This also implies that most of the decoherence process is accomplished within the previously computed acoustic decoherence time $`2T/d^{(R/R)}(\mathrm{})<<T`$. ## V Conclusion and discussion Within the bosonization framework, we have shown how the coupling to an external quantum electromagnetic field or to a two or three dimensional bath of longitudinal phonons can lead to decoherence of a Schrődinger cat state formed of localized elementary excitations in a Luttinger liquid. Three different phases build the decoherence scenario for Schrődinger cat states of the Luttinger liquid (see 7 for the electromagnetic case): * At a very short time, because of the time/energy uncertainty relation, energy exchanges between the environment and the Luttinger system are not conservative. In this regime, high frequencies take part in the decoherence process and non-Markovian effects are important. The precise time evolution of the system is strongly cut-off dependent. * After a transitory regime, the decoherence exponent becomes linear in time. This is the “Golden Rule” regime: energy conservation between the environment field and the Luttinger system is satisfied with a spectral width going down to its natural value (spontaneous photon or phonon emission). Only frequencies in resonance with the Luttinger eigenfrequencies contribute to dissipation. The structure of spectral densities for the environmental modes leads to a hierarchy of decoherence times corresponding to the multipolar expansion of the “radiation” emitted by the Luttinger system. In the acoustic case, the various decoherence times do not increase as much with increasing $`l`$. In the electromagnetic case, they increase with $`l`$. The total decoherence time is therefore much smaller in the acoustic than in the electromagnetic case. * At longer times, decoherence tends to saturation. In this regime, an effective Caldeira-Leggett model can be used to describe the dominating decoherence processes: in the electromagnetic case, one can keep only $`l=1`$ modes, corresponding to dipolar electromagnetic radiation. In the 2D acoustic case, one can use an effective Caldeira-Leggett model for all modes. Caldeira-Leggett computations are non-perturbative since they take into account all orders of the coupling between the quantum environment and the oscillator. The infinite time decoherence depends on two factors. The first one is, as expected, the distance between the two quantum states which depends on the dimension of the operators used in these states. The second one reflects the relative weight of each hydrodynamic mode in the decoherence process. In the electromagnetic case, since $`l=1`$ modes dominate, it gives a geometrical factor depending on the relative position $`\sigma _{12}`$. In the acoustic case, spatial dependence is almost always lost and we are left with an important mode number factor. That’s why, although $`\psi _R^{}(\sigma _1)|0`$ and $`\psi _R^{}(\sigma _1)|0`$ can be considered as “mesoscopically close” with respect to their “distance”, decoherence is much faster than dissipation in the acoustic case. This is a major difference with single mode decoherence studies CL:1985-1 where the ratio between decoherence and dissipation times is only due the distance between states entering the Schrődinger cat. Non-linearities in the spectrum of low energy excitations may also contribute to decoherence. As showed by Haldane Haldane:1981-1 , non-linearities in the spectrum couple the bosonic modes of the theory, turning the simple free model used in bosonization into an interacting theory. Coupling between modes should also play an important role in the decoherence properties of Schrődinger cat states. Indeed, the model presented here provides an upper limit for decoherence times. Non-linearities should be also taken into account when investigating the resistive behavior of small metallic loops induced by inelastic collisions Landauer:1985-1 . In the present work, photons or longitudinal phonons initially at equilibrium were used as the thermal bath for the system, but of course, one could imagine various extensions. One could use another description for quantum fluctuations of the environment. For example, one may think about changing the state of the environment, taking for example into account an external microwave radiation. Increasing the incoming radiation power within the range of resonant frequencies should increase decoherence of Schrődinger cat states (enhancement of dissipation by stimulated emission of radiation). With such environmental states, one expects to meet also the problem of “decoherence repartition” between all the modes of the Luttinger system (even in the electromagnetic case). Although this makes computations much harder to control, it may lead to more interesting behaviors. Finally, we are also investigating the extension of these ideas to two or three dimensional systems. ###### Acknowledgements. We acknowledge many useful discussions and encouragements with Ch. Chaubet, L. Cugliandolo, B. Douçot, J. Kurchan and L. Saminadayar. A. De Martino, R. Mélin and P. Pujol are warmly thanked for their careful reading of the manuscript and their numerous remarks and questions. D. Mouhanna has kindly sent us a copy of C. Cohen-Tannoudji lectures on decoherence at Collège de France. Part of this work was performed during the Spring workshop on Strongly correlated fermions at the Institut Henri Poincaré (Apr. - July 1999). The organizers of the XXIVth Rencontres de Moriond on Quantum Physics at mesoscopic scale are thanked for giving us the opportunity to present this work. We also thank D. Loss and Th. Martin for stimulating conversations during this meeting and communication of many useful references. ## Appendix A Keldysh’s Green’s functions Let us recall the well known expressions for a single harmonic mode of frequency $`\omega `$ at temperature $`\beta `$ (unit mass) Mahan : $`G_>(t,s)`$ $`=`$ $`{\displaystyle \frac{i}{2\omega }}\left(\mathrm{coth}({\displaystyle \frac{\beta \mathrm{}\omega }{2}})\mathrm{cos}(\omega (ts))i\mathrm{sin}(\omega (ts))\right)`$ (80) $`G_<(t,s)`$ $`=`$ $`{\displaystyle \frac{i}{2\omega }}\left(\mathrm{coth}({\displaystyle \frac{\beta \mathrm{}\omega }{2}})\mathrm{cos}(\omega (ts))+i\mathrm{sin}(\omega (ts))\right)`$ (81) $`G_T(t,s)`$ $`=`$ $`{\displaystyle \frac{i}{2\omega }}\left(\mathrm{coth}({\displaystyle \frac{\beta \mathrm{}\omega }{2}})\mathrm{cos}(\omega (ts))i\mathrm{sin}(\omega |ts|)\right)`$ (82) $`G_{\stackrel{~}{T}}(t,s)`$ $`=`$ $`{\displaystyle \frac{i}{2\omega }}\left(\mathrm{coth}({\displaystyle \frac{\beta \mathrm{}\omega }{2}})\mathrm{cos}(\omega (ts))+i\mathrm{sin}(\omega |ts|)\right)`$ (83) The retarded and advanced Green functions can be related to these expressions by $$G_R=G_TG_<\mathrm{and}G_A=G_TG_>$$ (85) For the electromagnetic field, a mode decomposition can be used. Using results of appendix B, we obtain: $$𝑫^{\alpha \beta }((𝒙,t);(𝒚,s))=\underset{I}{}𝑨_I^\alpha (𝒙)𝑨_I^\beta (𝒚)𝑮_{\omega _I}(t,s)$$ (86) The noise and dissipation kernels 4 and 5 can be expressed as: $`\nu ^{\alpha \beta }(x,y)`$ $`=`$ $`{\displaystyle \underset{I}{}}𝑨_I^\alpha (𝒙)𝑨_I^\beta (𝒚)\mathrm{coth}\left({\displaystyle \frac{\beta \mathrm{}\omega }{2}}\right){\displaystyle \frac{\mathrm{cos}(\omega _I(ts))}{2\omega _I}}`$ (87) $`\eta ^{\alpha \beta }(x,y)`$ $`=`$ $`{\displaystyle \underset{I}{}}𝑨_I^\alpha (𝒙)𝑨_I^\beta (𝒚){\displaystyle \frac{\mathrm{sin}(\omega _I(ts))}{2\omega _I}}`$ (88) ## Appendix B Electromagnetic modes in a cylindrical cavity We shall decompose the transverse vector potential in orthonormal modes: $$𝑨(𝒓,t)=\underset{I}{}\psi _I(t)𝑨_I(𝒓)$$ (89) These modes can be real, or complex. In the latter case, we assume the existence of an involution $`I\widehat{I}`$ over the set of indices implementing complex conjugation: $`(𝑨_I(𝒓))^{}=𝑨_{\widehat{I}}(𝒓)`$. Modes are normalized with respect of the volume $`𝒱`$ of the cavity by imposing an orthogonality condition: $$d^3𝒓𝑨_I(𝒓).𝑨_J(𝒓)=\delta _{I,\widehat{J}}.$$ The $`\psi _I(t)`$ are coordinates for a set of independent harmonic oscillators of frequency $`\omega _I`$. Expressions are given in cylindrical coordinates $`(r,\vartheta ,z)`$. $`R`$ denotes the radius of the cavity. It is useful to introduce: $`𝑨(𝒓)=𝓐(𝒓)/\sqrt{𝒱}`$. Let $`l`$ be the angular momentum of the mode around the $`Oz`$ axis and $`k_k`$ its momentum along this axis. Let $`u_{l,n}^{}`$ denote the $`n`$-th zero of the derivative of the $`l`$-th order Bessel function $`J_l`$. Then, the orthoradial component of the TE $`(l,k_z,n)`$ complex mode is given by: $$𝓐_\vartheta =\frac{u_{l,n}^{}}{\sqrt{(u_{l,n}^{})^2l^2}}\frac{J_l^{}(\frac{r}{R}u_{l,n}^{})}{J_l(u_{l,n}^{})}e^{i(l\vartheta k_zz)}$$ (90) We have $`\omega =c\sqrt{k_z^2+(u_{l,n}^{}/R)^2}`$. Here $`u_{l,n}`$ denotes the $`n`$-th zero of the $`l`$-th order Bessel function $`J_l`$. Then, the orthoradial component is given by: $$𝓐_\vartheta =\frac{k_zR}{\sqrt{u_{l,n}^2+(k_zR)^2}}\frac{l}{u_{l,n}}\frac{J_l(\frac{r}{R}u_{l,n})}{J_l^{}(u_{l,n})}e^{i(l\vartheta k_zz)}$$ (91) We have $`\omega =c\sqrt{k_z^2+(u_{l,n}/R)^2}`$. Normalized real modes are inferred from these complex modes by: $$𝓐_I^{(+)}=\frac{1}{\sqrt{2}}\left(𝓐_I+𝓐_{\widehat{I}}\right)\mathrm{and}𝓐_I^{()}=\frac{i}{\sqrt{2}}\left(𝓐_I𝓐_{\widehat{I}}\right)$$ (92) Expressions 90 and 91 show that: $$\{\begin{array}{cc}𝓐_{I,\vartheta }^{(+)}=𝒟_I\mathrm{cos}(l\vartheta k_zz)\hfill & \\ 𝓐_{I,\vartheta }^{()}=𝒟_I\mathrm{sin}(l\vartheta k_zz)\hfill & \end{array}$$ (93) where $`mathcalD_I`$ contains the $`r`$ dependence and depends on the mode characteristics $`TE/TM`$ and $`(l,k_z,n)`$. More precisely we have $`TE\mathrm{modes}:𝒟_I`$ $`=`$ $`{\displaystyle \frac{u_{l,n}^{}\sqrt{2}}{\sqrt{(u_{l,n}^{})^2l^2}}}{\displaystyle \frac{J_l^{}(\frac{r}{R}u_{l,n}^{})}{J_l(u_{l,n}^{})}}`$ (94) $`TM\mathrm{modes}:𝒟_I`$ $`=`$ $`{\displaystyle \frac{Rk_z\sqrt{2}}{\sqrt{u_{l,n}^2+(k_zR)^2}}}{\displaystyle \frac{l}{u_{l,n}}}{\displaystyle \frac{J_l(\frac{r}{R}u_{l,n})}{J_l^{}(u_{l,n})}}`$ (95) ## Appendix C Longitudinal acoustic modes in a cylindrical cavity The displacement field $`𝒖(𝒙,t)`$ contains a gradient part: $$𝒖(𝒙,t)=\underset{I}{}\frac{Q_I(t)}{\sqrt{𝒱}}(\mathbf{}\phi _I)(𝒙)$$ where the $`\phi _I`$ functions are eigenvalues of the Laplacian with von Neuman’s boundary conditions. The modes $`𝒖_I=\mathbf{}\phi _I`$ are normalized so that $$d^3𝒓𝒖_I(𝒓).𝒖_J(𝒓)=\delta _{I,\widehat{J}}𝒱.$$ The $`Q_I(t)`$ are coordinates for a set of independent harmonic oscillators with given frequency $`\omega _I`$. Longitudinal acoustic modes are indexed by $`I=(l,k_z,n)`$ for $`d=3`$ and $`I=(l,n)`$ for $`d=2`$ and are typically of the form $`e^{i(l\theta k_zz)}𝒩_I`$ where $`𝒩_I`$ contains the $`r`$ dependence ($`\omega _I=c_S\sqrt{k_z^2+(u_{l,n}^{}/R)^2}`$ for $`d=3`$ and $`\omega _I=c_Su_{l,n}^{}/R`$ for $`d=2`$): $`d=2:𝒩_I`$ $`=`$ $`{\displaystyle \frac{2u_{l,n}^{}}{(\omega _I\tau _S)\sqrt{(u_{l,n}^{})^2l^2}}}{\displaystyle \frac{J_l(\frac{r}{R}u_{l,n}^{})}{J_l(u_{l,n}^{})}}`$ (96) $`d=3:𝒩_I`$ $`=`$ $`{\displaystyle \frac{u_{l,n}^{}\sqrt{2}}{(\omega _I\tau _S)\sqrt{(u_{l,n}^{})^2l^2}}}{\displaystyle \frac{J_l(\frac{r}{R}u_{l,n}^{})}{J_l(u_{l,n}^{})}}`$ (97) ## Appendix D Asymptotic behavior of spectral densities Asymptotic expressions of the spectral densities corresponding to the infinite-cavity limit are useful to study the long time behavior of the decoherence and dissipation processes. The radial quantization is of $`\mathrm{\Delta }\omega _{}=\mathrm{\Delta }u_{l.n}^{(^{})}c/R\pi c/R`$ and the longitudinal one $`\mathrm{\Delta }\omega _{}=2\pi c/h`$. We shall therefore use the following expressions for the extremas of the Bessel functions: $$J_l(u_{l,n}^{})\sqrt{\frac{2}{\pi u_{l,n}^{}}}\text{and}J_l^{}(u_{l,n})\sqrt{\frac{2}{\pi u_{l,n}}}$$ (98) and the following low- and high- frequency expansions $$J_l(z)\frac{z^l}{2^ll!}\mathrm{and}J_l(z)\sqrt{\frac{2}{\pi z}}\mathrm{cos}(zl\pi /2\pi /4)$$ (99) A straightforward algebra gives us: 1. In the low-frequency regime ($`\omega \tau _{EM}<<1`$) $`𝒥_{l0}^{TE}(\omega )`$ $`{\displaystyle \frac{g}{\pi }}{\displaystyle \frac{l}{(2l1)!}}\left({\displaystyle \frac{\omega \tau _{EM}}{2\pi }}\right)^{2l1}`$ (100) $`𝒥_0^{TE}(\omega )`$ $`{\displaystyle \frac{1}{3\pi }}{\displaystyle \frac{v_S}{c}}{\displaystyle \frac{\alpha _{QED}}{\alpha }}\left({\displaystyle \frac{\omega \tau _{EM}}{2\pi }}\right)^3`$ (101) $`𝒥_{l0}^{TM}(\omega )`$ $`{\displaystyle \frac{g}{\pi }}{\displaystyle \frac{2l^2}{(2l+1)!}}\left({\displaystyle \frac{\omega \tau _{EM}}{2\pi }}\right)^{2l1}`$ (102) $`𝒥_0^{TM}(\omega )`$ $`=0`$ (103) Henceforth, the total spectral densities behave as: $`𝒥_{l0}(\omega )`$ $`{\displaystyle \frac{g}{\pi }}{\displaystyle \frac{l^2(l+1)}{(2l+1)!}}\left({\displaystyle \frac{\omega \tau _{EM}}{2\pi }}\right)^{2l1}`$ (104) $`𝒥_0(\omega )`$ $`{\displaystyle \frac{1}{3\pi }}{\displaystyle \frac{v_S}{c}}{\displaystyle \frac{\alpha _{QED}}{\alpha }}\left({\displaystyle \frac{\omega \tau _{EM}}{2\pi }}\right)^3`$ (105) 2. In the high-frequency regime ($`\omega \tau _{EM}>>1`$) $`𝒥_l^{TE}(\omega )`$ $`{\displaystyle \frac{lg}{2\pi }}`$ (106) $`𝒥_l^{TM}(\omega )`$ $`\omega ^1`$ (107) These analytic results agree with the numerics, which are represented in figure 3. The main result is that all the modes are supraohmic at low frequency, expect the mode $`l=1`$ which shows ohmic behavior. The same kind of expansion can be performed for the acoustic spectral densities. In this case, one obtains ($`d\{2,3\}`$): * In the low frequency regime: $$𝒥_l(\omega )g_{Ph}(L)(\omega \tau _S)^{2l+d}$$ (108) * In the high frequency regime: $$𝒥_l(\omega )g_{Ph}(L)\frac{l}{2^{2(d2)}\pi }(\omega \tau _S)^{d1}$$ (109) Only the latter will be used to compute decoherence properties since $`\omega _l`$ falls into the high frequency regime since $`v_S>>c_S`$. ## Appendix E Decoherence matrix computations for a Caldeira-Leggett solution The appendix presents details of the computation of the decoherence matrix $`D`$ (35) in the case of a damped oscillator solution. This has direct relevance for the Caldeira-Leggett model but also within the framework of the Breit-Wigner approximation in more general environments. These results have been discussed in CL:1985-1 ; Hu:1995-1 but their derivation is recalled here in a simpler way. Strictly speaking the approximate master equation approach described in section IV.6 is not valid for temperatures below the $`\mathrm{}\gamma /k_B`$. Therefore, this appendix aims at finding ultra-low temperature corrections to decoherence arising from the non-Markovian effects arising from the $`T0`$ behavior of the reservoir symmetric two point correlation function. Let $`\mathrm{\Omega }_R`$ and $`\gamma `$ denote the renormalized frequency of the oscillator and $`\gamma `$ the damping coefficient. We shall work in the weakly damped case, defined by $`\mathrm{\Omega }_1^2=\mathrm{\Omega }_R^2\gamma ^2/40`$, and measure the strength of dissipation by $`\varphi `$ such that $`\mathrm{tan}(\varphi )=\gamma /2\mathrm{\Omega }_R`$. The decoherence matrix can then be expressed in terms of the following auxiliary functions: $`Z_\pm (t,\omega )`$ $`=`$ $`{\displaystyle _0^t}e^{\gamma s/2+i(\omega \pm \mathrm{\Omega }_1)s}𝑑s,`$ (110) $`S(t)`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}(|Z_+|^2+|Z_{}|^2)(t,\omega )𝒥(\omega )\mathrm{coth}\left({\displaystyle \frac{\beta \mathrm{}\omega }{2}}\right)𝑑\omega ,`$ (111) $`P(t)`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}(Z_+\overline{Z_{}})(t,\omega )𝒥(\omega )\mathrm{coth}\left({\displaystyle \frac{\beta \mathrm{}\omega }{2}}\right)𝑑\omega ,`$ (112) We have: $`D`$ $`=`$ $`{\displaystyle \frac{1}{d}}\left(\begin{array}{cc}D_{11}& D_{12}\\ D_{12}& D_{22}\end{array}\right),`$ (115) $`d`$ $`=`$ $`4(1+{\displaystyle \frac{2e^{\gamma t}}{\mathrm{\Omega }_R(S^24P\overline{P})}}(S({\displaystyle \frac{\mathrm{\Omega }_R}{\mathrm{\Omega }}}+{\displaystyle \frac{\mathrm{\Omega }}{\mathrm{\Omega }_R}})+2\mathrm{}\left(Pe^{2i\mathrm{\Omega }_1t}({\displaystyle \frac{\mathrm{\Omega }_R}{\mathrm{\Omega }}}e^{2i\varphi }{\displaystyle \frac{\mathrm{\Omega }}{\mathrm{\Omega }_R}})\right))+`$ (116) $`+`$ $`{\displaystyle \frac{16\mathrm{cos}^2(\varphi )e^{2\gamma t}}{\mathrm{\Omega }_R^2(S^24P\overline{P})}}),`$ (117) and the coefficients are given by: $`D_{11}`$ $`=`$ $`\mathrm{\Omega }+{\displaystyle \frac{2e^{\gamma t}}{S^24P\overline{P}}}(S+2\mathrm{}(Pe^{2i(\mathrm{\Omega }_1t+\varphi )})),`$ (118) $`D_{12}=D_{21}`$ $`=`$ $`{\displaystyle \frac{2e^{\gamma t}}{\mathrm{\Omega }_R(S^24P\overline{P})}}(S\mathrm{sin}(\varphi )+2\mathrm{}(Pe^{i(2\mathrm{\Omega }_1t+\varphi )})),`$ (119) $`D_{22}`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{\Omega }}}+{\displaystyle \frac{2e^{\gamma t}}{\mathrm{\Omega }_R(S^24P\overline{P})}}(S+2\mathrm{}(Pe^{2i\mathrm{\Omega }_1t})).`$ (120) The $`S`$ and $`P`$ functions can be computed by the residue theorem: $`S`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}\left({\displaystyle \frac{1+e^{\gamma t}2e^{\gamma t/2}e^{i(\omega +\mathrm{\Omega }_1)t}}{(\omega +\mathrm{\Omega }_1)^2+\gamma ^2/4}}+{\displaystyle \frac{1+e^{\gamma t}2e^{\gamma t/2}e^{i(\omega \mathrm{\Omega }_1)t}}{(\omega \mathrm{\Omega }_1)^2+\gamma ^2/4}}\right)𝒥(\omega )\mathrm{coth}\left({\displaystyle \frac{\beta \mathrm{}\omega }{2}}\right)𝑑\omega ,`$ (121) $`P`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{1+e^{\gamma t}e^{2i\mathrm{\Omega }_1t}2e^{\gamma t/2}e^{i\mathrm{\Omega }_1t}e^{i\omega t}}{(\omega +\mathrm{\Omega }_1+i\gamma /2)(\omega \mathrm{\Omega }_1i\gamma /2)}}𝒥(\omega )\mathrm{coth}\left({\displaystyle \frac{\beta \mathrm{}\omega }{2}}\right)𝑑\omega .`$ (122) The main contribution to $`S`$ and $`P`$ is due to the poles $`\pm \mathrm{\Omega }_1+i\gamma /2`$: $`S^{(Main)}`$ $`=`$ $`{\displaystyle \frac{2\pi }{\gamma }}(1e^{\gamma t})\mathrm{}\left(𝒥(\mathrm{\Omega }_1+i\gamma /2)\mathrm{coth}\left({\displaystyle \frac{\beta \mathrm{}}{2}}(\mathrm{\Omega }_1+i\gamma /2)\right)\right),`$ (123) $`P^{(Main)}`$ $`=`$ $`{\displaystyle \frac{i\pi }{2}}(1e^{\gamma t}e^{2i\mathrm{\Omega }_1t}){\displaystyle \frac{𝒥(\mathrm{\Omega }_1+i\gamma /2)}{\mathrm{\Omega }_1+i\gamma /2}}\mathrm{coth}\left({\displaystyle \frac{\beta \mathrm{}}{2}}(\mathrm{\Omega }_1+i{\displaystyle \frac{\gamma }{2}})\right),`$ (124) In this expression cut-off dependent quantities have been discarded since they can be shown to be of order $`(\gamma /2\mathrm{\Omega }_1)\mathrm{log}(\mathrm{\Lambda }/\mathrm{\Omega }_R)`$, i.e. much smaller than the oscillator eigenfrequency’s renormalization. The temperature dependent part also contain poles which give a series of exponentially decreasing terms of the form $`\mathrm{exp}(2\pi nt/\tau _{Th})`$ ($`n1`$). At vanishing temperature, this series can be resummed into an algebraically decreasing correction as follows (here $`𝒥(\omega )=\gamma \omega /\pi \mathrm{\Omega }_1^2`$): $`S^{(Corr.)}`$ $`=`$ $`{\displaystyle \frac{2e^{\gamma t/2}}{\pi \mathrm{\Omega }_1^2t}}\mathrm{}\left(e^{i\mathrm{\Omega }_1t}(z_t𝒮(z_t)+\overline{z_t}𝒮(\overline{z_t}))+e^{i\mathrm{\Omega }_1t}(z𝒮(z_t)+\overline{z_t}𝒮(\overline{z_t}))\right),`$ (125) $`P^{(Corr.)}`$ $`=`$ $`{\displaystyle \frac{\varphi e^{\gamma t/2}e^{i\mathrm{\Omega }_1t}}{\pi \mathrm{\Omega }_1}}\left(𝒮(z_t)+𝒮(z_t)\overline{𝒮(\overline{z_t})}\overline{𝒮(\overline{z_t})}\right)`$ (126) where $`𝒮(z)=e^z\mathrm{Ei}(1,z)`$ and $`z=(\gamma /2i\mathrm{\Omega }_1)t`$. $`𝒮(z_t)`$ has an asymptotic expansion in $`1/t`$ for $`t+\mathrm{}`$. At zero temperature, these corrective terms 125 and 126 dominate 123 and 124 for $`\gamma t>>\mathrm{log}(2\mathrm{\Omega }_1/\gamma )`$. The very long time asymptotic of $`S`$ is therefore given by: $`S{\displaystyle \frac{2}{\mathrm{\Omega }_1}}\left(1+{\displaystyle \frac{4\varphi e^{\gamma t/2}\mathrm{cos}(\mathrm{\Omega }_1t)}{\pi (\mathrm{\Omega }_1t)^2}}\right)`$ (127) In case of weak damping $`\gamma <<\mathrm{\Omega }_R`$, the following approximations can be made: $`\varphi <<1`$, $`\mathrm{\Omega }\mathrm{\Omega }_R\mathrm{\Omega }_1`$ and $`f(\mathrm{\Omega }_1+i\gamma /2)f(\mathrm{\Omega }_1)`$. We will neglect $`P`$ since, in full generality, $`P\varphi S`$. Plugging everything in $`D`$’s expression, one finally ends up with a very long time asymptotics: $$D(T=0)\left(1e^{\gamma t}+\frac{4\varphi }{\pi }e^{3\gamma t/2}\frac{\mathrm{cos}(\mathrm{\Omega }_1t)}{(\mathrm{\Omega }_1t)^2}\right).\left(\begin{array}{cc}\mathrm{\Omega }_1/4& 0\\ 0& 1/4\mathrm{\Omega }_1\end{array}\right).$$ (128) Henceforth, in the very weak damping situation, master equation results for the decoherence matrix can safely be extrapolated down to $`T=0`$, even if strictly speaking non-Markovian effects should be taken into account.
warning/0001/hep-th0001036.html
ar5iv
text
# References 1. An interesting phenomenon has been observed in $`N=2`$ supersymmetric $`SU(2)`$ gauge theories with various flavors and with adjoint mass perturbation : confinement is caused by condensation of magnetic monopoles carrying nontrivial flavor quantum numbers (see also for further details): spontaneous flavor symmetry breaking is caused by the same dyamical mechanism responsible for confinement in these models. We wish to know what happens in more general classes of models, and through a systematic analysis, to gain a more microscopic understanding of these phenomena and related ones in Quantum Chromodynamics. As we see below, the generalization from $`SU(2)`$ to higher-rank gauge groups turns out to be quite subtle. We discuss here models constructed from exactly solvable $`N=2`$ $`SU(n_c)`$ and $`USp(2n_c)`$ gauge theories with all possible numbers of flavor compatible with asymptotic freedom, by perturbing them with a small adjoint mass (reducing supersymmetry to $`N=1`$) and keeping small, generic bare hypermultiplet (quark) masses. The advantage of doing so is that the only vacua retained are those in which the gauge coupling constant grows in the infrared. Another advantage is that in this way all flat directions are eliminated and one is left with a finite number of isolated vacua; keeping track of this number allows us to perform highly nontrivial checks of our analyses at various steps. Our analysis heavily relies on the breakthrough works by Seiberg and Witten , and those which followed them . Also crucial will be Seiberg’s $`N=1`$ electromagnetic duality , and newly discovered universal classes of (super) conformally invariant theories -. The special cases of $`SU(2)=USp(2)`$ theories with $`n_f=1,2,3,4`$ were studied in . For $`n_f=1,4`$, there is no dynamical flavor symmetry breaking. For $`n_f=2`$, monopoles in the $`(\underset{¯}{2},1)+(1,\underset{¯}{2})`$ (spinor) representation of the flavor $`[SU(2)\times SU(2)]/Z_2=SO(4)`$ group is found to condense after $`N=1`$ perturbation $`\mu \text{Tr}\mathrm{\Phi }^2`$: the flavor symmetry is necessarily broken to $`U(2)`$. For $`n_f=3`$, monopoles in the $`\underset{¯}{4}`$ (spinor) representation of the flavor $`SO(6)`$ group condense with $`\mu 0`$ and the flavor symmetry is broken to $`U(3)`$ while there is another vacuum where a flavor-singlet dyon condenses and the flavor symmetry is unbroken. This result naturally leads to a conjecture that the condensation of monoples with non-trivial flavor transformation property explains the confinement à la ‘t Hooft and the flavor symmetry breaking simultaneously. However, a simple thought reveals a problem with this picture. As we will see later, the monopoles in $`USp(2n_c)`$ theories transform under the spinor representation of $`SO(2n_f)`$ flavor symmetry, and their effective low-energy Lagrangian coupled to the magnetic $`U(1)`$ gauge group would have an accidental $`SU(2^{n_f1})`$ flavor symmetry, and their condensation would lead to far too many Nambu–Goldstone multiplets. The case of $`SU(2)`$ gauge theories was special because the flavor symmetries of the monopole action precisely coincide with the symmetry of the microscopic theories due to the small number of flavors. This argument suggests that the phenomenon of flavor symmetry breaking is richer in higher rank theories. Argyres, Plesser and Seiberg studied higher-rank $`SU(n_c)`$ theories with $`n_f2n_c1`$ (asymptotically free) in detail. They showed how the non-renormalization theorem of the hyperKähler metric on the Higgs branch could be used to show the persistence of unbroken non-abelian gauge group at the “roots” of the Higgs branches (non-baryonic and baryonic branches) where they intersect the Coulomb branch. Some isolated points on the non-baryonic roots with $`SU(r)`$ ($`r[n_f/2]`$) gauge group as well as the baryonic root (single point) with $`SU(\stackrel{~}{n}_c)=SU(n_fn_c)`$ gauge group were found to survive the $`\mu 0`$ perturbation. Their main focus, however, was the attempt to “derive” Seiberg’s duality between $`SU(n_c)`$ and $`SU(\stackrel{~}{n}_c)`$ gauge theories relying on the baryonic root,<sup>1</sup><sup>1</sup>1This “derivation,” however, was incomplete as it did not produce all components of the “meson” superfield. Moreover, the effective low-energy theory was perturbed by a relevant operator (the mass term for the mesons) and did not flow to the Seiberg’s magnetic theory correctly. We thank P. Argyres for discussions on this point. and the issue of flavor symmetry breaking was not studied at any depth. The analysis also left a puzzle why there were “extra” theories at the non-baryonic roots which seemingly had nothing to do with Seiberg’s dual theories. Another paper by Argyres, Plesser and Shapere addressed similar questions in $`SO(n_c)`$ and $`USp(2n_c)`$ theories . In the present paper, we find that the flavor $`U(n_f)`$ symmetry in $`SU(n_c)`$ theories can be dynamically broken to various $`U(r)\times U(n_fr)`$ groups. We find that in the $`r=1`$ vacua the dynamical symmetry breaking is indeed caused by the condensation of monopoles in the $`\underset{¯}{n_f}`$ representation. For general $`r`$, however, the monopoles in the $`\underset{¯}{{}_{n_f}{}^{}C_{r}^{}}`$ representation, whose condensation could have explained the flavor symmetry breaking but would have produced too-many Nambu–Goldstone multiplets, actually “break up” into “magnetic quarks” whose baryonic composites under the unbroken $`SU(r)`$ gauge group match the monopoles. The baryonic roots are shown always to coincide with the non-baryonic roots with $`r=\stackrel{~}{n}_c`$. The non-baryonic roots are shown to be necessary ingredients of the Seiberg’s dual theories rather than being “extra.” The vacua with unbroken flavor symmetries are associated with the baryonic roots. The situation with $`USp(2n_c)`$ theories is even less trivial. The low-energy theories are non-trivial superconformal theories with no description in terms of a weakly coupled local field theory. In obtaining these results, counting of the number of vacua proved to be an extremely useful tool. The counting was done in the semi-classical limit, large $`\mu `$ limit, using the curve, as well as using low-energy effective Lagrangians and they all agree with each other. 2. First we perform a preparatory analysis, by minimizing the scalar potential following from the Lagrangian valid in the semi-classical regime (when both $`\mu `$ and $`m`$ are large). $`N=1`$ supersymmetry and holomorphy guarantee the absence of phase transitions between large $`\mu `$, $`m`$ to small $`\mu `$, $`m`$. Therefore these vacua are related to quantum vacua in other regimes one by one. The Lagrangian of the models has the structure $$=\frac{1}{8\pi }\text{Im}\tau _{cl}\left[d^4\theta \mathrm{\Phi }^{}e^V\mathrm{\Phi }+d^2\theta \frac{1}{2}WW\right]+^{(quarks)}+\mathrm{\Delta },$$ (1) where $$\mathrm{\Delta }=d^2\theta \mu \text{Tr}\mathrm{\Phi }^2$$ (2) is the adjoint mass breaking the supersymmetry to $`N=1`$ and $$^{(quarks)}=\underset{i}{}\left[d^4\theta \{Q_i^{}e^VQ_i+\stackrel{~}{Q}_ie^V\stackrel{~}{Q}_i^{}\}+d^2\theta \{\sqrt{2}\stackrel{~}{Q}_i\mathrm{\Phi }Q^i+m_i\stackrel{~}{Q}_i\mathrm{\Phi }Q^i\}\right]$$ (3) describes the $`n_f`$ flavors of hypermultiplets (“quarks”), and $`\tau _{cl}\theta _0/\pi +8\pi i/g_0^2`$ is the bare $`\theta `$ parameter and coupling constant. The $`N=1`$ chiral and gauge superfields $`\mathrm{\Phi }=\varphi +\sqrt{2}\theta \psi +\mathrm{}`$, and $`W_\alpha =i\lambda +\frac{i}{2}(\sigma ^\mu \overline{\sigma }^\nu )_\alpha ^\beta F_{\mu \nu }\theta _\beta +\mathrm{}`$ are both in the adjoint representation of the gauge group, while the quarks are taken in the fundamental representation. In the limit $`m_i0,`$ and $`\mu 0`$, these models possess an exact flavor symmetry, $`U(n_f)\times Z_{2n_cn_f}`$ or $`SO(2n_f)\times Z_{2n_c+2n_f}`$, for $`SU(n_c)`$ or $`USp(2n_c)`$ gauge groups, respectively. In the equal quark mass limit, the symmetry of the symplectic gauge theory is reduced to $`U(n_f).`$ The models are asymptotically free as long as $`n_f<2n_c`$ (for $`SU(n_c)`$ gauge theory) or $`n_f<2n_c+2`$ (for $`USp(2n_c)`$). We find $$𝒩=\underset{r=0}{\overset{\mathrm{min}\{n_f,n_c1\}}{}}(n_cr){}_{n_f}{}^{}C_{r}^{}$$ (4) semi-classical solutions for $`SU(n_c)`$ gauge theory with $`n_f`$ flavors, while for $`Usp(2n_c)`$ theory with $`n_f`$ flavors, the number of $`N=1`$ vacua is $$𝒩=\underset{r=0}{\overset{\mathrm{min}\{n_c,n_f\}}{}}(n_cr+1){}_{n_f}{}^{}C_{r}^{}.$$ (5) The factor $`n_cr`$ or $`n_cr+1`$ appearing in the sum originates from Witten’s index for unbroken gauge group. For small number of flavors, these expressions simplify somewhat: $$𝒩_1=(2n_cn_f)\mathrm{\hspace{0.17em}2}^{n_f1},(SU(n_c)\text{with}n_fn_c);$$ (6) $$𝒩_1=(2n_c+2n_f)\mathrm{\hspace{0.17em}2}^{n_f1},(USp(2n_c)\text{with}n_fn_c+1).$$ (7) It is amusing that these different expressions all reproduce correctly the number of $`N=1`$ vacua in the case of $`SU(2)`$ theory (which is a special case, both of $`SU(n_c)`$ and of $`USp(2n_c)`$) with $`n_f=04,`$ $$𝒩=n_f+2.$$ (8) 3. We next determine the possible patterns of dynamical flavor symmetry breaking in these theories. This is done most easily by studying these theories at large fixed $`\mu \mathrm{\Lambda }`$, $`m_i0`$.<sup>2</sup><sup>2</sup>2We also investigated the limit $`\mu \mathrm{}`$ while $`m_i\mathrm{\Lambda }`$ fixed, which is suited for studying the decoupling of the adjoint fields. We checked this way the consistency with the known results about $`N=1`$ theories. Such an analysis is possible since at large adjoint mass the low-energy effective superpotential can be read off from the bare Lagrangian by integrating out the heavy adjoint field and by adding to it the known exact instanton–induced superpotentials of the corresponding $`N=1`$ theories. By minimizing the superpotential, we found in all cases the correct number of vacua Eqs.(4)-(7). $`N=1`$ supersymmetry kept intact throughout guarantees that there are no phase transitions as $`\mu `$ is varied; we can thus determine the symmetry breaking pattern in each $`N=1`$ vacuum from the first principles. The analysis is straightforward, but is not entirely trivial for large $`n_f`$ because the non-perturbative effects among the low-energy degrees of freedom (dual quarks and mesons) have to be correctly taken into account despite the fact that they are in a “free magnetic phase”. For instance, for $`SU(n_c)`$ theory with $`n_f<n_c`$ the effective superpotential reads $$W=\frac{1}{2\mu }\left[\text{Tr}M^2\frac{1}{n_c}(\text{Tr}M)^2\right]+\text{Tr}(mM)+\frac{\mathrm{\Lambda }_1^{(3n_cn_f)/(n_cn_f)}}{(detM)^{1/(n_cn_f)}},$$ (9) where $`M_i^j\stackrel{~}{Q}_i^aQ_a^j`$, and $`\mathrm{\Lambda }_1=(\mu ^{n_c}\mathrm{\Lambda }^{2n_cn_f})^{\frac{1}{3n_cn_f}}`$ is the scale of the $`N=1`$ theory. The minima of the potential are characterized by the set of vacuum expectation values (in the $`m_i0`$ limit), $$M=\text{diag}(\lambda _1,\lambda _2,\mathrm{},\lambda _{n_f}),$$ (10) $$\lambda _1=\mathrm{}=\lambda _r=(n_c+rn_f)Z,\lambda _{r+1}=\mathrm{}=\lambda _{n_f}=(n_cr)Z,$$ (11) where $$Z=C\left(\mu ^{n_cn_f}\mathrm{\Lambda }_1^{3n_cn_f}\right)^{1/(2n_cn_f)}\omega ^k,(k=1,2,\mathrm{}2n_cn_f;\omega =e^{2\pi i/(2n_cn_f)}),$$ (12) with $`CO(1)`$ a constant that depends on $`n_f`$, $`n_c`$ and $`r`$. In a vacuum characterized by $`r`$, the flavor symmetry of the model is broken spontaneously as $$U(n_f)U(r)\times U(n_fr).$$ (13) To avoid double counting, we can restrict $`r[n_f/2]`$ with all $`k`$, and for the special case of $`r=n_f/2`$ (possible only when $`n_f`$ is even), $`k=1,\mathrm{},n_cn_f/2`$, for each choice of $`r`$ flavors out of $`n_f`$. We find a total $$𝒩_1=(2n_cn_f)2^{n_f1}$$ (14) of such vacua, after summation over $`r`$. The number for $`n_f=n_c`$ is given by the same formula using the “quantum modified constraint” among the mesons and baryons following Seiberg. For $`n_fn_c`$ the above exhausts the number of the vacua. In the case $`n_f=n_c+1`$ we used the appropriate effective Lagrangian involving mesons and baryons to find that there are $`𝒩_1`$ vacua with various symmetry breaking (13) plus one vacuum with no flavor symmetry breaking. The total number $`𝒩_1+1`$ reproduces (4) correctly. The situation for larger numbers of flavor ($`n_f>n_c+1`$) is more subtle. The effective low-energy action in these cases has the form (we set the “matching scale” to unity to simplify expressions) $$W=\stackrel{~}{q}Mq+\text{Tr}(mM)\frac{1}{2\mu }\left[\text{Tr}M^2\frac{1}{n_c}(\text{Tr}M)^2\right],$$ (15) where $`q`$’s are $`n_f`$ flavors of dual quarks in the fundamental representation of the dual gauge group $`SU(\stackrel{~}{n}_c)`$, where $`\stackrel{~}{n}_c=n_fn_c`$. The minima of the potential following from Eq.(15) can be found straightforwardly, and gives $$𝒩_2=\underset{r=0}{\overset{\stackrel{~}{n}_c1}{}}{}_{n_f}{}^{}C_{r}^{}(\stackrel{~}{n}_cr)$$ (16) vacua. The solutions have a color-flavor diagonal form for $`q`$’s and $`\stackrel{~}{q}`$’s, with $`r`$ nonzero elements, $`d_i,\stackrel{~}{d}_i`$, where $$d_i=\stackrel{~}{d}_i=\left[m_i\frac{1}{n_c+rn_f}\underset{j=r+1}{\overset{n_f}{}}m_j\right]^{1/2},i=1,2,\mathrm{},r.$$ (17) The meson vacuum expectation value (VEV) is orthogonal to the squark VEVS, $$M=\text{diag}(0,0,\mathrm{},0,\lambda _{r+1},\lambda _{r+2},\mathrm{},\lambda _{n_f}),\lambda _i=\mu \left[m_i+\frac{1}{n_c+rn_f}\underset{j=r+1}{\overset{n_f}{}}m_j\right].$$ (18) All VEVS of fields carrying flavor quantum numbers thus vanish in the limit $`m_i0`$, showing that the flavor symmetry remains unbroken in this class of vacua. The problem is that the number of vacua found this way is too small, since we know that the exact number of vacua is $`𝒩`$ (Eq.(4)), $`𝒩>𝒩_2.`$ Where are other vacua? This apparent puzzle can be solved once the nontrivial $`SU(\stackrel{~}{n}_c)`$ instanton effects are taken into account properly.<sup>3</sup><sup>3</sup>3In fact, a related puzzle is how Seiberg’s dual Lagrangian \- the first two terms of Eq. (15) - can give rise to the right number of vacua for the massive $`N=1`$ SQCD with $`n_f>n_c+1`$. By following the same method as below but with $`\mu =\mathrm{}`$, we do find the correct number ($`n_c`$) of vacua. If the meson vacum expectation values have rank $`n_f`$, the dual quarks can be integrated out, leaving the effective superpotential, $$W_{\mathrm{𝑒𝑓𝑓}}=\frac{1}{2\mu }\left[\text{Tr}M^2\frac{1}{n_c}(\text{Tr}M)^2\right]+\text{Tr}(Mm)+\mathrm{\Lambda }_1^{(3n_cn_f)/(n_cn_f)}(detM)^{1/(n_fn_c)}.$$ (19) Minimization of this effective action gives $`𝒩_1=(2n_cn_f)2^{n_f1}`$ solutions, having the same forms as Eq.(10)-Eq.(12). At this point, one can make a highly nontrivial consistency check: by changing $`rn_fr`$ and rearranging terms, one shows that the total number of quantum vacua is equal to $$𝒩_1+𝒩_2=\underset{r=0}{\overset{n_c1}{}}(n_cr){}_{n_f}{}^{}C_{r}^{}=𝒩,$$ (20) i.e., equal to the total number of semi-classical vacua. We find therefore that there are two types of vacua: the first of them, with finite VEVS of mesons (in $`m_i0`$ limit), are present for all values of flavors. They are classified by an integer $`r[n_f/2]`$, and the flavor symmetry is spontaneously broken as $`U(n_f)U(r)\times U(n_fr)`$. In the second type of vacua, present only for large flavors ($`n_fn_f+1`$), the flavor symmetry remains unbroken. The second type of vacua are closely related to the emergence of the dual gauge group of Seiberg. The analysis in the case of $`USp(2n_c)`$ models is similar, but the result is qualitatively different. We find again two types of $`N=1`$ vacua. The first type of vacua has finite meson vacuum expectation values $`M^{ij}J^{ij}`$ (symplectic matrix) with the flavor $`SO(2n_f)`$ symmetry broken as $$SO(2n_f)U(n_f),$$ (21) in all vacua of this class. This phenomenon is quite reminiscent of what is believed to occur in the standard QCD. The number of this type of vacua is given by $$𝒩_1=(2n_c+2n_f)\mathrm{\hspace{0.17em}2}^{n_f1},$$ (22) which is the number of vacua for ($`n_f<n_c+2`$). As in the $`SU(n_c)`$ case, when the number of the flavor is sufficiently large ($`n_fn_c+2`$) we find also another class of vacua in which the flavor $`SO(2n_f)`$ symmetry is unbroken. There are $$𝒩_2=\underset{r=0}{\overset{n_fn_c2}{}}(n_fn_c1r){}_{n_f}{}^{}C_{r}^{}$$ (23) of them, and together with those of the first group, they make up the total number $$𝒩=𝒩_1+𝒩_2=\underset{r=0}{\overset{n_c}{}}(n_c+1r){}_{n_f}{}^{}C_{r}^{}$$ (24) which is the correct number of $`N=1`$ vacua for $`USp(2n_c)`$ (see Eq.(5)). 4. We now seek for a microscopic understanding of the mechanism of dynamical flavor symmetry breaking. We do so by studying the $`N=2`$ vacua on the Coulomb branch which survive $`\mu 0`$ perturbation. We start from the auxiliary genus $`n_c1`$ ($`n_c`$) curves for $`SU(n_c)`$ ($`USp(2n_c)`$) theories $$y^2=\underset{k=1}{\overset{n_c}{}}(x\varphi _k)^2+4\mathrm{\Lambda }^{2n_cn_f}\underset{j=1}{\overset{n_f}{}}(x+m_j),SU(n_c),n_f2n_c2,$$ (25) with $`\varphi _k`$ subject to the constraint $`_{k=1}^{n_c}\varphi _k=0`$, and $$xy^2=\left[x\underset{a=1}{\overset{n_c}{}}(x\varphi _a^2)^2+2\mathrm{\Lambda }^{2n_c+2n_f}m_1\mathrm{}m_{n_f}\right]^24\mathrm{\Lambda }^{2(2n_c+2n_f)}\underset{i=1}{\overset{n_f}{}}(x+m_i^2),USp(2n_c).$$ (26) The VEVS of $`a_{Di},a_i`$ are constructed as integrals over the non-trivial cycles of the meromorphic differentials on the curves. We require that the curve is maximally singular, i.e. $`n_c1`$ (or $`n_c`$ for $`USp(2n_c)`$) pairs of branch points to coincide: this determines the possible values of $`\{\varphi _a\}`$’s. These correspond to the $`N=1`$ vacua, with the particular $`N=1`$ perturbation, Eq.(2). Note that as we work with generic and nonvanishing quark masses, this is an unambiguous procedure to identify all the $`N=1`$ vacua of our interest. <sup>4</sup><sup>4</sup>4There are other kinds of singularities of $`N=2`$ QMS at which, for instance, three of the branch points meet. These correspond to $`N=1`$ vacua, selected out by different types of perturbations such as $`\text{Tr}\mathrm{\Phi }^3`$, which are not considered here. We find in this way precisely the same number ($`𝒩`$) of $`N=1`$ vacua, where $`𝒩`$ was determined earlier by the semi-classical and large $`\mu `$ analyses. In each vacuum, there are $`n_c1`$ (or $`n_c`$) different kinds of massless magnetic monopoles, corresponding to maximal Abelian subgroup of $`SU(n_c)`$ or of $`USp(2n_c)`$. At small generic quark masses, we observe that these singularities group into approximate multiplets of vacua, with multiplicities $`{}_{n_f}{}^{}C_{r}^{}`$, $`r=0,1,2,\mathrm{},[n_f/2]`$, in the case of $`SU(n_c)`$, while they appear in $`2^{n_f1}`$-plets plus certain number of other vacua, in the case of $`USp(2n_c)`$ theories. Their positions are compatible with the approximate discrete symmetries, $`Z_{2n_cn_f}`$ or $`Z_{2n_c+2n_f}`$. We have made an extensive numerical study in the case of rank two gauge groups with all possible numbers of flavor, as well as general analytical study of these phenomena for higher-rank groups. As $`m_i0`$ (or equal mass limit in the case of $`SU(n_c)`$) each multiplet of vacua collapse into one multiple vacuum. This behavior might suggest a more or less straightforward generalization of what occurs in $`SU(2)`$ gauge theories, mentioned at the beginning. Indeed, monopoles can acquire nontrivial flavor quantum numbers as shown by Jackiw and Rebbi through the fermion zero modes. In $`SU(n_c)`$ theories, by acting fermion zero mode operators $`d_i`$, $`d_j^{}`$ on the monopole state $`|\mathrm{\Omega }`$, such as $$d_i^{}|\mathrm{\Omega },d_{i_1}^{}d_{i_2}^{}|\mathrm{\Omega },\mathrm{},d_{i_1}^{}\mathrm{}d_{i_{n_f}}^{}|\mathrm{\Omega },$$ (27) we find semi-classical monopoles belonging to anti-symmetric tensor representations of $`U(n_f)`$. It might appear then the dynamical flavor symmetry breaking (13) is caused by the condensation of such monopoles. As mentioned earlier, however, this picture would lead to far too many Nambu–Goldstone multiplets (except for $`r=0,1`$). The same analysis for $`USp(2n_c)`$ case shows that the semi-classical monopoles are in the spinor representation of $`SO(2n_f)`$, and their condensation would give the symmetry breaking (21) and the number of vacua $`2^{n_f1}`$. We would again run into a paradox of having a too-large $`SU(2^{n_f1})`$ symmetry. Actually the theories avoid falling into this kind of paradox, but do so in a subtle way. Let us discuss below physics of $`SU(n_c)`$ and $`USp(2n_c)`$ gauge theories separately. 5. In the $`SU(n_c)`$ case, the $`N=1`$ vacua can all be generated from the various classes of superconformal theories with $`m_i=\mu =0`$, by perturbation by masses $`m_i`$. The first type of vacua (with multiplicity $`𝒩_1`$) correspond to the curves $$y^2x^{2r}(x\alpha _1)^2\mathrm{}(x\alpha _{n_cr1})^2(x\beta )(x\gamma ),r=0,1,2,\mathrm{},[n_f/2],$$ (28) that is $$\text{diag}\varphi =(\underset{r}{\underset{}{0,0,\mathrm{},0}},\varphi _1,\mathrm{}\varphi _{n_cr}),\underset{a=1}{\overset{n_cr}{}}\varphi _a=0,$$ (29) with $`\varphi _a`$’s chosen such that the nonzero $`2(n_cr1)`$ branch points are paired. These correspond to the so-called class 1 ($`r<n_f/2`$) and 3 ($`r=n_f/2`$, with $`n_cn_f/2`$ odd) superconformal theories , while the case, $`r=n_f/2`$, $`n_cn_f/2`$ even, may be interpreted as belonging to class 4. Since these adjoint VEVS break the discrete symmetry spontaneously, they appear in $`2n_cn_f`$ copies.<sup>5</sup><sup>5</sup>5There is an exception to this. In the case of $`r=n_f/2`$ with $`n_f`$ even, the explicit configuration of $`\varphi _a`$’s can be found by using the Chebyshev polynomials. This vacuum respects $`Z_2`$ subgroup of the $`Z_{2n_cn_f}`$ symmetry, showing that it appears in $`n_cn_f/2`$ copies rather than $`2n_cn_f.`$ This fact is crucial in the vacuum counting below Eq.(38). When (generic) quark masses are turned on, these vacua split into $`{}_{n_f}{}^{}C_{r}^{}`$-plet of single vacua. The second class of vacua stem from the (trivial) superconformal theory $$y^2x^{2\stackrel{~}{n}_c}(x^{n_c\stackrel{~}{n}_c}\mathrm{\Lambda }^2)^2,\stackrel{~}{n}_c=n_fn_c,$$ (30) corresponding to the singularity $$\text{diag}\varphi =(\underset{\stackrel{~}{n}_c}{\underset{}{0,0,\mathrm{},0}},\mathrm{\Lambda }\omega ,\mathrm{},\mathrm{\Lambda }\omega ^{n_c\stackrel{~}{n}_c})$$ (31) with $`\omega =e^{2\pi i/(n_c\stackrel{~}{n}_c)}`$. Actually there is no vacuum of the first type with $`r=\stackrel{~}{n}_c`$. The most detailed description of these $`N=1`$ vacua comes from the considerations based on nonrenormalization theorem of the Higgs branch metric . The first class of vacua with given $`r`$ is an $`SU(r)\times U(1)^{n_cr}`$ gauge theory with $`n_f`$ “quarks” and $`n_cr`$ singlet monopoles $`e_k`$’s, with an effective Lagrangian, $$W_{\mathrm{𝑛𝑜𝑛𝑏𝑎𝑟}}=\sqrt{2}\text{Tr}(q\varphi \stackrel{~}{q})+\sqrt{2}\psi _0\text{Tr}(q\stackrel{~}{q})+\sqrt{2}\underset{k=1}{\overset{n_cr1}{}}\psi _ke_k\stackrel{~}{e}_k+\mu \left(\mathrm{\Lambda }\underset{i=0}{\overset{n_cr1}{}}x_i\psi _i+\frac{1}{2}\text{Tr}\varphi ^2\right),$$ (32) where $`\varphi `$ and $`\psi _k`$’s are part of the $`SU(r)\times U(1)^{n_cr}`$ $`N=2`$ vector multiplets and $`x_iO(1)`$ constants. These are at the roots of the so-called “non-baryonic” branches , where they meet the Coulomb branch. They describe an infrared-free (i.e., non asymptotic free) effective theory for $`r<n_f/2`$. We now add the mass terms $$\text{Tr}(mq\stackrel{~}{q})\underset{k,i}{}S_k^im_ie_k\stackrel{~}{e}_k$$ (33) and minimize the potential. We find $`{}_{n_f}{}^{}C_{r}^{}`$ solutions characterized by the vacuum expectation values ($`q`$ and $`\stackrel{~}{q}`$ having color-flavor diagonal form, with nonvanishing elements, $`d_i`$ and $`\stackrel{~}{d}_i`$)<sup>6</sup><sup>6</sup>6Actually, Eq.(32) and Eq.(33) allow for a number of other solutions in which the vev of $`\psi _0`$ is of $`O(\mathrm{\Lambda })`$; these are the first group of $`N=1`$ vacua found in . Such solutions, involving fluctuations much larger than both $`m_i`$ and $`\mu `$, however, lie beyond the validity of the low-energy effective Lagrangian. They should therfore be regarded as an artefact of the approximation and must be discarded. $$\psi _0=\frac{1}{\sqrt{2}r}\underset{i=1}{\overset{r}{}}m_i,\psi _k=O(m_i),$$ (34) $$d_i\stackrel{~}{d}_i=\mu \left(m_i\frac{1}{r}\underset{j=1}{\overset{r}{}}m_j\right)\frac{1}{\sqrt{2}r}\mu \mathrm{\Lambda }x_0;e_k\stackrel{~}{e}_k\mu \mathrm{\Lambda }.$$ (35) The multiplicity $`{}_{n_f}{}^{}C_{r}^{}`$ arises from the choice of $`r`$ (out of $`n_f`$) quark masses used to construct the solution. In the massless limit we find $$d_i=\stackrel{~}{d}_i\sqrt{\mu \mathrm{\Lambda }},i=1,2,\mathrm{}r:$$ (36) this leads to the correct symmetry breaking pattern, $`U(n_f)U(r)\times U(n_fr)`$. For $`r=n_f/2`$, the theory at the singularity becomes a non-trivial superconformal theory. There is no description of this singularity in terms of weakly coupled local field theory. The monodromy around the singularity shows that the theory is indeed superconformal (we checked this explicitly for $`n_c=3`$ and $`n_f=4`$). Careful perturbation of the curve by the quark masses shows that there are $`(n_cn_f/2){}_{n_f}{}^{}C_{n_f/2}^{}`$ vacua.<sup>7</sup><sup>7</sup>7Due to some reason, however, the naive application of the effective Lagrangian Eq. (32,33) gives the correct number. The total number of the vacua of this type is ($`n_fn_c`$): $$(2n_cn_f)\underset{r=0}{\overset{(n_f1)/2}{}}{}_{n_f}{}^{}C_{r}^{}=(2n_cn_f)\mathrm{\hspace{0.17em}2}^{n_f1},(n_f=\text{odd})$$ (37) $$(2n_cn_f)\underset{r=0}{\overset{n_f/21}{}}{}_{n_f}{}^{}C_{r}^{}+\frac{2n_cn_f}{2}{}_{n_f}{}^{}C_{n_f/2}^{}=(2n_cn_f)\mathrm{\hspace{0.17em}2}^{n_f1},(n_f=\text{even}),$$ (38) which exhausts $`𝒩`$, Eq.(6). In Eq.(38) we have taken into account the fact that for even $`n_f`$, the vacua with $`r=n_f/2`$ do not transform under $`Z_{2n_cn_f}`$ but only under $`Z_{n_cn_f/2}.`$ When $`n_f>n_c`$, we need to exclude the term $`r=\stackrel{~}{n}_c=n_fn_c`$ from the sum because it gives the second type of vacua. We obtain therefore $`𝒩_1(2n_cn_f){}_{n_f}{}^{}C_{\stackrel{~}{n}_c}^{}`$ vacua. As for the second group of vacua, Eq.(30), Eq.(31), they are an $`SU(\stackrel{~}{n}_c)\times U(1)^{n_c\stackrel{~}{n}_c}`$ gauge theory with $`n_f`$ “quarks” and $`n_c\stackrel{~}{n}_c`$ singlet monopoles $`e_k`$’s . The effective low-energy Lagrangian for this theory is given by $$W_{\mathrm{𝑏𝑎𝑟}}=\sqrt{2}\text{Tr}(q\varphi \stackrel{~}{q})+\frac{\sqrt{2}}{\stackrel{~}{n}_c}\text{Tr}(q\stackrel{~}{q})\left(\underset{k=1}{\overset{n_c\stackrel{~}{n}_c}{}}\psi _k\right)\sqrt{2}\underset{k=1}{\overset{n_c\stackrel{~}{n}_c}{}}\psi _ke_k\stackrel{~}{e}_k+\mu \left(\mathrm{\Lambda }\underset{i=1}{\overset{n_c\stackrel{~}{n}_c}{}}x_i\psi _i+\frac{1}{2}\text{Tr}\varphi ^2\right).$$ (39) where $`\varphi `$ and $`\psi _k`$’s are now in $`SU(\stackrel{~}{n}_c)\times U(1)^{n_c\stackrel{~}{n}_c}`$ $`N=2`$ vector multiplets. We add the mass terms $$\text{Tr}(mq)\underset{k,i}{}S_k^im_ie_k\stackrel{~}{e}_k.$$ (40) We find two types of vacua of the effective low-energy Lagrangian. The first type has $`e_k=\stackrel{~}{e}_k=(\mu \mathrm{\Lambda }x_k/\sqrt{2})^{1/2}`$ for all $`k=1,\mathrm{}n_c\stackrel{~}{n}_c`$. Minimizing the potential in this case, we find $$𝒩_2=\underset{r=0}{\overset{\stackrel{~}{n}_c1}{}}(\stackrel{~}{n}_cr){}_{n_f}{}^{}C_{r}^{}$$ (41) $`N=1`$ vacua, characterized by the vevs $$\varphi =\frac{1}{\sqrt{2}}\text{diag}(m_1,\mathrm{},m_r,c,\mathrm{},c);c=\frac{1}{\stackrel{~}{n}_cr}\underset{k=1}{\overset{r}{}}m_k$$ (42) $$d_i,\stackrel{~}{d}_i\sqrt{\mu m}\stackrel{m_i0}{}0,e_k,\stackrel{~}{e}_k\sqrt{\mu \mathrm{\Lambda }}.$$ (43) The unbroken $`SU(\stackrel{~}{n}_cr)`$ gauge group gives $`\stackrel{~}{n}_cr`$ vacua each. These vacua describe the vacua with unbroken $`SU(n_f)`$ symmetry, which are known to exist from the large $`\mu `$ analysis. The second type of vacua in Eqs. (39,40) has one of the $`e_k=\stackrel{~}{e}_k=0`$ (hence $`n_c\stackrel{~}{n}_c=2n_cn_f`$ choices) while $`W/\psi _k=0`$ requires quarks to condense with $`q=\stackrel{~}{q}\sqrt{\mu \mathrm{\Lambda }}`$. Dropping $`e_k=\stackrel{~}{e}_k=0`$ from the Lagrangian, it becomes the same as that of the non-baryonic root Eqs. (32,33) and gives $`(2n_cn_f){}_{n_f}{}^{}C_{\stackrel{~}{n}_c}^{}`$ vacua. This precisely compensates the exclusion of $`r=\stackrel{~}{n}_c`$ in the sum for the non-baryonic roots and the correct total number of vacua $`𝒩_1+𝒩_2`$ is obtained. We thus find that both the number and the symmetry properties of the $`N=1`$ theories at small adjoint mass $`\mu `$ exactly match those found at large $`\mu ,`$ without encountering any paradoxical situation. We postpone a discussion of physical aspects of $`SU(n_c)`$ theories to the end (point 7 below). 6. In the $`USp(2n_c)`$ gauge theories, the first type of vacua can be identified more easily by first considering the equal but novanishing quark masses. The adjoint vevs in the curve Eq.(26) can be chosen so as to factor out the behavior $$y^2=(x+m^2)^{2r}[\mathrm{}],r=1,2,\mathrm{}$$ (44) which describes an $`SU(r)\times U(1)`$ gauge theory with $`n_f`$ quarks. These (trivial) superconformal theories belong in fact to the same universality classes as in the $`SU(n_c)`$ gauge theory as pointed out by . They are therefore described by exactly the same Lagrangian Eq.(32). At each vacuum with $`r`$, the symmetry (of equal mass theory, $`U(n_f)`$) is broken spontaneously as $$U(n_f)U(r)\times U(n_fr):$$ (45) as in Eq.(13). When a small mass splitting is added among $`m_i`$’s, each of the $`r`$ vacuum split into $`{}_{n_f}{}^{}C_{r}^{}`$ vacua, leading to the total of $$(2n_c+2n_f)\underset{r=0}{\overset{(n_f1)/2}{}}{}_{n_f}{}^{}C_{r}^{}=(2n_c+2n_f)\mathrm{\hspace{0.17em}2}^{n_f1},(n_f=\text{odd})$$ (46) $$(2n_c+2n_f)\underset{r=0}{\overset{n_f/21}{}}{}_{n_f}{}^{}C_{r}^{}+\frac{2n_c+2n_f}{2}{}_{n_f}{}^{}C_{n_f/2}^{}=(2n_c+2n_f)\mathrm{\hspace{0.17em}2}^{n_f1},(n_f=\text{even}),$$ (47) vacua of this type, consistently with Eq. (7).<sup>8</sup><sup>8</sup>8These $`N=1`$ vacua seem to have been overlooked in . In the massless limit the underlying theories possess a larger, flavor $`SO(2n_f)`$ symmetry. We know also from the large $`\mu `$ analysis that in the first group of vacua (with finite vevs), this symmetry is broken spontaneously to $`U(n_f)`$ symmetry always. How can such a result be consistent with Eq.(45) of equal (but nonvanishing) mass theory? What happens is that in the massless limit various $`N=1`$ vacua with different symmetry properties Eq.(45) (plus eventually other singularities) coalesce. The location of this singularity can be obtained exactly in terms of Chebyshev polynomials. At the singularity there are mutually non-local dyons and hence the theory is at a non-trivial infrared fixed point (in the example of $`USp(4)`$ theory with $`n_f=4`$, we have explicitly verified this by determining the singularities and branch points at finite equal mass $`m`$ and by studying the limit $`m0`$.) There is no description in terms of a weakly coupled local field theory, just as in the case $`r=n_f/2`$ for $`SU(n_c)`$ theories. Since the global flavor symmetry is $`SO(2n_f)`$ these superconformal theories belong to different universality classes as compared to those at finite mass. We find this behavior resonable because the semi-classical monopoles are in the spinor representation of the $`SO(2n_f)`$ flavor group and, in contrast to the situation in $`SU(n_c)`$ theories, cannot “break up” into quarks in the vector representation. They are therefore likely to persist at the singularity and makes the theory superconformal. Once the quark masses are turned on, however, the flavor group reduces to (at least) $`U(n_f)`$ and it becomes possible for monopoles to break up into quarks; this explains the behavior in the equal mass case. As for the second group of vacua, the situation is more analogous to the case of $`SU(n_c)`$ theories. The superpotential reads in this case (by adding mass terms to Eq.(5.10) of ): $`W`$ $`=`$ $`\mu \left(\text{Tr}\varphi ^2+\mathrm{\Lambda }{\displaystyle \underset{a=1}{\overset{2n_c+2n_f}{}}}x_a\psi _a\right)+{\displaystyle \frac{1}{\sqrt{2}}}q_a^i\varphi _b^aq_c^iJ^{bc}+{\displaystyle \frac{m_{ij}}{2}}q_a^iq_b^jJ^{ab}`$ (48) $`+`$ $`{\displaystyle \underset{a=1}{\overset{2n_c+2n_f}{}}}\left(\psi _ae_a\stackrel{~}{e}_a+S_a^im_ie_a\stackrel{~}{e}_a\right),`$ where $`J=i\sigma _2\mathrm{𝟏}_{n_c}`$ and $$m=i\sigma _2\text{diag}(m_1,m_2,\mathrm{},m_{n_f}).$$ (49) By minimizing the potential, we find $$𝒩_2=\underset{r=0}{\overset{\stackrel{~}{n}_c}{}}(\stackrel{~}{n}_cr+1){}_{n_f}{}^{}C_{r}^{}$$ (50) vacua, which precisely matches the number of the vacua of the second group, with squark vevvs behaving as $$q_i,\stackrel{~}{q}_i\sqrt{\mu m_i}\stackrel{m_i0}{}0.$$ (51) These are the desired $`SO(2n_f)`$ symmetric vacua. 7. To summarize, we have studied the dynamics of $`N=1`$ $`SU(n_c)`$ and $`USp(2n_c)`$ gauge theories obtained by perturbing $`N=2`$ theories with $`n_f`$ hypermultiplets in the fundamental representation with a finite adjoint mass $`\mu \mathrm{Tr}\mathrm{\Phi }^2`$, determining the possible flavor symmetry breaking patterns. There are vacua in confinement phase with symmetry breaking $`U(n_f)U(r)\times U(n_fr)`$ ($`r=0,1,\mathrm{},[n_f/2]`$) and $`SO(2n_f)U(n_f)`$, respectively. There also are non-confining vacua with no flavor symmetry breaking for $`n_fn_c+1`$, $`n_fn_c+2`$ for $`SU(n_c)`$ and $`USp(2n_c)`$ theories, respectively. With small but generic quark masses, the order parameter of confining vacua is indeed the condensation of magnetic monopoles for every $`U(1)`$ factor on the Coulomb branch à la ’t Hooft, in both types of gauge theories. The massless limit, however, is non-trivial and much more interesting. In $`SU(n_c)`$ theories, in vacua with $`r=1`$ magnetic monopoles are in the fundamental representation of $`U(n_f)`$ flavor group, and are charged under one of the $`U(1)`$’s: flavor-singlet monopoles are charged under other $`U(1)`$’s. Their condensation realizes the confinement and the flavor symmetry breaking at the same time. In vacua labelled by $`r`$, $`2r<n_f/2`$ but $`rn_fn_c`$, the grouping of the associated singularities on the Coulomb branch might suggest the condensation of monopoles in the rank-$`r`$ anti-symmetric tensor representation. Actually, this does not occur. The correctness of the effective action Eq.(32) shows that the low-energy degrees of freedom of these theories are (magnetic) quarks plus a number of singlet monopoles of an effective $`SU(r)\times U(1)^{n_cr}`$ gauge theory. Monopoles in a higher representation of $`SU(n_f)`$ flavor group probably exist semi-classically as seen in a Jackiw–Rebbi type analysis . Such monopoles can be interpreted as “baryons” made of the magnetic quarks, which, interactions being infrared-free, break up before they become massless at singularities on the Coulomb branch. The condensation of the magnetic quarks induces the confinement and flavor symmetry breaking, $`U(n_f)U(r)\times U(n_fr)`$, at the same time. This is how the system avoids falling into a paradox of having too many Nambu-Goldstone multiplets. In the special cases with $`r=n_f/2`$, the interactions among the monopoles are so strong that the low-energy theory describing them is a nontrivial superconformal theory (conformal invariance explicitly broken by the adjoint or quark masses). Although the symmetry breaking pattern is known ($`U(n_f)U(n_f/2)\times U(n_f/2)`$), the low energy degrees of freedom are fields whose interactions are not described by a local action. Finally, in the group of vacua labelled by $`r=n_fn_c`$, the interactions among monopoles are described by an effective infrared-free $`SU(n_fn_c)`$ gauge theory. There are two physically distinct groups of vacua in this case: one in which the magnetic quarks condense (i.e. confinement phase) with the unbroken symmetry $`U(n_fn_c)\times U(n_c)`$, and the other with no magnetic-quark condensation and hence with unbroken $`U(n_f)`$ symmetry (i.e. the free magnetic phase). In $`USp(2n_c)`$ theories, physics for non-vanishing and equal quark masses resembles that in the vacua with generic $`r`$ of $`SU(n_c)`$ theory. In the massless limit, however, where the flavor group enlarges to $`SO(2n_f)`$, the situation is more similar to the $`r=n_f/2`$ case of $`SU(n_c)`$: the low-energy degrees of freedom are fields with relatively non-local interactions, and the effective theory is a non-trivial superconformal one. Although no local effective Lagrangian is available, we know that the flavor $`SO(2n_f)`$ symmetry is spontaneously broken to diagonal $`U(n_f)`$ symmetry in all confining vacua. For large number of flavor, there are also vacua in free-magnetic phase. A more extensive account of our analysis will appear elesewhere. Acknowledgment One of the authors (K.K.) thanks Lawrence Berkeley National Laboratory, University of California, Berkeley, and ITP, University of California Santa Barbara, for their warm hospitalities. Part of the work was done during the workshop, “Supersymmetric Gauge Dynamics and String Theory” at ITP, UCSB to which two of us (G.C. and K.K.) participated. The authors acknowledge useful discussions with Philip Argyres, Prem Kumar, Misha Shifman and Arkady Vainshtein. This research was supported in part by the National Science Foundation under Grant No. PHY-94-07194, PHY-95-14797, and in part by the Director, Office of Science, Office of High Energy and Nuclear Physics, Division of High Energy Physics of the U.S. Department of Energy under Contract DE-AC03-76SF00098.
warning/0001/math-ph0001018.html
ar5iv
text
# Propagation of Molecular Chaos by Quantum Systems and the Dynamics of the Curie-Weiss Model ## 1 Introduction The infinite-particle dynamics of spin models with finite-range interactions — such as the Ising model — can be defined without difficulty in the norm limit of the local (finite-particle) dynamics \[Section 7.6\]. For mean-field spin models such as the Curie-Weiss model, the infinite-particle dynamics can only be defined in certain representations of the infinite-particle algebra, as the limit in the strong operator topology of the local dynamics . The purpose of this note is to introduce a new approach to the quantum mean-field dynamics via the propagation of quantum molecular chaos. The concept of quantum molecular chaos enables us to comprehend the infinite-particle limit of the Curie-Weiss dynamics without constructing an infinite-particle dynamics. For classical mean-field systems, the theory of the propagation of molecular chaos enables one to study the effective dynamics of finite groups of particles without defining dynamics of infinite particle states. We can achieve the same end in the quantum context by utilizing the analog for quantum systems of theory of the propagation of molecular chaos. This device is exploited in , where a quantum version of propagation of molecular chaos is used to derive the Vlaosov equation from the dynamics of quantum particles in the continuum. Their approach was inspired by , wherein the Vlasov equation was derived from the propagation of molecular chaos by mean-field systems of classical particles. The concept of molecular chaos dates back to Boltzmann . In order to derive the fundamental equation of the kinetic theory of gases, Boltzamnn assumed that the molecules of a nonequilibrium gas were in a state of “molecular disorder.” Nowadays, the term “molecular chaos” connotes a system of classical particles that may be regarded as having stochastically independent and identically distributed random positions and momenta. The state of any molecularly chaotic system is characterized by the probability law of a single particle of the system, and so the temporal evolution of any system that is at all times in a state of molecular chaos reduces to that of the probability law of a single particle. Both Boltzmann’s equation for dilute gases and Vlasov’s equation for dilute plasmas may be interpreted as equations that describe the dynamics of the position-velocity distribution $`f(𝐱,𝐯)d𝐱d𝐯`$ of a single particle in a molecularly chaotic gas or plasma. Because gases and plasmas remain in a molecularly chaotic state once they have entered one, i.e., because they “propagate molecular chaos,” the kinetic equations of Boltzmann and Vlasov can be thought of as evolution equations for a single-particle distribution $`f(𝐱,𝐯)d𝐱d𝐯`$. The concept of propagation of molecular chaos is due to Kac , who called it “propagation of the Boltzmann property” and used it to derive the homogenous Boltzmann equation from the infinite particle asymptotics of certain Markovian gas models. This idea was further developed in work by . McKean proved the propagation of chaos by systems of interacting diffusions. See for more recent work on and some generalizations of McKean’s propagation of chaos. For two good surveys of propagation of chaos and its applications, see . This paper is organized as follows. Quantum molecular chaos is defined in Section 2, and related to classical molecular chaos. Examples of quantum molecular chaos are provided; it is shown that sequences of canonical states are often molecularly chaotic. In Section 3 we define the propagation of quantum molecular chaos. We then prove that the Curie-Weiss model propagates molecular chaos and solve the mean-field dynamical equation for the single-particle state. ## 2 Quantum Molecular Chaos The definition of molecular chaos current in the probability literature is equivalent to the following : ###### Definition 1 Let $`S`$ be a separable metric space. For each $`n`$, let $`\rho _n`$ be a symmetric probability measure on $`S^n`$, the $`n`$-fold Cartesian power of $`S`$. (“Symmetric” means that the measures of rectangles are invariant under permutations of the coordinate axes.) Let $`\rho `$ be a probability measure on $`S`$. The sequence $`\{\rho _n\}`$ is $`\rho `$-chaotic if the $`k`$-dimensional marginal distributions $`\rho _n^{(k)}`$ converge (weakly) to $`\rho ^k`$ as $`n\mathrm{}`$, for each fixed $`k`$. The quantum analog of a probability measure is a state on a C\*-algebra with identity. A state on a C\*-algebra with identity $`𝒜`$ is a positive linear functional on $`𝒜`$ that equals $`1`$ at the identity element. The space of states on $`𝒜`$ endowed with the weak\* topology will be denoted $`𝒮(𝒜)`$. Molecular chaos is an attribute of certain sequences of symmetric probability measures; we now define quantum molecular chaos to be an attribute of certain sequences of symmetric states. ###### Definition 2 (Quantum Molecular Chaos) Let $`𝒜`$ be a C\*-algebra with identity, and denote the $`n`$-th (spatial) tensor power of $`𝒜`$ by $`^n𝒜`$. Let $`\rho `$ be a state on $`𝒜`$. For each $`n`$, let $`\rho _n`$ be a symmetric state on $`^n𝒜`$, that is, a state on $`^n𝒜`$ that satisfies $$\rho _n(A_1\mathrm{}A_n)=\rho _n(A_{\pi (1)}𝒜_{\pi (2)}\mathrm{}A_{\pi (n)})$$ for all permutations $`\pi `$ of $`\{1,2,\mathrm{},n\}`$. For each $`kn`$, let $`\rho _n^{(k)}𝒮(^k𝒜)`$ be defined by $$\rho _n^{(k)}(B)=\rho _n(B\mathrm{𝟏}\mathrm{𝟏}\mathrm{}\mathrm{𝟏}),$$ for all $`B^k𝒜`$, and let $`\rho ^k`$ be defined by the condition that, for all $`A_1,A_2,\mathrm{},A_k𝒜`$, $$\rho ^k(A_1A_2\mathrm{}A_k)=\rho (A_1)\rho (A_2)\mathrm{}\rho (A_k).$$ The sequence $`\{\rho _n\}`$ is $`\rho `$-chaotic if, for each $`k`$, the states $`\rho _n^{(k)}`$ converge weakly\* to $`\rho ^k`$ in $`𝒮(^k𝒜)`$ as $`n\mathrm{}`$. The sequence $`\{\rho _n\}`$ is molecularly chaotic if it is $`\rho `$-chaotic for some state $`\rho `$ on $`𝒜`$. Suppose $`𝒜`$ is the algebra generated by the observables for a single particle of a certain species. The algebra generated by all single-particle observables in a system of $`n`$ distinguishable particles of the same species is $`^n𝒜`$, and states on $`^n𝒜`$ correspond to statistical ensembles of those $`n`$-particle systems. Quantum molecular chaos of a sequence of $`n`$-particle states expresses a condition of quasi-independence of the particles when the number of particles is very large. Quantum molecular chaos is related to classical molecular chaos as follows: ###### Theorem 1 Let $`𝒜`$ be a C\* algebra with identity $`\mathrm{𝟏}`$ and for each $`n`$ let $`\rho _n`$ be a symmetric state on $`^n𝒜`$. The following are equivalent: * (i) The sequence $`\{\rho _n\}`$ is $`\rho `$-chaotic in the sense of Definition 2. * (ii) For each pair of positive elements $`Q_0`$ and $`Q_1`$ satisfying $`Q_0+Q_1=\mathrm{𝟏}`$, the sequence of probability measures $`\{P_n\}`$ on $`\{0,1\}^n`$ defined by $$P_n(j_1,j_2,\mathrm{},j_n)=\rho _n(Q_{j_1}Q_{j_2}\mathrm{}Q_{j_n}),$$ is $`P`$-chaotic in the classical sense of Definition 1, where $`P`$ is the probability measure on $`\{0,1\}`$ defined by $$P(j)=\rho (Q_j).$$ Proof: It is clear from Definition 2 that (i) $``$ (ii). The rest of this proof is devoted to showing that (i) $``$ (ii). We first establish the following claim: Let $`𝒫(𝒮(𝒜))`$ denote the space of regular Borel probability measures on the state space of $`𝒜`$, endowed with the weak\* topology as the dual of $`C(𝒮(𝒜))`$. Let $`\sigma `$ be any state on $`𝒜`$. The measure $`\delta _\sigma `$, a point mass at $`\sigma `$, is is the only measure $`\mu 𝒫(𝒮(𝒜))`$ such that $$\mu \{\tau 𝒮(𝒜):|\tau (Q)\sigma (Q)|ϵ\}=0$$ (1) for every $`ϵ>0`$ and $`Q𝒜`$ with $`\mathrm{𝟎}Q\mathrm{𝟏}`$. To prove this claim, first note that (1) holds for every element of $`𝒜`$ if it holds for those elements $`Q`$ with $`\mathrm{𝟎}Q\mathrm{𝟏}`$, since every element of $`𝒜`$ is a linear combination of such positive elements. Since (1) holds for every $`Q𝒜`$, the Borel measure $`\mu `$ is supported on arbitrarily small basic open neigborhoods of $`\sigma 𝒮(𝒜)`$, whence it follows that $`\mu (\{\sigma \})=1`$, since $`\mu `$ is a regular measure. Next we define a couple of homeomorphisms: Let $`^{\mathrm{}}𝒜`$ denote the inductive limit of the spatial tensor products $`^n𝒜`$. A state $`\sigma ^{\mathrm{}}𝒜`$ is called symmetric if $`\sigma (A_1A_2\mathrm{}A_n\mathrm{𝟏}\mathrm{𝟏}\mathrm{})`$ $`=`$ $`\sigma (A_{\pi (1)}A_{\pi (2)}\mathrm{}A_{\pi (k)}A_{k+1}\mathrm{}A_n\mathrm{𝟏}\mathrm{𝟏}\mathrm{})`$ for all elements $`A_1,A_2,\mathrm{},A_n`$ of $`𝒜`$, and all permutations $`\pi `$ of $`\{1,2,\mathrm{},k\}`$, $`kn`$. Denote the space of symmetric states on $`^{\mathrm{}}𝒜`$ by $`𝒮_{sym}(^{\mathrm{}}𝒜)`$. Størmer’s theorem \[Theorem 2.8\] states that there exists an affine homeomorphism $`\mathrm{\Phi }`$ from the space $`𝒫(𝒮(𝒜))`$ of regular probability measures on $`𝒮(𝒜)`$ to $`𝒮_{sym}(^{\mathrm{}}𝒜)`$, such that $`\mathrm{\Phi }(\delta _\mu )=\mu ^{\mathrm{}}`$. The classical precursor of Størmer’s theorem, de Finetti’s theorem , states that there exists an affine homeomorphism $`\mathrm{\Xi }:𝒫(𝒫(\{0,1\}))𝒫_{sym}(\{0,1\}^{\mathrm{}})`$ such that $`\mathrm{\Xi }(\delta _p)=p^{\mathrm{}}`$. Now assume that condition (ii) holds. Let $`\tau `$ be an arbitrary but fixed state on $`𝒜`$, and for each $`n`$, extend $`\rho _n^n𝒜`$ to the state $$\stackrel{~}{\rho }_n=\rho _n\tau \tau \tau \tau \tau \mathrm{}$$ in $`𝒮(^{\mathrm{}}𝒜)`$ Since $`𝒮(^{\mathrm{}}𝒜)`$ is weak\* compact, every subsequence of $`\{\stackrel{~}{\rho }_n\}`$ has cluster points; condition (ii) will used to prove that $`\rho ^{\mathrm{}}`$ is the only cluster point of $`\{\stackrel{~}{\rho }_n\}`$. It will follow that $`\{\stackrel{~}{\rho }_n\}`$ converges to $`\rho ^{\mathrm{}}`$, which implies that $`\{\rho _n\}`$ is $`\rho `$-chaotic. Let $`\mu 𝒮(^{\mathrm{}}𝒜)`$ be any cluster point of $`\{\stackrel{~}{\rho }_n\}`$, the limit of the subsequence $`\{\stackrel{~}{\rho }_{n_k}\}`$. Because of the increasing symmetry of the $`\stackrel{~}{\rho }_{n_k}`$, the state $`\mu `$ is symmetric: $`\mu 𝒮_{sym}(^{\mathrm{}}𝒜)`$. Suppose that $`\mu \rho ^{\mathrm{}}`$ (this assumption will lead to a contradiction). Then $`\mathrm{\Phi }^1(\mu )\delta _\rho `$ and we have shown that there must exist $`\mathrm{𝟎}<Q<\mathrm{𝟏}`$ and $`ϵ>0`$ such that $$\mathrm{\Phi }^1(\mu )\{\sigma :|\sigma (Q)\rho (Q)|ϵ\}>0.$$ (2) Set $`Q_0=Q`$ and $`Q_1=\mathrm{𝟏}Q`$. Define $`P:𝒮(𝒜)𝒫(\{0,1\})`$, mapping $`\sigma `$ to $`P_\sigma `$, by $$P_\sigma (j)=\sigma (Q_j);j\{0,1\}.$$ Define $`P^{\mathrm{}}:𝒮(^{\mathrm{}}𝒜)𝒫(\{0,1\}^{\mathrm{}})`$, mapping $`\sigma `$ to $`P_\sigma ^{\mathrm{}}`$, by $$P_\sigma ^{\mathrm{}}\{(x_1,x_2,\mathrm{}):x_1=j_1,\mathrm{},x_n=j_n\}=\sigma (Q_{j_1}\mathrm{}Q_{j_n}\mathrm{𝟏}\mathrm{}).$$ Condition (ii) implies that $`P^{\mathrm{}}(\stackrel{~}{\rho }_{n_k})(P_\rho )^{\mathrm{}}`$. Since $`P^{\mathrm{}}`$ is continuous and $`\stackrel{~}{\rho }_{n_k}\mu `$, it follows that $$P^{\mathrm{}}(\mu )=(P_\rho )^{\mathrm{}}.$$ (3) The composite map $`\mathrm{\Xi }^1P^{\mathrm{}}\mathrm{\Phi }`$ is affine and continuous, and it maps $`\delta _\sigma `$ to $`\delta _{P_\sigma }`$ for every $`\sigma 𝒮(𝒜)`$. The map $`\stackrel{~}{P}:𝒫(𝒮(𝒜))𝒫(𝒫(\{0,1\}))`$ induced by $`P:𝒮(𝒜)𝒫(\{0,1\})`$ is also affine and continuous, and also maps $`\delta _\sigma `$ to $`\delta _{P_\sigma }`$ for every $`\sigma 𝒮(𝒜)`$, so $`\stackrel{~}{P}`$ must equal $`\mathrm{\Xi }^1P^{\mathrm{}}\mathrm{\Phi }`$ by the Krein-Milman theorem. That is, the following diagram commutes: $$\begin{array}{ccc}𝒫(𝒮(𝒜))& \stackrel{\mathrm{\Phi }}{}& 𝒮_{sym}(^{\mathrm{}}𝒜)\\ \stackrel{~}{P}& & P^{\mathrm{}}& & \\ 𝒫(𝒫(\{0,1\}))& \stackrel{\mathrm{\Xi }}{}& 𝒫_{sym}(\{0,1\}^{\mathrm{}})\end{array}$$ By equation (3) and the commutativity of the preceding diagram, $$\stackrel{~}{P}(\mathrm{\Phi }^1(\mu ))=\mathrm{\Xi }^1(P^{\mathrm{}}(\mu ))=\delta _{P_\rho }.$$ (4) But equation (2) implies that $$\stackrel{~}{P}(\mathrm{\Phi }^1(\mu ))\{p𝒫(\{0,1\}):|p(0)P_\rho (0)|ϵ\}>0,$$ and this contradicts (4). $`\mathrm{}`$ ###### Corollary 1 Let $`𝒜`$ be a C\* algebra with identity, and for each $`n`$ let $`\rho _n`$ be a symmetric state on $`^n𝒜`$. If $`\rho _n^{(2)}`$ converges to $`\rho \rho `$ then $`\{\rho _n\}`$ is $`\rho `$-chaotic. Proof: Let $`Q_0`$ and $`Q_1`$ be any positive elements of $`𝒜`$ such that $`Q_0+Q_1=\mathrm{𝟏}`$, and let $`P_n`$ and $`P`$ be as in the statement of Theorem 1. The measures $`P_n`$ are symmetric, and $`P_n^{(2)}`$ converges to $`PP`$ since $`\rho _n^{(2)}`$ converges to $`\rho \rho `$. This suffices to imply that $`\{P_n\}`$ is $`P`$-chaotic . The $`\rho `$-chaos of $`\{\rho _n\}`$ now follows from Theorem 1. $`\mathrm{}`$ The following theorem shows that sequences of canonical states for mean-field systems are often molecularly chaotic: ###### Theorem 2 Let $`V`$ be an operator on $`^d`$ such that $`V(xy)=V(yx)`$ for all $`x,y^d`$. Let $`V_{1,2}^n`$ denote the operator on $`^n^d`$ defined by $$V_{1,2}^n(x_1x_2\mathrm{}x_n)=V(x_1x_2)x_3\mathrm{}x_n,$$ and for each $`i<jn`$, define $`V_{ij}^n`$ as acting similarly on the $`i^{th}`$ and $`j^{th}`$ factors of each simple tensor. Define the states $`\rho _n𝒮(^n^d)`$ by $$\begin{array}{ccc}\rho _n(A)=\frac{1}{Z}\mathrm{Tr}\left(e^{H_n}A\right);& H_n=\frac{1}{n}\underset{i<j}{}V_{ij}^n;& Z=\mathrm{Tr}\left(e^{H_n}\right).\end{array}$$ The sequence $`\{\rho _n\}`$ is $`\rho `$-chaotic if the density operator for $`\rho `$ is the unique minimizer of the free energy $$F[D]=\frac{1}{2}\mathrm{Tr}((DD)V)+\mathrm{Tr}(D\mathrm{log}D).$$ If $`V`$ is positive definite then $`F`$ has a unique minimizer. Proof Sketch: This theorem is the quantum version of Theorems 2 and 4 of . The properties of classical entropy that Messer and Spohn used to prove those theorems are equally true for the von Neumann entropy $`\mathrm{Tr}(D\mathrm{log}D)`$ of density operators, at least when $`D`$ operates on $`^d`$. The necessary properties of von Neumann entropy are proved in . $`\mathrm{}`$ ## 3 The Curie-Weiss Model Propagates Chaos A sequence of $`n`$-particle dynamics “propagates chaos” if molecularly chaotic sequences of initial distributions remain molecularly chaotic for all time under the $`n`$-particle dynamical evolutions. In the classical context, the $`n`$-particle dynamics are Markovian. Accordingly, in , we defined propagation of chaos in terms of Markov transition functions: ###### Definition 3 For each $`n`$, let $`K_n:S^n\times \sigma (S^n)\times [0,\mathrm{})[0,1]`$ be a Markov transition function which commutes with permutations in the sense that $$K_n(𝐱,E,t)=K_n(\pi 𝐱,\pi E,t)$$ for all permutations $`\pi `$ of the $`n`$ coordinates of $`𝐱`$ and the points of $`ES^n`$, and for all $`t0`$. (Here, $`\sigma (S^n)`$ denotes the Borel $`\sigma `$-field of $`S^n`$.) The sequence $`\{K_n\}_{n=1}^{\mathrm{}}`$ propagates chaos if, for all $`t0`$, the molecular chaos of a sequence $`\{\rho _n\}`$ entails the molecular chaos of the sequence $`\left\{_{S^n}K_n(𝐱,,t)\rho _n(d𝐱)\right\}`$. The quantum analog of a Markov transition function is a completely positive unital map. A linear map $`\varphi :𝒜_1𝒜_2`$ of C\* algebras is completely positive if, for each $`n`$, the map from $`𝒜_1(^n)`$ to $`𝒜_2(^n)`$ that sends $`AB`$ to $`\varphi (A)B`$ is positive . Propagation of chaos is an attribute of certain sequences of Markov transition functions; we now define quantum propagation of chaos to be an attribute of certain sequences of completely positive unital maps. ###### Definition 4 (Propagation of Quantum Molecular Chaos) For each $`n`$, let $`\varphi _n`$ be a completely positive map from $`^n𝒜`$ to itself that fixes the unit $`\mathrm{𝟏}\mathrm{}\mathrm{𝟏}^n𝒜`$ and which commutes with permutations, i.e., such that $$\varphi _n(A_{\pi (1)}A_{\pi (2)}\mathrm{}A_{\pi (n)})=\pi \varphi _n(A_1A_2\mathrm{}A_n)$$ (5) for all permutations $`\pi `$ of $`\{1,2,\mathrm{},n\}`$, where $`\pi `$ denotes the operator on $`^n𝒜`$ defined by $$\pi (B_1B_2\mathrm{}B_n)=B_{\pi (1)}B_{\pi (2)}\mathrm{}B_{\pi (n)}$$ for all $`B_1,B_2,\mathrm{},B_n𝒜`$. The sequence $`\{\varphi _n\}`$ propagates chaos if the molecular chaos of a sequence of states $`\{\rho _n\}`$ entails the molecular chaos of the sequence $`\{\rho _n\varphi _n\}`$. Consider the case where $`𝒜`$ is the algebra generated by the observables for a single particle of a certain species, so that $`^n𝒜`$ is the algebra generated by all single-particle observables in a system of $`n`$ distinguishable particles of that species. For each $`n`$, let the dynamics of the $`n`$-particle system be given by a Hamiltonian operator $`H_n^n𝒜`$. In the Heisenberg version of quantum dynamics, if $`A`$ is the operator that corresponds to measurement of a certain observable quantity at $`t=0`$, the operator corresponding to the measurement of the same quantity at time $`t>0`$ equals $`e^{iH_nt/\mathrm{}}Ae^{iH_nt/\mathrm{}}`$. The maps $`\varphi _{n,t}:^n𝒜^n𝒜`$ defined by $$\varphi _{n,t}(A)=e^{iH_nt/\mathrm{}}Ae^{iH_nt/\mathrm{}}$$ (6) are completely positive, and if they satisfy condition (5) one may ask whether the sequence $`\{\varphi _{n,t}\}`$ propagates chaos. We conjecture that chaos always propagates when the $`n`$-particle Hamiltonians $`H_n`$ are as follows: Let $`𝒜=(^d)`$, the algebra of all bounded operators on $`^d`$. The algebra $`^n𝒜`$ is isomorphic to $`(^n^d)`$, and states $`\tau 𝒮(^n𝒜)`$ correspond to density operators $`D_\tau `$ on $`(^n^d)`$ via the equation $`\tau (A)=\mathrm{Tr}(D_\tau A)`$. Suppose that $`V`$ is a Hermitian operator on $`^d^d`$ that is symmetric in the sense that $`V(xy)=V(yx)`$ for all $`x,y^d`$. Let $`V_{1,2}^n`$ denote the operator on $`^n^d`$ defined by $$V_{1,2}^n(x_1x_2\mathrm{}x_n)=V(x_1x_2)x_3\mathrm{}x_n,$$ and for each $`i<jn`$, define $`V_{ij}^n`$ similarly (as acting on the $`i^{th}`$ and $`j^{th}`$ factors of each simple tensor). Define the $`n`$-particle Hamiltonian $`H_n`$ as the sum of the pair potentials $`V_{ij}^n`$, with a $`1/n`$ scaling of the coupling constant: $$H_n=\frac{1}{n}\underset{i<j}{}V_{ij}^n.$$ (7) ###### Conjecture 1 The sequence $`\{\varphi _{n,t}\}`$ defined in equation (6) propagates chaos: If $`\{\rho _n\}`$ is $`\rho `$-chaotic then $`\{\rho _n\varphi _{n,t}\}`$ is $`\rho _t`$-chaotic, where the density operator for $`\rho _t`$ is the solution at time $`t`$ of $`{\displaystyle \frac{}{t}}D`$ $`=`$ $`{\displaystyle \frac{i}{\mathrm{}}}[V,DD]^{(1)}`$ $`D(0)`$ $`=`$ $`D_\rho .`$ Here $`[V,DD]^{(1)}`$ denotes a contraction of $`[V,DD]`$: if $`\{y_i\}`$ is any orthonormal basis of $`^d`$ and $`x^d`$, $$[V,DD]^{(1)}(x)=\underset{i}{}<[V,DD](xy_i),(xy_i)>.$$ This conjecture will now be verified for the Curie-Weiss model of ferromagnetism. In this model, the ferromagnetic material is modelled by a crystal in which the spin angular momentum of each atom is coupled to the average spin and to an external magnetic field. In case the applied magnetic field is directed along the $`z`$-axis, we may make the approximation that only the $`z`$-components of the spins are coupled to each other and the external field. As in , we consider the case of spin-$`\frac{1}{2}`$ atoms, so that the space of pure spin states of a single particle is $`^2`$, and the observables corresponding to the measurement of the $`x,y`$ and $`z`$ components of spin are the Pauli spin operators $$\begin{array}{ccc}\sigma ^x=\frac{\mathrm{}}{2}\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)& \sigma ^y=\frac{\mathrm{}}{2}\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right)& \sigma ^z=\frac{\mathrm{}}{2}\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right).\end{array}$$ The space of pure states of an $`n`$-spin system is $`^n^2`$. For each $`in`$, if $`A`$ is an operator on $``$, let $`A_i`$ denote the operator on $`^n^2`$ defined by $$A_i(v_1v_2\mathrm{}v_n)=v_1\mathrm{}A(v_i)\mathrm{}v_n.$$ If $`A`$ is Hermitian, $`A_i`$ corresponds to the measurement of the spin observable $`A`$ at the $`i^{th}`$ lattice site. The Hamiltonian for the $`n`$-site Curie-Weiss model is $$_n=J\underset{i=1}{\overset{n}{}}\left(\sigma _i^z\frac{_j\sigma _j^z}{n}\right)H\underset{i=1}{\overset{n}{}}\sigma _i^z=\frac{1}{n}\underset{i,j=1}{\overset{n}{}}\left(J\sigma _i^z\sigma _j^zH\sigma _i^z\right),$$ where $`J`$ is a positive coupling constant and $`H`$ is another constant proportional to the strength of the external magnetic field. Since $`^n(^2)`$ is isomorphic to $`(^n^2)`$, states $`\rho _n𝒮(^n(^2))`$ correspond to density operators $`D_{\rho _n}`$ in $`^n^2`$. ¿From definition (6) and the fact that $`\mathrm{Tr}(AB)=\mathrm{Tr}(BA)`$, $`\rho _n\varphi _{n,t}(A)`$ $`=`$ $`\mathrm{Tr}(D_{\rho _n}e^{i_nt/\mathrm{}}Ae^{i_nt/\mathrm{}})`$ $`=`$ $`\mathrm{Tr}(e^{i_nt/\mathrm{}}D_{\rho _n}e^{i_nt/\mathrm{}}A).`$ Therefore, $`\rho _n\varphi _{n,t}`$ has the density operator $$D_{\rho _n\varphi _{n,t}}=e^{i_nt/\mathrm{}}D_{\rho _n}e^{i_nt/\mathrm{}}.$$ (9) For any state $`\rho `$ on $`(^2)`$, let $`D_\rho `$ denote the corresponding density operator, and let $`\left[D_{\rho (t)}\right]`$ denote a $`2\times 2`$ matrix that represents $`D_\rho `$. ###### Theorem 3 The sequence of Hamiltonians $`\{_n\}`$ propagates chaos. If $`\{\rho _n\}`$ is a $`\rho `$-chaotic sequence of states with $`\left[D_\rho \right]=\left(\begin{array}{cc}a& c\\ \overline{c}& d\end{array}\right)`$, then for each $`t0`$, $`\{\rho _n\varphi _{n,t}\}`$ is $`\rho (t)`$-chaotic, where $$\left[D_{\rho (t)}\right]=\left(\begin{array}{cc}a& ce^{it(H+\mathrm{}J(ad))}\\ \overline{c}e^{it(H+\mathrm{}J(ad))}& d\end{array}\right).$$ Proof of Theorem 3: If $`𝐱\{0,1\}^n`$ for some $`n`$, let $`𝒩(𝐱)`$ denote the number of $`1`$s in $`𝐱`$: $$𝒩(𝐱)=\underset{i=1}{\overset{n}{}}x_i$$ if $`𝐱=(x_1,x_2,\mathrm{},x_n)`$. If $`g:\{0,1\}^n`$ is a symmetric function, then $`g(𝐱)`$ depends on $`𝐱`$ only through $`𝒩(𝐱)`$, so that $`g(𝐱)=g(𝐲)`$ if $`𝒩(𝐱)=𝒩(𝐲)`$. ###### Lemma 1 For each $`n`$, let $`f_n:\{0,1\}^n`$ be a symmetric function. Suppose that * (a) there exists $`B<\mathrm{}`$ such that $`_{𝐬\{0,1\}^n}\left|f_n(𝐬)\right|B`$ for all $`n`$, and * (b) there exists $`c`$ and $`f:\{0,1\}`$ such that, for $`k=0,1,2,\mathrm{}`$, $$\underset{z_1,z_2,\mathrm{},z_{nk}}{}f_n(x_1,\mathrm{},x_k,z_1,\mathrm{},z_{nk})cf(x_1)f(x_2)\mathrm{}f(x_k)$$ as $`n\mathrm{}`$. Then * (i) For all $`GC_{}([0,1])`$, $$\underset{n\mathrm{}}{lim}\underset{𝐱\{0,1\}^n}{}f_n(𝐱)G\left(\frac{𝒩(𝐱)}{n}\right)=cG(f(1))$$ * (ii) If $`c0`$ then $`0f(0),f(1)1`$ and $`f(0)+f(1)=1`$. Proof of Lemma: Let $$F_n(j)=\underset{𝒩(𝐱)=j}{}f_n(𝐱).$$ (10) Condition (b) implies that $$\underset{n\mathrm{}}{lim}\underset{z_1,z_2,\mathrm{},z_{nk}}{}f_n(1,1,\mathrm{},1,z_1,\mathrm{},z_{nk})=c\left(f(1)\right)^k$$ for all $`k`$. Grouping the summands for which exactly $`j`$ coordinates of $`(1,1,\mathrm{},1,z_1,\mathrm{},z_{nk})`$ equal $`1`$, we find that $$\underset{n\mathrm{}}{lim}\underset{j=k}{\overset{n}{}}F_n(j)\frac{\left(\begin{array}{c}nk\\ jk\end{array}\right)}{\left(\begin{array}{c}n\\ j\end{array}\right)}=c\left(f(1)\right)^k.$$ (11) Now $$\underset{j=k}{\overset{n}{}}F_n(j)\frac{\left(\begin{array}{c}nk\\ jk\end{array}\right)}{\left(\begin{array}{c}n\\ j\end{array}\right)}=\underset{j=k}{\overset{n}{}}F_n(j)\frac{j(j1)\mathrm{}(jk+1)}{n(n1)\mathrm{}(nk+1)},$$ while $`\left|{\displaystyle \underset{j=k}{\overset{n}{}}}F_n(j){\displaystyle \frac{j(j1)\mathrm{}(jk+1)}{n(n1)\mathrm{}(nk+1)}}{\displaystyle \underset{j=0}{\overset{n}{}}}F_n(j)\left({\displaystyle \frac{j}{n}}\right)^k\right|`$ $``$ $`{\displaystyle \underset{j=0}{\overset{n}{}}}\left|F_n(j)\right|\left(\left({\displaystyle \frac{k}{n}}\right)^n+\underset{jk}{\mathrm{max}}\left\{\left|{\displaystyle \frac{j(j1)\mathrm{}(jk+1)}{n(n1)\mathrm{}(nk+1)}}\left({\displaystyle \frac{j}{n}}\right)^k\right|\right\}\right)`$ $``$ $`B\left[\left({\displaystyle \frac{k}{n}}\right)^n+\underset{jk}{\mathrm{max}}\left\{\left({\displaystyle \frac{j}{n}}\right)^k\left({\displaystyle \frac{jk+1}{n}}\right)^k\right\}\right]`$ $``$ $`B\left[\left({\displaystyle \frac{k}{n}}\right)^n+{\displaystyle \frac{1}{n^k}}\underset{jk}{\mathrm{max}}\left\{\left(\begin{array}{c}k\\ 1\end{array}\right)j^{k1}(k1)+\left(\begin{array}{c}k\\ 2\end{array}\right)j^{k2}(k1)^2+\mathrm{}+(k1)^k\right\}\right]`$ $``$ $`B\left[\left({\displaystyle \frac{k}{n}}\right)^n+{\displaystyle \frac{1}{n}}\left\{\left(\begin{array}{c}k\\ 1\end{array}\right)k+\left(\begin{array}{c}k\\ 2\end{array}\right)k^2+\mathrm{}+k^k\right\}\right]`$ $``$ $`B\left[\left({\displaystyle \frac{k}{n}}\right)^n+{\displaystyle \frac{1}{n}}(1+k)^k\right],`$ and therefore $$\underset{n\mathrm{}}{lim}\underset{j=k}{\overset{n}{}}F_n(j)\frac{\left(\begin{array}{c}nk\\ jk\end{array}\right)}{\left(\begin{array}{c}n\\ j\end{array}\right)}=\underset{n\mathrm{}}{lim}\underset{j=0}{\overset{n}{}}F_n(j)\left(\frac{j}{n}\right)^k.$$ (14) Condition (a) implies that $`_j\left|F_n(j)\right|B`$ for all $`n`$. This bound, equations (11) and (14), and the fact that any continuous function on $`[0,1]`$ can be approximated uniformly by polynomials, imply that $$\underset{n\mathrm{}}{lim}\underset{j=0}{\overset{n}{}}F_n(j)G\left(\frac{j}{n}\right)=cG\left(f(1)\right)$$ for all $`GC_{}([0,1])`$. This establishes (i), in view of the definition (10) of $`F_n`$. If $`c0`$, conclusion (i) implies that $`f(1)[0,1]`$. Condition (b) for $`k=0`$ and $`k=1`$ states that $`\underset{n\mathrm{}}{lim}{\displaystyle \underset{z_1,z_2,\mathrm{},z_n}{}}f_n(z_1,z_2,\mathrm{},z_n)`$ $`=`$ $`c`$ $`\underset{n\mathrm{}}{lim}{\displaystyle \underset{z_1,z_2,\mathrm{},z_{n1}}{}}f_n(1,z_1,\mathrm{},z_{n1})`$ $`=`$ $`cf(1)`$ $`\underset{n\mathrm{}}{lim}{\displaystyle \underset{z_1,z_2,\mathrm{},z_{n1}}{}}f_n(0,z_1,\mathrm{},z_{n1})`$ $`=`$ $`cf(0),`$ so it follows that $`f(0)+f(1)=1`$ if $`c0`$. This establishes (ii), concluding the proof of the lemma. $`\mathrm{}`$ Let $`e_1=(1,0)`$ and $`e_2=(0,1)`$. For each $`n`$, let $$_n=\{e_{j_1}e_{j_2}\mathrm{}e_{j_n}^n^2|j_1,j_2,\mathrm{},j_n\{1,2\}\}.$$ Denote by $`[A]_{j_1,\mathrm{},j_n}^{k_1,\mathrm{},k_n}`$ the matrix elements for an operator $`A`$ on $`^n^2`$ relative to the basis $`_n`$, that is, for $`j_1,k_1,\mathrm{},j_n,k_n\{1,2\}`$, let $$[A]_{j_1,\mathrm{},j_n}^{k_1,\mathrm{},k_n}=A(e_{j_1}e_{j_2}\mathrm{}e_{j_n}),e_{k_1}e_{k_2}\mathrm{}e_{k_n}.$$ The sequence $`\{\rho _n\}`$ is $`\rho `$-chaotic. This implies that, for each $`k`$, $$\underset{z_1,z_2,\mathrm{},z_{nk}\{1,2\}}{}\left[D_{\rho _n}\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_{nk}}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_{nk}}\left[D_\rho \right]_{x_1}^{y_1}\left[D_\rho \right]_{x_2}^{y_2}\mathrm{}\left[D_\rho \right]_{x_k}^{y_k},$$ as $`n\mathrm{}`$. To show that $`\{\rho _n\varphi _{n,t}\}`$ is $`\rho (t)`$-chaotic, it suffices to show that for all $`k`$, $$\underset{z_1,z_2,\mathrm{},z_{nk}\{1,2\}}{}\left[D_{\rho _n\varphi _{n,t}}\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_{nk}}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_{nk}}\left[D_{\rho (t)}\right]_{x_1}^{y_1}\left[D_{\rho (t)}\right]_{x_2}^{y_2}\mathrm{}\left[D_{\rho (t)}\right]_{x_k}^{y_k}$$ (15) as $`n\mathrm{}`$. We proceed to verify (15). The operator $`_n`$ is diagonalized by the basis $`_n`$, and its diagonal entries are $$[H_n]_{k_1,\mathrm{},k_n}^{k_1,\mathrm{},k_n}=\frac{1}{n}\underset{r,s=1}{\overset{n}{}}\left(J\eta (e_{k_r})\eta (e_{k_s})H\eta (e_{k_r})\right),$$ (16) where $`\eta (e_1)`$ $`=`$ $`+{\displaystyle \frac{\mathrm{}}{2}}`$ $`\eta (e_2)`$ $`=`$ $`{\displaystyle \frac{\mathrm{}}{2}}.`$ Abbreviating $`J\eta (x)\eta (y)H\eta (x)`$ by $`w(x,y)`$, equations (9) and (16) imply that $`[D_{\rho _n\varphi _{n,t}}]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_{nk}}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_{nk}}=`$ $`\mathrm{exp}\left({\displaystyle \frac{it}{\mathrm{}n}}{\displaystyle \underset{r,s=1}{\overset{k}{}}}(w(y_r,y_s)w(x_r,x_s))\right)\mathrm{exp}\left(H{\displaystyle \frac{it}{\mathrm{}}}{\displaystyle \frac{nk}{n}}{\displaystyle \underset{r=1}{\overset{k}{}}}(\eta (x_r)\eta (y_r))\right)\times `$ $`\left[D\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_{nk}}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_{nk}}\mathrm{exp}\left(J{\displaystyle \frac{2it}{\mathrm{}n}}{\displaystyle \underset{r=1}{\overset{k}{}}}{\displaystyle \underset{s=k+1}{\overset{n}{}}}\eta (z_s)(\eta (x_r)\eta (y_r))\right).`$ Therefore, $`\underset{n\mathrm{}}{lim}{\displaystyle \underset{z_1,\mathrm{},z_{nk}\{1,2\}}{}}\left[D_{\rho _n\varphi _{n,t}}\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_{nk}}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_{nk}}`$ $`=`$ $`\mathrm{exp}\left(H{\displaystyle \frac{it}{\mathrm{}}}{\displaystyle \underset{r=1}{\overset{k}{}}}(\eta (x_r)\eta (y_r))\right)\times `$ $`\underset{n\mathrm{}}{lim}{\displaystyle \underset{z_1,\mathrm{},z_{nk}\{0,1\}}{}}\left[D\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_{nk}}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_{nk}}\mathrm{exp}\left(J{\displaystyle \frac{2it}{\mathrm{}n}}{\displaystyle \underset{r=1}{\overset{k}{}}}{\displaystyle \underset{s=k+1}{\overset{n}{}}}\eta (z_s)(\eta (x_r)\eta (y_r))\right)`$ $`=`$ $`\mathrm{exp}\left(H{\displaystyle \frac{it}{\mathrm{}}}{\displaystyle \underset{r=1}{\overset{k}{}}}(\eta (x_r)\eta (y_r))\right)\times `$ $`\underset{n\mathrm{}}{lim}{\displaystyle \underset{z_1,\mathrm{},z_{nk}\{1,2\}}{}}\left[D\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_{nk}}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_{nk}}\mathrm{exp}\left(J{\displaystyle \frac{2it}{\mathrm{}}}{\displaystyle \underset{r=1}{\overset{k}{}}}(\eta (x_r)\eta (y_r)){\displaystyle \frac{1}{n}}{\displaystyle \underset{s=k+1}{\overset{n}{}}}\eta (z_s)\right).`$ The last limit in the preceding equation may be calculated thanks to Lemma 1. To apply the lemma, fix $`x_1,y_1,\mathrm{},x_k,y_k\{1,2\}`$ and define $`f_n:\{1,2\};`$ $`f_n(z_1,z_2,\mathrm{},z_n)=\left[D\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_n}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_n}`$ $`G:[0,1];`$ $`G(s)=\mathrm{exp}\left(Jit{\displaystyle \underset{r=1}{\overset{k}{}}}(\eta (x_r)\eta (y_r))(s(1s))\right).`$ The functions $`f_n`$ are symmetric and satisfy condition (b) of the lemma, with $`f(1)`$ $`=`$ $`\left[D_\rho \right]_{e_1}^{e_1}`$ $`f(2)`$ $`=`$ $`\left[D_\rho \right]_{e_2}^{e_2}`$ $`c`$ $`=`$ $`\left[D_\rho \right]_{x_1}^{y_1}\left[D_\rho \right]_{x_2}^{y_2}\mathrm{}\left[D_\rho \right]_{x_k}^{y_k}.`$ The $`f_n`$ also satisfy condition (a) of the lemma, for $$\left|\left[D\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_n}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_n}\right|\frac{1}{2}\left(\left[D\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_n}^{x_1,\mathrm{},x_k,z_1,\mathrm{},z_n}+\left[D\right]_{y_1,\mathrm{},y_k,z_1,\mathrm{},z_n}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_n}\right)$$ by the positivity of $`D_{\rho _n}`$, so that $$\underset{z_1,\mathrm{},z_{nk}\{1,2\}}{}\left|\left[D\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_n}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_n}\right|\mathrm{tr}(D_{\rho _n})=1.$$ Conclusion (i) of Lemma 1 now reveals that $$\underset{z_1,\mathrm{},z_{nk}\{1,2\}}{}\left[D\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_{nk}}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_{nk}}\mathrm{exp}\left(J\frac{2it}{\mathrm{}}\underset{r=1}{\overset{k}{}}(\eta (x_r)\eta (y_r))\frac{1}{n}\underset{s=k+1}{\overset{n}{}}\eta (z_s)\right)$$ converges to $$\left[D_\rho \right]_{x_1}^{y_1}\left[D_\rho \right]_{x_2}^{y_2}\mathrm{}\left[D_\rho \right]_{x_k}^{y_k}\mathrm{exp}\left(Jit\left(\left[D_\rho \right]_{e_1}^{e_1}\left[D_\rho \right]_{e_2}^{e_2}\right)\underset{r=1}{\overset{k}{}}(\eta (x_r)\eta (y_r))\right),$$ whence $`\underset{n\mathrm{}}{lim}{\displaystyle \underset{z_1,\mathrm{},z_{nk}\{1,2\}}{}}\left[D_{\rho _n\varphi _{n,t}}\right]_{x_1,\mathrm{},x_k,z_1,\mathrm{},z_{nk}}^{y_1,\mathrm{},y_k,z_1,\mathrm{},z_{nk}}`$ $`=`$ $`\mathrm{exp}\left(H{\displaystyle \frac{it}{\mathrm{}}}{\displaystyle \underset{r=1}{\overset{k}{}}}(\eta (x_r)\eta (y_r))\right)\times `$ $`\left[D_\rho \right]_{x_1}^{y_1}\left[D_\rho \right]_{x_2}^{y_2}\mathrm{}\left[D_\rho \right]_{x_k}^{y_k}\mathrm{exp}\left(Jit\left(\left[D_\rho \right]_{e_1}^{e_1}\left[D_\rho \right]_{e_2}^{e_2}\right){\displaystyle \underset{r=1}{\overset{k}{}}}(\eta (x_r)\eta (y_r))\right)`$ $`=`$ $`{\displaystyle \underset{r=1}{\overset{k}{}}}\left[D_\rho \right]_{x_r}^{y_r}\mathrm{exp}\left\{it\left(\eta (x_r)\eta (y_r)\right)\left(H/\mathrm{}+J\left(\left[D_\rho \right]_{e_1}^{e_1}\left[D_\rho \right]_{e_2}^{e_2}\right)\right)\right\}.`$ This shows that $`\{\rho _n\varphi _{n,t}\}`$ is $`\rho (t)`$-chaotic, where $$\left[D_{\rho (t)}\right]_x^y=\left[D_\rho \right]_x^y\mathrm{exp}\left\{it\left(\eta (x)\eta (y)\right)\left(H/\mathrm{}+J\left(\left[D_\rho \right]_{e_1}^{e_1}\left[D_\rho \right]_{e_2}^{e_2}\right)\right)\right\}.$$ Finally, it may be verified that the density operators $`D_{\rho (t)}`$ satisfy equation (1) of Conjecture 1. $`\mathrm{}`$ ## 4 Future Work In future work, we hope to prove Conjecture 1, or at least to prove that the mean-field Heisenberg model propagates chaos. We shall also investigate the propagation of chaos by open systems (coupled to thermal baths) and prove the H-theorem for those processes. ## 5 Acknowledgements I am indebted to William Arveson, Sante Gnerre, Lucien Le Cam, Marc Rieffel, and Geoffrey Sewell for their advice and for their interest in this research. The formulation and proof of Theorem 1 was especially inspired by Elizabeth Kallman. I thank Alexandre Chorin for introducing me to statistical physics and other topics in applied mathematics. This work was supported in part by the Office of Science, Office of Advanced Scientific Computing Research, Mathematical, Information, and Computational Sciences Division, Applied Mathematical Sciences Subprogram, of the U.S. Department of Energy, under Contract No. DE-AC03-76SF00098.
warning/0001/math0001127.html
ar5iv
text
# An explicit formula for PBW quantization ## 0. Introduction We consider (possibly infinite dimensional) Lie algebras over a fixed field $`k`$ of characteristic zero. Let $`𝔤`$ be a Lie algebra, $`S=S𝔤`$ and $`U=U𝔤`$ the symmetric and universal enveloping algebras, and $`^0^1\mathrm{}U`$ the coalgebra filtration $`^n:=k+𝔤+𝔤^2+\mathrm{}+𝔤^n`$. Recall that the Poincaré-Birkhoff-Witt isomorphism between $`S`$ and the associated graded ring $`G_{}U`$ is induced by the symmetrization map $`e:S\stackrel{~}{}U`$ defined as $$e(g_1\mathrm{}g_p)=\frac{1}{p!}\underset{\sigma S_p}{}g_{\sigma (1)}\mathrm{}g_{\sigma (p)}$$ $`1`$ Thus the associative product $$B:SS\mathrm{@}>>>S,B(x,y)=e^1(exey)$$ $`2`$ maps $`S^n=_{p=0}^nS^p`$ into itself ($`n0`$), whence it can be written as $$B=\underset{p=0}{\overset{\mathrm{}}{}}B_p$$ $`3`$ where $`B_p`$ is homogeneous of degree $`p`$. We have $$B_0(x,y)=xy,B_1(x,y)=\frac{1}{2}\{x,y\}$$ $`4`$ Here $`\{,\}`$ is the Poisson bracket induced by the Lie bracket of $`𝔤`$. Thus if $`x_1,\mathrm{},x_n`$, $`y_1,\mathrm{},y_m𝔤`$, then $$B_1(x,y)=\frac{1}{2}\underset{i,j}{}x_1\mathrm{}\stackrel{𝑖}{}\mathrm{}x_ny_1\mathrm{}\stackrel{𝑗}{}\mathrm{}y_m[x_i,y_j]$$ In general we will have $$B_p(x_1\mathrm{}x_n,y_1\mathrm{}y_m)=\underset{i}{}z_{i,1}\mathrm{}z_{i,m+np}$$ $`5`$ for some elements $`z_{ij}𝔤`$. In this paper we prove a closed formula of the form 5, and an explicit description of the $`z_{i,j}`$ as Lie monomials in the $`x_r,y_s`$ (Theorem 1.1). As an application of our formula, we show that $`B_p`$ is a bidifferential operator of order $`p`$ (Theorem 2.2). The operators $`B_p`$ appear naturally in deformation theory and mathematical physics. One considers the family of products $$x_ty=\underset{p=0}{\overset{\mathrm{}}{}}B_p(x,y)t^p(tk)$$ as a one parameter deformation of the usual commutative product $`_0`$ of $`S`$ into the noncommutative product $`_1`$ of the enveloping algebra. If furthermore $`𝔤`$ is finite dimensional, then $`S`$ can be regarded as the ring of algebraic functions on the dual $`𝔤^{}`$, and the $`S_t=(S,_t)`$ as the rings of functions of a family of noncommutative varieties deforming or “quantizing” the Poisson variety $`𝔤^{}`$. We call this the PBW quantization because the $`B_p`$ are defined by means of the Poincaré-Birkhoff-Witt theorem. The idea of the proof of Theorem 1.1 is to use a well-known expression of the $`B_p`$ in terms of the Campbell-Hausdorff series () in combination with Dynkin’s explicit formula for the latter (, LA 4.17). Although particular cases of our formula were known (,) this paper is to our knowledge the first where it appears in its full generality. An analytic proof of the bidifferentiality of the $`B_p`$ was given in , but no estimate of its order is made there. Our proof of the bidifferentiality is combinatoric and is derived from the explicit formula of Theorem 1.1. The rest of this paper is organized as follows. The formula for $`B_p`$ is established in section 1 (theorem 1.1). Section 2 is devoted to the proof of its bidifferentiality (theorem 2.2). ## 1. A formula for $`B_p`$ In preparation for theorem 1.1 below, we introduce some notation. Let $`n,m1`$, $`X=\{x_1,\mathrm{},x_n\}`$, $`Y=\{y_1,\mathrm{},y_m\}`$ two sets of noncommuting indeterminates. If $`\alpha \{1,\mathrm{},n\}^p`$ ($`p1`$) is a multi-index, then we write: $$|\alpha |=p,\mathrm{Im}\alpha =\{\alpha (1),\mathrm{},\alpha (p)\},ad(x)^\alpha =ad(x_{\alpha (1)})\mathrm{}ad(x_{\alpha (p)})$$ $`6`$ The formula of theorem 1.1 below involves the following element of the free Lie algebra on the disjoint union $`XY`$: $$w(X,Y)=\frac{1}{n+m}(w^{}(X,Y)+w^{\prime \prime }(X,Y))$$ $`7`$ Here $`w^{}`$ and $`w^{\prime \prime }`$ are given by the sums 8 and 9; the restrictions in the summation indexes are explained below; see 11, 12. $`w^{}(X,Y)={\displaystyle \frac{(1)^{p+1}}{p}\frac{ad(x)^{\alpha _1}ad(y)^{\beta _1}\mathrm{}ad(x)^{\alpha _p}(y_k)}{|\alpha _1|!|\beta _1|!\mathrm{}|\alpha _p|!}}`$ $`8`$$`9`$ $`w^{\prime \prime }(X,Y)={\displaystyle \frac{(1)^{p+1}}{p}\frac{ad(x)^{\alpha _1}ad(y)^{\beta _1}\mathrm{}ad(y)^{\beta _{p1}}(x_k)}{|\alpha _1|!|\beta _1|!\mathrm{}|\beta _{p1}|!}}`$ In 8 the sum is taken over arbitrary $`p1`$ and all injective multi-indices $$\alpha _i\{1,\mathrm{},n\}^{|\alpha _i|},\beta _j\{1,\mathrm{},m\}^{|\beta _j|}$$ $`10`$ satisfying $$\begin{array}{c}|\alpha _1|+\mathrm{}+|\alpha _p|=n|\beta _1|+\mathrm{}+|\beta _{p1}|=m1\\ |\alpha _i|+|\beta _i|1(1ip1)|\alpha _p|1\\ \underset{i=1}{\overset{p}{}}\mathrm{Im}\alpha _i=\{1,\mathrm{},n\}\{k\}\underset{i=1}{\overset{p1}{}}\mathrm{Im}\beta _j=\{1,\mathrm{},m\}\end{array}$$ $`11`$ The sum in 9 is also taken over arbitrary $`p1`$, and all injective multi-indices 10, but now $$\begin{array}{c}|\alpha _1|+\mathrm{}+|\alpha _{p1}|=n1|\beta _1|+\mathrm{}+|\beta _{p1}|=m\\ |\alpha _i|+|\beta _i|1(1ip1)\\ \{k\}\underset{i=1}{\overset{p1}{}}\mathrm{Im}\alpha _i=\{1,\mathrm{},n\}\underset{i=1}{\overset{p1}{}}\mathrm{Im}\beta _j=\{1,\mathrm{},m\}\end{array}$$ $`12`$ In the theorem below the we consider the element $`w(A,B)`$ for $`AX`$, $`BY`$. If none of $`A`$, $`B`$ is empty, then $`w(A,B)`$ is already defined by 7; we further define $$w(\{a\},\mathrm{})=w(\mathrm{},\{a\})=a(aXY)$$ $`13`$ ###### Definition Definition 1.0 Let $`A_1`$, $`A_2`$ be sets, and $`𝒫(A_i)`$ the set of all subsets of $`A_i`$ ($`i=1,2`$). A bipartition of $`(A_1,A_2)`$ is a subset $`\pi 𝒫(A_1)\times 𝒫(A_2)`$ such that the following two conditions are satisfied: i) If $`S,T\pi `$ are distinct, then $`S_iT_i=\mathrm{}`$ ($`i=1,2`$). ii) $`A_i=_{S\pi }S_i`$ ($`i=1,2`$) A bipartition $`\pi `$ is called special if $`w(S,T)`$ is defined for all $`(S,T)\pi `$; that is if the following holds $$(\mathrm{},S)\text{ or }(S,\mathrm{})\pi \mathrm{\#}S=1$$ $`14`$ ###### Theorem 1.1 Let $`n,m,p1`$, $`X=\{x_1,\mathrm{},x_n\}`$, $`Y=\{y_1,\mathrm{},y_m\}`$ two sets of indeterminates, and $`B_p`$ the operator of 3 for the free Lie algebra on the disjoint union $`XY`$. Then $$B_p(x_1\mathrm{}x_n,y_1\mathrm{}y_m)=\underset{\pi }{}w(\pi _1^X,\pi _1^Y)\mathrm{}w(\pi _{n+mp}^X,\pi _{n+mp}^Y)$$ $`15`$ Here $`w`$ is as defined in 7, and the sum runs over all special bipartitions $`\pi =\{(\pi _1^X,\pi _1^Y),\mathrm{},(\pi _{n+mp}^X,\pi _{n+mp}^Y)\}`$ of cardinality $`m+np`$ of $`(X,Y)`$. In particular, $$B_{n+m1}(x_1\mathrm{}x_n,y_1\mathrm{}y_m)=w(X,Y)$$ $`16`$ ###### Demonstration Proof Write $`𝔤`$ for the free Lie algebra on $`XY`$. Let $`t_1,\mathrm{},t_n`$, $`u_1,\mathrm{},u_m`$ be commuting algebraically independent variables. Put $$x(t)=\underset{i=1}{\overset{n}{}}t_ix_i,y(t)=\underset{i=1}{\overset{m}{}}u_iy_i$$ One checks that $`B(x_1\mathrm{}x_n,y_1\mathrm{}y_m)`$ is the coefficient of $`t_1\mathrm{}t_nu_1\mathrm{}u_m`$ in the product of the exponential series $$exp(x(t))exp(y(u))=exp(z(t,u))$$ $`17`$ Here $`z(x(t),y(u))`$ is the Campbell-Hausdorff series. Consider the expansion of $`z`$ as a series in $`t,u`$. In order to compute the coefficient of $`t_1\mathrm{}t_nu_1\mathrm{}u_m`$ in 17, all terms in the expansion of $`z`$ corresponding to monomials in which any of the $`t_i,u_j`$ has exponent $`2`$ may be discarded. Each of the remaining terms is an element of $`𝔤`$ times a monomial of the form: $$t_Au_B:=t_{a_1}\mathrm{}t_{a_r}u_{b_1}\mathrm{}u_{b_s}$$ for some subsets $`A=\{a_1,\mathrm{},a_r\}\{1,\mathrm{},n\}`$, $`B=\{b_1,\mathrm{},b_s\}\{1,\mathrm{},m\}`$. One checks, using Dynkin’s formula (,LA 4.17), that the coefficient of $`t_Au_B`$ in $`z`$ is precisely the element $`w(A,B)`$. The theorem follows from this and the definition of the symmetrization map 1.∎ In the course of the proof of the theorem above we introduced a notation which shall be used often in what follows. If $`𝒜`$ is a $`k`$-algebra, $`a_1,\mathrm{},a_n𝒜`$ and $`S=\{i_1,\mathrm{},i_r\}\{1,\mathrm{},n\}`$ is a subset of $`r`$ elements, then we write: $$a_S:=a_{i_1}\mathrm{}a_{i_r}$$ $`18`$ In particular $$a_{\mathrm{}}=1$$ ## 2. The bidifferentiality of $`B_p`$ Let $`q,k0`$; define inductively $$c_0(q)=1,c_k(q)=1\underset{l=0}{\overset{k1}{}}c_l(q)\left(\genfrac{}{}{0pt}{}{q+k}{kl}\right)(k1)$$ $`19`$ ###### Lemma 2.0 Let $`𝔤`$ be a Lie algebra, $`p1`$, $`q0`$, $`r1`$, $`x_1,\mathrm{},x_{p+q}`$, $`y_1,\mathrm{},y_r𝔤`$ and $`c_k(q)`$ as in 19 above. Then, with the notation of 18, we have $$B_p(x_1\mathrm{}x_{p+q},y_1\mathrm{}y_r)=\underset{k=0}{\overset{r1}{}}c_k(q)(\underset{\mathrm{\#}S=q+k}{}x_SB_p(x_{S^c},y_1\mathrm{}y_r))$$ $`20`$ Here $`S\{1,\mathrm{},p+q\}`$, and $`S^c`$ is the complement of $`S`$. The symmetric formula holds for $`B_p(y_1\mathrm{}y_r,x_1\mathrm{}x_{p+q})`$. ###### Demonstration Proof We may assume that the $`x_i,y_j`$ are indeterminates and that $`𝔤`$ is the free Lie algebra. Apply theorem 1.1 to write the left hand side of 20 as a sum of terms indexed by all special bipartitions $`\pi `$ of $`(\{1,\mathrm{},p+q\},\{1,\mathrm{},r\})`$ of $`q+r`$ elements. It follows from the definition of a special bipartition that the number $`d_\pi `$ of empty sets in the list $`\pi _1^Y,\mathrm{},\pi _{q+r}^Y`$ of subsets of $`\{1,\mathrm{},r\}`$ is at least $`q`$ and at most $`q+r1`$. Write $`b_k`$ for the sum of those terms whose indexing bipartition has $`d_\pi =q+k`$ ($`0kr1`$). By definition, $$B_p(x_1\mathrm{}x_{p+q},y_1\mathrm{}y_r)=\underset{l=0}{\overset{r1}{}}b_l$$ Moreover, a counting argument shows that for $`0kr1`$ $$\underset{\mathrm{\#}S=q+k}{}x_SB_p(x_{S^c},y_1\mathrm{}y_r)=\underset{i=0}{\overset{r1k}{}}\left(\genfrac{}{}{0pt}{}{q+k+i}{i}\right)b_{k+i}$$ Hence $$\begin{array}{c}\underset{k=0}{\overset{r1}{}}c_k(q)(\underset{\mathrm{\#}S=q+k}{}x_SB_p(x_{S^c},y_1\mathrm{}y_r))=\underset{k=0}{\overset{r1}{}}\underset{i=0}{\overset{r1k}{}}c_k(q)\left(\genfrac{}{}{0pt}{}{q+k+i}{i}\right)b_{k+i}\\ =\underset{l=0}{\overset{r1}{}}(\underset{k=0}{\overset{l}{}}c_k(q)\left(\genfrac{}{}{0pt}{}{q+l}{lk}\right))b_l=\underset{l=0}{\overset{r1}{}}b_l\text{ (by 19)}\\ =B_p(x_1\mathrm{}x_{p+q},y_1\mathrm{}y_r)\mathit{}\end{array}$$ ###### Lemma 2.1 Let $`q1`$, $`m0`$. Then: $$0=(1)^m+\underset{t=1}{\overset{q}{}}\left(\genfrac{}{}{0pt}{}{m+q}{m+t}\right)(1)^tc_m(t)$$ ###### Demonstration Proof Fix $`q1`$; the proof is by induction on $`m0`$. The case $`m=0`$ is immediate. Assume $`m1`$ and by induction that the lemma holds for $`s<m`$. Then $$\begin{array}{c}\underset{t=1}{\overset{q}{}}\left(\genfrac{}{}{0pt}{}{m+q}{m+t}\right)(1)^tc_m(t)=\\ =\underset{t=1}{\overset{q}{}}\left(\genfrac{}{}{0pt}{}{m+q}{m+t}\right)(1)^t\underset{t=1}{\overset{q}{}}\underset{s=0}{\overset{m1}{}}(1)^t\left(\genfrac{}{}{0pt}{}{m+q}{m+t}\right)\left(\genfrac{}{}{0pt}{}{t+m}{ms}\right)c_s(t)\text{ (by 19)}\\ =(1)^m\underset{t=m+1}{\overset{m+q}{}}\left(\genfrac{}{}{0pt}{}{m+q}{t}\right)(1)^t\underset{s=0}{\overset{m1}{}}\underset{t=1}{\overset{q}{}}(1)^t\left(\genfrac{}{}{0pt}{}{m+q}{ms}\right)\left(\genfrac{}{}{0pt}{}{s+q}{s+t}\right)c_s(t)\\ =(1)^{m+1}(\underset{t=0}{\overset{m}{}}(1)^t\left(\genfrac{}{}{0pt}{}{m+q}{t}\right))+\underset{s=0}{\overset{m1}{}}(1)^s\left(\genfrac{}{}{0pt}{}{m+q}{ms}\right)\text{ (by inductive assumption)}\\ =(1)^{m+1}+(1)^{m+1}(\underset{t=1}{\overset{m}{}}(1)^t\left(\genfrac{}{}{0pt}{}{m+q}{t}\right))+\underset{t=1}{\overset{m}{}}(1)^{m+t}\left(\genfrac{}{}{0pt}{}{m+q}{t}\right)=(1)^{m+1}\mathit{}\end{array}$$ Recall that a $`k`$-linear endomorphism $`F`$ of a commutative associative algebra $`𝒜`$ is a differential operator of order $`p`$ if $$\underset{S\{1,\mathrm{},p\}}{}(1)^Sa_{S^c}F(a_Sb)=0(a_1,\mathrm{},a_p,b𝒜)$$ $`21`$ One checks that if $`𝒜`$ is generated as a $`k`$-algebra by a set $`X𝒜`$ then $`F`$ satisfies 21 if and only if it satisfies $$\underset{S\{1,\mathrm{},p+q\}}{}(1)^Sx_{S^c}F(x_S)=0(x_1,\mathrm{},x_{p+q}X,q0)$$ $`22`$ ###### Theorem 2.2 Let $`𝔤`$ be a Lie algebra, $`S=S𝔤`$ the symmetric algebra, $`aS`$, $`p1`$. Also let $`F:S\mathrm{@}>>>S`$ be one of $`B_p(a,)`$ or $`B_p(,a)`$. Then $`F`$ is a differential operator of order $`p`$. ###### Demonstration Proof It suffices to show the theorem for $`a`$ homogeneous. Assume $`a=y_1\mathrm{}y_r`$, $`y_i𝔤`$. We shall show that the identity 22 holds for $`X=𝔤`$. Put $`K=\{1,\mathrm{},p+q\}`$. The left hand side of 22 is the sum of $$\underset{l=pr+1}{\overset{p}{}}\underset{\mathrm{\#}S=l}{}(1)^lx_{K\backslash S}F(x_S)$$ $`23`$ and of $$\begin{array}{c}\underset{t=1}{\overset{q}{}}\underset{\mathrm{\#}S=p+t}{}(1)^{p+t}x_{K\backslash S}F(x_S)=\\ \underset{t=1}{\overset{q}{}}\underset{\mathrm{\#}S=p+t}{}\underset{k=0}{\overset{r1}{}}\underset{\begin{array}{c}TS\\ \mathrm{\#}T=pk\end{array}}{}(1)^{p+t}c_k(t)x_{K\backslash S}x_{S\backslash T}F(x_T)\text{(by Lemma 2.0)}\\ =\underset{l=pr+1}{\overset{p}{}}\underset{\mathrm{\#}S=l}{}(\underset{t=1}{\overset{q}{}}\left(\genfrac{}{}{0pt}{}{p+ql}{p+tl}\right)(1)^{p+t}c_{pl}(t))x_{K\backslash S}F(x_S)\end{array}$$ By lemma 2.1, the sum of the last expression with that of 23 is zero. ∎ ### Acknowledgements I wish to thank Boris Tsygan, from whom I learned of the relation between the Campbell-Hausdorff series and the $`B_p`$. Victor Ginzburg told me that no general formula as I was looking for seemed to be known. His comment encouraged me to carry out this research, and I am thankful to him for it. Thanks also to Alexei Davydov for a useful discussion.
warning/0001/cond-mat0001182.html
ar5iv
text
# Unconventional strong pinning in the low temperature phase of U.9725Th.0275Be13 ## Abstract We investigated low field vortex dynamics in a single crystal of U<sub>.9725</sub>Th<sub>.0275</sub>Be<sub>13</sub>.We found a sharp transition in the vortex creep rate at the lower transition temperature $`T_{c2}`$, coincident with the second jump in the specific heat. In the high-temperature phase, rather strong creep rates are observed. In the low temperature phase, the rates drop to undetectabely low levels. This behaviour indicates that a very strong pinning mechanism is present in the low temperature phase of U<sub>.9725</sub>Th<sub>.0275</sub>Be<sub>13</sub>, which could be explained by the existence of domain walls, separating discreetly degenerate states of a superconductor, that can sustain fractional vortices and thus act as very strong pinning centers. , , and thanks: Corresponding author. Present address: Laboratorium für Festkörperpkysik, ETH Hönggerberg, CH 8093 Zürich, Switzerland. E-mail: dumont@solid.phys.ethz.ch If a small amount of uranium is substituted for thorium in the heavy fermion superconductor UBe<sub>13</sub>, the superconducting transition temperature shows a non-monotonic dependence on thorium concentration $`x`$. Moreover, in a small critical range $`0.019<x<0.045`$ a second transition occurs below the superconducting transition. The nature of this phase transition is still a matter of controversy. In zero field $`\mu `$SR measurements, a continuous increase of $`\mu `$-spin-relaxation rate, which sets in at $`T_{c2}`$ has been observed. This has been interpreted as originating from a very weak spontaneous magnetic field. Superconducting origin of the transition at $`T_{c2}`$ has been deduced from the observation of a distinct increase in the slope of the lower critical field $`H_{c1}`$ at $`T_{c2}`$. Recently, the low-field magnetic properties of the heavy fermion superconductor UPt<sub>3</sub> have been investigated . In the low–field, low–temperature B–phase, bulk vortices remain so strongly pinned that the creep rates drop from rather high finite rates observed in the A–phase to undetectabely low levels at $`T_c^{}`$. The observation of zero initial creep in the B–phase of UPt<sub>3</sub> was then interpreted as originating from an intrinsic pinning mechanism in multicomponent superconducting phases. In such phases fractional vortices get trapped in domain walls separating discreetly degenerate states of the superconductor. These experimental results show that vortex creep measurements can be used as a powerful probe to give information about the character of a superconducting phase. We present here relaxation measurements of the remanent magnetization on a single crystal of U<sub>.9725</sub>Th<sub>.0275</sub>Be<sub>13</sub>(size: $`1.00\times 0.89\times 2.25`$ mm<sup>3</sup>) prepared in Los Alamos National Laboratory. The experimental arrangement has been described earlier . The single crystal has a transition temperature $`T_{c1}=523`$ mK. Specific heat measurements show a second jump at $`T_{c2}=350`$ mK. Typical relaxation measurements of the remanent magnetization are shown in fig. 1. All the values of $`M_{rem}`$ are taken with the sample in the fully critical state. We observe two different regimes of vortex creep, separated by a dramatic change in creep rates at $`T=T_{c2}`$. For temperatures above $`T_{c2}`$, we observe strong vortex creep and a logarithmic time dependence of the decays. For temperatures below $`T_{c2}`$, the initial vortex creep drops to zero. At longer times the decays start deviating from logarithmic behaviour. This acceleration of vortex creep at long times has also been observed in UPt<sub>3</sub>. Here we only discuss the initial slope of the decays. The normalized creep rates ($`\mathrm{ln}M/\mathrm{ln}t`$) are plotted in Fig.2 as a function of temperature. At $`T=T_{c2}`$, we observe a sharp transition in creep rates. They drop by 3 orders of magnitude to zero within our sensitivity ($`\mathrm{ln}M/\mathrm{ln}t10^5`$). Also shown in Fig.2 are measurements of the lower critical field $`H_{c1}`$ on the same crystal. $`H_{c1}`$ was obtained as the field where the first deviation from the initial slope of isothermal magnetization–curves occurs. In conclusion, we have observed a sharp transition in vortex pinning in U<sub>.9725</sub>Th<sub>.0275</sub>Be<sub>13</sub>at $`T_{c2}`$, similar to the one observed in UPt<sub>3</sub> at $`T_c^{}`$. This might hint towards a new type of strong pinning structures and/or a new type of vortices being present in the low–T phases of these two multiphase superconductors.
warning/0001/hep-ph0001164.html
ar5iv
text
# Hard Loop Approach to Anisotropic Systems ## I Introduction The state of equilibrium being static and homogeneous is sometimes anisotropic. This may happen when the system of quantum fields, which is of interest here, is under influence of an external force. A relativistic plasma, which is anisotropic due to a magnetic field, often occurs in astrophysical situations as the early Universe or Supernovae . Anisotropic states are also common for systems which are out of equilibrium. Sometimes such states can be treated as static and homogeneous, but only for sufficiently short time and space intervals. How short the intervals should be depends on the specific problem under consideration. The parton system generated at the early stage of ultrarelativistic heavy ion collisions at RHIC or LHC is of particular interest for us. The parton momentum distribution is not istotropic but strongly elongated along the beam . Therefore, specific color fluctuations, instead of being damped, can exponentially grow and noticeably influence the temporal evolution of the system. In a series of papers of one of us it has been argued that there are indeed very fast unstable plasma modes in such a parton system. The stability analysis has been performed within the semiclassical transport theory of quarks and gluons . Since the theory has been proven till now to be fully consistent with the QCD dynamics only for quasiequilibrium systems one wonders to what extend the results from are reliable. Thus, a QCD diagrammatic analysis is desirable. Perturbative approaches within the real time field theory provides a natural framework to study weakly interacting quantum field systems in and out of equilibrium. However, the naive perturbative expansion, when applied to gauge fields, suffers from various singularities and some physical quantities are even gauge dependent. These problems have been partly resolved for equilibrium systems by using an effective perturbative expansion where the so-called Hard Thermal Loops are resummed . The Hard Thermal Loop resummation technique within the finite-temperature QCD has been shown to be equivalent to the approach based on the classical transport equations, where color is treated as a classical variable, or on the semiclassical one, where the color degrees of freedom emerge due to the matrix structure of the parton distribution function. The Hard Thermal Loop approach has been generalized to nonequilibrium systems, but only very specific forms of deviations from the equilibrium have been discussed so far: systems out of chemical equilibrium, which are important in the context of heavy-ion collisions , and such where the momentum distribution is isotropic but not of the Bose-Einstein or Fermi-Dirac form . As observed in , the Hard Thermal Loop approach can be applied to any momentum distribution of hard particles which is static and homogeneous. This is evident when the Hard Thermal Loop effective action is derived within the transport theory . The term ‘thermal’ is then rather misleading and for this reason we shall omit it in the following. In this paper we discuss the applicability of the Hard Loop technique for systems with anisotropic momentum distributions. The technique has been earlier applied to the equilibrium QED plasma in a magnetic field . Our aim is to consider a general situation with an arbitrary momentum distribution. We analyse the problem from the point of view of the transport theory and the diagrammatic approach. Using the semiclassical kinetic equations we derive the Hard Loop induced current paying much attention to the gauge aspects of the procedure. We also explicitly demonstrate that the gluon polarization tensors found by means of the two approaches are identical. In this way, the applicability of the kinetic theory beyond the equilibrium is substantiated and more specifically, the reliability of the results from is shown. The Hard Loop diagrammatic technique has the advantage over the semiclassical transport theory approach that it can be naturally extended to fermionic self energies and to higher order diagrams beyond the semiclassical approximation. In this way the dispersion relations of quarks and other observables of the quark-gluon plasma, such as the energy loss of energetic partons, transport coefficients, or photon and dilepton production rates , can be calculated systematically in the case of anisotropic distributions. We take a first step in this direction computing the quark self energy for an arbitrary momentum distribution. The self energy controls the particle dispersion relation which provides an essential dynamical information about the system. We discuss therefore the general dispersion relation of gluons (plasmons) and quarks in the anisotropic quark-gluon plasma. Finally, we briefly consider possible applications of the formalism developed in this paper. ## II Transport Theory Approach In this section we first introduce the semiclassical transport theory of quarks and gluons . Then, applying the linear response method, the Hard Loop induced current is derived. Finally, we compute the gluon polarization. ### A Transport equations The distribution function of hard (anti-)quarks $`Q(𝐩,x)(\overline{Q}(𝐩,x))`$ is a hermitian $`N_c\times N_c`$ matrix in color space (for a $`SU(N_c)`$ color group); $`x`$ denotes the space-time quark coordinate and $`𝐩`$ its momentum. The four-momentum $`p=(E,𝐩)`$ is assumed to satisfy the mass-shell constraint. Since both quarks and gluons are treated as massless particles the constraint is $`p^2=0`$. We also mention here that the spin of quarks and gluons is taken into account as an internal degree of freedom. The distribution function transforms under local gauge transformation $`M`$ as $$Q(𝐩,x)M(x)Q(𝐩,x)M^{}(x).$$ (1) The color indices are here and in the most cases below suppressed. The distribution function of hard gluons is a hermitian $`(N_c^21)\times (N_c^21)`$ matrix which transforms as $$G(𝐩,x)(x)G(𝐩,x)^{}(x),$$ (2) where $$_{ab}(x)=\mathrm{Tr}[\tau _aM(x)\tau _bM^{}(x)]$$ with $`\tau _a,a=1,\mathrm{},N_c^21`$ being the $`SU(N_c)`$ group generators in the fundamental representation. The color current is expressed in the fundamental representation as $`j^\mu (x)=g{\displaystyle }{\displaystyle \frac{d^3p}{(2\pi )^32E}}p^\mu [Q(𝐩,x)\overline{Q}(𝐩,x){\displaystyle \frac{1}{N_c}}\mathrm{Tr}[Q(𝐩,x)`$ $``$ $`\overline{Q}(𝐩,x)]`$ (3) $`+`$ $`2i\tau _af_{abc}G_{bc}(𝐩,x)],`$ (4) where $`g`$ is the QCD coupling constant, $`f_{abc}`$ are the structure constants of the $`SU(N_c)`$ group. The distribution functions of quarks and gluons are assumed to satisfy the following collisionless transport equations: $`p^\mu D_\mu Q(𝐩,x)+gp^\mu {\displaystyle \frac{}{p_\nu }}{\displaystyle \frac{1}{2}}\{F_{\mu \nu }(x),Q(𝐩,x)\}`$ $`=`$ $`0,`$ (5) $`p^\mu D_\mu \overline{Q}(𝐩,x)gp^\mu {\displaystyle \frac{}{p_\nu }}{\displaystyle \frac{1}{2}}\{F_{\mu \nu }(x),\overline{Q}(𝐩,x)\}`$ $`=`$ $`0,`$ (6) $`p^\mu 𝒟_\mu G(𝐩,x)+gp^\mu {\displaystyle \frac{}{p_\nu }}{\displaystyle \frac{1}{2}}\{_{\mu \nu }(x),G(𝐩,x)\}`$ $`=`$ $`0,`$ (7) where $`\{\mathrm{},\mathrm{}\}`$ denotes the anicommutator; $`D_\mu `$ and $`𝒟_\mu `$ are the covariant derivatives which act as $$D_\mu =_\mu ig[A_\mu (x),\mathrm{}],𝒟_\mu =_\mu ig[𝒜_\mu (x),\mathrm{}],$$ with $`A_\mu `$ and $`𝒜_\mu `$ being the mean-field or background four-potentials; $$A^\mu (x)=A_a^\mu (x)\tau _a,𝒜_{ab}^\mu (x)=if_{abc}A_c^\mu (x);$$ $`F_{\mu \nu }`$ and $`_{\mu \nu }`$ are the mean-field stress tensors with a color index structure analogous to that of the four-potentials. The background field is generated by the color current (3) and the respective equation is $$D_\mu F^{\mu \nu }(x)=j^\nu (x).$$ (8) We note that the set of transport equations (5, 8) is covariant with respect to the gauge transformations (1, 2). ### B Plasma color response We discuss here how the plasma, which is (on average) colorless, homogeneous and stationary, responds to small color fluctuations. The distribution functions are assumed to be of the form $`Q_{ij}(𝐩,x)`$ $`=`$ $`n(𝐩)\delta _{ij}+\delta Q_{ij}(𝐩,x),`$ (9) $`\overline{Q}_{ij}(𝐩,x)`$ $`=`$ $`\overline{n}(𝐩)\delta _{ij}+\delta \overline{Q}_{ij}(𝐩,x),`$ (10) $`G_{ab}(𝐩,x)`$ $`=`$ $`n_g(𝐩)\delta _{ab}+\delta G_{ab}(𝐩,x),`$ (11) where the functions describing the deviation from the colorless state are assumed to be much smaller than the respective colorless functions. The same is assumed for the momentum gradients of these functions. The (anti-)quark and gluon distribution functions $`n(𝐩)`$, $`\overline{n}(𝐩)`$, $`n_g(𝐩)`$, reduce in equilibrium to the Fermi-Dirac or Bose-Einstein form i.e. $`n(𝐩)`$ $`=`$ $`{\displaystyle \frac{2}{\mathrm{exp}(|𝐩|\mu )/T+1}},`$ (12) $`\overline{n}(𝐩)`$ $`=`$ $`{\displaystyle \frac{2}{\mathrm{exp}(|𝐩|+\mu )/T+1}},`$ (13) $`n_g(𝐩)`$ $`=`$ $`{\displaystyle \frac{2}{\mathrm{exp}(|𝐩|/T)1}},`$ (14) where $`T`$ and $`\mu `$ denote the temperature and quark chemical potential, respectively, while the factor of 2 occurs due to the spin degrees of freedom. The number of quark flavours is assumed to be equal to one. Substituting (9) in (3) one gets $`j^\mu (x)=g{\displaystyle }{\displaystyle \frac{d^3p}{(2\pi )^32E}}p^\mu [\delta Q(𝐩,x)\delta \overline{Q}(𝐩,x){\displaystyle \frac{1}{N_c}}Tr[\delta Q(𝐩,x)`$ $``$ $`\delta \overline{Q}(𝐩,x)]`$ (15) $`+`$ $`2i\tau _af_{abc}\delta G_{bc}(𝐩,x)].`$ (16) As seen, the current occurs due to the deviation from the colorless state. Let us also observe here that not only (anti-)quarks but also gluons contribute to the current (15). Thus, the current is of essentially non-Abelian nature. Now, we substitute the distribution functions (9) into the transport equations (5). Assuming that the stress tensor is of the same order as $`\delta Q`$, $`\delta \overline{Q}`$ or $`\delta G`$ and linearizing the equations with respect to $`\delta Q`$, $`\delta \overline{Q}`$ and $`\delta G`$ we get $`p^\mu D_\mu \delta Q(𝐩,x)`$ $`=`$ $`gp^\mu F_{\mu \nu }(x){\displaystyle \frac{n(𝐩)}{p_\nu }},`$ (17) $`p^\mu D_\mu \delta \overline{Q}(𝐩,x)`$ $`=`$ $`gp^\mu F_{\mu \nu }(x){\displaystyle \frac{\overline{n}(𝐩)}{p_\nu }},`$ (18) $`p^\mu 𝒟_\mu \delta G(𝐩,x)`$ $`=`$ $`gp^\mu _{\mu \nu }(x){\displaystyle \frac{n_g(𝐩)}{p_\nu }}.`$ (19) We keep here the covariant derivatives to maintain the gauge covariance of the equations. To solve the equations such as Eqs. (17) one usually uses, see e.g. , the gauge parallel transporter defined in the fundamental representation as $$U(x,y)=𝒫\mathrm{exp}\left[ig_x^y𝑑z_\mu A^\mu (z)\right],$$ where $`𝒫`$ denotes the ordering along the path from $`x`$ to $`y`$. There is analogous formula of the gauge transporter $`𝒰(x,y)`$ in the adjoint representation. Using $`U`$ and $`𝒰`$ one finds the solutions of Eqs. (17) as $`\delta Q(𝐩,x)`$ $`=`$ $`g{\displaystyle d^4yG_p(xy)U(x,y)p^\mu F_{\mu \nu }(y)U(y,x)\frac{n(𝐩)}{p_\nu }},`$ (20) $`\delta \overline{Q}(𝐩,x)`$ $`=`$ $`g{\displaystyle d^4yG_p(xy)U(x,y)p^\mu F_{\mu \nu }(y)U(y,x)\frac{\overline{n}(𝐩)}{p_\nu }},`$ (21) $`\delta G(𝐩,x)`$ $`=`$ $`g{\displaystyle d^4yG_p(xy)𝒰(x,y)p^\mu _{\mu \nu }(y)𝒰(y,x)\frac{n_g(𝐩)}{p_\nu }},`$ (22) where $`G_p(x)`$ is the retarded Green’s function which satisfies the equation $$p_\mu ^\mu G_p(x)=\delta ^{(4)}(x)$$ and equals $$G_p(x)=E^1\mathrm{\Theta }(t)\delta ^{(3)}(𝐱𝐯t),$$ with $`t`$ being the 0-th component of $`x`$ ($`x^\mu (t,𝐱)`$), and $`𝐯`$ denoting the parton velocity i.e. $`𝐯𝐩/E`$. Substituting the solutions (20) in Eq. (15) one finds the color current of the gauge covariant form which reads $`j^\mu (x)=g^2{\displaystyle \frac{d^3p}{(2\pi )^32E}p^\mu p^\lambda d^4yG_p(xy)U(x,y)F_{\lambda \nu }(y)U(y,x)\frac{f(𝐩)}{p_\nu }}`$ (23) where $`f(𝐩)n(𝐩)+\overline{n}(𝐩)+2N_cn_g(𝐩)`$. Now, we are going to perform the Fourier transform of the induced current (23). Before this step however, we neglect the terms which are not of the leading order in $`g`$. Then, the transporters $`U`$ are approximated by unity and the stress tensor $`F_{\mu \nu }`$ by $`_\mu A_\nu _\nu A_\mu `$. Within such an approximation, the Fourier transformed induced current (23), which is no longer gauge covariant, equals $$j^\mu (k)=g^2\frac{d^3p}{(2\pi )^32E}p^\mu \frac{f(𝐩)}{p_\lambda }\left[g^{\lambda \nu }\frac{k^\lambda p^\nu }{p^\sigma k_\sigma +i0^+}\right]A_\nu (k).$$ (24) The induced current $`j^\mu (k)`$ can be expressed as $$j_a^\mu (k)=\mathrm{\Pi }_{ab}^{\mu \nu }(k)A_\nu ^b(k),$$ with $`\mathrm{\Pi }^{\mu \nu }`$ being the gluon polarization tensor. Transforming Eq. (24) to the adjoint representation one finds $$\mathrm{\Pi }^{\mu \nu }(k)=g^2\frac{d^3p}{(2\pi )^32E}p^\mu \frac{f(𝐩)}{p_\lambda }\left[g^{\lambda \nu }\frac{k^\lambda p^\nu }{p^\sigma k_\sigma +i0^+}\right].$$ (25) It should be noted here that the polarization tensor is proportional to a unit matrix in the color space. Now we are going to show that the polarization tensor is transversal i.e. $`k_\mu \mathrm{\Pi }^{\mu \nu }(k)=0`$. Let us first consider $`k_\mu \mathrm{\Pi }^{\mu 0}(k)`$. One immediately finds from Eq. (25) that $$k_\mu \mathrm{\Pi }^{\mu 0}(k)=\frac{g^2}{2}k^l\frac{d^3p}{(2\pi )^3}\frac{f(𝐩)}{p^l}.$$ The indices $`l,m,n=1,2,3`$ refer to the coordinates of three-vectors. The energy density carried by partons is expected to be finite. Therefore, $`f(𝐩=\mathrm{})`$ must vanish. Consequently, the above integral vanishes as well. Performing partial integration and demanding that $`f(𝐩=\mathrm{})=0`$ one also proves that $`k_\mu \mathrm{\Pi }^{\mu m}(k)=0`$. Analogously it can be also shown that $`\mathrm{\Pi }^{\mu \nu }(k)=\mathrm{\Pi }^{\nu \mu }(k)`$. ## III Diagrammatic approach In this section we consider the diagrammatic Hard Loop approach to anisotropic systems. Specifically, we compute the QCD polarization tensor and the quark self energy for an arbitrary parton momentum distribution. ### A Polarization tensor The contribution from the quark loop to the gluon self energy in the case of one quark flavor is of the form $$\mathrm{\Pi }_{ab}^{\mu \nu }(k)=\frac{i}{2}\delta _{ab}g^2\frac{d^4p}{(2\pi )^4}\mathrm{Tr}[\gamma ^\mu S(q)\gamma ^\nu S(p)],$$ (26) where $`qpk`$ and $`S`$ is the bare quark propagator. Since we are dealing with a non-equilibrium situation we adopt the real time formalism. Within the Keldysh representation , which has been shown to be especially convenient in the Hard Loop approximation , there are retarded $`(R)`$, advanced $`(A)`$, and symmetric $`(F)`$ propagators which in the case of massless quarks are given by $`S_{R,A}(p)`$ $`=`$ $`{\displaystyle \frac{p/}{p^2\pm i\mathrm{sgn}(p_0)0^+}},`$ (27) $`S_F(p)`$ $`=`$ $`2\pi ip/\left([1n(𝐩)]\mathrm{\Theta }(p_0)+[1\overline{n}(𝐩)]\mathrm{\Theta }(p_0)\right)\delta (p^2),`$ (28) where $`n(𝐩)`$ ($`\overline{n}(𝐩)`$) is, as previously, the (anti-)quark distribution function that reduces in equilibrium to the form (12). Performing the trace in (26) and suppressing the color indices, we find the retarded gluon self energy as $$\mathrm{\Pi }^{\mu \nu }(k)=ig^2\frac{d^4p}{(2\pi )^4}[q^\mu p^\nu +p^\mu q^\nu g^{\mu \nu }(qp)][\stackrel{~}{\mathrm{\Delta }}_F(q)\stackrel{~}{\mathrm{\Delta }}_R(p)+\stackrel{~}{\mathrm{\Delta }}_A(q)\stackrel{~}{\mathrm{\Delta }}_F(p)],$$ (29) where $`S_{R,A,F}(p)=p/\stackrel{~}{\mathrm{\Delta }}_{R,A,F}(p)`$. Terms containing $`\stackrel{~}{\mathrm{\Delta }}_A(q)\stackrel{~}{\mathrm{\Delta }}_A(p)`$ and $`\stackrel{~}{\mathrm{\Delta }}_R(q)\stackrel{~}{\mathrm{\Delta }}_R(p)`$ have been neglected as they vanish after integrating over $`p_0`$. First, we will consider the spatial components of the polarization tensor. The other components follow from it using the transversality of the polarization tensor, as we will discuss below. After performing the integration over $`p_0`$, we obtain $`\mathrm{\Pi }^{lm}(k)={\displaystyle \frac{g^2}{2}}{\displaystyle \frac{d^3p}{(2\pi )^3}\frac{f_q(𝐩)}{|𝐩|}}`$ $`[{\displaystyle \frac{2p^lp^mk^lp^mp^lk^m+\delta ^{lm}(\omega |𝐩|+𝐤𝐩)}{2\omega |𝐩|+2𝐤𝐩+k^2i\mathrm{sgn}(|𝐩|\omega )0^+}}`$ (31) $`+{\displaystyle \frac{2p^lp^mk^lp^mp^lk^m+\delta ^{lm}(\omega |𝐩|+𝐤𝐩)}{2\omega |𝐩|+2𝐤𝐩+k^2i\mathrm{sgn}(|𝐩|\omega )0^+}}],`$ where $`\omega `$ is the 0-th component of $`k`$ i.e. $`k=(\omega ,𝐤)`$ and $`f_q(𝐩)n(𝐩)+\overline{n}(𝐩)`$. Here the vacuum part has been neglected because it is suppressed compared to the matter part in the Hard Loop approximation. Adopting the Hard Loop approximation we assume that the internal momenta are much larger than the external one, i.e. $`\omega `$, $`k_lp_l`$. Note that for arbitrary anisotropic distributions we have to require the Hard Loop condition for each component of the momenta, whereas in the isotropic case $`\omega `$, $`|𝐤||𝐩|`$ suffices. Expanding the expression in the square brackets for small external momenta yields $`{\displaystyle \frac{p^lp^m}{\omega |𝐩|+𝐤𝐩i0^+}}+{\displaystyle \frac{p^lp^m}{\omega |𝐩|+𝐤𝐩+i0^+}}+{\displaystyle \frac{k^lp^mp^lk^m+\delta ^{lm}(\omega |𝐩|+𝐤𝐩)}{2(\omega |𝐩|+𝐤𝐩i0^+)}}`$ (32) $`+{\displaystyle \frac{k^lp^mp^lk^m+\delta ^{lm}(\omega |𝐩|+𝐤𝐩)}{2(\omega |𝐩|+𝐤𝐩+i0^+)}}{\displaystyle \frac{p^lp^mk^2}{2(\omega |𝐩|+𝐤𝐩i0^+)^2}}{\displaystyle \frac{p^lp^mk^2}{2(\omega |𝐩|+𝐤𝐩+i0^+)^2}}.`$ (33) In equilibrium the first two terms vanish after integrating over $`𝐩`$. This also holds out of equilibrium if we assume $`f_q(𝐩)=f_q(𝐩)`$. Then, we arrive at the final result $$\mathrm{\Pi }^{lm}(k)=\frac{g^2}{2}\frac{d^3p}{(2\pi )^3}\frac{f(𝐩)}{|𝐩|}\frac{(k^lp^m+p^lk^m)(\omega |𝐩|𝐤𝐩)+\delta ^{lm}(\omega |𝐩|𝐤𝐩)^2p^lp^m(\omega ^2|𝐤|^2)}{(\omega |𝐩|𝐤𝐩+i0^+)^2},$$ (34) where we replaced $`f_q(𝐩)`$ by $`f(𝐩)n(𝐩)+\overline{n}(𝐩)+2N_cn_g(𝐩)`$. The point is that in the Hard Loop limit the gluonic contributions to the polarization tensor have the same structure as the quark ones and only the distribution function and the color factor change. For essentially the same reason, the QCD polarization tensor, computed even with the complete expression (32) without assuming $`f(𝐩)=f(𝐩)`$, is gauge independent in the Hard Loop approximation. Indeed, the gluon polarization tensor has the same structure as the photon one in the Hard Loop limit. Since the one-loop photon polarization tensor contains no gauge boson propagator it is gauge independent. Consequently, the same holds for the gluon polarization tensor. The result (34) is fully equivalent to Eq. (25) obtained within the semiclassical kinetic theory. In order to show the equivalence, one performs a partial integration in (25) and immediately gets Eq. (34). However, we do not need to assume the reflection symmetry of the distribution function to derive Eq. (25). Two more comments are in order here. First, if we do not assume the reflection symmetry of $`f(𝐩)`$ the first two terms in (32) will contribute, leading to contributions in the polarisation tensor that dominate over the Hard Loop result (34) and are absent in the semiclassical approximation. As explained above, the extra contribution appears to be gauge independent. Second, for equilibrium distribution functions (12) the integrals over $`|𝐩|`$ and over the angle in (34) factorize. Then, it is easy to show that (34) reduces to the well known Hard Thermal Loop result where the polarization tensor has only two independent components and depends on $`\omega `$ and $`|𝐤|`$. Owing to transversality, the time-like components of $`\mathrm{\Pi }^{\mu \nu }`$ follow from $`\mathrm{\Pi }^{lm}`$. Indeed, $`\mathrm{\Pi }^{0m}(k)=k^l\mathrm{\Pi }^{lm}(k)/\omega `$ and $`\mathrm{\Pi }^{00}(k)=k^lk^m\mathrm{\Pi }^{lm}(k)/\omega ^2`$. In order to prove the transversality of the hard loop polarization tensor in the case of anisotropic distributions, we compute $`k_\mu \mathrm{\Pi }^{\mu \nu }(k)`$. Considering first the quark loop contribution, we get $$k_\mu \mathrm{\Pi }^{\mu \nu }(k)=ig^2\frac{d^4p}{(2\pi )^4}\left[2(kp)p^\nu p^2k^\nu k^2p^\nu \right]\left[\stackrel{~}{\mathrm{\Delta }}_F(q)\stackrel{~}{\mathrm{\Delta }}_R(p)+\stackrel{~}{\mathrm{\Delta }}_A(q)\stackrel{~}{\mathrm{\Delta }}_F(p)\right].$$ (35) After integrating over $`p_0`$ we find $`k_\mu \mathrm{\Pi }^{\mu 0}(k)={\displaystyle \frac{g^2}{2}}{\displaystyle \frac{d^3p}{(2\pi )^3}\frac{f_q(𝐩)}{|𝐩|}}`$ $`[{\displaystyle \frac{2(\omega |𝐩|𝐤𝐩)|𝐩|k^2|𝐩|}{2\omega |𝐩|+2𝐤𝐩+k^2i\mathrm{sgn}(|𝐩|\omega )0^+}}`$ (37) $`+{\displaystyle \frac{2(\omega |𝐩|𝐤𝐩)|𝐩|+k^2|𝐩|}{2\omega |𝐩|+2𝐤𝐩+k^2i\mathrm{sgn}(|𝐩|\omega )0^+}}]`$ and $`k_\mu \mathrm{\Pi }^{\mu m}(k)={\displaystyle \frac{g^2}{2}}{\displaystyle \frac{d^3p}{(2\pi )^3}\frac{f_q(𝐩)}{|𝐩|}}`$ $`[{\displaystyle \frac{2(\omega |𝐩|𝐤𝐩)p^mk^2p^m}{2\omega |𝐩|+2𝐤𝐩+k^2i\mathrm{sgn}(|𝐩|\omega )0^+}}`$ (39) $`+{\displaystyle \frac{2(\omega |𝐩|𝐤𝐩)p^mk^2p^m}{2\omega |𝐩|+2𝐤𝐩+k^2i\mathrm{sgn}(|𝐩|\omega )0^+}}].`$ Expanding the integrands in the these expressions for small external momenta analogously to (32), it is easy to show that (37) and (39) vanish in the Hard Loop approximation. This also holds if the gluon loop contribution is added as they have the same structure in the Hard Loop approximation. ### B Fermion self energy As mentioned in the Introduction, the diagrammatic technique has the advantage over the semiclassical transport theory approach that it can be easily extended to fermionic self energies. Therefore, we discuss the Hard Loop quark self energy for anisotropic momentum distributions. Using the Feynman gauge, the one-loop quark self energy is found as $$\mathrm{\Sigma }_{ij}(k)=2iC_F\delta _{ij}g^2\frac{d^4p}{(2\pi )^4}S(p)\mathrm{\Delta }(q),$$ (40) where $`C_F(N_c^21)/N_c`$ and now $`qkp`$. Adopting again the Keldysh representation, the gluon propagators in the Feynman gauge are given by $`\mathrm{\Delta }_{R,A}(q)`$ $`=`$ $`{\displaystyle \frac{1}{q^2\pm i\mathrm{sgn}(q_0)0^+}},`$ (41) $`\mathrm{\Delta }_F(q)`$ $`=`$ $`2\pi i\left[1+n_g(𝐪)\right]\delta (q^2).`$ (42) Suppressing the color indces, we find for the retarded quark self energy as $$\mathrm{\Sigma }(k)=ig^2C_F\frac{d^4p}{(2\pi )^4}p/\left[\stackrel{~}{\mathrm{\Delta }}_R(p)\mathrm{\Delta }_F(q)+\stackrel{~}{\mathrm{\Delta }}_R(p)\mathrm{\Delta }_A(q)+\stackrel{~}{\mathrm{\Delta }}_F(p)\mathrm{\Delta }_R(q)+\stackrel{~}{\mathrm{\Delta }}_A(p)\mathrm{\Delta }_R(q)\right].$$ (43) The matter part of (43) can be decomposed in two contributions which read after integrating over $`p_0`$ $`\mathrm{\Sigma }_1(k)={\displaystyle \frac{g^2}{8}}C_F{\displaystyle \frac{d^3p}{(2\pi )^3}\frac{n_g(𝐩)}{|𝐩|}}`$ $`[{\displaystyle \frac{(\omega p)\gamma _0(𝐤𝐩)𝜸}{2\omega |𝐩|+2𝐤𝐩+k^2+i\mathrm{sgn}(\omega |𝐩|)0^+}}`$ (45) $`+{\displaystyle \frac{(\omega +p)\gamma _0(𝐤𝐩)𝜸}{2\omega |𝐩|+2𝐤𝐩+k^2+i\mathrm{sgn}(\omega +|𝐩|)0^+}}]`$ and $`\mathrm{\Sigma }_2(k)={\displaystyle \frac{g^2}{16}}C_F{\displaystyle \frac{d^3p}{(2\pi )^3}\frac{n(𝐩)+\overline{n}(𝐩)}{|𝐩|}}`$ $`[{\displaystyle \frac{p\gamma _0+𝐩𝜸}{2\omega |𝐩|+2𝐤𝐩+k^2+i\mathrm{sgn}(\omega |𝐩|)0^+}}`$ (47) $`+{\displaystyle \frac{p\gamma _0+𝐩𝜸}{2\omega |𝐩|+2𝐤𝐩+k^2+i\mathrm{sgn}(\omega +|𝐩|)0^+}}].`$ In contrast to the polarization tensor, we need to expand the square brackets in (45) only to the first order for small external momenta, leading to $$\frac{p\gamma _0+𝐩𝜸}{2\omega |𝐩|+2𝐤𝐩i0^+}+\frac{p\gamma _0+𝐩𝜸}{2\omega |𝐩|+2𝐤𝐩+i0^+}.$$ (48) Assuming again the reflection symmetry for the distribution functions, we obtain the final gauge independent result in the Hard Loop approximation $$\mathrm{\Sigma }(k)=\frac{g^2}{16}C_F\frac{d^3p}{(2\pi )^3}\frac{2n_g(𝐩)+n(𝐩)+\overline{n}(𝐩)}{|𝐩|}\frac{\gamma _0+𝐯𝜸}{\omega +𝐯𝐤+i0^+}.$$ (49) In the case of isotropic distributions (49) reduces to the well known Hard Thermal Loop result , where the self energy for massless quarks contains only two independent scalar functions depending on $`\omega `$ and $`|𝐤|`$. Giving up the reflection symmetry of the distribution functions does not introduces new dominant terms in this case since the self energy follows already from the lowest order terms (48). We have adopted this symmetry to treat $`\mathrm{\Pi }`$ and $`\mathrm{\Sigma }`$ in exactly the same way. ## IV Dispersion relations The gluon polarization tensor and quark self energy can be used to determine the dispersion relations of gluons (plasmons) and quarks in the quasistatic and quasihomogeneous but anisotropic state of the quark-gluons plasma. ### A Gluon dispersion equation The background gluon field $`A^\mu (k)`$ satisfies the following equation of motion $$\left[k^2g^{\mu \nu }k^\mu k^\nu \mathrm{\Pi }^{\mu \nu }(k)\right]A_\nu (k)=0.$$ Therefore, the general plasmon dispersion equation is of the form $$\mathrm{det}\left[k^2g^{\mu \nu }k^\mu k^\nu \mathrm{\Pi }^{\mu \nu }(k)\right]=0.$$ (50) Equivalently, the dispersion relations are given by the positions of the pole of the effective gluon propagator. Due to the transversality of $`\mathrm{\Pi }^{\mu \nu }`$ not all components of $`\mathrm{\Pi }^{\mu \nu }`$ are independent from each other and consequently the dispersion equation (50) can be simplified. For this purpose we introduce the color permittivity tensor $`ϵ^{lm}(k)`$. Because of the relation $$ϵ^{lm}(k)E^l(k)E^m(k)=\mathrm{\Pi }^{\mu \nu }(k)A_\mu (k)A_\nu (k),$$ where $`𝐄`$ is the chromoelectric vector, the permittivity can be expressed through the polarization tensor as $$ϵ^{lm}(k)=\delta ^{lm}+\frac{1}{\omega ^2}\mathrm{\Pi }^{lm}(k).$$ (51) There are two other equalities which follow from the transversality of $`\mathrm{\Pi }^{\mu \nu }`$. Namely, $$\mathrm{\Pi }^{00}(k)=(ϵ^{lm}(k)\delta ^{lm})k^lk^m,\mathrm{\Pi }^{l0}(k)=(ϵ^{lm}(k)\delta ^{lm})\omega k^m.$$ Using the permittivity tensor the dispersion equation gets the form $$\mathrm{det}\left[𝐤^2\delta ^{lm}k^lk^m\omega ^2ϵ^{lm}(k)\right]=0$$ (52) with $$ϵ^{lm}(k)=\delta ^{lm}+\frac{g^2}{2\omega }\frac{d^3p}{(2\pi )^3}\frac{v^l}{\omega 𝐤𝐯+i0^+}\frac{f(𝐩)}{p^n}\left[\left(1\frac{𝐤𝐯}{\omega }\right)\delta ^{nm}+\frac{k^nv^m}{\omega }\right].$$ (53) In the isotropic state there are only two independent components of the permittivity tensor $$ϵ^{lm}(k)=ϵ_T(k)(\delta ^{lm}k^lk^m/𝐤^2))+ϵ_L(k)k^lk^m/𝐤^2,$$ and the dispersion equation (52) splits into two equations $$ϵ_T(k)=𝐤^2/\omega ^2,ϵ_L(k)=0.$$ The permittivity tensor (53) was calculated for the strongly elongated parton momentum distribution $`f(𝐩)`$ and it was found that there are unstable solutions of the dispersion equation (52). ### B Quark dispersion equation The quark dispersion relations are determined by the poles of the hard loop resummed quark propagator or equivalently are found as solutions of the equation $$\mathrm{det}[p/\mathrm{\Sigma }(p)]=0.$$ (54) One sees in Eq. (49) that the spinor structure of $`\mathrm{\Sigma }`$ is very simple: $`\mathrm{\Sigma }(p)=\gamma ^\mu \mathrm{\Sigma }_\mu (p)`$. However, we also include here the scalar part which is relevant for the massive quarks. Then, $$\mathrm{\Sigma }(p)=\gamma ^\mu \mathrm{\Sigma }_\mu (p)+C(p).$$ (55) Substituting the expression (55) into Eq. (54) and computing the determinant as explained in Appendix 1 of , we get $$\left[\left(p^\mu \mathrm{\Sigma }^\mu (p)\right)\left(p_\mu \mathrm{\Sigma }_\mu (p)\right)C^2(p)\right]^2=0.$$ (56) When the momentum distribution is isotropic, the structure of $`\mathrm{\Sigma }`$ further simplifies : $$\mathrm{\Sigma }(p)=A(p)p_0\gamma ^0+B(p)𝐩𝜸+C(p).$$ Then, the dispersion equation reads $$\left(1A(p)\right)^2p_0^2\left(1B(p)\right)^2𝐩^2C^2(p)=0.$$ ## V Discussion In the present work we have considered an anisotropic relativistic plasma which is either in equilibrium and the anisotropy is caused by external fields or the plasma is out of equilibrium. In the first case we deal with the homogeneous and static systems while in the second one it can be treated as quasihomogenous and quasistatic for sufficiently short space-time intervals. An example of the first case is the magnetized plasma while of the second one the parton system from the early stage of relativistic heavy-ion collisions where we encounter a strong anisotropy in the momentum distribution. The QCD polarization tensor has been computed in two ways. We have first applied the linear response method within the semiclassical transport theory and then the diagrammatic Hard Loop approach. The two methods are equivalent (but the distribution functions have to possess a reflection symmetry, i.e. $`f(𝐩)=f(𝐩)`$). When using the diagrammatic approach we have referred to the real time formalism since the systems under consideration are, in general, out of equilibrium. According to the Hard Loop approximation, we have used bare propagators for the internal lines of the polarization tensor which exhibit an explicit anisotropic momentum distribution. Another method has been used in to study anisotropic relativistic QED plasmas in a strong magnetic field. The system has been assumed there to be in equilibrium and dressed propagators corresponding to electrons in Landau levels have been adopted. In this way, anisotropic distributions arise although the distribution functions depend only on the energy. As already mentioned, the semiclassical kinetic theory of quarks and gluons has been shown so far to be fully consistent with the QCD dynamics only for quasiequilibrium . The considerations presented here demonstrate that the equivalence holds for the systems which are far from equilibrium although the space-time homogeneity must be invoked. Thus, the reliability of the kinetic theory methods is improved and in particular, the stability analysis of the parton system form the early stage of ultrarelativistic heavy-ion collisions, which has been based on the linearized kinetic equations, is substantiated. The main advantage of the diagrammatic approach over the transport one is that it allows for a systematic perturbative extension to higher order effects. Also the fermion self energy for anisotropic systems can be calculated in this way. Having the QCD polarization tensor and quark self energy derived here, one can construct effective gluon and quark propagators from the Dyson-Schwinger equation. The poles of the effective propagators determine (via Eqs. (52, 56)) the parton dispersion relations in an anisotropic quark-gluon plasma. In the isotropic plasma, the dispersion relations for gluons and quarks show two branches and start from the same energy at zero momentum . The point is that in this case there is no direction preferred and the longitudinal and transverse components of the dielectric function are identical when the momentum vanishes. For the anisotropic systems with a preferred direction even at zero momentum, we expect additional branches and the degeneracy at zero momentum to be removed. In equilibrium all modes are stable or damped due to the Landau mechanism. In the case of anisotropic systems growing modes, i.e. instabilities, are possible. The unstable modes were argued to occur in the parton system from the early stage of ultrarelativistic heavy-ion collisions . Since the characteristic time of instability development was estimated to be rather small (below 1 fm/c) these instabilities can significantly influence the temporal evolution of the parton system. Also the quark dispersion relations following from the effective quark propagator are of physical relevance, as they lead in equilibrium to interesting structures, e.g. van Hove peaks, in the dilepton production rate , which might serve as a signature for the quark-gluon plasma formation. It has to be seen whether these structures also survive in the nonequilibrium case. When the plasma is in the (isotropic) equilibrium state the zero frequency limit of the longitudinal component of the polarization tensor in the Hard Thermal Loop limit ($`\mathrm{\Pi }_L=\mathrm{\Pi }_{00}`$), which is identified with the lowest order Debye screening mass, is finite. The transverse component ($`\mathrm{\Pi }_T=(\delta _{lm}k_lk_m/|𝐤|^2)\mathrm{\Pi }_{lm}/2`$), on the other hand, shows no static magnetic screening. The situation is much more complicated in the anisotropic plasma. The screening length depends on the orientation of the vector $`𝐤`$ , see also . The diagrammatic approach, following the Hard Loop resummation technique , allows for a systematic calculation of observables, such as the energy loss of energetic partons or the production of photons and dileptons . Now, the program can be extended to the anisotropic quark-gluon plasma although one has to choose a specific form of the parton momentum distribution. We leave all these issues for future investigations. Here we have intended to provide only the general formalism to study anisotropic systems of quantum fields. ###### Acknowledgements. We are very grateful to the National Institute of Nuclear Theory at Seattle where this project was initiated during the Workshop Non-equilibrium Dynamics in Quantum Field Theory.
warning/0001/gr-qc0001073.html
ar5iv
text
# On the variational principle for dust shells in General Relativity ## I INTRODUCTION A thin spherically-symmetric dust shell is among the simplest popular models of collapsing gravitating configurations. The equations of motion of these objects are obtained in , . The construction of a variational principle for such systems was discussed from different points of view in -. There are a number of problems here, most basic of which is the dependence on the choice of the evolution parameter (internal, external, proper). The choice of time coordinate, in turn, affects the choice of a particular quantization scheme, leading, in general, to quantum theories which are not unitarily equivalent. In most of these papers the variational principle for shells is usually constructed in a comoving frame of reference, or in one of variants of freely falling frames of reference. However, use of such frames of reference frequently leads to effects unrelated to the object under consideration. The essential physics involves a picture of a gravitational collapse from the point of view of an infinitely remote stationary observer. In quantum theory this point of view enables us to treat bound states in terms of asymptotic quantities and to build the relevant scattering theory correctly. On the other hand, to treat primordial black holes in the theory of self-gravitating shells it is convenient to take the viewpoint of a central stationary observer. In the approach related to proper time of the shell reduction of the system leads to complicated Lagrangians and Hamiltonians which creates difficulties on quantization. In particular it leads to theories with higher derivatives or to finite difference equations. In our opinion, the choice of the exterior or interior stationary observers is most natural and corresponds to the real physics. To provide the necessary properties of invariance, specification of the canonical transformations in an extended phase space which translate the corresponding dynamical systems into one another enough. In addition the action for a shell should satisfy some natural requirements. In the absence of self-forces it should pass into the action for a geodesic motion. Further, according to the correspondence principle, at small velocities and masses of the shell, and also in the absence of other sources of the gravitational field, we should obtain the action for a self-gravitating Newtonian shell (see Appendix C). The natural Hamiltonian formulation of a self-gravitating shell was considered in works , . However this formulation was not obtained by a variational procedure from some initial action containing the standard Einstein-Hilbert term. The action for such a self-gravitating spherical shell of mass $`m`$ can be introduced with the help of a naive “relativization” of the Newtonian action. It is carried out by simple replacement of the kinetic energy $`mv^2/2`$ by the relativistic expression $`mc\sqrt{(1v^2/c^2)}`$ (see below Lagrangians (C19) and (93)). If there is an exterior gravitational field then the kinetic and potential energy of the shell have as their general relativistic analog the geodesic Lagrangian $`mc{}_{}{}^{(2)}ds_\pm /dt_\pm `$. The subsigns “<sub>±</sub>” correspond to exterior and interior observers. The gravitational self-action of the shell is the same for all cases, and its sign depends on whether stationary observer can exist inside and outside the shell. The above Hamiltonian formulation for the shell, as well as the procedure of “relativization” follows from the Lagrange formalism of dust shells constructed in the present paper. We view the system as a compound configuration consisting of two vacuum regions with a spatially-closed boundary surface formed by the shell. The initial action we take as the sum of actions of York type for either region and the action for dust matter. For the complete action introduced in this way the variational principle is compatible with the boundary-value problem of the corresponding Euler-Lagrange equations for either region of the configuration, and leads to “natural boundary conditions” on the shell. The missing boundary conditions are obtained by consideration of the variations with respect to normal displacements of the shell. The obtained conditions coincide with the known Israel matching conditions at singular hypersurfaces and are considered as constraints. Together with the equations of the gravitational field they are used to eliminate of the gravitational degrees of freedom. The tangential variations of thus-obtained action with constraints lead to the known equations of motion of the Israel . The problem of the complete reduction of the action is solved for spherically-symmetric systems. By transforming the variational formula and using the constraints the obtained action is reduced to two variants of the effective action. One of these variants describes the shell from an interior stationary observer’s point of view, and the other from the exterior one. Then we go over from the Lagrangian to the Hamiltonian description. The conditions of isometry of the exterior and interior sides of the shell lead to the momentum and Hamiltonian constraints. The canonical equivalence of these two variants of the description of the shells in the extended phase space indicates the existence of a “discrete gauge” transformation associated with the transition from the interior observer to the exterior one. The paper is organized as follows. In Sec.II the full action is constructed for a compound, piecewise smooth Lorentz manifold with a four-dimensional spatially-closed boundary surface between two vacuum regions, corresponding to the world sheet of the shell. From here the Einstein equations for regions outside the shell and surface equations follow. Further, the action for the shell and the equations of motion are constructed. In Sec.III spherically-symmetric relativistic dust shells are considered. The Lagrangians and Hamiltonians describing the shell from the point of view of the interior or exterior observer are obtained. Then momentum and Hamiltonian constraints are found. They emerge from independent consideration of the interior and exterior faces of the shell using the conditions of isometry of its two faces. In Sec.IY special cases of dust shells and configurations of several shells are briefly considered. In Appendix A it is shown that the surface equations, obtained in Sec.II, reduce to the known equations for jumps of the extrinsic curvature tensor of the shell. In Appendix B we show the canonical equivalence of the actions for the dust spherically-symmetric shell written relative to the interior and exterior observers. This equivalence is thought of as operating in the extended phase space of the corresponding dynamical system. In Appendix C the action for an arbitrary nonrelativistic gravitating dust shell is constructed. The Lagrangian for the spherical gravitating nonrelativistic dust shell is found. It was deemed worthwhile to consider the nonrelativistic case because it clarifies the interpretation of the results and allows comparisons with the general relativistic approach. In this work we consider both relativistic and non-relativistic systems. In this connection, we shall keep all the dimensional constants. Here $`c`$ is the velocity of light, $`\gamma `$ is the gravitational constant, $`\chi =8\pi \gamma /c^2`$, $`\mathrm{}`$ is Planck’s constant. The metric tensor $`g_{\mu \nu }(\mu ,\nu =0,1,2,3)`$ has the signature $`(+)`$. ## II The variational principle and equations of motion for relativistic dust shells Consider a time-like spatially-closed hypersurface $`\mathrm{\Sigma }_t^{(3)}`$ into some region $`D^{(4)}`$ of the space-time $`V^{(4)}`$. Let it be the world sheet of the infinitely thin dust shell with the surface density of dust $`\sigma `$. This shell divides the region $`D^{(4)}`$ into the interior and exterior ones, $`D_{}^{(4)}`$ and $`D_+^{(4)}`$. Introduce the general coordinate map $`x^\mu `$ on our compound manifold $`D^{(4)}=D_{}^{(4)}\mathrm{\Sigma }_t^{(3)}D_+^{(4)}`$ and the metrics $`g_{\mu \nu }^\pm `$ on $`D_\pm ^{(4)}`$, so that $`g_{\mu \nu }^{}|_{\mathrm{\Sigma }_t^{(3)}}=g_{\mu \nu }^{}|_{\mathrm{\Sigma }_t^{(3)}}`$. One defines the elements of the four-volume $`d^4\mathrm{\Omega }`$ on $`D_\pm ^{(4)}`$ and three-volume $`d^3\mathrm{\Omega }`$ on $`\mathrm{\Sigma }_t^{(3)}`$ according to the formulas $`d^4\mathrm{\Omega }`$ $`=`$ $`\sqrt{g}d^4x=\sqrt{g}dx^0dx^1dx^2dx^3,`$ (1) $`d^3\mathrm{\Omega }`$ $`=`$ $`\sqrt{g}n^\mu d\mathrm{\Sigma }_\mu =\sqrt{g}d\mathrm{\Sigma },`$ (2) where $`n^\mu `$ is the unit normal to $`\mathrm{\Sigma }_t^{(3)}`$, directed from $`D_{}^{(4)}`$ to $`D_+^{(4)}`$ $`(n_\mu n^\mu =1,u_\mu n^\mu =0)`$, $`g=det|g_{\mu \nu }|`$. Three-forms $`d\mathrm{\Sigma }_\mu `$ and $`d\mathrm{\Sigma }`$ are determined by the relations $$dx^\mu d\mathrm{\Sigma }_\nu =\delta _\nu ^\mu d^4x,\eta d\mathrm{\Sigma }=d^4x(d\mathrm{\Sigma }_\mu =n_\mu d\mathrm{\Sigma }),$$ (3) where “$``$” denotes the exterior product, and $`\eta =n_\mu dx^\mu `$ is a normal covector. Now let us fix coordinate system $`x^\mu `$ so that the coordinates $`x^a(a=2,3)`$ be Lagrange coordinates of particles on the shell $`\mathrm{\Sigma }_t^{(3)}`$. Then $`u^\mu x_{,\mu }^a=n^\mu x_{,\mu }^a=0`$, where “$`_{,\mu }`$” is derivative with respect to the coordinate $`x^\mu `$. Hence it follows $`u^a=n^a=0`$. The equations $`x^a=\text{const}`$ determine the world line $`\gamma `$ of some particle of dust on $`\mathrm{\Sigma }_t^{(3)}`$. The set $`\{\gamma \}=\{x^a,x^a+dx^a\}`$ of the world lines forms the elementary stream tube of dust. On the shell $`\mathrm{\Sigma }_t^{(3)}`$ we shall introduce the basis of one-forms $$\{e^0\omega =u_\mu dx^\mu ,e^a=dx^a\}(a,b=2,3)$$ (4) and the dual vector basis $`\{e_0u=u^\mu _\mu ,e_a\},e^i(e_k)=\delta _k^i(i,k=0,2,3).`$ (5) In the basis $`\{e^i\}=\{\omega ,dx^a\}`$ the metric tensor and three-form of volume on $`\mathrm{\Sigma }_t^{(3)}`$ are $`{}_{}{}^{(3)}g`$ $`=`$ $`\omega \omega q_{ab}dx^adx^b,`$ (6) $`d^3\mathrm{\Omega }=`$ $``$ $`\omega d^2\mathrm{\Omega },d^2\mathrm{\Omega }=\sqrt{q}dx^2dx^3.`$ (7) Here “$``$” is the sign of a tensor product, $`d^2\mathrm{\Omega }`$ is the surface element of the area for the section which is orthogonal to the elementary stream tube of dust, $`q_{ab}`$ is the metric on these sections, $`q=det|q_{ab}|`$. In the neighbourhood of the hypersurface $`\mathrm{\Sigma }_t^{(3)}`$ the metric tensor $`{}_{}{}^{(4)}g`$ and four-form of volume $`d^4\mathrm{\Omega }`$ can be expressed in the form $`{}_{}{}^{(4)}g`$ $`=`$ $`{}_{}{}^{(3)}g\eta \eta ,`$ (8) $`d^4\mathrm{\Omega }`$ $`=`$ $`\eta d^3\mathrm{\Omega }=\omega \eta d^2\mathrm{\Omega }.`$ (9) We introduce the two-form of mass on $`\mathrm{\Sigma }_t^{(3)}`$ by the formula $`d^2m=\sigma d^2\mathrm{\Omega }`$, then $`\sigma d^3\mathrm{\Omega }=\omega d^2m`$. Now we take the full action of the compound configuration in the form $$I_{tot}^{(g)}=I_{EH}c\underset{\mathrm{\Sigma }_t^{(3)}}{}\left(\sigma n^\mu +\frac{1}{2\chi }[\omega ^\mu ]\right)\sqrt{g}𝑑\mathrm{\Sigma }_\mu +I_{D^{(4)}}+I_0.$$ (10) It is the functional of the metric $`g_{\mu \nu }`$, density of the dust $`\sigma `$ and hypersurface $`\mathrm{\Sigma }_t^{(3)}`$: $`I_{tot}^{(g)}I_{tot}^{(g)}(g_{\mu \nu },\sigma ,\mathrm{\Sigma }_t^{(3)})`$. The first term in the right side of (10) $$I_{EH}=\frac{c}{2\chi }\underset{D_{}^{(4)}D_+^{(4)}}{}{}_{}{}^{(4)}Rd^4\mathrm{\Omega }$$ (11) is the Einstein-Hilbert action for the regions $`D_\pm ^{(4)}`$, where $`{}_{}{}^{(4)}R`$ is the curvature scalar. The second term in the right side (10) contains the matter term $`c\sigma d^3\mathrm{\Omega }`$ and matching term. The symbol $`[\omega ^\mu ]=\omega ^\mu |_+\omega ^\mu |_{}`$ denotes the jump of the quantity $`\omega ^\mu `$ $`=`$ $`g^{\sigma \rho }\mathrm{\Gamma }_{\sigma \rho }^\mu g^{\mu \rho }\mathrm{\Gamma }_{\sigma \rho }^\sigma ,`$ (12) $`\mathrm{\Gamma }_{\sigma \rho }^\mu `$ $`=`$ $`{\displaystyle \frac{1}{2}}g^{\mu \nu }(g_{\nu \rho ,\sigma }+g_{\nu \sigma ,\rho }g_{\rho \sigma ,\nu }),`$ (13) on $`\mathrm{\Sigma }_t`$. The sign “$`|_+`$” or “$`|_{}`$” indicates the marked values to be calculated as a limiting magnitude when approaching the boundary $`\mathrm{\Sigma }_t`$ from outside or inside respectively. In Appendix A it will be shown, that the relation $$[\omega ^\mu ]n_\mu =2[K]$$ (14) takes place. Here $`K=g^{\mu \nu }K_{\mu \nu }`$ is the trace of the extrinsic curvature tensor $$K_{\mu \nu }=n_{\mu ;\rho }h_\nu ^\rho (h_\nu ^\rho =\delta _\nu ^\rho +n^\rho n_\nu ),$$ (15) where $`_{;\rho }`$ is covariant derivative with respect to the coordinate $`x^\mu `$. The third term $$I_{D^{(4)}}=\frac{c}{2\chi }\underset{D^{(4)}}{}\omega ^\mu \sqrt{g}𝑑\mathrm{\Sigma }_\mu $$ (16) contains the surface terms which are introduced to fix the metric on the boundary $`D^{(4)}`$ of the region $`D^{(4)}`$. Note, that the boundary $`D^{(4)}`$ consists of the pieces of time-like as well as space-like hypersurfaces. The last term $`I_0`$ in (10) contains the boundary terms on the time-like infinitely remote hypersurfaces, necessary for normalization of the action. The relation $$\sqrt{g}{}_{}{}^{(4)}R=\sqrt{g}G+(\sqrt{g}\omega ^\mu )_{,\mu }$$ (17) takes place, where $$G=g^{\mu \nu }\left(\mathrm{\Gamma }_{\mu \sigma }^\rho \mathrm{\Gamma }_{\nu \rho }^\sigma \mathrm{\Gamma }_{\mu \nu }^\sigma \mathrm{\Gamma }_{\rho \sigma }^\rho \right)$$ (18) contains only the first derivatives of the metric. Therefore the action (10) can be rewritten in a more compact form $$I_{tot}^{(g)}=I_g+I_m+I_0,$$ (19) where $$I_g=\frac{c}{2\chi }\underset{D_{}^{(4)}D_+^{(4)}}{}\sqrt{g}Gd^4x=\underset{D_{}^{(4)}D_+^{(4)}}{}L_gd^4x$$ (20) is the gravitational action of the first order, and $$I_m=c\underset{\mathrm{\Sigma }_t^{(3)}}{}\sigma d^3\mathrm{\Omega }=c\underset{S_t^{(2)}}{}d^2m\underset{\gamma }{}\omega $$ (21) is the action for the dust. The first and the penultimate terms in (10) form the action which can be ascribed to that of the York’s type $`I_Y=I_{EH}+I_{D^{(4)}}`$. It is used in variational problems with the fixed metric on the boundary $`D^{(4)}`$ of the region $`D^{(4)}`$. It can also be used in variational problems with the general relativistic version of “natural boundary conditions” for “free edge” . In this case the metric on the boundary is arbitrary and the corresponding momenta vanishes. Together with $`I_0`$ it forms the York-Gibbons-Hawking action $`I_{YGH}=I_Y+I_0`$ for a free gravitational field. In our case of the compound configuration we also fix the metric on boundary $`D^{(4)}`$, as it was done in variational problem for action $`I_Y`$. In addition, inside the system there is the boundary surface $`\mathrm{\Sigma }_t^{(3)}`$, with singular distribution of matter on it. One can interpret this configuration as the two vacuum regions $`D_\pm ^{(4)}`$ with a common “loaded edge” (or with a “massive edge”). The sum of the actions of type $`I_Y`$ for these regions and of the action for matter $`I_m`$ and normalizing term $`I_0`$ do leads to the action $`I_{tot}^{(g)}`$. If there is no dust, $`\sigma =0`$, the common boundary is not “loaded”. Then, the requirement $`\delta I_{tot}^{(g)}=0`$, at arbitrary, everywhere continuous variations of the metric, gives generalization of the above “natural boundary conditions” for free hypersurface $`\mathrm{\Sigma }_t^{(3)}`$. They coincide with the condition of continuity for the extrinsic curvature on $`\mathrm{\Sigma }_t^{(3)}`$, i.e., with ordinary matching conditions. If the edge, being matched, is “loaded” by some surface distribution of matter, then we obtain the corresponding surface equation or the boundary conditions for $`D_\pm ^{(4)}`$. They are the analog of the generalized “natural boundary conditions” for “loaded edges”. The initial action is chosen so, that the surface equations on $`\mathrm{\Sigma }_t^{(3)}`$ following from the requirement $`\delta I_{tot}^{(g)}=0`$, coincide with the matching conditions on singular hypersurfaces . In this case, the variational principle for the action $`I_{tot}^{(g)}`$ will be compatible with the boundary-value problem of the corresponding Euler-Lagrange equations , . Note, that, as a rule, the boundary terms are formulated in terms of the extrinsic curvature of the corresponding hypersurfaces. For the configuration which contains the boundary hypersurface dividing the domain $`D^{(4)}`$ into parts and the whole boundary consisting of several pieces of edge, initial, and eventual hypersurfaces, it is more convenient to use the covariant approach. In order to calculate $`\delta I_{tot}^{(g)}`$ we use the complete action in the form (10). According to we have $$\delta \left(\sqrt{g}{}_{}{}^{(4)}R\right)=\sqrt{g}{}_{}{}^{(4)}G_{}^{\mu \nu }\delta g_{\mu \nu }+\left(\sqrt{g}\mathrm{\Omega }^\mu \right)_{,\mu },$$ (22) where $$\mathrm{\Omega }^\mu =g^{\sigma \rho }\delta \mathrm{\Gamma }_{\sigma \rho }^\mu g^{\mu \rho }\delta \mathrm{\Gamma }_{\sigma \rho }^\sigma ,$$ (23) and $`{}_{}{}^{(4)}G_{}^{\mu \nu }={}_{}{}^{(4)}R_{}^{\mu \nu }\frac{1}{2}{}_{}{}^{(4)}Rg^{\mu \nu }`$ is the Einstein tensor. In addition, we shall use the following conditions: the boundary of the configuration $`D^{(4)}`$, the metric on it, and the normal vector are fixed. Then $`\delta d\mathrm{\Sigma }_\mu |_{D^{(4)}}=0,\delta g_{\mu \nu }|_{D^{(4)}}=0,\delta n_\mu |_{D^{(4)}}=0`$. The hypersurface $`\mathrm{\Sigma }_t^{(3)}`$ is fixed, and the metric and its variations are continuous on $`\mathrm{\Sigma }_t^{(3)}`$: $`[g_{\mu \nu }]_{\mathrm{\Sigma }_t^{(3)}}=0`$, $`[\delta g_{\mu \nu }]_{\mathrm{\Sigma }_t^{(3)}}=0`$, $`[n_\mu ]_{\mathrm{\Sigma }_t^{(3)}}=0,[\delta n_\mu ]_{\mathrm{\Sigma }_t^{(3)}}=0`$. For the variation $`\delta I_m`$ according to the formula (21) we have $`\delta I_m=cd^2m\delta \omega =cd^2m\delta \omega _\gamma `$. Here, the quantity $`d^2m`$ is considered as a stationary value at variations of the metric . The sign “$`_{|\gamma }`$” designates restriction of the one-forms on the world line $`\gamma `$ so, that $$\delta \omega _{|\gamma }=\delta ds=\frac{1}{2}u^\mu u^\nu ds\delta g_{\mu \nu }=\frac{1}{2}u_\mu u_\nu \omega _{|\gamma }\delta g^{\mu \nu }.$$ (24) If all these conditions are satisfied, then from the requirement $`\delta I_{tot}^{(g)}=0`$ one obtains the vacuum Einstein equations $${}_{}{}^{(4)}G_{}^{\mu \nu }=0,D_\pm ^{(4)}$$ (25) and the surface equations on $`\mathrm{\Sigma }_t^{(3)}`$ $$Q_{\mu \nu }\frac{1}{2}Qg_{\mu \nu }=\chi \sigma u_\mu u_\nu ,$$ (26) where $`Q=g^{\mu \nu }Q_{\mu \nu }`$, and $$Q_{\sigma \rho }=n_\mu [\mathrm{\Gamma }_{\sigma \rho }^\mu ]\frac{1}{2}\left(n_\sigma [\mathrm{\Gamma }_{\mu \rho }^\mu ]+n_\rho [\mathrm{\Gamma }_{\mu \sigma }^\mu ]\right).$$ (27) It is shown in Appendix A that the surface equations (26) reduce to the known equations for the jump discontinuity of the extrinsic curvature tensor of the hypersurface $`\mathrm{\Sigma }_t^{(3)}`$ $`[K_{\mu \nu }][K]h_{\mu \nu }=\chi \sigma u_\mu u_\nu ,`$ (28) where $`h_{\mu \nu }=g_{\mu \nu }+n_\mu n_\nu `$ is the metric on $`\mathrm{\Sigma }_t^{(3)}`$. From the relations (28) it follows $$[K_{\mu \nu }]u^\mu u^\nu =\frac{\chi }{2}\sigma ,$$ (29) The missing equation for the average tensor of the extrinsic curvature $$\overline{K}_\nu ^\mu =\frac{1}{2}(K_{\nu |_+}^\mu +K_{\nu |_{}}^\mu )$$ (30) can be obtained by considering the variations of $`I_{tot}^{(g)}`$ with respect to normal displacements of the hypersurface $`\mathrm{\Sigma }_t^{(3)}`$. For this purpose we define some one-parameter family of time-like hypersurfaces in a neighbourhood of $`\mathrm{\Sigma }_t^{(3)}`$ so that $`\mathrm{\Sigma }_t^{(3)}`$ is included in this family. The family induces (3+1)-decomposition of the objects in the neighbourhood of $`\mathrm{\Sigma }_t^{(3)}`$. Thus for the four-curvature scalar one has $${}_{}{}^{(4)}R={}_{}{}^{(3)}R+K_\nu ^\mu K_\mu ^\nu K^2+\frac{2}{\sqrt{g}}\left\{\sqrt{g}(Kn^\mu a^\mu )\right\}_{,\mu },$$ (31) where $`a^\mu =n_{;\nu }^\mu n^\nu `$ and $`{}_{}{}^{(3)}R`$ is the curvature scalar of hypersurfaces of the family. Substituting (31) into (11) and taking into account the relations $`a^\mu n_\mu =0`$ and (14), one obtains the action (10) in the form $$I_{tot}^{(g)}=\widehat{I}_g+I_m+\widehat{I}_{D^{(4)}}+I_0,$$ (32) where $$\widehat{I}_g=\underset{D_{}^{(4)}D_+^{(4)}}{}\widehat{L}_gd^4x=\frac{c}{2\chi }\underset{D_{}^{(4)}D_+^{(4)}}{}\left({}_{}{}^{(3)}R+K_\nu ^\mu K_\mu ^\nu K^2\right)\sqrt{g}d^4x$$ (33) is the gravitational action, containing normal derivatives up to the first order, and $`\widehat{I}_{D^{(4)}}`$ and $`I_0`$ contain the boundary terms, which are unessential here. Now let every point $`p\mathrm{\Sigma }_t^{(3)}`$ be translated at a coordinate distance $`\delta x^\mu (p)=n^\mu \delta \lambda (p)`$ in the normal direction. As a result of the displacement one gets a new hypersurface $`\stackrel{~}{\mathrm{\Sigma }}_t^{(3)}`$. The initial and eventual positions of a shell are fixed, therefore $`\delta \lambda (p)=0,p\mathrm{\Sigma }_t^{(3)}D^{(4)}=\stackrel{~}{\mathrm{\Sigma }}_t^{(3)}D^{(4)}`$. In addition, we fix the metric $`g_{\mu \nu }`$ and all the quantities on $`\mathrm{\Sigma }_t^{(3)}`$, so that $`\delta I_m=0`$. As a result of the displacement of the hypersurface $`\mathrm{\Sigma }_t^{(3)}`$, the initial regions $`D_+^{(4)}`$ and $`D_{}^{(4)}`$ are transformed into new ones $`\stackrel{~}{D}_+^{(4)}`$ and $`\stackrel{~}{D}_{}^{(4)}`$, so that, $`\stackrel{~}{D}_{}^{(4)}\stackrel{~}{\mathrm{\Sigma }}_t^{(3)}\stackrel{~}{D}_+^{(4)}=D_{}^{(4)}\mathrm{\Sigma }_t^{(3)}D_+^{(4)}=D^{(4)}`$. Then, for example, the variation of the region $`D_{}^{(4)}`$ can be expressed in the form $`\delta D_{}^{(4)}=\stackrel{~}{D}_{}^{(4)}\backslash D_{}^{(4)}=D_+^{(4)}\backslash \stackrel{~}{D}_+^{(4)}`$. The variation of the action (33), under the above conditions, proves to be equal $$\delta I_{tot}^{(g)}=\delta \widehat{I}_g=\underset{\stackrel{~}{D}_{}^{(4)}\stackrel{~}{D}_+^{(4)}}{}\widehat{L}_gd^4x\underset{D_{}^{(4)}D_+^{(4)}}{}\widehat{L}_gd^4x\underset{\delta D_{}^{(4)}}{}\left(\widehat{L}_g^+\widehat{L}_g^{}\right)d^4x.$$ (34) Here $`\widehat{L}_g^+`$ and $`\widehat{L}_g^{}`$ are Lagrangians defined by the relation (33) and calculated as a limiting magnitude when approaching the hypersurface $`\mathrm{\Sigma }_t^{(3)}`$ from outside or inside respectively. Under the infinitesimal normal displacement of the hypersurface $`\mathrm{\Sigma }_t^{(3)}`$, the full action is variated by the formula $`\delta I_{tot}^{(g)}={\displaystyle \underset{\mathrm{\Sigma }_t^{(3)}}{}}\left(\widehat{L}_g^+\widehat{L}_g^{}\right)\delta x^\mu 𝑑\mathrm{\Sigma }_\mu ={\displaystyle \underset{\mathrm{\Sigma }_t^{(3)}}{}}[\widehat{L}_g]\delta \lambda 𝑑\mathrm{\Sigma }.`$ (35) Hence, from arbitrariness of $`\delta \lambda (p)`$ and the requirement $`\delta I_{tot}^{(g)}=0`$, one finds $$[\widehat{L}_g]=\widehat{L}_g^+\widehat{L}_g^{}=[K_\nu ^\mu K_\mu ^\nu K^2]=2\overline{K}_\nu ^\mu ([K_\mu ^\nu ][K]\delta _\mu ^\nu )=0.$$ (36) Here we considered that $`[^{(3)}R]=0`$ on $`\mathrm{\Sigma }_t^{(3)}`$. Then, using (28), from (36) we obtain $$\overline{K}_{\mu \nu }u^\mu u^\nu =0.$$ (37) The relations (28) and (37) form the necessary complete set of algebraic conditions or constraints for the extrinsic curvature tensor $`K_{\nu |_\pm }^\mu `$ of the hypersurface $`\mathrm{\Sigma }_t^{(3)}`$. Now we can eliminate gravitational degrees of freedom in the action $`I_{tot}^{(g)}`$ and construct the action for the shell. For this purpose it is necessary to calculate $`I_{tot}^{(g)}`$ on the solutions of the vacuum Einstein equations (25) taking into account the constraints (28) and (37). Note, first, that on this stage we use explicitly only the following results of these equations: $`{}_{}{}^{(4)}R`$ $`=`$ $`0,[\omega ^\mu ]n_\mu =2[K]=\chi \sigma .`$ (38) Substituting these relations for the corresponding terms in (10) one finds $$I_{tot}^{(g)}{}_{|\{equations(\text{38})\}}{}^{}=I_{sh}+I_{D^{(4)}}+I_0,$$ (39) where $$I_{sh}=\frac{1}{2}\underset{\mathrm{\Sigma }_t^{(3)}}{}c\sigma d^3\mathrm{\Omega }=\frac{c}{2}\underset{S_t^{(2)}}{}d^2m\underset{\gamma }{}\omega $$ (40) is the reduced action for the dust shell. This action must be considered together with constraints (28) and (37). The action $`I_{sh}^{(g)}`$ is quite certain if the gravitational fields in the neighbourhood of $`\mathrm{\Sigma }_t^{(3)}`$ are determined as the solutions of the vacuum Einstein equations (25) which satisfy the boundary conditions (28) and (37). That is the finding of these fields that completes the construction of the action for the shell. At this stage all the equations (25) and constraints (28), (37) are already used. Note, that one usually comes to the action for the shell in the other form. In our approach the action can be obtained at the partial reduction of initial action $`I_{tot}^{(g)}`$, when the constraint in (38) is not taken into account. As a result we come to the action of the type $$\stackrel{~}{I}_{sh}=c\underset{\mathrm{\Sigma }_t^{(3)}}{}\left(\sigma \frac{1}{\chi }[K]\right)\omega d^2\mathrm{\Omega }.$$ (41) or to some its modification. In the spherically-symmetric case from here follows the Lagrangian of the shell in a frame of reference of the comoving observer. However quantity $`[K]`$ contains second derivatives with respect to proper time of the shell. When eliminating them, through the integration by parts, one comes to rather complicated Lagrangians and Hamiltonians. To find the equations of motion for particles of the shell from action $`I_{sh}`$ (40) one should introduce the independent coordinates $`x_\pm ^\mu `$ in each of the regions $`D_\pm ^{(4)}`$, and the interior coordinates $`y^i(i,k=0,2,3)`$ on $`\mathrm{\Sigma }_t^{(3)}`$. Let the equations of embedding of $`\mathrm{\Sigma }_t^{(3)}`$ into $`D_\pm ^{(4)}`$ have the form $`x_\pm ^\mu =x_\pm ^\mu (y^i)`$. Then we can write the relations $`{}_{}{}^{(4)}ds_\pm ^2=g_{\mu \nu }^\pm dx_\pm ^\mu dx_\pm ^\nu ,{}_{}{}^{(3)}ds^2=g_{\mu \nu }^\pm x_{\pm ,i}^\mu x_{\pm ,k}^\mu dy^idy^k=h_{ik}dy^idy^k,`$ (42) $`\omega =\omega ^\pm =u_\mu ^\pm dx_\pm ^\mu ,\omega _{|\gamma }^\pm =ds_\pm ,u_\pm ^\mu =dx_\pm ^\mu /ds_\pm ,`$ (43) $`{}_{}{}^{(3)}\omega =u_\mu ^\pm x_{\pm ,i}^\mu dy^i=u_idy^i,{}_{}{}^{(3)}\omega _{|\gamma }^{}={}_{}{}^{(3)}ds,u^i=dy^i/{}_{}{}^{(3)}ds.`$ (44) Non-gravitational interaction between particles of the dust is absent. Therefore we consider quantity $`d^2m`$ to be unchanged when a flow line is varied. First, consider variations $`I_{sh}`$ with respect to the internal coordinates $`y^i`$. In this case $`_\gamma \omega =_\gamma ^{(3)}\omega _{|\gamma }=_\gamma ^{(3)}𝑑s`$. Then the metric $`h_{ik}(y^i)`$ is given on $`\mathrm{\Sigma }_t^{(3)}`$ and the variation of $`I_{sh}`$ leads to the equations of three-dimensional geodesic on the hypersurface $`\mathrm{\Sigma }_t^{(3)}`$ $$u_{;k}^iu^k=0.$$ (45) Here“$`_{;k}`$” denotes the covariant derivative with respect to the coordinate $`y^k`$ calculated with the help of the metric $`h_{ik}`$. The consideration of the variational principle $`\delta I_{sh}^{(g)}=0`$ with respect to the exterior coordinates $`x_\pm ^\mu `$ is more interesting treatment. In this case $`_\gamma \omega =_\gamma \omega _{|\gamma }^\pm =_\gamma ^{(4)}𝑑s_\pm `$ Then the metrics $`g_{\mu \nu }^\pm (x^\rho )`$ are given in a neighbourhood of the shell. Since the normal variations of the shell are already used, it is possible to consider the variations of dynamical quantities, generated only by the tangential to $`\mathrm{\Sigma }_t^{(3)}`$ variations of the coordinates $`x^\mu `$. These variations of the values will be denoted by the sign $`\stackrel{~}{\delta }`$. Thus, omitting for simplicity signs “$`\pm `$”, we have $$\stackrel{~}{\delta }x^\mu =\delta x^\mu +n^\mu n_\nu \delta x^\nu h_\nu ^\mu \delta x^\nu (n_\mu \stackrel{~}{\delta }x^\mu =0,h_\nu ^\mu =\delta _\nu ^\mu +n^\mu n_\nu ),$$ (46) where $`\delta x^\mu `$ are arbitrary values. Then we find $$\stackrel{~}{\delta }\omega _{|\gamma }=\stackrel{~}{\delta }^{(4)}ds=\stackrel{~}{\delta }\sqrt{g_{\mu \nu }dx^\mu dx^\nu }=u_{\mu ;\nu }u^\nu h_\rho ^\mu \delta x^\rho {}_{}{}^{(4)}ds+d(u_\mu \delta x^\mu ).$$ (47) Supposing that $`\delta x^\mu =0`$ on $`\mathrm{\Sigma }_t^{(3)}D^{(4)}`$, from the requirement $`\stackrel{~}{\delta }I_{sh}^{(g)}=0`$ we obtain the three-dimensional geodesic equations on $`\mathrm{\Sigma }_t^{(3)}`$, but, here, in the four-dimensional form $$u_{\mu ;\nu }u^\nu h_\rho ^\mu =0.$$ (48) This equation cab be rewritten as $$u_{\rho ;\nu }u^\nu =u_{\mu ;\nu }u^\nu n^\mu n_\rho .$$ (49) Hence, using the definition of $`K_{\mu \nu }`$ (15) one obtains $$u_{\rho ;\nu }u^\nu =n_\rho n_{\mu ;\nu }u^\mu u^\nu =n_\rho K_{\mu \nu }u^\mu u^\nu .$$ (50) Here we again introduce signs “$`\pm `$” and use the relations $`K_{\mu \nu |_\pm }=\overline{K}_{\mu \nu }\pm \frac{1}{2}[K_{\mu \nu }]`$. Then, taking into account constraints (29) and (37), we come to the equations of motion for the shell’s particles with respect to the exterior coordinates $$\left(u_{\mu ;\nu }u^\nu \right)_{|\pm }=\pm \frac{\chi }{4}\sigma n_\mu .$$ (51) For completeness one should add the unused constraints $$[K_{\mu \nu }]u^\mu e_i^\nu =0,[K_{\mu \nu }]e_a^\mu e_b^\nu =\frac{\chi \sigma }{2}h_{ab},$$ (52) where $`e_a^\mu =x^\mu /y^a`$. From (51) it follows the well-known Israel equations $`n^\mu {\displaystyle \frac{Du_\mu }{ds}}|_+`$ $`+`$ $`n^\mu {\displaystyle \frac{Du_\mu }{ds}}|_{}=0,e_i^\nu {\displaystyle \frac{Du_\mu }{ds}}|_\pm =0,`$ (53) $`n^\mu {\displaystyle \frac{Du_\mu }{ds}}|_+`$ $``$ $`n^\mu {\displaystyle \frac{Du_\mu }{ds}}|_{}={\displaystyle \frac{\chi \sigma }{2}},`$ (54) where $`Du_\mu =u_{\mu ;\nu }dx^\nu `$ is the covariant differential. The equations of motion of the dust shell (51) can immediately be found from the action $`I_{sh}`$. Indeed, acting in the same manner as when deducing the equations of motion (51), the variational formula (47) can be transformed to the form $$\stackrel{~}{\delta }^{(4)}ds_{|_\pm }=u_{\mu ;\nu }u^\nu \delta x_{|_\pm }^\mu {}_{}{}^{(4)}ds_{|_\pm }\frac{1}{2}[K_{\mu \nu }]u^\mu u^\nu n_\rho \delta x^\rho ds_{|_\pm }+d(u_\mu \delta x^\mu )_{|_\pm }.$$ (55) or $$\stackrel{~}{\delta }\omega _{|\gamma }^\pm =\stackrel{~}{\delta }^{(4)}ds_{|_\pm }=\left\{\left(u_{\mu ;\nu }u^\nu \pm \frac{\chi \sigma }{4}n_\mu \right)\delta x^\mu {}_{}{}^{(4)}ds+d(u_\mu \delta x^\mu )\right\}_{|\pm }.$$ (56) From here, under the above conditions, the equations of motion follow. The proposed variational deducing of the equations of motion makes the problem of construction of the effective action for the dust shell free from constrains (28) and (37). It turns out that it is possible for some special class of the configurations. To show it, we shall choose such interior coordinates $`y^i`$, which at $`i=a=2,3`$ are the Lagrange coordinates of particles on the shell $`\mathrm{\Sigma }_t^{(3)}`$. In addition, we introduce the coordinates $`x_{|\pm }^\mu `$ in the regions $`D_\pm ^{(4)}`$ so that, when $`\mu =a=2,3`$ the equalities $`x_+^a|_{\mathrm{\Sigma }_t^{(3)}}=x_{}^a|_{\mathrm{\Sigma }_t^{(3)}}=y^a`$ are satisfied. These coordinates are arbitrary in any other respect. Then the formulas of embedding of $`\mathrm{\Sigma }_t^{(3)}`$ into $`D_\pm ^{(4)}`$ have the form $`x_\pm ^n=x_\pm ^n(y^0)(n=0,1)`$ or $`f_\pm (x^0,x^1)=0`$. Therefore we have $`u_\pm ^\mu =\{u_\pm ^0,u_\pm ^1,0,0\}`$ and $`n_\mu ^\pm =\{n_0^\pm ,n_1^\pm ,0,0\}`$. Using the conditions $`(u_\mu u^\mu )_{|\pm }=(n_\mu n^\mu )_{|\pm }=1`$ and $`(u_\mu n^\mu )_{|\pm }=0`$ one finds $`n_0^\pm =u_\pm ^1,n_1^\pm =u_\pm ^0`$. Hence it follows $$n_\mu \delta x^\mu ds_{_{|\pm }}=(u^1\delta x^0u^0\delta x^1)ds_{_{|\pm }}=(dx^1\delta x^0dx^0\delta x^1)_{_{|\pm }}.$$ (57) Therefore the variational formula (56) has the form $$\stackrel{~}{\delta }\omega _{|\gamma }^\pm =\stackrel{~}{\delta }^{(4)}ds_{|_\pm }=\left\{\delta ^{(4)}ds\pm \frac{1}{4}\chi \sigma (dx^1\delta x^0dx^0\delta x^1)+d(u_\mu \delta x^\mu )\right\}_{|\pm }.$$ (58) Now we introduce the vector potential $`U_n=U_n(x^0,x^1)`$ by the relation $$d(U_ndx^n)G_{01}dx^0dx^1=\frac{1}{4}\chi \sigma dx^0dx^1,$$ (59) where $`G_{nm}U_{m,n}U_{n,m}(n,m=0,1)`$. Hence it follows that the configurations, being considered, admit such motions of matter for which $`\sigma =\sigma (x^0,x^1)`$. Using the definition (59) and the relation $$\delta (U_ndx^n)d(U_n\delta x^n)=G_{10}(dx^0\delta x^1dx^1\delta x^0),$$ (60) the variational formula (58) can be rewritten in the following form $`\stackrel{~}{\delta }\omega _{|\gamma }^\pm =\stackrel{~}{\delta }^{(4)}ds_{|_\pm }=\left\{\delta (dsU_ndx^n)+d[(u_n\pm U_n)\delta x^n+u_a\delta y^a]\right\}_{|\pm }.`$ (61) Returning to action for the shell (40), we conclude, that in the case under consideration we have $$\delta I_{sh}=\delta I_{sh}^\pm \frac{c}{2}\underset{S_t^{(2)}}{}d^2m\{(u_n\pm U_n)x_{,\mathrm{\hspace{0.17em}0}}^n\delta y^0+u_a\delta y^a\}_\pm |_A^B,$$ (62) where $$I_{sh}^\pm =\frac{c}{2}\underset{S_t^{(2)}}{}d^2m\underset{\gamma }{}\left(dsU_ndx^n\right)_{|\pm },(n=0,1)$$ (63) is the effective action for the shell written in terms of the exterior coordinates. Indices $`A`$ and $`B`$ indicate that the corresponding quantities are taken in initial and final positions of the shell. Since at fixed initial and final positions of particles $`\delta y^i|_{A,B}=0`$, then it follows $`\delta I_{sh}=\delta I_{sh}^\pm `$. In such away, under the above conditions, the action of the shell (40) with the constraints (28), (37) and the action (62) without these constraints are equivalent. The actions $`I_{sh}^+`$ and $`I_{sh}^{}`$ are equivalent in the same sense. Let us note, that in the considered above independent treatment of the interior and exterior faces of the shell there are new constraints following from isometry conditions of these faces. ## III Effective action for the spherical dust shell Let us consider spherically-symmetric compound region $`D^{(4)}=D_{}^{(4)}\mathrm{\Sigma }_t^{(3)}D_+^{(4)}V^{(4)}`$ into the spherically-symmetric space-time $`V^{(4)}`$, where $`D_{}^{(4)}`$ are exterior and interior regions separated from each other by spherically-symmetric time-like hypersurface $`\mathrm{\Sigma }_t^{(3)}`$. By using the curvature coordinates we can choose common in $`D_\pm ^{(4)}`$, spatial, spherical coordinates $`\{r,\theta ,\alpha \}`$, and individual time coordinates $`t_\pm `$ for $`D_\pm ^{(4)}`$ respectively. Then the world sheet for the shell $`\mathrm{\Sigma }_t^{(3)}`$ respectively the interior and exterior coordinates is determined by the equations $`r=R_{}(t_{})`$ and $`r=R_+(t_+)`$. Under appropriate choice of $`t_\pm `$ we have $`R_{}(t_{})=R_+(t_+)`$. Thus, the interior and exterior regions are determined by the relations $$D_{}^{(4)}=\{t_{},r,\theta ,\alpha :r_0<r<R_{}(t_{})\},D_+^{(4)}=\{t_+,r,\theta ,\alpha :R_+(t_+)<r<\mathrm{}\}$$ for all $`\{\theta ,\alpha \}\{0\theta \pi ,0\alpha <2\pi ,\}`$ and for all admissible $`t_\pm `$. The particles of the shell are described by one collective dynamical coordinate $`R=R_\pm (t_\pm )`$ and by the two fixed individual (Lagrange) angular coordinates $`\theta `$ and $`\alpha `$. The minimal value of $`r_0`$ is limited by the domain of definition of the curvature coordinates. The gravitational fields into the regions $`D_\pm ^{(4)}`$ are given by the metrics $${}_{}{}^{(4)}ds_\pm ^2=f_\pm c^2dt_\pm ^2f_\pm ^1dr^2r^2(d\theta ^2+\mathrm{sin}^2\theta d\alpha ^2),$$ (64) where $$f_\pm =1\frac{2\gamma M_\pm }{c^2r},$$ (65) and $`M_\pm `$ are the Schwarzschild masses $`(M_+>M_{})`$. Owing to the spherical symmetry $`\sigma =\sigma (t_\pm ,R)`$. Therefore the conditions of applicability of the modified action (62) are satisfied. In this case we have $`d^2m=\sigma d^2\mathrm{\Omega }=\sigma R^2\mathrm{sin}\theta d\theta d\alpha ,`$ (66) $`U_ndx^n=c\phi (t_\pm ,R)dt_\pm +U_R(t_\pm ,R)dR.`$ (67) Using the gauge condition $`U_R(t_\pm ,r)=0`$, the action (62) can be written in the form $$I_{sh}^\pm =\frac{c}{2}\underset{S_t^{(2)}}{}\sigma R^2\mathrm{sin}\theta d\theta d\alpha \underset{\gamma _\pm }{}\left({}_{}{}^{(2)}dsc\phi dt\right)_{|\pm }.$$ (68) Since the particles move only radially $`(\theta =\text{const},\phi =\text{const})`$ we shall use the truncate interval $${}_{}{}^{(2)}ds_\pm ^2=f_\pm c^2dt_\pm ^2f_\pm ^1dR^2.$$ (69) Further, from the formula (59) it follows $$\frac{1}{4}\chi \sigma =\frac{\gamma m}{2c^2R^2}=\frac{\phi }{R},$$ (70) where $`m=4\pi \sigma R^2`$ is the rest mass of the shell. Hence, up to an additive constant, one finds $$\phi =\frac{\gamma m}{2c^2R}.$$ (71) Finally, integrating in (68) over the angles $`\theta `$ and $`\alpha `$ and making use of (69) and (71), the effective action of the shell can be expressed in the form $$I_{sh}^\pm =\frac{1}{2}\underset{\gamma _\pm }{}L_{sh}^\pm 𝑑t_{|\pm }=\frac{1}{2}\underset{\gamma _\pm }{}\left(mc{}_{}{}^{(2)}ds\pm \frac{\gamma m^2}{2R}dt\right)_{|\pm },$$ (72) where $$L_{sh}^\pm =mc^2\sqrt{f_\pm f_\pm ^1R_{t\pm }^2/c^2}\pm U$$ (73) are the Lagrangians of the dust shell with respect stationary observes into the regions $`D_\pm ^{(4)},(R_{t\pm }=dR/dt_\pm )`$, and $$U^{(G)}=\frac{\gamma m^2}{2R}$$ (74) is the effective potential energy of the gravitational self-action of the shell. It is important that the self-action (74) has the same form as that in the Newtonian theory (formula (C14) in Appendix C). The Lagrangians (73) themselves can be obtained from the corresponding Newtonian analogs (see Appendix C, formulas (C15) and (C16)) by the formal replacement of the first and second terms describing the kinetic and potential energies of the shell into the external field by their general relativistic analog, the geodesic Lagrangian $`mc{}_{}{}^{(2)}ds_\pm /dt_\pm `$. It is natural that the Lagrangians (C15), (C16), up to an additive constant, are the Newtonian limits of the relativistic Lagrangians (73). It is easy to see that the actions (72) transform each into other under the discrete gauge transformation $$M_\pm \stackrel{}{}M_{}(f_\pm \stackrel{}{}f_{}),U^{(G)}\stackrel{}{}U^{(G)},t_\pm \stackrel{}{}t_{}.$$ This transformation generalize the corresponding transformation of the Newtonian theory of shells (see Appendix C) and reduce to the transformation from the interior observer to the exterior one and otherwise. Note, that despite the equivalence of the actions $`I_{sh}^\pm `$, similar to Newtonian case (C15), (C16), they can be considered quite independently. We also can consider the regions $`D_\pm ^{(4)}`$ together with the corresponding gravitational fields (64) separately and independently, as manifolds with the edge $`\mathrm{\Sigma }_{t\pm }^{(3)}`$. The edges $`\mathrm{\Sigma }_{t\pm }^{(3)}`$ acquire the physical meaning of different faces of the shell with the world sheet $`\mathrm{\Sigma }_t^{(3)}`$, provided the regions $`D_\pm ^{(4)}`$ are joined along these edges $`\mathrm{\Sigma }_{t\pm }^{(3)}`$. This can be performed only if the conditions of isometry of the edges $`\mathrm{\Sigma }_{t\pm }^{(3)}`$ are satisfied $$f_+c^2dt_+^2f_+^1dR^2=f_{}c^2dt_{}^2f_{}^1dR^2=c^2d\tau ^2,$$ (75) where $`\tau `$ is the proper time of the shell. In addition we have $`\mathrm{\Sigma }_{t+}^{(3)}=\mathrm{\Sigma }_t^{(3)}=\mathrm{\Sigma }_t^{(3)}`$, $`\gamma _+(t_+)=\gamma _{}(t_+)=\gamma `$. Now we study some results following from the isometry conditions of the edges. First, we obtain the relations between the velocities $$c^2\frac{f_+}{R_{t+}^2}\frac{1}{f_+}=c^2\frac{f_{}}{R_t^2}\frac{1}{f_{}},$$ (76) $$R_\tau ^2\left(\frac{dR}{d\tau }\right)^2=\frac{c^2R_{t\pm }^2}{c^2f_\pm f_\pm ^1R_{t\pm }^2},R_{t\pm }^2\left(\frac{dR}{dt_\pm }\right)^2=\frac{c^2f_\pm ^2R_\tau ^2}{c^2f_\pm +R_\tau ^2}.$$ (77) Then from the Lagrangians $`L_{sh}^\pm `$ (73) one finds the momenta and Hamiltonians of the shell $`P_\pm ={\displaystyle \frac{L_{sh}^\pm }{R_{t\pm }}}={\displaystyle \frac{mR_{t\pm }}{f_\pm \sqrt{f_\pm f_\pm ^1R_{t\pm }^2/c^2}}}={\displaystyle \frac{m}{f_\pm }}R_\tau ,`$ (78) $`H_{sh}^\pm ={\displaystyle \frac{mc^2f_\pm }{\sqrt{f_\pm f_\pm ^1R_{t\pm }^2/c^2}}}U=mc^2f_\pm {\displaystyle \frac{dt_\pm }{d\tau }}U`$ (79) or $`H_{sh}^\pm =c\sqrt{f_\pm (m^2c^2+f_\pm P_\pm ^2)}U=mc^2\sqrt{f_\pm +R_\tau ^2/c^2}U=E_\pm ,`$ (80) where $`E_\pm `$ are the energies of the shell which are conjugated to the time $`t_\pm `$ respectively and conserve with respect to the corresponding interior or exterior stationary observers’ point of view. After elimination of velocity $`R_\tau `$ from (78) and (80), the isometry conditions of the edges can be expressed in the form $`f_+P_+=f_{}P_{},`$ (81) $`\left(E_{}U\right)^2m^2c^4f_{}=\left(E_++U\right)^2m^2c^4f_+.`$ (82) Substituting $`U`$ and $`f_\pm `$ from (65) and (74) for those in the last relation and equating the coefficients at the same power of $`R`$ we obtain the relations between the Hamiltonian $`H_{sh}^\pm `$ and the Schwarzschild masses $`M_\pm `$ $$H_{sh}^+=H_{sh}^{}=(M_+M_{})c^2=E.$$ (83) Here $`E=E_\pm `$ is the full energy of the shell. This energy is conjugated to the coordinate time $`t_+`$ and $`t_{}`$ as well, and does not depend on the position of the stationary observer (inside or outside the shell). We shall interpret the relations (81) and (83) following from the above independent consideration of the shell faces, as momentum and Hamiltonian constraints. The Lagrangians $`L_{sh}^\pm `$ (73), as well as the relations (76) - (83), are valid only in a limited domain, since the used curvature coordinates are valid outside the event horizon only. Therefore $`L_{sh}^{}`$ can be used when $`R>2\gamma M_{}/c^2`$, and $`L_{sh}^+`$ when $`R>2\gamma M_+/c^2(M_+>M_{})`$. As is known, the complete description of the shells can be performed in the Kruskal-Szekeres coordinates. With respect to these coordinates the full Schwarzschild geometry consists of the four regions $`R^+,T^{},R^{},T^+`$, detached by the event horizons. Our above consideration concerned with the $`R^+`$ region only. Supposing $`r`$ to be a time coordinate, we can formally use the action in the form (72) under the horizon. However, here we encounter the ambiguity when choosing the sign before $`{}_{}{}^{(2)}ds`$. It is usually ascribed to ambiguity of the radial component direction of the unit normal to $`\mathrm{\Sigma }_t^{(3)}`$. The point is that in the curvature coordinates the regions $`T^{}`$ and $`T^+`$ coincide. Hence the time singularity $`r=0`$ contains the two singularities: past singularity and future singularity. Therefore, for instance, the movement of a test particle with the energy $`E=0`$ consists of the two stages. At the first stage the particle begins to move into the expanding region $`T^+`$ from the past singularity $`r=0`$ and reaches the horizon at a moment when $`r`$ reaches $`2\gamma M/c^2`$. Then it goes over into the contracting $`T^{}`$ region and moves from the horizon to the future singularity $`r=0`$. In the coordinates $`\{r,t\}`$, where $`r`$ is the time coordinate, the latter stage looks like the movement directed backwards in time. Similarly, in the curvature coordinates the regions $`R^{}`$ and $`R^+`$ of the Kruskal-Szekeres diagram coincide and ordinary movement of particles into the future of the $`R^{}`$-region looks as the movement directed backwards in time which corresponds to the change $`ds\stackrel{}{}ds`$. It means that ordinary particles moving into the $`R^{}`$-region are mapped into the $`R^+`$-region as antiparticles (remember the Feynman’s interpretation of antiparticles as ordinary particles moving backwards in time). Such trajectories can be taken into account by the change of the sign before $`mc{}_{}{}^{(2)}ds_\pm `$ in the expression for the action (72) of the shell. In order to use simplicity and convenience of the curvature coordinates and, at the same time, to keep information about shells into the $`R^{}`$-region we introduce an auxiliary discrete variable $`\epsilon =\pm 1`$ and make a change $`{}_{}{}^{(2)}ds_\pm \stackrel{}{}\epsilon _\pm {}_{}{}^{(2)}ds_\pm `$ in $`I_{sh}^\pm `$ (72). Herewith $`\epsilon _\pm =1`$ correspond to the shell into the $`R^+`$-region, and $`\epsilon _\pm =1`$ to the shell into the $`R^{}`$-region. Then, we introduce the quantities $`\mu _\pm =\epsilon _\pm m`$. As a result the extended action has the form $`I_{sh}^\pm (\mu _\pm )={\displaystyle \frac{1}{2}}{\displaystyle \underset{\gamma _\pm }{}}L_{sh}^\pm (\mu _\pm )𝑑t_{|\pm }={\displaystyle \frac{1}{2}}{\displaystyle \underset{\gamma _\pm }{}}\left(\mu c{}_{}{}^{(2)}dsU^Gdt\right)_{|\pm },`$ (84) where $$L_{sh}^\pm (\mu _\pm )=\mu _\pm c^2\sqrt{f_\pm f_\pm ^1R_{t\pm }^2/c^2}\pm U$$ (85) are the generalized Lagrangians describing the shell inside any of the $`R^\pm `$-regions with respect to the curvature coordinates of the interior $`\{t_{},R\}`$ or exterior $`\{t_+,R\}`$ regions. The event horizons $`R_g=2\gamma M_\pm /c^2`$ are, still, singular points of the dynamical systems (84) and must be excluded from consideration. For the extended system (84) the Hamiltonian has the form $$H_{sh}^\pm (\mu _\pm )=c\epsilon _\pm \sqrt{f_\pm (m^2c^2+f_\pm P_\pm ^2)}U=\mu _\pm c^2\sqrt{f_\pm +R_\tau ^2/c^2}U.$$ (86) Hence, taking into account the Hamiltonian constraints (83) one finds the standard relations of the theory of dust spherical shells . We shall rewrite them in terms of new designations $`\mu _{}\sqrt{f_{}+R_\tau ^2/c^2}`$ $``$ $`\mu _+\sqrt{f_++R_\tau ^2/c^2}={\displaystyle \frac{\gamma \mu ^2}{Rc^2}},`$ (87) $`\mu _{}\sqrt{f_{}+R_\tau ^2/c^2}`$ $`+`$ $`\mu _+\sqrt{f_++R_\tau ^2/c^2}=2(M_+M_{}).`$ (88) In the end of the section we write out the Hamilton-Jacobi equations corresponding to the Hamiltonians (86) and to the constraints (81), (83) for truncated actions $`S_0^\pm =S_0^\pm (R)`$ $$\frac{1}{f_\pm }\left(M_+M_{}\frac{U}{c^2}\right)^2\frac{f_\pm }{c^2}\left(\frac{dS_0^\pm }{dR}\right)^2=m^2.$$ (89) $$f_+dS_0^+=f_{}dS_0^{}.$$ (90) Then, the complete actions are determined by the formula $`S^\pm =c^2(M_+M_{})t_\pm +S_0^\pm `$. ## IV Particular cases of spherical dust configurations $``$ Self-gravitating dust shell. In this case $`M_{}=0`$. Denote $`M_+=M`$ and consider the shell moving into the $`R_+`$-region. Then with respect to the exterior coordinates, the Lagrangian and the Hamiltonian of the shell have the form $$L_{sh}^+=mc^2\sqrt{1\frac{2\gamma M}{c^2R}\left(1\frac{2\gamma M}{c^2R}\right)^1\frac{R_{t+}^2}{c^2}}\frac{\gamma m^2}{2R},$$ (91) $$H_{sh}^+=c\sqrt{1\frac{2\gamma M}{c^2R}}\sqrt{m^2c^2+\left(1\frac{2\gamma M}{c^2R}\right)P_+^2}+\frac{\gamma m^2}{2R}.$$ (92) The same shell with respect to the interior coordinates is described by the Lagrangian and the Hamiltonian $$L_{sh}^{}=mc^2\sqrt{1R_t^2/c^2}+\frac{\gamma m^2}{2R},$$ (93) $$H_{sh}^{}=c\sqrt{m^2c^2+P_{}^2}\frac{\gamma m^2}{2R}.$$ (94) This Hamiltonian was considered in the works , . The dynamical systems with $`L_{sh}^\pm `$ obey momentum and Hamiltonian constraints $`P_{}=f_+P_+`$$`H_{sh}^+=H_{sh}^{}=Mc^2`$ and they are canonically equivalent (see Appendix B). $``$ The dust shell with vanishing full energy. Now we consider the shell for which the binding energy $`E_b=(m+M_{}M_+)c^2`$ coincides with the rest energy $`mc^2`$. Denote $`M_+=M_{}M`$, $`f_+=f_{}f=12\gamma M/c^2R`$, $`t_+=t_{}t`$. Then for the full energy we have $`E=0`$. This is possible, as it follows from (87), (88 ), only when $`\mu _+=\mu _{}<0`$, i.e. for the wormhole. Such a shell can be considered as a classical model for “zeroth oscillations” of dust matter with bare mass $`m`$ under the gravitational field with $`f=12\gamma M/c^2R`$. In terms $`\{t,R\}`$ the trajectories of “zeroth oscillations” are determined by the equation $$\frac{dR}{dt}=\frac{2c^3}{\gamma m}\left(1\frac{2\gamma M}{c^2R}\right)\sqrt{\frac{\gamma ^2m^2}{4c^4}+\frac{2\gamma M}{c^2}RR^2}.$$ (95) Hence for the turning radius we have $$R_m=\frac{\gamma }{c^2}\left(M+\sqrt{M+\frac{m^2}{4}}\right).$$ (96) In the case of the flat space when $`M=0`$, from (95)and (96) we find $`{\displaystyle \frac{dR}{dt}}=c\sqrt{1{\displaystyle \frac{R^2}{R_{m0}^2}}},R_{m0}={\displaystyle \frac{\gamma m}{2c^2}}.`$ (97) The equations of motion of such “zeroth” shells coincide with those for the oscillator $$\frac{d^2R}{dt^2}+\omega ^2R=0.$$ (98) Its oscillations $`R(t)=R_{m0}\mathrm{cos}\omega (tt_0)`$ occur with the amplitude $`R_{m0}`$ and frequency $`\omega =c/R_{m0}=2c^3/\gamma m`$. Hence we find the time of life of these shells into the flat space-time as a half-period of the oscillation $$T=\frac{\pi }{\omega }=\frac{\pi \gamma m}{2c^3}=\frac{\pi }{c}R_{m0}.$$ (99) For the shell with mass equal to the mass of the Earth we have $`R_g=2\gamma M/c^24cm`$, $`R_{m0}=R_g/41cm,T10^{10}c`$. For the shells with Planck’s mass $`m=m_{pl}=\sqrt{\mathrm{}c/\gamma }`$ the time of life equals $`T=\pi T_{pl}/2`$, where $`T_{pl}=\sqrt{\mathrm{}\gamma /c^5}`$ is Planck’s time. We underline, that the “zero” shells are characterized by that their gravitational binding energy completely compensate proper energy, leaving their total energy to be equal to zero. These shells can be thought of as a classical prototype of the Wheeler’s space-time foam . $``$ The set of concentric dust shells. Now, consider briefly configurations consisting from the set of $`N`$ concentric dust shells. Let $`R_a,m_a,\tau _a`$ be the radius, bare mass and proper time of an $`a`$-th shell, respectively $`(a=1,2,\mathrm{},N)`$. For simplicity we suppose that $`R_a>R_b`$ if $`a>b`$. Then let $`M_a`$ be the Schwarzschild mass determining the gravitational field $`f_a=12\gamma M_a/c^2r`$ on the right side of an $`a`$-th shell, into the region $`R_a<r<R_{a+1}`$. Suppose $`f_a^{}=12\gamma M_{a1}/c^2R_a`$ and $`f_a^+=12\gamma M_a/c^2R_a`$. Let $`P_a^\pm =m_adR_a/f_a^\pm d\tau _a`$ be momenta of $`a`$-th shell, and $`U_a^{(G)}=\gamma m_a^2/2R_a`$ be its potential energy of the self-action. Then $$H_a^\pm =c\epsilon _a^\pm \sqrt{f_a^\pm \left(m_a^2c^2+f_a^\pm (P_a^\pm )^2\right)}U_a$$ (100) is the Hamiltonians of an $`a`$-th shell. They, similarly to momenta $`P_a^\pm `$, are considered from the stationary observers’ points of view, into the interior $`R_{a1}<r<R_a`$ and exterior $`R_a<r<R_{a+1}`$, regions respectively. They satisfy the momentum and Hamiltonian constraints $$f_a^+P_a^+=f_a^{}P_a^{},H_a^+=H_a^{}=(M_aM_{a1})c^2.$$ (101) Now we are ready to determine the full Hamiltonian of this configuration $$H_N=\underset{a=1}{\overset{N}{}}H_a^\pm .$$ (102) For the self-gravitating configuration $`M_0=0`$. Then $`H_1^\pm =M_1c^2`$ and the full Hamiltonian of the configuration satisfies the constrain $$H_N=Mc^2.$$ (103) Here $`M=M_N`$ is the Schwarzschild mass of the configuration. The system admits the discrete gauge transformations $$M_aM_{a1},U_aU_a,t_at_{a1}(a=1,2,\mathrm{},N),$$ where $`t_a`$ is coordinate time determined on the right from an $`a`$-th shell. The choice of sides (left or right) of the shells is not fixed beforehand and can be made by the reason of convenience. ###### Acknowledgements. I would like to acknowledge M.Korkina and S.Stepanov for helpful discussions of problems, touched in this paper. ## A Transformations of the surface equations We show that the surface equations (26) reduce to the known equations for the jumps of the extrinsic curvature tensor on the shell . First, we shall calculate $`n_\mu [\omega ^\mu ]`$. We suppose that the following conditions are satisfied on the hypersurface $`\mathrm{\Sigma }_t^{(3)}`$ $$[n_\mu ]=0,[n_{\mu ,\nu }]h_\sigma ^\nu =0,[n_{,\nu }^\mu ]h_\sigma ^\nu =0,[g_{\mu \nu ,\rho }]h_\sigma ^\rho =0.$$ (A1) Hence it follows $$[\mathrm{\Gamma }_{\rho \sigma }^\mu ]h_\mu ^\nu h_\beta ^\rho h_\alpha ^\sigma =0.$$ (A2) Then from the definition (15) one finds $`[\mathrm{\Gamma }_{\alpha \nu }^\sigma ]n_\sigma h_\beta ^\nu =[K_{\alpha \beta }],[\mathrm{\Gamma }_{\alpha \nu }^\sigma ]n_\sigma h^{\alpha \nu }=[K],`$ (A3) $`[\mathrm{\Gamma }_{\alpha \nu }^\sigma ]n^\alpha h_\beta ^\nu =[K_\beta ^\sigma ],[\mathrm{\Gamma }_{\alpha \nu }^\sigma ]n^\alpha h_\sigma ^\nu =[K],`$ (A4) $`[\mathrm{\Gamma }_{\alpha \nu }^\sigma ]n_\sigma n^\alpha h_\beta ^\nu =0.`$ (A5) According to (C14) we find $$n_\mu \omega ^\mu =n_\mu g^{\sigma \rho }\mathrm{\Gamma }_{\sigma \rho }^\mu n^\rho \mathrm{\Gamma }_{\rho \sigma }^\sigma =n_\mu h^{\sigma \rho }\mathrm{\Gamma }_{\rho \sigma }^\mu n^\mu h_\rho ^\sigma \mathrm{\Gamma }_{\mu \sigma }^\rho .$$ Therefore, making use of Eq. (A4) and (A5) we obtain the sought result (14). Then, projecting the equation (26) into the hypersurface $`\mathrm{\Sigma }_t^{(3)}`$ and into the normal $`n^\rho `$ one finds $`Q_{\sigma \rho }n^\rho {\displaystyle \frac{1}{2}}Qn_\sigma `$ $`=`$ $`0,`$ (A6) $`Q_{\sigma \rho }h_\alpha ^\sigma h_\beta ^\rho {\displaystyle \frac{1}{2}}Qh_{\alpha \beta }`$ $`=`$ $`\chi \sigma u_\alpha u_\beta .`$ (A7) Using the definitions (27) and Eqs. (A2)–(A5) we obtain $`Q=g_{\mu \nu }Q^{\mu \nu }=n_\mu [\omega ^\mu ]=2[K],`$ (A8) $`Q_{\sigma \rho }n^\rho =n_\mu n^\rho [\mathrm{\Gamma }_{\rho \sigma }^\mu ]{\displaystyle \frac{1}{2}}n_\sigma n^\rho [\mathrm{\Gamma }_{\rho \mu }^\mu ]+{\displaystyle \frac{1}{2}}[\mathrm{\Gamma }_{\sigma \mu }^\mu ]=`$ (A9) $`n_\sigma (n_\mu [\mathrm{\Gamma }_{\alpha \beta }^\mu ]n^\alpha n^\beta +[\mathrm{\Gamma }_{\mu \alpha }^\mu ]n^\alpha )=n_\sigma n^\rho [\mathrm{\Gamma }_{\rho \nu }^\mu ]h_\mu ^\nu =[K]n_\sigma ,`$ (A10) $`Q_{\sigma \rho }h_\alpha ^\sigma h_\beta ^\rho =n_\mu [\mathrm{\Gamma }_{\sigma \rho }^\mu ]h_\alpha ^\sigma h_\beta ^\rho =[K_{\alpha \beta }].`$ (A11) Thus the equation (A6) is satisfied identically, and the equations (A7) yield the sought relations (28). ## B On the canonical equivalence of the actions $`I_{sh}^\pm `$ for the dust spherical shell In order to show the canonical equivalence of the actions $`I_{sh}^\pm `$ in the extended phase space we write the variational principle (72) in the form $$\delta I_{sh}^\pm =\delta (P_\pm dRH_\pm dt_\pm )=0,$$ (B1) where $$P_\pm =\frac{1}{cf_\pm }\sqrt{(H_\pm \pm U)^2m^2c^4f_\pm }.$$ (B2) The dynamical systems with the actions $`I_{sh}^\pm `$ are restricted by momentum and Hamiltonian constraints (81) and (83), which follow from the independent consideration of the faces of the shell. The systems $`I_{sh}^\pm `$ will be canonically equivalent in the extended phase space of variables $`\{P_\pm ,H_\pm ,R,t_\pm \}`$, if $$dI_{sh}^+=dI_{sh}^{}+dF,$$ (B3) or $$P_+dRH_+dt_+=P_{}dRH_{}dt_{}+dF,$$ (B4) where $`F=F(R,t_+,t_{})`$ is the generating function of the canonical transformation $`\{P_+=P_+(P_{},t_{},R),t_+=t_+(P_{},t_{},R)\}`$. From (B4) we find $$H_+=\frac{F}{t_+},H_{}=\frac{F}{t_{}},P_+=P_{}\frac{F}{R}.$$ (B5) Using these relations the constraints (81) and (83) can be rewritten in the following way $$\frac{F}{t_+}=\frac{F}{t_{}}=E,\frac{F}{R}=P_{}\left(1\frac{f_{}}{f_+}\right).$$ (B6) From here we find $$F=E(t_{}t_+)+\sigma (R,E),$$ (B7) where $$\sigma (R,E)=\frac{1}{c}\left(\frac{1}{f_{}}\frac{1}{f_+}\right)\sqrt{(E\pm U)^2m^2c^4f_\pm }𝑑R.$$ (B8) The expression under the radical is invariant with respect to the replacement $`f_\pm f_{},UU`$. Then differentiating the expression (B7) over $`E`$ one finds the relation between $`t_+`$ and $`t_{}`$ $$\frac{F}{E}=t_{}t_++\frac{\sigma }{E}=\alpha ,$$ (B9) where the constant $`\alpha `$ cab be omitted. Thus, the transformations $`\{\begin{array}{ccc}t_+\hfill & =\hfill & t_{}+\frac{\sigma (R,E)}{E},\hfill \\ & & \\ P_+\hfill & =\hfill & P_{}+\frac{\sigma (R,E)}{R},\hfill \end{array}`$ (B13) are the sought canonical transformations of the extended phase space $`\{P_\pm ,H_\pm ,R,t_\pm \}`$ of the system. Herewith, the corresponding presymplectic form $$dP_+dRdEdt_+=dP_{}dRdEdt_{}$$ (B14) is invariant under these transformations. The difference of the shell actions, which are considered from the point of view of the exterior and interior observers, up to an additive constant is $`I_{sh}^+(R,t_+)I_{sh}^{}(R,t_{})=F(R,t_+,t_{})`$, where $`F=F(R,t_+,t_{})`$ is the generating function, is calculated by the elimination of the energy $`E`$ from the relations (B7) and (B9). ## C Non-relativistic dust shell Let us consider an infinitely thin dust layer in the Euclidean space $`R^{(3)}`$ in the form of the closed surface $`\mathrm{\Sigma }_t`$ moving in its own Newtonian gravitational field $`\phi =\phi (\stackrel{}{r})`$. The full action for this configuration has the form $`I_{tot}^{(N)}={\displaystyle \underset{t_1}{\overset{t_2}{}}}𝑑t\left\{{\displaystyle \underset{\mathrm{\Sigma }_t}{}}\left({\displaystyle \frac{1}{2}}\sigma \stackrel{}{v}^{\mathrm{\hspace{0.17em}2}}\sigma \phi \right)d^2f{\displaystyle \frac{1}{8\pi \gamma }}{\displaystyle \underset{D_{}D_+}{}}(\phi )^2𝑑V\right\}.`$ (C1) Here $`\mathrm{\Sigma }_t(t_1tt_2)`$ is a one-parameter family of the closed surfaces, $`D_{}`$ and $`D_+`$ are interior and exterior regions of the shell $`\mathrm{\Sigma }_t`$ at a moment $`t`$, $`d^2f`$ is the surface element on $`\mathrm{\Sigma }_t`$, $`dV`$ is the volume element in $`R^3`$, $`\stackrel{}{v}`$ is the velocity of particles of the shell, $``$ is the nabla operator, $`\sigma `$ is surface mass density of dust on $`\mathrm{\Sigma }_t`$. By virtue of the mass conservation law the value $`dm\sigma d^2f`$ conserves along the stream tube and in case of dust can be considered as a stationary value under arbitrary variations. Note also, that, we require that the potential $`\phi `$ be continuous, and, together with all its derivatives, vanish at infinity. The requirement of extremity of the action $`\delta I_{tot}^{(N)}=0`$ with respect to everywhere continuous variations $`\delta \phi `$, vanishing on infinity, leads to the Laplace equation $`\mathrm{\Delta }\phi (\stackrel{}{r})=0,\stackrel{}{r}D_{}D_+`$ (C2) with the boundary conditions for normal derivatives of $`\phi `$ on $`\mathrm{\Sigma }_t`$. The latter, for completeness, will be written out together with the continuity conditions for $`\phi `$ on $`\mathrm{\Sigma }_t`$: $`[\phi ]\phi _{|_+}\phi _|_{}=0,\left[{\displaystyle \frac{\phi }{\eta }}\right]{\displaystyle \frac{\phi }{\eta }}|_+{\displaystyle \frac{\phi }{\eta }}|_{}=4\pi \gamma \sigma ,\stackrel{}{r}\mathrm{\Sigma }_t(t=\text{const}).`$ (C3) Here $`/\eta =(\stackrel{}{n})`$ is the derivative with respect to the exterior normal $`\stackrel{}{n}`$ to $`\mathrm{\Sigma }_t`$, vector $`\stackrel{}{n}(\stackrel{}{n}^2=1)`$ is directed from $`D_{}`$ to $`D_+`$. The solution of the equations (C2), (C3) is the potential of a “simple layer” $`\phi _\sigma (\stackrel{}{r})=\gamma {\displaystyle \underset{\mathrm{\Sigma }_t}{}}{\displaystyle \frac{\sigma (\stackrel{}{r}^{})d^2f^{}}{|\stackrel{}{r}\stackrel{}{r}^{}|}}.`$ (C4) Now we can calculate $`I_{tot}^{(N)}`$ on the solutions of the equations (C2), (C3). Thereby, the potential $`\phi `$ is excluded from the full action (C1). Note that owing to (C2) we have $`(\phi )^2=(\phi \phi )`$. This allows to transform the volume integral of (C1) into the surface one on the boundaries of the regions $`D_\pm `$. Taking into account the boundary conditions (C3) and an asymptotic behaviour of $`\phi `$, we find the reduced action $`I_{tot}^{(N)}`$, as the value of the initial action on the solution (C4) of the equations (C2) and (C3) $$I_{tot}^{(N)}{}_{|\{solutionseq.(\text{C2}),(\text{C3})\}}{}^{}=I_{sh}^{(N)}+I_0^{(N)}.$$ (C5) Here $`I_0^{(N)}`$ contains the surface term, which is unessential for further consideration, and $`I_{sh}^{(N)}={\displaystyle \underset{t_1}{\overset{t_2}{}}}L_{sh}^{(N)}𝑑t`$ (C6) is the effective action for the shell with the Lagrangian $`L_{sh}^{(N)}={\displaystyle \frac{1}{2}}{\displaystyle \underset{\mathrm{\Sigma }_t}{}}\sigma \stackrel{}{v}^{\mathrm{\hspace{0.17em}2}}𝑑fU,`$ (C7) where $`U={\displaystyle \frac{1}{2}}{\displaystyle \underset{\mathrm{\Sigma }_t}{}}\sigma \phi _\sigma 𝑑f={\displaystyle \frac{\gamma }{2}}{\displaystyle \underset{\mathrm{\Sigma }_t}{}}{\displaystyle \underset{\mathrm{\Sigma }_t}{}}{\displaystyle \frac{\sigma (\stackrel{}{r})\sigma (\stackrel{}{r}^{})}{|\stackrel{}{r}\stackrel{}{r}^{}|}}𝑑f𝑑f^{}`$ (C8) is the functional of the potential energy of the gravitational self-action of the shell. The Lagrangian of the shell in an external gravitational field $`\phi _0=\phi _0(\stackrel{}{r})`$ has the form $`L_{sh}^{(N)}={\displaystyle \underset{\mathrm{\Sigma }_t}{}}\left({\displaystyle \frac{1}{2}}\sigma \stackrel{}{v}^{\mathrm{\hspace{0.17em}2}}\sigma \phi _0\right)𝑑fU.`$ (C9) Now consider the spherical non-relativistic dust shell. Let $`R=R(t)`$ be the radius of the spherical shell at a moment $`t`$. With respect to the spherical coordinates $`\{r,\theta ,\alpha \}`$, we have $`\sigma =\sigma (r),d^2f=R(t)\mathrm{sin}\theta d\theta d\alpha ,\stackrel{}{v}^{\mathrm{\hspace{0.17em}2}}=\dot{R}^2=(dR/dt)^2`$. The mass of the shell is $`m=4\pi \sigma R^2=\text{const}`$. Potential of the external field $`\phi _0`$ on the shell has the value $$\phi _0\phi _{}=\frac{\gamma m_{}}{R(t)},$$ (C10) where $`m_{}`$ is the total mass of the interior source. The potential $`\phi _\sigma `$ and the self-action energy for the shell $`U`$ prove to be equal $`\phi _\sigma (r)`$ $`=`$ $`\{\begin{array}{ccc}\gamma m/r,\hfill & rR(t)\hfill & \\ \gamma m/R(t),\hfill & r<R(t)\hfill & \end{array},`$ (C13) $`U`$ $`=`$ $`{\displaystyle \frac{1}{2}}m\phi _\sigma ={\displaystyle \frac{\gamma m^2}{2R(t)}}.`$ (C14) Replacing the corresponding terms in (C9) by those of (C10) - (C14), we obtain the Lagrangian of the spherically-symmetric dust shell in Newtonian theory of gravity $`L_{sh}^{(N)}={\displaystyle \frac{1}{2}}m\dot{R}^2+{\displaystyle \frac{\gamma mm_{}}{R}}U.`$ (C15) A distinctive feature of spherical shell is that the two-valued description of the shell dynamic becomes possible with respect to the observer’s position. From an interior observer’s point of view, except for the force of self-action $`U/R`$, the shell is effected by the external force $`F_{}=md\phi _{}/dr`$, which determines an interior gravitational field. This situation corresponds to the Lagrangian $`L_{sh}^{(N)}`$, therefore the latter can be interpreted as the Lagrangian describing the non-relativistic shell from an interior observer’s point of view. An exterior observer $`(r>R(t))`$ determines the field, judging by the force $`F_+=md\phi _+/dr`$ acting on the shell in the field $`\phi _+=\phi _{}+\phi _\sigma =\gamma m_+/R(t)`$. This field is generated by the total mass of the system $`m_+=m_{}+m`$. If, by making use of this relation, we eliminate $`m_{}`$ from $`L_{sh}^{(N)}`$ we obtain the following Lagrangian $`L_{sh+}^{(N)}={\displaystyle \frac{1}{2}}m\dot{R}^2+{\displaystyle \frac{\gamma mm_+}{R}}+U.`$ (C16) It can be interpreted as the Lagrangian describing the Newtonian shell from an exterior observer’s point of view. In such a way, transformation from an exterior observer to an interior one stipulates the discrete transformation $`m_\pm \stackrel{}{}m_{}=m_\pm m`$. Its can be interpreted as both the gravitational potential transformation $`\phi _\pm \stackrel{}{}\phi _{}=\phi _\pm \pm \gamma m/R`$ and the change of sign of the self-action potential $`U^{(N)}\stackrel{}{}U^{(N)}`$. The above two-valued description of spherical shell in the Newtonian theory has a formal character. This ambiguity, arising when describing spherically-symmetric shell, is a matter of principle in General Relativity. Note other feature of the shell, which has non-trivial meaning in General Relativity. The Lagrangians $`L_{sh\pm }^{(N)}`$ completely and closely determine the motion of boundaries of regions $`D_\pm `$. That is why they can be thought of as independent systems with their momenta and Hamiltonians $`P_\pm =m\dot{R}_\pm ,H_\pm ={\displaystyle \frac{P_\pm ^2}{2m}}{\displaystyle \frac{\gamma mm_\pm }{R_\pm }}U_\pm =E_\pm .`$ (C17) Here $`E_\pm `$ are the total energies of these boundaries, $`U_\pm ^{(N)}=\gamma m^2/2R_\pm `$ are their potential energies of self-action, and $`R_\pm =R_\pm (t)`$ are the radiuses of the regions’ boundary $`D_\pm `$. The systems (C17) describe the same shell provided the regions $`D_\pm `$ have a common boundary $`R_+(t)=R_{}(t)R(t)`$ for all moments $`t`$. In this case, eliminating the momentum $`PP_+=P_{}`$ from the equations $`H_\pm =E_\pm `$ one has $$E_+E_{}=\frac{\gamma m}{R}(m_+mm_{}).$$ (C18) Hence it follows an ordinary equality of energies and the additivity of masses $`E_+=E_{},m_+=mm_{}`$. In General Relativity, a similar but not trivial procedure follows from the isometry conditions for the boundaries of the corresponding four-dimensional regions. Finally we shall write the corresponding relations for a self-gravitating shell, when $`m_{}=0`$: $`L_{sh}^{(N)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}m\dot{R}^2+{\displaystyle \frac{\gamma m^2}{2R}},`$ (C19) $`P=m\dot{R},H`$ $`=`$ $`{\displaystyle \frac{P^2}{2m}}{\displaystyle \frac{\gamma m^2}{2R}}=E.`$ (C20)
warning/0001/nlin0001038.html
ar5iv
text
# Difference equations for the higher rank XXZ model with a boundary ## 1 Introduction Representation theory of the affine quantum group plays an important role in the description of solvable lattice models and massive integrable quantum field theories in two dimentions . For models with the affine quantum group symmetry the difference analogue of the Knizhnik-Zamolodchikov equations (quantum Knizhnik-Zamolodchikov equations) are satisfied by both correlation functions and form factors . Integrable models with boundary reflection have been also studied in lattice models and massive quantum theories. The boundary interaction is specfied by the reflection matrix $`K`$ for lattice models , and by the boundary $`S`$-matrix for massive quantum theories . It is shown in that the space of states of the boundary XXZ model can be described in terms of vertex operators associated with the bulk XXZ model . The explicit bosonic formulae of the boundary vacuum of the boundary XXZ model were obtained by using the bosonization of the vertex operators . This approach is also relevant for other various models . It is shown in that correlation functions and form factors in semi-infinite XXZ/XYZ spin chains with integrable boundary conditions satisfy the boundary analogue of the quantum Knizhnik-Zamolodchikov equation. In this paper we establish the similar results for the $`U_q(\widehat{sl_n})`$ -analogue of XXZ spin chain with a boundary magnetic field $`h`$: $`_B={\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}\left\{q{\displaystyle \underset{\genfrac{}{}{0pt}{}{a,b=0}{a>b}}{\overset{n1}{}}}e_{aa}^{(k+1)}e_{bb}^{(k)}+q^1{\displaystyle \underset{\genfrac{}{}{0pt}{}{a,b=0}{a<b}}{\overset{n1}{}}}e_{aa}^{(k+1)}e_{bb}^{(k)}{\displaystyle \underset{\genfrac{}{}{0pt}{}{a,b=0}{ab}}{\overset{n1}{}}}e_{ab}^{(k+1)}e_{ba}^{(k)}\right\}`$ (1.1) $`+{\displaystyle \frac{1q^2}{2q}}\left\{{\displaystyle \underset{a=0}{\overset{L1}{}}}e_{aa}^{(1)}{\displaystyle \underset{a=M}{\overset{n1}{}}}e_{aa}^{(1)}\right\}+h{\displaystyle \underset{a=L}{\overset{M1}{}}}e_{aa}^{(1)},`$ where $`1<q<0`$ and $`0LMn1`$. On the basis of the boundary vacuum states constructed in we derive the boundary analogue of the quantum Knizhnik-Zamolodchikov equations for the correlation functions in the higher rank XXZ model with a boundary. We also obtain the two point functions by solving the simplest difference equations for free boundary condition. The rest of this paper is organized as follows. In section 2 we review the vertex operator approach for the higher rank XXZ model with a boundary. In section 3 we derive the boundary quantum Knizhnik-Zamolodchikov equations for the $`2N`$-point correlation functions. In section 4 we obtain the two point functions by solving the difference equation with $`N=1`$ for free boundary condition. In Appendix A we summarize the results of the bosonization of the vertex operators in $`U_q(\widehat{sl_n})`$ . In Appendix B we summarize the bosonic formulae of the boundary vacuum states . ## 2 Formulation The higher rank XXZ model with boundary reflection was formulated in in terms of the vertex operators of the quantum affine group $`U_q(\widehat{sl_n})`$. For readers’ convenience let us briefly review the results in . Throughout this paper we fix $`n_2`$, and also fix $`q`$ such that $`1<q<0`$. The model is labeled by the three parameters $`i,L,M`$ such that $`0LMn1`$ and $`i\{L,M\}`$. In this paper we consider the following three cases: $$\begin{array}{cc}\hfill (C1)& 0L=M=in1,\hfill \\ \hfill (C2)& 0L=i<Mn1,\hfill \\ \hfill (C3)& 0L<M=in1.\hfill \end{array}$$ In what follows we denote the $`q`$-integer $`(q^kq^k)/(qq^1)`$ by $`[k]`$, and we use the following simbols: $`b(z)={\displaystyle \frac{qq^1z}{1z}},c(z)={\displaystyle \frac{qq^1}{1z}}.`$ (2.1) The nonzero entries of the R-matrix $`R^{(i)VV}(z)`$ are given by $`R^{(i)VV}(z)_{j_1,j_2}^{k_1,k_2}=r^{(i)VV}(z)\times \{\begin{array}{cc}1,& j_1=j_2=k_1=k_2\\ b(q^2z),& j_1=k_1j_2=k_2,\\ qc(q^2z),& j_1=k_2<j_2=k_1,\\ qzc(q^2z),& j_1=k_2>j_2=k_1.\end{array}`$ (2.6) Here the scalar functions are $`r^{(i)VV}(z)=z^{\delta _{i,0}}{\displaystyle \frac{(q^2z^1;q^{2n})_{\mathrm{}}(q^{2n}z;q^{2n})_{\mathrm{}}}{(q^2z;q^{2n})_{\mathrm{}}(q^{2n}z^1;q^{2n})_{\mathrm{}}}},`$ (2.7) where $$(z;p_1,\mathrm{},p_m)_{\mathrm{}}=\underset{k_1,\mathrm{},k_m=0}{\overset{\mathrm{}}{}}(1zp_1^{k_1}\mathrm{}p_m^{k_m}).$$ The boundary K-matrix $`K^{(i)}(z)`$ is a diagonal matrix, whose diagonal elements are given by $`K^{(i)}(z)_j^j={\displaystyle \frac{\phi ^{(i)}(z)}{\phi ^{(i)}(1/z)}}\times \{\begin{array}{cc}z^2,& 0jL1,\\ \frac{1rz}{1r/z},& LjM1,\\ 1,& Mjn1,\end{array}`$ (2.11) where we have set $`\phi ^{(i)}(z)=z^{\delta _{i,0}1}{\displaystyle \frac{(q^{2n+2}z^2;q^{4n})_{\mathrm{}}}{(q^{4n}z^2;q^{4n})_{\mathrm{}}}}\times \{\begin{array}{cc}1,& \mathrm{for}(C1),\\ \frac{(rq^{2n}z;q^{2n})_{\mathrm{}}}{(rq^{2n2M+2L}z;q^{2n})_{\mathrm{}}},& \mathrm{for}(C2),\\ \frac{(r^1z;q^{2n})_{\mathrm{}}}{(r^1q^{2M2L}z;q^{2n})_{\mathrm{}}},& \mathrm{for}(C3).\end{array}`$ (2.15) They satisfy the boundary Yang-Baxter equation: $`K_2^{(i)}(z_2)R_{21}^{(i)}(z_1z_2)K_1^{(i)}(z_1)R_{12}^{(i)}(z_1/z_2)=R_{21}^{(i)}(z_1/z_2)K_1^{(i)}(z_1)R_{12}^{(i)}(z_1z_2)K_2^{(i)}(z_2).`$ (2.16) Let $`V=v_0\mathrm{}v_{n1}`$ be the basic representation of $`U_q(sl_n)`$, and let $`V_z`$ be the evaluation representation of $`U_q(\widehat{sl_n})`$ in the homogeneous picture. Let $`V(\mathrm{\Lambda }_i)`$ be the irreducible highest weight module with the level 1 highest weight $`\mathrm{\Lambda }_i(i=0,\mathrm{},n1)`$. The type-I vertex operator $`\mathrm{\Phi }^{(i,i+1)}(z)`$ is an intertwining operator of $`U_q(\widehat{sl_n})`$ defined by $`\mathrm{\Phi }^{(i,i+1)}(z):V(\mathrm{\Lambda }_{i+1})V(\mathrm{\Lambda }_i)\widehat{}V_z,`$ (2.17) where the superscripts $`i,i+1`$ should be interpreted as elements in $`_n`$. Let us define the component of the vertex operators $`\mathrm{\Phi }_j^{(i,i+1)}(z)`$ as follows. $`\mathrm{\Phi }^{(i,i+1)}(z)|u={\displaystyle \underset{j=0}{\overset{n1}{}}}\mathrm{\Phi }_j^{(i,i+1)}(z)|uv_j,\mathrm{for}|uV(\mathrm{\Lambda }_{i+1}).`$ (2.18) The dual type-I vertex operator $`\mathrm{\Phi }^{(i+1,i)}(z)`$ is an intertwining operator of $`U_q(\widehat{sl_n})`$ defined by $`\mathrm{\Phi }^{(i+1,i)}(z):V(\mathrm{\Lambda }_i)V_z\widehat{V}(\mathrm{\Lambda }_{i+1}).`$ (2.19) Let us define the components of the dual vertex operators $`\mathrm{\Phi }_j^{(i+1,i)}(z)`$ as follows. $`\mathrm{\Phi }^{(i+1,i)}(z)(|uv_j)=\mathrm{\Phi }_j^{(i+1,i)}(z)|u,\mathrm{for}|uV(\mathrm{\Lambda }_i).`$ (2.20) Let us summarize here the properties of the vertex operators: Commutation relations The vertex operators satisfy the following commutation relation: $`\mathrm{\Phi }_{j_2}^{(i2,i1)}(z_2)\mathrm{\Phi }_{j_1}^{(i1,i)}(z_1)={\displaystyle \underset{\genfrac{}{}{0pt}{}{j_1^{},j_2^{}=0}{j_1^{}+j_2^{}=j_1+j_2}}{\overset{n1}{}}}R^{(i)VV}(z_1/z_2)_{j_1,j_2}^{j_1^{},j_2^{}}\mathrm{\Phi }_{j_1^{}}^{(i2,i1)}(z_1)\mathrm{\Phi }_{j_2^{}}^{(i1,i)}(z_2).`$ (2.21) Concerning other commutation relations (3.62), (3.64) and (3.76), see section 3. Normalizations We adopt the following normalizations: $`\mathrm{\Phi }^{(i,i+1)}(z)|\mathrm{\Lambda }_{i+1}=|\mathrm{\Lambda }_iv_i+\mathrm{},\mathrm{\Phi }^{(i+1,i)}(z)|\mathrm{\Lambda }_iv_i=|\mathrm{\Lambda }_{i+1}+\mathrm{},`$ (2.22) where $`|\mathrm{\Lambda }_i`$ is the highest weight vector of $`V(\mathrm{\Lambda }_i)`$. Invertibility They satisfy the following inversion relation: $`g_n\mathrm{\Phi }_j^{(i1,i)}(z)\mathrm{\Phi }_j^{(i,i1)}(z)=\mathrm{id},`$ (2.23) where $$g_n=\frac{(q^2;q^{2n})_{\mathrm{}}}{(q^{2n};q^{2n})_{\mathrm{}}}.$$ We define the normarized transfer matrix by $`T_B^{(i)}(z)=g_n{\displaystyle \underset{j=0}{\overset{n1}{}}}\mathrm{\Phi }_j^{(i,i1)}(z^1)K^{(i)}(z)_j^j\mathrm{\Phi }_j^{(i1,i)}(z),`$ (2.24) Let the space $`^{(i)}`$ be the span of vectors $`|p=_{k=1}^{\mathrm{}}v_{p(k)}`$, where $`p:/n`$ satisfies the asymptotic condition $`p(k)=k+i/n,\mathrm{for}k1.`$ (2.25) As usual, the transfer matrix (2.24) and the Hamiltonian (1.1) are related by $`{\displaystyle \frac{d}{dz}}T_B^{(i)}(z)|_{z=1}={\displaystyle \frac{2q}{1q^2}}_B+const,\mathrm{for}h={\displaystyle \frac{r+1}{r1}}\times {\displaystyle \frac{1q^2}{2q}}.`$ (2.26) Note that the left hand side act on the space $`V(\mathrm{\Lambda }_i)`$ while the right hand side acts on the space $`^{(i)}`$. Thus we can make the following identification: $`V(\mathrm{\Lambda }_i)^{(i)}.`$ (2.27) The boundary ground state and the dual boundary ground state are characterized by $`T_B^{(i)}(z)|i_B=|i_B,(i=0,\mathrm{},n1),`$ (2.28) and $`{}_{B}{}^{}i|T_B^{(i)}(z)={}_{B}{}^{}i|,(i=0,\mathrm{},n1).`$ (2.29) Using the inversion relation, the eigenvalue problems (2.28) and (2.29) are reduced to $`K^{(i)}(z)_j^j\mathrm{\Phi }_j^{(i1,i)}(z)|i_B=\mathrm{\Phi }_j^{(i1,i)}(z^1)|i_B,`$ (2.30) and $`K^{(i)}(z)_j^j{}_{B}{}^{}i|\mathrm{\Phi }_j^{(i,i1)}(z^1)={}_{B}{}^{}i|\mathrm{\Phi }_j^{(i,i1)}(z).`$ (2.31) The bosonizations of vertex operators are given in . The bosonic formulae of the boundary vacuum are given in . For readers’ convenience we summarize the bosonizations of vertex operators in Appendix A and the bosonic formula of the boundary vacuum in Apendix B. ## 3 Boundary quantum Knizhnik-Zamolodchikov equations The purpose of this section is to derive the $`q`$-difference equations for the correlation function of the higher rank XXZ spin chain with a boundary magnetic field. For $`U_q(\widehat{sl_2})`$ case , the said difference equations are based on the duality relation of vertex operators $$\mathrm{\Phi }_ϵ^{}(\zeta )=\mathrm{\Phi }_ϵ(q^1\zeta ),$$ in addition to (2.21), (2.30) and (2.31). For $`n>2`$ case, however, the dual vertex operator $`\mathrm{\Phi }_j^{}(z)`$ is written in terms of $`(n1)`$-st determinant of $`\mathrm{\Phi }_j(z)`$’s. Thus it is not convenient to use the duality relation for the present case. For $`n>2`$, we use the explicit formulae of the boundary states to derive the boundary quantum Knizhnik-Zamolodchikov equations. In this section we establish the following simple relations: $$\begin{array}{ccc}\hfill \mathrm{\Phi }_j^{(i+1,i)}(q^nz)|i_B& =& K^{(i)}(z)_j^j\mathrm{\Phi }_j^{(i+1,i)}(q^n/z)|i_B,(j=0,\mathrm{},n1),\hfill \\ \hfill {}_{B}{}^{}i|\mathrm{\Phi }_j^{(i,i+1)}(1/(q^nz))& =& K^{(i)}(z)_j^j{}_{B}{}^{}i|\mathrm{\Phi }_j^{(i,i+1)}(z/q^n),(j=0,\mathrm{},n1),\hfill \end{array}$$ (3.1) where the functions $`K^{(i)}(z)_j^j`$ are given by (3.5), (3.18) and (3.34). The relations (3.1) in addition to the commutation relations (2.21), (3.62), (3.64) and (3.76) imply the $`q`$-difference equations of the present model. ### 3.1 Boundary state In this subsection we use the symbols $`P^{}(z),Q^{}(z),R^{}(z),S^{}(z)`$, which are bosons defined in Appendix A. See Appendix A as for the definitions. Let us first consider consider $`\mathrm{`}\mathrm{`}(C1)0L=M=in1^{\prime \prime }`$-case. Let us show the following relation: $`\mathrm{\Phi }_j^{(i+1,i)}(q^nz)|i_B=K^{(i)}(z)_j^j\mathrm{\Phi }_j^{(i+1,i)}(q^n/z)|i_B,(j=0,\mathrm{},n1),`$ (3.2) where $`K^{(i)}(z)_j^j={\displaystyle \frac{\phi ^{(i)}(z)}{\phi ^{(i)}(1/z)}}\times \{\begin{array}{cc}z^2,& (0jL1=i1),\hfill \\ 1,& (i=Ljn1),\hfill \end{array}\mathrm{for}(C1),`$ (3.5) and $`\phi ^{(i)}(z)=z^{\delta _{i,0}}{\displaystyle \frac{(q^{4n}z^2;q^{4n})_{\mathrm{}}}{(q^{2n+2}z^2;q^{4n})_{\mathrm{}}}}.`$ (3.6) Multiply the both sides of (3.2) by $`k_j^{(i)}(z)\phi ^{(i)}(1/z)`$, where $`k_j^{(i)}(z)=\{\begin{array}{cc}z^1,& 0ji1,\\ 1,& ijn1.\end{array}`$ (3.9) Then the RHS of (3.2) is obtained from the LHS by changing $`z1/z`$. Bosonization formulae of $`P^{}(z),Q^{}(z)`$ and $`|i_B`$ imply the identity $`e^{Q^{}(q^nz)}|i_B={\displaystyle \frac{(q^{2n+2}z^2;q^{4n})_{\mathrm{}}}{(q^{4n}z^2;q^{4n})_{\mathrm{}}}}e^{P^{}(q^n/z)}|i_B.`$ (3.10) By using this identity we have $$k_0^i(z)\phi ^{(i)}(1/z)\mathrm{\Phi }_0^{(i+1,i)}(q^nz)|i_B=c_0^{}e^{P^{}(q^nz)+P^{}(q^n/z)}e^{\overline{\mathrm{\Lambda }}_1}|i_B,$$ (3.11) where $`c_0^{}`$ is some constant. The relation (3.2) with $`j=0`$ follows form the fact that RHS of (3.11) is symmetric under $`z1/z`$. Invoking the bosonization of the dual vertex operators, we also have for $`j>0`$ as follows: $$\begin{array}{cc}& k_j^{(i)}(z)\phi ^{(i)}(1/z)\mathrm{\Phi }_j^{(i+1,i)}(q^nz)|i_B\hfill \\ =& c_j^{}\frac{dw_1}{w_1}\mathrm{}\frac{dw_j}{w_j}k_j^{(i)}(z)\mathrm{Int}(z,w_1,w_2,\mathrm{},w_j)e^{P^{}(q^nz)+P^{}(q^n/z)}\hfill \\ \times & e^{R_1^{}(q^{n+1}w_1)+R_1^{}(q^{n+1}/w_1)+\mathrm{}+R_j^{}(q^{n+1}w_j)+R_j^{}(q^{n+1}/w_j)}e^{\overline{\mathrm{\Lambda }}_1}|i_B,\hfill \end{array}$$ (3.12) where $`c_j^{}`$’s are some constants. Here we set the integrand: $`\mathrm{Int}(w_0,w_1,\mathrm{},w_j)={\displaystyle \frac{w_j_{k=1}^j\left\{(1w_k^2)w_k^{\delta _{k,i}}(1qw_{k1}w_k)\right\}}{_{k=1}^jD(w_{k1},w_k)}},`$ (3.13) where $$D(w_1,w_2)=(1qw_1w_2)(1qw_1/w_2)(1qw_2/w_1)(1q/(w_1w_2)).$$ Thus the relation (3.2) with $`j>0`$ follows from the identities $`{\displaystyle \underset{ϵ_1=\pm ,\mathrm{},ϵ_j=\pm }{}}\left\{k_j^{(i)}(z^1)\mathrm{Int}(z,w_1^{ϵ_1},\mathrm{},w_j^{ϵ_j})k_j^{(i)}(z)\mathrm{Int}(z^1,w_1^{ϵ_1},\mathrm{},w_j^{ϵ_j})\right\}=0.`$ (3.14) Let us consider $`\mathrm{`}\mathrm{`}(C2)0L=i<Mn1^{\prime \prime }`$-case. From the same arguments as for $`(C1)`$, we have $`K^{(i)}(z)_j^j={\displaystyle \frac{\phi ^{(i)}(z)}{\phi ^{(i)}(1/z)}}\times \{\begin{array}{cc}z^2,& (0jL1),\\ \frac{1q^{n2M+2L}rz}{1q^{n2M+2L}rz^1},& (LjM1),\\ 1,& (Mjn1),\end{array}\mathrm{for}(C2),`$ (3.18) where we have set $`\phi ^{(i)}(z)=z^{\delta _{i,0}}{\displaystyle \frac{(q^{4n}z^2;q^{4n})_{\mathrm{}}(rq^nz;q^{2n})_{\mathrm{}}}{(q^{2n+2}z^2;q^{4n})_{\mathrm{}}(rq^{n+2L2M}z;q^{2n})_{\mathrm{}}}}.`$ (3.19) In this case the following relations are useful: $`e^{Q^{}(q^nz)}|0_B={\displaystyle \frac{(rq^{3n2M}z^1;q^{2n})_{\mathrm{}}(q^{2n+2}z^2;q^{4n})_{\mathrm{}}}{(rq^nz^1;q^{2n})_{\mathrm{}}(q^{4n}z^2;q^{4n})_{\mathrm{}}}}e^{P^{}(q^n/z)}|0_B,`$ (3.20) $`e^{Q^{}(q^nz)}|i_B={\displaystyle \frac{(rq^{n+2L2M}z^1;q^{2n})_{\mathrm{}}(q^{2n+2}z^2;q^{4n})_{\mathrm{}}}{(rq^nz^1;q^{2n})_{\mathrm{}}(q^{4n}z^2;q^{4n})_{\mathrm{}}}}e^{P^{}(q^n/z)}|i_B,(i1),`$ (3.21) and $`e^{S_j^{}(w)}|i_B=g_j^{(i)}(w)e^{R_j^{}(q^{2(n+1)}/w)}|i_B,`$ (3.22) where $`g_j^{(0)}(q^{n+1}w)=\{\begin{array}{cc}(11/w^2)(1q^{n+2ML}/(rw)),& j=L,\\ (11/w^2)(1q^{nM}r/w),& j=M,\\ (11/w^2),& jL,M,\end{array}`$ (3.26) and $`g_j^{(i)}(q^{n+1}w)=\{\begin{array}{cc}\frac{\left(11/w^2\right)}{\left(1rq^{n2M+L}/w\right)},& j=L,\\ (11/w^2)(1q^{nM}r/w),& j=M,\\ (11/w^2),& jL,M,\end{array}(i1).`$ (3.30) Let us consider $`\mathrm{`}\mathrm{`}(C3)0L<M=in1^{\prime \prime }`$-case. Repeating the same procedure as in $`(C1)`$, we have $`K^{(i)}(z)_j^j={\displaystyle \frac{\phi ^{(i)}(z)}{\phi ^{(i)}(1/z)}}\times \{\begin{array}{cc}z^2,& (0jL1),\\ \frac{1q^{n+2M2L}r^1z}{1q^{n+2M2L}r^1z^1},& (LjM1),\\ 1,& (Mjn1),\end{array}\mathrm{for}(C3),`$ (3.34) where we set $`\phi ^{(i)}(z)={\displaystyle \frac{(q^{4n}z^2;q^{4n})_{\mathrm{}}(r^1q^nz;q^{2n})_{\mathrm{}}}{(q^{2n+2}z^2;q^{4n})_{\mathrm{}}(r^1q^{n+2M2L}z;q^{2n})_{\mathrm{}}}}.`$ (3.35) In this case the following relations are useful: $`e^{Q^{}(q^nz)}|i_B={\displaystyle \frac{(r^1q^{2Mn}z^1;q^{2n})_{\mathrm{}}(q^{2n+2}z^2;q^{4n})_{\mathrm{}}}{(r^1q^nz^1;q^{2n})_{\mathrm{}}(q^{4n}z^2;q^{4n})_{\mathrm{}}}}e^{P^{}(q^n/z)}|i_B,(L=0),`$ (3.36) $`e^{Q^{}(q^nz)}|i_B={\displaystyle \frac{(r^1q^{2M2L+n}z^1;q^{2n})_{\mathrm{}}(q^{2n+2}z^2;q^{4n})_{\mathrm{}}}{(r^1q^nz^1;q^{2n})_{\mathrm{}}(q^{4n}z^2;q^{4n})_{\mathrm{}}}}e^{P^{}(q^n/z)}|i_B,(L1),`$ (3.37) and $`e^{S_j^{}(w)}|i_B=g_j^{(i)}(w)e^{R_j^{}(q^{2(n+1)}/w)}|i_B,`$ (3.38) where $`g_j^{(i)}(q^{n+1}w)=\{\begin{array}{cc}(11/w^2)(1q^{n+2ML}/(rw)),& j=L,\\ \frac{\left(11/w^2\right)}{\left(1q^{Mn}/\left(rw\right)\right)},& j=M,\\ (11/w^2),& jL,M,\end{array}`$ (3.42) ### 3.2 Dual boundary state From the same arguments as for the boundary state case, we can show the following relation: $`{}_{B}{}^{}i|\mathrm{\Phi }_j^{(i,i+1)}(1/(q^nz))=K^{(i)}(z)_j^j{}_{B}{}^{}i|\mathrm{\Phi }_j^{(i,i+1)}(z/q^n).`$ (3.43) For each case the following relations are useful: $`(C1)`$-case: $`0L=M=in1`$ $$\begin{array}{ccc}\hfill {}_{B}{}^{}i|e^{P(z/q^n)}& =& \frac{(q^{2n+2}z^2;q^{4n})_{\mathrm{}}}{(q^{4n}z^2;q^{4n})_{\mathrm{}}}{}_{B}{}^{}i|e^{Q(1/(q^nz))},\hfill \\ \hfill {}_{B}{}^{}i|e^{S_j^{}(w)}& =& g_j^{(i)}(w){}_{B}{}^{}i|e^{R_j^{}(q^2/w)},\hfill \end{array}$$ (3.44) where $`g_j^{(i)}(qw)=(1w^2).`$ (3.45) $`(C2)`$-case: $`0L=i<Mn1`$ $$\begin{array}{ccc}\hfill {}_{B}{}^{}i|e^{P(z/q^n)}& =& \frac{(q^{2n+2}z^2;q^{4n})_{\mathrm{}}(rq^{n+2L2M}z;q^{2n})_{\mathrm{}}}{(q^{4n}z^2;q^{4n})_{\mathrm{}}(rq^nz;q^{2n})_{\mathrm{}}}{}_{B}{}^{}i|e^{Q(1/(q^nz))},\hfill \\ \hfill {}_{B}{}^{}i|e^{S_j^{}(w)}& =& g_j^{(i)}(w){}_{B}{}^{}i|e^{R_j^{}(q^2/w)},\hfill \end{array}$$ (3.46) where $`g_j^{(0)}(qw)=\{\begin{array}{cc}\frac{\left(1w^2\right)}{\left(1q^Lw/r\right)},& j=L,\\ \frac{\left(1w^2\right)}{\left(1q^{2LM}rw\right)},& j=M,\\ (1w^2),& jL,M,\end{array}`$ (3.50) and $`g_j^{(i)}(qw)=\{\begin{array}{cc}(1w^2)(1q^Lrw),& j=L,\\ \frac{\left(1w^2\right)}{\left(1q^{2LM}rw\right)},& j=M,\\ (1w^2),& jL,M,\end{array}(i1).`$ (3.54) $`(C3)`$-case: $`0L<M=in1`$ $$\begin{array}{ccc}\hfill {}_{B}{}^{}i|e^{P(z/q^n)}& =& \frac{(q^{2n+2}z^2;q^{4n})_{\mathrm{}}(q^{n+2M2L}r^1z;q^{2n})_{\mathrm{}}}{(q^{4n}z^2;q^{4n})_{\mathrm{}}(q^nr^1z;q^{2n})_{\mathrm{}}}{}_{B}{}^{}i|e^{Q(1/(q^nz))}\hfill \\ \hfill {}_{B}{}^{}i|e^{S_j^{}(w)}& =& g_j^{(i)}(w){}_{B}{}^{}i|e^{R_j^{}(q^2/w)},\hfill \end{array}$$ (3.55) where $`g_j^{(i)}(qw)=\{\begin{array}{cc}\frac{\left(1w^2\right)}{\left(1q^Lw/r\right)},& j=L,\\ (1w^2)(1q^{M2L}w/r),& j=M,\\ (1w^2),& jL,M,\end{array}`$ (3.59) ### 3.3 Correlation functions and difference equations Let us consider the $`2N`$-point correlation function: $`G^{(i)}(z_1,\mathrm{},z_N|z_{N+1},\mathrm{},z_{2N})`$ $`=`$ $`{\displaystyle \underset{j_1=0}{\overset{n1}{}}}\mathrm{}{\displaystyle \underset{j_N=0}{\overset{n1}{}}}{\displaystyle \underset{j_{N+1}=0}{\overset{n1}{}}}\mathrm{}{\displaystyle \underset{j_{2N}=0}{\overset{n1}{}}}v_{j_1}^{}\mathrm{}v_{j_N}^{}v_{j_{N+1}}\mathrm{}v_{j_{2N}}`$ $`\times `$ $`G^{(i)}(z_1,\mathrm{},z_N|z_{N+1},\mathrm{},z_{2N})_{j_{N+1}\mathrm{}j_{2N}}^{j_1\mathrm{}j_N},`$ where $`G^{(i)}(z_1,\mathrm{},z_N|z_{N+1},\mathrm{},z_{2N})_{j_{N+1}\mathrm{}j_{2N}}^{j_1\mathrm{}j_N}`$ (3.61) $`=`$ $`{}_{B}{}^{}i|\mathrm{\Phi }_{j_1}^{(i,i1)}(z_1)\mathrm{}\mathrm{\Phi }_{j_N}^{(iN+1,iN)}(z_N)\mathrm{\Phi }_{j_{N+1}}^{(iN,iN+1)}(z_{N+1})\mathrm{}\mathrm{\Phi }_{j_{2N}}^{(i1,i)}(z_{2N})|i_{B}^{}.`$ In order to derive $`q`$-difference equations, we use the commutation relations of vertex operators and the action formulae of vertex operators to the boundary state. In what follows we assume that $`K^{(i)}(z)`$ is a diagonal matrix whose diagonal elements are given by (3.5), (3.18) and (3.34). The commutation relations between vertex operators of different types are given as follows : $`\mathrm{\Phi }_j^{(i,i+1)}(z_2)\mathrm{\Phi }_j^{(i+1,i)}(z_1)`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{n1}{}}}R^{(i)V^{}V}(z_1/z_2)_{j,j}^{k,k}\mathrm{\Phi }_k^{(i,i1)}(z_1)\mathrm{\Phi }_k^{(i1,i)}(z_2),`$ (3.62) $`\mathrm{\Phi }_k^{(i,i+1)}(z_2)\mathrm{\Phi }_j^{(i+1,i)}(z_1)`$ $`=`$ $`r^{(i)V^{}V}(z_1/z_2)\mathrm{\Phi }_j^{(i,i1)}(z_1)\mathrm{\Phi }_k^{(i1,i)}(z_2),(jk),`$ (3.63) and $`\mathrm{\Phi }_j^{(i,i1)}(z_2)\mathrm{\Phi }_j^{(i1,i)}(z_1)`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{n1}{}}}R^{(i)VV^{}}(z_1/z_2)_{j,j}^{k,k}\mathrm{\Phi }_k^{(i,i+1)}(z_1)\mathrm{\Phi }_k^{(i+1,i)}(z_2),`$ (3.64) $`\mathrm{\Phi }_k^{(i,i1)}(z_2)\mathrm{\Phi }_j^{(i1,i)}(z_1)`$ $`=`$ $`r^{(i)VV^{}}(z_1/z_2)\mathrm{\Phi }_j^{(i,i+1)}(z_1)\mathrm{\Phi }_k^{(i+1,i)}(z_2),(jk).`$ (3.65) Here the nonzero components are $`R^{(i)V^{}V}(z)_{j,j}^{k,k}`$ $`=`$ $`r^{(i)V^{}V}(z)\times \{\begin{array}{cc}b(z),& j=k,\\ c(z),& j>k,\\ zc(z),& j<k,\end{array}`$ (3.69) and $`R^{(i)VV^{}}(z)_{j,j}^{k,k}`$ $`=`$ $`r^{(i)VV^{}}(z)\times \{\begin{array}{cc}b(q^{2n}z),& j=k,\\ q^{2n}zc(q^{2n}z)q^{2(kj)},& j>k,\\ c(q^{2n}z)q^{2(kj)},& j<k,\end{array}`$ (3.73) where $`r^{(i)V^{}V}(z)=qz^{\delta _{i,0}}{\displaystyle \frac{(z;q^{2n})_{\mathrm{}}(q^{2n+2}z^1;q^{2n})_{\mathrm{}}}{(q^2z;q^{2n})_{\mathrm{}}(q^{2n}z^1;q^{2n})_{\mathrm{}}}},`$ (3.74) $`r^{(i)VV^{}}(z)=q^1z^{\delta _{i,0}}{\displaystyle \frac{(q^{2n}z;q^{2n})_{\mathrm{}}(q^2z^1;q^{2n})_{\mathrm{}}}{(q^{2n+2}z;q^{2n})_{\mathrm{}}(z^1;q^{2n})_{\mathrm{}}}}.`$ (3.75) The commutation relations between the dual vertex operators are given as $`\mathrm{\Phi }_{j_2}^{(i+2,i+1)}(z_2)\mathrm{\Phi }_{j_1}^{(i+1,i)}(z_1)={\displaystyle \underset{\genfrac{}{}{0pt}{}{k_1,k_2=0}{k_1+k_2=j_1+j_2}}{\overset{n1}{}}}R^{(i)V^{}V^{}}(z_1/z_2)_{j_1,j_2}^{k_1,k_2}\mathrm{\Phi }_{k_1}^{(i+2,i+1)}(z_1)\mathrm{\Phi }_{k_2}^{(i+1,i)}(z_2).`$ (3.76) Here the nonzero components are $`R^{(i)V^{}V^{}}(z_1/z_2)_{j_1,j_2}^{k_1,k_2}`$ $`=`$ $`r^{(i)V^{}V^{}}(z)\times \{\begin{array}{cc}1,& j_1=j_2=k_1=k_2,\\ b(q^2z),& j_1=k_1j_2=k_2,\\ qzc(q^2),& j_1=k_2<j_2=k_1,\\ qc(q^2),& j_1=k_2>j_2=k_1,\end{array}`$ (3.81) where $`r^{(i)V^{}V^{}}(z)=r^{(i)VV}(z).`$ (3.82) Now we are in a position to derive boundary quantum Knizhnik-Zamolodchikov equations, which is a version of Cherednik’s equation . From the commutation relations (2.21), (3.62), (3.64), (3.76) and the boundary state ideitities (3.1) we obtain the following $`q`$-difference equations: $`G^{(i)}(z_1\mathrm{}q^{2n}z_j\mathrm{}z_N|z_{N+1}\mathrm{}z_{2N})`$ $`=`$ $`R_{jj1}^{V^{}V^{}}(z_j/(q^{2n}z_{j1}))\mathrm{}R_{j1}^{V^{}V^{}}(z_j/(q^{2n}z_1))K_j^{(i)}(z_j/q^{2n})`$ $`\times `$ $`R_{1j}^{V^{}V^{}}(z_1z_j/q^{2n})\mathrm{}R_{j1j}^{V^{}V^{}}(z_{j1}z_j/q^{2n})R_{j+1j}^{V^{}V^{}}(z_{j+1}z_j/q^{2n})\mathrm{}R_{Nj}^{V^{}V^{}}(z_Nz_j/q^{2n})`$ $`\times `$ $`R_{N+1j}^{VV^{}}(z_{N+1}z_j/q^{2n})\mathrm{}R_{2Nj}^{VV^{}}(z_{2N}z_j/q^{2n})K_j^{(i)}(q^n/z_j)R_{j2N}^{V^{}V}(z_j/z_{2N})\mathrm{}R_{jN}^{V^{}V}(z_j/z_N)`$ $`\times `$ $`R_{jN+1}^{V^{}V^{}}(z_j/z_{N+1})\mathrm{}R_{jj+1}^{V^{}V^{}}(z_j/z_{j+1})G^{(i)}(z_1\mathrm{}z_N|z_{N+1}\mathrm{}z_{2N}),`$ and $`G^{(i)}(z_1\mathrm{}z_N|z_{N+1}\mathrm{}q^{2n}z_j\mathrm{}z_{2N})`$ $`=`$ $`R_{jj1}^{VV}(z_j/(q^{2n}z_{j1}))\mathrm{}R_{jN+1}^{VV}(z_j/(q^{2n}z_{N+1}))R_{jN}^{VV^{}}(z_j/(q^{2n}z_N))\mathrm{}R_{j1}^{VV^{}}(z_j/(q^{2n}z_1))`$ $`\times `$ $`K_j^{(i)}(q^n/z_j)R_{1j}^{V^{}V}(z_1z_j)\mathrm{}R_{Nj}^{V^{}V}(z_Nz_j)R_{N+1j}^{VV}(z_{N+1}z_j)\mathrm{}R_{j1j}^{VV}(z_{j1}z_j)`$ $`\times `$ $`R_{j+1j}^{VV}(z_{j+1}z_j)\mathrm{}R_{2Nj}^{VV}(z_{2N}z_j)K_j^{(i)}(z_j)`$ $`\times `$ $`R_{j2N}^{VV}(z_j/z_{2N})\mathrm{}R_{jj+1}^{VV}(z_j/z_{j+1})G^{(i)}(z_1\mathrm{}z_N|z_{N+1}\mathrm{}z_{2N}).`$ Here the coefficient matrces are given by (2.6),(2.11),(3.5), (3.18),(3.34),(3.69), (3.73) and (3.81). For $`N=1`$, the equations (3.3) and (3.3) are as follows: $`G^{(i)}(q^{2n}z_1|z_2)`$ $`=`$ $`K_1^{(i)}(z_1/q^{2n})R_{21}^{VV^{}}(z_2z_1/q^{2n})K_1^{(i)}(q^n/z_1)R_{12}^{V^{}V}(z_1/z_2)G^{(i)}(z_1|z_2),`$ (3.85) $`G^{(i)}(z_1|q^{2n}z_2)`$ $`=`$ $`R_{21}^{VV^{}}(z_2/(q^{2n}z_1))K_2^{(i)}(q^n/z_2)R_{12}^{V^{}V}(z_1z_2)K_2^{(i)}(z_2)G^{(i)}(z_1|z_2).`$ (3.86) ## 4 Two point functions The purpose of this section is to perform explicit calculations of two point functions for free boundary condition. In what follows we consider the case $`i=L=M=0`$ and $`N=1`$. In this case the boundary K-matrices $`K^{(0)}(z)`$ and $`K^{(0)}(z)`$ become scalar matrices, i.e. $$K^{(0)}(z)=\frac{\phi ^{(0)}(z)}{\phi ^{(0)}(z^1)}\times \mathrm{id},K^{(0)}(z)=\frac{\phi ^{(0)}(z)}{\phi ^{(0)}(z^1)}\times \mathrm{id}.$$ The boundary quantum Knizhnik-Zamolodchikov equations thus reduces to: $`G^{(0)}(q^{2n}z_1|z_2)`$ $`=`$ $`{\displaystyle \frac{\phi ^{(0)}(z_1/q^{2n})}{\phi ^{(0)}(q^{2n}/z_1)}}{\displaystyle \frac{\phi ^{(0)}(q^n/z_1)}{\phi ^{(0)}(z_1/q^n)}}R_{21}^{VV^{}}(z_2z_1/q^{2n})R_{12}^{V^{}V}(z_1/z_2)G^{(0)}(z_1|z_2),`$ (4.1) $`G^{(0)}(z_1|q^{2n}z_2)`$ $`=`$ $`{\displaystyle \frac{\phi ^{(0)}(z_2)}{\phi ^{(0)}(1/z_2)}}{\displaystyle \frac{\phi ^{(0)}(q^n/z_2)}{\phi ^{(0)}(z_2/q^n)}}R_{21}^{VV^{}}(z_2/(q^{2n}z_1))R_{12}^{V^{}V}(z_1z_2)G^{(0)}(z_1|z_2).`$ (4.2) Let us now introduce the scalar function $`r(z_1|z_2)`$ by $`r(z_1|z_2)=A(z_1)A(q^nz_2)B(z_1z_2)B(z_1/z_2),`$ (4.3) where $`A(z)`$ $`=`$ $`{\displaystyle \frac{(q^{2n+2}z^2;q^{2n},q^{4n})_{\mathrm{}}(q^{4n+2}/z^2;q^{2n},q^{4n})_{\mathrm{}}}{(q^{4n}z^2;q^{2n},q^{4n})_{\mathrm{}}(q^{6n}/z^2;q^{2n},q^{4n})_{\mathrm{}}}},`$ (4.4) $`B(z)`$ $`=`$ $`{\displaystyle \frac{(q^{2n}z;q^{2n},q^{2n})_{\mathrm{}}(q^{2n}/z;q^{2n},q^{2n})_{\mathrm{}}}{(q^{2n+2}z;q^{2n},q^{2n})_{\mathrm{}}(q^{2n+2}/z;q^{2n},q^{2n})_{\mathrm{}}}}.`$ (4.5) Note that the function $`r(z_1|z_2)`$ satisfies $`r(q^{2n}z_1|z_2)=q^{2n}z_1^2r^{VV^{}}(z_1z_2/q^{2n})r^{V^{}V}(z_1/z_2){\displaystyle \frac{\phi ^{(0)}(z_1/q^{2n})}{\phi ^{(0)}(q^{2n}/z_1)}}{\displaystyle \frac{\phi ^{(0)}(q^n/z_1)}{\phi ^{(0)}(z_1/q^n)}}\times r(z_1|z_2),`$ (4.6) $`r(z_1|q^{2n}z_2)=q^{2n}z_2^2r^{VV^{}}(z_2/(q^{2n}z_1))r^{V^{}V}(z_1z_2){\displaystyle \frac{\phi ^{(0)}(z_2)}{\phi ^{(0)}(1/z_2)}}{\displaystyle \frac{\phi ^{(0)}(q^n/z_2)}{\phi ^{(0)}(z_2/q^n)}}\times r(z_1|z_2).`$ (4.7) Let $`\overline{G}(z_1|z_2)_j`$ be the auxiliary function defined by $`\overline{G}(z_1|z_2)_j=r(z_1|z_2)^1G^{(0)}(z_1|z_2)_j^j.`$ (4.8) Then we have $`{\displaystyle \underset{j=0}{\overset{n1}{}}}\overline{G}(q^{2n}z_1|z_2)_j`$ $`=`$ $`{\displaystyle \frac{1q^{2n}/(z_1z_2)}{1z_1z_2}}{\displaystyle \frac{1q^{2n}z_2/z_1}{1z_1/z_2}}{\displaystyle \underset{j=0}{\overset{n1}{}}}\overline{G}(z_1|z_2)_j,`$ (4.9) $`{\displaystyle \underset{j=0}{\overset{n1}{}}}\overline{G}(z_1|q^{2n}z_2)_j`$ $`=`$ $`{\displaystyle \frac{1q^{2n}/(z_1z_2)}{1z_1z_2}}{\displaystyle \frac{1q^{2n}z_1/z_2}{1z_2/z_1}}{\displaystyle \underset{j=0}{\overset{n1}{}}}\overline{G}(z_1|z_2)_j.`$ (4.10) From these we obtain $`{\displaystyle \underset{j=0}{\overset{n1}{}}}{}_{B}{}^{}0|\mathrm{\Phi }_j^{(0,1)}(z_1)\mathrm{\Phi }_j^{(1,0)}(z_2)|0_{B}^{}=c_0r(z_1|z_2)\times `$ $`\times `$ $`\left\{(q^{2n}z_1/z_2;q^{2n})_{\mathrm{}}(q^{2n}z_2/z_1;q^{2n})_{\mathrm{}}(q^{2n}z_1z_2;q^{2n})_{\mathrm{}}(q^{2n}/(z_1z_2);q^{2n})_{\mathrm{}}\right\}^1,`$ where $`c_0`$ is a constant independent of spectral parameters $`z_1,z_2`$. By specializing the spectral parameters $`z_1=z_2`$, we have $`c_0=g_n{}_{}{}^{1}\times {}_{B}{}^{}0|0_{B}^{}\times \left\{{\displaystyle \frac{(q^{2n+2};q^{2n},q^{2n})_{\mathrm{}}}{(q^{4n};q^{2n},q^{2n})_{\mathrm{}}}}\right\}^2,`$ (4.12) where the norm $`{}_{B}{}^{}0|0_{B}^{}`$ is given as follows $`{}_{B}{}^{}0|0_{B}^{}={\displaystyle \frac{1}{\sqrt{(q^{4n};q^{4n})_{\mathrm{}}}}}{\displaystyle \underset{j=1}{\overset{n1}{}}}\left\{{\displaystyle \frac{\sqrt{(q^{4n+22j};q^{4n})_{\mathrm{}}(q^{4n22j};q^{4n})_{\mathrm{}}}}{(q^{4n2j};q^{4n})_{\mathrm{}}}}\right\}^{j(nj)}.`$ Let $`\omega `$ satisfy $`\omega ^n=1`$ and $`\omega 1`$. Then we have $`{\displaystyle \underset{j=0}{\overset{n1}{}}}(q^2\omega )^j\overline{G}(q^{2n}z_1|z_2)_j`$ $`=`$ $`q^{2n}z_1^2{\displaystyle \underset{j=0}{\overset{n1}{}}}(q^2\omega )^j\overline{G}(z_1|z_2)_j,`$ (4.13) $`{\displaystyle \underset{j=0}{\overset{n1}{}}}(q^2\omega )^j\overline{G}(z_1|q^{2n}z_2)_j`$ $`=`$ $`q^{2n}z_2^2{\displaystyle \underset{j=0}{\overset{n1}{}}}(q^2\omega )^j\overline{G}(z_1|z_2)_j.`$ (4.14) From these we obtain $`{\displaystyle \underset{j=0}{\overset{n1}{}}}(q^2\omega ^k)^j{}_{B}{}^{}0|\mathrm{\Phi }_j^{(0,1)}(z_1)\mathrm{\Phi }_j^{(1,0)}(z_2)|0_{B}^{}`$ (4.15) $`=`$ $`c_kr(z_1|z_2)\times \{(q^{2n}z_1^2;q^{4n})_{\mathrm{}}(q^{2n}/z_1^2;q^{4n})_{\mathrm{}}(q^{2n}z_2^2;q^{4n})_{\mathrm{}}(q^{2n}/z_2^2;q^{4n})_{\mathrm{}}\}^1.`$ Here $`c_k`$ are constants independent of spectral parameters $`z_1,z_2`$. Acknoeledgements. We wish to thank Prof. A. Kuniba for his interest to this work. TK was partly supported by Grant-in-Aid for Encouragements for Young Scientists (A) from Japan Society for the Promotion of Science. (11740099) ## Appendix A Bosonization of vertex operators in $`U_q(\widehat{sl_n})`$ For readers’ convenience, we summarize the results of bosonizations of the vertex operators . Let $`[\overline{P}]`$ be the $``$-algebra generated by the symbols $`\{e^{\alpha _2},\mathrm{},e^{\alpha _{n1}},e^{\overline{\mathrm{\Lambda }}_{n1}}\}`$ which satisfy the following defining relations: $`e^{\alpha _i}e^{\alpha _j}=(1)^{(\alpha _i|\alpha _j)}e^{\alpha _j}e^{\alpha _i},(2i,jn1),`$ $`e^{\alpha _i}e^{\overline{\mathrm{\Lambda }}_{n1}}=(1)^{\delta _{i,n1}}e^{\overline{\mathrm{\Lambda }}_{n1}}e^{\alpha _i},(2in1).`$ For $`\alpha =m_2\alpha _2+\mathrm{}+m_{n1}\alpha _{n1}+m_n\overline{\mathrm{\Lambda }}_{n1}`$, we denote $`e^{m_2\alpha _2}\mathrm{}e^{m_{n1}\alpha _{n1}}e^{m_n\overline{\mathrm{\Lambda }}_{n1}}`$ by $`e^\alpha `$. Let $`((\alpha _s|\alpha _t))_{1s,tn1}`$ stand for the A-type Catran matrix whose matrix element $`(\alpha _s|\alpha _t)`$ is an integer. Let $`[\overline{Q}]`$ be the $``$-subalgebra of $`[\overline{P}]`$ generated by the symbols $`\{e^{\alpha _1},\mathrm{},e^{\alpha _{n1}}\}`$ which satisfy the following defining relations: $`e^{\alpha _i}e^{\alpha _j}=(1)^{(\alpha _i|\alpha _j)}e^{\alpha _j}e^{\alpha _i},(1i,jn1).`$ Note that $`\alpha _1={\displaystyle \underset{r=2}{\overset{n1}{}}}r\alpha _r+n\overline{\mathrm{\Lambda }}_{n1},\overline{\mathrm{\Lambda }}_i={\displaystyle \underset{r=i+1}{\overset{n1}{}}}(ri)\alpha _r+(ni)\overline{\mathrm{\Lambda }}_{n1}.`$ Let us consider the $``$-algebra generated by the bosons $`a_s(k)`$ $`(s\{1,\mathrm{},n1\},k)`$ which satisfy the following defining relations: $$[a_s(k),a_t(l)]=\delta _{k+l,0}\frac{[(\alpha _s|\alpha _t)k][k]}{k}.$$ The highset weight module $`V(\mathrm{\Lambda }_i)`$ is realized as $$V(\mathrm{\Lambda }_i)=[a_s(k),(s\{1,\mathrm{},n1\},k0)][\overline{Q}]e^{\overline{\mathrm{\Lambda }}_i}.$$ We consider $`[\overline{Q}]e^{\overline{\mathrm{\Lambda }}_i}`$ as a subspace of $`[\overline{P}]`$. Here the actions of the operators $`a_s(k),_\alpha ,e^\alpha `$ on $`V(\mathrm{\Lambda }_i)`$ are defined as follows: $`a_s(k)fe^\beta =\{\begin{array}{cc}a_s(k)fe^\beta ,& (k<0),\\ [a_s(k),f]e^\beta ,& (k>0),\end{array}`$ $`_\alpha fe^\beta `$ $`=`$ $`(\alpha |\beta )fe^\beta .`$ $`e^\alpha fe^\beta `$ $`=`$ $`fe^\alpha e^\beta .`$ The inner product is explicitly given as follows: $`(\alpha _i|\overline{\mathrm{\Lambda }}_j)=\delta _{i,j},(\overline{\mathrm{\Lambda }}_i|\overline{\mathrm{\Lambda }}_j)={\displaystyle \frac{i(nj)}{n}},(1ijn1).`$ $`\mathrm{\Phi }_{n1}^{(i,i+1)}(z)`$ $`=`$ $`e^{P(z)}e^{Q(z)}e^{\overline{\mathrm{\Lambda }}_{n1}}(q^{n+1}z)^{_{\overline{\mathrm{\Lambda }}_{n1}}+\frac{ni1}{n}}(1)^{(_{\overline{\mathrm{\Lambda }}_1}\frac{ni1}{n})(n1)+\frac{1}{2}(ni)(ni1)},`$ $`\mathrm{\Phi }_0^{(i+1,i)}(z)`$ $`=`$ $`e^{P^{}(z)}e^{Q^{}(z)}e^{\overline{\mathrm{\Lambda }}_1}((1)^{n1}qz)^{_{\overline{\mathrm{\Lambda }}_1}+\frac{i}{n}}q^i(1)^{in+\frac{1}{2}i(i+1)},`$ $`\mathrm{\Phi }_j^{(i1i)}(z)`$ $`=`$ $`c_j{\displaystyle \mathrm{}_{C_j}\frac{dw_{j+1}}{2\pi iw_{j+1}}\mathrm{}\frac{dw_{n1}}{2\pi iw_{n1}}\frac{w_{j+1}}{z}\frac{1}{(1qw_{n1}/z)(1qz/w_{n1})}}`$ (A.2) $`\times `$ $`{\displaystyle \frac{1}{(1qw_{n1}/w_{n2})(1qw_{n2}/w_{n1})\mathrm{}(1qw_{j+2}/w_{j+1})(1qw_{j+1}/w_{j+2})}}`$ $`\times `$ $`:\mathrm{\Phi }_{n1}^{(i1i)}(z)X_{n1}^{}(q^{n+1}w_{n1})\mathrm{}X_{j+1}^{}(q^{n+1}w_{j+1}):,`$ and $`\mathrm{\Phi }_j^{(ii+1)}(z)`$ $`=`$ $`c_j^{}{\displaystyle \mathrm{}_{C_j^{}}\frac{dw_1}{2\pi iw_1}\mathrm{}\frac{dw_j}{2\pi iw_j}\frac{w_j}{z}\frac{1}{\left(1qz/w_1\right)\left(1qw_1/z\right)}}`$ (A.3) $`\times `$ $`{\displaystyle \frac{1}{(1qw_1/w_2)(1qw_2/w_1)\mathrm{}(1qw_{j1}/w_j)(1qw_j/w_{j1})}}`$ $`\times `$ $`:\mathrm{\Phi }_0^{(ii+1)}(z)X_1^{}(qw_1)\mathrm{}X_j^{}(qw_j):,`$ where $`c_j,c_j^{}`$ are appropriate constants. The contours $`C_j,C_j^{}`$ encircle $`w_l=0`$ anti-clockwise in such a way that $`C_j:`$ $`|q|<|w_{n1}/z|<|q^1|,|q|<|w_l/w_{l+1}|<|q^1|,(l=j+1,\mathrm{},n2),`$ $`C_j^{}:`$ $`|q|<|w_1/z|<|q^1|,|q|<|w_{l+1}/w_l|<|q^1|,(l=1,\mathrm{},j1).`$ Here we have used $`X_j^{}(w)=e^{R_j^{}(w)}e^{S_j^{}(w)}e^{\alpha _j}w^{_{\alpha _j}},`$ $`P(z)={\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}a_{n1}^{}(k)q^{\frac{2n+3}{2}k}z^k,`$ $`Q(z)={\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}a_{n1}^{}(k)q^{\frac{2n+1}{2}k}z^k,`$ $`P^{}(z)={\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}a_1^{}(k)q^{\frac{3}{2}k}z^k,`$ $`Q^{}(z)={\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}a_1^{}(k)q^{\frac{1}{2}k}z^k,`$ $`R_j^{}(w)={\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{a_j(k)}{[k]}}q^{\frac{k}{2}}w^k,`$ $`S_j^{}(w)={\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{a_j(k)}{[k]}}q^{\frac{k}{2}}w^k,`$ $`a_{n1}^{}(k)={\displaystyle \underset{l=1}{\overset{n1}{}}}{\displaystyle \frac{[lk]}{[k][nk]}}a_l(k),`$ $`a_1^{}(k)={\displaystyle \underset{l=1}{\overset{n1}{}}}{\displaystyle \frac{[(nl)k]}{[k][nk]}}a_l(k).`$ $`[a_j(k),a_{n1}^{}(k)]=\delta _{j,n1}{\displaystyle \frac{[k]}{k}},[a_j(k),a_1^{}(k)]=\delta _{j,1}{\displaystyle \frac{[k]}{k}}.`$ ## Appendix B Bosonization of the boundary vacuum states For readers’ convenience we summarize the bosonic formulae of the boundary vacuum . Let us set the symmetric matrix as $`\widehat{I}_{s,t}(k)=\{\begin{array}{cc}0,\hfill & (st=0),\hfill \\ \frac{[sk][(nt)k]}{[k]^2[nk]},\hfill & (1stn1),\hfill \\ \frac{[tk][(ns)k]}{[k]^2[nk]},\hfill & (1tsn1).\hfill \end{array}`$ (B.4) Let us consider the $``$-algebra generated by the bosons $`a_s(k)`$ $`(s\{1,\mathrm{},n1\},k)`$ which satisfy the following defining relations: $`[a_s(k),a_t(l)]=\delta _{k+l,0}{\displaystyle \frac{[(\alpha _s|\alpha _t)k][k]}{k}},`$ where I$`(\alpha _s|\alpha _t)`$ is an element of A-type Cartan matrix. The boundary state has the form $`|i_B=e^{F_i}|i,F_i={\displaystyle \underset{s,t=1}{\overset{n1}{}}}{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}\alpha _{s,t}(k)a_s(k)a_t(k)+{\displaystyle \underset{s=1}{\overset{n1}{}}}{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}\beta _s^{(i)}(k)a_s(k).`$ Here the coefficients of the quadratic part are given by $`\alpha _{s,t}(k)={\displaystyle \frac{kq^{2(n+1)k}}{2[k]}}\times \widehat{I}_{s,t}(k),`$ (B.5) and those of the linear part are given by $`\beta _j^{(i)}(k)`$ $`=`$ $`(q^{(n+3/2)k}q^{(n+1/2)k})\theta _k{\displaystyle \underset{s=1}{\overset{n1}{}}}\widehat{I}_{j,s}(k)`$ (B.6) $`+`$ $`\{\begin{array}{cc}0,\hfill & (C1),\hfill \\ \widehat{I}_{j,L}(k)q^{(2n2M+L+1/2)k}r^k\widehat{I}_{j,M}(k)q^{(2nM+1/2)k}r^k,\hfill & (C2),\hfill \\ \widehat{I}_{j,L}(k)q^{(2ML+1/2)k}r^k+\widehat{I}_{j,M}(k)q^{(M+1/2)k}r^k,\hfill & (C3),\hfill \end{array}`$ (B.10) where $`\theta _k=\{\begin{array}{cc}0,& k\text{ is odd},\\ 1,& k\text{ is even}.\end{array}`$ The dual boundary state has the form $`{}_{B}{}^{}i|=i|e^{G_i},G_i={\displaystyle \underset{s,t=1}{\overset{n1}{}}}{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}\gamma _{s,t}(k)a_s(k)a_t(k)+{\displaystyle \underset{s=1}{\overset{n1}{}}}{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}\delta _s^{(i)}(k)a_s(k).`$ (B.12) Here the coefficients of the quadratic part are given by $`\gamma _{s,t}(k)={\displaystyle \frac{kq^{2k}}{2[k]}}\times \widehat{I}_{s,t}(k),`$ (B.13) and those of the linear part are given by $`\delta _j^{(i)}(k)`$ $`=`$ $`(q^{k/2}q^{3k/2})\theta _k{\displaystyle \underset{s=1}{\overset{n1}{}}}\widehat{I}_{j,s}(k)`$ (B.14) $`+`$ $`\{\begin{array}{cc}0,\hfill & (C1),\hfill \\ q^{(L3/2)k}r^k\widehat{I}_{j,L}(k)q^{(2LM3/2)k}r^k\widehat{I}_{j,M}(k),\hfill & (C2),\hfill \\ q^{(L3/2)k}r^k\widehat{I}_{j,L}(k)+q^{(M2L3/2)k}r^k\widehat{I}_{j,M}(k),\hfill & (C3).\hfill \end{array}`$ (B.18)
warning/0001/astro-ph0001176.html
ar5iv
text
# 1 Introduction ## 1 Introduction X-ray observations have in the past played a minor, but important role, in cosmology. For example, the hot intergalactic medium prediction of the Steady State theory failed because the observed intensity was too low (Gould & Burbidge 1963). More recently, estimates of the gas fraction in clusters, which is dominated by the X-ray emitting intracluster medium, have shown that the matter density of the Universe is significantly below the critical value for closure (S White et al 1993; D White et al 1995). X-rays may not be the best means to quantify the cosmological parameters ($`H_0,\mathrm{\Omega }_\mathrm{m},q_0,\mathrm{\Lambda }`$ etc). Of course we should continue to do our best, but it is unlikely that X-ray results will ever supercede those expected from anticipated studies of the Cosmic Microwave Background. At present they supplement such work (e.g. Bridle et al 1999; Bahcall et al 1999). Cosmology for many of us is not, though, just the study of the geometry of the Universe, or of the inflation era, but involves the study of what has happened since redshifts of, say, 10. It involves the astrophysics of the baryons in this recent era, and working out how stars, black holes, galaxies and clusters arose and why. From that point of view, X-rays will be vital, since much of the dissipation in galaxy formation occurs at (soft) X-ray emitting temperatures; the X-ray emitting intracluster and intragroup medium is a living fossil, or calorimeter of the enrichment and energy injection history of galaxies; the central black holes in galaxies may only be clearly studied from their penetrating X-ray emission, and the isotropy of the X-ray Background may be guide to the development of structure. The deepest, and the most extended, gravitational potential wells in the Universe are potent X-ray emitters. I shall not in this introductory talk discuss the possible X-ray detection of most of the baryons in the local Universe through the study of groups, since this is covered by R Mushotzky; nor shall I explore the implications of the fluctuations in the X-ray Background, which are covered by X Barcons. Mainly I shall discuss some issues which I have been working on and which illustrate the main theme of this Conference. ## 2 Clusters and Cosmology Clusters are the most massive virialized objects in the Universe and so their number density is a key diagnostic of the cosmic power spectrum, on a scale of about 8 Mpc, $`\sigma _8`$. The evolution of their numbers, and particularly of the numbers above a particular intracluster gas (virial) temperature, is a good guide to the mean matter density in the Universe, $`\mathrm{\Omega }_\mathrm{m}`$ (Eke et al 1998). Current work suggests that if there is any evolution out to redshifts of about one, then it is very small (Ebeling 1999). The existence of the X-ray luminous EMSS cluster 1054-0321 provides a strong argument that $`\mathrm{\Omega }_\mathrm{m}`$ is significantly less than one (Gioia et al 1998). Together with radio or microwave observations, X-ray intensities yield interesting constraints on the Hubble constant through the S-Z effect. The common occurrence of cooling flows at the centres of clusters, and recent working on ageing the flows (Allen et al 1999), indicates the merger history of cluster cores and gives a clue to possible baryonic dark matter. Perhaps the most striking application of X-rays has been through the baryon fraction, $$f_\mathrm{b}=\frac{\mathrm{mass}\mathrm{of}\mathrm{baryons}}{\mathrm{total}\mathrm{mass}}=\frac{\mathrm{\Omega }_\mathrm{b}}{\mathrm{\Omega }_\mathrm{m}}.$$ Since $`\mathrm{\Omega }_\mathrm{b}`$ is known fairly well from the combination of cosmic nucleosynthesis and deuterium abundance work, $`\mathrm{\Omega }_\mathrm{m}`$ is well determined from X-ray cluster studies (White & Frenk 1991). The value found is about 0.3. Many clusters have now been carefully observed and their gas fractions determined out to $`r_{500}`$ or more. Specifying a radius is important as the fraction appears to rise with radius, from about 10 per cent in the core to about 20 per cent at $`r_{500}`$, within which the mean cluster density is 500 times that of the Universe. In a sample studied by Ettori and myself (1999), we find that the gas fraction drops with redshift as $`\left(1+z\right)^{1.5}`$, which is surprisingly steep given the lack of evolution seen in other properties. It is explained by our use of $`q_0=0.5`$ when determining $`r_{500}`$ (Sasaki 1996: for an isothermal cluster the gas mass $`M_\mathrm{g}nr^3`$, which since $`Ln^2r^3`$ means that $`M_\mathrm{g}L^{1/2}r^{3/2}D_\mathrm{A}`$, where $`D_\mathrm{A}`$ is the angular diameter distance of the cluster, then since the total mass $`M_\mathrm{T}TrD_\mathrm{A},`$ we have $`M_\mathrm{g}/M_\mathrm{T}D_\mathrm{A}^{3/2}`$ ). If we adjust $`q_0`$ (thus $`D_\mathrm{A}`$) so that the gas fraction is constant with redshift, then we find that $`q_0<0.1`$, in other words it provides a strong case for a low matter density Universe ($`\mathrm{\Omega }_\mathrm{m}<0.2`$). Why it gives a lower value than that found via cosmic nucleosynthesis is not clear, however. A further cosmologically interesting result from cluster studies with X-rays is that of scaling, particularly for the luminosity – temperature relation. From gravitational collapse alone, we expect that $`L_\mathrm{x}T_\mathrm{x}^2`$ (the virial radius $`r_\mathrm{v}r_{200}`$, defined as for $`r_{500}`$ above, within which the mean density of all clusters is the same at the formation redshift, therefore $`Mr_\mathrm{v}^3=\mathrm{const}`$ and since $`T_\mathrm{x}Mr_\mathrm{v}^1`$ from the virial theorem, then $`T_\mathrm{x}r_\mathrm{v}^2`$; then since $`L_\mathrm{x}n^2T_\mathrm{x}^{1/2}V`$ and the mean density of clusters $`n`$ is constant, then $`L_\mathrm{x}T_\mathrm{x}^2)`$. The observations however indicate $`T_\mathrm{x}^3`$ over the range $`kT_\mathrm{x}28`$ keV. At higher temperatures it appears to level off more to $`T_\mathrm{x}^2`$ (Allen & Fabian 1998) and is steeper at lower temperatures. A major implication of this departure from the expected scaling is that significant amounts of heat have been injected into clusters. Wu, Nulsen and I (1999) find it difficult to explain this in terms of supernova heating (but see Loewenstein 1999). Of course supernovae have enriched the intracluster medium in metals, but is likely that much of their heat was radiated from the interstellar medium of their host galaxies, If it was not then supernovae can hardly provide the feedback required for galaxy formation to result in the galaxies seen today rather than many small dense ones. We also investigated widespread cooling as a way of removing low entropy gas from cluster cores and also preheating to give an entropy floor to cluster gas Ponman et al 1999). None appears to be sufficient. We also investigated the limits that the $`1/4`$ keV intensity of the X-ray Background gives to galaxy formation (Wu, Fabian & Nulsen 1999; see also Pen 1997). The intensity predicted without any heat source apart from gravity is about an order of magnitude greater than that observed. Again a significant heat source is required, this time in all objects exceeding a few times $`10^{12}`$ solar masses. The total heat requirement to solve both the soft X-ray background problem and the scaling one is about 3 keV per particle, which is high. A possible and perhaps plausible heat source is winds from active galaxies, an issue I return to near the end. If a significant fraction of the power expected from the formation of the local mass density of massive black holes emerged in winds, as well as radiation, then the problems are solved. An implication of this is that the intergalactic medium may also have this mean energy, either in heat or potential energy, and so be much more difficult to detect then predicted by Cen & Ostriker (1999). ## 3 Active Galaxies and the X-ray Background The X-ray Background (XRB) is the sum of all the X-ray emission in the recent Universe. Since the spectrum in the 2–10 keV band is flatter than known classes of source, it is most probable that it is the sum of many obscured active galaxictic nuclei (AGN). This explanation arose over ten years ago (Setti & Woltjer 1989) and has been explored in detail since (Madau, Ghisellini & Fabian 1994; Celotti et al 1995; Matt & Fabian 1994; Comastri et al 1995; Wilman & Fabian 1999). A simple comarison of the XRB spectrum with that of an unobscured AGN, with a typical photon index of 2, demonstrates that most accretion in the Universe is obscured (Fabian et al 1998; Fabian & Iwasawa 1999). A robust estimate of the accretion power in the Universe, assumed to be from AGN, can be obtained by assuming that the intensity of the XRB at 30 keV is emitted by an underlying power-law and yet least affected by photoelectric absorption (Fabian & Iwasawa 1999). Matching that to the spectral energy distribution of an unobscured AGN (see e.g. Elvis et al 1994), then allows the absorption-corrected energy density from accretion to be determined, $`\epsilon _{\mathrm{AGN}}`$. It can be increased by a correction for Compton thick objects (where $`N_\mathrm{H}>1.5\times 10^{24}`$), by a factor of 1.3 (Maiolino et al 1998) to perhaps as much as 2. Then Soltan’s (1982) argument can be used to convert to an expected mean mass density in black holes now; $$\rho _{\mathrm{BH}}=\frac{\left(1+\overline{z}\right)\epsilon _{\mathrm{AGN}}}{0.1c^2},$$ where $`z2`$ is the mean redshift of the AGN, and an accretion efficiency of $`0.1`$ has been assumed. The result is about half the mass density found by Magorrian et al (1998) and is in good agreement with van der Marel’s (1999) value (see also Salucci et al 1999). It means that most of the mass of black holes is due to radiatively efficient, but obscured, accretion. Most, about 85 per cent, of the accretion power has been absorbed and presumably reradiated in the infrared. The total radiated power from AGN can be seen from the IR backgrounds (e.g. Fixsen et al 1998) to then be about one quarter (give or take a factor of two) of that from stars. Note here that X-rays are the only radiation which penetrates the absorber directly and so can discriminate and inform us of the actual evolution of AGN and their black holes. The radiation is absorbed at other wavelengths and so provides no direct information on the central source. It is likely that this also applies to the cores of most galaxies, which are the oldest parts. X-ray observations are likely to be crucial to understanding the accretion history of the Universe and of the evolution of the dense cores of galaxies. If the black holes in them formed very early and have always been obscured, then X-rays may become important for studying the earliest objects. There may be an intimate connection between the formation of a galaxy spheroid and its central black hole (Silk & Rees 1998; Fabian 1999). If part of the gas forming the spheroid remains as cold obscuring gas clouds, instead of rapidly forming stars, and if the central black hole grows by accretion and blows a wind, then it may blow away the gas, and end the growth of both the black hole and the spheroid when it becomes massive and powerful enough. This scenario means that the main growth phase of the black hole is obscured, and it only becomes unobscured when the fuel supply is blown away. It then lasts for a disk emptying time, which may be say 10 per cent of its original lifetime. As the gas is blown away it may be seen as a broad absorption line quasar. The implication here is that AGN have powerful winds, especially when they accrete near the Eddington limit. We have seen from the cluster discussion that such powerful winds can provide and explanation for the excess energy in clusters. Thus the total power of AGN and the excess heat in clusters may be intimately linked in this way. ## 4 Conclusions X-ray observations enable the thermal content and enrichment of the baryons in the Universe to be studied. The cooling of gas in the potential wells of galaxies, groups and clusters also emits predominantly in X-rays. They are the best direct probe of the accretion history of AGN, the integrated energy of which may be up to 50 per cent of that from stars. Acknowledgements. I am grateful to the organisers of this meeting for the opportunity to visit Santorini and hear the many interesting talks and discussions, as well as to my collaborators Stefano Ettori, Kazushi Iwasawa, Paul Nulsen and Kelvin Wu. The Royal Society is thanked for support.
warning/0001/astro-ph0001023.html
ar5iv
text
# Studies of Low-Mass Star Formation with ALMA ## 1. Introduction The focus in this paper will be on regions forming low mass stars in relative isolation, as other papers are covering issues of massive star formation and clustered star formation. Isolated star formation is interesting for several reasons. An empirical evolutionary scheme (Lada 1987, André et al. 1993) is generally accepted. There is a well-established theory (Shu et al. 1987) and variations (e.g., Foster & Chevalier 1993; Henriksen, André, & Bontemps 1997; McLaughlin & Pudritz 1997) that can be tested by observations. It is particularly suited to studies connecting star formation to planet formation. The two primary probes of the conditions in star-forming cores are continuum emission from dust and spectral lines from molecules. These are complementary in many ways. The dust emission is not affected by molecular depletion and traces column density very effectively, but dust grain sizes may be a function of the environment, or gas and dust distributions may differ because of ambipolar diffusion. In principle, molecular spectroscopy probes the local density and velocity fields, but it is sensitive to variations in chemical abundances. Together, these two probes can be very powerful. With ALMA, a hybrid probe will become widely available: molecular line absorption against emission from compact dust components, such as circumstellar disks. The exquisite sensitivity of ALMA to continuum emission will allow us to map the detailed structure of the dust column density in many cores, ranging along the evolutionary sequence, to trace the flow of matter from large scales to disks and stars. Together with information from other wavelength regions, a complete picture of the distribution of dust temperature and column density will result, along with information on possible changes in the grain size distribution. Maps of optically thin, thermally excited tracers will provide column densities of gas, for comparison to those of dust measured by the continuum emission. The molecular spectroscopy of lines with subthermal excitation will yield direct estimates of the local density. Combined studies of optically thick and thin lines will reveal the kinematics in detail. Finally, absorption spectroscopy of the material in front of opaque disks will be a new capability of ALMA that will help to unravel the complex velocity fields involved in forming stars. The dust continuum emission provides a probe of mass ($`M`$), total gas column density ($`N`$), dust temperature ($`T_D`$), grain properties, and the component of the magnetic field projected on the sky ($`B_{}`$). Molecular line emission can probe the gas kinetic temperature ($`T_K`$), volume density ($`n`$), velocity field ($`v`$), abundances ($`X`$), and the line-of-sight component of the magnetic field ($`B_{}`$). A more detailed discussion of these probes can be found in Evans (1999). ## 2. Dust Continuum Emission The dust continuum emission provides a probe of the total amount and distribution of the dust. If the dust and gas are well mixed, these quantities can be translated into the same information about the gas. One caveat is that the dust opacity may change from source to source or even within a single source (e.g., Visser et al. 1998). The primary tool that has been used in the past is the spectral energy distribution, or SED ($`S_\nu (\lambda )`$), which gives information on the source luminosity, by integrating under the SED. The mass can be determined by observing the flux in large beams at $`\lambda `$ large enough that the emission is optically thin. With suitable radiative transport codes (e.g. Egan, Leung, & Spagna 1988), source models constrained by the SED can yield the distribution of the dust temperature ($`T_D(r)`$) for a given set of grain properties. Matching different parts of the SED can constrain the choice of grain opacities (Adams, Lada, & Shu 1987; Butner et al. 1991; van der Tak et al. 1999). More recently, spatially resolved studies of the intensity of dust continuum emission have become a powerful probe. New instruments operating at submillimeter wavelengths have provided an enormous increase in this kind of data (e.g., Johnstone & Bally 1999; Motte, André, & Neri 1998). Maps of polarized dust emission are starting to provide maps of $`B_{}`$ (Greaves et al. 1999; Rao et al. 1998). By taking cuts through maps or by azimuthally averaging the intensity, one obtains $`I(b)`$, the intensity as a function of impact parameter ($`b`$), the separation of the beam from the center of emission. Plots of $`I(b)`$ provide probes of column density as a function of impact parameter $`N(b)`$, and with modeling, the density distribution $`n(r)`$, which is predicted by theories of isolated low-mass star formation. Recent examples of such studies are those of Chandler & Richer (2000), Hogerheijde & Sandell (2000) and Shirley et al. (2000). They reveal steeper density gradients than were apparent from most studies of molecular lines. An example of an SED and determinations of $`I(b)`$ are shown in Fig. 1 for B335, a well-studied region of low mass star formation. Interferometers naturally provide a complementary probe to $`I(b)`$, the spatial visibility function ($`S(uvdist)`$), which can also be compared to models. This approach is especially well suited to distinguishing compact structures, such as disks, from the envelope. An example of how this capability can separate components that would be otherwise indistinguishable is given in Fig. 2, taken from Looney, Mundy, & Welch (1997). To discuss the sensitivity of ALMA to dust continuum emission, we make some simplifying assumptions: the Rayleigh-Jeans limit is valid; the emission is optically thin; the dust opacity follows $$\kappa _\nu =9.0\times 10^{26}\mathrm{cm}^2\mathrm{H}_2^1\lambda _{\mathrm{mm}}^1;$$ (1) and the resolution is diffraction-limited ($`\theta _b=1.2\lambda /B`$, where $`B`$ is the maximum dimension of the array). Then the sensitivity to dust emission can be expressed in terms of the product of gas column density and dust temperature: $$NT_D=2.5\times 10^{24}\lambda _{\mathrm{mm}}B_{\mathrm{km}}^2\mathrm{\Delta }S_\nu (\mathrm{mJy})$$ (2) or, in terms of visual extinction, $$A_VT_D=2.5\times 10^3\lambda _{\mathrm{mm}}B_{\mathrm{km}}^2\mathrm{\Delta }S_\nu (\mathrm{mJy}).$$ (3) Alternatively, one can describe the sensitivity to gas mass in Earth masses given a distance ($`d`$): $$MT_D=470M_{}(d/140\mathrm{p}\mathrm{c})^2\lambda _{\mathrm{mm}}^3\mathrm{\Delta }\mathrm{S}_\nu (\mathrm{mJy}).$$ (4) The fiducial distance of 140 pc is a typical distance to nearby regions that are forming low-mass stars. These equations can be generalized to the case where the Rayleigh-Jeans approximation fails, at the cost of some clarity. These expressions will suffice to illustrate the main points, if we bear in mind that Rayleigh-Jeans failure will first tend to decrease the sensitivity to column density or mass at the shorter wavelengths. In calculating sensitivities, I have used the values of $`\mathrm{\Delta }S_\nu (\mathrm{mJy})`$ given by Butler & Wootten (1999), for 1.5 mm of precipitable water vapor (PWV). The values are for 1 $`\sigma `$ noise after 60 sec of integration. This would correspond to 22 $`\sigma `$ in an 8 hour integration. The resulting plots of $`NT_D`$ and $`MT_D`$ versus $`\lambda `$ are shown in Fig. 3, along with a plot of the spatial resolution in AU at a distance of 140 pc. The plot is truncated at 3 mm to allow the shorter wavelengths to be seen clearly. Note that $`\lambda 1`$ mm have comparable sensitivity to column density, with much better sensitivity in compact configurations. For the most compact configuration, ALMA could detect $`5.5\times 10^{20}`$ cm<sup>-2</sup> ($`A_V=0.55`$ mag) at $`T_D=10`$ K. The sensitivity to mass is best at $`\lambda =0.87`$ mm for the standard water vapor. The mass sensitivity is even better at the shortest wavelengths if the PWV drops to 0.35 mm (crosses in the top panel), but only if the dust is warm enough that the Rayleigh-Jeans limit applies ($`T_D>>41`$K at 850 GHz). To put things in perspective, ALMA can detect the dust emission from 0.3 $`M_{}`$ of gas at $`T_D=100`$ K at $`\lambda =0.87`$ mm! Another useful comparison is to the current state of the art. Table 1 shows the comparison between ALMA and SCUBA on the JCMT, assuming roughly comparable PWV, at $`\lambda =0.87`$ mm. It is clear from the figure and table that ALMA will take us into new regimes of sensitivity and resolution, allowing study of $`N(b)`$ and hence $`n(r)`$ to much finer scales than currently possible. The key requirement for this work is spatial dynamic range. Sensitivity to the largest relevant spatial scales is a challenge for any interferometer, and ALMA must solve this problem. ## 3. Molecular Line Emission Mapping molecular line emission provides a wealth of information about star forming regions. Transitions between levels in thermal equilibrium (CO, some transitions of H<sub>2</sub>CO, CH<sub>3</sub>CN, etc.) provide $`T_K`$, while transitions between levels not in thermal equilibrium allow a measure of $`n`$. The latter however is coupled to the abundance because of trapping, and multiple transitions of a single molecule are needed to separate these effects. In the simplest model, a homogeneous cloud, one derives a weighted mean density along each line of sight. For more sophisticated models (e.g., power laws and collapse models), the data constrain the model parameters. The line profiles contain vital information about the velocity field. While this has been difficult to extract, in some cases it is possible to learn about rotation (Goodman et al. 1993), infall (Myers, Evans, & Ohashi 2000), and outflow (Bachiller 1996). For a few transitions, the line profile, observed with suitable polarization, yields $`B_{}`$ (e.g., Crutcher 1999). The key instrumental parameters for molecular line emission are the beam size and the velocity resolution. Butler & Wootten (1999) give the formula for the line radiation temperature (proportional to intensity) for the case of diffraction-limited beams and velocity resolution of 1 km s<sup>-1</sup>: $$\mathrm{\Delta }T_R=0.32\mathrm{K}\mathrm{B}_{\mathrm{km}}^2\mathrm{\Delta }\mathrm{S}_1(\mathrm{mJy}).$$ (5) While bright lines may be observed with diffraction-limited beams, choosing a fixed beam of 1<sup>′′</sup> provides a convenient benchmark for weak lines. In this case, $$\mathrm{\Delta }T_R=0.013\mathrm{K}\frac{\lambda _{\mathrm{mm}}^2\mathrm{\Delta }\mathrm{S}_1(\mathrm{mJy})}{(\theta _\mathrm{b}/1{}_{}{}^{\prime \prime })^2\sqrt{\delta \mathrm{v}(\mathrm{km}\mathrm{s}^1)}}$$ (6) where $`\mathrm{\Delta }S_1(\mathrm{mJy})`$ is the sensitivity (1 $`\sigma `$ in 60 sec) in a 1 km s<sup>-1</sup> band from Butler & Wootten (1999), and $`\delta v(\mathrm{km}\mathrm{s}^1)`$ is the velocity resolution in km s<sup>-1</sup>. The current state of the art for interferometers using $`\theta _b=1^{\prime \prime }`$ is about 1000 mJy at 110 GHz. For ALMA, $`\mathrm{\Delta }S_1(\mathrm{mJy})`$ is less than 10 mJy up to 345 GHz, implying a gain of a factor of 100. Values of $`\mathrm{\Delta }T_R<0.5`$ K can be achieved from 0.35 to 2.7mm, even with PWV of 1.5 mm. Such sensitivity will allow detailed mapping of the temperature, density, and velocity field at an unprecedented scale. Figure 4 shows plots of $`\mathrm{\Delta }T_R`$ for $`\delta v=1`$ km s<sup>-1</sup> for both constant ($`\theta _b=1^{\prime \prime }`$) and for diffraction-limited beams. In addition to the ALMA requirements for sensitivity, spatial resolution, and spatial dynamic range established for continuum observations, line studies of molecular clouds add velocity resolution (must be easily variable and very fine) and frequency coverage (needed to cover the range of transitions and molecules needed for a full analysis). The baseline ALMA sensitivity for line radiation observed with diffraction-limited resolution and large baselines is marginal. Observing lines at very high spatial and spectral resolution will stretch ALMA to its limit in sensitivity. ## 4. Molecular Line Absorption Against Continuum Sources This technique, familiar at cm wavelengths, has only recently become possible in regions forming low mass stars (e.g., Choi, Panis, & Evans 1999). The background source is a compact dust continuum source, plausibly a circumstellar disk or perhaps the inner part of the envelope. With current instruments, only a few disks are strong enough to produce absorption lines from molecular gas in front. ALMA will make this a routine probe. Because the disk lies at the center of the infalling envelope, only gas in front of the disk will show absorption, while the rest of the cloud will produce emission. This selection provides a clear-cut way to resolve the infall-outflow ambiguity that plagues studies of cloud collapse. For a beam that includes only an opaque disk, only the front half of the cloud will be seen in absorption. For larger beams, the surrounding cloud will produce emission, resulting in an inverse P-Cygni profile for infall. The sensitivities for line emission are sufficient that most disks can be used in this way. ## 5. Summary and Requirements ALMA will provide a tremendous advance in capability for studies of isolated low-mass star formation. The key probes are dust continuum emission and molecular line emission. A new capability, molecular line absorption against circumstellar disks, will become routine, allowing clear-cut resolution of infall-outflow ambiguities. The requirements on ALMA for dust continuum emission are very low receiver temperatures and wide bandwidths, coverage of a wide range of wavelengths, and very good coverage of the uv plane. The last of these is particularly important, as star formation is intrinsically a multiscale problem, with essential information on scales ranging from a few AU to at least a few times $`10^4`$ AU. For spectral lines, the large bandwidth requirement is replaced by the need for a flexible correlator capable of velocity resolution as good as 0.01 km s<sup>-1</sup>. The requirement for a wide range of wavelengths is stiffened to essentially complete coverage of the bands that penetrate the atmosphere. This work has been supported by the State of Texas and NASA grant NAG5-7203. L. Looney and Y. Shirley provided figures. ## References Adams, F. C., Lada, C. J., & Shu, F. H. 1987, ApJ, 312, 788 André, P., Ward-Thompson, D., & Barsony, M. 1993, ApJ, 406, 122 Bachiller, R. 1996, ARA&A, 34, 111 Butler, B., and Wootten, A. 1999, ALMA Memo. No. 276 Butner, H. M., Evans, N. J., II, Lester, D. F., Levreault, R. M., & Strom, S. E. 1991, ApJ, 376, 636 Chandler, C. J., & Richer, J. S. 2000, ApJ, in press Choi, M., Panis, J-F., & Evans, N. J., II 1999, ApJS, 122, 519 Crutcher, R. M., 1999, ApJ, 520, 706 Egan, M. P., Leung, C. M., & Spagna, G. R., Jr. 1988, Computer Physics Communications, 48, 271 Evans, N. J., II 1999, ARA&A, 37, 311 Foster, P. N., & Chevalier, R. A. 1993, ApJ, 416, 303 Goodman, A. A., Benson, P. J., Fuller, G. A., & Myers, P. C. 1993, ApJ, 406, 528 Greaves, J. S., Holland, W. W., Friberg, P., & Dent, W. R. F. 1999, ApJ, 512, L139 Henriksen, R. N., André, P., & Bontemps, S. 1997, A&A, 323, 549 Hogerheijde, M. R., & Sandell, G. 2000, ApJ, in press Johnstone, D., & Bally, J. 1999, ApJ, 510, L49 Lada, C. J. 1987 in IAU Symp 115, Star Forming Regions, ed. M. Peimbert & J. Jugaku (Dordrecht: Reidel), 1 Looney, L. W., Mundy, L. G., & Welch, W. J. 1997, ApJ, 484, L157 McLaughlin, D. E., & Pudritz, R. E. 1997, ApJ, 476, 750 Motte, F., André, P., & Neri, R. 1998, A&A, 336, 150 Myers, P. C., Evans, N. J., II, & Ohashi, N. 2000, in Protostars and Planets IV, ed. V. Mannings, A. Boss, & S. Russell (Tucson: Univ. Arizona), in press Rao, R., Crutcher, R. M., Plambeck, R. L., & Wright, M. C. H. 1998, ApJ, 502, L75 Shirley, Y. L., Evans, N. J., II, Rawlings, J. M. C., Gregersen, E. M. 2000, ApJ, submitted Shu, F. H., Adams, F. C., & Lizano, S. 1987, ARA&A, 25, 23 van der Tak, F. F. S. van Dishoeck, E. F., Evans, N. J., II, Bakker, E., & Blake, G. A. 1999, ApJ, 522, 991 Visser, A. E., Richer, J. S., Chandler, C. J., & Padman, R. 1998, MNRAS, 301, 585
warning/0001/hep-th0001031.html
ar5iv
text
# Nt-yet ## 1 Introduction It has been shown (see for a recent review) that the dynamics of massless fields of all spins in $`AdS_4`$ can be described in terms of star-product algebras acting on the auxiliary spinor variables. In this formalism the part of the (nonlinear) equations of motion that contain space-time derivatives has a form of zero-curvature or covariant constancy conditions and therefore can be solved explicitly at least locally. The aim of this paper is to illustrate how this machinery can be used in practice to derive solutions of the massless free field equations in $`AdS_4`$. ### 1.1 AdS Geometry As is well-known, $`AdS_d`$ geometry is described by the zero-curvature equations for the $`AdS_d`$ algebra $`o(d1,2)`$ with the gauge fields $`A_{\underset{¯}{n}}{}_{}{}^{AB}(x)`$ ($`A,B=0÷d`$; $`\underset{¯}{m},\underset{¯}{n}=0÷d1`$; $`x^{\underset{¯}{n}}`$ are coordinates of $`AdS_d`$) identified with the $`AdS_d`$ gravitational fields according to $$\omega _{\underset{¯}{n}}{}_{}{}^{ab}=A_{\underset{¯}{n}}{}_{}{}^{ab},h_{\underset{¯}{n}}{}_{}{}^{a}=\lambda ^1A_{\underset{¯}{n}}{}_{}{}^{ad},$$ (1.1) where $`a,b=0÷d1`$ and $`\lambda `$ is a constant to be identified with the inverse $`AdS`$ radius. The $`o(d1,2)`$ field strengths are $$R_{\underset{¯}{m}\underset{¯}{n}}{}_{}{}^{ab}=_{\underset{¯}{m}}\omega _{\underset{¯}{n}}{}_{}{}^{ab}+\omega _{\underset{¯}{m}}{}_{}{}^{a}{}_{c}{}^{}\omega _{\underset{¯}{n}}^{}{}_{}{}^{cb}\lambda ^2h_{\underset{¯}{m}}{}_{}{}^{a}h_{\underset{¯}{n}}^{}{}_{}{}^{b}(\underset{¯}{m}\underset{¯}{n}),$$ (1.2) $$R_{\underset{¯}{m}\underset{¯}{n}}{}_{}{}^{a}=_{\underset{¯}{m}}h_{\underset{¯}{n}}{}_{}{}^{a}+\omega _{\underset{¯}{m}}{}_{}{}^{a}{}_{c}{}^{}h_{\underset{¯}{n}}^{}{}_{}{}^{c}(\underset{¯}{m}\underset{¯}{n}).$$ (1.3) Interpreting $`\omega _{\underset{¯}{n}}^{ab}`$ as Lorentz connection and $`h_{\underset{¯}{n}}^b`$ as the frame 1-form one observes that $`R_{\underset{¯}{m}\underset{¯}{n}}^a`$ identifies with the torsion tensor while the $`\lambda `$independent part of $`R_{\underset{¯}{m}\underset{¯}{n}}^{ab}`$ is the Riemann tensor. Setting $`R_{\underset{¯}{m}\underset{¯}{n}}{}_{}{}^{a}=0`$ one expresses $`\omega _{\underset{¯}{n}}^{ab}`$ in terms of $`h_{\underset{¯}{m}}^a`$. Imposing the equation $`R_{\underset{¯}{m}\underset{¯}{n}}{}_{}{}^{ab}=0`$ is then equivalent to the equation for $`AdS_d`$ described in terms of the frame field $`h_{\underset{¯}{n}}^a`$ which is required to be non-degenerate. Thus, $`AdS_d`$ geometry is described by the zero-curvature equation $$R^{AB}=0$$ (1.4) provided that $`det|h_{\underset{¯}{n}}{}_{}{}^{a}|0`$. From now on we focus on the particular case of $`AdS_4`$ using the well-known isomorphism $`o(3,2)sp(4|R)`$. The algebra $`sp(4|R)`$ admits the oscillator realization with the generators $$L_{\alpha \beta }=\frac{1}{4i}\{\widehat{y}_\alpha ,\widehat{y}_\beta \},\overline{L}_{\dot{\alpha }\dot{\beta }}=\frac{1}{4i}\{\widehat{\overline{y}}_{\dot{\alpha }},\widehat{\overline{y}}_{\dot{\beta }}\},P_{\alpha \dot{\beta }}=\lambda \frac{1}{2i}\widehat{y}_\alpha \widehat{\overline{y}}_{\dot{\beta }}$$ (1.5) realized as bilinears in two-component spinor oscillators satisfying the commutation relations<sup>1</sup><sup>1</sup>1One can equivalently use the Majorana spinor oscillators $`\widehat{Y}_\nu `$ $`(\nu =1÷4`$) with the commutation relations $`[\widehat{Y}_\mu ,\widehat{Y}_\nu ]=2iC_{\mu \nu },`$ where $`C_{\mu \nu }`$ is the charge conjugation matrix. The language of two-component spinors is however most useful for the analysis below. $$[\widehat{y}_\alpha ,\widehat{y}_\beta ]=2iϵ_{\alpha \beta },[\widehat{\overline{y}}_{\dot{\alpha }},\widehat{\overline{y}}_{\dot{\beta }}]=2iϵ_{\dot{\alpha }\dot{\beta }},[\widehat{y}_\alpha ,\widehat{\overline{y}}_{\dot{\beta }}]=0$$ (1.6) ($`\alpha ,\beta =1,2`$, $`\dot{\alpha },\dot{\beta }=1,2;\overline{y}_{\dot{\alpha }}=(y_\alpha )^+`$; for conventions see Appendix). The $`AdS_4`$ gravitational fields can now be identified with the 1-form bilinear in the oscillators $`w_0=dx^{\underset{¯}{n}}w_{0\underset{¯}{n}}={\displaystyle \frac{1}{4i}}dx^{\underset{¯}{n}}(\omega _{0\underset{¯}{n}}{}_{}{}^{\alpha \beta }\widehat{y}_{\alpha }^{}\widehat{y}_\beta +\overline{\omega }_{0\underset{¯}{n}}{}_{}{}^{\dot{\alpha }\dot{\beta }}\widehat{\overline{y}}_{\dot{\alpha }}^{}\widehat{\overline{y}}_{\dot{\beta }}+2\lambda h_{0\underset{¯}{n}}{}_{}{}^{\alpha \dot{\beta }}\widehat{y}_{\alpha }^{}\widehat{\overline{y}}_{\dot{\beta }})`$ (1.7) and satisfying the zero-curvature equation $$0=R_0dw_0w_0w_0.$$ (1.8) Here $`\omega _{0\underset{¯}{n}}{}_{}{}^{\alpha \beta }(x)`$ and $`\overline{\omega }_{0\underset{¯}{n}}{}_{}{}^{\dot{\alpha }\dot{\beta }}(x)`$ describe Lorentz connection while $`h_{0\underset{¯}{n}}{}_{}{}^{\alpha \dot{\beta }}(x)`$ describes vierbein in terms of two-component spinors. The curvature $`R=dwww`$ admits the expansion analogous to (1.7) with the components $`R_{\alpha \beta }=d\omega _{\alpha \beta }+\omega _\alpha {}_{}{}^{\gamma }\omega _{\beta \gamma }+\lambda ^2h_\alpha {}_{}{}^{\dot{\gamma }}h_{\beta \dot{\gamma }},`$ (1.9) $`\overline{R}_{\dot{\alpha }\dot{\beta }}=d\overline{\omega }_{\dot{\alpha }\dot{\beta }}+\overline{\omega }_{\dot{\alpha }}{}_{}{}^{\dot{\gamma }}\overline{\omega }_{\dot{\beta }\dot{\gamma }}+\lambda ^2h^\gamma {}_{\dot{\alpha }}{}^{}h_{\gamma \dot{\beta }},`$ (1.10) $`R_{\alpha \dot{\beta }}=dh_{\alpha \dot{\beta }}+\omega _\alpha {}_{}{}^{\gamma }h_{\gamma \dot{\beta }}+\overline{\omega }_{\dot{\beta }}{}_{}{}^{\dot{\gamma }}h_{\alpha \dot{\gamma }}.`$ (1.11) A particular solution of (1.8) can be chosen in the form $$h_{0\underset{¯}{n}}{}_{}{}^{\alpha \dot{\beta }}=z^1\sigma _{\underset{¯}{n}}{}_{}{}^{\alpha \dot{\beta }},$$ (1.12) $$\omega _{0\underset{¯}{n}}{}_{}{}^{\alpha \alpha }=\lambda ^2z^1\sigma _{\underset{¯}{n}}{}_{}{}^{\alpha \dot{\beta }}x_{}^{\alpha }{}_{\dot{\beta }}{}^{},$$ (1.13) $$\overline{\omega }_{0\underset{¯}{n}}{}_{}{}^{\dot{\beta }\dot{\beta }}=\lambda ^2z^1\sigma _{\underset{¯}{n}}{}_{}{}^{\alpha \dot{\beta }}x_{\alpha }^{}{}_{}{}^{\dot{\beta }},$$ (1.14) where $$x_{\alpha \dot{\beta }}=\sigma _{\alpha \dot{\beta }}^ax_a,x^2=x_ax^a=\frac{1}{2}x_{\alpha \dot{\beta }}x^{\alpha \dot{\beta }},z=1+\lambda ^2x^2,$$ (1.15) and sigma-matrices are Hermitian $`\overline{\sigma }_{\underset{¯}{n}}{}_{}{}^{\alpha \dot{\beta }}=\sigma _{\underset{¯}{n}}^{\beta \dot{\alpha }}`$ with the normalization $`\sigma _n{}_{}{}^{\alpha \dot{\beta }}\sigma _{m}^{}{}_{\alpha \dot{\beta }}{}^{}=2\eta _{nm}`$ where $`\eta _{nm}=\{1,1,1,1\}`$ is the flat Minkowski metric. Following to we use conventions with upper(lower) indices denoted by the same letter subject to symmetrization (see also Appendix). The equations (1.12)-(1.14) describe the vierbein and Lorentz connection of $`AdS_4`$ corresponding to the stereographic coordinates for the hyperboloid realization of $`AdS_4`$ via embedding into the 5d flat space with the signature $`(+,,,,+)`$. The $`AdS_4`$ metric tensor resulting from (1.12) is $$g_{0\underset{¯}{m}\underset{¯}{n}}=\frac{1}{2}h_{0\underset{¯}{m}}{}_{}{}^{\alpha \dot{\beta }}h_{0\underset{¯}{n}}^{}{}_{\alpha \dot{\beta }}{}^{}=\frac{\eta _{\underset{¯}{m}\underset{¯}{n}}}{(1+\lambda ^2x^2)^2}.$$ (1.16) Note that this form of the metric can be fixed by the requirement that it is conformally flat, regular at $`x=0`$ and depends on the coordinates $`x^{\alpha \dot{\beta }}`$ via the Lorentz covariant combination $`x^2`$. In the flat limit $`\lambda 0`$ this metric describes the Minkowski space. In these coordinates, the boundary of $`AdS_4`$ is identified with the hypersurface $`z=0`$. The “north pole” is realized as infinity $`x=\mathrm{}`$. It becomes a regular point in the inversed coordinates $$x^{\underset{¯}{n}}x_I^{\underset{¯}{n}}=\lambda ^2\frac{x^{\underset{¯}{n}}}{x^2}$$ (1.17) with the gravitational fields of the form: $$h_{0\underset{¯}{n}}^I{}_{}{}^{\alpha \dot{\beta }}=z_I^1\sigma _{\underset{¯}{n}}{}_{}{}^{\alpha \dot{\beta }},$$ (1.18) $$\omega _{0\underset{¯}{n}}^I{}_{}{}^{\alpha \alpha }=\lambda ^2z_I^1\sigma _{\underset{¯}{n}}{}_{}{}^{\alpha \dot{\beta }}x_{I}^{}{}_{}{}^{\alpha }{}_{\dot{\beta }}{}^{},$$ (1.19) $$\overline{\omega }_{0\underset{¯}{n}}^I{}_{}{}^{\dot{\beta }\dot{\beta }}=\lambda ^2z_I^1\sigma _{\underset{¯}{n}}{}_{}{}^{\alpha \dot{\beta }}x_{I\alpha }^{}{}_{}{}^{\dot{\beta }},$$ (1.20) This coordinate system describes the antipodal chart. ### 1.2 Star-Product and Free Massless Equations Instead of working in terms of the operators $`\widehat{y}_\alpha `$ and $`\widehat{\overline{y}}_{\dot{\alpha }}`$ it is convenient to use the star-product defined by the formula $`(fg)(y,\overline{y}|x)=(2\pi )^4{\displaystyle d^2ud^2\overline{u}d^2vd^2\overline{v}f(y+u,\overline{y}+\overline{u}|x)g(y+v,\overline{y}+\overline{v}|x)e^{i(u_\alpha v^\alpha +\overline{u}_{\dot{\alpha }}\overline{v}^{\dot{\alpha }})}}.`$ (1.21) It is elementary to see that this star-product is associative and well-defined for polynomial functions. The normalization is chosen such that 1 is the unit element of the algebra, i.e. $`1f=f1=f`$. The commutation relations for the generating elements $`y_\alpha `$ and $`\overline{y}_{\dot{\alpha }}`$ are $$[y_\alpha ,y_\beta ]_{}=2iϵ_{\alpha \beta },[\overline{y}_{\dot{\alpha }},\overline{y}_{\dot{\beta }}]_{}=2iϵ_{\dot{\alpha }\dot{\beta }},[y_\alpha ,\overline{y}_{\dot{\beta }}]_{}=0,$$ (1.22) where $`[a,b]_{}=abba.`$ The formula (1.21) is equivalent to the standard differential Moyal star-product by virtue of the Taylor expansion $$f(y,\overline{y})=exp(y^\alpha \frac{}{z^\alpha }+\overline{y}^{\dot{\alpha }}\frac{}{\overline{z}^{\dot{\alpha }}})f(z,\overline{z})|_{z=\overline{z}=0}.$$ (1.23) It describes the totally symmetric (i.e., Weyl) ordering of the operators $`\widehat{y}_\alpha ,\widehat{\overline{y}}_{\dot{\alpha }}`$ in terms of symbols of operators and was called “triangle formula” in . The star-product (1.21) acts on the auxiliary spinor variables $`y_\alpha ,\overline{y}_{\dot{\alpha }}`$ rather than directly on the space-time coordinates as in the non-commutative Yang-Mills limit of the string theory . It was argued however in that the dynamical field equations for higher spin massless fields transform the star-product nonlocality in $`y_\alpha ,\overline{y}_{\dot{\alpha }}`$ into the space-time nonlocality of the higher spin interactions. In terms of the star-product, the equation of the background $`AdS_4`$ space has the form: $$dw_0=w_0w_0.$$ (1.24) A less trivial fact shown in is that free equations for massless fields of all spins in $`AdS_4`$ can be cast into the form $`dw_1w_0w_1w_1w_0={\displaystyle \frac{i}{4}}[h_\alpha {}_{}{}^{\dot{\beta }}h^{\alpha \dot{\gamma }}{\displaystyle \frac{}{\overline{y}^{\dot{\beta }}}}{\displaystyle \frac{}{\overline{y}^{\dot{\gamma }}}}C(0,\overline{y})+h^\alpha {}_{\dot{\beta }}{}^{}h^{\gamma \dot{\beta }}{\displaystyle \frac{}{y^\alpha }}{\displaystyle \frac{}{y^\gamma }}C(y,0)],`$ (1.25) $$dC=w_0CC\stackrel{~}{w}_0.$$ (1.26) Here $`w_1(y,\overline{y}|x)`$ and $`C(y,\overline{y}|x)`$ are functions of spinor and space-time coordinates and tilde is defined according to $$\stackrel{~}{f}(y,\overline{y})=f(y,\overline{y}).$$ (1.27) Relativistic fields are identified with the coefficients in the Taylor expansions in powers of the auxiliary spinor variables $$w_1(y,\overline{y}x)=\underset{n,m=0}{\overset{\mathrm{}}{}}\frac{\lambda ^{1|\frac{nm}{2}|}}{2in!m!}w_1^{\alpha _1\mathrm{}\alpha _n}{}_{,}{}^{}{}_{}{}^{\dot{\beta }_1\mathrm{}\dot{\beta }_m}(x)y_{\alpha _1}\mathrm{}y_{\alpha _n}\overline{y}_{\dot{\beta }_1}\mathrm{}\overline{y}_{\dot{\beta }_m}$$ (1.28) and $$C(y,\overline{y}|x)=\underset{n,m=0}{\overset{\mathrm{}}{}}\frac{\lambda ^{2\frac{n+m}{2}}}{m!n!}C^{\alpha _1\mathrm{}\alpha _n}{}_{,}{}^{}{}_{}{}^{\dot{\alpha }_1\mathrm{}\dot{\alpha }_m}(x)y_{\alpha _1}\mathrm{}y_{\alpha _n}\overline{y}_{\dot{\alpha }_1}\mathrm{}\overline{y}_{\dot{\alpha }_m},$$ (1.29) Inserting (1.29) into (1.26), one arrives at the following infinite chain of equations $$D^LC_{\alpha (m),\dot{\beta }(n)}=ih^{\gamma \dot{\delta }}C_{\alpha (m)\gamma ,\dot{\beta }(n)\dot{\delta }}i\lambda ^2nmh_{\alpha \dot{\beta }}C_{\alpha (m1),\dot{\beta }(n1)},$$ (1.30) where $`D^L`$ is the Lorentz-covariant differential $$D^LA_{\alpha \dot{\beta }}=dA_{\alpha \dot{\beta }}+\omega _\alpha {}_{}{}^{\gamma }A_{\gamma \dot{\beta }}+\overline{\omega }_{\dot{\beta }}{}_{}{}^{\dot{\delta }}A_{\alpha \dot{\delta }}.$$ (1.31) The system (1.30) decomposes into a set of independent subsystems with $`nm`$ fixed. It turns out that the subsystem with $`|nm|=2s`$ describes a massless field of spin $`s`$ (note that the fields $`C_{\alpha (m),\dot{\beta }(n)}`$ and $`C_{\beta (n),\dot{\alpha }(m)}`$ are complex conjugated). Analogously, by the substitution of (1.28), the equation (1.25) amounts to $`R_{1\alpha (n),\dot{\beta }(m)}`$ $`D^Lw_{1\alpha (n)\dot{\beta }(m)}+n\gamma (n,m,\lambda )h_\alpha {}_{}{}^{\dot{\delta }}w_{1\alpha (n1),\dot{\beta }(m)\dot{\delta }}+m\gamma (m,n,\lambda )h^\gamma {}_{\dot{\beta }}{}^{}w_{1\gamma \alpha (n),\dot{\beta }(m1)}`$ (1.32) $`=\delta (m)h^{\gamma \dot{\delta }}h^\gamma {}_{\dot{\delta }}{}^{}C_{\alpha (n)\gamma (2)}^{}+\delta (n)h^{\eta \dot{\delta }}h_\eta {}_{}{}^{\dot{\delta }}\overline{C}_{\dot{\beta }(m)\dot{\delta }(2)}^{},`$ where $$\gamma (n,m,\lambda )=\theta (mn)+\lambda ^2\theta (nm2)+\lambda \delta (nm1).$$ (1.33) A spin $`s1`$ dynamical massless field is identified with the 1-form (potential) $$w_{\alpha (n),\dot{\beta }(n)}n=(s1),s1\text{integer},$$ (1.34) $$w_{\alpha (n),\dot{\beta }(m)}n+m=2(s1),|nm|=1s3/2\text{half-integer}.$$ (1.35) The matter fields are described by the 0-forms $$C_{\alpha (0),\dot{\alpha }(0)}s=0,$$ (1.36) $$C_{\alpha (1),\dot{\alpha }(0)}C_{\alpha (0),\dot{\alpha }(1)}s=1/2.$$ (1.37) All other components of the expansions (1.28) and (1.29) express in terms of the derivatives of physical fields by virtue of the equations (1.32), (1.30). For example, for 0-forms $`C`$ one has $$C_{\alpha (n),\dot{\beta }(m)}=\frac{1}{(2i)^{\frac{1}{2}(n+m2s)}}h_{\alpha \dot{\beta }}^{\underset{¯}{n}_1}D_{\underset{¯}{n}_1}^L\mathrm{}h_{\alpha \dot{\beta }}^{\underset{¯}{n}_{\frac{1}{2}(n+m2s)}}D_{\underset{¯}{n}_{\frac{1}{2}(n+m2s)}}^LC_{\alpha (2s)}nm,$$ (1.38) or $$C_{\alpha (n),\dot{\beta }(m)}=\frac{1}{(2i)^{\frac{1}{2}(n+m2s)}}h_{\alpha \dot{\beta }}^{\underset{¯}{n}_1}D_{\underset{¯}{n}_1}^L\mathrm{}h_{\alpha \dot{\beta }}^{\underset{¯}{n}_{\frac{1}{2}(n+m2s)}}D_{\underset{¯}{n}_{\frac{1}{2}(n+m2s)}}^LC_{\dot{\beta }(2s)}nm.$$ (1.39) Analogous formula holds for the 1-forms $`w`$ $$w_{\alpha (n),\dot{\beta }(m)}\left(\frac{}{x}\right)^{[\frac{|nm|}{2}]}w^{phys}+w^{gauge},$$ (1.40) where $`w^{phys}`$ denotes the appropriate field from the list (1.34), (1.35) while $`w^{gauge}`$ is a pure gauge part. As shown in the content of the equations (1.25) and (1.26) is equivalent to the relations (1.32), (1.30) and usual dynamical equations for the physical massless fields in $`AdS_4`$. The fields $$w^{\alpha (n),\dot{\beta }(m)}\text{with}n+m=2(s1)$$ (1.41) and $$C_{\alpha (m),\dot{\beta }(n)}\text{with}|nm|=2s$$ (1.42) are associated with the massless field of spin $`s`$ along with all its on-mass-shell nontrivial derivatives. For spin $`s1`$, the 0 forms $`C`$ describe gauge invariant field strengths generalizing the spin 1 Maxwell field strength and spin 2 Weyl tensor to an arbitrary spin. For example, in the spin 2 case (1.28) is equivalent to the linearized Einstein equations because it just tells us that torsion is zero and all the components of the linearized Riemann tensor are zero except for those which are described by the Weyl tensor $`C_{\alpha _1\mathrm{}\alpha _4}`$, $`C_{\dot{\alpha }_1\mathrm{}\dot{\alpha }_4}`$. The aim of this paper is to explore the fact that once the dynamical equations are reformulated in the form (1.25) and (1.26) one can write down their generic solution explicitly in terms of the star-product provided that the $`AdS_4`$ vacuum equations (1.24)-(1.26) are solved in the pure gauge form $$w_0=g^1dg,$$ (1.43) where $`g(y,\overline{y}|x)`$ is some invertible element of the star-product algebra, $`gg^1=g^1g=1`$ and $`d=dx^{\underset{¯}{n}}\frac{}{x^{\underset{¯}{n}}}`$ is the space-time differential. A form of the appropriate $`g(y,\overline{y}|x)`$ is found in the section 2. The covariant constancy equation (1.26) has the generic solution of the form $$C(x)=g^1(x)C_0\stackrel{~}{g}(x),$$ (1.44) where $`C_0(y,\overline{y})`$ is an arbitrary $`x`$independent element of the star-product algebra. For exponential functions $`C_0`$ we reproduce in the section 3 a set of particular solutions that in the flat limit tend to the plane wave solutions for massless fields of an arbitrary spin. A less trivial fact shown in the section 4 is that it is also possible to write down a generic solution of the equation (1.25) despite it does not have a zero-curvature form. Let us emphasize that this way of solving dynamical equations reduces to evaluation of some elementary Gaussian integrals originating from the star-product (1.21) (i.e. there is no need to solve any differential equations). The equation (1.44) can be interpreted as a covariantized Taylor expansion. Indeed, if $`g(x_0)=1`$ for some $`x_0`$ then $`C_0(y,\overline{y})=C(y,\overline{y}|x_0)`$. In accordance with (1.29), (1.38) and (1.39), $`C_0(y,\overline{y})`$ therefore describes all on-mass-shell nontrivial derivatives of the physical field at the point $`x_0`$. ## 2 Gauge Function In this section we find the function $`g(y,\overline{y}|x)`$ that reproduces the $`AdS_4`$ gravitational fields (1.12)-(1.14) in the star-product algebra version of the ansatz (1.7) $`w_{0\underset{¯}{n}}={\displaystyle \frac{1}{4i}}(\omega _{0\underset{¯}{n}}{}_{}{}^{\alpha \beta }y_{\alpha }^{}y_\beta +\overline{\omega }_{0\underset{¯}{n}}{}_{}{}^{\dot{\alpha }\dot{\beta }}\overline{y}_{\dot{\alpha }}^{}\overline{y}_{\dot{\beta }}+2\lambda h_{0\underset{¯}{n}}{}_{}{}^{\alpha \dot{\beta }}y_{\alpha }^{}\overline{y}_{\dot{\beta }})`$ (2.1) by virtue of the pure gauge representation (1.43). Note that the bilinear ansatz is consistent, because the bilinears form a closed $`sp(4)`$ subalgebra with respect to commutators<sup>2</sup><sup>2</sup>2It is a simple exercise with the star-product to check that this ansatz with the fields (1.12)-(1.14) satisfies (1.24). We therefore have to find such a function $`g`$ that the resulting connection (1.43) is bilinear in the auxiliary oscillators $`y`$ and $`\overline{y}`$. Note that if $`w_0`$ would contain some higher-order polynomials in oscillators, this would imply that some higher spin fields acquire nonvanishing vacuum expectation values thus making the physical interpretation of the corresponding solutions less straightforward. Let us look for the solution for $`g`$ in the Lorentz-covariant form $$g(y,\overline{y}|x)=e^{if(x^2)x^{\alpha \dot{\alpha }}y_\alpha \overline{y}_{\dot{\alpha }}+r(x^2)}$$ (2.2) with some functions $`f(x^2)`$ and $`r(x^2)`$. One readily shows that the inverse element $`g^1`$ is $$g^1(y,\overline{y}|x)=(1+f^2x^2)^2e^{if(x^2)x^{\alpha \dot{\alpha }}y_\alpha \overline{y}_{\dot{\alpha }}r(x^2)}.$$ (2.3) Direct computation using the star-product (1.21) leads by virtue of evaluation of elementary Gaussian integrals to the following result $`g^1dg=2(m_1+m_2x^{\alpha \dot{\beta }}y_\alpha \overline{y}_{\dot{\beta }})x_ndx^n+m_3y_\alpha \overline{y}_{\dot{\beta }}dx^{\alpha \dot{\beta }}+m_4(x^\gamma {}_{\dot{\beta }}{}^{}y_{\alpha }^{}y_\gamma +x_\alpha {}_{}{}^{\dot{\gamma }}\overline{y}_{\dot{\beta }}^{}\overline{y}_{\dot{\gamma }})dx^{\alpha \dot{\beta }},`$ (2.4) where $$m_1=r^{}q^1(2ff^{}x^2+f^2),$$ $$m_2=iq^1f^{}+iq^2f^3,$$ $$m_3=iq^2f(1f^2x^2),$$ $$m_4=iq^2f^2,$$ and we use notation $$q=1+f^2x^2$$ and $$h^{}(x^2)=\frac{h(x^2)}{x^2}.$$ In order to reproduce $`AdS_4`$ vierbein and connection (1.12)-(1.14) we have to set $`m_1=m_2=0`$. The condition $`m_2=0`$ is equivalent to $$f^3+f^{}+f^{}f^2x^2=0.$$ (2.5) Its general solution is $`f_\pm =\lambda ({\displaystyle \frac{1}{1\pm \sqrt{z}}}),`$ (2.6) where $`\lambda `$ is an arbitrary integration constant and $`z`$ is defined in (1.15). The solution $`f_+`$ is regular at $`x0`$. The solution $`f_{}`$ is singular at $`x0`$. In fact, the two gauge functions $`g`$ corresponding to these solutions are related to each other by the inversion (1.17) away from the north ($`x=\mathrm{}`$) and south ($`x=0`$) poles thus solving the problem for the two stereographic charts. In the rest of the paper we will for definiteness consider the solution regular at $`x=0`$ with $`f=f_+`$. Solving the condition $`m_1=0`$ for $`f_+`$, one finds up to the integration constant that does not affect $`w_0`$, that $$r=ln(\frac{2\sqrt{z}}{1+\sqrt{z}}).$$ (2.7) As a result $`g`$ takes the form $`g(y,\overline{y}|x)=2{\displaystyle \frac{\sqrt{z}}{1+\sqrt{z}}}\mathrm{exp}[{\displaystyle \frac{i\lambda }{1+\sqrt{z}}}x^{\alpha \dot{\alpha }}y_\alpha \overline{y}_{\dot{\alpha }}]`$ (2.8) with the inverse $$g^1(y,\overline{y}|x)=\stackrel{~}{g}(y,\overline{y}|x)=2\frac{\sqrt{z}}{1+\sqrt{z}}\mathrm{exp}[\frac{i\lambda }{1+\sqrt{z}}x^{\alpha \dot{\beta }}y_\alpha \overline{y}_{\dot{\beta }}].$$ (2.9) Inserting (2.6) and (2.7) into the expressions for $`m_3`$ and $`m_4`$ we find that the corresponding Lorentz connection and vierbein reproduce (1.12)-(1.14). The solution (1.18)-(1.20) corresponds to $`f_{}`$. ## 3 Plane Waves Having found the gauge function $`g`$ one solves the free field equation (1.26) for matter fields and Weyl tensors of an arbitrary spin in the form (1.44) equivalent to $`C(y,\overline{y}|x)=(2\pi )^4{\displaystyle _{\mathrm{}}^{\mathrm{}}}`$ $`d^2ud^2vd^2\overline{u}d^2\overline{v}\mathrm{exp}i(u_\alpha v^\alpha +\overline{u}_{\dot{\alpha }}\overline{v}^{\dot{\alpha }})`$ (3.1) $`\times g^1(y+u,\overline{y}+\overline{u})C_0(y+u+v,\overline{y}+\overline{u}+\overline{v})\stackrel{~}{g}(y+v,\overline{y}+\overline{v}).`$ As emphasized in the section 1.2, this formula reproduces the covariantized Taylor expansion with $`C_0`$ identified with all on-mass-shell nontrivial derivatives of $`C(x)`$ at $`x=x_0`$ with $`g(x_0)=I`$. Note that $`x_0=0`$ is the “south pole” for the solution (2.8). Let us now choose $`C_0`$ in the form $`C_0=\lambda ^2c_0\mathrm{exp}i\lambda ^{\frac{1}{2}}(y^\alpha \eta _\alpha +\overline{y}^{\dot{\alpha }}\overline{\eta }_{\dot{\alpha }}),`$ (3.2) where $`c_0`$ is an arbitrary constant and $`\eta _\alpha `$ and $`\overline{\eta }_{\dot{\alpha }}`$ are complex conjugated constant spinor parameters. Substitution to (3.1) leads to the following result upon evaluation of the elementary Gaussian integrals $$C(y,\overline{y}|x)=c_0\lambda ^2z\mathrm{exp}i\left[(\lambda y_\alpha \overline{y}_{\dot{\beta }}+\eta _\alpha \overline{\eta }_{\dot{\beta }})x^{\alpha \dot{\beta }}+\lambda ^{\frac{1}{2}}\sqrt{z}(y^\alpha \eta _\alpha +\overline{y}^{\dot{\alpha }}\overline{\eta }_{\dot{\alpha }})\right].$$ (3.3) Taking into account the identification of the particular components of the expansion (1.29) explained in the section 1.2 one identifies the matter fields and higher spin Weyl tensors with $$C_{\alpha _1\mathrm{}\alpha _n}(x)=\lambda ^{\frac{n}{2}2}\frac{}{y^{\alpha _1}}\mathrm{}\frac{}{y^{\alpha _n}}C(y,\overline{y}|x)|_{y=\overline{y}=0}$$ (3.4) and $$\overline{C}_{\dot{\alpha }_1\mathrm{}\dot{\alpha }_n}(x)=\lambda ^{\frac{n}{2}2}\frac{}{\overline{y}^{\dot{\alpha }_1}}\mathrm{}\frac{}{\overline{y}^{\dot{\alpha }_n}}C(y,\overline{y}|x)|_{y=\overline{y}=0}.$$ (3.5) From (3.3) one gets $$C_{\alpha \mathrm{}\alpha _{2s}}(x)=c_0z^{s+1}\eta _{\alpha _1}\mathrm{}\eta _{\alpha _{2s}}\mathrm{exp}ik_{\gamma \dot{\beta }}x^{\gamma \dot{\beta }}$$ (3.6) and $$\overline{C}_{\dot{\alpha }\mathrm{}\dot{\alpha }_{2s}}(x)=c_0z^{s+1}\overline{\eta }_{\dot{\alpha }_1}\mathrm{}\overline{\eta }_{\dot{\alpha }_{2s}}\mathrm{exp}ik_{\gamma \dot{\beta }}x^{\gamma \dot{\beta }},$$ (3.7) where $$k_{\alpha \dot{\beta }}=\eta _\alpha \overline{\eta }_{\dot{\beta }}.$$ (3.8) The expression (3.8) is the standard twistor representation for an arbitrary light-like vector $`k_{\alpha \dot{\beta }}`$ in terms of spinors. By construction, the formulae (3.6) and (3.7) describe solutions of the free equations of motion of massless fields of arbitrary spin in $`AdS_4`$. In the flat space $`\lambda 0`$ (i.e., $`z1`$) these formulae describe usual flat space plane waves. We therefore interpret the solution (3.3) as describing $`AdS_4`$ ”plane waves”. Let us note that, as expected, these solutions tend to zero at the boundary of $`AdS_4`$ $`z=0`$. ## 4 Higher Spin Potentials Once the 0-forms $`C`$ are found one can in principle solve the equation (1.25) for the gauge potentials $`w_1`$ modulo gauge transformation $$\delta w_1=dϵw_0ϵ+ϵw_0.$$ (4.1) where $`w_0`$ is the background $`AdS_4`$ gauge (gravitational) field (2.1) and $`ϵ(y,\overline{y}|x)`$ is an arbitrary gauge parameter. This problem is analogous to the reconstruction of the electromagnetic potential via the field strength or the metric tensor via the Riemann tensor. Since the equation (1.25) is formally consistent it admits some solution. Remarkably, this solution can also be found explicitly for the general gauge function $`g(y,\overline{y}|x)`$ and $`C_0(y,\overline{y})`$. The systematic derivation can be obtained with the help of a more sophisticated technics developed in for the analysis of the nonlinear problem. Here we announce the final result $`w_1(y,\overline{y}|x)`$ $`=`$ $`{\displaystyle \frac{1}{32\pi ^4}}(\omega _0{}_{}{}^{\alpha \beta }{\displaystyle \frac{}{y^\beta }}+\lambda h_0{}_{}{}^{\alpha \dot{\beta }}{\displaystyle \frac{}{\overline{y}^{\dot{\beta }}}}){\displaystyle _0^1}sds{\displaystyle _{\mathrm{}}^{\mathrm{}}}d^2ud^2vd^2\overline{u}d^2\overline{v}[(y_\alpha +2v_\alpha )`$ (4.2) $`\times g^1((1s)y+u,\overline{y}+\overline{u}|x)C_0((1s)y+v+u,\overline{y}+\overline{u}+\overline{v})`$ $`\times g((1s)y+(12s)v,\overline{y}+\overline{v}|x)\mathrm{exp}i(uv+\overline{u}\overline{v})]+c.c.,`$ leaving the details of its derivation to a more detailed future publication . Note that this expression admits no simple interpretation in terms of the star-product (1.21). This happens because it is derived as some projection from a larger star-product algebra. However, one can check directly that (4.2) solves (1.25). To this end one inserts this formula into the left hand side of (1.25) using the equations (1.26). Then one observes that the resulting terms in the integrand sum up to a total derivative in the integration variable $`s`$. The contribution at $`s=0`$ vanishes because of the factor of $`s`$ in the integration measure, while the contribution at $`s=1`$ reproduces the right hand side of (1.25). Note that only the terms independent of the Lorentz connection contribute. The computation sketched above is however tedious enough. Insertion of the expression (3.2) into (4.2) leads to the following result $`w_1={\displaystyle \frac{1}{2}}zc_0(\omega ^{\alpha \beta }{\displaystyle \frac{}{y^\beta }}+\lambda h^{\alpha \dot{\gamma }}{\displaystyle \frac{}{\overline{y}^{\dot{\gamma }}}}){\displaystyle _0^1}𝑑s{\displaystyle \frac{s}{(s+(1s)\sqrt{z})^3}}`$ $`\times (y_\alpha +\lambda ^{1/2}(1+\sqrt{z})\eta _\alpha +\lambda ^{1/2}x_{\alpha \dot{\beta }}\overline{\eta }^{\dot{\beta }}\lambda \overline{y}^{\dot{\beta }}x_{\alpha \dot{\beta }})`$ $`\times \mathrm{exp}i{\displaystyle \frac{\lambda ^{1/2}}{s+(1s)\sqrt{z}}}[(1s)(\eta _\alpha y^\alpha +\overline{\eta }_{\dot{\alpha }}\overline{y}^{\dot{\alpha }}+\lambda x^{\alpha \dot{\beta }}(y_\alpha \overline{\eta }_{\dot{\beta }}+\eta _\alpha \overline{y}_{\dot{\beta }}))`$ $`+s(\sqrt{z}\overline{\eta }_{\dot{\alpha }}\overline{y}^{\dot{\alpha }}\lambda ^{1/2}x^{\alpha \dot{\beta }}\eta _\alpha \overline{\eta }_{\dot{\beta }})]+c.c.`$ (4.3) Let us note that the quantity $`s+(1s)\sqrt{z}`$ is strictly positive in the region $`z>0`$ for any $`0s1`$. Therefore, the right hand side of the formula (4) is some entire function in the auxiliary spinor variables $`y`$ and $`\overline{y}`$. This means that it gives rise to the well-defined gauge potentials associated with the coefficients of the expansion (1.28) everywhere in the north pole chart. The south pole chart can be analyzed analogously. In the limit $`z0`$ the expression (4) acquires singularity at $`s=0`$. ## 5 Conclusion It is demonstrated that the general formalism developed originally for the formulation of the interacting higher spin gauge theories allows one to write down explicit solutions of the field equations in terms of the star-product algebras in auxiliary spinor variables. The true reason for this is that the formalism of star-product algebras makes infinite-dimensional higher spin symmetries explicit thus allowing one to formulate the field equations as certain covariant constancy conditions. For simplicity, in this paper we have focused on the most symmetric case with gravitational fields possessing explicit Lorentz symmetry. The developed approach can be applied in other coordinate systems. We believe that it will have a wide area of applicability and can be extended to dynamical systems in various dimensions, black hole and brane backgrounds as well as to the superspace. ## Acknowledgments The authors are grateful to R.Metsaev, S.Konstein and D.Sorokin for useful conversations. This research was supported in part by INTAS, Grant No.96-0308 and by the RFBR Grant No.99-02-16207. ## Appendix. Notation Following to , we use conventions with upper(lower) indices denoted by the same letter subject to symmetrization prior possible contractions with lower(upper) indices denoted by the same letter. With these conventions only a number of indices within any symmetrized group is important. This number is often indicated in brackets. A maximal possible number of lower and upper indices denoted by the same letter is supposed to be contracted. We choose mostly minus flat Minkowsi metric $`\eta ^{nm}=\{1,1,1,1\}`$. $`_{\underset{¯}{n}}=\frac{}{x^{\underset{¯}{n}}}`$. Two-component spinor indices are raised and lowered according to: $$A_\alpha =ϵ_{\beta \alpha }A^\beta A^\alpha =ϵ^{\alpha \beta }A_\beta ,$$ where $`e_{12}=e^{12}=1`$. Sigma matrices are expressed in terms of Pauli matrices according to $$\sigma _{\alpha \dot{\beta }}^m=(I,\sigma ^i)$$ $$\sigma _{\alpha \dot{\beta }}^n\sigma ^{m\alpha \dot{\beta }}=2\eta ^{nm}.$$ Also the following conventions are used $$x_{\alpha \dot{\beta }}=\sigma _{\alpha \dot{\beta }}^ax_a,x^2=x_ax^a=\frac{1}{2}x_{\alpha \dot{\beta }}x^{\alpha \dot{\beta }}.$$ $$\delta (n)=\{\begin{array}{cc}1,n=0\hfill & \\ 0,n0\hfill & \end{array},$$ $$\theta (n)=\{\begin{array}{cc}1,n0\hfill & \\ 0,n<0\hfill & \end{array}.$$
warning/0001/cond-mat0001034.html
ar5iv
text
# A hydrodynamic approach to the Bose-Glass transition ## Abstract Nonlinear hydrodynamics is used to evaluate disorder-induced corrections to the vortex liquid tilt modulus for finite screening length and arbitrary disorder geometry. Explicit results for aligned columnar defects yield a criterion for locating the Bose glass transition line at all fields. , thanks: Partially supported by the National Science Foundation through grants No. DMR97-30678 and DMR98-05818. The vortex phase diagram of high temperature superconductors shows a rich diversity of phases, including vortex lattices, liquids and glasses. Novel types of glasses are also possible because of pinning in disordered samples (for a review, see Refs. ). Much progress in the theoretical understanding of vortex behavior at low fields has been made by employing the formal analogy of the statistical mechanics of $`(2+1)`$-dimensional directed lines with the quantum mechanics of $`2d`$ bosons. In this mapping, the vortex lines traversing the sample along the direction $`z`$ of the external field, $`𝐇_0=\widehat{𝐳}H_0`$, correspond to the imaginary-time world lines of the quantum particles. The thickness, $`L,`$ of the superconducting sample is the inverse temperature, $`\beta _B\mathrm{}`$, of the bosons, the vortex line tension, $`\stackrel{~}{ϵ}_1`$, represents the boson mass, $`m`$, and thermal fluctuations in the vortex state $`k_BT`$ map onto quantum fluctuations $`\mathrm{}`$ in the boson system. This “boson mapping” has been particularly useful for understanding the properties of vortices in superconducting samples with aligned damage tracks from heavy-ion irradiation. This type of disorder yields a low-temperature Bose-glass phase where every vortex is trapped on a columnar defect and vortex pinning is strongly enhanced. The transition at $`T_{BG}`$ from the entangled vortex liquid to the Bose-glass is continuous and is signalled by the vanishing of the linear resistivity and of the inverse tilt modulus, $`c_{44}^1`$. The drawback of the boson mapping in the form used by Nelson and coworkers is that intervortex interactions are assumed to be strictly local in $`z`$, the magnetic field direction. This restricts the application of the results to low fields. It also renders problematic the evaluation of the tilt modulus $`c_{44}`$, which measures the linear response to a transverse tilting field $`𝐇_{}`$ normal to $`𝐇_0`$. The long-wavelength tilt modulus of the vortex liquid can be written as $`c_{44}=c_{44}^v+B^2/(4\pi )`$, where $`c_{44}^v=n_0\stackrel{~}{ϵ}_1`$, with $`n_0=B/\varphi _0`$ the vortex density, is the single vortex part and the second term represents a compressive contribution. By assuming local interactions along $`z`$, the compressive part of $`c_{44}`$, which dominates at high fields, is neglected entirely. With this approximation, the mathematical analogy between vortices and $`2d`$ bosons can be exploited further to show that the inverse tilt modulus maps onto the superfluid density, $`n_s`$, of the bosons, with $$\frac{n_0^2}{c_{44}^v}=\frac{n_s}{m}.$$ (1) The transition as $`TT_{BG}^+`$ from the entangled vortex liquid to the Bose glass corresponds then to the transition from a boson superfluid to a localized normal phase of bosons. Täuber and Nelson have evaluated perturbatively the reduction of $`n_s`$ due to various types of disorder. They also showed that for aligned columnar disorder such a perturbative calculation yields a useful criterion for locating the transition line $`T_{BG}(B)`$ at low fields. Larkin and Vinokur argued that a generalization of Eq. (1) that incorporates the compressive part of $`c_{44}`$ can be obtained by using the nonlocal mapping on vortex lines onto $`2d`$ bosons introduced some time ago by Feigel’man and collaborators. These authors showed that the fully nonlocal London model of interacting vortex lines can be mapped to a system of $`2d`$ charged bosons coupled to a massive photon field. The duality between vortices and bosons translates into $$c_{44}=\frac{B^2}{4\pi }+\frac{n_0^2\stackrel{~}{ϵ}_1}{n_s},$$ (2) where $`n_s`$ is defined here in terms of the polarization function of the fictitious gauge field. Near a Bose glass transition where $`n_s0`$ the second term of Eq. (2) dominates and results in the divergence of the tilt modulus. Feigel’man et al. carried out a perturbative calculation of the reduction of $`n_s`$ from intervortex interactions in a clean material (in the limit $`\lambda \mathrm{}`$) , but the calculation of the reduction of $`n_s`$ from disorder in the context of the charged boson model is cumbersome and does not provide much physical insight. An alternative approach for modeling interacting vortex arrays is hydrodynamics, which has proved very useful to describe the long wavelength properties of flux-line liquids at high fields. Hydrodynamics provides a physically transparent formulation that naturally incorporates the nonlocality of the intervortex interaction. The goal of this note is to show how hydrodynamics can be used in a transparent way to evaluate disorder-induced corrections to the wave-vector dependent tilt modulus for finite values of the screening length $`\lambda `$ and arbitrary disorder geometry. The connection of the hydrodynamic formulation to the charged boson formalism will also be discussed. Explicit results are presented for aligned columnar defects. The hydrodynamic free energy of the flux-line liquid is a functional of two coarse-grained fields, the local areal density of vortices, $`n(𝐫)`$, and the tilt field, $`𝐭(𝐫)`$, which measures the local deviation of a volume of flux liquid from the direction of the external field, $`𝐇_0`$. It is given by $`F=F_0+F_D`$, where $`F_0`$ is the free energy in the absence of disorder and $`F_D`$ describes the coupling to quenched defects. The disorder-free contribution is $`F_0={\displaystyle _𝐫}{\displaystyle \frac{\stackrel{~}{ϵ}_1|𝐭(𝐫)|^2}{2n(𝐫)}}+{\displaystyle \frac{1}{2n_0^2}}{\displaystyle _𝐪}\left\{c_{44}^{c0}(𝐪)|𝐭(𝐪)|^2+c_{11}^0(𝐪)|\delta n(𝐪)|^2\right\},`$ (3) where $`\delta n(𝐫)=n(𝐫)n_0`$. The density and tilt field are related by the familiar constraint that flux lines cannot start nor stop inside the sample, $$_zn(𝐫)+_{}𝐭(𝐫)=0.$$ (4) The bare elastic constants, $`c_{44}^{c0}`$ (the compressive part of the tilt modulus) and $`c_{11}^0`$ (the bare compressional modulus), are determined by the intervortex interaction. For isotropic materials they are simply $`c_{44}^{c0}(𝐪)=c_{11}^{c0}(𝐪)=(B^2/4\pi )/(1+q^2\lambda ^2)`$, with $`\lambda `$ the screening length. Quenched disorder from material defects couples to the flux-line density and gives a contribution $$F_D=_𝐫V_D(𝐫)\delta n(𝐫).$$ (5) The random potential $`V_D(𝐫)`$ is taken to be Gaussian, statistically homogeneous, and isotropic in the $`xy`$ plane so that $$\overline{\delta V_D(𝐪)\delta V_D(𝐪^{})}=\mathrm{\Delta }(q_{},q_z)\mathrm{\Omega }\delta _{𝐪_{}+𝐪_{}^{}{}_{}{}^{},0},$$ (6) where $`\delta V_D(𝐫)=V_D(𝐫)\overline{V_D(𝐫)}`$, and the overbar represents the disorder average. The correlator $`\mathrm{\Delta }(q_{},q_z)`$ depends on the geometry of disorder and will be specified below for the case of interest. The hydrodynamic free energy $`F_0`$ goes beyond the Gaussian hydrodynamic model commonly used in the vortex literature as the first term on the right hand side of Eq. (3), describing the “kinetic energy” part of the vortex interaction, incorporates non-Gaussian terms. In a recent publication we showed that this non-Gaussian hydrodynamics is precisely equivalent to the charged boson model of Feigel’man and coworkers. Such a “nonlinear hydrodynamics” was used in Ref. to evaluate perturbatively the enhancement of $`c_{44}`$ from interactions in a clean material. It was shown there that nonlocality is crucial to yield a correction to $`c_{44}`$ that remains finite for $`L\mathrm{}`$. The same model is used here to evaluate corrections to $`c_{44}`$ from disorder. In particular, for aligned columnar defects our calculation yields a criterion for locating the transition line $`T_{BG}(B)`$ at all fields. The tilt modulus, $`c_{44}`$, can be expressed in terms of the tilt field autocorrelation function, $`T_{ij}(𝐪)=\overline{t_i(𝐪)t_j(𝐪)}`$, where the brackets denote a thermal average with weight $`\mathrm{exp}[(F_0+F_D)/k_BT]`$, to be carried out subject to the constraint (4). It is given by $$\frac{n_0^2k_BT}{c_{44}}=\underset{q_z0}{lim}\underset{q_{}0}{lim}P_{ij}^T(\widehat{𝐪}_{})T_{ij}(𝐪),$$ (7) with $`P_{ij}^T(\widehat{𝐪}_{})=\delta _{ij}\widehat{q}_i\widehat{q}_j`$ the familiar transverse projection operator, and $`\widehat{𝐪}_{}=𝐪_{}/q_{}`$. When only quadratic terms in the fluctuations $`\delta n`$ and $`𝐭`$ are retained, the resulting Gaussian free energy, $`F_G`$, is $`F_G=F_{0G}+F_D`$, where $`F_{0G}`$ is obtained from Eq. (3) by the replacement $`|𝐭|^2/n(𝐫)|𝐭|^2/n_0`$. The first two terms on the RHS of Eq. (3) can then be combined to define a bare tilt modulus $`c_{44}^0=n\stackrel{~}{ϵ}_1+c_{44}^{0c}`$ and Eq. (7) is simply an identity. To Gaussian order there is no coupling between the density field and the transverse part of the tilt field that determines $`c_{44}`$. As a result, disorder, which couples to the density, does not change the tilt modulus. Non-Gaussian terms in the hydrodynamic free energy of Eq. (3) do introduce a coupling between transverse tilt and density, yielding a renormalization of the tilt modulus. Specifically, the free energy is written as $`F=F_G+\delta F`$, with $$\delta F=_𝐫\frac{\stackrel{~}{ϵ}_1|𝐭(𝐫)|^2}{2n_0}\frac{\delta n(𝐫)}{n(𝐫)}.$$ (8) By treating the non-Gaussian part $`\delta F`$ perturbatively, we obtain $$\frac{n_0^2}{c_{44}}=\frac{n_0^2}{c_{44}^0}\left[1\frac{\stackrel{~}{ϵ}_1/n_0}{c_{44}^0/n_0^2}\frac{n_n}{n_0}\right],$$ (9) with $$\frac{n_n}{n_0}=\frac{1}{n_0^2}_𝐪\left\{\left[1\frac{\overline{|𝐭(𝐪)|^2_G}}{2n_0\stackrel{~}{ϵ}_1k_BT}\right]\overline{|\delta n(𝐪)|^2_G}\frac{|\overline{\widehat{𝐪}_{}𝐭(𝐪)\delta n(𝐪)_G}|^2}{2n_0\stackrel{~}{ϵ}_1k_BT}\right\}.$$ (10) The brackets $`\overline{\mathrm{}_G}`$ in Eq. (10) denote a thermal average with weight $`\mathrm{exp}[F_G/k_BT]`$, subject to the constraint (4), followed by the average over quenched disorder. The correction has been denoted by $`n_n`$ because it has the suggestive interpretation of a normal-fluid density. Notice that its expression is formally identical to that obtained in in the absence of disorder. Disorder only enters here in the expressions for the Gaussian correlators, which can be found for instance in Ref. . A more general expression for the renormalized tilt modulus at finite wave vector can also be obtained by the same methods. In the dilute limit, $`\lambda <<a`$, with $`a=1/\sqrt{n_0}`$ the intervortex spacing, the $`z`$-nonlocality becomes insignificant and Eq. (10) reduces to the result obtained for instance by Täuber and Nelson using the local boson mapping. We now focus on the case of correlated disorder created by heavy ion irradiation yielding rectilinear damage tracks aligned with the field $`𝐇_0`$. This corresponds to $`\mathrm{\Delta }(𝐪)=\mathrm{\Delta }L\delta _{q_z,0}`$, where $`\mathrm{\Delta }U_0^2b^4/d^2`$, with $`U_0`$ the depth of the pinning potential well, $`b`$ its radius, and $`d`$ the average distance between columnar defects. In this case the disorder contribution to $`n_n`$ from disorder takes a particularly simple form, given by $$\frac{n_n^D}{n_0}=\frac{\mathrm{\Delta }}{2\stackrel{~}{ϵ}_1^2}_𝐪_{}\frac{q_{}^4}{[ϵ_B(q_{})/k_BT]^4}G(q_{}),$$ (11) where $`G(q_{})=1+c_{44}^{c0}(q_{},q_z=0)/c_{44}^0(q_{},q_z=0)`$, and $`ϵ_B(q_{})`$ corresponds to the spectrum of $`2d`$ bosons with screened interactions, $$\frac{ϵ_B(q_{})}{k_BT}=\sqrt{\frac{n_0q_{}^2V(q_{})}{\stackrel{~}{ϵ}_1}+\left(\frac{k_BTq_{}^2}{2\stackrel{~}{ϵ}_1}\right)^2},$$ (12) with $`V(q_{})=V_0/(1+\stackrel{~}{\lambda }_{}^2q_{}^2)`$ the screened boson interaction and $`V_0=\varphi _0^2/4\pi `$. The function $`G(q_{})`$ depends only weakly on $`q_{}`$, varying between $`1`$ and $`2`$, and is the only manifestation of the $`z`$-nonlocality of the intervortex interaction. If we let $`G(q_{})=1`$ and negelct the screening, i.e., $`V(q_{})=V_0`$, our result becomes identical to the disorder-induced renormalization of $`1/c_{44}`$ obtained by others. As expected, for the case of aligned columnar defects the $`z`$-nonlocality is not very important. The screening of the boson interaction incorporated in $`V(q_{})`$ is, however, important at high vortex densities. This becomes clear by plotting the boson spectrum as a function of wave vector, as shown in Fig. 1. At high density the screening of the interaction yields a flat region in the spectrum. The wave vector integral in Eq. (11) is dominated by $`q_{}k_{BZ}`$. At low density the main contribution to the integral comes therefore from a region where the spectrum is phonon-like, i.e., $`ϵ_B(q_{})/(k_BT)q_{}\sqrt{n_0\stackrel{~}{V}(q_{},T)/\stackrel{~}{ϵ}_1}`$, as obtained from the theory of uncharged boson superfluids. At high density, however, the main contribution to the integral comes from a region of wave vectors where the spectrum is plasmon-like, i.e., $`ϵ_B(q_{})/(k_BT)\sqrt{4\pi n_0}`$, as appropriate for a charged superfluid. This was actually recognized earlier by Larkin and Vinokur, who in order to use the results of the local boson theory at high fields proposed an ad hoc formula for the Bogoliubov spectrum that captures the dense liquid physics by interpolating between these two limits. Hydrodynamics naturally provides a simple and unified description of the behavior of flux-line liquids in the presence of disorder that applies over a wide range of densities. As discussed in Ref. , the perturbation theory breaks down at high fields and temperatures ($`k_BT/\stackrel{~}{ϵ}_1a>1`$). An approximate expression for the tilt moduls of the form proposed by Larkin and Vinokur can be obtained by treating the nonlinearities in a mean field approximation, as done by Feigel’man and collaborators for the charged boson superfluid. The tilt-tilt autocorrelation function is evaluated by retaining only transverse fields and using a Hartree-type approximation for the non-Gaussian terms. This yields $$\frac{n_0^2}{c_{44}}\frac{1}{c_{44}^{c0}/n_0^2+\stackrel{~}{ϵ}_1/(n_0n_n)},$$ (13) with $`n_n`$ given approximately by Eq. (10). The Bose glass transition line $`B_{BG}(T)`$ can now be evaluated as the locus of points in the $`(B,T)`$ plane where $`n_n^D/n_0=1`$, corresponding to $`c_{44}\mathrm{}`$. In general the integral in Eq. (11) has to be evaluated numerically. The resulting $`B_{BG}(T)`$ phase line is shown in Fig. 2. Analytical results can be obtained in the limit of low and high field. For a dilute liquid $`(\lambda >>a)`$, the screening of the interaction can be neglected and the Bogoliubov spectrum can be approximated as linear in $`q_{}`$. The integral can then be evaluated and gives $$B_{BG}(T)B_\varphi \frac{\mathrm{ln}(1/n\lambda ^2)}{8\pi ^2}\left(\frac{T^{}}{T}\right)^4\left(\frac{T_cT}{T}\right)^4,$$ (14) where $`B_\varphi =\varphi _0/d^2`$ is the matching field and $`k_BT^{}=b(\stackrel{~}{ϵ}_1U_0)^{1/2}`$ is a characteristic pinning energy scale, in agreement with earlier work by other authors. In dense liquids ($`\lambda >>a`$), the Boguliobov spectrum is approximately $`q_{}`$-independent. Again, the transition line can be obtained analytically, with $$B_{BG}(T)B_\varphi \frac{U_0b^2}{64\pi \stackrel{~}{ϵ}_1d^2}\left(\frac{T^{}}{T}\right)^6\left(\frac{T_cT}{T}\right)^6,$$ (15) which agrees with the earlier result of Larkin and Vinokur. Hydrodynamics provides a physically transparent framework for evaluating disorder-induced corrections to the elastic constants of the vortex liquid. It naturally incorporates the full nonlocality of the intervortex interaction and it allows us to compute finite-wavevector elastic constants for arbitrary disorder geometry. In particular, the nonlocality in the field ($`z`$) direction is expected to be important for splayed columnar defects, where terms coupling disorder directly to the tilt field may need to be incorporated in the coarse-grained theory. This work will be presented elsewhere. Figure Captions
warning/0001/math-ph0001003.html
ar5iv
text
# 1 Introduction ## 1 Introduction Analytic factorization for matrix valued functions, or the matrix Riemann-Hilbert problem, appears in problems of exactly integrable dynamical systems as well as in problems of elasticity and diffraction. In the context of exactly integrable models, the factorization problem for the R-matrix associated to an affine Lie algebra is the same as the matrix Riemann-Hilbert problem . In this case exists a general formula for the solution in terms of Riemann theta functions , derived from algebro-geometric methods of finite-band integration theory . This construction is based on the fact that the time evolution linearizes on the Jacobian of the spectral curve, and consequently one can express the solution in terms of the Baker-Akheizer functions, which are eigen-vectors of the Lax matrix. Various integrable dynamical systems can be formulated via such R-matrix , and most typical are the spinning tops. In this paper we derive the explicit expressions for the case of an $`SO(2)`$ spinning top, which is a simple but a non-trivial example of the two-by-two matrix factorization problem. This is done in order to illustrate the general theta-function formula from , and also in order to be able to compare to alternative methods of matrix factorization. The neccesity for alternative methods of factorization comes from the fact that the theta-function formula becomes difficult to use in the case when the spectral curve is non-hyperelliptic, which is mostly the case. Also there are cases when only a non-canonical factorization is possible, for which the theta-function formula is not valid. We illustrate this on the example of a non-compact extension of the $`SO(2)`$ spinning top. ## 2 Lax formulation and matrix Rieman-Hilbert problem Consider a rotator with a unit radius and angle $`\varphi `$ in a potential $`V(\varphi )=a^2\mathrm{cos}2\varphi `$, where $`a`$ is a real constant. The Hamiltonian is given by $$H=\frac{1}{2}l^2a^2\mathrm{cos}2\varphi ,$$ (1) where $`l=\dot{\varphi }=d\varphi /dt`$. As a one-dimensional Hamiltonian system, it is integrable, and it is the $`n=2`$ case of an $`n`$-dimensional integrable spinning top system given by the following $`n`$-by-$`n`$ Lax matrix $$L=𝐚\lambda +𝐥+𝐬\lambda ^1=𝐚\lambda +𝐥+R_\varphi ^1𝐟R_\varphi \lambda ^1,$$ (2) where $`\lambda `$ is a complex number called the spectral parameter, $`𝐚`$ and $`𝐟`$ are fixed symmetric matrices, $`𝐥`$ is antisymmetric and traceless matrix, while $`R_\varphi `$ is an $`SO(n)`$ matrix. Formula (2) gives a mapping from the phase space of the $`n`$-dimensional spinning top $`T^{}SO(n)=(R_\varphi ,𝐥)`$ to a Hamiltonian orbit in the affine $`gl(n)`$ algebra $`\widehat{g}`$ defined as $$\widehat{g}=\underset{j𝐙}{}g_j\lambda ^j,g_jgl(n).$$ Without loss of generality one can take that the $`𝐚`$ and $`𝐟`$ matrices are diagonal, and for simplicity we take $`𝐚=𝐟`$ and $`tr𝐚=0`$. The last condition implies $`trL=0`$, and this means that we are restricting $`\widehat{g}`$ to the affine $`sl(n)`$ sub-algebra. The Hamiltonian (1) is the $`n=2`$ case of the $`SO(n)`$ Hamiltonian $$H=\frac{1}{2}Restr(\lambda ^1L^2)=\frac{1}{2}tr𝐥^2tr(𝐚R_\varphi ^1𝐟R_\varphi ).$$ (3) The corresponding Lax equations follow from the R-matrix formulation , and they are given by $$\dot{L}=[L,M_\pm ],M_\pm =\frac{1}{2}(R\pm 1)d\phi (L)=\pm P_\pm L,$$ (4) where $`\phi (L)=\frac{1}{2}tr(L^2)`$ is an invariant function on $`\widehat{g}`$ and $`R=P_+P_{}`$ is the R-matrix. $`P_\pm `$ are the projectors onto $`\widehat{g}_\pm `$ subalgebras, which are defined as $$\widehat{g}_+=\underset{j0}{}g_j\lambda ^j,\widehat{g}_{}=\underset{j<0}{}g_j\lambda ^j,$$ (5) so that $`\widehat{g}=\widehat{g}_+\widehat{g}_{}`$, and therefore $$M_+=P_+L=𝐥+𝐚\lambda ,M_{}=P_{}L=𝐬\lambda ^1.$$ (6) The equations (4) are solved by $$L(\lambda ,t)=g_\pm (\lambda ,t)L(\lambda )g_\pm ^1(\lambda ,t)$$ (7) where $`L(\lambda )=L(\lambda ,t=0)`$ and $`g_\pm (\lambda ,t)`$ are solutions of the following matrix factorization problem $$e^{tL(\lambda )}=g_+^1(\lambda ,t)g_{}(\lambda ,t),$$ (8) such that $`g_+`$ is analytic for $`\lambda \mathrm{}`$ and $`g_{}`$ is analytic for $`\lambda 0`$. This is a Riemann-Hilbert problem for $`e^{tL(\lambda )}`$ and a curve which encircles the $`\lambda =0`$ point in the complex plain. Solution of the factorization problem (8) exists for $`t`$ that satisfies the Gohberg-Feldman bound, which is satisfied for sufficiently small times . Otherwise one has a non-canonical factorization $$e^{tL(\lambda )}=g_+^1(\lambda ,t)D(\lambda )g_{}(\lambda ,t),$$ (9) where $`D(\lambda )=diag(\lambda ^{k_1},\mathrm{},\lambda ^{k_n})`$. ## 3 Algebro-geometric factorization One way to solve the problem (8) is to use the connection with Riemann surfaces. The spectral curve is given by $$det(L(\lambda )\mu )=0,$$ (10) and it is time-independent. Its compact model defines the corresponding Riemann surface $`\mathrm{\Gamma }`$. Consider the line bundle $`E_L(p)𝐂^n`$ associated to the eigen-vectors of $`L`$ $$L(p)v(p)=\mu (p)v(p),p\mathrm{\Gamma }.$$ (11) Its time-evolution is linear on $`Jac\mathrm{\Gamma }`$, and it is given by $`E_t=EF_t`$, where $`F_t`$ is the line bundle on $`\mathrm{\Gamma }`$ determined by the transition function $`\mathrm{exp}t\mu (p)`$ with respect to the covering $`\{U_+,U_{}\}`$, where $`U_\pm =\{p|\lambda ^{\pm 1}(p)\mathrm{}\}`$. In order to solve (8), one needs to consider the holomorphic sections of the dual bundle $`E_L^{}`$, since the bundle $`E_L`$ does not have holomorphic sections. Denote these sections as $`\psi (p)`$, then they satisfy $$L(p)\psi (p)=\mu (p)\psi (p),$$ (12) so that $`\psi _\pm (p,t)=g_\pm (\lambda (p),t)\psi (p)`$ are eigenvectors of $`L(\lambda ,t)`$ which are regular in $`U_\pm `$. Hence $$g_\pm (\lambda ,t)=\mathrm{\Psi }_\pm (\lambda ,t)\mathrm{\Psi }_\pm ^1(\lambda ,t=0)$$ (13) where $`\mathrm{\Psi }_\pm (\lambda ,t)_{jk}=\psi ^j(p_k,t)`$. Hence if one knows the sections $`\psi _\pm `$, one can solve the factorization problem (8) as well as the dynamics. These sections can be represented via meromorphic functions $`\phi _\pm `$ on $`U_\pm `$ as $$\phi _\pm (p,t)=\frac{\psi _\pm (p,t)}{s_\pm (p)},$$ (14) where $`s_\pm (p)`$ are sections of $`E_L^{}`$. These functions are called Baker-Akheizer (BA) functions and satisfy $$(\phi _\pm )D^{}$$ (15) where $`D^{}`$ is a divisor of degree $`g+n1`$, where $`g`$ is the genus of $`\mathrm{\Gamma }`$. Given (14) and (15) and the following properties of $`\psi _\pm `$ $$\frac{d\psi _\pm (p,t)}{dt}=M_\pm (\lambda ,t)\psi _\pm (p,t)$$ (16) and $$\psi _+(p,t)=e^{\mu (p)t}\psi _{}(p,t),$$ (17) one can construct the BA functions in terms of the theta functions . Let $`P_{\mathrm{}}=_{j=1}^nP_j`$ be the divisor of poles of $`\lambda `$, let $`p_0`$ be a fixed point in $`\mathrm{\Gamma }`$ and let $`D`$ be a divisor of degree $`g`$ such that the line bundle $`E_L^{}`$ is associated to the divisor $$D^{}=P_{\mathrm{}}+Dp_0.$$ (18) Let $`\omega _j`$ be a set of normalized holomorphic Abelian differentials on $`\mathrm{\Gamma }`$ $$_{a_j}\omega _k=2\pi i\delta _{jk},_{b_j}\omega _k=B_{jk},$$ (19) where $`a_j`$ and $`b_j`$, $`j=1,\mathrm{},g`$ is a basis of homology cycles and $`B`$ is the period matrix. Note that the Riemann theta function is defined on $`𝐂^g`$ by quasi-periodicity conditions $$\theta (z+2\pi in)=\theta (z),\theta (z+Bn)=e^{\frac{1}{2}(Bn,n)(z,n)}\theta (z),n𝐙^g$$ (20) and can be written as $$\theta (z)=\underset{n𝐙^g}{}\mathrm{exp}\{\frac{1}{2}(Bn,n)+(z,n)\},$$ (21) where $`(X,Y)=_{j=1}^gX_jY_j`$. The theta function can be considered as a section on $`Jac\mathrm{\Gamma }=𝐂^g/(2\pi in+Bm)`$. Let $`A(p)`$ be the Abel transform $$A(p)=_{p_0}^p\omega ,$$ (22) and let $`d\mathrm{\Omega }`$ be an Abelian differential of the second kind, normalized by $$_{a_j}𝑑\mathrm{\Omega }=0,$$ (23) which is regular in $`U_{}`$ such that $`d\mathrm{\Omega }d\mu `$ is regular in $`U_+`$. Then $$\phi _{}^j(p,t)=\gamma _j(t)e^{t\mathrm{\Omega }(p)}\frac{\theta (A(p)cVt)\theta (A(p)c_0)}{\theta (A(p)c_1)\theta (A(p)c_j)},$$ (24) where $`\mathrm{\Omega }(p)=_{p_0}^p𝑑\mathrm{\Omega }`$, and the constants $`c_k`$ and the function $`\gamma _j(t)`$ are determined from the following requirements (a) $`\phi _{}^j`$ has no discontinuity across the cuts (b) $`\phi _{}^j`$ is subordinate to the divisor $`p_0P_jD`$ (c) $`Res_{P_j}(\lambda ^1\phi _{}^j)=d_j`$ where $`d_j`$ are time-independent constants. Condition (a) is satisfied if $`c+c_0=c_1+c_j`$ and $`V_j=_{b_j}𝑑\mathrm{\Omega }`$. Condition (b) is satisfied if the constants $`c_k`$ are chosen such that the divisor of zeros of $`\theta (A(p)c_0)`$ is $`p_0+\stackrel{~}{D}`$, the divisor of zeros of $`\theta (A(p)c_1)`$ is $`D+\stackrel{~}{D}`$ and the divisor of zeros of $`\theta (A(p)c_j)`$ is $`P_j+\stackrel{~}{D}`$, for some divisor $`\stackrel{~}{D}`$. Condition (c) serves to determine the functions $`\gamma _j(t)`$ and it follows from (16) when $`\lambda \mathrm{}`$. The following identity will be useful for our purposes $$V_j=_{b_j}𝑑\mathrm{\Omega }=\underset{k=1}{\overset{n}{}}Res_{P_k}(\mu \omega _k).$$ (25) ## 4 The SO(2) Case In the case of the rotator, $`𝐚=𝐟=diag(a,a)`$, $$𝐥=\left(\begin{array}{cc}0& l\\ l& 0\end{array}\right),$$ (26) and $$𝐬=a\left(\begin{array}{cc}\mathrm{cos}2\varphi & \mathrm{sin}2\varphi \\ \mathrm{sin}2\varphi & \mathrm{cos}2\varphi \end{array}\right),$$ (27) so that $$L=\left(\begin{array}{cc}a\lambda +a\lambda ^1\mathrm{cos}2\varphi & l+a\lambda ^1\mathrm{sin}2\varphi \\ l+a\lambda ^1\mathrm{sin}2\varphi & a\lambda a\lambda ^1\mathrm{cos}2\varphi \end{array}\right).$$ (28) The spectral curve (10) is then given by $$\mu ^2=a^2\lambda ^22E+a^2\lambda ^2,$$ (29) where $`E=H`$ is the energy. The equation (29) can be transformed into elliptic form $$w^2=\lambda ^4\frac{2E}{a^2}\lambda ^2+1=(\lambda ^2l_+^2)(\lambda ^2l_{}^2),$$ (30) so that the spectral surface is a torus. We take the $`a_1`$ cycle to encircle the $`[l_+,l_{}]`$ cut, while the $`b_1`$ cycle goes from the $`[l_+,l_{}]`$ cut to the $`[l_{},l_+]`$ cut, were it passes through to the second sheet and it goes back to the $`[l_+,l_{}]`$ cut. The normalized holomorphic Abelian differential is given by $$\omega =\frac{2\pi i}{𝒜}\frac{d\lambda }{w},𝒜=_{a_1}\frac{d\lambda }{w},$$ (31) so that the period matrix is given by $$B=_{b_1}\omega =2\pi i\frac{}{𝒜}=2\pi i\tau .$$ (32) The Abelian differential of the second kind $`d\mathrm{\Omega }`$ must satisfy for $`\lambda 0`$ $$d\mathrm{\Omega }(p)=d\mu +O(1)d\lambda =\left[\frac{a}{\lambda ^2}+O(1)\right]d\lambda ,$$ (33) and it is normalized by $$_{a_1}𝑑\mathrm{\Omega }=0.$$ (34) Hence the second-order poles are at $`p=0^+`$ and $`p=0^{}`$ and one can construct $`d\mathrm{\Omega }`$ in terms of the prime forms . Alternatively, from (33) it follows that the corresponding Abelian integral $`\mathrm{\Omega }(p)`$ behaves as $$\mathrm{\Omega }(p)=\pm \frac{a}{\lambda }+O(1),$$ (35) for $`p0^\pm `$, and hence it is a meromorphic function with simple poles at $`0^\pm `$, and with residues $`\pm a`$. A meromorphic function having only simple poles can be specified by the position of the poles and with the values of the residues, and it can be constructed in terms of the theta functions . This implies that $$\mathrm{\Omega }(p)=a\alpha \left[\frac{\theta _1^{}(A(p)A(0^+))}{\theta _1(A(p)A(0^+))}\frac{\theta _1^{}(A(p)A(0^{}))}{\theta _1(A(p)A(0^{}))}\right],$$ (36) where $`\alpha =2\pi i/𝒜`$ and $`\theta _1(z)`$ is the odd Jacobi theta function (40) and $`\theta ^{}(z)=\frac{d\theta (z)}{dz}`$. Also note that the identity (25) gives $$V=2a\frac{2\pi i}{𝒜}=2a\alpha .$$ (37) We take $`D=p_1e_i`$, where $`e_i`$ are the end-points of the brunch-cuts, and $`P_{\mathrm{}}=P_1+P_2=\mathrm{}^++\mathrm{}^{}`$ so that $`c_0`$ $`=`$ $`\pi i+{\displaystyle \frac{1}{2}}B`$ $`c_1`$ $`=`$ $`A(p_1)+\pi i+{\displaystyle \frac{1}{2}}B`$ (38) $`c_j`$ $`=`$ $`A(P_j)+\pi i+{\displaystyle \frac{1}{2}}B,`$ and hence $$c=c_1+c_jc_0=A(p_1)+A(P_j)+\pi i+\frac{1}{2}B.$$ (39) By using $$\theta (z+\pi i+\frac{1}{2}B)=e^{B/8z/2}\theta _1(z),$$ (40) where $`\theta _1(z)=i\theta [\frac{1}{2},\frac{1}{2}](z)`$ is the odd Jacobi theta function, we obtain $$\phi _{}^j(p,t)=\gamma _j(t)e^{t\stackrel{~}{\mathrm{\Omega }}(p)}\frac{\theta _1(A(p)+A(p_1)+A(P_j)+Vt)\theta _1(A(p))}{\theta _1(A(p)+A(p_1))\theta _1(A(p)+A(P_j))},$$ (41) and $`\stackrel{~}{\mathrm{\Omega }}(p)=\mathrm{\Omega }(p)V/2`$. The constants $`\gamma _j(t)`$ can be determined from the condition (c), and for this one needs the expansion of (41) in $`\zeta =\lambda ^1`$ for $`\lambda \mathrm{}`$. The following formulas will be useful $$A(p)=_{p_0}^\mathrm{}^\pm \omega +_\mathrm{}^\pm ^p\omega =A(P_j)z_j$$ (42) and $$z_j=\alpha _j\zeta +O(\zeta ^2)=\pm \alpha \zeta +O(\zeta ^2).$$ (43) One also has for $`pP_j`$ $$\mathrm{\Omega }(p)=\mathrm{\Omega }_j^0+\mathrm{\Omega }_j^1\zeta +O(\zeta ^2),$$ (44) so that when $`pP_j`$ we obtain $`\phi _{}^j(p,t)`$ $`=`$ $`\gamma _j(t)e^{t\stackrel{~}{\mathrm{\Omega }}_j^0}{\displaystyle \frac{\theta _1(A_1+Vt)\theta _1(A_j)}{\theta _1(A_j+A_1)\alpha _j\theta _1^{}(0)}}[\zeta ^1+\alpha _j({\displaystyle \frac{\mathrm{\Omega }_j^1}{\alpha _j}}t+{\displaystyle \frac{\theta _1^{}(A_1+Vt)}{\theta _1(A_1+Vt)}}`$ (45) $`+`$ $`{\displaystyle \frac{\theta _1^{}(A_j)}{\theta _1(A_j)}}{\displaystyle \frac{\theta _1^{}(A_j+A_1)}{\theta _1(A_j+A_1)}}{\displaystyle \frac{\theta _1^{\prime \prime }(0)}{2\theta _1^{}(0)}})+O(\zeta )],`$ where $`A(p_k)=A_k`$. Also when $`pP_j^{}`$, where $`p^{}=(w,\lambda )`$, we have $$\phi _{}^j(p,t)=\gamma _j(t)e^{t\stackrel{~}{\mathrm{\Omega }}_j^{}^0}\frac{\theta _1(A_1+A_jA_j^{}+Vt)\theta _1(A_j^{})}{\theta _1(A_j^{}+A_1)\theta _1(A_j^{}+A_j)}\left[1+O(\zeta )\right].$$ (46) From (45) it follows that $$d_j=\gamma _j(t)e^{t\stackrel{~}{\mathrm{\Omega }}_j^0}\frac{\theta _1(A_1+Vt)\theta _1(A_j)}{\theta _1(A_j+A_1)\alpha _j\theta _1^{}(0)}.$$ (47) As far as the solutions of the equations of motion are concerned, these can be obtained by inserting the $`\lambda \mathrm{}`$ expansions (45) and (46) of the BA functions into (16). Namely, for $`\lambda \mathrm{}`$ we have $$\phi _{}^j=\lambda \psi _{}^j=\lambda \psi _0^j+\psi _1^j+O(1/\lambda )$$ (48) so that (16) gives $$\frac{d\psi _0^j}{dt}=0,\frac{d\psi _1^j}{dt}=𝐬\psi _0^j.$$ (49) By using (49) together with (45) and (46) we get $`\psi _0^1=(d_1,0)^T`$, $`\psi _0^2=(0,d_2)^T`$ and $`\pm \alpha {\displaystyle \frac{d}{dt}}\left[{\displaystyle \frac{\theta _1^{}(A_1+Vt)}{\theta _1(A_1+Vt)}}\right]+\mathrm{\Omega }_1^j`$ $`=`$ $`\pm a\mathrm{cos}2\varphi `$ (50) $`{\displaystyle \frac{d}{dt}}\left[\gamma _j(t)e^{t\stackrel{~}{\mathrm{\Omega }}_j^{}^0}{\displaystyle \frac{\theta _1(A_jA_j^{}+A_1+Vt)\theta _1(A_j^{})}{\theta _1(A_1A_j^{})\theta _1(A_jA_j^{})}}\right]`$ $`=`$ $`ad_j\mathrm{sin}2\varphi .`$ (51) From (50) we obtain $$\mathrm{sin}^2\varphi =\frac{1}{2}(\mathrm{\Omega }_1/a1)\alpha \frac{V}{2a}\frac{d^2}{dz^2}\mathrm{log}\theta _1(z),$$ (52) where $`\mathrm{\Omega }_1=\mathrm{\Omega }_1^1=\mathrm{\Omega }_1^2`$ and $`z=Vt+A_1`$. Due to (25), $`V/2a=\alpha `$ so that (52) can be written as $$\mathrm{sin}^2\varphi =\mathrm{const}.+\alpha ^2\mathrm{}(2a\alpha t+A_1),$$ (53) where $`\mathrm{}`$ is the Weierstrass function. Since $$\mathrm{}(\alpha ^1u|\alpha ^1\omega _1,\alpha ^1\omega _2)=\alpha ^2\mathrm{}(u|\omega _1,\omega _2),$$ (54) where $`\omega _{1,2}`$ are the periods of the Weierstrass function, we get $$\mathrm{sin}^2\varphi =\mathrm{}\left(2at+\frac{A_1}{\alpha }\right)+\mathrm{const}..$$ (55) The expression (55) can be shown to coincide with the solution of the equations of motion obtained from the energy expression $$2E=\dot{\varphi }^22a^2\mathrm{cos}2\varphi .$$ (56) It follows from (56) that $$\stackrel{~}{\omega }t=\pm _0^\varphi \frac{d\varphi }{\sqrt{1k^2\mathrm{sin}^2\varphi }},$$ (57) where $`\stackrel{~}{\omega }=\sqrt{2(E+a^2)}`$, $`k^2=2a^2(E+a^2)^1`$ and $`Ea^2`$. The elliptic integral (57) implies $$\mathrm{sin}\varphi =\mathrm{sn}(\stackrel{~}{\omega }t,k).$$ (58) However, in order to compare to (55), we rewrite the integral (57) as $$2at=\pm _0^{\mathrm{sin}^2\varphi }\frac{dx}{\sqrt{4x(1x)(k^2x)}}.$$ (59) Since the Weierstrass function can be defined as $$u=_{\mathrm{}(u)}^{\mathrm{}}\frac{dy}{\sqrt{4y^3g_2yg_3}},$$ (60) from (59) it follows that $$\mathrm{sin}^2\phi =\mathrm{}(2at+u_0)+\mathrm{const}.,Imu_00,$$ (61) where $$u_0=_0^{\mathrm{}}\frac{dx}{\sqrt{4x(1x)(k^2x)}}.$$ (62) Clearly the expression (61) coincides with (55). Now define the matrices $`\mathrm{\Phi }_\pm (\lambda ,t)_{jk}=\phi _\pm ^j(p_k,t)`$, where $$\phi _+(p,t)=e^{\mu (p)t}\phi _{}(p,t),$$ then $$g_\pm (\lambda ,t)=\mathrm{\Phi }_\pm (\lambda ,t)\mathrm{\Phi }_\pm ^1(\lambda ,t=0).$$ (63) This formula follows from (13), due to relation (14). The relation (55) can be satisfied for all times if the choice of $`p_1`$ is such that $`Im(A_1/\alpha )0`$, in accordance with (61). This follows from $`\mathrm{}(u)=u^2+O(u^2)`$ so that the r.h.s. of (55) diverges for $$2at+\frac{A_1}{\alpha }=0.$$ (64) Therefore if $`Im(A_1/\alpha )0`$, one obtains a physical solution. If $`p_1`$ is chosen such that $`Im(A_1/\alpha )=0`$, then the factorization (8) will not be valid for $`t`$ close to $`A_1/(2a\alpha )`$. In this case one would have a non-canonical factorization given by (9). We will show in the next section that this case corresponds to a dynamical system based on a non-compact extension of the $`SO(2)`$ group. ## 5 Related problems Let us replace $`\varphi `$ with $`iq`$ in (1), and let $`l=\dot{q}`$. In this way we obtain a new dynamical system describing particle on a line moving in a potential $`V(q)=a^2\mathrm{cosh}2q`$. The solution of the equations of motion can be obtained from the energy $$2E=\dot{q}^22a^2\mathrm{cosh}2q,$$ and it is given by $$2at=\pm _{ϵ_k}^{\mathrm{sinh}^2q}\frac{dx}{\sqrt{4x(1+x)(k^2+x)}},$$ (65) where $`ϵ_k=0`$ for $`k^2=2a^2/(E+a^2)0`$, while $`ϵ_k=k^2`$ for $`k^2<0`$. Note that now there is no restriction $`E+a^20`$ as in the compact case, since the potential is not bounded from bellow. Then by using the same relations as in the compact case we get $$\mathrm{sinh}^2q=\mathrm{}(2at+u_0)+\mathrm{const}.,Imu_0=0,$$ (66) since $$u_0=_{ϵ_k}^{\mathrm{}}\frac{dx}{\sqrt{4x(1+x)(k^2+x)}}.$$ (67) This is an example of a dynamical system where $`q`$ diverges for finite times, and hence one can expect that the corresponding factorization will be non-canonical. The replacement $`\varphi iq`$ corresponds to replacing the $`SO(2)`$ group with the non-compact subgroup of $`SO(2,𝐂)`$. The corresponding Lax matrix can be obtained from (2) by taking $`𝐟=𝐚`$ and $`\varphi =iq`$, so that $$L=\left(\begin{array}{cc}a\lambda a\lambda ^1\mathrm{cosh}2q& lia\lambda ^1\mathrm{sinh}2q\\ lia\lambda ^1\mathrm{sinh}2q& a\lambda +a\lambda ^1\mathrm{cosh}2q\end{array}\right),$$ (68) and $`l=i\dot{q}`$. The spectral curve is (29) with $`EE`$, since $$l^22a^2\mathrm{cos}2\varphi =\dot{q}^22a^2\mathrm{cosh}2q=2E.$$ Consequently, the derivation of the BA functions is almost the same as in the section 4, and the only difference is in the equation (50), where $`aa`$ due to the fact that now $`𝐟=𝐚`$. This gives $$\mathrm{sinh}^2q=\mathrm{}(2at+A_1/\alpha )+\mathrm{const}..$$ (69) The physical solutions are obtained for $`ImA_1/\alpha =0`$, in agreement with (66). Hence the canonical factorization of $`e^{tL}`$ will not be possible for $`2atA_1/\alpha `$. Note that for $`n3`$ one obtains the spectral curves which are non-hyperelliptic, and in that case finding the BA functions gets difficult, because not much is known about the non-hyperelliptic curves, except in some special cases, like Kowalewski top (see for other examples). For example, for $`n=3`$ case the Lax matrix (2) gives a cubic spectral curve $$\mu ^3+L^{(2)}\mu detL=0,$$ (70) which is difficult to analyze. It would be interesting to see whether some alternative method of solving the corresponding factorization problem can be used, for example see . Preliminary study indicates that the techniques developed in can be used to solve the $`n=2`$ case , so that one can expect that the $`n=3`$ case could be also solved. Note that the Lax matrix (2) can be generalized to $$L=𝐚\lambda ^m+\underset{j=m1}{\overset{k}{}}L_j\lambda ^j,m,k𝐍,$$ (71) where $`𝐚`$ is a constant matrix while $`L_j`$ are the dynamical matrices . The dynamics is given by (4) and the corresponding factorization problem is given by (8). In the $`n=2`$ case ($`Lgl(2)`$ or $`sl(2)`$) the BA functions can be easily calculated since in that case one obtains a hyper-elliptic spectral curve $$w^2=P_{2m+2k}(\lambda ),$$ (72) where $`P_{2m+2k}(\lambda )`$ is a polynomial of order $`2m+2k`$. However, this system is not physically interesting, i.e. the corresponding Hamiltonian $$H=\frac{1}{2}Restr(\lambda ^1L^2),$$ (73) has no physical applications. Still, the two-by-two matrices (71) provide non-trivial solvable examples of the Riemann-Hilbert problem for $`e^{tL}`$, and one can compare the results to alternative methods of matrix factorization. ## Acknowledgements We would like to thank J. Mourao, A. Perelomov and D. Korotkin for helpful discussions. A.M. was supported by the grant PRAXIS/BCC/18981/98 from the Portugese Foundation for Science and Technology.
warning/0001/hep-ph0001056.html
ar5iv
text
# BUHEP-00-1 Fermilab-Pub-00/008-T hep-ph/0001056 Vacuum Alignment in Technicolor Theories I. The Technifermion Sector ## 1. Introduction One of the original motivations for the dynamical approach to electroweak and flavor symmetry breaking—specifically, technicolor and extended technicolor —was the belief that it would solve the problem of strong CP–violation in QCD . The idea was this: In a theory consisting only of gauge interactions of massless fermions, instanton angles such as $`\theta _{QCD}`$ may be freely rotated to zero. Purely dynamical masses, i.e., fermion bilinear condensates, may be assumed to be CP–conserving. And, the fermions’ hard masses are generated by the joint action of dynamical and explicit chiral symmetry breaking, all induced by gauge interactions alone. It was hoped that this combination naturally would produce a for which $`\overline{\theta }_q=\theta _{QCD}+\mathrm{arg}det(M_q)=0`$ without an axion. This is naive, especially if at least some of the observed CP–violation is to emerge from diagonalizing the quark mass matrix $`M_q`$. In fact, the way to determine the true status of CP symmetry in a superficially CP–invariant theory was prescribed long ago by Dashen . He studied the question of determining the correct perturbative ground state $`|\mathrm{\Omega }`$ upon which to begin an expansion about the chiral limit. This process is known as vacuum alignment. When the chiral symmetry of quarks is spontaneously broken, there are infinitely many degenerate vacua, parameterized by transformations corresponding to massless Goldstone bosons. Dashen showed that, if this chiral symmetry is also explicitly broken by $`_q^{}=\overline{q}_LM_qq_R+\mathrm{h}.\mathrm{c}.`$, the degeneracy is lifted and the correctly aligned zeroth–order ground state $`|\mathrm{\Omega }`$ is the one in which the expected value of $`_q^{}`$ is least. In practice, it is easier to fix $`|\mathrm{\Omega }`$ as a “standard vacuum” with simple condensates <sup>1</sup><sup>1</sup>1We assume that the quark chiral symmetry $`SU(2n)_LSU(2n)_R`$ is spontaneously broken to an $`SU(2n)`$ subgroup, in which case the quark condensates $`\overline{q}_{aL}q_{bR}`$ are proportional to an $`SU(2n)`$ matrix. In the standard vacuum, $`\mathrm{\Omega }|\overline{q}_{aL}q_{bR}|\mathrm{\Omega }\delta _{ab}`$. and chirally rotate $`_q^{}`$ to find the minimum vacuum energy. Dashen showed that, even if the original $`_q^{}`$ is CP–conserving, i.e., if $`M_q`$ is real, the Hamiltonian aligned with $`|\mathrm{\Omega }`$ may be CP–violating. This is spontaneous CP–violation. For real $`M_q`$, it occurs if $`\overline{\theta }_q=\pi `$. The aligned Hamiltonian has the CP–violating term $`i\nu _q\overline{q}\gamma _5q`$, where $`\nu _q`$ is of order the smallest eigenvalue of $`M_q`$ . Dashen’s study was made in the context of QCD, but it applies to a theory in which QCD is united with technicolor to generate quark masses by extended technicolor . In such a theory, the chiral symmetries of technifermions are spontaneously broken at $`\mathrm{\Lambda }_{TC}1\mathrm{TeV}`$, giving rise to massless technipions, $`\pi _T`$. All but the three $`\pi _T^{\pm ,0}`$ that become the longitudinal components of the $`W^\pm `$ and $`Z^0`$ bosons must get large masses, at least 50–70 GeV for the charged ones. Quark chiral symmetries are spontaneously broken at the much lower scale $`\mathrm{\Lambda }_{QCD}1\mathrm{GeV}`$<sup>2</sup><sup>2</sup>2Complications due to the top quark will be discussed below. All these symmetries, except electroweak $`SU(2)U(1)`$, are explicitly broken by ETC–boson exchange interactions. They are well–approximated at 1 TeV by four–fermion interactions, $`\overline{T}T\overline{T}T`$ and $`\overline{q}T\overline{T}q`$, suppressed by the square of $`M_{ETC}>100\mathrm{TeV}`$<sup>3</sup><sup>3</sup>3To a lesser extent, the electroweak interactions also contribute to explicit symmetry breaking; see Ref. . They are ignored in Eq. (2) below. It is natural to assume that ETC breaking is such that these four–fermion interactions have real coefficients and so are superficially CP–conserving. Vacuum alignment then has three possible outcomes: (1) the correct chiral–breaking perturbation, $`^{}`$, is still CP–conserving and, in particular, the Cabibbo–Kobayashi–Maskawa (CKM) matrix is real; (2) $`^{}`$ is CP–violating, but $`|\nu _q|m_u`$ is $`10^9`$ times too large; (3) $`^{}`$ is CP–violating, but $`|\nu _q|=0`$ or is at most is of order the ratio of condensates $`\overline{q}q/\overline{T}T<10^9`$. This last alternative, of course, is the desired one. Unfortunately, no physical criteria were found to lead to models of type 3. The matter rested there until the dynamical attempts known as topcolor–assisted technicolor (TC2) were made to deal with the large mass of the top quark . It has always been difficult for the dynamical approach, especially extended technicolor, to account for the top quark’s mass. Either the ETC scale generating the top mass must be near 1 TeV, leading to conflict with experimental measurements on the $`\rho `$ parameter and the $`Z\overline{b}b`$ decay rate , or, if it is made much higher, the coupling $`g_{ETC}`$ must be unnaturally fine–tuned. Hill circumvented these difficulties by invoking another strong interaction near 1 TeV, topcolor, to generate a large $`\overline{t}t`$ condensate and top mass. In TC2, ordinary technicolor remains responsible for the bulk of electroweak symmetry breaking. An important consequence of this scenario—and this is where vacuum alignment comes back in—is that top condensation implies a triplet of massless Goldstone “top–pions”, $`\pi _t^{\pm ,0}`$. These must acquire mass $`M_{\pi _t}>m_t=175\mathrm{GeV}`$; otherwise $`tb\pi _t^+`$ becomes a major decay mode. Extended technicolor interactions provide this mass by contributing 5–10 GeV to $`m_t`$ . At the same time, this ETC contribution must not induce appreciable mixing of top–pions with ordinary technipions . Some technipions may be as light as 100 GeV , so that large mixing would lead to substantial, and also unobserved, $`tb\pi _T^+`$. Balaji studied top–pion mass and mixing in a specific model, and he obtained encouraging results . However, his conclusions are preliminary because he was unable to execute vacuum alignment properly. This is understandable because vacuum alignment in TC2 models is very complicated. Now it involves at least two gauge interactions strong near 1 TeV—technicolor and topcolor—with some technifermions transforming under both. And, many technifermions are needed to accommodate various experimental constraints, making the chiral flavor group quite large; see Ref. for details. One of these experimental constraints is that no physical technipion be massless or very light. The criterion used in Refs. for deciding this was that no spontaneously broken chiral charge (other than the electroweak charges) can commute with the ETC–generated $`\overline{T}T\overline{T}T`$ interactions. We shall see in Section 3 that this criterion, which works in QCD, is insufficient to guarantee that all technipions are massive. The problem of vacuum alignment in technicolor theories is too complex for analytical treatment. Numerical methods are needed. We start the numerical analysis in this paper by considering the technifermion sector of a simple ETC model, one in which there are $`N`$ doublets of a single type of technifermion that transforms according to the complex fundamental representation of the technicolor gauge group $`SU(N_{TC})`$. The rest of this paper proceeds as follows: In Section 2 we define our simplified ETC model and present the formalism in first–order chiral perturbation theory for vacuum alignment and calculating technipion masses. There we illustrate the unexpected (to us, anyway) fact that chiral symmetries are not always manifest in the chiral–breaking perturbation $`^{}`$. We present in Section 3 the main results of vacuum alignment in the technifermion sector. We have found a quite surprising result: the phases in the technifermions’ unitary vacuum–alignment matrix $`W_0`$ may be integer multiples of $`\pi /N^{}`$ where $`N^{}N`$. If they are allowed by unitarity, these “rational phases” occur because the terms in $`^{}`$ make it energetically favorable for certain phases to be equal or to differ by $`\pi `$ and because $`W_0`$ is unimodular. If unitarity frustrates this alignment of phases in $`W_0`$, they are irrational. We shall see that the rational phases appear as islands in an irrational sea, the boundaries of which are defined by critical values of the parameters in $`^{}`$. Furthermore, a technipion becomes massless, a Goldstone boson to first order, at the island shore, where the ETC parameters become critical. This has the important phenomenological consequence that an exceptionally light technipion often accompanies the rational phases because generically chosen parameters are not far from the critical ones. Thus, some technipions may be even lighter than we expected , a fact which may be welcome and which, in any case, can be used to help choose among models. We conclude in Section 4 with a brief look ahead to vacuum alignment and CP–violation in the quark sector. ## 2. The Extended Technicolor Model To simplify our numerical studies, we consider models in which a single kind of technifermion interacts with quarks (but no leptons) via ETC interactions. There are $`N`$ technifermion doublets $`(U_{iL,R},D_{iL,R})`$, $`i=1,2,\mathrm{},N`$, all transforming according to the fundamental representation of the technicolor gauge group $`SU(N_{TC})`$. There are $`n`$ generations of $`SU(3)_C`$ triplet quarks $`(u_{aL,R},d_{aL,R})`$, $`a=1,2,\mathrm{},n`$. The left–handed fermions are electroweak $`SU(2)`$ doublets and the right–handed ones are singlets. Here and below, we exhibit only flavor, not technicolor nor color, indices. Although it is not essential for our studies, we shall assume that the technicolor gauge coupling runs slowly, or “walks” from the TC to the ETC scale . No provision to give a realistic top quark mass, such as topcolor–assisted technicolor , will be made in this paper. The technifermions are ordinary color–singlets, so the chiral flavor group of our model is $`G_f=\left[SU(2N)_LSU(2N)_R\right]\left[SU(2n)_LSU(2n)_R\right]`$. We have excluded anomalous $`U_A(1)`$’s strongly broken by TC and color instanton effects. When the TC and QCD couplings reach their required critical values, these symmetries are spontaneously broken to $`S_f=SU(2N)SU(2n)`$. We shall take this residual symmetry to be the diagonal vectorial one by adopting as our standard vacuum the state $`|\mathrm{\Omega }`$ in which the nonzero fermion bilinear condensates are diagonal: $`\mathrm{\Omega }|\overline{U}_{iL}U_{jR}|\mathrm{\Omega }`$ $`=`$ $`\mathrm{\Omega }|\overline{D}_{iL}D_{jR}|\mathrm{\Omega }=\delta _{ij}\mathrm{\Delta }_T`$ $`\mathrm{\Omega }|\overline{u}_{aL}u_{bR}|\mathrm{\Omega }`$ $`=`$ $`\mathrm{\Omega }|\overline{d}_{aL}d_{bR}|\mathrm{\Omega }=\delta _{ab}\mathrm{\Delta }_q.`$ (1) The condensates $`\mathrm{\Delta }_TN_{TC}\mathrm{\Lambda }_{TC}^3`$ and $`\mathrm{\Delta }_qN_C\mathrm{\Lambda }_{QCD}^3`$ when they are renormalized at their respective strong interaction scales. Of course, $`N_C=3`$. All of the $`G_f`$ symmetries except for the gauged electroweak $`SU(2)U(1)`$ must be explicitly broken by extended technicolor interactions . In the absence of a concrete ETC model, we write the interactions broken at the scale $`M_{ETC}/g_{ETC}=𝒪(100\mathrm{TeV})`$ in the phenomenological four-fermion form <sup>4</sup><sup>4</sup>4In Eq. (2), we have not made any assumption about the structure of ETC interactions vis–a–vis the electroweak ones. $`^{}`$ $``$ $`_{TT}^{}+_{Tq}^{}+_{qq}^{}`$ (2) $`=`$ $`\mathrm{\Lambda }_{ijkl}^{TT}\overline{T}_{iL}\gamma ^\mu T_{jL}\overline{T}_{kR}\gamma _\mu T_{lR}+\mathrm{\Lambda }_{iabj}^{Tq}\overline{T}_{iL}\gamma ^\mu q_{aL}\overline{q}_{bR}\gamma _\mu T_{jR}+\mathrm{h}.\mathrm{c}.`$ $`+`$ $`\mathrm{\Lambda }_{abcd}^{qq}\overline{q}_{aL}\gamma _\mu q_{bL}\overline{q}_{cR}\gamma _\mu q_{dR},`$ where $`T_{iL,R}`$ and $`q_{aL,R}`$ stand for all $`2N`$ technifermions and $`2n`$ quarks, respectively. Here, $`M_{ETC}`$ is a typical ETC gauge boson mass and the $`\mathrm{\Lambda }`$ coefficients are $`g_{ETC}^2/M_{ETC}^2`$ times mixing factors for these bosons and group theoretical factors. Typically, the $`\mathrm{\Lambda }`$’s are positive, though some may be negative. In our calculations, we choose the $`\mathrm{\Lambda }`$’s to avoid unwanted Goldstone bosons. Hermiticity of $`^{}`$ requires $$(\mathrm{\Lambda }_{ijkl}^{TT})^{}=\mathrm{\Lambda }_{jilk}^{TT},(\mathrm{\Lambda }_{iabj}^{Tq})^{}=\mathrm{\Lambda }_{aijb}^{Tq},(\mathrm{\Lambda }_{abcd}^{qq})^{}=\mathrm{\Lambda }_{badc}^{qq}.$$ (3) The assumption of time-reversal invariance for this theory before any potential breaking via vacuum alignment means that the angles $`\theta _{TC}=\theta _{QCD}=0`$ (at tree level) and that all the $`\mathrm{\Lambda }`$’s are real. Thus, e.g., $`\mathrm{\Lambda }_{ijkl}^{TT}=\mathrm{\Lambda }_{jilk}^{TT}`$. All the four–fermion operators in $`^{}`$ are renormalized at the ETC scale. Throughout this work, we shall assume that the ETC gauge symmetries commute with electroweak $`SU(2)`$, but not with weak hypercharge $`U(1)`$ (indeed, they must not; see Ref. ). The ETC interactions then take the form, e.g., $$_{TT}^{}=\left(\overline{U}_{iL}\gamma ^\mu U_{jL}+\overline{D}_{iL}\gamma ^\mu D_{jL}\right)\left(\mathrm{\Lambda }_{ijkl}^U\overline{U}_{kR}\gamma _\mu U_{lR}+\mathrm{\Lambda }_{ijkl}^D\overline{D}_{kR}\gamma _\mu D_{lR}\right).$$ (4) Having chosen a standard chiral–perturbative ground state, $`|\mathrm{\Omega }`$, vacuum alignment proceeds by minimizing the expectation value of the rotated Hamiltonian. This is obtained by making the $`G_f`$ transformation $`T_{L,R}W_{L,R}T_{L,R}`$ and $`q_{L,R}V_{L,R}q_{L,R}`$, where $`W_{L,R}SU(2N)_{L,R}`$ and $`V_{L,R}SU(2n)_{L,R}`$: $`^{}(W,V)`$ $`=`$ $`_{TT}^{}(W_L,W_R)+_{Tq}^{}(W,V)+_{qq}^{}(V_L,V_R)`$ $`=`$ $`\mathrm{\Lambda }_{ijkl}^{TT}\overline{T}_{i^{}L}W_{Li^{}i}^{}\gamma ^\mu W_{Ljj^{}}T_{j^{}L}\overline{T}_{k^{}R}W_{Rk^{}k}^{}\gamma ^\mu W_{Rll^{}}T_{l^{}L}+\mathrm{}.`$ Since $`T`$ and $`q`$ transform according to complex representations of their respective color groups, the four–fermion condensates in the $`S_f`$–invariant $`|\mathrm{\Omega }`$ have the form $`\mathrm{\Omega }|\overline{T}_{iL}\gamma ^\mu T_{jL}\overline{T}_{kR}\gamma _\mu T_{lR}|\mathrm{\Omega }`$ $`=`$ $`\mathrm{\Delta }_{TT}\delta _{il}\delta _{jk},`$ $`\mathrm{\Omega }|\overline{T}_{iL}\gamma ^\mu q_{aL}\overline{q}_{bR}\gamma _\mu T_{jR}|\mathrm{\Omega }`$ $`=`$ $`\mathrm{\Delta }_{Tq}\delta _{ij}\delta _{ab},`$ (6) $`\mathrm{\Omega }|\overline{q}_{aL}\gamma ^\mu q_{bL}\overline{q}_{cR}\gamma _\mu q_{dR}|\mathrm{\Omega }`$ $`=`$ $`\mathrm{\Delta }_{qq}\delta _{ad}\delta _{bc}.`$ The condensates are positive, renormalized at $`M_{ETC}`$ and, in the large–$`N_{TC}`$ and $`N_C`$ limits, they are given by $`\mathrm{\Delta }_{TT}(\mathrm{\Delta }_T(M_{ETC}))^2`$, $`\mathrm{\Delta }_{Tq}\mathrm{\Delta }_T(M_{ETC})`$ $`\times \mathrm{\Delta }_q(M_{ETC})`$, and $`\mathrm{\Delta }_{qq}(\mathrm{\Delta }_q(M_{ETC}))^2`$. In walking technicolor, $`\mathrm{\Delta }_T(M_{ETC})(M_{ETC}/\mathrm{\Lambda }_{TC})\mathrm{\Delta }_T(\mathrm{\Lambda }_{TC})`$ $`=10^2`$$`10^3\times \mathrm{\Delta }_T(\mathrm{\Lambda }_{TC})`$. In QCD, however, $`\mathrm{\Delta }_q(M_{ETC})(\mathrm{log}(M_{ETC}/\mathrm{\Lambda }_{QCD}))^{\gamma _m}\mathrm{\Delta }_q(\mathrm{\Lambda }_{ETC})\mathrm{\Delta }_q(\mathrm{\Lambda }_{QCD})`$, where $`\gamma _m2\alpha _C/\pi `$ for $`SU(3)_C`$ . Thus, the ratio $$r=\frac{\mathrm{\Delta }_{Tq}(M_{ETC})}{\mathrm{\Delta }_{TT}(M_{ETC})}\frac{\mathrm{\Delta }_{qq}(M_{ETC})}{\mathrm{\Delta }_{Tq}(M_{ETC})}$$ (7) is at most $`10^{10}`$. This is 10–100 times smaller than in a technicolor theory in which the coupling does not walk. With these condensates, the vacuum energy is a function only of $`W=W_LW_R^{}`$ and $`V=V_LV_R^{}`$, elements of the coset space $`G_f/S_f`$: $`E(W,V)=E_{TT}(W)+E_{Tq}(W,V)+E_{qq}(V)`$ (8) $`=\mathrm{\Lambda }_{ijkl}^{TT}W_{jk}W_{li}^{}\mathrm{\Delta }_{TT}(\mathrm{\Lambda }_{iabj}^{Tq}V_{ab}W_{ji}^{}+\mathrm{c}.\mathrm{c}.)\mathrm{\Delta }_{Tq}\mathrm{\Lambda }_{abcd}^{qq}V_{bc}V_{da}^{}\mathrm{\Delta }_{qq}.`$ Note that time–reversal invariance of the unrotated Hamiltonian $`^{}`$ implies that $`E(W,V)=E(W^{},V^{})`$. Hence, spontaneous CP–violation occurs if the solutions $`W_0`$, $`V_0`$ to the minimization problem are complex. Following Ref. , we define technifermion current mass matrices renormalized at the ETC scale as follows:<sup>5</sup><sup>5</sup>5These definitions differ from those in Ref. by a common vectorial transformation on the left and right–handed fields. This affects none of our discussion. $`M_{Tij}\mathrm{\Delta }_T(M_{ETC})`$ $`=`$ $`\left(W_{ik}^{}{\displaystyle \frac{E}{W_{jk}^{}}}\right)_{W_0,V_0}`$ (9) $`=`$ $`W_{0ik}^{}\left(\mathrm{\Lambda }_{klmj}^{TT}W_{0lm}\mathrm{\Delta }_{TT}+\mathrm{\Lambda }_{kabj}^{Tq}V_{0ab}\mathrm{\Delta }_{Tq}\right)`$ $`=`$ $`W_{0ik}^{}\mathrm{\Lambda }_{klmj}^{TT}W_{0lm}\mathrm{\Delta }_{TT}\left(1+𝒪(r)\right).`$ For quarks, $`M_{qab}\mathrm{\Delta }_q(M_{ETC})`$ $`=`$ $`\left(V_{ac}^{}{\displaystyle \frac{E}{V_{bc}^{}}}\right)_{W_0,V_0}`$ (10) $`=`$ $`V_{0ac}^{}\left(\mathrm{\Lambda }_{cijb}^{Tq}W_{0ij}\mathrm{\Delta }_{Tq}+\mathrm{\Lambda }_{cdeb}^{qq}V_{0de}\mathrm{\Delta }_{qq}\right)`$ $`=`$ $`V_{0ac}^{}\mathrm{\Lambda }_{cijb}^{Tq}W_{0ij}\mathrm{\Delta }_{Tq}\left(1+𝒪(r)\right).`$ Nuyts’ theorem generalized to technicolor states that, as a consequence of extremizing the energy, the imaginary parts of these matrices are proportional to the identity: $`\left(M_TM_T^{}\right)\mathrm{\Delta }_T(M_{ETC})=i\nu _T1_{2N},`$ $`\left(M_qM_q^{}\right)\mathrm{\Delta }_q(M_{ETC})=i\nu _q1_{2n}.`$ (11) The parameters $`\nu _T`$ and $`\nu _q`$ are Lagrange multipliers associated with the unimodularity constraints on $`W_0`$ and $`V_0`$, respectively. These equations imply that $`M_T`$ and $`M_q`$ are each diagonalized by a (different) single special unitary transformations. Taking the trace of both sides of Eqs. (2. The Extended Technicolor Model) and using Eqs. (9,10) gives $`2iN\nu _T`$ $``$ $`\mathrm{Tr}\left(M_TM_T^{}\right)\mathrm{\Delta }_T(M_{ETC})`$ (12) $`=`$ $`\mathrm{Tr}\left(M_qM_q^{}\right)\mathrm{\Delta }_q(M_{ETC})2in\nu _q`$ $`=`$ $`2i\mathrm{\Lambda }_{kabj}^{Tq}\mathrm{Im}\left(W_{0kj}^{}V_{0ab}\right)\mathrm{\Delta }_{Tq}.`$ This relation between $`\nu _T`$ and $`\nu _q`$ requires that $`SU(N_{TC})`$ and $`SU(3)_C`$ are embedded in a simple ETC group, so that $`\theta _{TC}=\theta _{QCD}`$. Strong CP–violation occurs if $`\nu _{T,q}0`$. The angle $`\overline{\theta }_q`$ characterizing this for quarks’ is defined by (for $`\theta _{QCD}=0`$<sup>6</sup><sup>6</sup>6See, e.g., Ref. for the relation between $`\theta _q`$ and $`\nu _q`$ for the case of three light quarks. $$\overline{\theta }_q=\mathrm{arg}det\left(M_q\right)\mathrm{\Delta }_q=\mathrm{arg}det\left(_q\right)\mathrm{\Delta }_q,$$ (13) where, up to $`\mathrm{\Lambda }^{qq}`$ terms of relative order $`r`$, $$_{qab}\mathrm{\Delta }_q(V_0M_q)_{ab}\mathrm{\Delta }_q=\mathrm{\Lambda }_{aijb}^{Tq}W_{0ij}\mathrm{\Delta }_{Tq}$$ (14) is the primordial quark mass matrix, i.e., the one before vacuum alignment in the quark sector. We see that strong CP–violation arises from a conflict between mass terms and a chiral symmetry constraint on the alignment matrix. This is what Dashen and Nuyts showed for quarks in QCD and what we found in Ref. for extended technicolor. In a world with just one type of fermion, say $`T_{iL,R}`$, with explicit flavor symmetry breaking due to gauge interactions alone, $`M_T=M_T^{}`$ and there is no strong CP–violation even if the aligning matrix $`W_0`$ is complex and CP symmetry is spontaneously broken, Suppose we found $`\overline{\theta }_q=0`$ up to the $`\mathrm{\Lambda }^{qq}`$ terms of order $`r`$. Are there larger contributions to $`\overline{\theta }_q`$? The first to worry about are two–loop ETC contributions to $`_q`$. There are two types of these: those with one technifermion dynamical mass insertion and those with three. The first are proportional to a single power of $`W_0`$ and, because the $`\mathrm{\Lambda }^{TT}`$’s are real, it is plausible that they will not change $`\overline{\theta }_q`$. This must be checked in specific models. The three–insertion graphs involve two $`W_0`$’s and one $`W_0^{}`$ convoluted with $`\mathrm{\Lambda }^{TT}`$’s and these are more likely to contribute to $`\overline{\theta }_q`$. Apart from any $`g_{ETC}^2/16\pi ^2`$ suppression these graphs may have, they are of relative order $`\mathrm{\Delta }_T^2(M_{ETC})/M_{ETC}^6<(\mathrm{\Lambda }_{TC}/M_{ETC})^4<10^{10}`$ in a walking technicolor theory. We tentatively conclude that the $`\overline{\theta }_q`$ defined in Eq. (13) is a reliable measure of strong CP–violation in extended technicolor models. We need only know $`W_0`$ and the $`\mathrm{\Lambda }^{Tq}`$ to determine it. Our strategy for vacuum alignment, which we carry out numerically, is the following: Because $`r`$ is small, we first minimize $`E_{TT}`$ to determine $`W_0`$. If we wish to determine $`W_{0L}`$ and $`W_{0R}`$ separately, we make vectorial transformations on $`T_{L,R}`$ that diagonalize $`M_{Tij}`$. Physical results such as technipion and quark masses are unchanged even if we use, for example, $`W_{0L}=W_0`$. The results of technifermion alignment are presented in the next section. Once $`W_0`$ is determined, it is inserted as a set of parameters into $`E_{Tq}`$ and this is minimized as a function of $`V`$. If there are several degenerate solutions $`W_0`$ minimizing $`E_{TT}`$, one should choose the one giving the deepest minimum $`E_{Tq}(W_0,V_0)`$. When $`V_0`$ is known, the matrices $`V_{0L}`$, $`V_{0R}`$ are determined by diagonalizing the matrix $`M_q`$ in Eq. (10). The quark CKM matrix is then obtained from $`V_{0L}`$. Finally, holding $`V_0`$ fixed, one can refine $`W_0`$ by reminimizing $`E_{TT}+E_{Tq}`$ as a function of $`W`$. This will induce corrections of $`𝒪(r)`$ in $`W_0`$ and $`\overline{\theta }_q`$. There is no point in refining $`V_0`$ by minimizing the full energy including $`E_{qq}`$. However, note that the rotated $`_{qq}^{}(V_0)`$ may contain sources of quark CP–violation not contained in the CKM matrix . These studies of CP–violation in the quark sector will be presented in our next paper. We are concerned in this paper with vacuum alignment in the technifermion sector, and we turn to this now. We will allow only models in which alignment conserves electric charge, i.e., does not induce $`\overline{U}_iD_j`$ condensates. Then, the matrix minimizing $`E_{TT}(W)`$ must be block–diagonal, $$W_0=\left(\begin{array}{cc}W_0^U& 0\\ 0& W_0^D\end{array}\right),$$ (15) where $`W_0^U`$, $`W_0^D`$ are $`U(N)`$ matrices satisfying $`det(W_0^U)det(W_0^D)=1`$. The phase indeterminacy of the individual $`U(N)`$ matrices corresponds to the electroweak $`T_3`$ symmetry. Thus, for admissible models, we can minimize $`E_{TT}`$ in the subspace of block–diagonal matrices. Using Eq. (4), the vacuum energy takes the form $`E_{TT}(W^U,W^D)`$ $`=`$ $`(\mathrm{\Lambda }_{ijkl}^UW_{jk}^UW_{li}^U+\mathrm{\Lambda }_{ijkl}^DW_{jk}^DW_{li}^D)\mathrm{\Delta }_{TT}`$ (16) $``$ $`E_U(W^U)+E_D(W^D).`$ Since this expression is bilinear in $`W^{U,D}W^{U,D}`$, without loss of generality we can determine $`W_0`$ by separately minimizing $`E_U`$ and $`E_D`$ in the space of $`SU(N)`$ matrices. We do this in the next section, taking care to ensure that no Goldstone bosons remain massless other than the three associated with electroweak $`SU(2)`$ symmetry. This means that $`\mathrm{\Lambda }_{ijkl}^{U,D}`$ must be chosen so that there are no massless $`SU(N)SU(N)`$ Goldstone boson in either $`U`$ or $`D`$ sector. We close this section with some remarks on calculating the pseudoGoldstone boson (technipion) mass spectrum in these technicolor models. In the standard chiral–perturbative ground state, $`|\mathrm{\Omega }`$, the spontaneously broken symmetries are formally generated by the $`G_f/S_f`$ charges $$Q_5^A=\frac{1}{2}d^3x\left(T_R^{}\lambda _AT_RT_L^{}\lambda _AT_L\right),$$ (17) Here, $`\lambda _A`$ are the $`4N^21`$ Gell-Mann matrices of $`SU(2N)`$. To first order in the chiral perturbation $`_{TT}^{}`$, technipion masses are given by Dashen’s formula , $$F_T^2M_{\pi AB}^2=i^2\mathrm{\Omega }\left|[Q_5^A,[Q_5^B,_{TT}^{}(W_{0L},W_{0R})]]\right|\mathrm{\Omega },$$ (18) where $`F_T=246\mathrm{GeV}/\sqrt{N}`$ is the technipion decay constant and $`_{TT}^{}`$ is given in Eq. (2. The Extended Technicolor Model). As noted above, we can determine $`W_{0L}`$ and $`W_{0R}`$ by diagonalizing the technifermion current–mass matrix, $`M_T`$. However, since $`M_{\pi AB}^2`$ is invariant under vectorial transformations of the technifermion fields, it is simpler to compute it using $`W_{0L}=W_0`$ and $`W_{0R}=1`$. The result is $`F_T^2M_{\pi AB}^2=\frac{1}{2}\mathrm{\Lambda }_{ijkl}[\left(\{\lambda _A,\lambda _B\}W_0^{}\right)_{li}W_{0jk}+\left(W_0\{\lambda _A,\lambda _B\}\right)_{jk}W_{0li}^{}`$ $`2\left(\lambda _AW_0^{}\right)_{li}\left(W_0\lambda _B\right)_{jk}2\left(\lambda _BW_0^{}\right)_{li}\left(W_0\lambda _A\right)_{jk}]\mathrm{\Delta }_{TT}.`$ (19) Note that, because the vector charges annihilate the standard vacuum, the axial charges in Eq. (2. The Extended Technicolor Model) may be replaced by left–handed or right–handed charges or by any linear combination that is not purely a vector charge. In using Eq. (2. The Extended Technicolor Model) in these models, we have seen examples in which a technipion’s mass vanishes without there being a corresponding conserved chiral charge, i.e., a linear combination of the $`Q_R^A`$ and $`Q_L^A`$ which commutes with $`_{TT}^{}(W_0)`$. A two–technidoublet, $`SU(4)SU(4)`$ example is provided by the following set of $`\mathrm{\Lambda }`$’s (whose scale is arbitrary): $`\mathrm{\Lambda }_{1111}^U`$ $`=`$ $`\mathrm{\Lambda }_{1111}^D=\mathrm{\Lambda }_{2222}^U=\mathrm{\Lambda }_{2222}^D=1`$ $`\mathrm{\Lambda }_{1112}^U`$ $`=`$ $`\mathrm{\Lambda }_{1121}^U=\mathrm{\Lambda }_{1112}^D=\mathrm{\Lambda }_{2221}^D=\frac{1}{2}`$ $`\mathrm{\Lambda }_{1211}^U`$ $`=`$ $`\mathrm{\Lambda }_{2111}^U=\frac{1}{2}.`$ (20) In addition to the three electroweak Goldstone bosons coupling to $$\frac{1}{2}d^3x\underset{i=1}{\overset{2}{}}(U_i^{},D_i)\gamma _5\tau _a\left(\begin{array}{c}U_i\\ D_i\end{array}\right),$$ there is a fourth one associated with the $`W_0`$–rotation of the axial charge $$\frac{1}{2}d^3x\left(D_1^{}\gamma _5D_1D_2^{}\gamma _5D_2\right).$$ However, the divergence of its current is manifestly of first order in $`_{TT}^{}(W)`$. This extra massless technipion is at first surprising when one recalls that Dashen proved that a zero eigenvalue of the Goldstone boson mass–squared matrix implies that the corresponding current is conserved . Furthermore, in QCD we have become used to a conserved current being associated with a massless Goldstone boson. There, the symmetry that leaves the boson massless is manifest in the mass matrix $`M_q`$ of $`_q^{}=\overline{q}_LM_qq_R+\mathrm{h}.\mathrm{c}.`$. The resolution of this puzzle is that Dashen’s proof applies to the matrix elements of double commutators in the exact ground state, $`|\mathrm{vac}`$, of the full Hamiltonian $`H=(_0+^{}(W))`$. The matrix in Eq. (2. The Extended Technicolor Model) is calculated in the perturbative ground state, $`|\mathrm{\Omega }`$, which is the limit of $`|\mathrm{vac}`$ as $`^{}(W)0`$. Consequently, all that can be proved for a “massless” Goldstone boson at the perturbative level at which we work is that all matrix elements of the divergence of the corresponding current must be of second order in $`^{}`$. We emphasize that, although the masslessness of this technipion may be an approximation, it is important phenomenologically. Corrections to its mass are likely to be so small that it is already ruled out experimentally. ## 3. Results From the Technifermion Sector The vacuum energy in the $`U`$ and $`D`$–technifermion sectors has the form $`E(W)`$ $`=`$ $`{\displaystyle \underset{ijkl=1}{\overset{N}{}}}\mathrm{\Lambda }_{ijkl}W_{jk}W_{li}^{}\mathrm{\Delta }_{TT}`$ (21) $``$ $`{\displaystyle \underset{ijkl=1}{\overset{N}{}}}\mathrm{\Lambda }_{ijkl}|W_{jk}||W_{il}|\mathrm{exp}[i(\varphi _{jk}\varphi _{il})]\mathrm{\Delta }_{TT},`$ where $`W=W_U`$ or $`W_DSU(N)`$, $`\mathrm{\Lambda }_{ijkl}=\mathrm{\Lambda }_{ijkl}^{}=\mathrm{\Lambda }_{jilk}`$, and $`\varphi _{jk}=\mathrm{arg}(W_{jk})`$. We remind the reader that we always choose the $`\mathrm{\Lambda }_{ijkl}`$ so that there is no $`SU(N)SU(N)`$ Goldstone boson in either $`U`$ or $`D`$ sector. Note that, if $`W_0`$ minimizes $`E(W)`$, then so do the matrices $`Z_N^{2m}W_0=\mathrm{exp}(2im\pi /N)W_0`$, $`m=1,2,\mathrm{},N`$, and their complex conjugates. This degeneracy may be lifted by the quark–technifermion interaction $`_{Tq}^{}`$. It is especially convenient to parameterize $`W`$ in the form $$W=D_LKD_R.$$ (22) Here, $`D_{L,R}`$ are diagonal unimodular matrices, each depending on $`N1`$ phases: $$D_{L,R}=\mathrm{diag}[\mathrm{exp}(i\chi _{L,R\mathrm{\hspace{0.17em}1}}),\mathrm{exp}(i\chi _{L,R\mathrm{\hspace{0.17em}2}}),\mathrm{},\mathrm{exp}(i(\chi _{L,R\mathrm{\hspace{0.17em}1}}+\mathrm{}+\chi _{L,RN1}))],$$ (23) and $`K`$ is an $`(N1)^2`$–parameter CKM matrix which we write in the standard Harari–Leurer form . This matrix depends on $`\frac{1}{2}N(N1)`$ angles $`\theta _{ij}`$, $`1i<jN`$, and $`\frac{1}{2}(N1)(N2)`$ phases $`\delta _{ij}`$ , $`1i<j1N1`$. We have discovered several remarkable properties of the matrices $`W_0`$ which minimize $`E(W)`$. They have to do with the fact that the coefficient $`\mathrm{\Lambda }_{ijkl}`$ tends to align the phases $`\varphi _{jk}`$ and $`\varphi _{il}`$: if $`\mathrm{\Lambda }_{ijkl}>0`$, its contribution to $`E(W)`$ is minimized if the phases can be equal; if $`\mathrm{\Lambda }_{ijkl}<0`$, the phases want to differ by $`\pi `$. Of course, because not all the $`N^2`$ phases $`\varphi _{jk}`$ are independent, unitarity can frustrate phase alignment. If the nonzero $`\mathrm{\Lambda }_{ijkl}`$ link all the $`\varphi _{jk}`$ together, then all of them will be equal, mod $`\pi `$, if that can be consistent with unitarity. Unimodularity then requires all $`\varphi _{jk}=2m\pi /N`$, mod $`\pi `$, with $`m=1,2,\mathrm{},N`$. We call this “complete phase alignment” and we say that the phases are “rational”. Rational phases may also occur when the nonzero $`\mathrm{\Lambda }_{ijkl}`$ link some, but not all, of the phases. If it is allowed by unitarity, we have found that the phases are multiples of $`\pi /N^{}`$ (modulo the $`Z_N`$ phase $`2m\pi /N`$) for one or more values of $`N^{}`$ between 1 and $`N`$. This case of partial phase alignment is very rich, with many possibilities and, sometimes, degenerate minima whose $`W_0`$’s are not unitarily equivalent nor related by conjugation or a $`Z_N`$ factor. Its implications for quark CP–violation will be studied in our next paper. A necessary condition for phase alignment is that the CKM matrix $`K`$ is real. The reason for this is seen by looking at a typical complex term in $`K`$ for the 3$`\times `$3 case, e.g., $`s_{12}s_{23}c_{12}c_{23}s_{13}\mathrm{exp}(i\delta _{13})`$, where $`s_{12}=\mathrm{sin}\theta _{12}`$, etc. The mixing angles $`\theta _{ij}`$ are determined by the $`\mathrm{\Lambda }`$’s that are dominant in minimizing the energy and by unitarity. Then, the overall phase of this term will be a random irrational number unless $`\delta _{13}=0`$ or $`\pi `$ or one of the $`\theta _{ij}=0`$. If $`K`$ is complex, it contains more random phases than can be made rational by choices of phases in $`D_L`$ and $`D_R`$, and so the $`\varphi _{jk}`$ will be randomly irrational. Note that the case $`N=2`$ is special because $`K`$ is always real. In that case, all phases in $`W_0`$ are $`0`$ or $`\pi /2`$, mod $`\pi `$. Suppose that completely or partially–aligned rational phases occur for some set of $`\mathrm{\Lambda }`$’s. Then we find that the nonzero $`\mathrm{\Lambda }`$’s may be varied over an appreciable range with no change whatever in the phases. Ultimately, a large enough excursion in the $`\mathrm{\Lambda }`$’s will make it impossible to maintain unitarity with aligned phases and, at certain critical values of the $`\mathrm{\Lambda }`$’s, they change continuously from rational to irrational (or, in the $`SU(2)`$ case, discontinuously from one rational set to another). A rational–to–irrational phase transition may also occur if vanishing $`\mathrm{\Lambda }`$’s are made nonzero. By further varying the $`\mathrm{\Lambda }`$’s, another, possibly inequivalent, set of rational phases may characterize the matrix $`W_0`$. Thus, the minima of $`E(W)`$ as one varies the $`\mathrm{\Lambda }`$’s are islands of rational aligned phases in a sea of irrational ones. A Goldstone boson appears whenever a transition occurs between different types of phases. As the critical $`\mathrm{\Lambda }`$’s are approached, one of the $`M_\pi ^2`$ decreases to zero and then increases again once the boundary is passed. What is happening is this: As the transition is approached, the ground states for a set of rational phases are becoming less stable and a technipion’s $`M_\pi ^2`$ is diving through zero to negative values. At the same time, the ground states for a nearby set of irrational phases are becoming more stable and the corresponding $`M_\pi ^2`$ is increasing from negative to positive values. The two types of phases coexist at the rational island shore, giving rise to infinitely many degenerate minima that are characterized by an indeterminacy in the phases of $`D_{L,R}`$ and $`K`$. Hence, $`M_\pi ^2=0`$ (to first order) there. This is another situation in which the massless state’s chiral charge does not commute with $`_{TT}^{}`$. This phenomenon may be important. The appearance of an exceptionally light technipion is not uncommon because typical rational–phase $`\mathrm{\Lambda }`$–parameters often are not far from critical ones. In Ref. we observed that, because the number of technidoublets in typical TC2 models is large, $`N10`$, the technihadron scale is low and technipion masses may be as light as 100 GeV. Now we see that some technipions may be even lighter than nominally expected from the $`\mathrm{\Lambda }`$’s. In a specific model, this may be a major prediction or it may be a show–stopper. Finally, another interesting property of the rational–phase minima is that the coefficients $`\stackrel{~}{\mathrm{\Lambda }}_{ijkl}=_{i^{}j^{}}\mathrm{\Lambda }_{i^{}j^{}kl}W_{0ii^{}}^{}W_{0j^{}j}`$ in the rotated Hamiltonian $$_{TT}^{}(W_0)=\stackrel{~}{\mathrm{\Lambda }}_{ijkl}\overline{T}_{Li}\gamma ^\mu T_{Lj}\overline{T}_{Rk}\gamma _\mu T_{Rl},$$ (24) also have rational phases. This follows directly from the fact that nonzero $`\mathrm{\Lambda }`$’s align phases. If the phases are rational and $`\mathrm{\Lambda }_{i^{}j^{}kl}0`$, then the CKM matrix $`K`$ is real and $`\varphi _{j^{}k}\varphi _{i^{}l}=\chi _{Lj^{}}\chi _{Rk}\chi _{Li^{}}+\chi _{Rl}=0`$ (mod $`\pi `$). The phase of an individual term in the sum for $`\stackrel{~}{\mathrm{\Lambda }}_{ijkl}`$ is then $`\varphi _{j^{}j}\varphi _{i^{}i}=\chi _{Lj^{}}\chi _{Rj}\chi _{Li^{}}+\chi _{Ri}=\chi _{Rk}\chi _{Rj}\chi _{Rl}+\chi _{Ri}`$ (mod $`\pi `$), a rational phase which is the same for all terms in the sum over $`i^{},j^{}`$. One example of these phenomena is provided by an $`SU(3)`$ model in which the nonzero $`\mathrm{\Lambda }`$’s are: $`\mathrm{\Lambda }_{1111}=\mathrm{\Lambda }_{1221}=\mathrm{\Lambda }_{2112}=\mathrm{\Lambda }_{1212}=\mathrm{\Lambda }_{2121}=1.0`$ $`\mathrm{\Lambda }_{1122}=1.5,\mathrm{\Lambda }_{1133}=1.4`$ $`\mathrm{\Lambda }_{1331}=\mathrm{\Lambda }_{3113}=1.6,\mathrm{\Lambda }_{1313}=\mathrm{\Lambda }_{3131}=1.8`$ $`\mathrm{\Lambda }_{1222}=\mathrm{\Lambda }_{2122}=\mathrm{\Lambda }_{2212}=\mathrm{\Lambda }_{2221}=0.501.1.`$ (25) These want to align $`\varphi _{11}=\varphi _{22}=\varphi _{33}=\varphi _{12}=\varphi _{21}`$ and $`\varphi _{13}=\varphi _{31}`$. Phases $`\varphi _{23}`$ and $`\varphi _{32}`$ are not linked by the $`\mathrm{\Lambda }`$’s. The effect of varying $`\mathrm{\Lambda }_{1222}`$ from 0.5 to 1.1 is illustrated in Fig. 1. The phases start out aligned and rational, indeed, $`W=\mathrm{exp}(2i\pi /3)\times 1`$, and the vacuum energy (in units of $`\mathrm{\Delta }_{TT}`$) remains constant at $`6.20`$. At $`\mathrm{\Lambda }_{1222}0.725`$, it becomes energetically favorable for $`W`$ to become nondiagonal. The phases are still aligned and rational, equal to $`2\pi /3`$ (mod $`\pi /2`$), and a technipion mass becomes zero here. Now the energy decreases as $`\mathrm{\Lambda }_{1222}`$ is increased. At $`\mathrm{\Lambda }_{1222}1.015`$, a second transition occurs in which rational phases are no longer possible, and a different technipion’s mass goes through zero. At $`\mathrm{\Lambda }_{1222}1.045`$, a transition occurs back to rational phases, all equal to $`\pi /3`$ (mod $`\pi )`$, and the same technipion’s mass vanishes again. Throughout this variation of $`\mathrm{\Lambda }_{1222}`$, the other six technipion squared masses remain fairly constant with values between 5 and 15. Thus, in this example, the two technipion masses shown in Fig. 1 are always quite light. ## 4. Summary and Outlook We have numerically studied vacuum alignment in a class of theories in which electroweak and flavor symmetries are dynamically broken by gauge interactions alone. To make these intial studies tractable, we considered extended technicolor with $`N`$ doublets of a single type of technifermion, $`T_{iL,R}`$, transforming according to a complex representation of $`SU(N_{TC})`$ but as $`SU(3)_C`$ singlets. These were coupled by ETC to $`n`$ quark doublets, $`q_{aL,R}`$. In the absence of an explicit model for ETC, we assumed its broken gauge interactions could produce any desired four–fermion interaction of the form $`^{}`$ in Eq. (2). As usual, we assumed that ETC commutes with electroweak $`SU(2)`$, but not $`SU(N_{TC})SU(3)_CU(1)`$ . We also assumed, quite naturally, that ETC breaking preserves CP–invariance so that the $`\mathrm{\Lambda }`$ parameters in $`^{}`$ are real. We focussed on the technifermion sector in this paper. This restriction determines the vacuum–aligning matrix $`W_0`$ of technifermions up to tiny, but potentially important corrections of order $`\overline{q}q_{ETC}/\overline{T}T_{ETC}10^{10}`$. The problem is then simplified both numerically and analytically to minimizing the vacuum energy $`E_{TT}(W)`$ in the subspace of up and down–block diagonal $`W`$–matrices which conserve electric charge. We need then only study the alignment problem in a single charge sector. To ensure that no technipions remain massless other than the three associated with electroweak symmetry, the ETC parameters $`\mathrm{\Lambda }_{ijkl}^{U,D}`$ in $`_{TT}^{}`$ must be chosen so that there are no $`SU(N)SU(N)`$ Goldstone boson in either $`U`$ or $`D`$ sector. We found several interesting features of vacuum alignment: 1. A technipion mass may vanish to first order in the symmetry breaking perturbation even if its chiral charge does not commute with $`_{TT}^{}(W_0)`$. This differs from what happens in QCD and $`\mathrm{\Sigma }`$–model–like effective Lagrangians where the symmetries of the perturbation are manifest. The reason for this difference is the four–fermion nature of $`_{TT}^{}`$ and the symmetries of the zeroth–order ground state $`|\mathrm{\Omega }`$. 2. The real parameter $`\mathrm{\Lambda }_{ijkl}^{U,D}`$ links the $`W^{U,D}`$ phases $`\varphi _{jk}^{U,D}`$ and $`\varphi _{il}^{U,D}`$. If allowed by unitarity of $`W`$, these phases are then equal or differ by $`\pi `$. If there is complete phase alignment, all phases are equal to integer multiples of $`2\pi /N`$, mod $`\pi `$. If only partially aligned, the phases are integer multiples of $`\pi /N^{}`$ for one or more $`N^{}N`$. If phase alignment is inconsistent with unitarity, the phases are irrational multiples of $`\pi `$. 3. Rational phase sets are natural in the sense that they remain unchanged for a finite range of $`\mathrm{\Lambda }`$ parameters. In $`\mathrm{\Lambda }`$–space, the rational phase solutions to vacuum alignment form discrete islands in a sea of irrational phase solutions. 4. A massless (to first order) Goldstone boson appears when the $`\mathrm{\Lambda }`$’s take on critical values defining the boundary between rational and irrational phases. Thus, exceptionally light technipions are not at all uncommon and are a new phenomenological consequence of vacuum alignment. Vacuum alignment in the quark sector and the central issue of quark CP–violation will be addressed in a subsequent paper. It is obvious from Eq. (12) that irrational phases in the technifermion matrix $`W_0`$ will induce strong CP–violation for quarks: $`\nu _q0`$. It is therefore fortunate that rational phases occur naturally. They may permit a dynamical theory of quark flavor in which only weak CP–violation occurs and in which there is no axion. ## Acknowledgements We thank Sekhar Chivukula for his helpful comments on the manuscript. The research of KL and TR is supported in part by the Department of Energy under Grant No. DE–FG02–91ER40676. The research of EE is supported by the Fermi National Accelerator Laboratory, which is operated by Universities Research Association, Inc., under Contract No. DE–AC02–76CHO3000.
warning/0001/cond-mat0001423.html
ar5iv
text
# Statistics and noise in a quantum measurement process ## Introduction. The long-standing interest in fundamental questions of the quantum measurement received new impetus by the experimental progress in mesoscopic physics and increasing activities in quantum state engineering. The basic idea is to use as meter a device, able to carry a macroscopic current, which is coupled to the quantum system such that the conductance depends on the quantum state. By monitoring the current one performs a quantum measurement, which, in turn, causes a dephasing of the quantum system . The dephasing has been demonstrated in an experiment of Buks et al. where a quantum dot is embedded in one arm of a ‘which-path’ interferometer. The current through a quantum point contact (QPC) in close proximity to the dot suppresses the interference. However, since passing electrons interact with the current only for a short dwell-time in the dot, the meter fails to distinguish between two possible paths of individual electron; only a reduction of interference has been observed. This situation is referred to as a weak measurement. For a strong quantum measurement a different setting is needed, where a closed quantum system is observed by a meter. Then a sufficiently long observation may provide information about the quantum state. This situation is realized when a single-electron transistor (SET) is coupled to a Josephson junction single-charge box, which for suitable parameters serves as a quantum bit (qubit) . The analysis of the time evolution of the density matrix of the coupled system demonstrates the mutual influence between qubit and meter, i.e. measurement and dephasing . The measurement process is characterized by three time scales. On the shortest, the dephasing time $`\tau _\phi `$, the phase coherence between two eigenstates of the qubit is lost, while their occupation probabilities remain unchanged. Later measurement-induced transitions mix the eigenstates, changing their occupation probabilities on a time scale $`t_{\mathrm{mix}}>\tau _\phi `$ and erasing information about the initial state of the qubit. The origin of the mixing is that the charge operator (the measured quantity) and the qubit’s Hamiltonian do not commute. The third time scale appears in the dynamics of the current in the SET. Consider the probability distribution $`P(m,t)`$ that $`m`$ electrons have tunneled through the SET by time $`t`$. It was shown that after a certain time $`t_{\mathrm{meas}}`$ information about the state of the qubit can be extracted by reading out $`m`$. As expected from the basic principles of quantum mechanics the observation of the qubit first of all disturbs its state. Hence $`t_{\mathrm{meas}}\tau _\phi `$. The measurement process is only effective if the mixing is slow, $`t_{\mathrm{mix}}t_{\mathrm{meas}}`$. In the opposite limit of strong mixing the information about the qubit’s state is lost before a read-out is achieved. The distribution function $`P(m,t)`$ describes the statistics of the charge which has tunneled. The distribution of the current in the SET $`p(I,t)`$ and current-current correlations require, furthermore, the knowledge of correlations of the values of $`m`$ at different times. In earlier papers on the statistics in a SET or a QPC the behavior of $`P(m,t)`$ at times shorter than $`t_{\mathrm{mix}}`$ was derived. Effects due to the additional knowledge acquired by an observer , and the possible influence of the wave-function collapse on the monitored current were also discussed. Here we develop a systematic approach, based on the time evolution (Schrödinger equation) of the density matrix of the coupled system. This approach allows us to study averages and correlators of the current and charge. Since, due to shot noise, instantaneous values of the current fluctuate strongly, we study the current $`\overline{I}`$, averaged over a finite time interval $`\mathrm{\Delta }t`$. We calculate the mixing time in a SET and derive analytic expressions for $`p(\overline{I},\mathrm{\Delta }t,t)`$, as well as $`P(m,t)`$, valid on both short and long time scales. We study the noise spectrum of the current and find that in the limit of strong measurement ($`t_{\mathrm{meas}}<t_{\mathrm{mix}}`$) the long-time dynamics is characterized by telegraph noise, with jumps between two possible current values, corresponding to two qubit’s eigenstates. The results are of immediate experimental interest. A recent experiment demonstrated the quantum coherence in a macroscopic superconducting electron box , but the coherence time was limited by the measuring device. The SET-based measurement should extend the coherence time, which combined with experimental progress in fast measurement techniques should increase the maximum number of coherent quantum manipulations. Master equation for the measurement by a SET. The system of a qubit coupled to a SET is shown in Fig. 1. The qubit is a Josephson junction in the Coulomb blockade regime. Its dynamics is limited to a two-dimensional Hilbert space spanned by two charge states, with $`n=0`$ or $`1`$ extra Cooper pair on a superconducting island. The island is coupled capacitively to the middle island of the SET, influencing the transport current. The SET is kept in the off-state during manipulations on the qubit , with no dissipative current and no additional decoherence. To perform the measurement, the transport voltage is switched to a sufficiently high value, so that the current starts to flow in the SET. The Hamiltonian of the system is given by $$=_{\mathrm{SET}}+_{\mathrm{qb}}^N+_\mathrm{T}+_\psi .$$ (1) The first term is the charging energy of the transistor, quadratic in the charge $`Ne`$ on the middle island, $`_{\mathrm{SET}}=E_{\mathrm{SET}}(NN_0)^2`$. The induced charge $`N_0e`$ is determined by the gate voltage $`V_\mathrm{g}`$ and other voltages in the circuit. The Hamiltonian of the qubit, $`_{\mathrm{qb}}^N=_{\mathrm{qb}}+N_{\mathrm{int}}`$, includes the Hamiltonian of the uncoupled qubit $`_{\mathrm{qb}}`$ and the Coulomb interaction with the SET $`N_{\mathrm{int}}`$ (split into two terms for later convenience.) In the basis of the qubit’s charge states they are given by $`_{\mathrm{qb}}=\frac{1}{2}(E_{\mathrm{qb}}\widehat{\sigma }_zE_\mathrm{J}\widehat{\sigma }_x)`$ and $`N_{\mathrm{int}}=E_{\mathrm{int}}N\widehat{\sigma }_z`$. The charging energy scales $`E_{\mathrm{SET}}`$, $`E_{\mathrm{qb}}`$ and $`E_{\mathrm{int}}`$ are determined by capacitances in the circuit, while $`E_\mathrm{J}`$ is the Josephson coupling. We consider the eigenstates of $`_{\mathrm{qb}}`$, $`|0`$ and $`|1`$, as the logical states of the qubit. In this basis, $`_{\mathrm{qb}}`$ is diagonal, with the level spacing $`\mathrm{\Delta }E(E_\mathrm{J}^2+E_{\mathrm{qb}}^2)^{1/2}`$, while $`_{\mathrm{int}}=\left(\begin{array}{cc}E_{\mathrm{int}}^{}& E_{\mathrm{int}}^{}\\ E_{\mathrm{int}}^{}& E_{\mathrm{int}}^{}\end{array}\right)`$, where $`E_{\mathrm{int}}^{}E_{\mathrm{int}}E_{\mathrm{qb}}/\mathrm{\Delta }E`$ and $`E_{\mathrm{int}}^{}E_{\mathrm{int}}E_\mathrm{J}/\mathrm{\Delta }E`$. The term $`_\psi `$ describes the Fermions in the island and electrodes of the SET, while $`_\mathrm{T}`$ governs the tunneling in the SET. Here we assume weak coupling to the environment, with relaxation slower than the SET-induced mixing. The opposite limit is dicussed in Ref. . The full density matrix can be reduced to $`\rho _{jN^{}m^{}}^{iNm}(t)`$ by tracing over microscopic degrees of freedom and keeping track only of $`N`$ and $`m`$, the number of electrons which have tunneled through the SET. Here $`i,j=0,1`$ refer to a qubit’s basis. A closed set of equations can be derived for $`\rho _N^{ij}(m)\rho _{jNm}^{iNm}`$, the diagonal entries of the density matrix in $`N`$ and $`m`$ . Solving these equations, we analyze the evolution of the reduced density matrix of the qubit, $`\varrho ^{ij}(t)_{N,m}\rho _N^{ij}(m,t)`$, as well as $`P(m,t)_{N,j}\rho _N^{jj}(m,t)`$ and other statistical characteristics of the current in the SET. At low temperatures and transport voltages only two charge states of the middle island of the SET, with $`N`$ and $`N+1`$ electrons, contribute to the dynamics. Translational invariance in $`m`$-space suggests the Fourier transformation $`\rho _N^{ij}(k)_me^{\mathrm{i}km}\rho _N^{ij}(m)`$. Expanding in the tunneling term to lowest order, we obtain the following master equation (cf. Refs. ): $`\mathrm{}{\displaystyle \frac{d}{dt}}\left(\begin{array}{c}\widehat{\rho }_N\\ \widehat{\rho }_{N+1}\end{array}\right)`$ $`+`$ $`\left(\begin{array}{c}\mathrm{i}[_{\mathrm{qb}},\widehat{\rho }_N]\\ \mathrm{i}[_{\mathrm{qb}}+_{\mathrm{int}},\widehat{\rho }_{N+1}]\end{array}\right)`$ (6) $`=`$ $`\left(\begin{array}{cc}\stackrel{ˇ}{\mathrm{\Gamma }}_L& e^{\mathrm{i}k}\stackrel{ˇ}{\mathrm{\Gamma }}_R\\ \stackrel{ˇ}{\mathrm{\Gamma }}_L& \stackrel{ˇ}{\mathrm{\Gamma }}_R\end{array}\right)\left(\begin{array}{c}\widehat{\rho }_N\\ \widehat{\rho }_{N+1}\end{array}\right).`$ (11) Here the operators $`\stackrel{ˇ}{\mathrm{\Gamma }}_L\widehat{\rho }`$ $``$ $`\mathrm{\Gamma }_L\widehat{\rho }+\pi \alpha _L[_{\mathrm{int}},\widehat{\rho }]_+,`$ (12) $`\stackrel{ˇ}{\mathrm{\Gamma }}_R\widehat{\rho }`$ $``$ $`\mathrm{\Gamma }_R\widehat{\rho }\pi \alpha _R[_{\mathrm{int}},\widehat{\rho }]_+.`$ (13) represent the tunneling rates in the left and right junctions, with $`\alpha _{\mathrm{L}/\mathrm{R}}R_\mathrm{K}/(8\pi ^3R_{\mathrm{L}/\mathrm{R}}^\mathrm{T})`$ being the tunnel conductance of the junctions in units of the resistance quantum $`R_\mathrm{K}=h/e^2`$. The rates are fixed by the potentials $`\mu _L`$, $`\mu _R=\mu _L+V_{\mathrm{tr}}`$ of the leads: $`\mathrm{\Gamma }_L=2\pi \alpha _L[\mu _L(12N_0)E_{\mathrm{SET}}]`$ and $`\mathrm{\Gamma }_R=2\pi \alpha _R[(12N_0)E_{\mathrm{SET}}\mu _R]`$. They define the tunneling rate $`\mathrm{\Gamma }=\mathrm{\Gamma }_L\mathrm{\Gamma }_R/(\mathrm{\Gamma }_L+\mathrm{\Gamma }_R)`$ through the SET. The last terms in Eqs. (12,13) make these rates sensitive to the qubit’s state. The initial condition at the beginning of the measurement, written down in the logical basis, $$\rho _N^{ij}(m,t=0)=\left(\begin{array}{cc}|a|^2& ab^{}\\ a^{}b& |b|^2\end{array}\right)w_N\delta _{m0},$$ (14) describes the qubit in a pure state $`a|0+b|1`$ and the SET in the zero-voltage equilibrium state. One can assume that $`w_N=1`$ and $`w_{N+1}=0`$. ## Reduction of the master equation. In general the dynamics of the qubit’s density matrix $`\widehat{\varrho }`$, described by the master equation (11), is complicated since dephasing (decay of the off-diagonal entries) and mixing (relaxation of the diagonal to their stationary values) may occur on similar time scales, $`\tau _\phi t_{\mathrm{mix}}`$. However, under suitable conditions the mixing is slow, which is the prerequisite for a measurement process. This is the case, if the qubit operates in the regime with dominant charging energy: $$|E_{\mathrm{qb}}|,|E_{\mathrm{qb}}+2E_{\mathrm{int}}|E_\mathrm{J}.$$ (15) Weak ($`E_{\mathrm{int}}\mathrm{\Delta }E`$) or strong ($`E_{\mathrm{int}}\mathrm{\Delta }E`$) coupling to the SET can be considered. In the latter case a faster measurement is achieved (see Eq. (29)). We first analyze the dynamics without mixing, i.e. we put $`E_{\mathrm{int}}^{}=0`$. In this case the time evolutions of $`\rho _N^{ij}`$ for the four different pairs of indices $`ij`$ are decoupled from each other, each being characterized by two eigenmodes. The absence of mixing, further, implies the conservation of occupations of the logical states $`\varrho ^{ii}`$, for $`i=0,1`$, and we find two ‘Goldstone’ modes for $`k1`$, with eigenvalues $$\lambda _+^{ii}(k)\mathrm{i}\mathrm{\Gamma }^ik\frac{1}{2}f^i\mathrm{\Gamma }^ik^2.$$ (16) Here $`\mathrm{\Gamma }^i\mathrm{\Gamma }_L^i\mathrm{\Gamma }_R^i/(\mathrm{\Gamma }_L^i+\mathrm{\Gamma }_R^i)`$ are the tunneling rates through the SET, $`\mathrm{\Gamma }_L^{0/1}=\mathrm{\Gamma }_L\pm \pi \alpha _LE_{\mathrm{int}}^{}`$ and $`\mathrm{\Gamma }_R^{0/1}=\mathrm{\Gamma }_R\pi \alpha _RE_{\mathrm{int}}^{}`$ are the tunneling rates in the junctions for two logical states (cf. Eqs. (12,13)), and $`f^i12\mathrm{\Gamma }^i/(\mathrm{\Gamma }_L^i+\mathrm{\Gamma }_R^i)`$ are the Fano factors responsible for the shot noise reduction. The other two eigenmodes decay fast, with the rates $`\lambda _{}^{ii}(\mathrm{\Gamma }_L^i+\mathrm{\Gamma }_R^i)`$. The analysis of the eigenvalues, $`\lambda _\pm ^{01}(k)=[\lambda _\pm ^{10}(k)]^{}`$, of the four off-diagonal modes in $`ij`$, reveals the dephasing time $`\tau _\phi `$ of the qubit by the measurement, i.e., the decay time of $`\varrho ^{01}=\rho _N^{01}(k=0)+\rho _{N+1}^{01}(k=0)`$. It is given by $`\tau _\phi ^1=4\mathrm{\Gamma }E_{\mathrm{int}}^{\mathrm{\hspace{0.33em}2}}/(\mathrm{\Gamma }_L+\mathrm{\Gamma }_R)^2`$ if $`E_{\mathrm{int}}^{}\mathrm{\Gamma }_L+\mathrm{\Gamma }_R`$, and $`\tau _\phi ^1=w_N\mathrm{\Gamma }_L+w_{N+1}\mathrm{\Gamma }_R`$ in the opposite case. The picture is modified by the mixing at finite $`E_{\mathrm{int}}^{}`$. We find that the mixing may be treated perturbatively if $`|\lambda _\pm ^{01}|E_{\mathrm{int}}^{}`$, which turns to be the case if the condition (15) holds. Then, in the second order, the degeneracy between the long-living modes (16) is lifted and the long-time evolution of the occupations $`\rho ^{ii}(k)=\rho _N^{ii}(k)+\rho _{N+1}^{ii}(k)`$ is given by a reduced master equation, $`{\displaystyle \frac{d}{dt}}\left(\begin{array}{cc}\rho ^{00}(k)& \\ \rho ^{11}(k)& \end{array}\right)=M_{\mathrm{red}}\left(\begin{array}{cc}\rho ^{00}(k)& \\ \rho ^{11}(k)& \end{array}\right),`$ (21) $`M_{\mathrm{red}}=\left(\begin{array}{cc}\lambda _+^{00}(k)& 0\\ 0& \lambda _+^{11}(k)\end{array}\right)+{\displaystyle \frac{1}{2}}\mathrm{\Gamma }_{\mathrm{mix}}\left(\begin{array}{cc}1& 1\\ 1& 1\end{array}\right).`$ (26) For the mixing rate, $`\mathrm{\Gamma }_{\mathrm{mix}}`$, we obtain $`\mathrm{\Gamma }_{\mathrm{mix}}`$ (27) $`{\displaystyle \frac{4\mathrm{\Gamma }E_\mathrm{J}^2E_{\mathrm{int}}^2}{\mathrm{\Delta }E^2(\mathrm{\Delta }E+2E_{\mathrm{int}}^{})^2+[\mathrm{\Gamma }_R\mathrm{\Delta }E+\mathrm{\Gamma }_L(\mathrm{\Delta }E+2E_{\mathrm{int}}^{})]^2}}.`$ (28) To understand the role of the mixing we assume first $`\mathrm{\Gamma }_{\mathrm{mix}}=0`$ in Eqs. (21,26). Then, for the initial condition (14) we obtain $`\rho ^{00}(k)=|a|^2e^{\lambda _+^{00}(k)t}`$, $`\rho ^{11}(k)=|b|^2e^{\lambda _+^{11}(k)t}`$, and $`P(k,t)_mP(m,t)e^{ikm}=\rho ^{00}(k)+\rho ^{11}(k)`$. From this we obtain the distribution $`P(m,t)`$, which evolves from a peak $`\delta (m)`$ at $`t=0`$ into two peaks with weights $`|a|^2`$ and $`|b|^2`$, moving in $`m`$-space with velocities $`\mathrm{\Gamma }^0`$ and $`\mathrm{\Gamma }^1`$, and with widths growing as $`\sqrt{2f^i\mathrm{\Gamma }^it}`$. The peaks separate after a time $$t_{\mathrm{meas}}=\left(\frac{\sqrt{2f^0\mathrm{\Gamma }^0}+\sqrt{2f^1\mathrm{\Gamma }^1}}{\mathrm{\Gamma }^0\mathrm{\Gamma }^1}\right)^2.$$ (29) Thus measuring the charge $`m`$ after $`t_{\mathrm{meas}}`$ constitutes a strong quantum measurement . However, at longer times $`t>\mathrm{\Gamma }_{\mathrm{mix}}^1`$ the mixing spoils this picture. In particular, the occupations of the logical states relax to the equal-weight distribution: $`\varrho ^{00}(t)\varrho ^{11}(t)\mathrm{exp}(\mathrm{\Gamma }_{\mathrm{mix}}t)`$. Therefore the two-peak structure appears only in the interval between $`t_{\mathrm{meas}}t\mathrm{\Gamma }_{\mathrm{mix}}^1`$. The measurement can be performed only if $`t_{\mathrm{meas}}\mathrm{\Gamma }_{\mathrm{mix}}^1`$ The measurement takes longer than the dephasing, $`t_{\mathrm{meas}}\tau _\phi `$. Such measurement can be called non-efficient : the information about the qubit is contained in the SET already after $`\tau _\phi `$, but can be read out from the current only later. The quantum measurement with a QPC can be described in a similar way. The Coulomb interaction of the qubit with the current results in two tunneling rates $`\mathrm{\Gamma }^{0/1}=\overline{\mathrm{\Gamma }}\pm \delta \mathrm{\Gamma }/2`$ for two qubit’s states. Tracing out microscopic degrees of freedom one arrives at the master equation for the Fourier transform of the density matrix $`\rho ^{ij}(m)`$, which can be rewritten as $`\mathrm{}{\displaystyle \frac{d}{dt}}\widehat{\rho }+\mathrm{i}[_{\mathrm{qb}},\widehat{\rho }]`$ $`=`$ $`\left[\overline{\mathrm{\Gamma }}\widehat{\rho }+{\displaystyle \frac{1}{2}}\delta \mathrm{\Gamma }[\widehat{\sigma }_z,\widehat{\rho }]_+\right](e^{\mathrm{i}k}1)`$ (31) $`(4\tau _\phi )^1e^{\mathrm{i}k}[\widehat{\sigma }_z,[\widehat{\sigma }_z,\widehat{\rho }]].`$ One can show that $`\tau _\phi ^1\frac{1}{2}(\sqrt{\mathrm{\Gamma }^0}\sqrt{\mathrm{\Gamma }^1})^2`$ is the decay rate of $`\rho ^{01}`$. For $`\rho ^{ii}`$ the eigenvalues are given by (16), without Fano factors. The measurement time (29) and the dephasing time coincide, implying the 100% efficiency. Under the condition $`E_\mathrm{J}\mathrm{max}(\mathrm{\Delta }E,\tau _\phi ^1)`$ the perturbative treatment produces the reduced master equation (21,26), with the mixing rate $$\mathrm{\Gamma }_{\mathrm{mix}}=E_\mathrm{J}^2\frac{\tau _\phi }{1+\mathrm{\Delta }E^2\tau _\phi ^2}.$$ (32) A phenomenon, termed the Zeno (or watchdog) effect, can be seen in the limit $`\tau _\phi ^1\mathrm{\Delta }E`$: the stronger is the measurement, quantified by $`\tau _\phi ^1`$, the weaker is the rate of jumps between the eigenstates, $`\mathrm{\Gamma }_{\mathrm{mix}}E_\mathrm{J}^2\tau _\phi `$. The analysis of the SET mixing rate (27) in terms of the Zeno effect is more complicated. The rates $`\tau _\phi ^1`$ and $`\mathrm{\Gamma }_{\mathrm{mix}}`$ depend in this case on several parameters and no simple relation between $`\mathrm{\Gamma }_{\mathrm{mix}}`$ and $`\tau _\phi `$ is found. However, in the regime $`E_{\mathrm{int}}\mathrm{\Gamma }_L\mathrm{\Gamma }_R|\mathrm{\Delta }E|`$, these two rates change in opposite directions as functions of $`\mathrm{\Gamma }_L\mathrm{\Gamma }_R`$, which is reminiscent of the Zeno physics. Statistics of charge and current. The results of this section apply to the SET and QPC alike. The statistical quantities studied depend on the initial density matrix (14): e.g., $`P(m,t)=P(m,t|\rho _0)`$. In the two-mode approximation (21,26) this reduces to a dependence on $`\gamma _0\varrho ^{00}\varrho ^{11}=|a|^2|b|^2`$. We solve Eq. (21) to obtain $`P(m,t|\gamma _0)=\mathrm{Tr}_{\mathrm{qb}}[U(m,t)\rho _0]`$, where $`U(m,t)`$ is the inverse Fourier transform of $`U(k,t)\mathrm{exp}[M_{\mathrm{red}}(k)t]`$. If $`\mathrm{\Gamma }^{0/1}=\overline{\mathrm{\Gamma }}\pm \delta \mathrm{\Gamma }/2`$ are close, the resulting distribution is $$P(m,t|\gamma _0)=\underset{\delta m}{}\stackrel{~}{P}(m\delta m,t|\gamma _0)\frac{e^{\delta m^2/2f\overline{\mathrm{\Gamma }}t}}{\sqrt{2\pi f\overline{\mathrm{\Gamma }}t}}.$$ (33) The first term in the convolution (33) contains two delta-peaks, corresponding to two qubit’s logic states: $`\stackrel{~}{P}(m,t|\gamma _0)=P_{\mathrm{pl}}({\displaystyle \frac{m\overline{\mathrm{\Gamma }}t}{\delta \mathrm{\Gamma }t/2}},{\displaystyle \frac{1}{2}}\mathrm{\Gamma }_{\mathrm{mix}}t\left|\gamma _0\right)`$ (34) $`+e^{\mathrm{\Gamma }_{\mathrm{mix}}t/2}\left[|a|^2\delta \left(m\mathrm{\Gamma }^0t\right)+|b|^2\delta \left(m\mathrm{\Gamma }^1t\right)\right].`$ (35) On the time scale $`t_{\mathrm{mix}}\mathrm{\Gamma }_{\mathrm{mix}}^1`$ the peaks disappear; instead a plateau arises. It is described by $`P_{\mathrm{pl}}(x,\tau |\gamma _0)=e^\tau {\displaystyle \frac{\mathrm{\Gamma }_{\mathrm{mix}}}{2\delta \mathrm{\Gamma }}}\{I_0\left(\tau \sqrt{1x^2}\right)`$ (36) $`+(1+\gamma _0x)I_1\left(\tau \sqrt{1x^2}\right)/\sqrt{1x^2}\},`$ (37) at $`|x|<1`$ and $`P_{\mathrm{pl}}=0`$ for $`|x|>1`$. Here $`I_0`$, $`I_1`$ are the modified Bessel functions. At longer times the plateau transforms into a narrow peak centered around $`m=\overline{\mathrm{\Gamma }}t`$. The Gaussian in Eq. (33) arises due to shot noise. Its effect is to smear out the distribution (see Fig. 2a). We also calculate $`P_2(m,t;m^{},t^{}|\rho _0)`$, the joint probability to have $`m`$ electrons at $`t`$ and $`m^{}`$ electrons at $`t^{}`$. This allows us to obtain the probability distribution $$p(\overline{I},\mathrm{\Delta }t,t|\rho _0)\underset{m}{}P_2(m+\overline{I}\mathrm{\Delta }t,t+\mathrm{\Delta }t;m,t|\rho _0)$$ (38) of the current $`\overline{I}_t^{t+\mathrm{\Delta }t}I(t^{})𝑑t^{}`$ averaged over the time interval $`\mathrm{\Delta }t`$. The evolution is Markovian, and we obtain: $`P_2(m,t;m+\mathrm{\Delta }m,t+\mathrm{\Delta }t)=\mathrm{Tr}_{\mathrm{qb}}\left[U(\mathrm{\Delta }m,\mathrm{\Delta }t)U(m,t)\rho _0\right]`$ for $`\mathrm{\Delta }t>0`$. The derivation of $`p(\overline{I},\mathrm{\Delta }t,t)`$ thus reduces to the calculation of the charge distribution (33) for different initial conditions: $$p(\overline{I},\mathrm{\Delta }t,t|\gamma _0)=P(m=\overline{I}\mathrm{\Delta }t,\mathrm{\Delta }t|e^{\mathrm{\Gamma }_{\mathrm{mix}}t}\gamma _0).$$ (39) The behavior of $`p(\overline{I},\mathrm{\Delta }t,t)`$ is shown in Fig. 2. A strong quantum measurement is achieved if $`t_{\mathrm{meas}}<\mathrm{\Delta }t<t_{\mathrm{mix}}`$, at times $`t<t_{\mathrm{mix}}`$ (see Fig. 2b). In this case the measured current is close to either $`\mathrm{\Gamma }^0`$ or $`\mathrm{\Gamma }^1`$, with probabilities $`|a|^2`$ and $`|b|^2`$, respectively. At longer times a typical current pattern is a telegraph signal jumping between $`\mathrm{\Gamma }^0`$ and $`\mathrm{\Gamma }^1`$ on a time scale $`t_{\mathrm{mix}}`$. If $`\mathrm{\Delta }tt_{\mathrm{meas}}`$ the meter does not have enough time to extract the signal from the shot-noise background. At larger $`\mathrm{\Delta }t`$ the meter-induced mixing erases the information, partially ($`\mathrm{\Delta }tt_{\mathrm{mix}}`$, Fig. 2c) or completely ($`\mathrm{\Delta }tt_{\mathrm{mix}}`$, Fig. 2d), before it is read out. The telegraph noise behavior is also seen in the current noise. Fourier transformation of the correlator $$I(t)I(t^{})_{\rho _0}=\underset{m,m^{}}{}mm^{}_t_t^{}P_2(m,t;m^{},t^{}|\rho _0)$$ (40) gives in the stationary case the noise spectrum, $$S_I(\omega )=2e^2f\overline{\mathrm{\Gamma }}+\frac{e^2\delta \mathrm{\Gamma }^2\mathrm{\Gamma }_{\mathrm{mix}}}{\omega ^2+\mathrm{\Gamma }_{\mathrm{mix}}^2},$$ (41) as the sum of the shot- and telegraph-noise contributions. At low frequencies $`\omega \mathrm{\Gamma }_{\mathrm{mix}}`$ the latter becomes visible on top of the shot noise as we approach the regime of the strong measurement: $`S_{\mathrm{telegraph}}/S_{\mathrm{shot}}4t_{\mathrm{mix}}/t_{\mathrm{meas}}`$. To conclude, we have developed the master equation approach to study the statistics of currents in a SET or a QPC as a quantum meter. We evaluate the probability distributions and the noise spectrum of the current. We acknowledge discussions with Y. Blanter, L. Dreher, S. Gurvitz, D. Ivanov, and A. Korotkov.
warning/0001/astro-ph0001502.html
ar5iv
text
# The radio luminosity of persistent X-ray binaries ## 1 Introduction Radio synchrotron emission is observed from $`20`$% of X-ray binaries (e.g. Hjellming & Han 1995). In several cases the radio emission has been resolved into jet-like outflows reminiscent of the jet/lobe structures in AGN (e.g. Fender, Bell Burnell & Waltman 1997; Mirabel & Rodriguez 1999, Fender 2000). Much recent work, both theoretical and observational, e.g. Hjellming & Johnston (1988), Penninx (1989), Hjellming & Han (1995), Falcke & Biermann (1996) and Livio (1997), has discussed not only these clearly-resolved relativistic outflows, but also weaker unresolved radio emission from X-ray binaries. It seems increasingly possible that all radio emission from X-ray binaries could arise in such outflows. However, little serious study has been made of the properties of radio emission from the persistent sources to see whether it is consistent with this wide-ranging model. In this paper we consider the radio luminosity of the persistently accreting BHC, Z-type and Atoll-type X-ray binaries and discuss whether observations are compatible with the generic jet picture. ## 2 The sample : types of X-ray binary In this work we are interested in the radio emission from persistently accreting X-ray binaries, i.e. X-ray binary systems in which we infer from their continual and reliable detection by X-ray satellite missions that they are in a quasi-steady state of stable accretion. Van Paradijs (1995; hereafter vP95), lists 193 X-ray binaries in the most up to date catalogue available. More than 70 of these systems are transients with unstable accretion and are not considered here. In addition, more than 15 new X-ray binaries have been discovered since the publication of the catalogue, but again they are all transient, and not under study in this work. We shall ignore the distinction between low- and high-mass X-ray binaries, which is of concern for the evolution of the systems but probably not for the physics of jet formation and its coupling to the inner accretion disc. This leaves four clear subclasses of X-ray binaries : * Black Hole candidates (BHCs) * Z-type (neutron star) X-ray binaries * Atoll-type (neutron star) X-ray binaries * X-ray pulsar (neutron star) X-ray binaries In addition, the majority of systems in the vP95 catalogue are unclassified beyond being indicated as Bursters and/or Dippers. However, advances in our knowledge and understanding of the properties of the neutron-star X-ray binaries indicate that the majority of the unclassified systems are likely to be Atoll-like, although a small subclass of lower-luminosity sources is possible (Ford, van der Klis and van Paradijs, private communication). So, the majority of the persistently detected X-ray binaries are likely to be Atoll-type neutron star low-mass binaries. X-ray transients are generally low-mass BHCs, although a small number are Atoll-like. The sample of persistent BHCs (four plus two in the LMC) and Z-type sources (six), both intrinsically very luminous, is likely to be more or less complete for our Galaxy and the Magellanic Clouds. We note that Smale & Kuulkers (1999) have recently claimed that LMC X-2 may be a Z source but do not consider it as such in this work and their interpretation has yet to be confirmed. Below we shall discuss the state of our knowledge about the radio emission in these four classes of system. ### 2.1 Black hole candidates Four binary systems in our Galaxy (Cyg X-1, GX 339-4, 1E1740.7-2942 and GRS 1758-258) and two in the LMC (LMC X-1 and LMC X-3) are considered to contain persistently accreting stellar mass black holes. The majority of other BHCs are X-ray transient systems which spend most of the time in the ‘off’ state in which the accretion rate is very low and the associated X-ray emission very weak. Two of the persistent sources (Cyg X-1 and GX 339-4) also undergo state changes, typically being observed in the ‘low/hard’ state but occasionally switching to the ‘high/soft’ state, sometimes via ‘intermediate’ states (e.g. Zhang et al. 1997; Mendez & van der Klis 1997) but the dynamic range in observed luminosity is less than that of the transient systems and they can (almost) always be detected by X-ray satellite missions. An ‘off’ state, corresponding to very low X-ray flux levels, has also been observed from GX 339-4. The radio emission from Cyg X-1 and GX 339-4 in the low/hard X-ray state is relatively well studied, being observed to be weak (in comparison to more extreme systems like Cyg X-3) and steady, and roughly correlated with the X-ray emission (Pooley, Fender & Brocksopp 1999; Brocksopp et al. 1999; Hannikainen et al. 1998; Corbel et al. 2000). In addition, the radio emission is observed to drop below detectable levels during transitions to intermediate or high/soft X-ray states (e.g. Tananbaum et al. 1972; Fender et al. 1999a). The radio spectra of these two systems are approximately flat between 2 – 15 GHz (Pooley et al. 1999; Fender et al. 1997, Corbel et al. 1997), with no observed cut-off at high or low frequencies. Recent mm-wavelength observations have shown that this flat spectrum continues to at least 220 GHz in Cyg X-1 (Fender et al. 1999b). The first imaging of a compact jet from Cyg X-1 has recently been reported (Stirling, Spencer & Garrett 1998). Fewer observations have been made of the Galactic Centre systems 1E1740.7-2942 and GRS 1758-258, but they too appear to be weak and steady, with approximately flat radio spectra, (Martí 1993; Mirabel 1994) similar to Cyg X-1 and GX 339-4. In addition, both have associated weak arcmin-scale radio lobes (Mirabel 1994) providing direct evidence for the action of a jet on the ISM in the past, if not at present. LMC X-1 and LMC X-3 are very luminous X-ray sources from which radio emission has never been detected despite several deep observations (Fender et al. 1998). Table 1 summarises the observed radio flux densities and distance estimates of the persistent BHC systems. ### 2.2 Z-type X-ray binaries The six Z-type X-ray binaries (Z-sources) are believed to be low magnetic field neutron stars accreting at, or just below, the Eddington limit, and as such are amongst the most luminous persistent X-ray sources in our Galaxy (e.g. van der Klis 1995, 1999). Their name arises from a characteristic pattern they trace out in X-ray colour-colour and hardness-intensity diagrams and they are further characterised by their X-ray timing properties (e.g. van der Klis 1995, 1999). Penninx (1989) reported that four of the six Z-sources had been detected at radio wavelengths, and suggested that they all had approximately the same radio luminosity when on the ‘Horizontal Branch’ of the X-ray colour-colour or hardness-intensity diagrams – see e.g. van der Klis (1995). Following his suggestion, the remaining two Z-sources were detected at the predicted levels (Cooke & Ponman 1991; Penninx et al. 1993). These systems are usually, but not always, detected by radio observations with a sensitivity of $`0.1`$ mJy. As noted above, the most persistent radio emission is associated with the ‘Horizontal Branch’ of the X-ray emission; weaker emission appears to be associated with the ‘Normal Branch’, and radio emission appears to be ‘off’ during the ‘Flaring Branch’ (Hjellming & Han 1995 and references therein). They show generally flat radio spectra except during occasional flaring events (most prominent in Sco X-1) during which time an optically thin component is superposed. Bradshaw, Geldzahler & Fomalont (1997) have found evidence for periodic flux density variations from Sco X-1, and more recently Bradshaw, Fomalont & Geldzahler (1999) have established the distance and verified the existence of radio-emitting outflows from the system. The radio flux densities and distance estimates of the Z-sources are summarised in table 2. ### 2.3 Atoll-type X-ray binaries The Atoll-type X-ray binaries (‘Atoll sources’), like the Z sources, are low mass X-ray binaries containing low magnetic field accreting neutron stars. However, they are believed to be accreting at around an order of magnitude lower rate than the Z sources (see e.g. van der Klis 1995). As discussed above, while vP95 listed only 11 Atoll sources, it now seems likely that the majority of other low-luminosity X-ray binaries classified as Bursters and/or Dippers (and/or occasionally Transient) are also Atoll sources. Only a very small number of these systems have reported detections at radio wavelengths (Hjellming & Han 1995). Grindlay & Seaquist (1986) report the detection of a weak ($`0.49\pm 0.12`$ mJy) radio signal from 4U 1820-30, which has not been confirmed. Martí et al. (1998) report the possible detection of variable radio emission from GX 354+0 which peaks at a level of $`0.5`$ mJy but which is undetected to $`0.3`$ mJy in the majority of their observations. Gaensler, Stappers & Getts (1999) also report a transient radio detection of the Atoll-like millisecond X-ray pulsar SAX 1808.4-3658. The only convincing and repeated detection of persistent radio emission from an Atoll source is that of GX 13+1, an unusual system sharing some of the properties of both Atoll and Z sources (Penninx 1990; Homan et al. 1998). As already stated, the majority of other, unclassified X-ray binaries are also likely to be Atoll sources. The X-ray binary radio surveys of Nelson & Spencer (1988 – northern hemisphere) and McKie (1997; see also Spencer et al. 1997 – southern hemisphere) typically reached flux density detection limits of 2.0 and 0.2 mJy respectively. Without doubt some of the ‘miscellaneous’ X-ray binaries observed (but not detected) were unclassified low-luminosity Atoll sources; none were detected as radio sources. In addition, some neutron-star transient systems (e.g. Aql X-1) are classified as Atoll-like and have been observed to produce radio emission during outbursts (presumably at high accretion rates and/or during state transitions). The radio flux densities and distance estimates of the persistent Atoll-sources are summarised in table 3. ### 2.4 X-ray pulsars More than 30 X-ray pulsar systems, containing a high magnetic field neutron star which disrupts the inner accretion disc and channels the accretion flow onto its magnetic poles (White, Nagase & Parmar 1995; Bildsten et al. 1997), are known. Many of these systems are in high mass X-ray binaries, suggesting that the neutron stars are relatively young. Not one of the high-field X-ray pulsars has ever been convincingly detected as a synchrotron radio source (Fender et al. 1997). Martí et al. (1997) report a marginal radio detection of the X-ray pulsar system GX 1+4 but this has yet to be confirmed, and in any case may be consistent with thermal free-free emission from the red giant wind in this system. Fender et al. (1997) showed that there was a significant anticorrelation between the properties of radio emission and X-ray pulsations from X-ray binaries, which they suggested arose because of the disruption of the inner accretion disc preventing the formation of an outflow from the system. Table 4 lists ten persistent X-ray pulsar systems for which there are good limits to the radio flux density and reasonable distance estimates. None of the systems are detected as radio sources. It is worth noting that (at least) an additional six transient X-ray pulsars (4U 0115+63, GS 0834-430, A 0535-66, A 1118-616, 4U 1145-619, 4U 1416-62) have not been detected to comparable limits (Fender et al. 1997; McKie 1997; some distance estimates in Negueruela 1998). As noted above, there is a reported detection of transient radio emission from the accretion-powered millisecond X-ray pulsar SAX 1808.4-3658 (Gaensler et al. 1999). However, in nearly all its properties this system is more like an Atoll source than an X-ray pulsar, due to the low ($`10^9`$G, c.f. $`10^{11}`$G for other X-ray pulsars) magnetic field. ### 2.5 Other, peculiar, systems In addition to the classes of system described above, there are several systems which are persistently detected by X-ray missions but which are not easily classified. These include Cyg X-3, Cir X-1 and, for the past 5 years, GRS 1915+105, although this last source was almost certainly ‘off’ prior to 1994. Several of these systems, notably those mentioned above, are bright and variable radio sources. However, while the quiescent state of their radio emission is relatively weak with a flat spectrum, their radio emission is often dominated by the superposition of many components which evolve from optically thick to optically thin. It seems that radio jet production in these systems is more sporadic and violent than in the persistent sources, and it is not within the scope of this paper to try and identify in detail areas of common astrophysics between the production of radio emission in persistent and transient sources. ## 3 A common radio luminosity for persistent BHC and Z sources Fig 1 plots the mean radio flux densities and best distance estimates for the persistent BHC and Z sources, plus the anomalous Atoll source GX 13+1, from the data listed in Tables 1-4. The limits on radio emission from the other Atoll sources and the persistent X-ray pulsar systems are also indicated. The weakness of the radio emission from the X-ray pulsar and Atoll sources compared to the persistent BHC and Z sources is immediately apparent. In fact the data suggest a common mean radio luminosity for the persistent BHC and Z sources, which is far above that of the other persistent X-ray binaries. It is worth stressing that this radio luminosity is still orders of magnitude below that of the more extreme and poorly classified systems such as Cyg X-3, SS 433 etc. In order to test the goodness of fit of a common observed radio luminosity for all the sources, a straight line was fitted to the relationship between the (base ten) logarithm of centimetre radio flux density, $`S_{\mathrm{cm}}`$ (in mJy) and estimated distance, $`d`$ (in kpc). Both the intercept and gradient were treated as free parameters and our least squares fitting procedure allowed for errors on both flux and distance – the appropriate errors on the log quantities having been derived from the tabulated flux and distance errors using Monte Carlo simulations. We included only the ten galactic sources in our fits, since the two LMC sources had only upper limits on the observed flux. The resulting best fit relation was $$\mathrm{log}_{10}S_{\mathrm{cm}}=(2.03\pm 0.40)(2.46\pm 0.51)\mathrm{log}_{10}d$$ (1) ($`\chi _{\mathrm{red}}^2=0.514`$) which is clearly consistent with a $`d^2`$ relation and therefore a common luminosity. Fixing the slope of the relation to be equal to $`2`$ produced a similar fit, although now with a smaller error on the intercept: $$\mathrm{log}_{10}S_{\mathrm{cm}}=(1.71\pm 0.10)2\mathrm{log}_{10}d$$ (2) ($`\chi _{\mathrm{red}}^2=0.527`$) which is shown by the dotted line on Fig 1. Expressed in terms of flux and distance (in kpc) this relation corresponds to $$S_{\mathrm{cm}}=\frac{55\pm 13}{d^2}\mathrm{mJy}$$ (3) A possible explanation for the low values of $`\chi _{\mathrm{red}}^2`$ may be that some of the flux and distance errors tabulated in Tables 1 & 2 have been over-estimated: if the quoted errors – particularly those in distance – were correct then the data would be unlikely to lie so close to the best fit straight line shown in Fig 1. We discuss this point further below, but remark for the moment that the agreement of the data with an inverse square relation between flux and distance is clearly convincing. This is a surprising result given that both different accretion structures and outflow velocities (see below) in the neutron star and black hole systems might be expected. For a flat spectrum from 30 to 2 cm (1 to 15 GHz), this corresponds to an observed radio synchrotron luminosity of $`10^{30}`$ erg s<sup>-1</sup> ($`10^{23}`$ W) for these sources. It should be stressed however that a high-frequency cut-off in synchrotron emission has yet to be found in any of these sources, and the total (ie. radio – mm – infrared) synchrotron luminosity is likely to be orders of magnitude higher (see for example Fender et al. 1997c and Mirabel et al. 1998 where a synchrotron luminosity $`10^{36}`$ erg s<sup>-1</sup> is inferred for GRS 1915+105 from the observation of the flat spectral component to 2 $`\mu `$m). ### 3.1 The sample and selection effects As discussed in the introduction, the sources under discussion, the neutron star Z-sources and persistent BHCs, are unique in their relatively steady, bright X-ray emission, implying steady accretion at high luminosity. Given the number and sensitivity of X-ray missions over the past 30+ yr, the sample is likely to be more or less complete for the entire Galaxy, LMC and SMC. The sample is clearly dominated by sources in the vicinity of the Galactic centre, with distances between 7 – 10 kpc. This clustering of data points is suggestive of selection effects dominating the fit to the data, as a result of small numbers in the sample. However, as discussed above, the sample is effectively volume-limited and cannot be expanded. One selection effect which could significantly bias the fit would be if the Galactic centre systems were being detected only when they came up above the detection limits of typical radio observations. In this case, a mean flux density calculated from an unrepresentative sample of positive detections could be very similar for all the sources, regardless of their true mean radio flux. However, this does not appear to be the case : comprehensive high-sensitivity observations of Cyg X-2 and GX 17+2 (Hjellming et al. 1990; Penninx et al. 1988), among the most distant sources in the sample, consistently detect these sources and accurately measure the mean flux density. ## 4 Beaming and the luminosity function It has been speculated that low-level radio emission from the persistent BHC and Z-source X-ray binaries could arise in a continuous jet (e.g. Hjellming & Han 1995). It has also been suggested that the velocities of jets from accretion discs should approximately reflect the escape velocity of the central object, i.e. jets from black holes will have velocities $`0.9c`$, those from neutron stars $`0.3c`$ etc. (e.g. Livio 1997). Combining these ideas, a scenario can be envisaged whereby the low-level radio emission from persistent BHCs and Z-sources originates in compact jets of velocities $`0.9c`$ and $`0.3c`$ respectively. In the light of our result that all persistent BHC and Z-source X-ray binaries have the same mean radio luminosity at centimetre wavelengths, however, there are problems with this interpretation, based upon the Doppler boosting associated with a relativistic jet. Relativistic jets are significantly Doppler boosted; for a continuous jet at a given angle to the line of sight, $`\theta `$, the rest-frame flux of the source, $`S`$, is boosted to an apparent value, $$S^{}=SD^{2\alpha }$$ (4) where $`\alpha `$ is the spectral index of the radio emission, and the relativistic Doppler factor $`D`$ is defined as $$D=[\gamma (1\beta \mathrm{cos}\theta )]^1$$ (5) ($``$ for approaching and receding components respectively) and $`\gamma `$ is the Lorentz factor $$\gamma =(1\beta ^2)^{1/2}$$ (6) where $`\beta `$ is the velocity of the jet expressed as a fraction of $`c`$. As an example, at an angle to the line of sight of 30 degrees, a flat-spectrum ($`\alpha =0`$) symmetric jet of velocity $`0.9c`$ would appear 2.4 times brighter than the same jet at a velocity of $`0.3c`$, and 3.9 times brighter than in the rest frame. The total flux observed will be a sum of the approaching and receding jets. Note that for jets near to the plane of the sky both approaching and receding jets can be de-boosted (as is the case for GRS 1915+105; Mirabel & Rodriguez 1994; Fender et al. 1999b). Figure 2 illustrates the ratio of observed to intrinsic flux expected from symmetric jets at $`0.3c`$ and $`0.9c`$ for all inclinations. Assuming naively that any jet is approximately perpendicular to the orbital plane of the system, then the angle, $`\theta `$, of the jet to the line of sight would be equal to the inclination, $`i`$, of the orbital plane to the line of sight. Estimates of the orbital inclinations of the systems in question are very limited. A survey of the literature does not reveal any particular bias in inclination estimates for any of the classes of source, so we will assume that the inclinations are uniformly distributed in $`\mathrm{cos}i`$. It certainly appears likely that the systems show a significant spread in inclination. This immediately presents a problem for the hypothesis that the radio emission from these systems originates in Doppler boosted jets. The problem may be stated qualitatively as follows. If the ten galactic systems are viewed at a range of different inclinations, their observed fluxes will be boosted by a range of different Doppler factors. Thus, if their observed fluxes obey the inverse-square relationship with distance given by Eq. (3), their intrinsic fluxes will not in general obey this relationship. It would, therefore, seem unlikely that one should obtain such a relation between distance and observed flux by chance, since it would require a series of fortunate coincidences in order that the intrinsic fluxes and inclinations yield observed fluxes in agreement with the fitted relation. One resolution of this problem would be if all the systems were observed at approximately the same inclination, since the sources would then all be Doppler boosted by the same factor. As we have already remarked, however, this possibility appears incompatible with the inclination estimates reported in the literature. Even if we choose to regard these estimates as unreliable, it seems reasonable to suppose that the orbital planes of the ten systems should be randomly sampled from a uniform distribution over all possible orientations. It is then straightforward to show that the probability of drawing a sample of ten sources, with inclinations all lying within, say, an interval of 5 degrees, is less than $`10^{10}`$. Is it possible that the inclinations of the observed systems are selected to lie within a narrow range? One plausible mechanism for such a selection effect might be if their intrinsic fluxes were too faint to be detected, but there exists a critical inclination at which the Doppler boosting factor is sufficient to raise the observed fluxes above the detection limits of typical radio observations. As remarked above, however, our sample is not dominated by systems whose flux lies close to the detection limit. The observed fluxes listed in Tables 1 & 2 span a range of more than a factor of 20, arguing in a favour of a wide range of different inclinations consistent with the estimated limits. ### 4.1 Monte Carlo simulations We have used extensive Monte Carlo simulations to investigate the effect of beaming and the width of the intrinsic radio luminosity function (LF) of the persistent BHC and Z sources. The width is defined such that the LF is uniform in $`\mathrm{log}L`$ between $`\mathrm{log}(L/width)`$ and $`\mathrm{log}(L\times width)`$. We have defined a critical width for which, after running $`10^4`$ simulations, 90% of the sample have $`\chi _{\mathrm{red}}^2>3`$. Not surprisingly, the higher values of $`\beta `$ we consider for the jets, the narrower the intrinsic LF must be in order to produce the observed inverse square relationship. The results of the simulations are listed in Table 5. In addition to considering the same value of $`\beta `$ for both jets from black hole and neutron star systems, we also consider what may be considered the ‘canonical’ model, where jets from black hole systems have $`\beta =0.9`$ and jets from neutron star systems have $`\beta =0.3`$. It is clear from the results of the MC simulations presented in table 5 that for jet velocities in the range 0 – 0.7$`c`$ the intrinsic radio luminosity of the BHC and Z-sources must be the same to within a factor of 25 or so. This is a result of the relative unimportance of Doppler boosting at these velocities. For higher jet velocities the intrinsic luminosities need to be even closer together: for jets in both types of systems with velocities of 0.9$`c`$ they must be intrinsically within a factor of 15 in radio luminosity in order to produce the observed $`d^2`$ relation. Thus we can conclude that the classes of BHC (in the low/hard X-ray state) and Z sources have a common intrinsic radio luminosity within an order of magnitude or so. We can also use the MC simulations to quantify how much weaker radio sources the Atoll and X-ray pulsar systems are, as a class, compared to the BHC and Z-sources. We use the observed upper limits on radio emission and distance estimates given in tables 3 & 4. The maximum mean luminosities of these two classes, compared to that obtained for the combined BHC and Z-sources, are given in table 6. From this we can assert that, as a class, the mean radio luminosity of Atoll sources is more than a factor of five below that of the BHC and Z sources. For the X-ray pulsar systems the limits are even stronger; they are at least an order of magnitude fainter, as radio sources, than the BHC and Z sources. ## 5 Discussion We have found that the BHC and Z-source X-ray binaries share a common mean radio luminosity to within a factor of 15–25, depending on the velocity of the inferred outflows. The Atoll sources are $`5`$ times fainter; the X-ray pulsar systems $`10`$ times so. One reason that these results are surprising is the different inferred accretion modes for the BHCs and Z-sources. In BHCs in the low/hard X-ray state the ‘standard’ (thin, cold, optically thick) accretion disc is believed to be truncated many Schwarzchild radii from the central black hole and replaced in the inner regions by an optically thin, radiatively inefficient quasi-spherical flow (Advection-dominated accretion flows; see Svensson 1998 and references therein). However in Z and Atoll sources the ‘standard’ accretion disc is believed to reach to almost the surface of the neutron star (e.g. van der Klis 1999). ### 5.1 Jets ? We have established that Doppler boosting is unlikely to affect the observed radio luminosities of the BHC and Z sources by more than an order of magnitude. Thus, assuming the emission is incoherent, we can apply the limiting brightness temperature of $`10^{12}`$ K which results from second-order inverse Compton losses. As a result we find that the emission at 2 GHz must arise in a region $`10^{12}`$ cm (for a spherical emitting region). This is a significant size scale, larger than the inferred binary separations of most, maybe all, of the systems (e.g. $`3\times 10^{11}`$ cm for the Z source Sco X-1). A cone of opening half-angle 10 degrees in the plane of the sky would require a length of $`10^{13}`$ cm to produce the same observed surface area; angling the jet more towards the line of sight or making the opening angle smaller only increases this dimension. Similarly, attributing at least some of the observed radio flux to optically thin emission, less efficient than optically thick, also increasing the required emitting volume. Coupled with the recent imaging of collimated outflows from both Sco X-1 (Bradshaw et al. 1999) and Cyg X-1 (Stirling et al. 1998; de la Force et al. in prep) the observational evidence seems to point to extended radio-emitting outflows in all BHCs, Z sources and GX 13+1. ### 5.2 Why not Atoll sources ? Why are Z sources so much brighter radio emitters than the Atoll sources ? the inferred differences between the two classes are a stronger magnetic field and higher accretion rate in the Z sources. We do not believe that the magnetic field plays much of a role in this difference: * The inferred magnetic field in Z sources lie between those of the Atoll sources and the X-ray pulsars, both of which we shown to be significantly less luminous radio sources. * The presence of kilohertz quasi-periodic oscillations in both Z and Atoll sources (van der Klis 1999 and references therein) implies that the accretion flow in both classes is not truncated by the magnetic field and instead reached almost to the surface of the neutron star. Instead, it seems likely that it is the accretion rate which is the origin of the difference. This is supported by the radio detections of the Atoll source GX 13+1 at a similar level to the Z sources; this system is believed to be accreting at a higher rate ($`10^{17}`$ g s<sup>-1</sup> $`0.1`$ Eddington) than the other Atoll sources. In addition, we can imagine that the occasional detections of Atoll-type sources at radio wavelengths are associated with transient periods of high accretion rates comparable to those continuously occuring in the Z sources. Alternatively, or perhaps additionally, transient radio emission seems to be produced at points of change in the X-ray ‘state’ of a system. Perhaps GX 13+1 and the Z sources change X-ray ‘state’ more often, or physically in a more dramatic way, than the other Atoll sources, and hence are more prone to significant mass ejections. ### 5.3 Why not X-ray pulsars ? As mentioned in section 2.4, no X-ray pulsar has ever been detected as a synchrotron radio source. As originally suggested by Fender et al. (1997b) we believe this is due to the truncation of the inner accretion flow by strong neutron star magnetic fields which force the accreting material to flow along the field lines towards the magetic poles. We can now definitively state that the X-ray pulsars are at least one order of magnitude fainter than the BHC and Z sources as a population of radio emitters. The apparent exception to this rule is the recent detection of transient radio emission from SAX 1808.4-3658 (Gaensler et al. 1999). In fact, this observations seems to confirm, rather than violate the above hypothesis, as in SAX 1808.4-3658 the magnetic field appears to be so weak ($`10^9`$G) as to allow the nearly-Keplerian flow of material almost to the neutron star surface (Wijnands & van der Klis 1998). In this case the source is in nearly all respects Atoll-like and the detection of transient, weak, radio emission is consistent with this picture. It appears that somewhere in the range $`10^9`$$`10^{11}`$ G, the magnetic field of a neutron star becomes so strong that its affect on the inner disc structure is enough to prevent the formation of a radio-emitting outflow. ## 6 Conclusions We have investigated the radio detections and upper limits on the radio emission from persistent (i.e. non transient) X-ray binaries. Whilst always bearing in mind that the sample is not large, our conclusions are summarised in Table 6, and are : * The BHCs (in the low/hard state) and the Z sources (on the horizontal branch) share a common mean observed radio luminosity corresponding to $`(55\pm 13)/d^2`$ mJy, where $`d`$ is the distance to the source in kpc. * Depending on the degree of Doppler boosting of the radio emission, this implies a common intrinsic radio luminosity to within a factor of 25 (decreasing as Doppler boosting becomes more important to e.g. a factor of 15 if both BHCs and Z sources have jets with $`v=0.9c`$). * Upper limits on radio emission from Atoll and X-ray pulsar populations as a whole show that they are in general at least 5 and 10 times fainter, respectively, than the BHC/Z systems. * Assuming that the radio emission from BHC/Z systems arised in jets for which Doppler boosting is not very significant, we find that all these systems are likely to be generating radio-emitting outflows or jets whose physical scales are significantly larger than the binary orbits. Combining these results with knowledge of the nature of accretion in different types of X-ray binaries, we can surmise that the following physical conditions are required for formation of a radio jet : * A dipole magnetic field of $`10^{10}`$ G associated with the accreting compact object, allowing the formation of an accretion flow to $`1000`$ km which is not channeled onto the magnetic poles of the neutron star. * A high accretion rate ($`0.1`$ Eddington) and/or dramatic physical changes in the accretion mode which result in the ejection of disc material. and further that the coupling between accretion and outflow in persistent systems (excluding X-ray pulsars) is comparable for both neutron stars and black holes, and therefore probably does not require the presence of either a surface or an event horizon. Observation of exactly what causes the Atoll sources to occasionally produce radio emission, and determination of the high-frequency spectrum of the radio emission from the Z sources (to see if they, like the BHCs, possess a flat spectrum through mm wavelengths) are amongst the many important future observations to be made in this field. ## acknowledgements RPF wishes to thank Eric Ford, Guy Pooley, Michiel van der Klis, Jan van Paradijs and Ben Stappers for useful discussions. This work was supported in part by ASTRON grant 781-76-017, and in part by EC Marie Curie Fellowship ERBFMBICT 972436.
warning/0001/hep-ph0001174.html
ar5iv
text
# References The Electric Dipole Moment and CP Violation in $`BX_sl^+l^{}`$ in SUGRA Models with Nonuniversal Gaugino Masses Chao-Shang Huang, Liao Wei Institute of Theoretical Physics, Academia Sinica, P.O.Box 2735, Beijing 100080,P.R.China ## Abstract The constraints of electric dipole moments (EDMs) of electron and neutron on the parameter space in supergravity (SUGRA) models with nonuniversal gaugino masses are analyzed. It is shown that with a light sparticle spectrum , the sufficient cancellations in the calculation of EDMs can happen for all phases being order of one in the small tan$`\beta `$ case and all phases but $`\varphi _\mu `$ ($`|\varphi _\mu |\stackrel{<}{}\pi /6`$) order of one in the large tan$`\beta `$ case. This is in contrast to the case of mSUGRA in which in the parameter space where cancellations among various SUSY contributions to EDMs happen $`|\varphi _\mu |`$ must be less than $`\pi /10`$ for small $`tan\beta `$ and $`𝒪(10^2)`$ for large $`tan\beta `$. Direct CP asymmetries and the T-odd normal polarization of lepton in $`BX_sl^+l^{}`$ are investigated in the models. In the large tan$`\beta `$ case, $`A_{CP}^2`$ and $`P_N`$ for l=$`\mu `$ ( $`\tau `$) can be enhanced by about a factor of ten ( ten) and ten (three) respectively compared to those of mSUGRA. Recent observation of Re($`ϵ^{}`$/$`ϵ`$) by KTev collaboration definitely confirms the earlier NA31 experiment. This direct CP violation measurement in the Kaon system can be accommodated by the CKM phase in standard model (SM) within the theoretical uncertainties. However, the CKM phase is not enough to explain the matter-antimatter asymmetry in the universe and gives the contribution to EDMs much smaller than the limits of EDMs of electron and neutron. One needs to have new sources of CP violation and examine their phenomenological effects. There exist new sources of CP violation in SUSY theories which come from the phases of soft SUSY breaking parameters. It is well-known for a long time that in order to satisfy the current experimental limits on EDMs of electron and neutron SUSY CP-violating phases have to be much smaller($`\stackrel{<}{}10^2`$) unless sfermion masses of the first and second generations are very large ($`>`$ 1 Tev) . Recently it has been shown that various contributions to EDMs cancel with each other in significant regions of the parameter space so that the current experimental limits on the EDM of electron (EDME) and the EDM of neutron (EDMN) , $`|d_e|<4.3\times 10^{27}ecm`$ (1) and $`|d_n|<6.3\times 10^{26}ecm,`$ (2) can be satisfied for SUSY models with SUSY phases of order one and relatively light sparticle spectrum ($`<`$ 1 Tev) . In mSUGRA even in the parameter space where cancellation among various SUSY contributions to neutron EDM(EDMN) happens $`|\varphi _\mu |`$ must be less than $`\pi /10`$ for small $`tan\beta `$ and $`𝒪(10^2)`$ for large $`tan\beta `$ while the allowed range of $`\varphi _{A_0}`$ is almost unconstrained. Brhlik et al. pointed out that more sufficient cancellations happen in MSSM if gaugino masses are complex . In the letter we consider cancellation phenomena in SUGRA with nonuniversal gaugino masses. CP violation has so far only been observed in K system. It is one of the goals of the B factories presently under construction to discover and examine CP violation in the B system. Direct CP violation in $`BX_sl^+l^{}`$ in SM has been examined and the result is that it is unobservablly small . In mSUGRA the CP assymmetry of branching ratio on $`BX_sl^+l^{}`$has been given in . A detailed analysis of SUSY contributions to CP Violation in semileptonic B decays has been performed using the mass insertion approximation in . Direct CP asymmetries and the T-odd normal polarization of lepton in $`BX_sl^+l^{}`$ in mSUGRA with CP-violating phases are investigated in our previous paper . In the letter we extend the investigation to SUGRA models with nonuniversal gaugino masses after studying the allowed regions of the parameter space in the models by EDM data. In order to concentrate on the effects of the phases arising from complex gaugino masses we limit ourself to a class of SUGRA models with nonuniversal gaugino masses in which gaugino masses at high energy scale (GUT scale) are nonuniversal but scalar masses and trilinear couplings at GUT scale are still universal. Such a class of effective SUGRA models can naturally arise from string models . In this class of models, compared to the mSUGRA, there are two more new independent phases which we choose to be $`\varphi _1`$ and $`\varphi _3`$, the phases of gaugino masses $`M_1`$ and $`M_3`$, in addition to the phases $`\varphi _\mu `$ and $`\varphi _{A_0}`$. From the one loop renormalization group equations (RGEs) for $`M_i`$ (i=1,2,3) $`{\displaystyle \frac{dM_i}{dt}}={\displaystyle \frac{1}{4\pi }}b_i\alpha _iM_ii=1,2,3`$ (3) where $`\alpha _i=\frac{g_i^2}{4\pi }`$, $`t=ln(Q^2/M_{GUT}^2)`$, we know that the phases of $`M_i`$ do not run, like the phase of $`\mu `$. Let us recall the cancellation mechanism for EDME. There are only two contributions, the chargino (-sfermion loop) and neutrilino (-sfermion loop) contributions, to the EDME. The chargino contribution involves gaugino -Higgsino(g-h) mixing. The neutrilino contribution involves not only gaugino-Higgsino mixing but also gaugino-gaugino(g-g) mixing. The chargino contribution and the part of the neutrilino contribution which comes from g-h mixing have automatically opposite sign because of the opposite sign of $`\mu `$ in chargino and neutrilino mass matrices. In general, the chargino contribution in magnitude is significantly larger than the part of neutrilino contribution. Therefore, as pointed out in ref. , a cancellation can happen only if the another part of the neutrilino contribution which comes from the g-g mixing can balance some of the difference between the two contributions. For EDME the neutrilino contribution which comes from the g-g mixing is proportional to $`{\displaystyle \frac{1}{m_{\stackrel{~}{e}}^2}}m_e[A_esin(\varphi _{A_e}\varphi _1)+|\mu |tan\beta sin(\varphi _\mu +\varphi _1)]`$ (4) Therefore, given $`\varphi _\mu `$, the sign of the contribution can be controlled by choosing $`\varphi _1`$ and $`\varphi _{A_0}`$ and the magnitude of the contribution can increase by increasing $`\mu `$ and/or $`A_e`$. Because the chargino contribution is dependent on $`\mu `$ and independent on $`\varphi _{A_e}`$, it is sufficient to have a cancellation that the magnitude of the $`A_e`$ term (first term) in eq.(4) is comparable to that of the $`\mu tan\beta `$ term (second term) in eq.(4). This is easy to be down in MSSM in which $`A_e`$ and $`\mu `$ are free parameters. Thus, an almost exact cancellation can occur for the whole range of $`\varphi _\mu `$. That is exactly what happens in MSSM . However, in SUGRA models low energy properties are determined by running RGEs from the high energy scale to the electroweak(EW) scale and the radiative breaking mechanism of the EW symmetry puts constraints on CP- violating phases. As long as we limit our discussion to mass spectra less than than 1 Tev, $`M_3`$ and $`A_0`$(hence $`A_e`$) can not be too large. For small tan $`\beta `$(say $`\stackrel{<}{}2`$), the sufficient condition (i.e., the two terms in eq.(4) have size of the same order) can easily be realized in the almost whole range of $`\varphi _\mu `$ by choosing $`\varphi _{A_0}`$ and $`\varphi _1`$. For moderate and large tan$`\beta `$, it is difficult for the condition to be realized due to the limited values of $`A_0`$ (hence $`A_e`$) so that only for some limited ranges of $`\varphi _\mu `$ the EDM constraint can be satisfied. The similar (but more complicated) situation occurs for EDMN with appropriate $`\varphi _3`$ as well as $`\varphi _{A_0}`$ chosen. In order to show the important role of $`\varphi _{A_0}`$ played in the cancellation mechanism, in fig.1a and 1b we display EDME as function of $`\varphi _1`$ for $`\varphi _{A_0}`$=0 and different $`\varphi _\mu `$ for both small tan$`\beta `$ (2) and large tan$`\beta `$ (30) cases. We can see from the fig.1 that most of the range of $`\varphi _\mu `$ is excluded by EDME in both cases. For EDMN as function of $`\varphi _3`$, similar results are obtained. That is, like $`\varphi _1`$, for positive $`\varphi _\mu `$ cancellations happen in some narrow ranges within \[$`\pi `$,$`2\pi `$\] of $`\varphi _3`$ and within \[0,$`\pi `$\] for negative $`\varphi _\mu `$. When we vary the values of $`\varphi _{A_0}`$ we achieve the above mentioned results: almost whole range of $`\varphi _\mu `$ is allowed by EDME and EDMN for small tan$`\beta `$ and $`|\varphi _\mu |\stackrel{<}{}\pi /6`$ for large tan$`\beta `$(see fig.2). Moreover, because $`\varphi _1(\varphi _3)`$ is correlated with $`\varphi _\mu `$, we find that with varying $`\varphi _\mu `$ the whole range of $`\varphi _1`$ and $`\varphi _3`$ can be allowed by EDM constraints. For large tan$`\beta `$ (30) case largest $`|\varphi _\mu |`$(about $`\pi /6`$) correspond to $`\varphi _1`$ and $`\varphi _3`$ around $`\pi /2\pi /6`$, while $`\varphi _1`$ and $`\varphi _3`$ are around $`\pm \pi /2`$ when $`\varphi _\mu `$ about $`0.4`$, and when $`\varphi _\mu `$ is about $`\pm 0.2`$ they are around $`\pi /4`$. The correlated values of $`\varphi _3`$ and $`\varphi _\mu `$ are needed in analyses of $`BX_sl^+l^{}`$ and $`BX_s\gamma `$ below. Correlation between $`\varphi _\mu `$ and tan$`\beta `$, with the absolute value of soft breaking terms chosen as those in fig1 and appropriate phases chosen, is shown in fig.2 where all of the points are allowed by the experimental bounds on EDME and EDMN. One can see in the figure that $`\varphi _\mu `$ becomes more constrained as $`tan\beta `$ is increased. Nevertherless, for $`tan\beta `$ larger than 6 the allowed regions of $`\varphi _\mu `$ are almost unchanged, which means that effects of the $`A_e`$ term (and $`A_d`$ term in the case of EDMN) in eq.(4) are relatively small and the balance is provided by the $`\mu tan\beta `$ term in eq.(4). Since we also consider the large $`tan\beta `$ case, we include the two loop contribution given by D. Chang et.al. But the numerical calculations in the regions of the parameter space in which one loop EDMs satisfy the current experimental limits due to the cancellation mechanism show that it is very small compared to one loop contributions. We now turn to the calculations of the CP violation in $`BX_sl^+l^{}`$. The direct CP asymmetries in decay rate and backward-forward assymmetry for $`BX_sl^+l^{}`$ and $`\overline{B}\overline{X}_sl^+l^{}`$ are defined by $`A_{CP}^1(\widehat{s})={\displaystyle \frac{\mathrm{d}\mathrm{\Gamma }/\mathrm{d}\widehat{s}\mathrm{d}\overline{\mathrm{\Gamma }}/\mathrm{d}\widehat{s}}{\mathrm{d}\mathrm{\Gamma }/\mathrm{d}\widehat{s}+\mathrm{d}\overline{\mathrm{\Gamma }}/\mathrm{d}\widehat{s}}}={\displaystyle \frac{D(\widehat{s})\overline{D}(\widehat{s})}{D(\widehat{s})+\overline{D}(\widehat{s})}},`$ $`A_{CP}^2(\widehat{s})={\displaystyle \frac{A(\widehat{s})\overline{A}(\widehat{s})}{A(\widehat{s})+\overline{A}(\widehat{s})}}`$ (5) where $`A(\widehat{s})=3\sqrt{1{\displaystyle \frac{4t^2}{\widehat{s}}}}{\displaystyle \frac{E(\widehat{s})}{D(\widehat{s})}},`$ $`D(\widehat{s})=4\left|C_7\right|^2(1+{\displaystyle \frac{2}{\widehat{s}}})(1+{\displaystyle \frac{2t^2}{\widehat{s}}})+\left|C_{8}^{}{}_{}{}^{eff}\right|^2(2\widehat{s}+1)(1+{\displaystyle \frac{2t^2}{\widehat{s}}})+\left|C_9\right|^2\left[1+2\widehat{s}+(14\widehat{s}){\displaystyle \frac{2t^2}{\widehat{s}}}\right]`$ $`+12Re(C_{8}^{}{}_{}{}^{eff}C_7^{})(1+{\displaystyle \frac{2t^2}{\widehat{s}}})+{\displaystyle \frac{3}{2}}\left|C_{Q_1}\right|^2(1{\displaystyle \frac{4t^2}{\widehat{s}}})\widehat{s}+{\displaystyle \frac{3}{2}}\left|C_{Q_2}\right|^2\widehat{s}+6Re(C_9C_{Q_2}^{})t`$ $`E(\widehat{s})=Re(C_8^{eff}C_9^{}\widehat{s}+2C_7C_9^{}+C_8^{eff}C_{Q_1}^{}t+2C_7C_{Q_1}^{}t)`$ (6) Another observable related to CP violating effects in $`BX_sl^+l^{}`$ is the normal polarization of the lepton in the decay, $`P_N`$, which is the T-odd projection of the lepton spin onto the normal of the decay plane, i.e $`P_N\stackrel{}{s}_l(\stackrel{}{p}_s\times \stackrel{}{p}_l^{})`$ . A straightforward calculation leads to $`P_N={\displaystyle \frac{3\pi }{4}}\sqrt{1{\displaystyle \frac{4t^2}{\widehat{s}}}}\widehat{s}^{\frac{1}{2}}Im\left[2C_8^{eff}C_9t+4C_9C_7^{}{\displaystyle \frac{t}{\widehat{s}}}+C_8^{eff}C_{Q_1}+2C_7^{}C_{Q_1}+C_9^{}C_{Q_2}\right]/D(\widehat{s})`$ (7) The Wilson coefficients $`C_i`$ and $`C_{Q_i}`$ in eqs.(6) and (7) have been given in ref.. Since only $`C_8^{eff}`$ contains the non-trivial strong phase, $`A_{CP}^1`$ is determined by Im$`C_7`$ and $`A_{CP}^2`$ by Im$`C_{Q_1}`$ and Im$`C_7`$. Although $`P_N`$ depends on all the relevant Wilson coefficients a large $`P_N`$ does require relatively large values of Im$`C_{Q_i}(i=1,2)`$ . With the main contributions coming from exchanging chargino-stop loop with neutral Higgs coupled to external b quark , imaginary parts of $`C_{Q_i}`$s come mainly from terms proportional to (unitarity condition for stop mixing matrix has been used) $`{\displaystyle \frac{m_{\chi _i}m_t}{m_W^2sin\beta cos\beta }}U(i,2)V(i,2)D_{t21}D_{t11}^{},i=1,2`$ (8) i.e CP violating phases enter into the imaginary parts of $`C_{Q_i}`$ through g-h mixings (U, V) and chiral mixing ($`D_t`$) of stops. From the chargino mass matrix $`M_C=\left(\begin{array}{cc}M_2& \sqrt{2}m_Wsin\beta \\ \sqrt{2}m_Wcos\beta & \mu \end{array}\right)`$ (9) and stop mass matrix $`M_{\stackrel{~}{t}}^2=\left(\begin{array}{cc}M_{\stackrel{~}{Q}}^2+m{}_{t}{}^{}{}_{}{}^{2}+M_z^2(\frac{1}{2}Q_u\mathrm{sin}^2\theta _W)\mathrm{cos}2\beta & m_t(A_t^{}\mu \mathrm{cot}\beta )\\ m_t(A_t\mu ^{}\mathrm{cot}\beta )& M_{\stackrel{~}{U}}^2+m{}_{t}{}^{}{}_{}{}^{2}+M_z^2Q_u\mathrm{sin}^2\theta _W\mathrm{cos}2\beta \end{array}\right),`$ (10) we know that $`_{i=1}^2m_{\chi _i}U(i,2)V(i,2)=\mu `$ and $`D_{t21}D_{t11}^{}=\frac{m_t}{m_{\stackrel{~}{t}_1}^2m_{\stackrel{~}{t}_2}^2}(A_t\mu ^{}cot\beta )`$. Therefore, $`A_t`$ itself is as important as $`\mu `$ for providing imaginary contributions to $`C_{Q_i}`$, in particular, for large tan$`\beta `$. The similar conclusion holds also for $`C_7`$. We have known well that $`A_t`$ at the EW scale mainly depends on $`M_3`$ at the GUT scale through RGE effects. In fact there exists a quasi fixed point which shows the ratio of $`A_t`$ at $`m_Z`$ scale to $`M_3`$ at GUT scale to be about $`1.6`$ provided the Yukawa couplings of the third generation are large enough . Hence it is possible for $`A_t`$ to achieve a very large imaginary parts in the non-universal gaugino mass models, in contrast to the case of mSUGRA. Especially in large $`tan\beta `$ case, $`|\varphi _\mu |`$ is limited by EDM data to be less than $`\pi /6`$, so $`A_t`$ plays a more important role in CP Violation than $`\mu `$. To study the effect of large $`\varphi _3`$ (hence large $`\varphi _{A_t}`$) on $`C_7`$, we notice that essentially $`A_t`$ is multiplied by $`\mu `$. Changing the sign of $`A_t`$ has the same effects of switching the sign of $`\mu `$ and switching the sign of $`\mu `$ results in a sign change in $`C_7`$ (because in most of the parameter space $`\mu `$ are much larger than the non-diagonal terms in eq.(9)), so if $`\varphi _3`$ is in \[$`\pi /2`$, $`3\pi /2`$\] (hence $`\varphi _{A_0}`$ in \[$`\pi /2,\pi /2`$\]) and $`\varphi _\mu `$ in \[$`\pi /2,\pi /2`$\]), supersymmetry contributions give enhancement to $`ReC_7`$ so that it is hard to satisfy the $`BX_s\gamma `$ constraints: $`2.\times 10^4<Br(BX_s\gamma )<4.5\times 10^4`$ . Similar situation occurs for $`\varphi _3`$ being in \[$`\pi /2,\pi /2`$\]) and $`\varphi _\mu `$ in \[$`\pi /2`$,$`3\pi /2`$\]. Since supersymmetry gives large contributions to $`C_7`$ only when $`tan\beta `$ is large we shall focus on large $`tan\beta `$ case in the following. From the above analysis of EDMN we know that for somewhat large $`\varphi _\mu `$(say around $`0.4`$) cancellations happen for $`\varphi _3`$ near $`\pm \pi /2`$. With this kind of phases of $`M_3`$ and $`\mu `$ SUSY contributions to $`C_7`$ are almost totally imaginary. So the value of Im$`C_7`$ is constrained to be very small by the branching ratio of $`BX_s\gamma `$. One way to avoid the $`BX_s\gamma `$ constraint is to make use of the cancellation happened at $`\varphi _3`$ around $`\pm \pi /4`$ and $`\varphi _\mu `$ about $`0.2`$, With such kind of phases and a low mass spectrum ($`|M_2|`$ and $`|M_3|`$ around 150 GeV) the real part of SUSY contributions to $`C_7`$ cancels those from W-top and charged Higgs-top loops. The real part can even be cancelled to be near zero and only a large imaginary part of $`C_7`$ remains. Another way is to just suppress the total SUSY contributions to $`C_7`$, i.e., to make the mass spectrum heavier but still less than 1 Tev ($`|M_2|`$ and $`|M_3|`$ larger than about 300 GeV and less than about 500 GeV). In the former case $`A_{CP}^1`$ can reach order $`1\%`$ for $`BX_se^+e^{}`$. For $`BX_s\mu ^+\mu ^{}`$ and $`BX_s\tau ^+\tau ^{}`$, because of their larger Yukawa coupling there are great enhancement of branching ratio so that $`A_{CP}^1`$ are smaller than that for $`BX_se^+e^{}`$. In the later case(we shall call it as region A hereafter) it can only be a few thousandth at most, i.e., the same order as that in SM. As pointed out above, the effects of large $`\varphi _3`$ on $`C_{Q_i}`$ are similar to those on $`C_7`$. In large tan$`\beta `$ case, for $`\varphi _\mu \pm 0.2`$ and $`\varphi _3\pi /4`$ or $`\varphi _\mu \pi \pm 0.2`$ and $`\varphi _3\pi \pi /4`$ and with $`|M_2|`$ and $`|M_3|`$ around 150 GeV(which we shall call as region B for simplicity), which are allowed by the EDME and EDMN limits, Im$`C_{Q_i}`$ reachs maxima. In small tan$`\beta `$ case, although the constraints of EDMs on $`\varphi _\mu `$ and $`\varphi _3`$ are relaxed the magnitude of Im$`C_{Q_i}`$ is very small since $`C_{Q_i}`$ is proportional to $`m_ltan^2\beta `$(even $`m_ltan^3\beta `$ in some regions of the parameter space). Therefore, we expect the significant CP violation in large tan$`\beta `$ case. From eqs.(5) and (6), $`A_{CP}^2`$ can be rewritten as $`A_{CP}^2={\displaystyle \frac{E(\widehat{s})\overline{D}(\widehat{s})\overline{E}(\widehat{s})D(\widehat{s})}{E(\widehat{s})\overline{D}(\widehat{s})+\overline{E}(\widehat{s})D(\widehat{s})}}`$ For l=e, the difference between $`E(\widehat{s})`$ and $`\overline{E}(\widehat{s})`$ can be neglected(as it is proportional to lepton mass square, see eq.(6)). So $`A_{CP}^2`$ for l=e can be reduced to $`A_{CP}^2{\displaystyle \frac{\overline{D}(\widehat{s})D(\widehat{s})}{\overline{D}(\widehat{s})+D(\widehat{s})}}`$ that is exactly the opposite of $`A_{CP}^1`$. The same conclusion can be drawn for l=$`\mu ,\tau `$ in small tan$`\beta `$ case due to smallness of $`C_{Q_1}`$. On the other hand, for l=$`\tau `$ in large tan$`\beta `$ case, $`|E(\widehat{s})\overline{E}(\widehat{s})|`$ can be more important than $`|D\overline{D}|`$ and consequently one has approximately $`A_{CP}^2{\displaystyle \frac{E(\widehat{s})\overline{E}(\widehat{s})}{E(\widehat{s})+\overline{E}(\widehat{s})}}`$ Thus, it is propotional to t Im$`C_{Q1}`$. Therefore, in region B where $`C_{Q_i}`$s reach the maxima, $`A_{CP}^2`$ can be over $`50\%`$. In region A, $`C_{Q_i}`$s are less important and $`A_{CP}^2`$ can reach about $`5\%`$ at most. The correlation between $`A_{CP}^2`$ and EDME (or EDMN) in region A is plotted in fig.3 (note that we choose $`\widehat{s}=0.76`$ as representative in the figure). For l=$`\mu `$, the magnitude of $`A_{CP}^2`$ can be estimated to be of order $`1\%`$ at most in the large tan$`\beta `$ case. Numerical calculations prove this estimate. Fig.4 shows the correlation of EDM constraints and $`P_N`$ of $`BX_s\tau ^+\tau ^{}`$in region A. We can see in this figure that $`P_N`$ can reach more than 15 percent. In region B, as $`C_{Q_i}`$s are much larger the numerator in eq.(7) is increased a lot. But the denominator in eq.(7) (and consenquently the branching ratio of $`BX_s\tau ^+\tau ^{}`$) is also greatly enhanced in this region, so $`P_N`$ is just about $`15\%`$, i.e., not larger than the magnitude that can be achieved in region A. Situations for muon are similar and because of its much smaller Yukawa coupling the magnitude of $`P_N`$ can only reach about $`6\%`$. For electron $`P_N`$ is negligibly small, due to its neglegibly small mass. An important feature that can be seen from fig.3 and fig.4 is that the magnitudes of $`A_{CP}^2`$ and $`P_N`$ will not be reduced if EDM constraints improved. That is because the regions of parameter space where EDM constraints are satisfied are of width about $`\pi /20`$ for $`\varphi _3`$ and about $`\pi /4`$ for $`\varphi _{A_0}`$(adjustment needed), while $`C_{Q_i}`$s do not change sharply within these regions. In summary, we have analyzed the constraints of electric dipole moments of electron and neutron on the parameter space in supergravity models with nonuniversal gaugino masses. It is shown that with a light sparticle spectrum, the sufficient cancellations in the calculation of EDMs can happen due to the presence of the two new phases arising from complex gaugino masses, in addition to the phases $`\varphi _\mu `$ and $`\varphi _{A_0}`$. With appropriate correlation between $`\varphi _\mu `$ and $`\varphi _1`$ (for EDME) or $`\varphi _3`$ (for EDMN) as well as an appropriate choice of $`\varphi _{A_0}`$, cancellations can occur and all phases can be order of one in the small tan$`\beta `$ case and all phases but $`\varphi _\mu `$ ($`|\varphi _\mu |\stackrel{<}{}\pi /6`$) order of one in the large tan$`\beta `$ case. This is in contrast to the case of mSUGRA where in the parameter space where cancellations among various SUSY contributions to EDMs happen $`\varphi _\mu `$ must be less than $`\pi /10`$ for small $`tan\beta `$ and $`𝒪(10^2)`$ for large $`tan\beta `$. And our analysis show that the branching ratio of $`BX_s\gamma `$ gives an extra constraint on the phases for large tan$`\beta `$ case with light mass spectrum. We have calculated direct CP asymmetries and the T-odd normal polarization of lepton in $`BX_sl^+l^{}`$ in the regions of the parameter space in the models where the constraints from EDMs as well as $`BX_s\gamma `$ are satisfied. It is shown that the results for $`A_{CP}^1`$ are similar to those in mSUGRA if the mass sprectrum is relatively heavier($`|M_2|\&|M_3|\stackrel{>}{}300GeV`$) and it is also true for $`A_{CP}^2`$ in the small tan$`\beta `$ case. The former is due to the constraint from $`BX_s\gamma `$ and the latter is due to smallness of the contributions from exchanging neutral Higgs bosons in the small tan$`\beta `$ case. However, in the large tan$`\beta `$ case, $`A_{CP}^2`$ can reach $`1\%`$ for $`l=\mu `$ and is a few percent in most of allowed regions and can reach $`50\%`$ in some allowed regions for l=$`\tau `$. $`A_{CP}^2`$ for $`l=e`$ is approximately equal to $`A_{CP}^1`$ even in the large tan$`\beta `$ case. $`P_N`$ can reach $`6\%`$ for $`l=\mu `$ and is in the range from $`1\%`$ to $`15\%`$ in most of the allowed regions for l=$`\tau `$ in the large tan$`\beta `$ case. That is, there is a significant enhancement compared to the mSUGRA in which $`P_N`$ only can reach about $`0.5\%`$ for l=$`\mu `$ and about $`5\%`$ for l=$`\tau `$. In the small tan$`\beta `$ case the results are similar to those in mSUGRA. For l=e, $`P_N`$ is negligibly small, as it should be. This research was supported in part by the National Natural Science Foundation of China. Thanks to T. Nath and A. Pilaftsi for their helpful communications.
warning/0001/hep-ph0001182.html
ar5iv
text
# TAUP-2610-2000 hep-ph/0001182 Production of fermions in models of string cosmology ## 1 Introduction It is well known that particles are produced in cosmological backgrounds by amplification of vacuum fluctuations during inflationary epochs . In models of string cosmology , most of the produced particle spectra were found to be very different from the spectra in other cosmological models. First, in string cosmology models, gravitational waves may be generated in a substantial amount, in contrast to slow-roll inflation which predicts a “desert”. Furthermore, axions and other scalar particles have phenomenologically interesting spectra \[6-10\]. Finally, it is important to note that in addition to scalar particles and gravitational waves, photons can be produced . In general relativity, gauge field perturbations are not amplified in conformally flat cosmological backgrounds (even if inflationary) because of the scale invariant coupling of gauge fields in four dimensions. However, in string cosmology, the presence of a time-dependent dilaton prefactor in front of the gauge-field kinetic term breaks scale invariance and results in photon production which may even lead to the creation of macroscopic magnetic fields . We study fermion production by the standard amplification of quantum fluctuation in models of string cosmology. Although expectations were that fermions are copiously produced, we have found that production of spin 1/2 and spin 3/2 fermions in models of string cosmology during the dilaton-driven inflationary (DDI) phase is highly suppressed. It is known that massless spin 1/2 fermions are not amplified in conformally flat cosmological backgrounds . This is a consequence of their conformally invariant interactions. We show that in models of string cosmology the presence of a time-dependent dilaton prefactor in front of the fermion kinetic term does not result in amplification of fermionic perturbations. The spin-3/2 gravitino, the supersymmetric partner of the graviton, can be produced in conformally flat cosmological backgrounds, since gravitinos do not have conformally invariant couplings. It was recently shown that if gravitinos have vanishing zero-temperature, flat background mass, then they are not produced \[14-18\]. We show that, in models of string cosmology, the appearance of the dilaton prefactor in front of the gravitino kinetic terms does not result in amplification of massless gravitino perturbations, and therefore that gravitino production during DDI is negligible. We consider fermion production in models of string cosmology which realize the pre-big-bang scenario . In this scenario the evolution of the universe starts from a state of very small curvature and coupling and then undergoes a long phase of DDI and at some later time smoothly joins standard radiation dominated (RD) cosmological evolution, thus giving rise to a singularity free inflationary cosmology. Particles are produced during the period of DDI by the standard mechanism of amplification of quantum fluctuations. We assume throughout an isotropic and homogeneous four dimensional flat universe, described by a FRW metric. As in , the mass of the fermions is assumed to vanish in the early universe, i.e. in the dilaton driven period, and to take a nonzero value from the start of the RD era and on. All fields, except the dilaton and metric, are assumed to have a trivial vacuum expectation value during the inflationary phase. Fermion production, and in particular gravitino production, has the potential of placing severe constraints on models of string cosmology. If it turned out that gravitinos are maximally produced during DDI as gravitons are (a reasonable assumption) then after supersymmetry is broken, gravitinos can become by far the dominant energy component in the universe. The universe would become matter dominated very early, and the resulting cosmology will be very different than the standard RD cosmology. The issue of gravitino production during the evolution following the DDI phase is still unresolved, and we do not discuss thermal production which may turn out to put tight constraints on models of string cosmology, as it does for other models of inflation. ## 2 Spin 1/2 fermions We show here that massless spin 1/2 particles are not produced during DDI in models of string cosmology. We first explicitly show using their equations of motion that they are not produced, and then show that they are not produced even for a space and time dependent dilaton by showing that the low energy string effective action of massless spin 1/2 fermions can be transformed into a conformally invariant action. In the low energy effective action, a dilaton prefactor appears in front of the kinetic term, $$𝒜=d^4x\sqrt{g}e^{l\varphi }\left\{\frac{i}{2}\left[\overline{\psi }\widehat{\gamma }^\mu D_\mu \psi \left(D_\mu \overline{\psi }\right)\widehat{\gamma }^\mu \psi \right]M\overline{\psi }\psi \right\}.$$ (1) Greek letters denote space-time indices and Latin letters denote tangent-space indices. Gamma matrices and the group generators for spin 1/2 irreducible representations with curved indices are defined by $$\widehat{\gamma }^\mu =e_a^\mu (x)\gamma ^a\text{}\widehat{\mathrm{\Sigma }}^{\lambda \nu }=\mathrm{\Sigma }^{ab}e_a^\lambda e_b^\nu $$ (2) where $`e_\alpha ^\mu `$ is the vierbein. $`D_\mu `$ is the covariant derivative with respect to the spinorial structure $$D_\mu =_\mu +\mathrm{\Gamma }_\mu =_\mu +\frac{1}{2}g_{\lambda \delta }\mathrm{\Gamma }_{\nu \mu }^\delta \widehat{\mathrm{\Sigma }}^{\lambda \nu }.$$ (3) Fermions have $`l=1`$, but we keep parameter $`l`$ general, since our results do not depend on its specific value. Variation of the action with respect to $`\overline{\psi }`$ yields the Dirac equation, $$i\widehat{\gamma }^\mu D_\mu \psi M\psi +\frac{i}{2}l\widehat{\gamma }^\mu _\mu \varphi \psi =0.$$ (4) In the simplified model of background evolution we adopt, the evolution of the universe is described by (conformal) time dependent dilaton $`\varphi (\eta )`$ and a spatially flat FRW metric, in which the line element can be written as $`d^2s=a^2(\eta )\left(d^2\eta d^2\stackrel{}{x}\right).`$ Here $`\stackrel{}{x}=(x^1,x^2,x^3)`$ are comoving space coordinates and $`x^0\eta `$ is the conformal time. Thus, the vierbein can be written as $`e_\mu ^a=a\delta _\mu ^a,`$ $`e_a^\mu =a^1\delta _a^\mu `$. In this case, the equation of motion (4) reduces to $$\gamma ^a_a\psi +\frac{1}{2}\frac{a^{}}{a}\stackrel{}{\gamma }\stackrel{}{\mathrm{\Sigma }}\psi +iMa\psi ++\frac{l}{2}_0\varphi \gamma ^0\psi =0,$$ (5) where $`\stackrel{}{\mathrm{\Sigma }}=(\mathrm{\Sigma }^{01},\mathrm{\Sigma }^{02},\mathrm{\Sigma }^{03}).`$ Choosing a gamma-matrix representation such that $$\gamma ^0=\left(\begin{array}{cc}I& 0\\ 0& I\end{array}\right);\stackrel{}{\gamma }=\left(\begin{array}{cc}0& \stackrel{}{\sigma }\\ \stackrel{}{\sigma }& 0\end{array}\right),$$ (6) the field $`\psi `$ satisfies the following equations of motion $`\left(_0+iMa+{\displaystyle \frac{l}{2}}_0\varphi +{\displaystyle \frac{3}{2}}{\displaystyle \frac{a^{}}{a}}\right)\psi _1+\sigma ^i_i\psi _2`$ $`=`$ $`0`$ (7) $`\left(_0iMa+{\displaystyle \frac{l}{2}}_0\varphi +{\displaystyle \frac{3}{2}}{\displaystyle \frac{a^{}}{a}}\right)\psi _2+\sigma ^i_i\psi _1`$ $`=`$ $`0.`$ Introducing the field $`\chi `$, $$\psi (\eta ,\stackrel{}{x})=a^{3/2}e^{\frac{l}{2}\varphi }\chi (\eta ,\stackrel{}{x}),$$ (8) equations (7) reduce to $`\left(_0+iMa\right)\chi _1+\sigma ^i_i\chi _2`$ $`=`$ $`0`$ (9) $`\left(_0iMa\right)\chi _2+\sigma ^i_i\chi _1`$ $`=`$ $`0.`$ (10) Substituting (9) into (10) and vice-versa, and writing $`\chi `$ as $`\chi (\eta ,\stackrel{}{x})=\chi (\eta )e^{ikx}`$, we arrive at $$\left(_0^2+k^2+M^2a^2\pm iMa\right)\left(\begin{array}{c}\chi _1\\ \chi _2\end{array}\right)=0.$$ (11) Finally, the equation of massless spin 1/2 fermions is reduced to the equation in flat space-time and therefore these particles are not produced. Let us turn now to the massless low energy effective action of spin 1/2 particles and show that it is conformally invariant, even for a space and time dependent dilaton and scale factor. Under the conformal transformation $$g^{\mu \nu }\stackrel{~}{g}^{\mu \nu }=\mathrm{\Omega }^2g^{\mu \nu },$$ (12) vierbeins, gamma matrices and covariant derivatives transforms as follows, $$e_a^\mu \stackrel{~}{e}_a^\mu =\mathrm{\Omega }e_a^\mu $$ $$\widehat{\gamma }^\mu \stackrel{~}{\widehat{\gamma }}{}_{}{}^{\mu }=\mathrm{\Omega }\widehat{\gamma }^\mu $$ $$D_\mu \stackrel{~}{D}_\mu =D_\mu +\frac{1}{2}\mathrm{\Sigma }^{ab}e_a^\nu g_{\rho \nu }e_b^\rho \frac{}{x^\mu }\left(\mathrm{ln}\mathrm{\Omega }\right),$$ and $`\psi `$ transform as follows $$e^{l\varphi (x)}e^{l\stackrel{~}{\varphi }(x)}=\phi \left(x\right)e^{l\varphi (x)}$$ $$\psi \stackrel{~}{\psi }=\mathrm{\Omega }^{3/2}(x)\phi ^{1/2}\left(x\right)\psi .$$ The low energy effective action of massless spin 1/2 particles (1) transforms into $$𝒜\stackrel{~}{𝒜}=d^4x\sqrt{\stackrel{~}{g}}e^{l\stackrel{~}{\varphi }}\frac{1}{2}i\left[\overline{\stackrel{~}{\psi }}\stackrel{~}{\widehat{\gamma }}{}_{}{}^{\mu }\stackrel{~}{D}_{\mu }^{}\stackrel{~}{\psi }\left(\stackrel{~}{D}_\mu \overline{\stackrel{~}{\psi }}\right)\stackrel{~}{\widehat{\gamma }}{}_{}{}^{\mu }\stackrel{~}{\psi }\right]=$$ (13) $$d^4x\sqrt{g}e^{l\varphi }\left[\frac{1}{2}i\left(\overline{\psi }\widehat{\gamma }^\mu D_\mu \psi (D_\mu \overline{\psi })\widehat{\gamma }^\mu \psi \right)\right]+$$ $$d^4x\sqrt{g}e^{l\varphi }\left[\frac{1}{4}i\overline{\psi }\left(\widehat{\gamma }^\mu \mathrm{\Sigma }^{ab}\mathrm{\Sigma }^{ab}\widehat{\gamma }^\mu \right)\psi e_a^\nu g_{\rho \nu }e_b^\rho \frac{}{x^\mu }\left(\mathrm{ln}\mathrm{\Omega }\right)\right].$$ Since $`\widehat{\gamma }^\mu \mathrm{\Sigma }^{ab}\mathrm{\Sigma }^{ab}\widehat{\gamma }^\mu `$ is antisymmetric in $`(a,b)`$ and $`e_a^\nu g_{\rho \nu }e_b^\rho `$ is symmetric in $`(a,b)`$, it follows that $$\left(\widehat{\gamma }^\mu \mathrm{\Sigma }^{ab}\mathrm{\Sigma }^{ab}\widehat{\gamma }^\mu \right)e_a^\nu g_{\rho \nu }e_b^\rho =0.$$ Thus, the low energy effective action (1) is conformally invariant in the massless limit. In other words, fluctuations of massless spin 1/2 fermions are not amplified in a conformally flat background, even if inflationary, in agreement with our previous calculation. Action (1) is valid during the DDI period, and since the fermions are assumed to be massless during that period they are not produced. Following that period, during the string phase, fermions are assumed massless, and it seems reasonable to assume that the fermions are not produced during the string phase as well (but see below), if so they may be produced only during the RD and MD epochs, by thermal and non-thermal production. Since in the RD and MD era the dilaton is assumed to be fixed and constant, we conclude that the spectra of the produced spin 1/2 particles in string cosmology models are similar to those in other models, and therefore the cosmological constraints and observable features are common. ## 3 Gravitinos - spin 3/2 fermions In this section we show that the appearance of a time-dependent dilaton prefactor in the low energy effective action does not change the amplification of fluctuation of massless gravitinos. Before studying the properties of the action, it is useful to recall a few known facts about gravitinos in conformally flat cosmological backgrounds. Massive gravitinos in conformally flat space time are described by four Majorana spinors $`\psi ^\mu `$, satisfying the equations $`\left[i\gamma ^0_0+i{\displaystyle \frac{5\dot{a}}{2a}}\gamma ^0ma+k\gamma ^3\right]\psi _{3/2}=0,`$ $`\left[i\gamma ^0_0+i{\displaystyle \frac{5\dot{a}}{2a}}\gamma ^0ma+k\left(A+iB\gamma ^0\right)\gamma ^3\right]\psi _{1/2}=0,`$ (14) $`A`$ $`=`$ $`{\displaystyle \frac{1}{3\left(\frac{\dot{a}^2}{a^4}+m^2\right)^2}}\left[2{\displaystyle \frac{\ddot{a}}{a^3}}\left(m^2{\displaystyle \frac{\dot{a}^2}{a^4}}\right)+{\displaystyle \frac{\dot{a}^4}{a^8}}4m^2{\displaystyle \frac{\dot{a}^2}{a^4}}+3m^44{\displaystyle \frac{\dot{a}}{a^3}}\dot{m}m\right]`$ $`B={\displaystyle \frac{2m}{3\left(\frac{\dot{a}^2}{a^4}+m^2\right)^2}}\left[2{\displaystyle \frac{\ddot{a}\dot{a}}{a^5}}{\displaystyle \frac{\dot{a}^3}{a^6}}+3m^2{\displaystyle \frac{\dot{a}}{a^2}}+{\displaystyle \frac{\dot{m}}{ma}}\left(m^2{\displaystyle \frac{\dot{a}^2}{a^4}}\right)\right].`$ Here $`\psi _{1/2}`$ and $`\psi _{3/2}`$ are defined such that $`\psi ^0`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{3}}}c\widehat{\gamma }_3d\widehat{\gamma }_1\psi _{1/2}`$ (15) $`\psi ^1`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{6}}}\psi _{1/2}+{\displaystyle \frac{1}{\sqrt{2}}}\psi _{3/2}`$ (16) $`\psi ^2`$ $`=`$ $`\widehat{\gamma }^2\widehat{\gamma }_1\left({\displaystyle \frac{1}{\sqrt{6}}}\psi _{1/2}{\displaystyle \frac{1}{\sqrt{2}}}\psi _{3/2}\right)`$ (17) $`\psi ^3`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{3}}}\left(d\widehat{\gamma }^3\right)\widehat{\gamma }_1\psi _{1/2}`$ (18) and $`c`$ $`=`$ $`{\displaystyle \frac{1}{3a\left(\frac{\dot{a}^2}{a^4}+m^2\right)}}\left[\left(2{\displaystyle \frac{\ddot{a}}{a^3}}+{\displaystyle \frac{\dot{a}^2}{a^4}}3m^2\right)\gamma ^0+2i{\displaystyle \frac{\dot{m}}{a}}\right]`$ (19) $`d`$ $`=`$ $`{\displaystyle \frac{|\stackrel{}{k}|}{a^2\left(\frac{\dot{a}^2}{a^4}+m^2\right)}}\left(i{\displaystyle \frac{\dot{a}}{a^2}}\gamma ^0+m\right),`$ (20) and $`\stackrel{}{k}=(0,0,k)`$. The fields $`\psi _{1/2}`$ and $`\psi _{3/2}`$ correspond to the $`\pm 1/2`$ and $`\pm 3/2`$ helicity states in the flat limit. In the massless limit, the field equation of motion in (14) becomes $`\left[i\gamma ^0_0+i{\displaystyle \frac{5\dot{a}}{2a}}\gamma ^0+k\gamma ^3\right]\psi _{3/2}=0`$ $`\left[i\gamma ^0_0+i{\displaystyle \frac{5\dot{a}}{2a}}\gamma ^0+k\left(A+iB\gamma ^0\right)\gamma ^3\right]\psi _{1/2}=0,`$ (21) $`A={\displaystyle \frac{1}{3}}\left[2{\displaystyle \frac{\ddot{a}}{\dot{a}}}{\displaystyle \frac{a}{\dot{a}}}+1\right]`$ $`B=0.`$ The pre-big bang model suggests that the evolution of the scale factor $`a(\eta )`$ is a power dependent function, i.e., $$a=a_s\left(\frac{\eta }{\eta _s}\right)^\alpha .$$ (22) In this case, $`A`$ becomes a constant; i.e., $`A=\left(2\alpha \right)/3\alpha `$, and eqs. (21) become: $`\left[i\gamma ^0_0+i{\displaystyle \frac{5\dot{a}}{2a}}\gamma ^0+k\gamma ^3\right]\psi _{3/2}=0`$ (23) $`\left[i\gamma ^0_0+i{\displaystyle \frac{5\dot{a}}{2a}}\gamma ^0+\overline{k}\gamma ^3\right]\psi _{1/2}=0,`$ where $`\overline{k}=\left(\frac{2\alpha }{3\alpha }\right)k.`$ As for spin 1/2 particles, one can reduce eqs. (23) to the field eqs. of motion in flat space-time with a suitable transformation. Thus, massless gravitino fluctuation are not amplified in a “power low” expanding universe and no production of gravitinos occurs during the DDI phase. Next, we show that the low limit effective action of a massless gravitino can be transformed to the action of massless gravitino in standard supergravity, even for a space and time dependent dilaton background. The low energy effective action of the gravitino in string theory is the following, $$𝒜=d^4xe^{l\varphi }ϵ^{\mu \nu \rho \sigma }\overline{\psi }_\mu \gamma _5\widehat{\gamma }_\nu D_\rho \psi _\sigma +h.c.,$$ (24) where $$D_\mu =_\mu +\frac{1}{2}\omega _{\mu ab}\sigma ^{ab},$$ (25) $`\omega _{\mu ab}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(C_{\mu ab}+C_{ab\mu }+C_{b\mu a}\right),`$ (26) $`C_{}^{a}{}_{\mu \nu }{}^{}`$ $`=`$ $`_\mu e_\nu ^a_\nu e_\mu ^a{\displaystyle \frac{1}{2M_P^2}}\overline{\psi }_\mu \gamma ^a\psi _\nu _\mu e_\nu ^a_\nu e_\mu ^a.`$ (27) Using the transformation $`\psi _\sigma =e^{l\varphi /2}\stackrel{~}{\psi }_\sigma ,`$ action (24) becomes $$d^4x\left[\frac{1}{2}lϵ^{\mu \nu \rho \sigma }\stackrel{~}{\psi }_\mu \gamma _5\widehat{\gamma }_\nu \stackrel{~}{\psi }_\sigma _\rho \varphi +ϵ^{\mu \nu \rho \sigma }\stackrel{~}{\psi }_\mu \gamma _5\widehat{\gamma }_\nu D_\rho \stackrel{~}{\psi }_\sigma \right]+h.c..$$ (28) The first term in (28) is canceled by its hermitian conjugate since using $`\gamma ^0\left[\gamma _5\widehat{\gamma }_\nu \right]^+\gamma ^0=\gamma _5\widehat{\gamma }_\nu `$, $$\left[ϵ^{\mu \nu \rho \sigma }\overline{\psi }_\mu \gamma _5\widehat{\gamma }_\nu \psi _\sigma _\rho \varphi \right]^{}=ϵ^{\mu \nu \rho \sigma }\overline{\psi }_\mu \gamma _5\widehat{\gamma }_\nu \psi _\sigma _\rho \varphi .$$ (29) Thus, $$d^4xe^{l\varphi }ϵ^{\mu \nu \rho \sigma }\overline{\psi }_\mu \gamma _5\widehat{\gamma }_\nu D_\rho \psi _\sigma +h.c.d^4xϵ^{\mu \nu \rho \sigma }\overline{\psi }_\mu \gamma _5\widehat{\gamma }_\nu D_\rho \psi _\sigma +h.c.$$ (30) when $`\psi _\sigma e^{l\varphi /2}\psi _\sigma `$, showing that gravitinos are not produced during the DDI phase, since, as we recall from eq.(23), there is no production of massless gravitino, and therefore, they will not be produced in string cosmology, as well. The mass of the gravitino is assumed to vanish in the early universe, i.e. in the DDI period, and then to take a nonzero value after the start of the RD era when supersymmetry is broken. Thus, using the same argument that we have used for spin 1/2 particles, we conclude that gravitinos can get produced only during the subsequent RD and MD era. ## 4 Summary and Conclusions We have studied fermion production during DDI in string cosmology models, and found that massless spin 1/2 fermions and gravitinos are not produced. This unexpected result can be contrasted with boson production. In particular, naive expectations would be that gravitinos are copiously produced. In general, bosons have different spectra in string cosmology and in standard slow-roll models, but this is not the case for fermions. Spin 1/2 particles preserves their conformal invariance, and are not produced during the DDI. They can be produced from the start of RD era and on, by thermal and non-thermal production, as in other cosmological models. Similarly, although the action of spin 3/2 particles is not conformally invariant, it is invariant, in the massless limit, under ”rescaling”. Thus, assuming massless gravitinos in the early universe, spin 3/2 particles have the same spectrum in string and standard cosmology. Our results have a limitation; we have neglected production of fermions during the string phase. As mentioned before, the string phase is a high-curvature phase of otherwise unknown properties. Usually, one could overcome lack of knowledge of string phase evolution by using the fact that fluctuations of fields are frozen when their wavelength is much larger than the curvature scale during the string phase. However, fermions seem to never freeze. Thus the unknown properties of the string phase may be important for fermions production. This is a puzzling issue which we do not understand. Causality requires that amplitudes of waves whose wavelength is larger than the cosmological horizon be constant, but fermions seem to evade this requirement, since they do not seem to feel the cosmological horizon. Finally, we note that our results may hold for more general string models. In this paper, as well as previous ones, we have assumed that the dilaton depends only on time. However the conformal invariance of the massless spin 1/2 action in the low energy effective string action is preserved also for a space and time dependent dilaton. As a matter of fact it is easy to eliminate a space and time dependent dilaton from the actions of spin 1/2 and spin 3/2 fermions. Thus our results are valid even in string cosmology models that involve a space and time dependent dilaton background. ## Acknoledgments We thank L. Kofman, M. Lemoine and A. Riotto for discussions, and A. Linde for useful comments on the manuscript.
warning/0001/cond-mat0001202.html
ar5iv
text
# Chiral Twisting of a Smectic-𝐴 Liquid Crystal ## I Introduction The addition of chirality to the molecular structure of smectic liquid crystals leads to unique materials properties which have allowed for the development of fast, high-resolution display devices . Molecular chirality results in ferroelectricity in the tilted smectic-$`C^{}`$ liquid crystalline phase . In the non-tilted smectic-$`A`$ phase, rotational symmetry about the molecular long axis rules out any spontaneous polarization in the bulk. However, application of an electric field parallel to the smectic layers breaks this symmetry due to the coupling of the field to the transverse dipole moment and leads to an induced tilt relative to the layer normal, the electroclinic effect . A similar symmetry-breaking effect can also arise due to surface interactions. Unidirectional rubbing of a cell surface is commonly used to align liquid crystals so that their long axes are along the rubbing direction. In the smectic-$`A`$ phase, the layer normal is also typically parallel to the rubbing direction resulting in the planar or bookshelf geometry. However, Nakagawa et al. observed that for a smectic-$`A`$ composed of chiral molecules, the optic axis is rotated with respect to the rubbing direction . This result was attributed to a polar interaction between the liquid crystal molecules and the surface which gives rise to a surface electroclinic effect . Subsequent studies on another material found that the rotation angle of the optic axis can be as large as 18 . Detailed studies by Chen and coworkers, using optical second-harmonic generation, found that the molecules anchored at the cell surface are oriented along the rubbing direction, while the optic axis of the bulk is rotated with respect to the rubbing direction by an angle $`\mathrm{\Psi }`$ . Taking the optic axis to be parallel to the layer normal in the bulk, this suggests that direction of the layer normal is determined by surface interactions and remains constant as one moves from the surface to the bulk. Thus, unlike the field induced electroclinic effect that results in a uniform tilt of the director over the entire thickness of the cell, the surface electroclinic effect is a localized effect which is restricted to a boundary layer near the cell surface. This layer is much thinner than the wavelength of visible light, since a uniform optic axis and good extinction are observed when a thick cell is viewed between crossed polarizers . In the interfacial region, the molecular director twists by $`\mathrm{\Psi }`$ from the rubbing axis until it reaches the layer normal direction as one moves away from the surface to the interior of the cell as illustrated in Fig. 1. The thickness of the boundary layer $`\xi `$ is determined by the correlation length of tilt fluctuations in the smectic-$`A`$ phase. Using ellipsometry, Chen et al. found the expected divergence of $`\xi `$ with decreasing temperature near the smectic-$`A`$ to smectic-$`C^{}`$ transition . In this paper, we use circular dichroism (CD) spectroscopy to directly probe the chiral twist between the surface and bulk states. Circular dichroism is a sensitive measure of the twisted orientation of molecules with an appropriate chromophore in the chemical structure of the molecule. Past studies have found large CD peaks in the ferroelectric smectic-$`C^{}`$ phase and pretransitional effects in smectic-$`A`$ phase when approaching smectic-$`C^{}`$ phase . This chiral order arises from helical variation of the azimuthal tilt direction from layer to layer and is observed for light propagating perpendicular, but not parallel, to the smectic layers. The present work represents the first observation of CD due to the chiral order induced by the surface electroclinic effect and is only observed when the light is propagating parallel to the smectic layers. Our results show the existence of a strong twist over a wide temperature range in the smectic-$`A`$ phase. We have also studied the effect of an applied electric field that shows a complex behavior depending upon the temperature. ## II Experimental Procedure The liquid crystal used in our studies, KN125, was synthesized using established procedures . The chemical structure of this molecule is shown in Fig. 2. This material exhibits a large electroclinic effect over a wide temperature range . The phase sequence for KN125 upon heating is Crystal – 29 C – Smectic-$`A`$ – 80 C – Isotropic. The smectic-$`A`$ phase can be supercooled below 25 C for several days without crystallizing. CD studies were performed on a Jasco J-720 spectropolarimeter. All data are given in units of molar circular dichroism $`\mathrm{\Delta }\epsilon \epsilon _L\epsilon _R`$, where $`\epsilon _{L(R)}`$ is the molar decadic absorption coefficient for left- (right-) circularly polarized light (e.g. the ratio of transmitted to incident intensity is $`I_t/I_i=10^{\epsilon cd}`$, where $`c`$ is the molar concentration and $`d`$ is the pathlength in cm). Studies in the planar geometry were performed using commercial liquid crystal cells (E.H.C. Co., Tokyo) with a nominal cell gap of 2.5 $`\mu `$m. These cells had a rubbed polyimide surface layer and were mounted in an Instec MK1 hot stage. The cells were oriented so that the sample optic axis was rotated by 45 deg. from the optic axis of the polarization modulator in the spectrometer to eliminate artifacts from the inherent birefringence of the sample (see discussion below). A 100 Hz bipolar square wave of variable amplitude was applied to 1 cm<sup>2</sup> ITO electrodes. All fields are given in RMS voltage. For probing the CD along the layer normal direction, a cell having a 12 $`\mu `$m gap and treated with octadecyltrichlorosilane to promote homeotropic alignment (molecules perpendicular to the surface) was used. ## III Results The CD spectra of KN125 measured in two geometries are shown in Figure 2(a). These spectra were both taken at 25 C with no applied field. It was ascertained using optical microscopy that the sample had a uniform optic axis and was in the smectic-$`A`$ phase, with no signs of crystallization. The spectra in Fig. 2(a) reveal a large peak around 362 nm when the molecules are aligned in the planar geometry (solid line), while such a peak is not seen for homeotropic alignment (dashed line). The small wiggles at longer wavelengths are interference effects due to the finite thickness of the sample. Two precursor compounds were also studied to establish the origin of the peak in the CD spectra. The first precursor, denoted HNBB, consisted of the core of KN125 without the acyl chains, and the second, HBB, consisted of KN125 without either the chains or the nitro substitution. Absorption spectra of these compounds in solution, along with their chemical structures, are shown in Fig. 2(b). A broad peak centered near 360 nm appears in the HNBB spectra (solid line) and is absent in the HBB spectra (dashed line). This shows that the peak in the CD spectra at 362 nm reflects an electronic transition related to the laterally substituted nitro group. The wavelength indicates that the $`n_a\pi ^{}`$ transition is slightly red-shifted due to coupling of the nitro electronic orbitals to those in the biphenyl group . It should be pointed out that the CD peak shown in Fig. 2 is more than 2000 times larger than the intrinsic CD of KN125 in solution ($`\mathrm{\Delta }\epsilon _{355}=0.035`$ cm<sup>-1</sup> M<sup>-1</sup> in acetonitrile). Comparison of the CD and UV spectra gives a dissymmetry factor $`g\mathrm{\Delta }\epsilon /\epsilon =0.36`$. An important test of chiral phenomena is to study the CD spectra from opposite enantiomers, since any effect which arises purely from the handedness of the molecule, should change sign, but not magnitude, when the chirality of the molecule is reversed. In Figure 3 we show the CD spectra from both enantiomers of KN125 at 25 C in planar cells. We clearly see that reversing the molecular chirality leads to a concurrent change in the sign of the CD peak, but no change in its magnitude. Although this large CD signal implies a twist of the chiral molecules, we have to carefully eliminate all possible experimental artifacts which could be contributing to the CD signal. One obvious problem can be caused by a misalignment of the rubbing directions at the two surfaces leading to a small twist through the cell. We have performed experiments using numerous cells, both commercially obtained and fabricated in-house, taking care in each case to see that the rubbing directions at the two surfaces are accurately antiparallel (within 1). We find the peak magnitude varies by less than 10% between cells under similar conditions (temperature, field, alignment layer). More importantly, we always observe a negative CD peak from the (R)-enantiomer of KN125, while the (S)-enantiomer always gives a positive peak. Another important factor to consider is the effect of linear dichroism. A number of authors have discussed the difficulties in deconvoluting the linear and circular contributions to signals obtained using polarization-modulation spectroscopy . In particular, Shindo and coworkers have elegantly addressed this problem using the Stokes-Mueller formalism . They find that the CD signal measured in a polarization-modulation spectrometer is given by $`\mathrm{\Delta }\epsilon _{\mathrm{meas}}`$ $`=`$ $`\mathrm{\Delta }\epsilon _{\mathrm{true}}+{\displaystyle \frac{1}{2}}\left(\text{LD}^{}\text{LB}\text{LD}\text{LB}^{}\right)`$ (2) $`+\left(\text{LD}^{}\mathrm{sin}2\theta \text{LD}\mathrm{cos}2\theta \right)\mathrm{sin}\alpha `$ where LD (LB) is the linear dichroism (birefringence) with respect to the optic axis of the material, LD (LB) is the linear dichroism (birefringence) with respect to an axis rotated 45 degrees from the optic axis around the direction of light propagation, $`\theta `$ is the angle between the optic axis of the sample and that of the spectrometer, as defined by its optical modulator, and $`\alpha `$ is an angle representing any imperfections in the optical modulator due to misalignment or residual birefringence. Note that in these units $`\text{LD}=\epsilon _{}\epsilon _{}`$ and $`\text{LB}=4\pi (n_{}n_{})/(c\lambda _0\mathrm{ln}10)`$, where $`c`$ is the molar concentration, $`\lambda _0`$ is the vacuum wavelength of the incident light, and $`n_{()}`$ is the index of refraction parallel (perpendicular) to the optic axis. Similar relations hold for LD and LB. The first component of the measured signal is the true CD signal which is independent of $`\theta `$. The second term is a combination of the linear dichroism and linear birefringence as measured both along the optic axis of the sample and at 45 degrees to it. It is also rotationally independent and is only nonzero for a biaxial sample where LD, LB$`{}_{}{}^{}0`$. Since we find the linear dichroism at 45 degrees to be at least $`10^3`$ times smaller than that measured along the optic axis, (LD/LD $`10^3`$), our sample is approximately uniaxial and we ignore the second term. The final term in Eqn. (2) is a rotationally dependent term due to the interaction between a nonideal optical modulator and a sample with linear dichroism. Indeed, we find that the measured CD signal has a small rotationally dependent component which does not change sign when the chirality of the molecule is switched . The angular dependence of this component is consistent with a small misalignment of the photoelastic modulator ($`\alpha =1.1^{}`$). Shindo et al. also consider additional artifacts due to a partially polarizing detector and imperfect harmonic response of a lock-in amplifier . A quartz depolarizer inserted before the detector produces no change in the measured CD ruling out the former. The latter is very difficult to quantify. However, no evidence of such artifacts is seen when an achiral, linearly anisotropic sample is studied. When the sample is aligned so that the optic axis of the liquid crystal is 45 degrees with respect to the modulator, the contribution to the CD due to linear effects is expected to vanish in a uniaxial sample. All measurements discussed in this paper were made in such a geometry. Having thus eliminated experimental artifacts, we shall now discuss the effect of temperature and applied electric field on the CD spectra. Figure 4 shows the temperature variation of the CD peak in the absence of an applied electric field. CD spectra at five different temperatures are shown in Fig. 4(a) for (R)-KN125. As the temperature increases, the magnitude of the CD peak continually decreases, but its position remains unchanged. This confirms that the observed effect is due to preferential absorption of right-circularly polarized light, as opposed to a preferential scattering effect, where the wavelength of the CD peak is expected to shift with temperature. Figure 4(b) shows the temperature dependence of the CD peak magnitude, $`\mathrm{\Delta }\epsilon _{pk}`$, for both enantiomers of KN125 at 25 C with no applied field. We see that both enantiomers show a similar decrease in $`\mathrm{\Delta }\epsilon _{pk}`$ with increasing temperature. However, the CD peak does not continually vanish as the isotropic transition is approached. Rather, there remains a small remnant value over a wide temperature range in the smectic-$`A`$ phase right up to the melting point. When an AC electric field is applied across the sample at 25 C, we also find a change in the magnitude of the CD peak, but not in its position. As the field is increased, the CD peak magnitude continually decreases, as shown in Figure 5. Both enantiomers show a similar field dependence of the peak CD. However, the CD peak does not completely disappear at high fields. Instead, we find a small residual CD signal at high field ($`\mathrm{\Delta }\epsilon _{pk}`$ = -14.7 at 25 C, 5.2 V/$`\mu `$m) which is similar to that seen in Fig. 4(b) at higher temperatures and no field ($`\mathrm{\Delta }\epsilon _{pk}`$ = -14.9 at 75 C, 0 V/$`\mu `$m). This remnant CD indicates the existence of a small twist which cannot be unwound by electric field or thermal effects. As the temperature is increased in the smectic-$`A`$ phase, the situation becomes more complex. The field dependence of the peak magnitude is shown for both enantiomers at 40 and 60 C in Fig. 6(a). Near 40 C, the peak magnitude is relatively field independent (open symbols). At higher temperatures, $`\mathrm{\Delta }\epsilon _{pk}`$ begins to increase with field, as seen in the data at 60 C (filled symbols). The temperature dependence of the peak circular dichroism of (R)-KN125 is shown for three different fields in Fig. 6(b). We find significant differences in the temperature dependence of the CD peak at different fields. At zero field, the peak decreases monotonically as the temperature is increased as also shown in Fig. 4(b). At an intermediate field of 2.6 V/$`\mu `$m, the peak CD is relatively temperature independent, while at large field, 5.2 V/$`\mu `$m, the peak initially increases with temperature to a maximum value around 60 C and then begins to decrease. At all fields the peak goes to zero at the isotropic phase transition. Similar behavior is also seen in the (S)-enantiomer. This complex dependence of the CD on both temperature and field provides further evidence that the observed results are not artifacts due to linear effects, since measurements of the linear dichroism along the optic axis reveal that it is weakly dependent on both temperature and an applied AC field . ## IV Discussion A number of possible explanations may describe the origin of the large CD peak observed in the planar, but not homeotropic, geometry. The possibility of experimental artifacts was discussed in detail above. Another possible explanation is in-plane chiral fluctuations of the type proposed by Lubensky, Kamien, and Stark . While the smectic-$`A`$ phase has no macroscopic chiral order, it is possible for the molecular chirality to bias the thermal fluctuations of the molecules giving the system local, or short-range, chiral order. However, since a smectic-$`A`$ cannot support long-range twist deformations , these fluctuations can only extend over a few molecules and are unlikely to be responsible for the large enhancement we see in the CD spectra. Instead, we believe the large chiral effect is due to a twisting of the director induced by the surface electroclinic effect discussed earlier. Here, the polar interactions at the cell wall-liquid crystal interface act like an applied field, causing the molecules to be tilted with respect to the layer normal by an angle $`\mathrm{\Psi }`$. Since the molecules are strongly anchored by the rubbing, the layer normal is then tilted by $`\mathrm{\Psi }`$ from the rubbing direction as depicted in Fig. 1. The direction of the layer normal established by the surface propagates unchanged into the bulk. The layer normals induced by opposing cell interfaces are rotated in opposite directions since the interfaces are mirror images of each other. This implies that the molecules in the two interfacial regions twist in the same direction so that the net twist induced by the two cell walls is twice that from a single interface. However, this also implies that the layer normal rotates by $`2\mathrm{\Psi }`$ between the two surfaces. Since uniform twisting of the layer normal is energetically disfavored in the smectic-$`A`$ phase , this rotation presumably occurs over a very small distance in a twist grain boundary inside the cell . Patel, Lee, and Goodby observed such defects in thin cells, but not in thicker cells where only one domain was seen . In the thick cells, the observation of a single domain with optic axis not parallel to the rubbing direction was attributed to preferential growth of one of the two possible layer normal directions into the bulk . We observe good optical extinction when the cell is viewed between crossed polarizers with one polarizer aligned along the optic axis and find the optic axis of the sample is rotated by about 10 from the rubbing direction. Interestingly, this 10 offset is similar to the zero field mosaicity measured in x-ray diffraction experiments on the same material . We find that the offset angle changes by less than a degree over the entire smectic-$`A`$ phase. Further support for the surface electroclinic effect model is obtained by studying a cell with single-sided alignment. Recent work has shown that KN125 can be well aligned in a cell where only one side has rubbed polyimide on it, the other surface being plain conducting (ITO) glass . The thickness of the cell was approximately the same as the double-sided alignment layer cell described earlier. Fig. 3 shows the CD spectra from such a single-sided alignment layer cell containing (R)-KN125. We find that the CD peak is decreased compared to the double-sided cell, but still quite large. In fact the peak magnitude is slightly larger than half that from the double-sided cell, possibly indicating a reduction in twist at the grain boundary. This observation rules out chiral fluctuations in the bulk as a major contributor to the observed CD, since such fluctuations should only depend on the thickness of the cell and not on whether one or both sides have an alignment layer. We also find the temperature and field dependence of the CD peak in the singled-sided alignment layer cell follows that of the double-sided cell, as shown in Figs. 4(b) and 5, respectively (diamonds). In order to explain the complex field dependence of the CD spectra with increasing temperature, we consider two temperature regions. At low temperature ($``$ 25-35 C), the CD peak decreases rapidly with applied field, while at high temperature ($``$ 50-80 C), the CD increases with field. In the low temperature region, the material responds as though it were near the smectic-$`A`$ to smectic-$`C^{}`$ transition. This is also seen in the optical response of KN125, where the electroclinic coefficient $`e_c`$ is found to increase sharply with decreasing temperature as shown in Fig. 4(b) (solid circles). The divergence of $`e_c`$ indicates an apparent transition to a tilted smectic phase at $`20^{}`$C, whose occurrence is preempted by crystallization. Similar pretransitional behavior is also expected to affect the surface properties of the liquid crystal. Near the apparent bulk transition, the surface region should freeze into an ordered phase — either a tilted, chiral smectic phase, with a twist of the molecular director from layer to layer, or the chiral-stripe phase discussed in Ref. , with a twist of the molecular director in the smectic layer plane. In either case, the chiral environment of the director should enhance the CD signal compared with the uniform smectic-$`A`$ phase. As the temperature decreases, the penetration depth of the ordered phase into the interior should increase, and hence the CD signal should increase. Indeed, Chen et al. confirmed that the size of the interfacial region diverges as the smectic-$`C^{}`$ transition temperature is approached . This is consistent with the result shown in Fig. 4(b) that $`|\mathrm{\Delta }\epsilon _{pk}|`$ increases sharply as the temperature decreases near the apparent transition. Furthermore, when a voltage is applied across the cell, the twist of the director in the ordered phase should be at least partially suppressed, leading to a decrease in the CD signal . This agrees with the observation in Fig. 5 that $`|\mathrm{\Delta }\epsilon _{pk}|`$ decreases as the electric field increases in this temperature range. However, the twist in the interfacial region cannot be unwound by the field, leaving a remnant CD at high field. At high temperature, the behavior should be quite different. In this temperature regime, surface freezing becomes unimportant, and the liquid crystal is purely in the smectic-A phase. For that reason, the chiral twisting of the director should be dominated by the surface electroclinic effect. To find the profile of the electroclinic tilt angle near a surface, we must minimize the free energy subject to the appropriate boundary conditions. The free energy can be written as $`F`$ $`={\displaystyle d^3x\left[\frac{1}{2}K_2|\theta |^2\frac{1}{2}a(𝐧𝐍)^2s(𝐄𝐧\times 𝐍)(𝐧𝐍)\right]}`$ (4) $`={\displaystyle d^3x\left[\frac{1}{2}K_2|\theta |^2\frac{1}{2}a\mathrm{cos}^2\theta sE\mathrm{sin}\theta \mathrm{cos}\theta \right]},`$ where $`𝐧`$ is the molecular director, $`𝐍`$ is the layer normal, $`\theta `$ is the tilt angle between $`𝐧`$ and $`𝐍`$, and $`K_2`$ is the twist elastic constant. Note that this free energy is invariant under the symmetries $`𝐧𝐧`$ and $`𝐍𝐍`$. The final term represents the bulk electroclinic coupling of the tilt angle to the field, with $`s`$ known as the structure coefficient . In the limit of small $`\theta `$, this free energy reduces to the more familiar expression $`F=d^3x[\frac{1}{2}K_2|\theta |^2+\frac{1}{2}a\theta ^2sE\theta ]`$, plus a constant. To minimize the free energy, we rewrite it in the form $$F=d^3x\left[\frac{1}{2}K_2|\theta |^2\frac{1}{4}a^{}\mathrm{cos}2(\theta \theta _0)\right],$$ (5) plus an irrelevant constant, where $`a^{}=[a^2+(2sE)^2]^{1/2},`$ (7) $`\theta _0={\displaystyle \frac{1}{2}}\mathrm{arctan}{\displaystyle \frac{2sE}{a}}.`$ (8) In the bulk, the minimum of the free energy is $`\theta =\theta _0`$; i.e. $`\theta _0`$ is the bulk electroclinic tilt. Near a surface defined by $`y=0`$, the minimum of the free energy is given by the Euler-Lagrange equation $$K_2\frac{d^2\theta }{dy^2}+\frac{1}{2}a^{}\mathrm{sin}2(\theta \theta _0)=0.$$ (9) This differential equation is a standard sine-Gordon soliton equation. Because of the surface electroclinic effect, we impose the boundary condition $`\theta =\mathrm{\Psi }`$ at $`y=0`$, where $`\mathrm{\Psi }`$ is the surface tilt. In the bulk, the boundary condition is $`d\theta /dy0`$ as $`y\mathrm{}`$. Solving the Euler-Lagrange equation with these boundary conditions gives the tilt profile $$\theta (y)=\theta _0+2\mathrm{arctan}\left[\left(\mathrm{tan}\frac{\mathrm{\Psi }\theta _0}{2}\right)e^{y/\xi }\right],$$ (10) where $`\xi =(K_2/a^{})^{1/2}`$ is the tilt correlation length. In principle, one might want to calculate the CD signal from the tilt profile $`\theta (y)`$. In practice, however, the CD signal is a complex molecular property that depends on the detailed interactions of the liquid crystal with light. Even without a theory for the optical properties of the molecules, we can say that the CD signal is a chiral optical property that depends on the chiral parameters of the system. From the tilt profile, we can determine how typical chiral parameters depend on the applied electric field. The simplest chiral parameter that one might calculate is the integrated twist $$Q_1(E)=_0^{\mathrm{}}𝑑y\frac{d\theta }{dy}.$$ (11) This integrated twist is fixed by the boundary conditions to be $`Q_1(E)=\theta _0\mathrm{\Psi }`$. Because the experiment averages over the front and back surfaces of the cell, and over the full period of an AC electric field, it is appropriate to calculate the average chiral parameter $$\overline{Q_1}(E)=\frac{Q_1(E)+Q_1(E)}{2}=\mathrm{\Psi }.$$ (12) Note that this chiral parameter is independent of $`E`$. However, there are many measures of chirality which can correlate with the CD signal . One possibility, allowed by symmetry, is $$Q_3(E)=_0^{\mathrm{}}𝑑y\left(\frac{d\theta }{dy}\right)^3,$$ (13) which is not fixed by the boundary conditions. From (10) we find the field-averaged result $$\overline{Q_3}(E)=\frac{Q_3(E)+Q_3(E)}{2}=\frac{1}{4\xi ^2}\left[2\mathrm{\Psi }+\mathrm{sin}2\mathrm{\Psi }\mathrm{cos}2\theta _0\right].$$ (14) This result shows that the average chiral parameter $`\overline{Q_3}(E)`$ is nonzero only when the system has a nonzero surface electroclinic tilt $`\mathrm{\Psi }`$. This average chiral parameter is nonzero at $`E=0`$ and grows larger in magnitude as $`E`$ increases. For that reason, in the high-temperature regime, we expect that an applied field should also give an increase in the CD signal. This is consistent with the experimental results of Fig. 6. Furthermore, we can estimate how much the applied field should increase the CD signal. The result above implies that $$\frac{\overline{Q_3}(E)\overline{Q_3}(0)}{\overline{Q_3}(0)}=\frac{\frac{1}{\mathrm{cos}2\theta _0}1}{1\frac{\mathrm{sin}2\mathrm{\Psi }}{2\mathrm{\Psi }}}\alpha E^2,$$ (15) with $`\alpha =3e_c^2/\mathrm{\Psi }^2`$, where $`e_cd\theta _0/dE|_{E0}=s/a`$ is the bulk electroclinic coefficient, and the approximation holds for small $`\theta _0`$ and $`\mathrm{\Psi }`$. We have $`\mathrm{\Psi }10`$ deg at all temperatures and $`e_c=0.78`$ deg (V/$`\mu `$m)<sup>-1</sup> at $`60^{}`$C, and hence we expect $`\alpha 0.018`$ (V/$`\mu `$m)<sup>-2</sup>. By comparison, the solid line in Fig. 6(a) shows a fit of the $`60^{}`$C data to the parabolic form $$\mathrm{\Delta }\epsilon _{pk}(E)=\mathrm{\Delta }\epsilon _{pk}(0)\left(1+\alpha E^2\right),$$ (16) which gives the coefficient $`\alpha =0.046`$ (V/$`\mu `$m)<sup>-2</sup>. Thus, our simple model gives at least the right order of magnitude for the sensitivity of the CD signal to applied electric field. Of course, $`\overline{Q_3}(E)`$ is not the only pseudoscalar measure of chirality. While $`\overline{Q_1}(E)`$ was independent of $`E`$, we could consider $$Q_1(E;Y)=_0^Y𝑑y\frac{d\theta }{dy}.$$ (17) where $`Y`$ would be the lengthscale over which surface effects modify the smectic order leading to an enhanced CD signal. Because of the nonlinear dependence of $`\theta (y)`$ on $`E`$ in Eq. (10), the chiral parameter $`\overline{Q_1}(E;Y)`$ will depend quadratically on $`E`$. Moreover, if $`Y\xi `$, the coefficient of the quadratic term will have the same order of magnitude as $`\alpha `$ calculated above. This will be a generic feature of any chiral parameter constructed from $`\theta (y)`$ – the quadratic coefficient will always be of order $`e_c^2/\mathrm{\Psi }^2`$, in agreement with the experimental result. In conclusion, molecular twisting within the layer of a chiral smectic-$`A`$ has been studied using circular dichroism spectroscopy. The results have been discussed in terms of a model based on the surface electroclinic effect. At high temperatures, where the bulk is purely smectic-$`A`$, an applied field leads to an increase in the chiral parameter $`Q_3`$ which can account for the observed field dependence of the CD spectra. At lower temperatures, surface effects enhance the chiral, pretransitional behavior. ## V Acknowledgments We thank Jawad Naciri and Anna Davis for synthesis of the samples used in our studies and T. C. Lubensky, H.-G. Kuball, and S. Chandrasekhar for helpful discussions. This work was supported by the Office of Naval Research. RDK was supported in part by NSF CAREER Grant DMR97-32963.
warning/0001/astro-ph0001312.html
ar5iv
text
# 1 Introduction ## 1 Introduction Clusters of galaxies offer many interesting opportunities to astrophysical and cosmological research. One of them is the study of the cosmic chemical evolution on the largest scale so far. In turn, the heavy element content of clusters results in interesting constraints on the formation and evolution of clusters as well as of their population of galaxies. Current estimates of iron content of the intracluster medium (ICM) are presented in Section 2, together with those of the abundance of the $`\alpha `$-elements. Section 2 also deals with the reported radial abundance gradients in cool clusters, as well as with the relative share of iron and the other heavy elements now in the ICM and locked into stars and galaxies. Section 3 deals with the origin of the ICM metals, identifying in the old, giant spheroids the main producers. Section 4 offers an estimate of the ICM preheating by early supernova driven galactic winds, while Section 5 provides references for further reading concerning the implications for the globla metallicity of the universe at low as well as high redshift. ## 2 Iron and $`\alpha `$-Elements in the ICM As Larson and Dinerstein (1975) predicted and X-ray observations confirmed (Mitchell et al. 1976), the ICM is rich in heavy elements. Actually, it has the larger share of them compared to cluster galaxies. ### 2.1 Iron Fig. 1 shows the iron abundance in the ICM of clusters and groups as a function of ICM temperature from an earlier compilation (Renzini 1997, hereafter R97). Data cluster along two sequences in Fig. 1: for $`kT\begin{array}{c}>\hfill \\ \hfill \end{array}3`$ keV the ICM iron abundance is constant, i.e. independent of cluster temperature. Abundances for clusters in this horizontal sequence come from the Iron-K complex at $`7`$ keV, which emissions are due to transitions to the K level of H-like and He-like iron ions. Conversely, at lower temperatures data delineate an almost vertical sequence, with a fairly strong abundance-temparature correlation reaching extremely low iron abundances in groups cooler than $`1`$ keV. Iron abundances in the vertical sequence from the Iron-L complex at $`1`$ keV, which emission lines are due to transitions to the L level of iron ions with three or more electrons. In these cooler clusters iron is indeed in such lower ionization stages, and the iron-K emission disappears. Such low abundances were regarded with suspicion in R97, being possibly the result of systematic inaccuracies in the atomic physics data used to model the iron-L complex. Fig. 2 is the same of Fig. 1 with one difference. Clusters and groups in R97 with $`kT<2`$ keV have been substituted by a set of 11 groups from Buote (1999). Buote argues that the atomic physics may be correct, but gets substantially higher (perhaps more realistic) abundances when using 2-temperature fits. It remains to be established whether the apparent correlation of the iron abundance with temperature is real or instead is an artifact of inadequate diagnostics at low temperatures. For this reason conclusions concerning abundances derived from iron-K should be regarded as more secure than those for cooler clusters. Fig. 3 shows the iron-mass-to-light-ratio (Fe$`M/L`$) of the ICM, i.e. the ratio $`M_{\mathrm{Fe}}^{\mathrm{ICM}}/L_\mathrm{B}`$ of the total iron mass in the ICM over the total $`B`$-band luminosity of the galaxies in the cluster. Again, objects with $`kT<2`$ keV in R97 have been omitted, with the exception of 4 objects from Buote (1999) for which a measure of the ICM mass was available from the R97 compilation. The drop of the Fe$`M/L`$ in poor clusters and groups (i.e. for $`kT\begin{array}{c}<\hfill \\ \hfill \end{array}2`$ keV) noticed in R97 can be traced back to a drop in both the iron abundance (which however may not be real) and in the ICM mass to light ratio. Such groups appear to be gas poor compared to clusters, which suggest that they may be subject to baryon and metal losses due to strong galactic winds driving much of the ICM out of them (Renzini et al. 1993; R97: Davis et al. 1998). For the rest of this paper I will concentrate on clusters with $`kT\begin{array}{c}>\hfill \\ \hfill \end{array}23`$ keV. Fig.s 1-3 show that both the iron abundance and the Fe$`M/L`$ in rich clusters ($`kT\begin{array}{c}>\hfill \\ \hfill \end{array}3`$ keV) are independent of cluster temperature, hence of cluster richness and optical luminosity. For these clusters one has $`Z_{\mathrm{ICM}}^{\mathrm{Fe}}=0.3\pm 0.1`$ solar, and $`M_{\mathrm{Fe}}^{\mathrm{ICM}}/L_\mathrm{B}=\left(0.02\pm 0.01\right)`$ for $`H_{}=50`$. The most straightforward interpretation is that clusters did not lose iron (hence baryons), nor differentially acquired pristine baryonic material, and that the conversion of baryonic gas into stars and galaxies has proceeded with the same efficiency and the stellar IMF in all clusters (R97). Otherwise, we should observe cluster to cluster variations of the iron abundance and of the Fe$`M/L`$. The absence of such variations also argues for the iron now residing in the ICM having been ejected from galaxies by supernova- (or AGN-) driven galactic winds, rather than having been stripped by ram pressure (Renzini et al. 1993; Dupke & White 1999). Indeed, ram pressure effects become much stronger with incereasing cluster richmess, hence galaxy velocity dispersion and ICM temperature. Correspondingly, if ram pressure would play a major role in getting iron out of galaxies one would expect the ICM abundance and Fe$`M/L`$ to increase with $`kT`$, which is not observed. ### 2.2 The $`\alpha `$-Elements From ASCA X-ray observations the ICM abundances of some of the $`\alpha `$-elements (O, Ne, Mg, and Si) have also beed derived, yielding an average $`<[\alpha `$/Fe\]$`>+0.2`$ (Mushotzky et al. 1996). This modest $`\alpha `$-element enhancement suggested an ICM enrichment dominated by SNII products. However, the $`\alpha `$-element overabundance vanishes when consistently adopting the “meteoritic” iron abundance for the sun, as opposed to the “photospheric” value (Ishimaru & Arimoto 1997), with a formal average $`<[\alpha `$/Fe\]$`>+0.04\pm 0.2`$ (R97). Clusters of galaxies as a whole are therefore nearly solar as far as the elemental ratios are concerned, which argues for stellar nucleosynthesis having proceeded in much the same way in the solar neighborhood as well as at the galaxy cluster scale. This argues for a similar ratio of the number of Type Ia to Type II SNs, as well as a similar IMF (R97), suggesting that the star formation process (IMF, binary fraction, etc.) is universal, with little or no dependence on the global characteristics of the parent galaxies in which molecular clouds are turned into stars. ### 2.3 Radial Abundance Gradients, Are They Real? The values of the Fe$`M/L`$ shown in Fig. 3 were derived as the product of the ICM mass times the central iron abundance. However, radial gradients in the iron abundance have been reported for several clusters (e.g. Fukazawa et al. 1994; Dupke & White 1999); Finoguenov, et al. 1999; White 1999), which would require a radial integration of $`\rho \left(r\right)Z_{\mathrm{Fe}}\left(r\right)dV`$ to derive the total iron mass in the ICM. This is not currently feasible because ASCA data lack of the sufficient image quality to do a proper job. Nevertheless, a cursory examination of the radial gradient so far reported reveals two intriguing trends. The first is that the presence of radial abundance gradients is confined to cool clusters and groups ($`kT\begin{array}{c}<\hfill \\ \hfill \end{array}3`$ keV, with hotter clusters – which incidentally are generally those with the best S/N ratios – never showing appreciable gradients. The second trend one can notice in the published data is that an abundance gradient is most often associated to a temperature gradient: iron abundances increase inwards as temperature drops towards the center to $`2`$ keV or below. These trends are illustrated in Fig. 4, showing the abundance-temperature relation for a representative set of clusters (Pellegrini 2000). Given the uncertainties affecting the low temperature abundances – resulting from either problems with the atomic physics or from the complexities of a multi-temperature ICM – it seems prudent to conclude that the reality of the reported abundance gradients is far from beeing established. Although more or less plausible mechanisms could be invented to produce radial abundance gradients in clusters, it is neverhteless worth entertaining the possibility of the reported gradients being a mere artifact of inadequate diagnostics. ### 2.4 Most Metals are not in Galaxies The metal abundance of the stellar component of cluster galaxies can only be inferred from integrated spectra coupled to synthetic stellar populations. Much of the stellar mass in clusters is confined to passively evolving spheroids (ellipticals and bulges) for which the iron abundance may range from $`1/3`$ solar to a few times solar. Following R97, the ratio of the iron mass in the ICM to the iron mass loked into stars and galaxies is given by: $$\frac{Z_{\mathrm{ICM}}^{\mathrm{Fe}}M_{\mathrm{ICM}}}{Z_{}^{\mathrm{Fe}}M_{}}1.65h^{3/2},$$ (1) or 4.6, 2.5, and 1.65, respectively for $`h=0.5`$, 0.75, and 1 ($`h=H_{}/100`$), and having adopted $`Z_{\mathrm{ICM}}^{\mathrm{Fe}}=0.3`$ solar and $`Z_{}^{\mathrm{Fe}}=1`$ solar. Note that with the adopted values for the quantities in equation (2) most of the cluster iron resides in the ICM, rather than being locked into stars, especially for low values of $`H_{}`$. Therefore, it appears that there is at least as much metal mass out of cluster galaxies (in the ICM), as there is inside them (locked into stars). This sets a strong constraint by models of the chemical evolution of galaxies. ## 3 Origin of the Metals in the ICM The constant Fe$`M/L`$ of clusters says that the total mass of iron in the ICM is proportional to the total optical luminosity of the cluster galaxies (Songaila et al. 1990; Ciotti et al. 1991; Arnaud et al. 1992; Renzini et al. 1993). This strongly argues for the iron (metals) now in the ICM having been produced by the (massive) stars of the same stellar population to which belong the low-mass stars now producing the bulk of the cluster optical light. As well known, much of the cluster light comes from the old spheroidals (ellipticals and bulges), hence one can say that the bulk of cluster metals were produced by ellipticals and bulges when they were young. It is now well established that the stellar populations in spheroidals – in clusters as well in the field – are very old, with the bulk of stars having formed at $`z\begin{array}{c}>\hfill \\ \hfill \end{array}3`$ (for a recent review with extensive references see Renzini & Cimatti 1999). Therefore, the simplest interpretation of the data suggests that the bulk of the heavy elements in the ICM were produced and expelled from galaxies a long time ago, i.e. at $`z\begin{array}{c}>\hfill \\ \hfill \end{array}3`$. If so, the ICM abundances should not show any significant evolution all the way to very high redshifts. Existing data are in agreement with this prediction (see Fig. 1), but do not reach much beyond $`z0.5`$. Future X-ray observatories could check this prediction. It is also interesting to address the question of which galaxies have produced the bulk of the iron and the other heavy elements, i.e. the relative contribution as a function of the present-day luminosity of cluster galaxies. This has been recently investigated by Thomas (1999), from whom I reproduce here Fig. 5. This shows that the bright galaxies (those with $`LL^{}`$) produce the bulk of the cluster light, while the dwarf contribute a negligible amount of light in spite of dominating the galaxy counts by a large margin. Hence, most galaxies don’t do much, while the brightest $`3\%`$ of all galaxies contribute $`97\%`$ of the whole cluster light. Given that it is most likely that the metals were produced by the same stellar population that now shines, one can safely conclude that also the bulk of the cluster metals have been produced by the giant galaxies (or – paying a tribute to a widespread belief – by the stars that are now in the giants). Dwarfs contribution to ICM metals may have been somewhat larger than their tiny contribution to the cluster light, since metals can more easily escape from their shallower potential wells (Thomas 1999), but this cannot alter the conclusion that the giants dominate metal production by a very large factor. ## 4 A Semiempirical Estimate of Preheating The total amount of iron in clusters represnts a record of the overall past supernova (SN) activity as well as of the past mass and energy ejected from cluster galaxies. The empirical values Fe$`M/L`$ was used to set a constraint on the so-called preheating of the ICM (Renzini 1994), a hot topic indeed at this meeting. The total SN heating is given by the kinetic energy released by one SN ($`10^{51}`$ erg) times the number of SNs that have exploded. It is convenient to express this energy per unit present optical light $`L_\mathrm{B}`$, i.e.: $$\frac{E_{\mathrm{SN}}}{L_\mathrm{B}}=10^{51}\frac{N_{\mathrm{SN}}}{L_\mathrm{B}}=10^{51}\left(\frac{M_{\mathrm{Fe}}}{L_\mathrm{B}}\right)^{\mathrm{TOT}}\frac{1}{<M_{\mathrm{Fe}}>}10^{50}\left(\mathrm{erg}/L_{}\right),$$ (2) where as total (ICM+galaxies) Fe$`M/L`$=0.03 $`M_{}/L_{}`$ is adopted, and the average iron mass release per SN event is assumed to be $`0.3M_{}`$. This is the appropriate average if $`3/4`$ of all iron is made by Type Ia SNs, and $`1/4`$ by Type II and other SN types (see R97). Were all the iron made instead by SNIIs, the resulting SN energy would go up by perhaps a factor of 3 or 4. This estimate should therefore be accurate to within a factor of 2 or 3. The presence of a large amount of iron in the ICM demonstrates that matter (and then energy) has been ejected from galaxies. The kinetic energy injected into the ICM by galactic winds, again per unit cluster light, is given by 1/2 the ejected mass times the typical wind velocity squared, i.e.: $$\frac{E_\mathrm{w}}{L_\mathrm{B}}=\frac{1}{2}\frac{M_{\mathrm{Fe}}^{\mathrm{ICM}}}{L_\mathrm{B}}\frac{v_\mathrm{w}^2}{Z_\mathrm{w}^{\mathrm{Fe}}}1.5\times 10^{49}\frac{Z_{}^{\mathrm{Fe}}}{Z_\mathrm{w}^{\mathrm{Fe}}}\left(\frac{v_w}{500\mathrm{km}\mathrm{s}^1}\right)^210^{49}\left(\mathrm{erg}/L_{}\right),$$ (3) where the empirical Fe$`M/L`$ for the ICM has been used, the average metallicity of the winds $`Z_\mathrm{w}^{\mathrm{Fe}}`$ is assumed to be twice solar, and the wind velocity $`v_\mathrm{w}`$ cannot be much different from the escape velocity from individual galaxies, as usual in the case of thermal winds. Again, this estimate may be regarded as accurate to within a factor of 2, or so. A first inference from these estimates is that $`510\%`$ of the kinetic energy released by SNs survives as kinetic energy of galactic winds, thus contributing to the heating of the ICM. A roughly similar amount should go into work to extract the gas from the potential well of individual galaxies, while the bulk $`8090\%`$ has to be radiated away locally and does not contribute to the feedback. The estimated $`10^{49}`$ erg/$`L_{}`$ correspond to a “preheating” of $`0.1`$ keV per particle, for a typical cluster $`M_{\mathrm{ICM}}/L_\mathrm{B}40M_{}/L_{}`$. This is $`20`$ times lower than the (AGN induced) preheating some model require to fit the cluster $`L_\mathrm{X}T`$ relation (Wu et al. 1999). Not even if all the SN energy were to go to increase the thermal energy of the ICM whould this extreme requirement be met. Crucial for the evolution of the ICM is however the entropy, hence the relative timing of galactic wind heating and cluster formation (Kaiser 1991; Cavaliere et al. 1993). The entropy increase associated to a given amount of preheating depends on the gas density when the heating takes place.. Most of star formation in cluster elliptical galaxies – and therefore most SN activity and galactic winds – appears to be confined at very early times, i.e., at $`z\begin{array}{c}>\hfill \\ \hfill \end{array}3`$. It remains to be seen whether such early preheating may render $`0.1`$ keV per particle sufficient to increase enough the entropy thus producing the observed $`L_\mathrm{X}T`$ relation. Alternative models less demanding than those of Wu et al. in terms of preheating are presented by Tozzi at this meeting (Tozzi & Norman 1999). Finally, if AGN preheating were as high as 2 keV/particle (averaged over the whole cluster) it may result in long term suppression of cluster cooling flows, in analogy with the scenario proposed for intermittent gas flows in ellipticals (Ciotti & Ostriker 1997). ## 5 Clusters as Fair Samples and the Metallicity of the High-$`z`$ Universe In R97 and in some its subsequent updates (e.g. Renzini 1998, 1999) it is argued that clusters are representative of the local universe as a whole, as far as the fraction of baryons already turned into stars is concerned, hence of the global metallicity of the local universe ($`1/3`$ solar). Moreover, since stars in spheroids formed at $`z\begin{array}{c}>\hfill \\ \hfill \end{array}3`$ and spheroids account for at least 1/3 of the present day total mass in stars, it is concluded that the metallicity of the whole universe at $`z3`$ had to be $`1/3`$ of the present global value, hence $`1/10`$ solar. This simple argument supports the notion of a prompt initial enrichment of the early universe. Acknowledgements. I am grateful to Silvia Pellegrini and Daniel Thomas for their permission to reproduce here Fig. 4 and Fig. 5, respectively.
warning/0001/astro-ph0001538.html
ar5iv
text
# Quasar Candidate Multicolor Selection Technique: a different approach. ## 1 Introduction Nowadays, very high redshift galaxies and quasars are being found. However, quasars still offer the opportunity of a (relatively) easy construction of high-z samples, due to their high luminosity, quasi-stellar shape (allowing accurate photometry), and to their emission lines (allowing spectroscopical identification). Therefore, many large optical quasar surveys are in progress or in project. Even if high energy emission is quite a common quasar property, X-ray samples still need the detection of emission lines in the optical domain for the purposes of redshift determination. To improve the efficiency of all quasar studies, one has to increase both the size and the upper redshift limit of the sample, and to achieve a maximum completeness over a redshift range as large as possible. These surveys are essentially based on the photometric preselection of quasar candidates, in order to optimize the efficiency of telescope time. Preselection makes the completeness questionable both at low and at high redshift (see discussion in e.g. Miller & Mitchell, 1998). Even multiple color-color diagrams Hall et al., 1996b still reveal a bias in the range $`2.2z3`$, exactly where a reversing of evolution is thought to occur, making completeness specially important. At low redshift, it has been shown that the usual morphologic prerequisite for quasi-stellar candidates results in a noticeable deficiency Wolf et al., (1998). This trend may increase the apparent evolution of quasars. The present paper describes a joint method both for distinguishing between quasars and stars/galaxies by their photometry and for obtaining an estimate of the photometric redshift of the quasar candidates as well. In §2 we present the current state of the art in the optical search for quasars. §3 gives a brief description of our method, which has been first validated with simulations, as shown in §4, before being applied to real samples (§5). The last section of this paper (§6) contains a summary of our conclusions and a brief discussion. ## 2 State of the art in the optical search for quasars One of the most common problems in optical surveys is the selection criteria one has to apply in order to distinguish between stars and quasars, since both categories of objects have (in most cases) a point–like morphology. Quasars are spatially resolved only at low redshifts and only when high quality imaging data are available. Selection of QSO candidates among stellar objects from photometric data is becoming a standard method. The UVX technique (UV excess) separates stars from quasars with a redshift $`z2.2`$. This technique is still currently being used, e.g. in the Chile-UK quasar survey, and the AAT 2dF QSO Survey. To avoid both the redshift limitation $`(z<2.2)`$ and the bias towards blue objects, other equivalent techniques are also being applied (for example BRX), with quite a high efficiency in other redshift ranges. Many Multicolor Surveys use an increased number of filters, especially towards the red, with special care paid to high-z quasars: examples are found in Boyle et al. (1991), Kennefick et al. (1997), Osmer et al. (1998) and the references therein, Jarvis & Mac Alpine (1998) and the DPOSS (Digital POSS II). In all these cases, the efficiency never exceeds 35% ($``$ 3 candidates selected per one real quasar), depending on magnitude and redshift. For all the selection methods described above, the working space is the N $`\times `$ 2D-manifold of N color-color planes. The locus of quasars at different redshifts in each of these planes is compared to the locus expected for Main Sequence stars. For quasars, the allowed multi-color space is defined through simulated template spectra, set to different redshifts. Basically all the selection procedures use the positions of candidates in the multi-color space either to compare with the expected locus for quasars or to compute the distance to the Main Sequence (Newberg & Yanny, 1997; Sloan Digital Sky Survey - Fan, 1999; Krisciunas et al., 1998). Quasar candidates are among the objects at the largest distances with respect to this Main Sequence “Snake”. However the selection of candidates with quasar spectra is basically a 1D problem: the fitting between the observed spectral energy distribution (SED) and the equivalent one obtained from templates (quasars/stars), all of them computed within the same photometric system. Two quasar surveys make use of such a fit to select candidates and to estimate their photometric redshifts. The Calar Alto Deep Imaging Survey (CADIS) uses specialized filters and a multicolor classification algorithm to distinguish between stars, galaxies and quasars, and the identification of high-z quasars in particular is based on the photometric redshift method Wolf et al., (1998). The Large Zenith Telescope, a 6-meter liquid-mirror telescope, will provide photometric redshifts for more than 1 million of galaxies, thanks to a series of 40 medium-band filters specially designed, and will be able to select quasar candidates quite efficiently. For the purposes of this paper, even a reduced set of broad band photometric data on a given object shall be considered as a very low resolution spectrum (R$``$ 1!), which could be fitted by synthetic templates according to the standard methods commonly used on photometric redshift estimates. Assumptions on the morphology of the objects could be avoided in principle, in such a way that both point-like and spatially resolved sources could be examined through the same pipeline. Furthermore, it is interesting to have a prior idea on the redshift range of each candidate, in order to perform the spectroscopic follow up in the suitable wavelength range (visible or IR). The software described in the present paper has been developed for the needs of preparing the quasars/AGN sample to be assembled in the future VIRMOS survey. It can however be applied to all multicolor catalogues. The forthcoming VIRMOS spectroscopic galaxy survey will consist in 2 surveys: a deep one and a shallow one. The VIRMOS-SHALLOW spectroscopic sample (statistically 1/4 to 1/3 complete to $`I22.5`$, no other preselection) will ensure random completeness, without any color or morphology preselection. The VIRMOS-DEEP survey (complete for all objects with $`22.5<I24`$) will provide the opportunity of assembling a quasar sample avoiding the drawbacks of any preselection, therefore basically free of the usual biases (colors, redshift etc). So far, very few such samples are available: three examples come from the CFRS (I=22.5, Schade et al., 1995), the Faint Galaxy Redshift Survey (B=24, Glazebrook et al., 1995) and more recently Cohen et al. (1999) to R=24 in the HDF, involving, however, very few quasars. We are examining various possibilities of assembling a third quasar sample in the limits of the VIRMOS-SHALLOW survey, trying to get as much as possible of the spectra of the several thousands of quasars present in this field by selecting photometric candidates. Comparing the three VIRMOS samples should provide a unique opportunity of examining the biases introduced by a photometric and/or morphological preselection on the quasar samples. ## 3 Description of the method Our method used here is closely based on hyperz, a public photometric redshift code presently under development at the Observatoire Midi-Pyrénées. Details on hyperz will be given in a forthcoming paper (Bolzonella et al., 2000; see also Miralles & Pelló, 1998; Pelló et al., 1999). It is basically a standard SED fitting procedure: the observed photometric SED of a given object, obtained through $`n`$ filters, is compared to the SED computed for a set of template spectra. The aim is to find the best fit between the observed and the model photometry through a standard $`\chi ^2`$ minimization procedure: $$\chi ^2=\underset{i=1}{\overset{n}{}}\frac{(F_{obs}^iF_{mod}^i(z))^2}{\sigma _i^2}$$ where $`F_{obs}^i`$ and $`F_{mod}^i`$ are respectively the observed and the template fluxes in the $`i`$ band, and $`\sigma _i`$ is the error on the observed flux in this band. Fluxes are normalized to an arbitrary reference filter. This is a quite general method. The photometric redshift is the redshift $`z`$ corresponding to the minimum $`\chi ^2`$ value. As expected, the accuracy on the photometric redshift determination depends strongly on the set of filters used (number and wavelength coverage); this is illustrated in the next paragraph. In order to apply this method to quasars, the template set is made by series of simulated quasar spectra. These spectra have been constructed by varying the slope of the power-law spectra (spectral index) in the optical (between 0.0 and 1.0, in 3 steps) while keeping the ultraviolet index, $`a_{UV}`$, constant at 1.76 (a value compatible with Wang et al., 1998). Emission lines are included (Ly<sub>α</sub>, Ly<sub>β</sub>, CIII, CIV, MgII, SiIV, H<sub>α</sub>, H<sub>β</sub> and H<sub>γ</sub>) with gaussian profiles and typical intensities (e.g. Peterson, 1997), as well as the small blue bump, centered at 3000 Å. The redshift value is set between 0 and 7. Totally, there are 213 quasar spectra, with 71 redshift steps (d$`z=0.1`$), with 3 simulated spectra on each step, obtained by varying the optical spectral index. The Ly<sub>α</sub> forest has been modelized according to Madau (1995), while no reddening has been included. We also included the stellar library of Pickles (1998), which contains 108 spectra of stellar types ranging from O to M, including dwarfs and giant stars. Additionally, we included 22 white dwarf and carbon stars spectra, which are known to contaminate the quasar candidates catalogues. When applying the method to a given sample, the discrimination between stars and quasars is performed in two steps. A first distinction between stars and quasars is made by identifying as star (quasars) the objects that show a better fit to one of the stellar (quasar) spectra. A further selection on the remaining sample can be made, based on the value of the reduced $`\chi ^2`$, with stars identified as objects excluded as quasars at the 95% or 99% confidence levels. In the present paper, the terms “$`\chi ^2`$ selection” or “no $`\chi ^2`$ selection” will apply for such a distinction. When an object is identified as quasar candidate, a photometric redshift estimation is automatically provided. The number of degrees of freedom, $`\nu `$, is equal to the number of filters minus 2: we lose one degree by normalizing the flux to a reference filter and one degree because of the “best fit” procedure. In this way, the separation between stars and quasars is reduced to a one–dimensional fit between the data (the observed SED) and the spectral templates. As expected, and as confirmed by the simulations, the quality of the quasar/star discrimination depends on the filter combination, as it happens for the photometric redshift estimate in the case of galaxies. Not only the number of filters is important, but also the wavelength range covered. Another important point is the width of the filters which must be suited to the width of quasar spectral features. A realistic determination of the photometric errors is the key to a successful issue for this method. ## 4 Simulations The first test of this method has been done on mixed (quasars + stars) simulated catalogues. In this paragraph we intend to show the importance of the set of filters used for optimizing the results. First, let us define an arbitrary and useful quantity which is aimed to quantify the efficiency of this procedure. Let $`N_c`$ be the number of quasar candidates resulting from the identification technique, $`N_f`$ the number of actual quasars found as candidates (real quasars effectively selected), and $`N_e`$ the number of expected quasars. In principle, $`N_e`$ is a quantity which can be estimated from previous surveys. We define the “efficiency” of the method as follows: efficiency=completeness $`\times `$ confirmation rate, where completeness$`=N_f`$/$`N_e`$ and confirmation rate $`=N_f`$/$`N_c`$. Obviously, both the completeness and the confirmation rate take values in the interval and so does the efficiency. The number of stars in our simulated catalogues was given by the Galaxy model of Robin et al., 1995 (R95). The star counts correspond to the six Deep Multicolor Survey (hereafter DMS) fields, since we primarily tested our results with the results of this survey. The fields with their surface (in square arc minutes) and the estimated star numbers are given in table 1. Quasar numbers were based on previous surveys (Hartwick & Schade, 1990, hereafter HS90). Catalogues contain $``$100 quasars and $``$3000 stars, with a magnitude limit of $`m_b`$=22.2 and photometric errors scaling with magnitude, with typical values of the order of 0.1 magnitudes. Quasars are uniformly distributed in redshift, in order to obtain reasonable statistical results at all redshifts. Photometry in broad-band filters has been simulated, covering the wavelength range between $``$ 3000 Å and $``$ 20000 Å: U, B, V, R, I, H, and K. Broad-band filters have been used for practical reasons, in particular because they directly apply to the VIRMOS survey, which will make use of wide-band filters only, although narrow-band filters would give better results (Peri, Iovino & Hickson, 1997). 40 different filter combinations have been tested in the aim of selecting quasar candidates. Fig. 1a and b illustrate the completeness (solid line), confirmation rate (dashed line) and efficiency (dash-dotted line) for these 40 combinations, without any $`\chi ^2`$ selection and with selection at 99% confidence level, respectively. The filter combinations in fig. 1a have been classified by increasing completeness. The same order has been used in fig. 1b, to demonstrate the effect of a $`\chi ^2`$ selection on the completeness and the other parameters. Table 2 shows the filter combination which corresponds to the abscissa values. A closer look at fig. 1 and Table 2 reveals that the filter combination is, most of the time, more important than the number of filters. In general, filters like U or the infrared filters H and/or K are very useful: they add much photometric information constraining the number of candidates. Neighboring filters like B, V, R, when used all together, decrease the efficiency because, as mentioned above, they increase the number of the degrees of freedom without adding any essential information. A comparison of the two figures reveals several trends. The completeness is almost not influenced by a $`\chi ^2`$ selection in the case of 3-filter combinations (such as UBV, BRI, BVR, BVI, or UBR). It is, however, strongly influenced by a $`\chi ^2`$ selection when the filter U is present, in sets of 4 or more filters. Note that this quantity decreases significantly , when a $`\chi ^2`$ selection is imposed, for filter combinations with index higher than 26, which are all sets containing the U-filter. Furthermore, the confirmation rate is generally increasing with increasing number of filters. The decision whether to make a $`\chi ^2`$ selection or not depends on the available time for spectroscopy and the scientific objectives of the project. For this purpose, we made an approximative estimation of the time needed for making the spectroscopy in order to cover an area of 1 square degree, $`T_{sp}`$, for the same 40 filter combinations, given a detector area, $`S`$, a magnitude limit, $`m_b`$, a quasar candidate density, $`n_q=N_{cand}/S`$ (depending on this limit as well as on the filter combination), a number of slits, $`G`$ and an exposure time, $`dT_{sp}`$, with and without any $`\chi ^2`$ selection. We calculate $`T_{sp}`$ as: $$T_{sp}=\left\{1+int\left(\frac{N_{cand}}{G}\right)\right\}\times \frac{dT_{sp}}{S}$$ Fig. 2 illustrates $`T_{sp}`$ (in hours) for $`m_b=22.2`$, S=0.06 square degrees, G=10 and $`dT_{sp}=`$3 hours, without any preselection (solid line) and with a $`\chi ^2`$ selection (dashed line). Note that the y-axis is logarithmic. As expected, there is an anticorrelation between $`T_{sp}`$ and the confirmation rate, as it appears from the comparison of fig. 1 and 2. The highest the confirmation rate, the less time is required. ### 4.1 Determination of the Photometric Redshifts The problem of the quasar photometric redshift (hereafter $`z_{phot}`$) estimation is at least as complex as the one for galaxies. Quasar spectral features (mainly emission lines) are rather narrow and the overall shape of the ultra-violet and optical continuum is a single power low, while the filters commonly used for photometry are wide-band filters, not very suitable for the quasar characteristics. All the above are in the origin of the dispersed values of the estimated quasar $`z_{phot}`$, when compared to the spectroscopic ones, especially for quasars with redshifts lower than $`2`$. The filter combination is very important for the accurate determination of the $`z_{phot}`$. Figure 3 shows the photometric redshift versus the model redshift, for a catalogue of 1540 simulated quasars, uniformly distributed over the redshift range , and for 4 different filter combinations. From top to bottom and left to right the following filter combinations are illustrated: BVRI (a), BVIK (b), UBVRI (c), UBVRIK (d). The use of the filter set BVRI is clearly a bad choice: these four filters cover a relatively short wavelength range and their effective wavelengths lie close to each other. The $`z_{phot}`$ in this case is poorly determined for $`z3`$, while there is also a large dispersion at $`z5`$. If we replace the R-filter by the K-filter (figure 3b), the dispersion becomes more important at redshifts higher than 4, since this is the redshift at which Ly<sub>α</sub> should enter the R-band, but it decreases the dispersion at redshifts lower than 3, due to the very large wavelength range covered. Adding the filter U to the BVRI combination decreases the dispersion at both high and low redshifts (figure 3c), but one can still feel the absence of an infrared filter, whose importance is seen in figure 3d. The periodic accumulations of points at different redshifts are due to the passage of Ly<sub>α</sub> from one filter to the other. As a general rule, the determination of the photometric redshifts should be rather accurate in the redshift range even without the U or an infrared filter, but their presence is crucial for all other redshift ranges. The quality of the $`z_{phot}`$ determination is likely to improve if real quasar spectra are used. However, this implies that a spectral library is required, containing some hundreds of quasar spectra, with redshifts spanning from 0 to $`5`$ and with a spectral coverage spanning form the ultra-violet to the infrared. ### 4.2 C, CR and E as a function of redshift The completeness, the confirmation rate and, therefore, the efficiency of the method depend on the redshift. The demonstration of this dependence is easy to show for the completeness, but not for the confirmation rate. The completeness within redshift bins has the same definition as for the entire sample: the ratio between the number of real quasars with spectroscopic redshifts (hereafter $`z_{spec}`$) within the examined bin, identified as candidates even if their $`z_{phot}`$ is significantly different from their $`z_{spec}`$, over the estimated total number of quasars. Figure 4 illustrates the completeness versus the redshift, without $`\chi ^2`$ selection, for 4 different filter sets: the solid, dotted, dashed and dashed-dotted lines correspond to BVIK, UBVRIH, UVK and BVRI, respectively. The completeness over the whole redshift range for this 4 filter combinations is in all cases more than 85%, as shown in figure 1a. Figure 4 reveals that, independently of the filter combination, the completeness is always lower in the redshift range due, as already mentioned, to the stellar–like colors of quasars. This fact could certainly increases the number of quasar candidates within this range but it could also be a possible source of incompleteness, since some quasars will inevitably be classified as stars. Furthermore, different filter combinations offer a different completeness at different redshifts. The filter set BVIK assures a completeness of over 70% in the critical redshift range , which however remains low up to $`z5.5`$. UVK offers the possibility of a completeness of $``$80% for $`z<2`$, while UBVRIH results give a completeness over 90% for high-$`z`$ quasar samples ($`z>3`$). Things become more complicated if one tries to make an equivalent evaluation of the confirmation rate within redshift bins. This time, the confirmation rate is the ratio between the number of real quasars with $`z_{phot}`$ in the examined redshift bin, over the number of quasar candidates (quasars + possibly stars) with $`z_{phot}`$ that belong to this particular redshift interval. Obviously, this definition depends on the accuracy of the determination of the $`z_{phot}`$ as well as on the width of the redshift bins. Therefore, the dependence of the confirmation rate versus the redshift can be evaluated qualitatively but not quantitatively, and so it goes for the efficiency as well. If the final aim of a survey is assembling a quasar sample with redshifts within a restrained interval, all candidates with a $`z_{phot}`$ that lies outside this interval will, in principle, be rejected. In the more realistic case of an all redshift survey, the most interesting strategy is to simply optimize the spectroscopy towards the optical or the infrared wavelengths. The spectroscopy of quasars with redshifts up to $``$3 is much more interesting in the optical band, while higher redshift quasars should be spectroscopically observed in the infrared, since most of their emission lines are shifted to these wavelengths. The simulations have shown that, although the determination of quasar $`z_{phot}`$ is sometimes dispersed at $`z2.5`$ and $`z4.5`$ (figure 3), this effect hardly translates into high-$`z`$ quasars ($`z>4`$) being wrongly identified as low-$`z`$ ones ($`z<3`$) or vice versa. ## 5 Applying the method to real samples The method presented above has been tested on a series of samples found in the literature, giving rather satisfactory results. Depending on the sample and the available information, the confirmation rate and/or the completeness have been calculated. ### 5.1 The Deep Multicolor Survey The Deep Multicolor Survey is an imaging survey covering 6 fields on a total area of 0.83 square degrees on the northern sky at high galactic latitude, carried out at the 4m Mayall telescope of the Kitt Peak National Observatory. 6 filters have been used, covering the wavelength range from 3000 to 10000 Å: U, B, V, R’, I75 and I86, and a photometric catalogue of $``$ 21000 stellar–like objects has been assembled, with the aim of conducting a multicolor quasar search (Osmer et al. (1998) and the references therein). 137 stars, 49 compact narrow emission line galaxies (hereafter CNELGs), and 54 quasars with redshifts spanning from 0.3 to 4.3 have already been spectroscopically identified. These objects were the 240 quasar candidates selected by Hall et al. according to their position on color-color plots, a high fraction of them (194 objects) selected by the UVX technique (Hall et al., 1996). This selection favored quasars with redshifts less than 2.2. The redshift distribution of these objects can be seen in fig. 5, where the spectroscopic sample is plotted by a dotted line. For our purposes, we cut the catalogue at B=22.3, a limit at which the sample is almost 100% complete and only marginally contaminated ($``$ 8%) by galaxies, and we only used objects for which photometry was available in all 6 filters. Hence, 3 of the spectroscopically confirmed quasars were excluded, and we ended up with a catalogue of 3720 stellar-like objects. For the magnitude limit and the area covered by the survey, approximately 106$`{}_{22}{}^{}{}_{}{}^{+27}`$ quasars are expected to be found among the objects of the catalogue (HS90), resulting a completeness of at least 80% for the usual multicolor selection techniques (Hall, private communication) and $``$88% for our method. Table 3 summarizes the characteristics of the different catalogues examined in this section: surface covered, magnitude limits, filters used, number of quasars, and number of candidates for different confidence levels. The expected number of stars on the DMS field has been computed from R95. In the case of the DMS field, the number of spectroscopically confirmed quasars is also given between brackets. Among the 3720 objects in this catalogue, we identified 3332 as stars because their photometry shows a better match to the stellar templates. With our method, 6 of the 51 spectroscopically confirmed quasars (DMS0059-0056, DMS1358-0054, DMS1714+5012, DMS1714+5003, DMS1714-4959 and DMS1358-0055) are misclassified as stars. Only 7 of the 97 B $`22.3`$ candidates, spectroscopically identified as stars, but all 34 B $`22.3`$ CNELGs belong to our candidate list, if no $`\chi ^2`$ selection is made. At a 99% confidence level, 25 CNELGs are classified as quasar candidates. To summarize, we find 388 quasar candidates for 106 actual quasars, with an expected completeness of 45/51=0.88 and an expected confirmation rate of (106 $`\times `$ 0.88)/388=0.24, if no $`\chi ^2`$ selection is made. The equivalent values with 99% confidence level selection are 0.88 and 0.27, respectively, for 348 quasar candidates (see also Table 4). Fig. 5 displays with a solid line the estimated redshift distribution of our 388 quasar candidates. The accumulation of identifications around $`z1`$ is due to the CNELGs, with $`z_{spec}`$ in the range \[0.1,0.8\] typically, identified by our procedure as quasar candidates with $`z_{phot}1`$. The broad distribution at $`z[2.0,3.0]`$ is probably due to the stellar–like colors of quasars at these redshifts, which makes the distinction between stars and quasars particularly difficult by photometric means, increasing the number of quasar candidates, but it also corresponds to the expected maximum of the quasar space distribution. For comparison, we plot the redshift distribution of the 51 spectroscopically confirmed quasars (dotted line). The above results are summarized in the first 6 columns of Table 4, where the values of the completeness (C), the confirmation rate (CR) and the efficiency (E) are shown in columns 3 and 4. The values predicted by the simulations are also given for comparison in columns 5 and 6, using the same filter combinations as for the real data, and keeping the filter responses as close as possible to those actually used. Column 2 gives the same quantities for the DMS and the usual multicolor selection techniques. Note that the values given for the completeness and the efficiency are the lower estimations. The two methods give comparable values for all 3 quantities (C, CR and E), with a slightly better efficiency for the multicolor technique when a $`\chi ^2`$ criterion is imposed. In order to estimate the contamination due to compact galaxies, we have added to the template library the spectra of blue starburst galaxies, obtained from the public model Starburst99 code by Leitherer et al. (1999), as well as a delta-burst model taken from the new Bruzual & Charlot evolutionary code (GISSEL98, Bruzual & Charlot 1993), and a set of 4 empirical SEDs compiled by Coleman, Wu and Weedman (1980) (hereafter CWW) to represent the local population of galaxies (E, Sbc, Scd and Im). Spectra from the Starburst99 code were selected with solar metallicity and Salpeter IMF, taking different ages for an instantaneous burst ranging from 1 to 20 Myr. In the case of the GISSEL98 delta-burst, the IMF is that of Miller & Scalo (1979), with solar metallicity and 51 different ages ranging from 0.001 to 20 Gyr. CWW spectra were extended to wavelengths $`\lambda 1400`$ Å and $`\lambda 10000`$ Å using the equivalent GISSEL98 spectra. When applying the method to the DMS sample of 3523 stellar objects (we remind here that we only use objects with photometry in all 6 filters), using the template library extended with galaxy spectra, we find that $`80\%`$ of them are better fitted by stellar or quasar templates rather than by galaxy spectra. This fraction is only 56% for the subsample of 97 objects spectroscopically identified as stars. Among the subsample of the 51 objects spectroscopically identified as quasars, there are 38% that show a better fit to the galaxy spectra, and 80% among them correspond in particular to genuine young starbursts according to the best-fit spectra, with ages ranging between 1 and 20 Myrs. The result is different when the same extended template library is applied to the 34 B $`22.3`$ DMS CNELGs. In this case, 85% of the sample (29 objects) is well fitted by galaxy spectra rather than quasar or stellar templates. The fraction of best-fit galaxy templates corresponding to young starbursts is $`2/3`$. Generally speaking, a typical fraction between $`1/4`$ and $`1/3`$ of the quasar candidates is also compatible with young starbursts, and CNELGs are hardly selected as quasars, when supplementary galaxy templates are used. To summarize, adding star-bursting and early-type galaxy templates does not give a better solution to the problem of selecting quasar candidates: if star-burst candidates are included in the list, the number of candidates is multiplied by a factor of 4 to 5. On the other side, objects like CNELGs will not be mixed with the quasar candidates but we take the risk to miss $`35\%`$ of the quasars, identified as star-bursting galaxies. However, for the purposes of a survey, all objects will be tested through the same, general pipeline. Some objects (like quasars and compact blue galaxies) will inevitably end up as candidates in more than one list, depending on the different selection criteria. #### 5.1.1 Quasar Photometric Redshifts Figure 6 left presents the $`z_{phot}`$ versus $`z_{spec}`$ for the 51 DMS quasars. For 23 of them, $`|z_{phot}z_{spec}|0.1`$. We note an important dispersion for redshifts lower than typically 2.5: quasars with a $`z_{spec}[1,2]`$ are better fitted by templates of quasars with $`z[0,1]`$. This is also the tendency arising from simulations and shown on figure 3 (upper right), for the filter combination UBVRI. Figure 6 right presents the $`z_{phot}`$ versus $`z_{spec}`$ for a new sample of 46 spectroscopically identified quasars (Hall et al., 2000). 17 of them have $`|z_{phot}z_{spec}|0.1`$. This time we note that quasars with a $`z_{spec}[0,1]`$ are better fitted by quasars with $`z[1,2]`$, and this trend is also seen on the simulations (figure 3 upper right) for the same set of filters. ### 5.2 HDF<sub>Bright</sub> and Parkes The completeness has also been tested on two recently published quasar catalogues. Liu et al. (1999) published a list of 30 spectroscopically confirmed quasars with their photometry in B, V and R. These objects have $`17.6B21.0`$ and $`0.44z2.98`$, and were selected on a one square degree field centered on the HDF North. They were firstly identified as quasar candidates by their (B-V) (V-R) colors, along with 31 other candidates that, after the spectroscopy, turned out to be stars. With our method we identified all 30 quasars as quasar candidates. However, the limited number of filters did not allow us to reduce the number of stars selected as quasar candidates, 30 out of 31 were in our own list. We also applied our method to the Parkes subsample published by Francis et al. (1999). Only point-like objects, with the photometry in all 7 filters (B, V, R, I, J, H, K<sub>n</sub>) were kept, thus 129 out of the 157 quasars of the initial sample remain in our catalogue. At a confidence level of 99%, 107 among these objects were classified as quasar candidates (completeness=0.83). If no $`\chi ^2`$ selection is applied, the completeness is equal to 0.97. Columns 7 to 14 in Table 4 summarize these results. Simulation results are also given for comparison. Note that in both cases, the completeness for the real samples is found to be even higher than expected according to the simulations. Confirmation rate and efficiency cannot be computed for these two samples, since the entire point-like object catalogues (including stars) are not available. ### 5.3 EIS-wide (patch B) An attempt to apply our method to the EIS–wide (patch B) data was made. The results are approximatively the same as those obtained in multicolor procedures. This was somewhat expected: the multicolor approach proposed here, when applied to a 3 filter data set, is equivalent to a color–selection technique. Note that 3 filters have been used to carry out this survey: B, V and I. On a (B-V) (V-I) diagram, quasar evolutionary tracks up to a redshift of $`3.3`$ and the Main Sequence stars mix together. ## 6 Discussion The quasar multicolor selection technique presented in this paper has some advantages as compared to the “traditional” color-color methods. First, it has the advantage of selecting quasar candidates even at redshifts where quasar and stellar colors are very much alike (2.5 – 3.0). However, the efficiency varies with the redshift. Independently of the filter set, some quasars with redshifts lying in the interval \[2.5,3\] will be skipped. Furthermore, the number of quasar candidates with estimated $`z_{phot}`$ within this range will be far higher in comparison to other redshift bins. The set of filters used plays an important role in these results. Adding a filter is not necessarily equivalent to adding information and does not always improve the results, but imposes more stringent constraints by increasing the number of degrees of freedom. For this reason, a $`\chi ^2`$ selection does not improve the results in all cases. This technique gives also an estimate of the (photometric) redshift of the candidates, and it is thus able to improve the spectroscopic follow up of a given sample. In some cases, which depend (again) on the filter combination and the redshift of the objects, this determination can be accurate ($`|z_{phot}z_{spec}|0.1`$). In most cases, quasars at redshift $`z3`$ will not be misidentified at $`z4`$ and vice versa, therefore the spectroscopy can be reliably targeted towards the optical or the infrared bands. For the purposes of a survey aiming in assembling a quasar catalogue with a minimum cost, adding star-bursting and early-type galaxy templates does not appear as a priority. On the contrary, this addition will cause either the increase of the candidates by a factor of almost 5 (if star-burst candidates are considered as possible quasars) or the decrease of the completeness $`40\%`$ (because an important fraction of real quasars will be better fitted by star-burst galaxies templates). For practical purposes (e.g the VIRMOS survey), when all the sources examined through the same pipeline, a fraction of objects identified as quasars will be also found within the “starburst” sample, but this does not affect the present selection. The same method can be extended towards different directions. First, it can be adapted to particular objects like very red quasars, by adding the suited templates. However, this has to be done thoroughly, to prevent from a possible and undesirable increase of the quasar candidates. It can also be used on spatially resolved objects, like Seyfert galaxies, because by construction, the software is independent of the morphology of the objects. This can be done by adding a stellar component to the AGN simulated spectra already used. Last but not least, it can be extended to very high redshifts (higher than 6), after an appropriate modeling of the intergalactic absorption. As already mentioned, this software has been developed for the needs of preparing the quasars/AGN sample assembled in the future VIRMOS survey. However, the results and conclusions of this paper apply to all multicolor surveys in general. A cross-identification between optical and X-ray sources can be extremely useful, when X-ray data are available, and can increase the efficiency, as quasar candidates with a compact X-ray counterpart will have an even higher chance of being real quasars. ###### Acknowledgements. We would like to thank M. Bolzonella, J. P. Picat and M. Elvis for useful discussions. We would also like to thank the referee, Dr. P. Hall for his very useful comments that led to the improvement of our work.
warning/0001/nlin0001028.html
ar5iv
text
# 1 Introduction ## 1 Introduction The Kadomtsev-Petviashvili (KP) hierarchy of integrable soliton nonlinear evolution equations is among the most important physically relevant integrable systems. Quite recently a new class of integrable systems motivated by the Toda field theory appeared both in the mathematical and in the physical literature . On the other side the applications of the supersymmetry (SUSY) to the soliton theory provide us a possibility of the generalization of the integrable systems. The supersymmetric integrable equations \[3-14\] have drawn a lot of attention in recent years for a variety of reasons. In order to get a supersymmetric theory we have to add to a system of k bosonic equations kN fermions and k(N-1) boson fields (k=1,2,..N=1,2,..) in such a way that the final theory becomes SUSY invariant. Interestingly enough, the supersymmetrizations, leads to new effects (not present in the bosonic soliton’s theory). In this paper we describe a new method of the $`N=2`$ supersymmetrization of the KP hierarchy, first time presented for (M)KdV equation in . Our method contains the $`N=1`$ supersymmetric KdV equation presented by Beckers as a special case. Here we would like to present new results on the supercomplexified method. We construct the supersymmetric Lax operator for our supercomplex KP hierarchy. This operator however does not generate the supersymmetric conserved currents defined in the whole superspace. We describe therefore special procedure which allows us to obtain whole chain of conserved currents for our supercomplex hierarchy. The supercomplex Korteweg de Vries equation, considered as a special element of our KP hierarchy, constitue a bi-hamiltonian equation and is completely integrable. The second Hamiltonian operator of the supercomplexified KdV equation generates some odd $`N=2`$ SUSY algebra in contrast to the even algebra SUSY $`N=2`$ Virasoro algebra considered in \[8-10\]. This supercomplexification is a general method. In order to obtain the $`N=4`$ supercomplex version of soliton’s equation it is possible to use two nonequivalent methods. In both cases we obtain two different generalizations of $`N=4`$ KdV equation. The model proposed in this paper, differs from the one considered in . The idea of introducing an odd hamiltonian structure is not new. Leites noticed almost 20 years ago , that in the superspace one can consider both even and odd sympletic structures, with even and odd Poisson brackets respectively. The odd brackets (also known as antibrackets) have recently drawn some interest in the context of BRST formalism in the Lagrangian framework , in the supersymmetrical quantum mechanics , and in the classical mechanics . ## 2 Supercomplexification We shall consider an $`N=2`$ superspace with the space coordinates $`x`$ and the Grassman coordinates $`\theta _1,\theta _2,\theta _2\theta _1=\theta _1\theta _2,\theta _1^2=\theta _2^2=0`$. The supersymmetric covariant derivatives are defined by $$=\frac{}{x},D_1=\frac{}{\theta _1}+\theta _1,D_2=\frac{}{\theta _2}+\theta _2,$$ (1) with the properties $$D_1^2=D_2^2=,D_1D_2+D_2D_1=0.$$ (2) $$D_1^1:=D_1^1,D_2^1:=D_2^1.$$ (3) We define the integration over the $`N=2`$ superspace to be $$𝑑XH(x,\theta _1,\theta _2)=𝑑x𝑑\theta _1𝑑\theta _2H(x,\theta _1,\theta _2),$$ (4) where Berezin’s convention are assumed $$𝑑\theta _i\theta _j:=\delta _{i,j},𝑑\theta _i:=0.$$ (5) We always assume that the components of the superfileds and their derivatives vanish rapidly enough. Let us now consider some classical evolution equation in the form $$u_t=F(u,u_x,u_{xx},\mathrm{})=PgradH(u,u_x,u_{xx}..),$$ (6) where u is considered evolution function, P is the Poisson tensor, H is Hamiltonian of some dynamical system and $`grad`$ denotes the functional gradient. In order to get the $`N=2`$ supersymmetric version of the equation (6) let us consider the following ansatz which in the next we will call as supercomplexificiation $$u=(D_1D_2U)+iU_x,$$ (7) where now $`U`$ is some $`N=2`$ superfield and $`i`$ is imaginary quantity $`i^2=1`$. Introducing this ansatz to equation (6) we obtain that superfield $`U`$ evolves in the time as $$U_t=G(U,U_x,U_{xx},(D_1U),(D_2U),\mathrm{}.)=D_2^1D_1^1Re(F)=^1Im(F),$$ (8) where $`Re`$ and $`Im`$ denotes the real and imaginary part respectively. In order to guarantee validity of (8) we assume that $`F`$ is chosen in such a way that $`(D_1Re(F))=(D_2Im(F)),`$ (9) $`(D_2Re(F))=(D_1Im(F)),`$ (10) always holds. It is a condition of solvability of our construction. Examples. Let us consider the famous KdV equation $$u_t=u_{xxx}6u_xu.$$ (11) The supercomplex version of this equation is $$U_t=U_{xxx}6(D_1D_2U)U_x,$$ (12) and takes the following form in the components $`f_t`$ $`=`$ $`f_{xxx}+6gf_x,`$ (13) $`g_t`$ $`=`$ $`(g_{xx}+3g^23f_x^2),`$ (14) $`\xi _{1t}`$ $`=`$ $`\xi _{1xxx}+6\xi _{1x}g+6\xi _{2x}f_x,`$ (15) $`\xi _{2t}`$ $`=`$ $`\xi _{2xxx}6\xi _{1x}f_x+6\xi _{2x}g,`$ (16) where $$U=f+\theta _1\xi _1+\theta _2\xi _2+\theta _2\theta _1g.$$ (17) Notice that the bosonic part of the equations (13,14) does not contain any fermions fields. Hence in some sense this ”supersymmetrization is without supersymmetry”. Now assuming that $$U=\mathrm{\Pi }+\theta _2\mathrm{\Phi },$$ (18) where $`\mathrm{\Pi }`$ and $`\mathrm{\Phi }`$ are $`N=1`$ superbosonic and superfermionic functions respectively, our supercomplex KdV equation (12) reduces to $`\mathrm{\Pi }_t`$ $`=`$ $`\mathrm{\Pi }_{xxx}6\mathrm{\Pi }_x(D_1\mathrm{\Phi }),`$ (19) $`\mathrm{\Phi }_t`$ $`=`$ $`\mathrm{\Phi }_{xxx}+6\mathrm{\Pi }_x(D_1\mathrm{\Pi }_x)6\mathrm{\Phi }_x(D_1\mathrm{\Phi }).`$ (20) Notice that when $`\mathrm{\Pi }=0`$ then our equations (20) reduces to the $`N=1`$ supersymmetric KdV equation considered by Beckers in . Let us consider the supercomplexified version of the Nonlinear Schrődinger equation as a second example $`f_t`$ $`=`$ $`f_{xx}2gf^2,`$ (21) $`g_t`$ $`=`$ $`g_{xx}+2fg^2,`$ (22) as the second example. Assuming that $`f`$ $`=`$ $`(D_1D_2F)+iF_x,`$ (23) $`g`$ $`=`$ $`(D_1D_2G)+iG_x,`$ (24) we obtain $`F_t`$ $`=`$ $`F_{xx}2^1\left(2(D_1D_2G)(D_1D_2F)F_x+G_x(D_1D_2F)^2G_xF_x^2\right),`$ (25) $`G_t`$ $`=`$ $`G_{xx}+2^1\left(2(D_1D_2G)(D_1D_2F)G_x+F_x(D_1D_2G)^2F_xG_x^2\right).`$ (26) ## 3 N=2 Supercomplexified Kadomtsev-Petviashvili hierarchy It well known that Korteweg de Vries equation as well as the Nonlinear Schrődinger equation are the memebrs of the KP hierarchy. This hierarchy is defined by the following Lax operator $$Lax:=+f^1g,$$ (27) which generates the equations by the so called Lax pair representation $$\frac{d}{dt}Lax=[Lax,(Lax)_+^n],$$ (28) where $`n`$ is an arbitrary natural number and $`(+)`$ denotes the projection onto purely superdifferential part of the pseudosuperdifferential element. In order to define the Lax operator which generate our supercomplexified solitonic equations let us first define the supercomplexified algebra of pseudosuperdifferential elements $`\mathrm{{\rm Y}}`$ as a set of the following elements $$\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}((D_1D_2F_n)+F_{nx}^1D_1D_2)^n,$$ (29) where $`F_n`$ is an arbitrary $`N=2`$ superfield. There are three different projection in this algebra $$P_k(g)=\underset{n=k}{\overset{+\mathrm{}}{}}((D_1D_2F_n)+F_{nx}^1D_1D_2)^n,$$ (30) where $`g`$ is an arbitrary element belonging to $`\mathrm{{\rm Y}}`$ while $`k`$ can take three values $`k=0,1,2`$ only. Now in ordedr to obtain the supercomplexified version of the Lax operator (27) it is enought to replace $`f`$ and $`g`$ by the following substitution $`f(D_1D_2F)+F_xD_1D_2^1,`$ (31) $`g(D_1D_2G)+G_xD_1D_2^1,`$ (32) in (27) and replace the projection ($`+`$) by the projection $`P_1`$ defined by (30) with $`k=0`$ . As an example let us consider the Lax pair representation for the supercomplexified N=2 KdV equation. The Lax operator for this case is $$Lax:=+^1((D_1D_2F)+F_xD_1D_2^1),$$ (33) and produces the supercomplexified KdV equation $$\frac{d}{dt}Lax:=[Lax,P_1(Lax^3)],$$ (34) where $$P_1(Lax^3)=^3+3((D_1D_2F)+F_xD_1D_2^1).$$ (35) It is interesting to notice that this Lax operator does not produce any conserved currents defined in the whole $`N=2`$ superspace. The traditional supersymmetric residual definition as the coefficients standing in $`D_1D_2^1`$ in $`Lax^n`$ gives us that this coefficient after integration over whole superspace is zero. ## 4 Conserved currents for the supercomplexified KdV equation It is easy to prove using the symbolic computer computations \[20-21\] that this equation does not possesses any superbosonic currents. On the other side using the same techinque it is easy to find superfermionic conserved currents. Let us explain the connections of these currents with the usual (classical) currents of the KdV equation. This connection is achieved in four steps. First step. Let us supercomplexify an arbitrary conserved current of the classical KdV equation. By $`H_{nr}`$ let us denote the real part of n-th conserved current after the supercomplexifications. Second step. We compute the usual integral of $`H_{nr}`$. This can be denoted as $$H_{nr}𝑑x=K_n^0+K_n^1𝑑x.$$ (36) Third step. Now we compute the supersymmetrical integral over first supersymmetrical variable from $`K_n^1`$. It can be symbolicaly denoted as $$(D_1K_n^1)𝑑x=H_{1n}+(D_1S_n^1𝑑x).$$ (37) Fourth step. Finally we compute the supersymmetrical integral over second supersymmetrical variable which is denoted as $$(D_2S_n^1)𝑑x=H_{2n}.$$ (38) $`H_{1n}`$ and $`H_{2n}`$ are just the conserved superfermionic currents of the supercomplexified KdV equation. Let us presents several superfermionic conserved currents of the supercomplexified KdV equation $`H_{12}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x𝑑\theta _1𝑑\theta _2U(D_1U_x)},`$ (39) $`H_{22}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x𝑑\theta _1𝑑\theta _2U(D_2U_x)},`$ (40) $`H_{13}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x𝑑\theta _1𝑑\theta _2((D_2U_{xxx})U+4(D_2U)(D_1D_2U)U_x)},`$ (41) $`H_{23}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x𝑑\theta _1𝑑\theta _2((D_1U_{xxx})U+4(D_1U)(D_1D_2U)U_x)}.`$ (42) Interestingly these currents does not contain the classical currents of the KdV equation in the bosonic or fermionic parts. We have checked, using the symbolic computations \[20-21\] thatthis procedure could be aplied to the whole supercomplexified KP hierarchy as well. ## 5 Odd Virasoro - like algebra It is well known that Korteweg - de Vries equation constitute the so called bihamiltonian structure $$u_t=grad\left(\frac{1}{2}𝑑x(u_xu_x2u^3)\right)=Virgrad\left(\frac{1}{2}𝑑xu^2\right),$$ (43) where $$Vir:=^32u2u.$$ (44) The operator $``$ and $`Vir`$ are two Poissons tensors which constitue with two different hamiltonians the so called bihamiltonian structure. Tensor $`Vir`$ is connected with the Virasoro algebra, realized as the Poison bracket algebra, through the Fourier decomposition of the field $`u`$\[\]. Indeed introducing $$u:=\underset{k=\mathrm{}}{\overset{+\mathrm{}}{}}T_kexp(ikx)\frac{1}{4},$$ (45) where $`T_k`$ is some generator, we obtain that $$\{u(x),u(y)\}=Vir\delta (xy),$$ (46) are equivalent with the following Virasoro algebra. $$\{T_n,T_m\}=(n+m)T_{n+m}+\delta _{n+m,0}(n^3n).$$ (47) We successed to find the supercomplexified version of the bihamiltonian structure of the supercomplexified KdV equation. Let us define four operators $`P_{21}`$ $`:=`$ $`D_12^1(D_1D_2U)D_12(D_1D_2U)^1D_1`$ (49) $`+2^1U_xD_2+2U_x^1D_2,`$ $`P_{22}`$ $`:=`$ $`D_22^1(D_1D_2U)D_22(D_1D_2U)^1D_2`$ (51) $`2^1U_xD_12U_x^1D_1,`$ $`P_{11}`$ $`:=`$ $`D_1^1,`$ (52) $`P_{12}`$ $`:=`$ $`D_2^1.`$ (53) These operators generate the supercomplexified KdV equation (12) $$U_t:=P_{21}\frac{\delta H_{12}}{\delta U}=P_{22}\frac{\delta H_{22}}{\delta U}=P_{11}\frac{\delta H_{23}}{\delta U}=P_{12}\frac{\delta H_{13}}{\delta U}$$ (54) where hamiltonians are defdined by (36-39). We can construct an $`O(2)`$ invariant bihamiltonian structure considering the linear combination of $`P_{11}\pm P_{12},P_{21}\pm P_{22}`$ with $`H_{12}\pm H_{22},H_{14}\pm H_{24}`$ respectively. These structures define the same supercomplexified KdV equation (12). Notice that the operators $`P_{21},P_{22}`$ or $`P_{21}\pm P_{22}`$ play the same role as the Virasoro algebra in the usual KdV equation. There is a basic difference - our Hamiltonian operators generates the odd Poisson brackets in the odd superspace. In order to obtain the explicit realization of this algebra we connect the Hamiltonian operator $`P_{21}P_{22}`$ with the Poisson bracket $$\{U(x,\theta _1,\theta _2),U(y,\theta _1^{^{}},\theta _2^{^{}}\}=(P_{21}P_{22})(\theta _1\theta _1^{^{}})(\theta _2\theta _2^{^{}})\delta (xy),$$ (55) where $$U(x,\theta _1,\theta _2)=u_0+\theta _1\xi _1+\theta _2\xi _2+\theta _2\theta _1u_1.$$ (56) Introducing the Fourier decomposition of $`u_0,\xi _1,\xi _2,u_1`$ $$\xi _j:=\underset{s=\mathrm{}}{\overset{\mathrm{}}{}}G_s^je^{isx},j:=1,2,$$ (57) $$u_0:=i\underset{s=\mathrm{}}{\overset{\mathrm{}}{}}L_se^{isx},u_1:=\underset{s=\mathrm{}}{\overset{\mathrm{}}{}}T_se^{isx}\frac{1}{4},$$ (58) in (64) we obtain $$\{T_n,T_m\}=\{L_n,L_m\}=\{L_n,T_m\}=0,$$ (59) $`\{T_n,G_m^i\}`$ $`=`$ $`(n^21)\delta _{n+m,0}+(1)^i2{\displaystyle \frac{nm}{m}}T_{n+m}2{\displaystyle \frac{n^2m^2}{m}}L_{n+m},`$ (60) $`\{G_n^i,L_m\}`$ $`=`$ $`(n{\displaystyle \frac{1}{n}})\delta _{n+m,0}+2{\displaystyle \frac{mn}{nm}}T_{n+m}+(1)^i2{\displaystyle \frac{m^2n^2}{nm}}L_{n+m},`$ (61) $`\{G_n^i,G_m^i\}`$ $`=`$ $`(1)^i2{\displaystyle \frac{m^2n^2}{nm}}\left(G_{n+m}^1+G_{n+m}^2\right),`$ (62) $`\{G_n^1,G_m^2\}`$ $`=`$ $`2{\displaystyle \frac{m^2n^2}{nm}}\left(G_{n+m}^1G_{n+m}^2\right).`$ (63) These formulae define the closed algebra with the graded Jacobi identity $$\underset{cycl(a,b,c)}{}(1)^{[1+a][1+c]}\{a,\{b,c\}\}=0,$$ (64) where $`[a]`$ denotes the parity of $`a`$. It is the desired odd Virasoro - like algebra. ## 6 Supercomplexified N=4 KdV equation It is possible to obtain in two different manner the supercomplexified N=4 version of the soliton equation. In the first manner we strictly speaking make duble supercomplexifications of the usual solitons equations while in the second method we supercomplexify the well know N=2 supersymmetric equations. Let us describe these methods. In the first method we assume that the superfunction $`U`$ which satisfy the equation (12) could be presented as $$U:=(D_3D_4W)+iW_x.$$ (65) where now $`W`$ is some $`N=4`$ bosonic superfunction. Introducing this ansatz to (12) we obtain that $`W`$ satisfy the dubly supercomplexified KdV equation $$W_t:=W_{xxx}6^1((D^4W)W_{xx}+(D_1D_2W)(D_3D_4W)),$$ (66) where $`D^4=D_1D_2D_3D_4`$. For the second method we consider the $`N=2`$ supersymmetric KdV equation $$\mathrm{\Phi }_t:=(\mathrm{\Phi }_{xx}+3\mathrm{\Phi }(D_1D_2\mathrm{\Phi })+\frac{1}{2}(\alpha 1)(D_1D_2\mathrm{\Phi }^2)+\alpha \mathrm{\Phi }^3).$$ (67) where $`\alpha `$ can take three values $`2,1,4`$ if we would like to consider the integrable extensions. We assume that the superfield $`\mathrm{\Phi }`$ satisfy the $`N=2,\alpha =4`$ SUSY KdV equation (62), takes after the supercomplexification the following form $$\mathrm{\Phi }:=(D_3D_4\mathrm{{\rm Y}})+i\mathrm{{\rm Y}}_x,$$ (68) where $`\mathrm{{\rm Y}}`$ is some $`N=4`$ superboson field. Substituting this form in (65) we obtain $`\mathrm{{\rm Y}}_t`$ $`:=`$ $`\mathrm{{\rm Y}}_{xxx}+3(D_1D_2(\mathrm{{\rm Y}}_x(D_3D_4\mathrm{{\rm Y}})))4\mathrm{{\rm Y}}_x^3+`$ (70) $`\left(\mathrm{{\rm Y}}_x(D^4\mathrm{{\rm Y}})+(D_3D_4\mathrm{{\rm Y}})(D_1D_2\mathrm{{\rm Y}}_x)\right)+12(D_3D_4\mathrm{{\rm Y}})^2\mathrm{{\rm Y}}_x.`$ It is the desired generalization of the $`N=4`$ SUSY KdV equation, which is different from the one considered in . Let us remark that usuing the similar methods descrbided earlier it is possible to obtain the supercomplexification of the bihamiltonians structures, the Lax pair representations and the conserved currents. As an example let me present the Lax operator for the equations (64) and (67). For equation (64) the Lax operator is $$L:=^2+(D^4W)+(D_1D_2W_x)i_2+(D_3D_4W_x)i_1+W_{xx}i_1i_2$$ (71) where $`i_1=^1D_1D_2`$, $`i_2=^1D_3D_4`$ and $`i_1^2=i_2^2=1`$. For equation (67) the Lax operator is $$L:=\left(D_1D_2+(D_3D_4\mathrm{{\rm Y}})+\mathrm{{\rm Y}}_x^1D_3D_4\right)^2$$ (72) Acknowledgement The author wish to thank organizer for the scientific discussion and warm hospitability during the conference. This paper has been supported by the KBN grant 2 P03B 136 16 and by Bogoliubow-Infeld found.
warning/0001/astro-ph0001064.html
ar5iv
text
# Photometric modal discrimination in 𝛿 Scuti and 𝛾 Doradus stars ## 1. Introduction In recent years the development of photometric multi-site campaigns organized by different teams: Delta Scuti Network (http://dsn.astro.univie.ac.at), STEPHI (http://dasgal.obspmfr/stephi/) and STACC (Frandsen et al. 1996) for studying multi-periodic $`\delta `$ Scuti stars has increased the number of detected frequencies up to a level at which we could think that direct fitting to a theoretical model could be relevant to test the modelization and finally to do real asteroseismology of these stars. See for instance Breger et al. (1999) for FG Vir and Pamyatnykh et al. (1998) for XX Pyx. However the number of theoretically excited and photometrically visible radial and non-radial periods in a $`\delta `$ Scuti is so large that it is impossible to match all theoretical frequencies to the observed ones. For this reason a method to identify modes is needed. One can think that the rotational splitting is the most simple method and to look for singlets (radial, $`l=0`$), triplets (dipole, $`l=1`$), etc., in the frequency spectrum. However the overlapping and non-linear interactions among them destroy the expected regular pattern as demonstrated in Breger et al (1999). In any case some residual regularity can be expected and used as an indication of the degree of the spherical harmonic $`l`$ of the observed modes, as explained there. Therefore identification of at least a few frequencies becomes necessary to attempt to make asteroseismological techniques available for $`\delta `$ Scuti stars. Another approach, better explained in other reviews of this conference, consists of the use of the line profile variations to model the surface velocity field and then deduce the corresponding spherical harmonic. Technical details are given elsewhere, Mantegazza in these proceedings and Aerts these proceedings as well, but I would like to note here the difficulty to get sufficient spectroscopic measurements, since both high S/N ratio and time coverage to resolve frequencies are required, so that medium size telescopes are needed. In fact it is quite difficult to organize multi-site spectroscopic campaigns for many days, but easier when the same is done for small photometric telescopes. However the enormous advantage of the spectroscopy is that one can obtain information not only on the $`l`$-values but also on the azimuthal order of the spherical harmonic $`m`$-values, which are not detectable from purely photometric observations. The main question we want to address here is whether multi-band photometry alone can be useful to discriminate $`l`$-values or not. Based on the linearization made by Dziembowski (1977) of the bolometric magnitude variations exhibited by a star undergoing non-radial oscillations, Balona & Stobie (1979a, 1979b) formulated an analytic expression which will be used throughout this review in the form given by Watson (1988). The equation, as we will see later on, contains evaluations of flux derivatives and of integrals over limb darkening coefficients. This requires a well-behaved model atmospheres not only describing mean values, from which we can deduce physical parameters, but also their variations as far as temperature and/or gravity varies. I will start by introducing the linearized equation and the physical conditions required for applicability, then I will discus the limitations of the available model atmospheres, basically Kurucz models and modifications, concerning partial derivatives and limb darkening coefficients. As we will see, there are two unknown quantities in the formula, the phase lag and a parameter related with departures from adiabaticity, which deserves a detailed discussion given in section 3. The next section is devoted to the application to real data mainly through the use of Strömgren photometry. Finally in the last section, I will present the expected possibilities of the method in the context of future space missions. ## 2. The linearized equation Thermal time scales for $`\delta `$ Scuti stars are at least an order of magnitude larger than the shortest observed period. Furthermore the relative radius variations, even for the High Amplitude $`\delta `$ Scuti Stars (HADS), are so small that second order terms can be neglected. Under such conditions the linearized equation governing the photometric variation of a pulsating star, $`\delta x(t)`$, can be expressed, following the formulation given in Watson (1988), as: $`\delta x(t)`$ $`=`$ $`1.086ϵP_{l,|m|}(\mathrm{cos}(\theta ))\times `$ (1) $`[(T_1+T_2)\mathrm{cos}(\omega t+\mathrm{\Psi }^T)+(T_3+T_4+T_5)\mathrm{cos}(\omega t)]`$ where the variables have the following definitions: * the $`x`$’s stand for the different photometric bands. Here I will use $`u,v,b`$ and $`y`$ of the Strömgren photometric system. * $`ϵ1`$ is an arbitrary small quantity. * $`P_{l,|m|}(\mathrm{cos}(\theta ))`$ is the associated Legendre polynomial of order $`l,m`$ to the pulsation frequency $`\omega `$, oriented with the angle $`\theta `$, i. e. the inclination of the stellar pulsation axis to the observer. * $`\mathrm{\Psi }^T`$ is the Phase Lag, i.e. the angle between maximum temperature and minimum radius, hence $`\mathrm{\Psi }^T=\pi `$ in perfect adiabatic conditions. * The $`T_1\mathrm{}T_5`$ can be written as follows: $`T_1`$ $`=`$ $`b_{l,x}B{\displaystyle \frac{x(t)}{\mathrm{log}T}},`$ (2) $`T_2`$ $`=`$ $`{\displaystyle \frac{B}{2.3026}}{\displaystyle \frac{b_{l,x}}{\mathrm{log}T}},`$ (3) $`T_3`$ $`=`$ $`b_{l,x}(2+l)(1l),`$ (4) $`T_4`$ $`=`$ $`b_{l,x}p^{}C{\displaystyle \frac{x(t)}{\mathrm{log}g}},`$ (5) $`T_5`$ $`=`$ $`{\displaystyle \frac{p^{}C}{2.3026}}{\displaystyle \frac{b_{l,x}}{\mathrm{log}g}}`$ (6) where $`b_{l,x}`$ are the weighted limb darkening integrals defined as: $`b_{l,x}`$ $`=`$ $`{\displaystyle _0^1}h_\lambda (\mu )\mu P_l(\mu )𝑑\mu `$ (7) where, using a quadratic limb darkening law, $`h_\lambda (\mu )`$ $`=`$ $`\chi __0+\chi __1\mu +\chi __2\mu ^2`$ (8) Partial derivatives and limb darkening coefficients have to be tabulated from theoretical model atmospheres. $`p^{}`$ is a measure of the variation of the pressure when a gravity variation occurs, evaluated at an optical layer where we observe the continuum flux. It is then $`p^{}=\left({\displaystyle \frac{\mathrm{log}g}{\mathrm{log}p}}\right)_{_{\tau =1}}`$ (9) From model atmospheres and in the $`\delta `$ Scuti regime this value can be considered as a constant equal to approximately 1.4. $`B`$ and $`C`$ are related by the equation: $`B=R\left(11/\mathrm{\Gamma }_2\right)C`$ (10) where $`C`$ is given by: $`C=(4+1/\alpha __H)l(l+1)\alpha __H`$ (11) and $`\alpha __H=G\rho _{_{}}Q^2/3\pi .`$ (12) $`Q`$ is calibrated, see for instance Breger et al. (1993), using the following equation: $`\mathrm{log}Q=6.456+\mathrm{log}P+0.5\mathrm{log}g+0.1M_{\mathrm{bol}}+\mathrm{log}T_{\mathrm{eff}}.`$ (13) Standard photometric indices are then used to calibrate gravity, absolute magnitude and temperature and, by knowing the pulsation period, we can obtain the observational value for the pulsation constant $`Q`$. $`R`$ is introduced as a free parameter to estimate deviations from adiabaticity, since for an adiabatic atmosphere: $`B=(11/\mathrm{\Gamma }_2)C`$ (14) then $`R`$ must be unity in that case and physical values for $`R`$ must be between $`0`$ and $`1`$. The adiabatic exponent $`\mathrm{\Gamma }_2`$ can show dramatic changes in the atmosphere of a pulsating star but in our case we assume a constant value of $`5/3`$. So any change of this parameter will change the value of $`R`$, which will remain as an unknown parameter in equation (1). The general procedure described here to obtain $`l`$-values from observed photometric indices is explained in the flow chart of Fig 1. It is based on the use of model atmospheres to calculate not only global physical quantities for a given star but also variations of the photometric indices with respect to temperature and gravity, together with the above defined integral of the limb darkening coefficients. However, and in order to know the corresponding $`l`$-value, we need to have an estimation of the two unknown variables $`\mathrm{\Psi }^T`$ and $`R`$. Before doing that in the next section we will discuss more in detail the precision of these quantities when model atmospheres from Kurucz (1993), i.e. ATLAS9 code, are used. Starting by the observables $`x`$, the calibrated photometric indices, $`P`$, the pulsation period and the time variations of the photometric indices $`x`$, $`\delta x`$ , which include phases and amplitudes, we are able to obtain for the $`\delta `$ Scuti and related stars the following: 1. Effective temperatures, gravity and metallicity from model atmospheres 2. Limb darkening coefficients and partial derivatives, also from model atmospheres 3. Pulsation constant from equation (13) and the previously calibrated quantities There are three quantities which remain unknown from equation (1): the phase lag, $`\mathrm{\Psi }^T`$, the adiabaticity parameter, $`R`$, and the $`l`$-value we are searching for. At this point two strategies exist: * to make an assumption with physical sense about $`\mathrm{\Psi }^T`$ and $`R`$. Usually $`R`$ is selected in the range $`0.25<R<1`$ only from theoretical considerations whereas $`\mathrm{\Psi }^T`$ is selected in the range of $`90{}_{}{}^{}<\mathrm{\Psi }^T<135^{}`$ from spectroscopic observations of $`\delta `$ Scuti stars. One can draw then the corresponding “regions of interest” for the different $`l`$-values. A comparison of the phase differences vs. amplitude ratios with observations is the approach followed by different authors (Watson, 1988 and Garrido et al., 1990). * to estimate $`\mathrm{\Psi }^T`$ and $`R`$ either from theoretical calculations (Balona and Evers, 1999) or by photometric observations as explained in detail in Garrido et al. (1990). In both cases the final estimation of the $`l`$-value is made by selecting, through a least squares procedure, the minimum distance between predictions and observations as we will see in section 5.2. Following the flow chart in Fig 1 we will next discuss in detail a practical application by using Strömgren data and Kurucz models. ## 3. Kurucz model atmospheres ### 3.1. Physical calibration The Strömgren system is a photometric system composed of 4 filters of an intermediate width centered at $`u:3500`$ Å, $`v:4100`$ Å, $`b:4670`$ Å and $`y:5470`$ Å. They are very well adapted to derive physical parameters for the spectral types we are dealing with: $`\delta `$ Scuti and related stars. A direct comparison with the Kurucz models, without any correction, of the standard stars as defined in Smalley and Kupka (1997) gives the following results for the 6 primary standard stars: Standard ATLAS9 $`T_{\mathrm{eff}}(\mathrm{model})T_{\mathrm{eff}}(\mathrm{observed})=88\pm 120`$ K Standard ATLAS9 $`\mathrm{log}g(\mathrm{model})\mathrm{log}g(\mathrm{observed})=0.23\pm 0.18`$ and for the 15 IRFM standard stars: Standard ATLAS9 $`T_{\mathrm{eff}}(\mathrm{model})T_{\mathrm{eff}}(\mathrm{observed})=164\pm 131`$ K Standard ATLAS9 $`\mathrm{log}g(\mathrm{model})\mathrm{log}g(\mathrm{observed})=0.30\pm 0.34`$ Using the modification indicated by Smalley and Kupka (1997) which consists basically in the use a new treatment of the turbulent convection developed by Canuto and Mazzitelli (1991, 1992), one finds for the 6 primary standard stars: Modified ATLAS9 $`T_{\mathrm{eff}}(\mathrm{model})T_{\mathrm{eff}}(\mathrm{observed})=75\pm 115`$ K Modified ATLAS9 $`\mathrm{log}g(\mathrm{model})\mathrm{log}g(\mathrm{observed})=0.02\pm 0.05`$ and for the 15 IRFM standards stars: Modified ATLAS9 $`T_{\mathrm{eff}}(\mathrm{model})T_{\mathrm{eff}}(\mathrm{observed})=109\pm 118`$ K Modified ATLAS9 $`\mathrm{log}g(\mathrm{model})\mathrm{log}g(\mathrm{observed})=0.10\pm 0.12`$ The small difference is due to different assumptions for the physical atmospheric values for Vega: Smalley and Kupka (1997) using spectrophotometry from Hayes (1985), obtain $`T_{\mathrm{eff}}`$=9550 K, $`\mathrm{log}g`$=3.95, $`[M/H]`$=-0.5 and a micro-turbulence of $`2.0`$ km s<sup>-1</sup> whereas Kurucz uses that of Hayes and Latham (1975) obtaining: $`T_{\mathrm{eff}}`$=9400 K, $`\mathrm{log}g`$=3.90, \[M/H\]=-0.5 and a micro-turbulence of 0.0 km s<sup>-1</sup>. However the Smalley and Kupka (1997) calibration seems to be better concerning gravities, where they obtain a lower dispersion. In any case, and for our purposes, we can establish that the best temperature calibration we can have is around 100 K in error, and the best gravity is good within around 0.1 dex in $`\mathrm{log}g`$. Metallicity as measured with the index $`m_0`$ can be calibrated using the procedure given in Smalley (1993) to which we refer for a detailed discussion. In any case the discrepancy shown by the models introduces an uncertainty of 0.5 in \[M/H\] which is not dramatic as compared to other uncertainties related with the derivatives as we will see in the following subsection. ### 3.2. Partial derivatives Once we have a calibrated value for $`T_{\mathrm{eff}}`$, $`\mathrm{log}g`$ and \[M/H\] we can select a grid and perform the needed partial derivatives with respect to $`T_{\mathrm{eff}}`$ and $`\mathrm{log}g`$ in equation (1). These grids are sampled at 500 K in $`T_{\mathrm{eff}}`$ and 0.5 dex in $`\mathrm{log}g`$. Cubic splines subroutines have been used to calculate them. In Fig 2 the partial derivatives of the original Pop I $`[M/H]=0`$ Kurucz models with respect to $`T_{\mathrm{eff}}`$ are plotted, models are with overshooting and $`[l/H]=1.25`$. They are marked with full lines, and those of the Kurucz models modified by Smalley and Kupka (1997; hereafter S&K), are marked with dots. The main difference is that S&K models are calculated with a different convection treatment, as explained in detail in their paper. Ranges in $`T_{\mathrm{eff}}`$ and $`\mathrm{log}g`$ are those for $`\delta `$ Scuti and related stars. They are clearly decreasing functions with increasing temperature for $`b,v`$ and $`y`$ filters and show a not very pronounced minimum in the ultraviolet $`u`$ band. There is also a general trend in the sense that blue bands are noisier than red ones, independent of the models used. Models are essentially the same for the $`y`$ visible band; however the modified S&K models are smoother than original Kurucz ones. For the $`u`$ band this effect is magnified but, in any case, at temperatures higher than 7500 K, even S&K models start to show some small discontinuities. Uncertainties in these derivatives can reach up to $`20\%`$ in mag/$`\mathrm{log}T`$ units, depending on the model and/or the ($`T_{\mathrm{eff}}`$, $`\mathrm{log}g`$) regime that we are selecting. There is basically no differences when a non solar metallicity is used (see Fig 3) but, as expected, partial derivatives of the blue bands are lower than for solar metallicity: a metal deficiency of the order of $`[M/H]=1`$ makes a significant contribution to the flux depletion at these wavelengths and consequently a lower dependence. Concerning the partial derivatives with respect to $`\mathrm{log}g`$ the behavior is very similar in the sense that they are noisier for blue colors and for the original Kurucz models, as can be seen in Fig 4. Furthermore they are all almost constant over the ($`T_{\mathrm{eff}}`$, $`\mathrm{log}g`$) range shown in the figure. However the amplitude of variations of these derivatives for the original Kurucz models can reach values of the order of $`100\%`$ for some temperatures and gravities! Here again the S&K modified models behave better, always in the sense that these are smoother than the original Kurucz ones. One can think that these uncertainties in the partial derivatives would make the method useless but we will see in the next section that the relative importance of each term in equation (1) enables some possibilities for discriminating $`l`$-values depending on the filters and physical conditions even considering these uncertainties. ### 3.3. Limb darkening integrals and derivatives Limb darkening integrals defined in (7) with quadratic limb darkening laws for the Strömgren photometric bands $`uvby`$ have been calculated using the coefficients given in Table II, page 262, of Watson (1988). Here the step in temperature is 250 K and I only show values for 3 different gravities, 3.5, 4 and 4.5 dex. Integrals corresponding to $`l=1,2,3`$ and$`4`$ for a solar metallicity atmosphere are shown in Fig 5, 6, 7 and 8. Radial values, $`l=0`$ are exactly unity from the definition given in (7). They are slowly decreasing functions as temperature increases and with a not very well defined trend for gravities ranging from 3.5 to 4.5 dex. It is also to be noted that there is a slight wavelength dependence which becomes more important for high $`l`$-values. This will be crucial for the method developed originally by Garrido et al. (1990) and explained in section 5. One important feature of these plots are the discontinuities appearing at around 7000–7500 K in the original Kurucz models. This effect, which can be also seen in Fig 5 of the Watson (1988) review, is also independent of the wavelength and describes some inconsistency in these models. I selected, for comparison purposes, the new models developed in the PHOENIX code (see Hauschildt 1992, 1993; Hauschildt and Baron 1995; Baron et al. 1996 for a detailed description). In these models no discontinuity is seen although some inconsistencies still persist, in particular concerning the gravity variation, (see for instance the difference between $`\mathrm{log}g=3.5`$ and the other two derivatives at $`\mathrm{log}g=4,4.5`$ in the $`u`$ band for $`l=1,2`$ or the small deviations at high temperatures depending on the gravity). Discontinuities appearing in the Kurucz models at 7000–7500 K seem to be produced by the arbitrary suppression of overshooting in these models. In any case PHOENIX does not use overshooting and the effect is not present. The effect is extraordinarily enhanced when we calculate derivatives, shown in Fig 9 and 10, which have also to be calculated as indicated in equation (1). It is important to note here that the inclusion of the limb darkening integrals variations into that equation is irrelevant if we use the standard original Kurucz models, since uncertainties are of the same order as the derivatives. The situation is improved when one use the PHOENIX models but even so there are some regions (high temperature regions, double value for the partial derivatives in the $`u`$ band for different gravities, …) where these uncertainties still remain. At first sight these large inconsistencies in the derivatives and in the limb darkening integrals and their derivatives seem to introduce large uncertainties in equation (1), but the situation is improved when we realize the different contributions from different terms in that equation. In Fig 11 I show these different contributions for a typical $`\delta `$ Scuti regime: ($`\mathrm{log}T_{\mathrm{eff}}=7500`$ K, $`\mathrm{log}g=4,[M/H]=0`$), and with a mean value for the adiabaticity parameter, i.e. $`R`$=0.5. The different filters and filter combinations are the classical ones for the Strömgren photometric system, i.e. $`(by)`$ as a temperature indicator and $`c`$ as a luminosity indicator (at least for stars close enough to the Main Sequence). As expected, the main contribution to any band and $`l`$-value comes from temperature variations. Geometry variations, given by the term defined in (4), is the second important parameter for low $`l`$-values but for $`l`$=3 and 4 can become as large as, or even larger than, the temperature term defined in (2). It is also to be noted the large effect of the gravity variations in the color index $`c`$, as properly defined by the Strömgren photometry. The basic idea is to use the geometry dependence, through the term defined in (4), for $`l`$=0, 1 and 2 because of the changing sign of this term when passing from $`l`$=0 to $`l`$=2 and the null contribution for $`l`$=1. So the main utility is not basically changed by the uncertainties of the other terms in equation (1) so allowing, for the lowest $`l`$-values, a proper discrimination as we will see later on. In any case the term defined by (6) is always very small and the term defined by (3), no very well known neither, is only important for $`l`$=4 when the smearing out effects along the stellar surface make the photometric method inapplicable. ## 4. “Regions of interest” Once $`T_{\mathrm{eff}},\mathrm{log}g,[M/H]`$, derivatives and limb darkening have been calculated for a given star two unknown quantities in equation (1), $`(\mathrm{\Psi }^T,R)`$, still remain in order to obtain the quantity we are looking for: $`l`$, the degree of the spherical surface harmonic. $`(\mathrm{\Psi }^T`$ is not a very well known quantity but it can be estimated from simultaneous photometric and radial velocities observations. Typical values for some $`\delta `$ Scuti given in Breger et al. (1976) range from $`90^{}`$ to $`140^{}`$ but phase lags for $`\gamma `$ Dor stars are completely unknown. On the other hand the adiabaticity parameter $`R`$ can not be known from pure observations and one must assume some reasonable value for it which, according to equation (11), is restricted to values between 0 and 1. In what follows I will adopt a constant value of 1.4 for $`p^{}`$ and 5/3 for $`\mathrm{\Gamma }_2`$, which represent very well the $`\delta `$ Scuti regime. When a range for the phase lag and for $`R`$ is assumed a diagram such as that shown in Fig 12 can be drawn, where amplitude ratios and phase differences are calculated for the range $`90{}_{}{}^{}\mathrm{\Psi }^T140^{}`$ and $`0.25R1`$ and for the discriminant color pairs formed by $`y`$ and $`b`$ bands. In this review I will give the most usual combination of a visual band, $`y`$, and a temperature indicator, $`(by)`$, but other combinations are also useful, e.g. $`v`$ and $`y`$ as shown in Garrido et al. (1990). Uncertainties on the borders of these zones can be of the order of $`20\%`$ because of the actual precision of model atmospheres previously discussed. As can be seen in Fig 12 and originally demonstrated by Stamford and Watson (1981), the regions of radial modes are clearly separated from the regions of non-radial ones which in turn can be even distinguished among them. The reason is that, in the $`\delta `$ Scuti regime, where the phase lags are close to $`100^{}`$, a clear phase shift is originated for different photometric bands for the lowest $`l`$-values, as previously discussed. For $`l`$=3 the amplitude ratio is very different from the other three $`l`$-values and this can be used to discriminate it. These phase differences are however very small and we need a very high precision photometry when determining phases and amplitudes from a classical Fourier fitting to the time series. On the other hand we do not know the phase lags for the $`\gamma `$ Dor stars and hence these diagrams are not very useful for them until reliable values for these stars are known. At the end of this review I will give a explanation on how to calculate $`\mathrm{\Psi }^T`$ and $`R`$ values from multi-band photometric observations and we will see that the values deduced for the phase lags for these variables are close to $`0^{}`$. The effect of using different model atmospheres is not relevant, discrimination is also attained in a clear way and the only remarkable difference is the location of the photometric temperature indicator $`(by)`$, indicating a slight different temperature calibration for these two model atmospheres. However the effect of the pulsation constant, is dramatic. When this value is lowered to a value of 0.015 days, corresponding to a radial 3rd or 4th overtone, then negative values are predicted also for radial modes and it is not true anymore that any negative value, in a $`(by)`$ vs. $`y`$ plot, corresponds to a non-radial mode. An example is given in Fig 13. When we go to higher temperatures the radial regions at positive phase differences begin to shrink and finally disappear at around 8500 K even for a radial value for $`Q`$. An intermediate case is plotted in Fig 14 where the temperature is 8000 K and a higher radial value is assumed for $`Q`$. Notice the difference in amplitude ratios with respect to previous figures for the case $`l`$=3 which could be due to the effect of the temperature derivatives in the model atmospheres. In conclusion, these “regions of interest” are very dependent on the chosen photometric bands and used model atmospheres. In particular, model atmospheres with smoother derivatives and without discontinuities are required. A good test for these model atmospheres would be to compare the observed photometric variations with theoretical variations induced by a change in temperature. In other words not only to use standard stars for calibrating photometric indices but to use pulsating stars, with the above explained cautions, as standard stars regarding photometric variations. $`l`$-values deduced from these considerations should be taken with caution given the errors of the photometric measurements and the inconsistencies of model atmospheres. However there is a way to alleviate this problem: to know the quantities $`(\mathrm{\Psi }^T,R)`$ and the to solve equation (1) for $`l`$. ## 5. Practical calculation of $`(\mathrm{\Psi }^T,R)`$ I review in this section the two different methods to calculate actual values for the unknown quantities $`(\mathrm{\Psi }^T,R)`$: A theoretical one given recently by Balona and Evers (1999) and another observational one given in Garrido et al. (1990). ### 5.1. Theoretical approach Balona and Evers (1999) derive, from theoretical considerations, the unknown quantities $`(\mathrm{\Psi }^T,R)`$. Although they use $`f`$ instead of $`R`$ those parameters are related through the equation: $`f`$ $`=`$ $`4RC{\displaystyle \frac{\mathrm{\Gamma }_21}{\mathrm{\Gamma }_2}}`$ (15) Unfortunately these parameters are very sensible to the treatment of the convection specially for cool $`\delta `$ Scuti models, as shown by the authors in their Fig 3. The onset of the convection occurs at $`T_{\mathrm{eff}}7900`$ K; for higher temperatures the models are radiative and $`(\mathrm{\Psi }^T,f)`$ are constant and independent of the mixing length parameter adopted. The authors develop a procedure which minimizes the distance between the observed and theoretically predicted color phases and amplitudes. These theoretical predictions come from non-adiabatic pulsation calculations of the couple $`(\mathrm{\Psi }^T,f)`$ from a model which also predicts the observed frequency of the pulsating star. They give some identification for the best observed stars in the Strömgren multicolor photometric system in their Table 4. They apply the method to three double mode $`\delta `$ Scuti stars with period ratios typical of radial pulsators, i.e. AE UMa: 0.773, BP Peg: 0.772 and RV Ari: 0.773, and they find good solutions in general but they note that discrimination between $`l`$=0 and $`l`$=1 is sometimes poor. For some other HADS they find no solutions and for another they have to select a $`g`$-mode as the best solution. The method can be limited by our ignorance of the convection but in principle could be very useful for radiative models, i.e. the hot regime of the $`\delta `$ Scuti stars. ### 5.2. Observational approach This method was developed by Garrido et al. (1990) and uses multicolor photometry to derive the three unknown parameters $`(\mathrm{\Psi }^T,R)`$ and $`l`$ in equation (1). At first sight a direct fitting to the formula appears to give very unstable values since its complexity prevents an inverse solution but, as indicated in that paper, if we assume no dependence on $`\lambda `$ for the limb darkening integrals, which is nearly true for the lowest $`l`$-values (see Fig 5, 6, 7 and 8), equation (1) becomes reversible for any $`l`$-value. Errors are of the order of $`1\%`$ for $`l`$=1, $`6\%`$ for $`l`$=2, $`18\%`$ for $`l`$=3 and $`30\%`$ for $`l`$=4. In particular we need at least three bands, in order to have at least two color indices, but in practice we have the four Strömgren bands so allowing to calculate 15 different values for the couple $`(\mathrm{\Psi }^T,R)`$. Couples of values for $`(\mathrm{\Psi }^T,R)`$ are plotted in Fig 15, for the $`\delta `$ Scuti stars, where the error bars refers to 1-sigma value for the 15 values calculated for each star of Table 1. Errors bars for the other values not plotted in Fig 15 are of the same order and are not shown for clarity. All the $`R`$ values fall in the range of $`0.25R1`$, which is just the expected range from theoretical arguments given before. Also $`80{}_{}{}^{}\mathrm{\Psi }^T180^{}`$ as indicated by the existing simultaneous radial velocity observations. Values based on Kurucz models are more concentrated than values for the S&K models and the reason is not clear. In order to see the effect of the large variations shown by the $`u`$ derivatives in Fig 2, 3 and 4 I plot in them an average of these derivatives over temperature and gravity. The result is a more concentrated range for the $`(\mathrm{\Psi }^T,R)`$ values, so indicating that the dispersion shown by the derivatives could be a source of noise for calculating these parameters, because of the non-smoothness of model atmospheres. In Fig 16 the couples $`(\mathrm{\Psi }^T,R)`$ are plotted for the $`\gamma `$ Dor stars. These values are more spread basically for two reason: the small amplitude of the variations for these stars and the small phase differences observed for the photometric indices. Some physical unfounded $`R`$ values, i.e. greater than unity, appear for some stars. In any case the values derived using the Kurucz models show larger dispersions. This fact seems to indicate that Kurucz models are less stable for lower temperatures, since $`\gamma `$ Dor stars are cooler than the majority of the $`\delta `$ Scuti stars. This fact could be also the explanation of the more concentrated values found before for $`\delta `$ Scuti stars. In any case the phase differences observed in the Strömgren bands for these stars seem to indicate that phase lags for these stars are close to zero. This is the first time that these values are calculated for these stars and, if confirmed, could be an important clue to determine the pulsational characteristics of these not very well understood variables. It is to be noted here that $`\mathrm{\Psi }^T=180^{}`$ for an adiabatic atmosphere and that $`\mathrm{\Psi }^T=0^{}`$ for a stellar spot! ## 6. Some examples of practical identifications ### 6.1. $`\delta `$ Scuti stars The next step following the flow chart of Fig 1 is to calculate, from equation (1) and with known phase lags and $`R`$ known, the $`l`$-value which minimizes the distances from the predicted values to the observed ones in all the photometric bands. To do that I have considered phases and amplitudes separately. The best fit, in the sense of minimum variance for amplitudes, is marked “Amplitude variance” in the plots 17, 18, 19 and 10 and the corresponding to the phases, is marked “Phase variance”. The reason is that when looking at the “regions of interest” there are some $`l`$-values which are separated in the phase difference axis, like $`l`$=0 and 1 in Fig 12 and other are separated in the amplitude ratio axis, like $`l`$=3. As already mentioned the best way to test these identifications is to apply the method to the well known double mode $`\delta `$ Scuti stars which are oscillating into two radial modes. In Fig 17 I plot the results for these stars for the first four $`l`$-values and for the two comparison axis: amplitudes and phases. As can be seen all the minima, both for fundamental and overtone modes, in the “phase variance” panels, are located always and in a very clear manner, except maybe for the first overtone of BP Peg, in the $`l`$=0 position as expected from its radial nature. However the minima regarding “amplitude variance” are not very different for the first $`l`$=0, 1, 2 values reflecting the behavior seen in Fig 12, in the sense that phases are discriminant in this regime for the lowest $`l`$-values and amplitudes are for the higher ones. A blind application of the method for phases and amplitudes together could introduce then some extra variance which could change the minimum. It is important to know, before doing these comparisons, where the star regime falls in the phase difference vs amplitude ratio diagram. When the method is applied to HADS the results are clear: all of them are pulsating in a radial mode (Fig 18). Nevertheless when applied to the low amplitude $`\delta `$ Scuti stars the results are less conclusive because of the small amplitudes and the corresponding larger relative errors in the determination of amplitudes and especially phases. The results for the three best studied stars (due to their monoperiodicity) plotted seem to indicate that at least 20 CVn is oscillating in a radial mode, but the other two seem to be pulsating in an $`l`$=1 or 2 mode (Fig 19). ### 6.2. $`\gamma `$ Doradus stars The results for the $`\gamma `$ Dor stars, as indicated in Fig 20, show that all frequencies for all stars seem to be $`l`$=1 non-radial modes if we accept the “phase variance” indications, whereas the discrimination is not so clear for the “amplitude variance” plots, indicating in this case the presence of modes with $`l`$=1, 2 but not certainly 3. As explained in the previous section it would be convenient, in order to select the most relevant discriminator, to construct a “region of interest” for these variables with the new phase lag derived before. Such a plot is shown in Fig 21, where the relevant zones have been calculated for phase lags close to zero and atmospheric characteristics different to those of a $`\delta `$ Scuti star plus a very different pulsation constant which for a $`\gamma `$ Dor star is an order of magnitude larger. It is clear from that figure that discrimination between $`l`$=1 and $`l`$=2 modes is based on phase differences and they are not distinguishable from amplitude ratios; $`l`$=3 is however well discriminated from the other two lower $`l`$-values. Going back to the interpretation of the results shown in Fig 20 for $`\gamma `$ Dor stars, it becomes clear from the amplitude diagrams that these stars are not oscillating in an $`l`$=3 mode and, from the phase diagrams, that they are probably oscillating in an $`l`$=1 mode. However the small amplitudes and small phase differences observed in these stars prevent me to be sure about these conclusions. It would be very interesting to perform simultaneous photometry and spectroscopy for these objects in order to have real measurements of their phase lags. ## 7. Color information in space missions The motivation to do asteroseismology from space is to detect and analyze solar like oscillations and to learn about the stellar interiors in other stars than the Sun. Color informations for these observations are not relevant since the modes can be easily identified from the rotational splitting. Due to the extremely small amplitudes expected to be measured, the number of photons is a crucial quantity in order to decrease the noise and therefore white light is the preferred solution. Modal identification is made by direct measuring of the rotational splitting. However it was previously demonstrated that multicolor information is essential to identify the modes for the stars we are interested in this conference, i.e. $`\delta `$ Scuti and $`\gamma `$ Dor variables. Asteroseismological techniques become therefore possible opening the possibility to test stellar interiors in this region of the HR diagram. COROT (http://www.astrsp-mrs.fr/www/ecorot.html) is a French asteroseismological and planet detection space mission which is now under study by the CNES. The scientific mission is explained in another part of this meeting. Basically there are two CCDs in the focal plane, one of them dedicated to asteroseismology, mainly for solar like oscillations, with no color information, and the other dedicated to planet detection by transit with color information in order to distinguish transits from other active phenomena in stars. This color information is achieved through the interposition of a prism in front of the CCD. The idea is to download two or three parts of this spectrum so giving different photometric colors. The field will contain several thousands of stars in order to be able to detect some events due to terrestrial planet transits during the two years for which the mission is expected to be operative. MONS is mission lead by the Danish community which is also under study and will probably offer some color information. The reader is referred to another part of this meeting. Although the mission is not yet decided, it would probably have a dichroic prism in order to measure simultaneously two colors. In any case these missions will provide data with very high precision and then much of the problems we have now for extracting information from photometry probably will disappear. We will be able to get data for multiperiodic $`\delta `$ Scuti stars and also for $`\gamma `$ Dor with large enough amplitude to make feasible the photometric techniques that I review in this report. I have made some simulations, constructing an arbitrary 12 color photometric system, for these two kind of pulsating stars in order to see the order of magnitude of the phases differences and amplitude ratios among the different photometric bands. These bands have arbitrarily 100 Å width distributed over the visible spectrum each 300 Å. I present the results in Fig 22 for a $`\delta `$ Scuti regime and in Fig 23 for a $`\gamma `$ Dor regime. As shown there the main discriminant factor for radial and non-radial modes in $`\delta `$ Scuti stars is the phase difference between two colors, being also possible, depending on the precision, to discriminate among the non-radial modes. The main contribution of the amplitude ratios is to separate the high order $`l`$=4 mode from the others. In any case a combination of the two diagrams will provide, at the precision we need, a clear identification for the lowest $`l`$-values in the sense that the larger the difference in wavelength the better the discrimination, at least up to 3700 Å. For $`\gamma `$ Dor stars the situation is slightly different mainly due the the very high pulsation constant and the very different phase lag. As shown before the amplitude ratios do not separate $`l`$=1 from $`l`$=2 but the other non-radial modes are more clearly discriminated. The same is true for $`l`$=2 and $`l`$=3 concerning phase differences but fortunately the other modes can be separated. The combined information from the two panels could be of capital importance to separate modes. With the precision we expect to have in COROT, from 100 to 10 ppm depending on the magnitude in the exoplanet focal plane, I estimate here phases and amplitudes will be so small that the method described becomes realistically applicable. I think that under these conditions an asteroseismological study of these variables is possible without the ambiguities of miss-identifications of the modes. ## 8. Conclusions We already know that simple matching of observed to theoretically calculated frequency spectra for the best observed $`\delta `$ Scuti variables, such as FG Vir in Breger et al. (1999), do not give a unique solution. Possible combinations varying some physical inputs of the static models do not allow us to constraint the theoretical model. An identification procedure is then required for the observed modes. We have seen that the success of the photometric method to identify the degree of the spherical harmonic $`l`$ for $`\delta `$ Scuti and $`\gamma `$ Dor stars depends on the model atmospheres we are using and on the precision on the photometric data we are analysing. Although the global uncertainties bound to the photometric calibrations, i.e. calculation of $`T_{\mathrm{eff}},\mathrm{log}g,[M/H]`$and$`Q`$, do not affect dramatically to the discrimination procedure, the model atmospheres present some subtle characteristics. Kurucz models present some inconsistencies, at the level of the required continuities in the fluxes and derivatives which, in the HR region where the stars we are investigating fall, seem to be related with the treatment of convection and in particular of overshooting. Other improved models, such as those described in Smalley and Kupka (1997), present a smoother behavior at these temperatures and gravities but some inconsistencies still remain regarding continuities in the flux derivatives. An appropriate comparison with very accurate photometric data on the variables could be useful to improve these discontinuities in the models. Photometric precision is now sufficient only for the highest amplitude pulsating variables. In particular, nowadays existing HADS photometric data with precisions of around 1 mmag, are useful to identify their oscillation modes. From observations of other variables with lower amplitudes I estimate that we need at least signal to noise ratios of the order of 20–30 in order to have a stable Fourier solution and then make the photometric method feasible. The method supplies reasonable $`l`$-values for HADS, i.e. all of them are found to be radial pulsators, and for some other $`\delta `$ Scuti stars with lower amplitudes non-radial modes seem to be identified. In any case when applied to known radial pulsators, as the double mode $`\delta `$ Scuti stars, the results are consistent with their radial nature. When applied to the new discovered $`\gamma `$ Dor stars the method is able to give estimations of two not very well known physical quantities, $`\mathrm{\Psi }^T`$ and $`R`$. The phase lags for these stars, if confirmed, would be very different from the classical value observed in $`\delta `$ Scuti stars. Furthermore the existing multi-band photometry for some of them seems to indicate that these stars are very probably oscillating in an $`l`$=1 mode. It is also shown that, within the uncertainties, the method provides estimations of $`(\mathrm{\Psi }^T,R)`$ which can be used to compare with theoretical models. These quantities depend very much on the convection parameters which are not very well known and could be useful to modelize the convection in these stars, which in turn could give important clues to the understanding of the onset of the convection in this region of the HR diagram. The new generation of asteroseismological space missions will provide in a next future a huge quantity of very high quality data for these stars. As demonstrated in this review and in order to understand and disentangle the complicated frequency spectrum of multiperiodic $`\delta `$ Scuti stars we need colored information as demonstrated in this review. Actual model atmospheres need to be improved, specially the smoothness in the derivatives and limb darkening variations, in order to extract the whole potential of the photometric discrimination method. #### Acknowledgments. I am very grateful to Enrique Solano for supplying me the Kurucz fluxes and limb darkening coefficients and to Antonio Claret for the PHOENIX limb darkening coefficients. I am also grateful to Chris Sterken for many helpful comments, and to the editors for making possible a more readable version of the manuscript. ## References Balona, L. A. and Stobie, R. S. 1979a, MNRAS, 189, 627 Balona, L. A. and Stobie, R. S. 1979b, MNRAS, 189, 649 Balona, L. A. and Evers, E. A. 1999, MNRAS, 302, 349 Breger, M. et al. 1993, A&A, 271, 482 Breger et al. 1976, ApJ, 210, 163 Breger, M. et al. 1999, A&A, 341, 151 Canuto, V. M. and Mazzitelli, I. 1991, ApJ, 370, 295 Canuto, V. M. and Mazzitelli, I. 1992, ApJ, 389, 724 Dziembowski, W. 1977, Acta Astronomica, 27, 203 Garrido, R. et al. 1990, A&A, 234, 262 Frandsen, S. et al. 1996, A&A, 308, 132 Hauschildt, P. H. 1992, J. Quant. Spectrosc. Radiat. Transfer, 47, 433 Hauschildt, P. H. 1993, J. Quant. Spectrosc. Radiat. Transfer, 50, 301 Hauschildt, P. H. and Baron, E. 1995, J. Quant. Spectrosc. Radiat. Transfer, 54, 987 Hauschildt et al. 1996, ApJ, 462, 386 Hayes, D. S. 1985, ApJ, 289, 726 Hayes, D. S. and Latham, D. W. 1975, ApJ, 197, 593 Hintz, E. G. et al. 1998, PASP, 110, 689 Kurucz, R. L. 1993, Kurucz CD-ROM 13:ATLAS9, SAO, Cambridge, USA Pamyatnykh, A. A. et al. 1998, A&A, 331, 141 Rodriguez, E. et al. 1988a, Rev. Mex. Astron. Astrofis., 16, 7 Rodriguez, E. et al. 1989, Thesis, Univ. Granada Rodriguez, E. et al. 1990, Rev. Mex. Astron. Astrofis., 20, 37 Rodriguez, E. et al. 1992a, A&AS, 93, 189 Rodriguez, E. et al. 1992b, A&AS, 96, 429 Rodriguez, E. et al. 1993a, A&A, 273, 473 Rodriguez, E. et al. 1993b, A&AS, 100, 571 Rodriguez, E. et al. 1993c, A&AS, 101, 421 Rodriguez, E. et al. 1994, A&AS, 106, 21 Rodriguez, E. et al. 1995, MNRAS, 277, 965 Rodriguez, E. et al. 1998, A&A, 331, 171 Rolland, A. et al. 1991, A&AS, 91, 347 Smalley, B. and Kupka, F. 1997, A&A, 328, 349 Smalley, B. 1993, A&A, 274, 391 Stamford, P. A. and Watson, R.D. 1981, Ap&SS, 77, 131 Watson, R. D. 1988, Ap&SS, 140, 225
warning/0001/astro-ph0001401.html
ar5iv
text
# Power Density Spectra of Gamma-Ray Bursts ## 1. Introduction The light curves of gamma-ray bursts (GRBs) typically have many random peaks and in spite of extensive statistical studies (e.g., Nemiroff et al. 1994; Norris et al. 1996; Stern 1996), the temporal behavior of GRBs remains a puzzle. Contrary to the complicated diverse behavior in the time domain, long GRBs show a simple behavior in the Fourier domain (Beloborodov, Stern, & Svensson 1998). Their PDS is a power law of index $`\alpha 5/3`$ (with a break at $`1`$ Hz) plus standard (exponentially distributed) statistical fluctuations superimposed onto the power law. The PDS slope and the break characterize the process randomly generating the diverse light curves of GRBs. Intriguingly, the PDS slope coincides with the slope of the Kolmogorov law. The power-law behavior is seen in individual bursts (we illustrate this in §3) and may provide a clue to the nature of GRBs. In the present paper, we study in detail the PDSs of gamma-ray bursts. In our analysis we use 527 GRB light curves with 64 ms resolution obtained by the Burst and Transient Source Experiment (BATSE) in the four LAD energy channels, I–IV: (I) $`2050`$ keV, (II) $`50100`$ keV, (III) $`100300`$ keV, and (IV) $`h\nu >300`$ keV. The background is subtracted in each channel using linear fits to the 1024 ms data. The method of data analysis is described in §2. In §3, we study a sample of the 4 brightest and longest bursts. In §4 the average PDS for the full sample of 527 GRBs is calculated and discussed. In §5, we calculate the PDSs in the separate energy channels and quantify the difference of the temporal structure between the channels in terms of the PDS slope. We also calculate the autocorrelation function (ACF) in the separate channels and compare the results with previous studies of the ACF. In §6, we address dim GRBs and compare them to the bright ones. The results are discussed in §7. ## 2. Data Analysis ### 2.1. The Sample Long bursts are of particular interest since their internal temporal structure can be studied by spectral analysis over a larger range of time scales. We choose bursts with durations $`T_{90}>20`$ s where $`T_{90}`$ is the time it takes to accumulate from 5% to 95% of the total fluence of a burst summed over all the four channels, I+II+III+IV. Hereafter, we measure the brightness of a burst by its peak count rate, $`C_{\mathrm{peak}}`$, in channels II+III. We analyze bursts with $`C_{\mathrm{peak}}>100`$ counts per time-bin, $`\mathrm{\Delta }t=0.064`$ s. This condition coupled with the duration condition, $`T_{90}>20`$ s, gives a sample of 559 GRBs. The sample still contains GRBs with low fluence, which are not good for Fourier analysis. We exclude bursts with fluences $`\mathrm{\Phi }<32C_{\mathrm{peak}}`$. The resulting sample contains 527 bursts. ### 2.2. PDS Calculation We calculate the Fourier transform, $`C_f`$, of each light curve, $`C(t)`$, using the standard Fast Fourier Transform method. The power density spectrum, $`P_f`$, is given by $`P_f=(C_fC_f^{}+C_fC_f^{})/2=C_fC_f^{}`$ since $`C(t)`$ is real. The Fourier transform is calculated on a standard grid with a time bin $`\mathrm{\Delta }t=64`$ ms and a total number of bins $`N_{\mathrm{bin}}=2^{14}`$, which corresponds to a total time $`T1048`$ s since the trigger time. The light curve of each burst is considered in its individual time window $`(t_1,t_2)`$ (see §2.4). In the calculations of the PDS, we extend the time interval to $`(0,T)`$ by adding zeros at $`(0,t_1)`$ and $`(t_2,T)`$. The adding zeros introduces random small-scale fluctuations in the PDS (see, e.g., Bracewell 1967). We, however, study the PDSs averaged over adjacent frequencies and/or over a sample of GRBs. Then the fluctuations in $`P_f`$ associated with a specific choice of the grid disappear and do not affect the results. ### 2.3. The Poisson Level Poisson noise in the measured count rate affects the light curve on short time scales, and, correspondingly, affects the PDS at high frequencies. The Poisson noise has a flat spectrum, $`P_ff^0`$, introducing the “Poisson level”, $`P_0`$, in a PDS. This level equals the burst total fluence including the background in the considered time window. The power spectrum above this level displays the intrinsic variability of the signal (but see §4.2 for the time-window effects). We calculate the individual Poisson level for each burst and subtract it from the burst PDS. ### 2.4. The Time Window One would like to see the whole burst in the window. However, (1) it is difficult to determine exactly the end of a burst because the burst can always have a weak “tail” hidden in the background; (2) the window should not be very large because inclusion of long weak tails leads to an increase of the background fluence, $`\mathrm{\Phi }_b`$, without a substantial increase in the signal fluence, $`\mathrm{\Phi }`$. The resulting low ratio $`\mathrm{\Phi }/\mathrm{\Phi }_b`$ implies a high Poisson level, which makes the quality of the PDS worse. To reduce the Poisson level, we cut off the light curves at a time $`t_2`$ defined so that the signal count rate $`C(t)`$ does not exceed $`\epsilon C_{\mathrm{peak}}`$ at $`t>t_2`$ and $`C(t_2)=\epsilon C_{\mathrm{peak}}`$. Keeping in mind bursts with a weak beginning, we define the starting window time, $`t_1`$, so that $`C(t)<\epsilon C_{\mathrm{peak}}`$ at $`t<t_1`$ and $`C(t_1)=\epsilon C_{\mathrm{peak}}`$. In our analysis, $`\epsilon =0.05`$ is chosen. We checked that varying $`\epsilon `$ does not significantly affect the results unless $`\epsilon >0.1`$. Note that, for dim bursts, Poisson fluctuations of the background (which are imprinted on $`C(t)`$ even after subtraction of the average background level) may exceed $`\epsilon C_{\mathrm{peak}}`$. Then the window $`(t_1,t_2)`$ is determined by the background fluctuations rather than by the signal. ## 3. Individual Bright Bursts The brightest and longest bursts are the best ones for Fourier analysis. In this section, we study the four brightest bursts with $`T_{90}>100`$ s. They have trigger numbers # 2156 (GRB 930201), 2856 (GRB 940302), 3227 (GRB 941008), and 6472 (GRB 941110). We take the light curves, $`C(t)`$, summed over channels II and III, in which the signal is strongest. To simplify the comparison of different bursts, we take peak-normalized light curves. Their Fourier transform, $`C_f`$, is therefore normalized by $`C_{\mathrm{peak}}`$, and the PDS, $`P_f=C_fC_f^{}`$, is normalized by $`C_{\mathrm{peak}}^2`$. The four peak-normalized light curves and their PDSs are shown in Figure 1 (we smooth the PDSs on the scale $`\mathrm{\Delta }\mathrm{log}f=0.04`$ before plotting). The light curves are very different while the PDSs are similar. They can be described as a single power law, $`\mathrm{log}P_f=A+\alpha \mathrm{log}f`$ ($`A1`$ and $`\alpha 1.5`$) with super-imposed fluctuations $`\mathrm{\Delta }P_f/P_f1`$. In spite of the large $`\mathrm{\Delta }P_f`$, the power-law behavior can be seen in each burst due to the power-law extending over more than two decades in frequency. The presence of an underlying power law is an interesting feature of the GRB temporal behavior. A possible way to extract it from the noisy individual PDSs is to take the average PDS over a sample of long GRBs. Then the fluctuations affecting each individual PDS tend to cancel each other and one can see the power-law behavior. The average of the four PDSs is plotted in Figure 2. One can see that the amplitude of irregularities is reduced after the averaging and one can measure the slope of the resulting PDS. The power-law fit in the range $`2.0<\mathrm{log}f<0.2`$ gives $`\mathrm{log}P_f=1.251.45\mathrm{log}f`$. (We use the standard $`\chi ^2`$ fitting procedure with equal weights for each frequency bin.) The PDS analysis was previously performed for a number of individual bursts. Belli (1992), for example, analyzed GRBs detected by the Konus experiment onboard the Soviet probes Venera 11 and Venera 12, and Giblin et al. (1998) studied individual BATSE bursts. The power-law PDS can be observed in their longest complex bursts as well. When turning to GRBs with relatively modest durations, $`T_{90}20100`$ s, it becomes more difficult to see the power-law behavior in individual bursts. The statistical fluctuations, $`\mathrm{\Delta }P_f/P_f1`$, significantly affect the apparent PDS slope derived for the available range of frequencies, which is limited by $`T_{90}^1`$. However, having a large number of bursts allows one to study the PDS by averaging over the sample. ## 4. The Average PDS In most of the GRB models (e.g., in the internal shock model), different bursts are produced by one physical mechanism of stochastic nature, i.e., an individual burst is a random realization of the same standard process. Stern & Svensson (1996) suggested an empirical pulse-avalanche model of this type. They showed that near-critical avalanches reproduce the statistical characteristics of GRBs as well as their extreme diversity. With such an approach, individual GRBs are like pieces of a “puzzle”, and the features of the standard engine can be probed with statistical methods applied to a sufficiently large ensemble of GRBs. ### 4.1. The averaging The simplest statistical characteristic of the PDS is the average PDS, $`\overline{P}_f`$. The averaging means that we sum up the PDSs of individual bursts with some weights (normalization) and divide the result by the number of bursts in the sample. If no normalization is employed, then the brightest bursts strongly dominate $`\overline{P}_f`$, and their individual fluctuations lead to strong fluctuations in $`\overline{P}_f`$. The PDSs of relatively weak bursts are suppressed proportionally to $`C_{\mathrm{peak}}^2`$ and they are lost in the averaging. A normalization procedure is thus needed to increase the weight of relatively weak bursts in the sample. We prefer the peak-normalization because: (i) The amplitude of the resulting PDS is practically independent of the burst brightness (see, e.g., Fig. 1). (ii) The fluctuations in $`\overline{P}_f`$ turn out to be minimal and the best accuracy of the slope is achieved. The distribution of the individual $`P_f`$ around $`\overline{P}_f`$ follows a standard exponential law (see Beloborodov et al. 1998), and the amplitude of fluctuations in $`\overline{P}_f`$ is given by a simple formula, $`\mathrm{\Delta }\overline{P}_f/\overline{P}_fN^{1/2}`$, where $`N`$ is the number of bursts in the sample. The averaging procedure is the following. (1) For each burst we first determine its individual Poisson level. (2) We determine the peak, $`C_{\mathrm{peak}}`$, of the light curve, $`C(t)`$, with the background subtracted. (3) We normalize the light curve to its peak. (4) We calculate the PDS of the normalized light curve. (5) We normalize the Poisson level by $`C_{\mathrm{peak}}^2`$ and subtract it from the PDS. (6) We sum up the resulting PDSs over the sample and divide by the number of bursts. The resulting average PDS, $`\overline{P}_f`$, for the 527 peak-normalized light curves in channels (II+III) is shown in Figure 3. For comparison, we also plot the average PDS with the $`\sqrt{\mathrm{\Phi }}`$-normalization, where $`\mathrm{\Phi }`$ is the burst fluence excluding the background. Each light curve is then normalized by $`\sqrt{\mathrm{\Phi }}`$, and its PDS (and its Poisson level) is normalized by $`\mathrm{\Phi }`$. This normalization gives different weights (the weights of dim bursts are reduced as compared to the peak-normalization). The resulting average PDS has a different amplitude. Nevertheless, $`\overline{P}_f`$ again follows a power law with approximately the same slope. This provides evidence that the self-similar behavior with $`\alpha 5/3`$ is an intrinsic property of GRBs, rather than an artifact of the averaging procedure. This interpretation is also supported by the fact that we observe the power law in the longest individual bursts. When comparing individual PDSs, $`P_f`$, with the peak-normalized $`\overline{P}_f`$, one sees that $`P_f`$ is exponentially distributed around $`\overline{P}_f`$, and the distribution is self-similar with respect to shifts in $`f`$. The power-law fit to $`\overline{P}_f`$ in the range $`1.4<\mathrm{log}f<0.1`$ gives $`\alpha 1.75`$ for the peak-normalized bursts, and $`\alpha 1.67`$ for $`\sqrt{\mathrm{\Phi }}`$-normalized bursts. The formal error found from the $`\chi ^2`$ fitting is very small, $`\mathrm{\Delta }\alpha \pm 0.01`$ (we take the variance, $`\mathrm{\Delta }\overline{P}_f/\overline{P}_f=N^{1/2}`$). Note that the slope in the peak-normalized case is different from $`1.67`$ reported in Beloborodov et al. (1998). This change is caused by the fact that we have extended the sample to smaller brightnesses (see §6). The deviation from the power law at the low-frequency end is due to the finite duration of bursts. At the high frequency end, there is a break at $`1`$ Hz. The break is observed in the brightest GRBs even without subtracting the Poisson level (see Beloborodov et al. 1998 and the top panel in Fig. 8). Note that the break position is the same for the peak-normalized and $`\sqrt{\mathrm{\Phi }}`$-normalized bursts (see Fig. 3b). It stays the same in the separate energy channels (see Fig. 5) and does not depend on the Poisson level. One may also see the break in individual long bursts (Figs. 1 and 2). The break is far too sharp to be explained as an artifact of the 64 ms time binning, which suppresses the PDS by a factor of $`[\mathrm{sin}(\pi f\mathrm{\Delta }t)/(\pi f\mathrm{\Delta }t)]^2`$ where $`\mathrm{\Delta }t=64`$ ms is the time bin (cf. van der Klis 1989). ### 4.2. The Effects of Finite Signal Duration Fourier analysis was designed for physical problems dealing with linear differential equations. For example, it is usually applied to small perturbations above a given background solution. The Fourier power spectrum is also commonly used in the temporal studies of long signals or noises, e.g., in persistent astrophysical sources. By contrast, GRBs have strongly non-linear signals with short durations. The typical number of BATSE time-bins ($`\mathrm{\Delta }t=64`$ ms) in a long GRB is a few $`\times 10^3`$. Do the effects of finite duration (i.e., time-window effects) strongly affect the measured PDS? The issue is illustrated in Figure 4. We prepared an artificial long signal with $`\overline{P}_ff^\alpha `$ and exponentially distributed $`P_f`$, i.e., the probability to detect $`P_f`$ at a given $`f`$ is proportional to $`\mathrm{exp}(P_f/\overline{P}_f`$). The phase structure was taken to be Gaussian (i.e., random). The signal duration is $`T_0=2^{17}\mathrm{\Delta }t8400`$ s. Then, we cut the long signal into 64 pieces of equal length $`t_0=2^{11}\mathrm{\Delta }t130`$ s. We thus get 64 short signals, each is a random realization/fragment of the same stationary process characterized by the index $`\alpha `$. Analogously to our analysis of real GRBs, we normalize each signal to its peak and add zeros up to $`T=2^{14}\mathrm{\Delta }t`$ (our standard grid). Then we calculate the average PDS, $`\overline{P}_f`$, for the 64 artificial signals. The result is shown in Figure 4 for the three cases: $`\alpha =0`$ (Poisson), $`\alpha =1`$ (flicker), and $`\alpha =5/3`$ (Kolmogorov). One can see that the time-window effects are strong in the case $`\alpha =0`$. The Poisson signal is roughly constant on large time scales. As a result, at modest $`f`$, $`\overline{P}_f`$ is just the power spectrum of the $`t_0`$-window. By contrast, in the case $`\alpha =5/3`$, we have large-amplitude variations in the signal on all time scales. The average PDS of the 64 short signals then reproduces well the intrinsic PDS slope throughout the whole range of frequencies, down to $`ft_0^1`$ (see bottom panel in Fig. 4). Hence, the power law we observe in the average PDS of GRBs can be interpreted as that GRBs are random short realizations of a standard process which is characterized by the PDS slope $`\alpha 5/3`$. Note, however, that the power spectrum does not provide a complete description of the signal since the phase structure is not considered. ## 5. The PDS and the ACF in Channels I, II, III, IV What does the average PDS look like in the separate energy channels? The signal to noise ratio is low in channel I and especially in channel IV. The number of GRBs for which good PDSs can be obtained in channels I–IV is therefore limited to the brightest bursts. We choose a sample of bursts with $`C_{\mathrm{peak}}>500`$ counts/bin (in channels II+III), which contains 152 bursts. The burst analysis is now performed in each channel separately. We determine the Poisson level of a burst in each channel, $`P_0^i`$, and find the peak of the light curve, $`C_{\mathrm{peak}}^i`$, where $`i=`$ I, II, III, IV. Then we perform the peak-normalization: $`C^i(t)C^i(t)/C_{\mathrm{peak}}^i`$ and $`P_0^iP_0^i/(C_{\mathrm{peak}}^i)^2`$. We set the time window in each channel $`(t_1^i,t_2^i)`$ as described in §2.4. The resulting average PDSs in channels I–IV are shown in Figure 5. The differences in the slopes are clearly seen. We fitted the PDSs by power laws, $`\mathrm{log}\overline{P}_f=A+\alpha \mathrm{log}f`$, in the range $`1.6<\mathrm{log}f<0`$. Channel IV is fitted in the range $`1.3<\mathrm{log}f<0.1`$. $`A`$ and $`\alpha `$ of the fits are listed in Table 1. | | Table 1. | | | | | --- | --- | --- | --- | --- | | Channel | I | II | III | IV | | $`\alpha `$ | $`1.72`$ | $`1.67`$ | $`1.60`$ | $`1.50`$ | | $`A`$ | $`1.03`$ | $`1.05`$ | $`1.06`$ | $`1.07`$ | The observed behavior can be briefly described as follows: The “red” power ($`\overline{P}_f`$ at low $`f`$) decreases at high photon energies and increases at low photon energies. It should be compared with the well-known fact that the pulses in a GRB are narrower in the hard channels (e.g., Norris et al. 1996). The hardness of emission varies dramatically during a burst and this leads to different temporal structure in different channels. E.g., pulses observed in the soft channels may be suppressed in the hard channels. One therefore could expect that the PDS has different slopes in different channels. Note that, typically, most of the GRB energy is released in channels II+III, and the average PDS of bolometric light-curves approximately follows the $`5/3`$ law. In principle, the autocorrelation function (ACF) contains the same information as the PDS, since one is the Fourier transform of the other (the Wiener-Khinchin theorem). In practice, the two are not completely equivalent because of the time-window effects and the presence of a noisy background. On modest time scales, $`<30`$ s, the direct ACF calculation and the calculation via the Fourier transform of the PDS give the same result to within a few percent (we have calculated the ACF by both methods). The average ACF, $`\overline{A}(\tau )`$, for our sample of 152 bright GRBs is shown in Figure 6 for each of the four channels. The ACF gets narrower at high energies, in agreement with previous studies (Fenimore et al. 1995), except that our ACF is $`2`$ times wider as compared to that obtained by Fenimore et al. (1995). The ACF width averaged over the channels is in approximate agreement with that calculated for the bolometric light curves by Stern & Svensson (1996). Note that the average ACF is obtained by summing up the individual ACFs normalized to unity, i.e., $`A(0)=1`$. Recalling that the ACF is the Fourier transform of the PDS, one can see that this normalization is equivalent to the normalization of the light curve by $`\sqrt{P_{\mathrm{tot}}}`$ where $`P_{\mathrm{tot}}=[C(t)]^2𝑑t=P_f𝑑f`$ is the total power. This normalization is different from the peak-normalization we use in the calculations of the average PDS, i.e., we prescribe different weights to individual bursts when averaging the PDS and the ACF, respectively. Therefore, the average ACF is not the Fourier transform of the average PDS shown in Figure 5. This relation holds when the average PDS is calculated with the $`\sqrt{P_{\mathrm{tot}}}`$-normalization. The ACF in each channel is perfectly fitted by the stretched exponential: $`\overline{A}(\tau )=\mathrm{exp}([\tau /\tau _0]^\beta )`$ (see Fig. 7). The parameters $`\tau _0`$ and $`\beta `$ are listed in Table 2. The index $`\beta `$ is related to the PDS slope $`\alpha `$ by the simple relation $`\beta (1+\alpha )`$. | | Table 2. | | | | | --- | --- | --- | --- | --- | | Channel | I | II | III | IV | | $`\beta `$ | $`0.73`$ | $`0.67`$ | $`0.63`$ | $`0.6`$ | | $`\tau _0`$ | $`14.0`$ | $`10.7`$ | $`7.3`$ | $`5.1`$ | The changing $`\beta `$ demonstrates that the ACF shape changes from channel to channel, as it should do since the PDS changes. As a first approximation, the PDS slope is equal to $`5/3`$, and the ACF index $`\beta 2/3=5/31`$. The values of $`\tau _0`$ quantify the ACF width in different channels. The parameter $`\tau _0`$ is better defined as compared to measuring the ACF width at a certain level, e.g., at $`e^{0.5}`$ of the maximum level, as done in Fenimore et al. (1995). Nevertheless, the scaling of $`\tau _0`$ with energy is approximately the same, $`\tau _0E^{0.4}`$ where $`E`$ is the photon energy of the low energy boundary of the channel, see Fenimore et al. (1995). It should be stressed, however, that the stretching of $`\tau _0`$ is not related to the stretching of any physical time scale of intrinsic correlations during the burst. The ACF width, $`\tau _0`$, is rather related to the position of the breaks in the PDS, especially the low frequency break, which in turn is determined by the burst duration. For an infinitely long signal with a power law PDS, $`\tau _0`$ would tend to infinity. This is natural since the power law behavior implies self-similarity, i.e., the absence of any preferred time scale. Specific time scales are introduced only by the breaks in the PDS. We therefore have only two physical time scales in long GRBs: The first one is associated with the $`1`$ Hz break and the second one is associated with the low frequency break due to the finite burst duration. ## 6. Dim versus Bright Bursts Now we address the following question: Is there any correlation between the PDS slope and the burst brightness? ### 6.1. PDS Slope Correlates with the Burst Brightness We take the full sample of 527 light curves in channels II+III and divide it into 3 groups of different brightnesses: (A) $`C_{\mathrm{peak}}>800`$, (B) $`300<C_{\mathrm{peak}}<800`$, and (C) $`100<C_{\mathrm{peak}}<300`$. Group A contains 91 bursts, group B 222 bursts, and group C 214 bursts. We have calculated the average PDS for each group separately employing the peak-normalization. The results are presented in Figure 8. We fitted the average PDSs by power laws, $`\mathrm{log}P_f=A+\alpha \mathrm{log}f`$. The parameters $`A`$ and $`\alpha `$ of the fits in the range $`1.5<\mathrm{log}f<0.1`$ are listed in Table 3. We conclude that the average PDS of dim bursts gets markedly steeper. | | Table 3. | | | | --- | --- | --- | --- | | group | A | B | C | | $`\alpha `$ | $`1.63`$ | $`1.74`$ | $`1.82`$ | | $`A`$ | $`1.04`$ | $`0.98`$ | $`0.91`$ | ### 6.2. Subtraction of the Poisson Level In the PDS calculations we subtracted the Poisson level, which is quite high for dim bursts (see Fig. 8). One therefore should address a technical question: How well is the intrinsic PDS restored after subtracting the Poisson level? To investigate this issue we have performed the following test. We take the sample of 91 bright bursts (group A) and add a strong Poisson noise (with an average level of 25000 counts/bin) to each burst in the group. We thus artificially increase the Poisson level by two orders of magnitude. Alternatively, one may consider this procedure as a rescaling of bright bursts to a smaller brightness while keeping the Poisson background at the same level. Then, we analyze the artificial bursts in the same way as we did with the real dim GRBs. Note that even after subtraction of the time-averaged background, its Poisson noise strongly affects the signal and may create artificial peaks dominating the true peak of the signal. For the artificial bursts, we know the true peak (which is the peak of the original bright burst). In real dim bursts we do not know the position of the true peak and employ the peak search scheme described in Stern, Poutanen, & Svensson (1999). The result is compared with the average PDS of the original sample in Figure 9. We find that the subtraction of the Poisson level allows one to restore the original PDS, $`P_f`$, also at frequencies where $`P_f`$ is well below the Poisson level. Even the 1 Hz break remains present. We conclude that the high Poisson level is unlikely to significantly affect the measured PDS slope. ## 7. Discussion ### 7.1. Relation to the Average Time Profile The average time profile (ATP) of GRBs was found to follow a stretched exponential of index $`1/3`$ (Stern 1996). Is there any relation between the ATP and the average PDS? One should note two important differences between the ATP and the PDS studies: (1) the $`5/3`$ PDS, though affected by the statistical fluctuations, is observed in individual bursts, while the ATP is a purely statistical property of a large sample of bursts; (2) the ATP is constructed for GRBs of all durations (and only in this case it displays the perfect stretched exponential) while the PDS is studied for long bursts only. Nevertheless, both the ATP and the PDS characterize the stochastic process generating GRBs. Stern (1999) shows that both can be reproduced simultaneously with the pulse-avalanche model of Stern & Svensson (1996). ### 7.2. The 1 Hz Break The sharpness of the break in the average PDS appears to be an important feature that constrains models of GRBs. If the signal is produced in the rest frame of a relativistic outflow then variations of the outflow Lorentz factor, $`\mathrm{\Gamma }`$, from burst to burst would smear out the break. The sharpness of the break then implies a narrow dispersion of $`\mathrm{\Gamma }`$, $`\mathrm{\Delta }\mathrm{\Gamma }/\mathrm{\Gamma }<2`$, which appears to be unlikely. Alternatively, the GRB variability may come from the central engine. The break in the PDS might then be related to a typical time scale, $`1`$ s, in the central engine. Finally, one may associate the break with the dynamical time scale corresponding to the inner radius of the optically thin zone of the outflow, $`R_{}`$. Then the variability on time scales shorter than $`t_{}=R_{}/c\mathrm{\Gamma }^2`$ can be suppressed. In this case, however, one should explain why $`t_{}`$ stays the same in different bursts. ### 7.3. Dim GRBs It has often been hypothesized that dim bursts are at high cosmological redshifts. For instance, it is necessarily the case if GRBs have approximately the same intrinsic luminosity, the so-called ”standard candle” hypothesis. We can test this hypothesis using the power spectrum analysis. Suppose that the dim bursts are intrinsically the same as the bright ones. Then any difference in their average PDS should be due to a cosmological redshift. First consider the bolometric light curves assuming that their average PDS follows the $`5/3`$ power law. It is easy to see that a redshift, $`z`$, will not change the PDS slope. As we normalize each bursts to its peak before the averaging, the effect of a redshift is just stretching the light curve in time, i.e., precisely the time dilation effect. This will lead to an increase of the net normalization of the PDS by a factor of $`(1+z)^{1/3}`$. The slope does not change since the dilation factor $`(1+z)`$ is the same for each time scale. One should, however, recall that we observe bursts in a limited spectral interval. A redshift then implies a shift of the signal with respect to our spectral window. E.g., photons detected in channel III would originally have been emitted in channel IV. As we know from §5, the PDSs are different in different energy channels. Therefore, one expects that the PDS slope will change for redshifted GRBs. We have seen in §5 that the PDS of bright bursts flattens in the hard channels. Hence, a redshift of the bright bursts must be accompanied by a flattening of the PDS. Contrary to this behavior, we observe that the PDS of dim bursts steepens. Hence, the evolution of the PDS with brightness is inconsistent with the standard candle hypothesis. It implies that the burst luminosity function is broad and dim bursts are intrinsically weak. Evidence for a broad luminosity function is also found when looking at the isotropic luminosities of the bursts with measured redshifts. Note, however, that the differences in the apparent luminosities could be caused by orientation effects if GRBs are beamed. One should not therefore exclude that the total intrinsic luminosities are approximately the same. The different temporal structure of dim bursts may be an important fact in this respect. In particular, it suggests that the observed time dilation of dim bursts (e.g., Norris et al. 1994) may be caused mainly by physical processes occurring in the bursts rather than by a cosmological redshift. Note that the intrinsic difference of the temporal structure of dim GRBs was also found when studying their average time profile (see Stern et at. 1997). The rising part of the ATP does not change with decreasing brightness while the decaying part suffers from time dilation. This behavior is inconsistent with a cosmological time dilation which should apply equally to both parts of the ATP. ### 7.4. Internal Shocks A likely scenario of GRBs involves internal shocks in a relativistic outflow with a Lorentz factor $`\mathrm{\Gamma }10^2`$ (see, e.g., Piran 1999 for a review). The shock develops when an inner faster shell of the outflow tries to overtake the previous slower shell. The pulses in a burst are then associated with collisions between the shells. The PDSs predicted by the model were recently tested against the observed $`5/3`$ law (Panaitescu, Spada, & Mészáros 1999; Spada, Panaitescu, & Mészáros 2000; Beloborodov 1999,2000). The self-similar temporal structure with $`\alpha =5/3`$ can be reproduced by the model. Further constraints should, however, be imposed by the observed dependence of $`\alpha `$ on photon energy. Besides, the phase structure of the signal, neglected so far, should be taken into consideration (Beloborodov 2000). This research has made use of data obtained through the HEASARC Online Service provided by NASA/GSFC. We thank A.F. Illarionov, I. Panchenko, and J. Poutanen for discussions. This work was supported by the Swedish Natural Science Research Council, the Swedish Royal Academy of Science, the Wennergren Foundation for Scientific Research, a NORDITA Nordic Project grant, the Swedish Institute, RFBR grant 97-02-16975, and NSF grant PHY94-07194.
warning/0001/astro-ph0001215.html
ar5iv
text
# 𝐻₃^(limit-from2+) molecular ion in a strong magnetic field: a triangular configuration ## Abstract The bound state in the system of three protons and an electron $`(pppe)`$ under a homogeneous strong magnetic field where the protons are situated in the vertices of an equilateral triangle perpendicular to the magnetic field lines is found. It is shown that for magnetic fields $`B=10^{11}4.414\times 10^{13}G`$ the potential energy curves as a function of the internuclear distance $`R`$ have an explicit minimum. For all magnetic fields studied, the binding energy of the triangular configuration is less than the binding energy of the linear parallel configuration (A. Turbiner et al. JETP Lett. 69, p. 844). In the contrary to the linear case, the binding energy decreases with a magnetic field growth, while the equilibrium internuclear distance slowly increases. preprint: México ICN-UNAM 99-03 Recently a quantitative study of the system of three protons and an electron $`(pppe)`$ in a strong magnetic field gave theoretical evidence of the existence of a bound state for magnetic fields $`B>10^{11}G`$ . It manifests the existence of the exotic molecular ion $`H_3^{(2+)}`$. Many years ago it was indicated by Kadomtsev and Kudryavtsev , and Ruderman the possible existence of linear exotic molecular systems in presence of a strong magnetic field. Due to the enormous Lorentz force acting on the electronic cloud, it shrinks in the direction transversal to the magnetic field lines leading to an effective quasi-one-dimensionality of the studied systems. This makes the electron-nuclei attraction more effective to compensate the Coulombic repulsion of nuclei. Such a fact suggests the possibility of the existence of exotic systems which do not exist without a magnetic field. It has been shown some time ago that a strong magnetic field can lead to the formation of linear hydrogenic chains $`H_n,n>2`$ situated along magnetic lines . More recently the first quantitative study about the possible existence of an exotic bound state in the system of three protons and an electron $`(pppe)`$ in a strong magnetic field was carried out and it confirmed existence of $`H_3^{(2+)}`$ ion. Although the linear parallel configuration (where protons are situated on a line parallel to the magnetic field line) for the system $`(pppe)`$ seems to be optimal for very strong magnetic fields, other possible configurations have to be considered in order to have a better understanding of the properties of such systems. In this article we explore the possibility of the existence of the molecular ion $`H_3^{(2+)}`$, where the protons are situated in the vertices of an equilateral triangle perpendicular to a homogeneous magnetic field. We limit our present investigation to magnetic fields $`B=10^{11}4.414\times 10^{13}G`$ for which a non-relativistic approach is still valid (for a detailed discussion see and references therein). Our study is restricted to the ground state within the Born-Oppenheimer approximation. The present calculations are carried out in the framework of a variational method. Trial functions are chosen in accordance with the following criterion (see ): (i) the trial function $`\mathrm{\Psi }_t(x)`$ should include all symmetry properties of the problem in hand; (ii) the trial function for the ground state should not vanish inside the domain where the problem is defined; (iii) the potential $`V_t(x)=\frac{^2\mathrm{\Psi }_t}{\mathrm{\Psi }_t}`$, for which the trial function is an exact eigenfunction, should reproduce the original potential behavior near singularities as well as its asymptotic behavior. This prescription has been successfully applied to the study of the $`H_2^+`$ ion in strong magnetic fields giving the more accurate results for $`B>0`$ , as well as to the study of the linear parallel configuration of the $`H_3^{(2+)}`$ and $`H_4^{(3+)}`$ ions in a strong magnetic fields . We consider a system of one electron and three identical infinitely-heavy centers of unit charge situated on the vertices of an equilateral triangle with internuclear distance $`R`$ (which is the length of the triangle side) lying in a plane perpendicular to the magnetic field of strength $`B`$ directed along the $`z`$ axis, $`\stackrel{}{B}=(0,0,B)`$. The potential corresponding to the system we study is given by $$V=\frac{6}{R}\frac{2}{r_1}\frac{2}{r_2}\frac{2}{r_3}+\frac{B^2\rho ^2}{4}+B\widehat{\mathrm{}}_z$$ (1) where the quantity $`\rho =\sqrt{x^2+y^2}`$ is the distance from the electron position to the $`z`$-axis, $`r_{1,2,3}`$ are the distances from the electron to the first, second and third centers, respectively, $`R`$ is the distance between centers (see Fig. 1 for notations) and $`\widehat{\mathrm{}}_z`$ is the $`z`$-component of the angular momentum operator. The linear Zeeman effect term in the potential $`(B\mathrm{}_z)`$ can be dropped off since we use real trial functions <sup>*</sup><sup>*</sup>*For a real trial function the expectation value of the $`z`$-component of the angular momentum $`\mathrm{}_z`$ equals to zero and leads to a vanishing contribution to the variational energy emerging from the linear Zeeman effect term . (see below). Spin Zeeman effects are neglected. Through the paper the Rydberg is used as the energy unit. For the other quantities standard atomic units are used. The trial functions we are going to use for the present problem are similar to those were exploited in for the study of the linear parallel configuration. These functions were built according to the criterion described in . They contain the basic features of Coulomb systems in a magnetic field, as well as the symmetry property under permutation of the three charged centers. In particular, it implies that the potentials corresponding to those functions reproduce the Coulomb-like behavior near the centers and the two-dimensional oscillator behavior in the $`(x,y)`$ plane at large distances (see below). The simplest of those functions is a function of the Heitler-London type multiplied by the lowest Landau orbital: $$\mathrm{\Psi }_1=e^{\alpha _1(r_1+r_2+r_3)\beta _1B\rho ^2/4},$$ (2) (cf. Eq. (2) in Ref. ), where $`\alpha _1`$ and $`\beta _1`$ are variational parameters. Since we consider the distance between centers as an extra variational parameter, it has in total three variational parameters. This function gives an adequate description of the covalent coupling of the system near equilibrium. As illustration we show the potential corresponding to this function: $`V_1={\displaystyle \frac{^2\mathrm{\Psi }_1}{\mathrm{\Psi }_1}}`$ $`=`$ $`3\alpha _1^2\beta _1B2\alpha _1{\displaystyle \underset{i=1}{\overset{3}{}}}{\displaystyle \frac{1}{r_i}}+2\alpha _1^2{\displaystyle \underset{i<j}{}}(\widehat{n}_i\widehat{n}_j)+\alpha _1\beta _1B{\displaystyle \underset{i=1}{\overset{3}{}}}{\displaystyle \frac{\rho ^2(xx_i+yy_i)}{r_i}}+{\displaystyle \frac{\beta _{1}^{}{}_{}{}^{2}B^2\rho ^2}{4}},`$ (3) $`\widehat{n}_i\widehat{n}_j`$ $`=`$ $`{\displaystyle \frac{1}{r_ir_j}}\left[(xx_i)(xx_j)+(yy_i)(yy_j)+(zz_i)(zz_j)\right],`$ (4) where $`\widehat{n}_i(i=1,2,3)`$ is the unit vector in the direction of the vector pointing from the position of the $`i`$-th center to the position of the electron. A second trial function is a Hund-Mulliken type function multiplied by the lowest Landau orbital: $$\mathrm{\Psi }_2=\left(e^{\alpha _2r_1}+e^{\alpha _2r_2}+e^{\alpha _2r_3}\right)e^{\beta _2B\rho ^2/4},$$ (5) (cf. Eq. (4) in Ref. ). Here $`\alpha _2,\beta _2`$ are variational parameters. This function describes a ionic coupling of a hydrogen atom with two charged centers. Another suitable trial function which describes a ionic coupling between an $`H_2^+`$ ion and a proton is given by $$\mathrm{\Psi }_3=\left(e^{\alpha _3(r_1+r_2)}+e^{\alpha _3(r_1+r_3)}+e^{\alpha _3(r_2+r_3)}\right)e^{\beta _3B\rho ^2/4},$$ (6) (cf. Eq. (5) in Ref. ) where $`\alpha _3,\beta _3`$ are variational parameters. Since the $`H_2^+`$ ion has the lowest total energy among the one-electron systems for $`B10^{13}G`$ , an important contribution coming from this trial function is expected. In order to include in a single trial function the different physical behavior near equilibrium and at large distances, appropriate interpolations of the trial functions (2), (5) and (6) are done: * A natural interpolation is given by a non-linear superposition of the form: $$\mathrm{\Psi }_{4nls}=\left(\underset{\{\alpha _4,\alpha _5,\alpha _6\}}{}e^{\alpha _4r_1\alpha _5r_2\alpha _6r_3}\right)e^{\beta _4B\rho ^2/4},$$ (7) (cf. Eq. (6) in Ref. ) where $`\alpha _4,\alpha _5,\alpha _6`$ and $`\beta _4`$ are variational parameters, and the sum is over all permutations of the parameters $`\{\alpha _4,\alpha _5,\alpha _6\}`$. If all parameters coincide ($`\alpha _4=\alpha _5=\alpha _6=\alpha _1`$), the function (7) reduces to the Heitler-London type function (2). Function (7) also reduces to the Hund-Mulliken type wave function (5) when only one parameter is non-zero say $`\alpha _4=\alpha _2`$ and $`\alpha _5=\alpha _6=0`$. If two parameters are non-zero, and equal, say, $`\alpha _4=\alpha _5=\alpha _3`$, and $`\alpha _6=0`$ it reduces to the trial function (6). When all parameters are different among them and different from zero the function (7) provides us with a $`3`$-center modification of a Guillemin-Zener type function. The function (7) has in total five variational parameters (including the internuclear distance $`R`$ as one of the parameters). * Another more immediate interpolation is given by a linear superposition of the functions (2), (5) and (6). $$\mathrm{\Psi }_{5ls}=A_1\mathrm{\Psi }_1+A_2\mathrm{\Psi }_2+A_3\mathrm{\Psi }_3,$$ (8) where $`A_1,A_2,A_3`$ are taken as extra variational parameters. Since $`\mathrm{\Psi }_{1,2,3}`$ are not orthogonal, the parameters $`A_{1,2,3}`$ do not have the usual meaning of weight factors. For the present calculations we use a linear superposition of the above interpolations (7), (8) given by: $$\mathrm{\Psi }_6=A_1\mathrm{\Psi }_1+A_2\mathrm{\Psi }_2+A_3\mathrm{\Psi }_3+A_4\mathrm{\Psi }_{4nls},$$ (9) where $`A_1,A_2,A_3,A_4`$ are considered again as extra variational parameters with no meaning of weight factors. The function (9) combines in a single function form, a suitable description of the system for the different physical regimes (near equilibrium and large internuclear distances for strong and very strong magnetic fields). Therefore we guess it should provide a relevant approximation for the ground state of the system. This was indeed the case of the linear parallel configuration of $`H_3^{(2+)}`$ (see Ref. ). This function has in total fourteen variational parameters. The variational procedure is carried out using the standard minimization package MINUIT from CERN-LIB. We use the integration routine D01FCF from NAG-LIB. All integrals are calculated with relative accuracy $`10^8`$. The results of our calculations are presented in Table I. For all considered magnetic fields $`B=10^{11}4.414\times 10^{13}G`$ we found a minimum in the potential energy curve as a function of the internuclear distance $`R`$ (see below Fig. 3) indicating the formation of a bound state. For such magnetic fields the total energy of the triangular configuration is always larger than the total energy of the corresponding linear parallel configuration. Indeed this shows that for very strong magnetic fields the linear configuration is the most favored. However, a new striking unexpected phenomenon appears when one considers a triangular configuration. While for the linear configuration the binding energyThe binding energy is defined as the affinity to keep the electron bound $`E_b=BE_T`$. $`B`$ is given in $`Ry`$ and thus has a meaning of the energy of free electron in magnetic field. increases and the internuclear distance decreases as the magnetic field grows, the opposite is true for the triangular configuration. Namely, the binding energy is slowly decreasing and the internuclear distance is slowly increasing (!) both being almost constant. A similar behavior occurs with the longitudinal localization length of the electron $`|z|`$ which also slowly increases in the triangular configuration with a magnetic field growth, opposite to what happens in the linear configuration where it decreases rather sharply. The transversal size of the electronic cloud $`\rho `$ is very close to the corresponding cyclotron radius $`B^{1/2}(a.u.)`$ for both cases. We do not have a clear physical picture of this phenomenon and it needs further consideration. A surprising feature is revealed if we study different trial functions. In Table II it is shown the variational results for the total energy $`E_T`$ for two quite arbitrarily chosen values of the internuclear distance, $`R=0.45`$ and $`R=2.527`$. For magnetic fields $`10^{11}B10^{12}G`$ an adequate description of the system for internuclear distances $`R0.45`$ is given by the trial function (6) modelling a ionic coupling, $`H_2^++`$proton. In this case, the single function (6) gives the lowest total energy among the trial functions $`\mathrm{\Psi }_{1,2,3,4}`$. On the other hand, the trial function (2) modelling a covalent coupling gives the lowest total energy for $`R2.527`$ giving a hint that a covalent coupling dominates when the internuclear distance is large compared to the equilibrium distance. Amazingly, for $`B10^{13}G`$ the variational energy seems to be insensitive to the specific form of the trial function (!). The electronic density distribution $`|\mathrm{\Psi }|^2`$ in the plane $`z=0`$ is shown in Fig. 2. For all magnetic fields $`B=10^{11}4.414\times 10^{13}G`$ the electronic density distribution $`|\mathrm{\Psi }|^2`$ exhibits a single maximum located at the center of the triangle $`(x=0,y=0)`$. As the magnetic field grows the electronic cloud becomes more concentrated around $`x=0,y=0`$, while its distance to the protons is slowly increased. From our results we can draw the conclusion that the exotic molecular ion $`H_3^{(2+)}`$ can exist in the presence of a strong magnetic field in a triangular configuration of the charged centers which is unstable towards a decay mode to the linear configuration. In particular, in the present article we have shown that the molecular ion $`H_3^{(2+)}`$ can also exist in the equilateral (fixed nuclei) triangular configuration perpendicular to the magnetic field. The results of our variational study for the ground state show that, contrary to the case of the linear parallel configuration, the binding energy of the system decreases, and the internuclear distance increases, as the magnetic field grows from $`B=10^{11}G`$ up to $`4.414\times 10^{13}G`$. A more extended study of this as well as other configurations is needed in order to have a better understanding of the properties of the exotic system $`H_3^{(2+)}`$. In particular, a proper contribution coming from the linear Zeeman effect term in (1) has to be taken into account. However, the fact that the linear Zeeman effect contribution vanishes in our consideration (which occurs when real trial functions are considered) looks relevant in the region of very strong magnetic fields $`B=10^{11}4.414\times 10^{13}G`$ where this term can be neglected anyhow. Although in general it is not clear so far (see discussion in ) how an adiabatic separation of the electronic and the nuclear motion can be performed in the presence of a strong external magnetic field, our fixed-nuclei approach is a good starting point to study of the exotic molecular ion $`H_3^{(2+)}`$. ###### Acknowledgements. The author wish to express his deep gratitude to Alexander Turbiner for introducing to the subject and for numerous valuable discussions. This work is supported in part by DGAPA Grant # IN105296 (México).
warning/0001/hep-ph0001067.html
ar5iv
text
# hep-ph/0001067 Electromagnetic properties of a neutrino stream ## I Introduction and Conclusions From a modern point of view, the methods of finite temperature field theory (FTFT) provide a natural setting to treat the problems related to the propagation of photons and neutrinos through a dense medium. This view has been largely stimulated by the work of Weldon, who emphasized the convenience of carrying out covariant, real-time calculations in this kind of problem. The work of Weldon demonstrated that the real-time formulation of FTFT is well suited to the study of systems involving gauge fields and/or chiral fermions at finite temperature, which can be extended in an efficient and transparent way to realistic situations involving, for example, photons and/or neutrinos in a matter background. The electromagnetic properties of neutrinos in a medium, besides their intrinsic interest, are relevant in many physical applications. For example, the induced electromagnetic couplings of a neutrino propagating in a background of electrons and nucleons is responsible for the plasmon decay process $`\gamma \nu \overline{\nu }`$ in stars, and modify the MSW resonant condition in the presence of an external magnetic field. A neutrino gas also exhibits the phenomemon of optical activity as a result of the chiral nature of the neutrino interactions. The covariance in this type of calculation is implemented by introducing the velocity four-vector $`u^\mu `$ of the medium, in terms of which the thermal propagators are written in a manifestly covariant form. Therefore, covariance is maintained, but quantities such as the photon self-energy or the neutrino electromagnetic form factors depend not just on the kinematical momentum variables of the problem, but also on $`u^\mu `$. In practice the vector $`u^\mu `$ is in the end set to $`(1,\stackrel{}{0})`$, which is equivalent to having carried out the calculation from the start with respect to a frame of reference in which the medium is at rest. This is usually the relevant situation. Therefore, while generally useful, the covariant nature of these methods has not been of particular importance in the applications mentioned. However, there is a class of problem in which setting $`u^\mu =(1,\stackrel{}{0})`$ is not possible. These are problems that involve one stream of particles (which we can think of as a moving medium) flowing through a background medium, which we can take to be stationary. If we denote by $`u^\mu `$ the velocity four-vector of the stationary medium, and by $`u^\mu `$ the corresponding one for the moving background, then we can set $`u^\mu =(1,\stackrel{}{0})`$, but we cannot take both to be $`(1,\stackrel{}{0})`$ simultaneously. Thus, for example, if we were to calculate the self-energy of the photon propagating through such media, it will depend on the momentum and velocity four-vector $`u^\mu `$ as usual, and in addition on $`u^\mu `$. This additional dependence can have consequences that are as important as the effects of the stationary background itself. For example, it is well known that in a plasma in which a bunch of electrons move, as a whole, relative to a plasma at rest, besides the usual dispersion relation of the longitudinal photon mode, another branch appears whose dispersion relation depends on the velocity of the beam. Under some conditions, the sign of the imaginary part of this dispersion relation is such that the corresponding amplitude grows exponentially, which signifies an instability of the system against the excitation of those modes. This kind of system is familiar in plasma physics research, and examples of them are discussed in textbooks on the subject. Recently, it has been suggested that a similar kind of streaming stability might be driven by a flow of neutrinos through a matter background. Because the neutrino acquires an effective electromagnetic coupling as it traverses a medium, the propagation of a photon in a medium that contains a drifting neutrino background may be affected in a way similar to the case mentioned above. As argued in Ref. , such effects can appear under the conditions of realistic situations such as those in a supernova explosion, gamma ray bursts, or the Early Universe. Similarly, other neutrino processes that have been studied previously, such as those mentioned above, may be modified in important ways if the neutrino gas is moving as a whole relative to the matter background. Motivated by all these considerations, in this work we calculate the neutrino contribution to the photon self-energy in a medium in which the neutrino gas moves as a whole relative to a matter background which we take to be at rest. The calculation is based on the application of FTFT to this problem in the manner that has been suggested above. The implicit assumption is that, in the rest frame of the stream, the neutrino background is characterized by a momentum distribution function that is parametrized in the usual way. Although our focus is the case in which the neutrino background constitutes the stream, largely motivated by the potential applications that have been mentioned, the calculation and the formulas for the photon self-energy are presented in such a way that they can be adapted to other cases as well. Therefore, they complement the existing calculations of the photon self-energy in which all the particles form a common background with a unique velocity four-vector. The results for the photon self-energy can be equivalently interpreted in terms of the dielectric constant of the system, and in that way we show that the well known textbooks results for the stream stabilities are reproduced when the appropriate limits are taken. On the other hand, the results we obtain are valid for general conditions (whether they are relativistic and/or degenerate or the converse) of the gases that form the plasma at rest as well as the stream, hold for general values of the velocity of the stream, and are valid also for general values of the photon momentum and not necessarily for some particular limit. Therefore, they are useful also in the study of similar processes that may occur in other contexts, such as high-temperature $`QCD`$, heavy ion collisions or other similar environments in which the methods of FTFT are applicable. In Section II, we give the general one-loop formulas for the generic contribution of a moving fermion background to the photon self-energy. The contribution from any given fermion can be written in terms of a few independent functions, which are expressed as integrals over the momentum distribution functions of the fermion. Explicit formulas are given for various limiting cases of physical interest, which also serve to show how some of the results derived in textbooks for simple cases are reproduced in the appropriate limit. The case of the system that is composed of a matter background made of electrons and nucleons (and possibly their antiparticles), and a neutrino gas that propagates as a whole relative to the matter background, is considered in Section III. We begin by reviewing the one-loop formulas for the effective electromagnetic vertex of the neutrino in a matter background, in the way that will be used in the calculation of the photon self-energy. The neutrino background contribution to the photon self-energy is then determined. It depends on the momentum distribution functions of the matter particles and well as those for the neutrinos. As an application of the formulas obtained, the dispersion relation for the longitudinal modes is considered, with attention to the possible effect of the neutrino contribution to the instability of the system. In that context, our results do not indicate the existence of unstable modes, and therefore we do not find support for the idea that stream instabilities due to the presence of the neutrino background can develop in such systems. In Section III we also consider the dispersion relations for the transverse modes. The chiral nature of the neutrino interactions gives rise to the phenomenon of optical activity, which had been studied earlier. Here we show how the results of Refs. are modified when the neutrino gas is moving relative to the matter background. The main effect is that the dispersion relations for the two circularly polarized modes are not isotropic. As a consequence, the frequency splitting between them, which is the measure of the rotation of the plane of polarization, depends on the direction of propagation of the mode relative to the velocity of the neutrino gas. In particular, under the appropriate conditions, the frequency difference is the opposite to what is found if the neutrino gas is not moving relative to the matter background. Section IV contains our outlook, where other possible effects and applications are also mentioned. ## II Photon self-energy in a fermion background We will consider a medium that consists of a gas of nucleons, electrons, neutrinos and their antiparticles. Each fermion gas gives a contribution to the elements of the $`2\times 2`$ photon self-energy matrix, that are determined by calculating the diagram shown in Fig. 1. In particular, the contribution to the $`\pi _{11\mu \nu }`$ element from each fermion $`f`$ in the loop is $$i\pi _{11\mu \nu }^{(f)}=(1)(i)^2\text{Tr}\frac{d^4p}{(2\pi )^4}j_{f\mu }^{(em)}(q)iS_{F11}^{(f)}(p+q)j_{f\nu }^{(em)}(q)iS_{F11}^{(f)}(p),$$ (1) where $`j_{f\mu }^{(em)}(q)`$ is the electromagnetic current of the fermion. It is defined in such a way that the on-shell matrix element of the electromagnetic current operator $`J_\mu ^{(em)}`$ is given by $$f(p^{})|J^{(em)}(0)_\mu |f(p)=\overline{u}(p^{})j_{f\mu }^{(em)}(q)u(p),$$ (2) where $`q=pp^{}`$ and $`u(p)`$ is a Dirac spinor. For the electron it is simply $`e\gamma _\mu `$, for the nucleons it must in principle include the magnetic moment term, and for the neutrino we must use the effective electromagnetic neutrino vertex in the medium. The fermion propagator that appears in Eq. (1) is given by $$S_{F11}^{(f)}(p)=(p\text{/}+m_f)\left[\frac{1}{p^2m_f^2+iϵ}+2\pi i\delta (p^2m_f^2)\eta _f(pu^{(f)})\right]$$ (3) where $$\eta _f(pu^{(f)})=\frac{\theta (pu^{(f)})}{e^{\beta _fpu^{(f)}\alpha _f}+1}+\frac{\theta (pu^{(f)})}{e^{\beta _fpu^{(f)}+\alpha _f}+1},$$ (4) with $`\beta _f`$ being the inverse temperature and $`\alpha _f`$ the fermion chemical potential. The vector $`u^{(f)\mu }`$ is the velocity four-vector of the fermion gas as a whole, so that $`u^{(f)\mu }=(1,\stackrel{}{0})`$ if the fermion background is at rest. In Eq. (4) we are allowing for the possibility that the different fermion gases of the background may be at different temperatures and, most importantly for our purposes later, that each gas has a velocity four-vector that is not necessarily the same for all of them. The implicit assumption here is that, in the rest frame of each fermion background, the corresponding particles have an isotropic thermal distribution characterized by a temperature and chemical potential $`1/\beta _f`$ and $`\alpha _f`$. The dispersion relations of the propagating photon modes are obtained by solving the equation $$(q^2g_{\mu \nu }q^\mu q^\nu \pi _{\mu \nu }^{(eff)})A^\nu =0,$$ (5) where $$\text{Re}\pi _{\mu \nu }^{(eff)}=\underset{f}{}\text{Re}\pi _{\mu \nu }^{(f)},$$ (6) and we have denoted by $`\pi _{\mu \nu }^{(f)}`$ the background-dependent term of Eq. (1). In the rest of this paper we will focus only on the real part of the dispersion relations, but similar considerations could be used to calculate the imaginary part as well. In order to calculate $`\text{Re}\pi _{\mu \nu }^{(eff)}`$, and thus determine the dispersion relations, we must know the composition of the medium and the formulas for the electromagnetic current that must be substituted in Eq. (1). To proceed, we consider the various cases separately. ### A Matter background We consider first an isotropic medium composed of nucleon and electron gases, with a common velocity four-vector $`u^\mu `$. The most general form of the physical self-energy function in this case, which we denote by $`\pi _{\mu \nu }^{(m)}`$ is $$\pi _{\mu \nu }^{(m)}=\pi _T^{(m)}R_{\mu \nu }(q,u)+\pi _L^{(m)}Q_{\mu \nu }(q,u)+\pi _P^{(m)}P_{\mu \nu }(q,u),$$ (7) where $`R_{\mu \nu }(q,u)`$ $`=`$ $`g_{\mu \nu }{\displaystyle \frac{q^\mu q^\nu }{q^2}}Q_{\mu \nu }(q,u)`$ (8) $`Q_{\mu \nu }(q,u)`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{u}_\mu \stackrel{~}{u}_\nu }{\stackrel{~}{u}^2}}`$ (9) $`P_{\mu \nu }(q,u)`$ $`=`$ $`{\displaystyle \frac{i}{𝒬}}ϵ_{\mu \nu \alpha \beta }q^\alpha u^\beta ,`$ (10) with $$\stackrel{~}{u}_\mu \left(g_{\mu \nu }\frac{q_\mu q_\nu }{q^2}\right)u^\nu .$$ (11) Although we have not indicated it explicitly, in general, $`\pi _{T,L,P}^{(m)}`$ are functions of the scalar variables $`\mathrm{\Omega }`$ $`=`$ $`qu`$ (12) $`Q`$ $`=`$ $`\sqrt{\mathrm{\Omega }^2q^2},`$ (13) which have the interpretation of being the photon energy and momentum in the rest frame of the medium. A detailed calculation of the photon self-energy in such a medium was carried out in Ref. . For our present purposes, it is useful to summarize those results as follows. The nucleon magnetic moment term contribution is not important for practical purposes. Therefore, we use here $`j_{f\mu }^{(em)}=e_f\gamma _\mu `$, so that the neutron contribution is being neglected. Substituting in Eq. (1) the formula for $`S_{F11}^{(f)}`$, the contribution from each fermion in the loop can be expressed in the form $`\text{Re}\pi _{\mu \nu }^{(f)}`$ $`=`$ $`4e^2\left[{\displaystyle \frac{1}{2}}\left(A_f(\mathrm{\Omega },𝒬){\displaystyle \frac{B_f(\mathrm{\Omega },𝒬)}{\stackrel{~}{u}^2}}\right)R_{\mu \nu }(q,u)+{\displaystyle \frac{B_f(\mathrm{\Omega },𝒬)}{\stackrel{~}{u}^2}}Q_{\mu \nu }(q,u)\right],`$ (14) with the coefficients $`A_f`$ and $`B_f`$ defined as $`A_f(\mathrm{\Omega },𝒬)`$ $`=`$ $`{\displaystyle \frac{d^3p}{(2\pi )^32E_f}\left(f_f(pu)+f_{\overline{f}}(pu)\right)\left[\frac{2m_f^22pq}{q^2+2pq}+(qq)\right]}`$ (15) $`B_f(\mathrm{\Omega },𝒬)`$ $`=`$ $`{\displaystyle \frac{d^3p}{(2\pi )^32E_f}\left(f_f(pu)+f_{\overline{f}}(pu)\right)\left[\frac{2(pu)^2+2(pu)(qu)pq}{q^2+2pq}+(qq)\right]}.`$ (16) In these formulas, $$p^\mu =(E,\stackrel{}{p}),E_f=\sqrt{\stackrel{}{p}^{\mathrm{\hspace{0.17em}2}}+m_f^2},$$ (17) and $`f_{f,\overline{f}}`$ denote the particle and antiparticle number density distributions $$f_{f,\overline{f}}()=\frac{1}{e^{\beta _f\alpha _f}+1}$$ (18) with the minus(plus) sign holding for the particles(antiparticles), respectively. Comparing Eqs. (7) and (14) we can identify the contribution of any fermion to the real part of the transverse and longitudinal components of the self-energy, and therefore obtain $`\text{Re}\pi _T^{(m)}`$ $`=`$ $`2e^2{\displaystyle \underset{f}{}}\left(A_f(\mathrm{\Omega },𝒬)+{\displaystyle \frac{q^2}{Q^2}}B_f(\mathrm{\Omega },𝒬)\right),`$ (19) $`\text{Re}\pi _L^{(m)}`$ $`=`$ $`4e^2{\displaystyle \underset{f}{}}{\displaystyle \frac{q^2}{Q^2}}B_f(\mathrm{\Omega },𝒬),`$ (20) where the relation $`\stackrel{~}{u}^2=𝒬^2/q^2`$ has been used. ### B Matter background and a stream of charged particles We now consider a medium that contains, in addition to the background as has been considered above, another gas of particles with a velocity four-vector $`u_\mu ^{}`$. We will refer to them as the matter background and the stream, respectively, and we assume that the latter consists of only one specie of fermions $`f^{}`$ with an electromagnetic coupling $`j_{f^{}\mu }^{(em)}=e_f^{}\gamma _\mu `$. The fermion $`f^{}`$ could be, for example, the electron or any other charged particle. We will denote by $`U^{\mathrm{\hspace{0.17em}0}}`$ and $`\stackrel{}{U}^{}`$ the components of $`u^\mu `$ in the rest frame of the matter background so that, in that frame, $$u^\mu =(1,\stackrel{}{0}),u^\mu =(U^{\mathrm{\hspace{0.17em}0}},\stackrel{}{U}^{}).$$ (21) The contribution from $`f^{}`$ to the photon-self-energy is given by a formula analogous to Eq. (14), $$\pi _{\mu \nu }^{(f^{})}=\pi _T^{(f^{})}R_{\mu \nu }(q,u^{})+\pi _L^{(f^{})}Q_{\mu \nu }(q,u^{}),$$ (22) where $`\text{Re}\pi _T^{(f^{})}`$ $`=`$ $`2e_f^{}^2\left(A_f^{}(\mathrm{\Omega }^{},𝒬^{})+{\displaystyle \frac{q^2}{𝒬^{\mathrm{\hspace{0.17em}2}}}}B_f^{}(\mathrm{\Omega }^{},𝒬^{})\right)`$ (23) $`\text{Re}\pi _L^{(f^{})}`$ $`=`$ $`4e_f^{}^2{\displaystyle \frac{q^2}{𝒬^{\mathrm{\hspace{0.17em}2}}}}B_f^{}(\mathrm{\Omega }^{},𝒬^{}).`$ (24) In Eq. (23) the functions $`A_f^{}`$ and $`B_f^{}`$ are given by formulas analogous to Eq. (15), but with the replacement $`u^\mu u^\mu `$, and we have used $`\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}=𝒬^{\mathrm{\hspace{0.17em}2}}/q^2`$ where, similarly to Eq. (11), $$\stackrel{~}{u}_\mu ^{}=\left(g_{\mu \nu }\frac{q_\mu q_\nu }{q^2}\right)u^\nu .$$ (25) In analogy with Eq. (12), the variables $`\mathrm{\Omega }^{},𝒬^{}`$ are defined by $`\mathrm{\Omega }^{}`$ $`=`$ $`qu^{}`$ (26) $`Q^{}`$ $`=`$ $`\sqrt{\mathrm{\Omega }^{\mathrm{\hspace{0.17em}2}}q^2},`$ (27) and they are expressed in terms of $`\mathrm{\Omega }`$ and $`𝒬`$ by the relations $`\mathrm{\Omega }^{}`$ $`=`$ $`U^{\mathrm{\hspace{0.17em}0}}\mathrm{\Omega }\stackrel{}{U}^{}\stackrel{}{𝒬}`$ (28) $`𝒬^{}`$ $`=`$ $`\sqrt{\left(U^{\mathrm{\hspace{0.17em}0}}\mathrm{\Omega }\stackrel{}{U}^{}\stackrel{}{𝒬}\right)^2\mathrm{\Omega }^2+𝒬^2}.`$ (29) The total photon self-energy is given by $`\pi _{\mu \nu }^{(eff)}`$ $`=`$ $`\pi _T^{(m)}R_{\mu \nu }(q,u)+\pi _L^{(m)}Q_{\mu \nu }(q,u)+\pi _T^{(f^{})}R_{\mu \nu }(q,u^{})+\pi _L^{(f^{})}Q_{\mu \nu }(q,u^{}).`$ (30) Eq. (30) can be written in a convenient form by the following procedure. From the definition of $`R_{\mu \nu }`$ given in Eq. (8) we have $$R_{\mu \nu }(q,u^{})=R_{\mu \nu }(q,u)+Q_{\mu \nu }(q,u)Q_{\mu \nu }(q,u^{}).$$ (31) We now define the vectors $$e_1^\mu \frac{R^{\mu \nu }(q,u)u_\nu ^{}}{\sqrt{N_1}},e_2^\mu iP^{\mu \nu }(q,u)e_{1\nu }.$$ (32) with $`N_1`$ $`=`$ $`u_\mu ^{}u_\nu ^{}R^{\mu \nu }(q,u)`$ (33) $`=`$ $`{\displaystyle \frac{(\stackrel{~}{u}u^{})^2}{\stackrel{~}{u}^2}}\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}},`$ (34) which can be expressed in the form $$N_1=U^{\mathrm{\hspace{0.17em}2}}𝒬^2\left(\stackrel{}{U}^{}\stackrel{}{𝒬}\right)^2.$$ (35) It is easy to verify that $`e_{1,2}^\mu `$ are mutually orthogonal and satisfy $$e_{1,2}q=e_{1,2}\stackrel{~}{u}=0,e_{1,2}^2=1.$$ (36) Thus, together with $`\stackrel{~}{u}^\mu `$, they form a complete set transverse to $`q^\mu `$, and therefore it is possible to express $`\stackrel{~}{u}^\mu `$ in terms of them. The desired relation, which follows from substituting Eq. (8) into Eq. (32), is $$\stackrel{~}{u}_\mu ^{}=\sqrt{N_1}e_{1\mu }+\left(\frac{\stackrel{~}{u}u^{}}{\stackrel{~}{u}^2}\right)\stackrel{~}{u}_\mu ,$$ (37) which substituting in the definitions given in Eq. (8) yields the convenient formulas $`Q_{\mu \nu }(q,u^{})`$ $`=`$ $`{\displaystyle \frac{N_1}{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}e_{1\mu }e_{1\nu }+\left({\displaystyle \frac{N_1}{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}+1\right)Q_{\mu \nu }(u,q)+\sqrt{N_1}{\displaystyle \frac{\stackrel{~}{u}u^{}}{\stackrel{~}{u}^2\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}(e_{1\mu }\stackrel{~}{u}_\nu +\stackrel{~}{u}_\mu e_{1\nu })`$ (38) $`P_{\mu \nu }(q,u^{})`$ $`=`$ $`{\displaystyle \frac{𝒬\stackrel{~}{u}u^{}}{𝒬^{}\stackrel{~}{u}^2}}P_{\mu \nu }(q,u)+{\displaystyle \frac{i𝒬\sqrt{N_1}}{𝒬^{}\stackrel{~}{u}^2}}\left(\stackrel{~}{u}_\mu e_{2\nu }\stackrel{~}{u}_\nu e_{2\mu }\right).`$ (39) On the other hand, as shown in Ref. , $`R_{\mu \nu }`$ can be decomposed in the form $$R_{\mu \nu }(q,u)=(e_1^\mu e_1^\nu +e_2^\mu e_2^\nu ).$$ (40) In this way, using Eqs. (31) and (38) in Eq. (30) together with the decomposition given in Eq. (40), we finally arrive at $`\pi _{\mu \nu }^{(eff)}`$ $`=`$ $`e_{1\mu }e_{1\nu }\left[\pi _T^{(m)}+\pi _T^{(f^{})}{\displaystyle \frac{N_1}{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}\left(\pi _L^{(f^{})}\pi _T^{(f^{})}\right)\right]e_{2\mu }e_{2\nu }\left(\pi _T^{(m)}+\pi _T^{(f^{})}\right)`$ (43) $`+{\displaystyle \frac{\stackrel{~}{u}_\mu \stackrel{~}{u}_\nu }{\stackrel{~}{u}^2}}\left[\pi _L^{(m)}+\pi _L^{(f^{})}+{\displaystyle \frac{N_1}{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}\left(\pi _L^{(f^{})}\pi _T^{(f^{})}\right)\right]+\sqrt{N_1}{\displaystyle \frac{(\stackrel{~}{u}u^{})}{\stackrel{~}{u}^2\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}\left(\pi _L^{(f^{})}\pi _T^{(f^{})}\right)\left(e_{1\mu }\stackrel{~}{u}_\nu +\stackrel{~}{u}_\mu e_{1\nu }\right).`$ Eq. (43), besides unfolding the main structure of the modes in a particularly simple way, is a useful formula that allows us to obtain the dispersion relation of the modes under a variety of conditions. In the absence of the stream, the solutions consist of one longitudinal mode with polarization vector $`e_3^\mu \stackrel{~}{u}^\mu `$, and two degenerate transverse modes with polarization vectors $`e_{1,2}^\mu `$ that satisfy $`Q_{\mu \nu }e_{1,2}^\nu =0`$. Their dispersion relations are determined by solving the equations $`q^2=\text{Re}\pi _{L,T}^{(m)}`$ for the longitudinal and transverse modes, respectively. The presence of the stream breaks the degeneracy of the transverse modes, and in general causes a mixing between them with the longitudinal mode. In those cases in which it is permissible to treat the mixing term (the last term in Eq. (43)) as a perturbation (e.g., the number density in the stream is sufficiently smaller than those in the matter background), the dispersion relations are obtained approximately by solving $`q^2`$ $`=`$ $`\pi _T^{(m)}+\pi _T^{(f^{})}+{\displaystyle \frac{N_1q^2}{𝒬^{\mathrm{\hspace{0.17em}2}}}}\left(\pi _L^{(f^{})}\pi _T^{(f^{})}\right)`$ (44) $`q^2`$ $`=`$ $`\pi _T^{(m)}+\pi _T^{(f^{})}`$ (45) $`q^2`$ $`=`$ $`\pi _L^{(m)}+\pi _L^{(f^{})}{\displaystyle \frac{N_1q^2}{𝒬^{\mathrm{\hspace{0.17em}2}}}}\left(\pi _L^{(f^{})}\pi _T^{(f^{})}\right).`$ (46) with corresponding polarization vectors $`e_{1,2}`$ and $`e_3\stackrel{~}{u}`$, respectively. In Eq. (44) we have used the relation $`\stackrel{~}{u}^2=𝒬^2/q^2`$ and the corresponding one for $`\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}`$. If the mixing term is not sufficiently small so that the higher order terms are important, then the full $`2\times 2`$ problem in the $`e_1\stackrel{~}{u}`$ plane must be considered which, although tedious, is straightforward. In the equations Eq. (44), it is understood that the variables $`\mathrm{\Omega }^{},𝒬^{}`$ are to be expressed in terms of $`\mathrm{\Omega },𝒬`$ by means of the relations given in Eq. (28). They thus become implicit equations for $`\mathrm{\Omega },𝒬`$, whose solutions determine the dispersion relations $`\mathrm{\Omega }(𝒬)`$ of the various modes. ### C Discussion For illustrative purposes, we consider the specific case of a stream of electrons and a matter background that consists of an electron gas and a non-relativistic proton gas. We borrow from Ref. \[see Eqs. (A5) and (A9)\] the following results $`B_f(\mathrm{\Omega },𝒬)`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^3𝒫}{(2\pi )^3}\frac{\stackrel{}{𝒬}_𝒫(f_f()+f_{\overline{f}}())}{\mathrm{\Omega }\stackrel{}{v}_𝒫\stackrel{}{𝒬}}},`$ (47) $`A_f(\mathrm{\Omega },𝒬)`$ $`=`$ $`B_f(\mathrm{\Omega },𝒬)+{\displaystyle \frac{\mathrm{\Omega }}{2}}{\displaystyle \frac{d^3𝒫}{(2\pi )^3}\frac{\stackrel{}{v}_𝒫_𝒫(f_f()+f_{\overline{f}}())}{\mathrm{\Omega }\stackrel{}{v}_𝒫\stackrel{}{𝒬}}},`$ (48) where $`=\sqrt{\stackrel{}{𝒫}^2+m_f^2}`$, $`_𝒫`$ is the gradient operator with respect to the momentum $`\stackrel{}{𝒫}`$ and $`\stackrel{}{v}_𝒫=\stackrel{}{𝒫}/`$. As shown there, they are valid for values of $`q`$ such that $$q/1,$$ (49) where $``$ stands for a typical average value of the energy of the fermions in the gas. For distribution functions that depend on $`\stackrel{}{𝒫}`$ only through $``$, we can replace $`_𝒫\stackrel{}{v}_𝒫\frac{}{}`$ in Eq. (47). Several useful formulas follow from Eq. (47) in particular cases. For example, if the fermions are relativistic, $`A_f(\mathrm{\Omega },𝒬)`$ $`=`$ $`3\omega _{0f}^2`$ (50) $`B_f(\mathrm{\Omega },𝒬)`$ $`=`$ $`3\omega _{0f}^2\left(1{\displaystyle \frac{\mathrm{\Omega }}{2𝒬}}\mathrm{ln}\left|{\displaystyle \frac{\mathrm{\Omega }+𝒬}{\mathrm{\Omega }𝒬}}\right|\right).`$ (51) Eq. (50) holds for a degenerate or non-degenerate gas. For the non-relativistic and non-degenerate case we use $`A_f(\mathrm{\Omega },𝒬)`$ $`=`$ $`3\omega _{0f}^2+{\displaystyle \frac{𝒬^2\omega _{0f}^2}{\mathrm{\Omega }^2}}`$ (52) $`B_f(\mathrm{\Omega },𝒬)`$ $`=`$ $`{\displaystyle \frac{𝒬^2\omega _{0f}^2}{\mathrm{\Omega }^2}},`$ (53) which are valid if, in addition to Eq. (49), $$\mathrm{\Omega }\overline{v}_f𝒬,$$ (54) where $`\overline{v}_f1/\sqrt{\beta _fm_f}`$ is a typical value of the velocity of the fermions in the gas. The quantity $`\omega _{0f}^2`$, which is related to the plasma frequency in the gas, is given by $$\omega _{0f}^2=\frac{d^3𝒫}{(2\pi )^32}(f_f()+f_{\overline{f}}())\left[1\frac{𝒫^2}{3^2}\right].$$ (55) In Eqs. (50) and (52) we have used its form in the relativistic (ER) limit and non-relativistic (NR) limits, $$\omega _{0f}^2=\{\begin{array}{cc}\frac{1}{6\pi ^2}_0^{\mathrm{}}𝑑𝒫𝒫(f_f(𝒫)+f_{\overline{f}}(𝒫))\hfill & \text{(ER)}\hfill \\ \frac{n_f}{4m_f}\hfill & \text{(NR)},\hfill \end{array}$$ (56) where $`n_f`$ is the fermion number density in the frame in which the background is at rest. The corresponding formulas for the non-relativistic and degenerate case are given in Ref. . Under most circumstances, the protons make a negligible contribution to the dispersion relations. The conditions under which those terms can be important are given in Ref. . Here we do not include those special cases and therefore we will not consider the protons further. Lastly, we assume that the stream is not moving too fast as a whole, so that the term that is proportional to $`N_1`$ in Eq. (44) can be neglected, since according to Eq. (35) it is or the order of the velocity squared of the stream. With these assumptions, the dispersion relations become $`q^2`$ $`=`$ $`2e^2\left[\left(A_e(\mathrm{\Omega },𝒬)+{\displaystyle \frac{q^2}{𝒬^2}}B_e(\mathrm{\Omega },𝒬)\right)+\left(A_e^{}(\mathrm{\Omega }^{},𝒬^{})+{\displaystyle \frac{q^2}{𝒬^{\mathrm{\hspace{0.17em}2}}}}B_e^{}(\mathrm{\Omega }^{},𝒬^{})\right)\right]`$ (57) $`q^2`$ $`=`$ $`4e^2\left[{\displaystyle \frac{q^2}{𝒬^2}}B_e(\mathrm{\Omega },𝒬)+{\displaystyle \frac{q^2}{𝒬^{\mathrm{\hspace{0.17em}2}}}}B_e^{}(\mathrm{\Omega }^{},𝒬^{})\right],`$ (58) for the transverse and longitudinal modes, respectively. We now consider several cases separately. #### 1 Non-relativistic matter and stream electrons For the electrons in the matter background we use Eq. (52). Similarly, for the stream $`A_e^{}(\mathrm{\Omega }^{},𝒬^{})`$ $`=`$ $`3\omega _{0e^{}}^2+{\displaystyle \frac{𝒬^{\mathrm{\hspace{0.17em}2}}\omega _{0e^{}}^2}{\mathrm{\Omega }^{\mathrm{\hspace{0.17em}2}}}}`$ (59) $`B_e^{}(\mathrm{\Omega }^{},𝒬^{})`$ $`=`$ $`{\displaystyle \frac{𝒬^{\mathrm{\hspace{0.17em}2}}\omega _{0e^{}}^2}{\mathrm{\Omega }^{\mathrm{\hspace{0.17em}2}}}},`$ (60) which, as we will see, are suitable for finding the long wavelength limit of the dispersion relations. Eq. (59) can be used if the solution is such that $$\mathrm{\Omega }^{}\overline{v}_e^{}𝒬^{}.$$ (61) The conditions under which the solution thus found is valid can be ascertained afterwards. The formula for $`\omega _{0e^{}}^2`$ is the same expression given in Eq. (55), but with replacement $`f_{f,\overline{f}}()f_{e^{},\overline{e}^{}}()`$, where $$f_{e^{},\overline{e}^{}}()=\frac{1}{e^{\beta _e^{}\alpha _e^{}}+1}.$$ (62) As we have indicated previously, the implicit assumption that is being made here is that the electrons that compose the stream have, in the rest frame of the stream, an isotropic thermal distribution characterized by a temperature and chemical $`1/\beta _e^{}`$ and $`\alpha _e^{}`$. Let us consider the dispersion relation for the longitudinal mode. From Eqs. (19) and (23) this yields $$1=4e^2\left(\frac{\omega _{0e}^2}{\mathrm{\Omega }^2}+\frac{\omega _{0e^{}}^2}{\mathrm{\Omega }^{\mathrm{\hspace{0.17em}2}}}\right).$$ (63) From Eq. (26) $$\mathrm{\Omega }^{}=\mathrm{\Omega }\stackrel{}{𝒬}\stackrel{}{U}^{}$$ (64) using $`U^{\mathrm{\hspace{0.17em}0}}1`$, and therefore the dispersion relation is $$(\mathrm{\Omega }\stackrel{}{𝒬}\stackrel{}{U}^{})^2(\mathrm{\Omega }^24e^2\omega _{0e}^2)4e^2\omega _{0e^{}}^2\mathrm{\Omega }^2=0.$$ (65) The salient feature of Eq. (65) is that, besides the usual solution $`\mathrm{\Omega }_L^2(𝒬0)4e^2\omega _{0e}^2`$, there is another one with $`\mathrm{\Omega }_L\stackrel{}{U}^{}\stackrel{}{𝒬}`$. The standard way to find this second solution is to substitute $$\mathrm{\Omega }=\stackrel{}{U}^{}\stackrel{}{𝒬}+\delta _L$$ (66) in Eq. (65) and determine $`\delta _L`$ approximately by taking it to be a small quantity. In this fashion, we find $$\delta _L=\pm \frac{|2e\omega _{0e^{}}\stackrel{}{U}^{}\stackrel{}{𝒬}|}{\sqrt{(\stackrel{}{U}^{}\stackrel{}{𝒬})^24e^2\omega _{0e}^2}},$$ (67) which shows the well-known instability of this system. For values of $`\stackrel{}{U}^{}\stackrel{}{𝒬}`$ such that $$0<|\stackrel{}{U}^{}\stackrel{}{𝒬}|<2|e|\omega _{0e},$$ (68) the dispersion relation has a solution with a positive imaginary part, which signals that the system is unstable against oscillations with those values of $`\stackrel{}{U}^{}\stackrel{}{𝒬}`$. The condition that $`\delta _L`$ be small relative to $`\stackrel{}{U}^{}\stackrel{}{𝒬}`$ is satisfied for sufficiently small values of $`\omega _{0e^{}}/\omega _{0e}`$. On the other hand notice that, for this solution, $`\mathrm{\Omega }^{}=\delta _L`$, and $`𝒬^{}=\sqrt{\mathrm{\Omega }^{\mathrm{\hspace{0.17em}2}}\mathrm{\Omega }^2+𝒬^2}𝒬`$. The conditions given in Eqs. (49) and (61) are then equivalent to $`|\stackrel{}{U}^{}\stackrel{}{𝒬}|\overline{v}_e𝒬`$ and $`\delta \overline{v}_e^{}𝒬`$ which, for sufficiently small values of the thermal velocities, are satisfied as well. Turning now the attention to the transverse dispersion relation, we substitute the formulas for $`A_{e,e^{}}`$ and $`B_{e,e^{}}`$ given in Eqs. (52) and (59) into Eq. (57). This yields simply $$q^2=4e^2\omega _{0e}^2+4e^2\omega _{0e^{}}^2,$$ (69) which shows that in this case the presence of the stream perturbs somewhat the transverse dispersion relation by shifting the value of the plasma frequency, but it does not produce a significant effect otherwise. Eqs. (63) and (69) reproduce the well-known results found in textbooks, which are derived by kinetic theory or similar semi-classical methods. However, the results that we have obtained, and which are summarized in Eqs. (43) and (44), go farther. Together with the expressions for the self-energy functions in terms of the coefficients $`A_f`$ and $`B_f`$ \[Eqs. (19) and (23)\] they allow us to study systems under a wider variety of conditions, including those for which the semi-classical approaches, and the simple formula given in Eq. (63) in particular, are not be applicable. #### 2 Non-relativistic matter electrons and relativistic stream electrons For the electrons in the stream we must use in this case $`A_e^{}(\mathrm{\Omega }^{},𝒬^{})`$ $`=`$ $`3\omega _{0e^{}}^2`$ (70) $`B_e^{}(\mathrm{\Omega }^{},𝒬^{})`$ $`=`$ $`3\omega _{0e^{}}^2\left(1{\displaystyle \frac{\mathrm{\Omega }^{}}{2𝒬^{}}}\mathrm{ln}\left|{\displaystyle \frac{\mathrm{\Omega }^{}+𝒬^{}}{\mathrm{\Omega }^{}𝒬^{}}}\right|\right),`$ (71) while the matter electron formulas are the same as the previous ones. The dispersion relations are then determined by $`\mathrm{\Omega }^24e^2\omega _{0e}^2`$ $`=`$ $`4e^2\omega _{0e^{}}^2f_L`$ (72) $`\mathrm{\Omega }^2𝒬^24e^2\omega _{0e}^2`$ $`=`$ $`4e^2\omega _{0e^{}}^2f_T`$ (73) for the longitudinal and transverse modes, respectively, where we have defined $`f_L`$ $`=`$ $`{\displaystyle \frac{3\mathrm{\Omega }^2}{𝒬^{\mathrm{\hspace{0.17em}2}}}}\left[{\displaystyle \frac{\mathrm{\Omega }^{}}{2𝒬^{}}}\mathrm{ln}\left|{\displaystyle \frac{\mathrm{\Omega }^{}+𝒬^{}}{\mathrm{\Omega }^{}𝒬^{}}}\right|1\right]`$ (74) $`f_T`$ $`=`$ $`{\displaystyle \frac{3}{2}}\left\{1+{\displaystyle \frac{q^2}{𝒬^{\mathrm{\hspace{0.17em}2}}}}\left[1{\displaystyle \frac{\mathrm{\Omega }^{}}{2𝒬^{}}}\mathrm{ln}\left|{\displaystyle \frac{\mathrm{\Omega }^{}+𝒬^{}}{\mathrm{\Omega }^{}𝒬^{}}}\right|\right]\right\}.`$ (75) Let us consider the longitudinal dispersion relation. With the help of Fig. 2 it is easy to see that, besides the usual solution $`\mathrm{\Omega }_L4e^2\omega _{0e}`$, any other solution to Eq. (72) must have $`\mathrm{\Omega }𝒬`$ (which implies that $`\mathrm{\Omega }^{}𝒬^{}`$ and the function $`f_L`$ can be large) so that the stream term can compete with the matter term in that equation. Substituting $$\mathrm{\Omega }=𝒬+\delta _L$$ (76) in Eq. (72) we find $$\delta _L=\pm 2𝒬e^{2(𝒬^24e^2\omega _{0e}^2)/12e^2\omega _{0e^{}}^2}$$ (77) for $`𝒬^2>4e^2\omega _{0e}^2`$. In contrast to the case considered in Section II C 1, there is no sign that a stream instability may develop in the present one. For the transverse dispersion relations the situation is different. The function $`f_T`$ is not larger than a number of order 1, as shown in Fig. 2, so that the stream contribution in Eq. (73) only introduces a perturbation in the usual solution. In summary, when the stream consists of a relativistic electron gas, there is no sign that any stream instabilities may develop. This result will be a useful reference point when we consider in Section III the case in which the stream consists of neutrinos. ## III Photon self-energy in a neutrino stream In this section we consider a system composed of a matter background as in Section II B, and a neutrino stream with a velocity four-vector $`u^\mu `$. Our immediate task is to determine the neutrino stream contribution to the photon self-energy, for which we must calculate a diagram similar to the one in Fig. 1, but with a neutrino as the fermion in the loop. Denoting the effective electromagnetic vertex of the neutrino in the matter background by $`\mathrm{\Gamma }_\mu ^{(\nu )}(q)`$, then $$i\pi _{11\mu \nu }^{(\nu )}=(1)(i)^2\text{Tr}\frac{d^4p}{(2\pi )^4}\mathrm{\Gamma }_\mu ^{(\nu )}(q)iS_{F11}^{(\nu )}(p+q)\mathrm{\Gamma }_\nu ^{(\nu )}(q)iS_{F11}^{(\nu )}(p),$$ (78) where the neutrino propagator $`S_{F11}^{(\nu )}`$ is is given by Eq. (3) (with $`m_\nu `$ = 0). The neutrino effective vertex can be expressed in the form $$\mathrm{\Gamma }_\mu ^{(\nu )}(q)=T_{\mu \nu }(q)\gamma ^\nu L,$$ (79) where $`L=\frac{1}{2}(1\gamma _5)`$ as usual, and $`T_{\mu \nu }`$ can be decomposed as $$T_{\mu \nu }=T_TR_{\mu \nu }(q,u)+T_LQ_{\mu \nu }(q,u)+T_PP_{\mu \nu }(q,u).$$ (80) A detailed calculation of the various terms in Eq. (80) was carried out in Ref. . For our purposes here, we can summarize the main results obtained there as follows. For practical purposes, the contribution to $`T_{T,L}`$ due to the anomalous magnetic moment couplings of the nucleons in the background is negligible, so that $`T_T`$ $`=`$ $`2\sqrt{2}|e|G_Fa_p\left(A_p(\mathrm{\Omega },𝒬){\displaystyle \frac{B_p(\mathrm{\Omega },𝒬)}{\stackrel{~}{u}^2}}\right)+T_T^{(e)},`$ (81) $`T_L`$ $`=`$ $`4\sqrt{2}|e|G_Fa_p{\displaystyle \frac{B_p(\mathrm{\Omega },𝒬)}{\stackrel{~}{u}^2}}+T_L^{(e)},`$ (82) while $$T_P=T_P^{(e)}4\sqrt{2}G_Fb_p𝒬(|e|+2m_p\kappa _p)C_p(\mathrm{\Omega },𝒬)8m_n\kappa _n\sqrt{2}G_Fb_n𝒬C_n(\mathrm{\Omega },𝒬).$$ (83) In these formulas $`\kappa _p`$ $`=`$ $`1.79\left({\displaystyle \frac{|e|}{2m_p}}\right),`$ (84) $`\kappa _p`$ $`=`$ $`1.91\left({\displaystyle \frac{|e|}{2m_n}}\right),`$ (85) are the anomalous magnetic moment of the nucleons, the coefficients $`a_f`$ and $`b_f`$ are the neutral current couplings of the fermion $`f`$, while $`A_p`$ and $`B_p`$ are defined in Eq. (15) and $$C_f(\mathrm{\Omega },𝒬)=\frac{d^3p}{(2\pi )^32E}\left(\frac{\stackrel{~}{u}p}{\stackrel{~}{u}^2}\right)(f_ff_{\overline{f}})\left[\frac{1}{q^2+2pq}+(qq)\right].$$ (87) The electron terms $`T_X^{(e)}`$ were calculated in Ref. , and are given by $`T_T^{(e)}`$ $`=`$ $`2\sqrt{2}eG_F\left(A_e(\mathrm{\Omega },𝒬){\displaystyle \frac{B_e(\mathrm{\Omega },𝒬)}{\stackrel{~}{u}^2}}\right)\{\begin{array}{cc}a_e+1\hfill & (\nu _e)\hfill \\ & \\ a_e\hfill & (\nu _{\mu ,\tau })\hfill \end{array}`$ (91) $`T_L^{(e)}`$ $`=`$ $`4\sqrt{2}eG_F{\displaystyle \frac{B_e(\mathrm{\Omega },𝒬)}{\stackrel{~}{u}^2}}\{\begin{array}{cc}a_e+1\hfill & (\nu _e)\hfill \\ & \\ a_e\hfill & (\nu _{\mu ,\tau })\hfill \end{array}`$ (95) $`T_P^{(e)}`$ $`=`$ $`4\sqrt{2}eG_F𝒬C_e(\mathrm{\Omega },𝒬)\{\begin{array}{cc}b_e1\hfill & (\nu _e)\hfill \\ & \\ b_e\hfill & (\nu _{\mu ,\tau })\hfill \end{array}.`$ (99) Substituting Eq. (80) in Eq. (78), we then obtain for the neutrino stream contribution to the photon self-energy $$\text{Re}\pi _{\mu \nu }^{(\nu )}=2T_{\mu \alpha }(q)T_{\nu \beta }(q)J^{\alpha \beta }$$ (100) where $`J^{\alpha \beta }`$ $``$ $`{\displaystyle }{\displaystyle \frac{d^3p}{(2\pi )^32E}}\{(f_\nu (pu^{})+f_{\overline{\nu }}(pu^{}))[{\displaystyle \frac{2p^\alpha p^\beta p^\alpha q^\beta q^\alpha p^\beta +g^{\alpha \beta }pq}{q^22pq}}+(qq)]`$ (102) $`+(f_\nu (pu^{})+f_{\overline{\nu }}(pu^{}))iϵ^{\alpha \beta \lambda \rho }q_\lambda p_\rho [{\displaystyle \frac{1}{q^22pq}}+{\displaystyle \frac{1}{q^22pq}}]\}.`$ The transversality and symmetry properties of $`J^{\alpha \beta }`$ imply that it is expressible in terms of the tensors $`R^{\alpha \beta }(q,u^{})`$, $`Q^{\alpha \beta }(q,u^{})`$ and $`P^{\alpha \beta }(q,u^{})`$, with coefficients that can be determined by projecting Eq. (102) along these tensors. Thus we find $`J^{\alpha \beta }`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(A_\nu (\mathrm{\Omega }^{},𝒬^{}){\displaystyle \frac{B_\nu (\mathrm{\Omega }^{},𝒬^{})}{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}\right)R^{\alpha \beta }(q,u^{})`$ (105) $`+{\displaystyle \frac{B_\nu (\mathrm{\Omega }^{},𝒬^{})}{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}Q^{\alpha \beta }(q,u^{})+𝒬^{}C_\nu (\mathrm{\Omega }^{},𝒬^{})P^{\alpha \beta }(q,u^{}),`$ with the coefficients $`A_\nu (\mathrm{\Omega }^{},𝒬^{})`$, $`B_\nu (\mathrm{\Omega }^{},𝒬^{})`$ and $`C_\nu (\mathrm{\Omega }^{},𝒬^{})`$ defined in Eqs. (15) and (87). In Eq. (105) the tensors $`R_{\mu \nu }(q,u^{})`$, $`Q_{\mu \nu }(q,u^{})`$ and $`P_{\mu \nu }(q,u^{})`$ are eliminated using Eqs. (31) and (38), and $`R_{\mu \nu }(q,u)`$ is decomposed as in Eq. (40). In this way, $`J_{\alpha \beta }`$ $`=`$ $`\left\{{\displaystyle \frac{1}{2}}\left(A_\nu {\displaystyle \frac{B_\nu }{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}\right){\displaystyle \frac{N_1}{2\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}\left[{\displaystyle \frac{3B_\nu }{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}A_\nu \right]\right\}e_{1\alpha }e_{1\beta }{\displaystyle \frac{1}{2}}\left(A_\nu {\displaystyle \frac{B_\nu }{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}\right)e_{2\alpha }e_{2\beta }`$ (108) $`+\left\{{\displaystyle \frac{B_\nu }{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}+{\displaystyle \frac{N_1}{2\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}\left[{\displaystyle \frac{3B_\nu }{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}A_\nu \right]\right\}Q_{\alpha \beta }(q,u)+C_\nu 𝒬\left({\displaystyle \frac{\stackrel{~}{u}u^{}}{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}\right)P_{\alpha \beta }(q,u)`$ $`+{\displaystyle \frac{\sqrt{N_1}\stackrel{~}{u}u^{}}{2\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}\stackrel{~}{u}^2}}\left[{\displaystyle \frac{3B_\nu }{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}A_\nu \right]\left(e_{1\alpha }\stackrel{~}{u}_\beta +\stackrel{~}{u}_\alpha e_{1\beta }\right)+{\displaystyle \frac{iC_\nu \sqrt{N_1}𝒬}{\stackrel{~}{u}^2}}\left(\stackrel{~}{u}_\alpha e_{2\beta }e_{2\alpha }\stackrel{~}{u}_\beta \right).`$ By substituting Eq. (108) in Eq. (100) we finally obtain the formula for the neutrino contribution which, when it is added to $`\pi _{\mu \nu }^{(m)}`$ to obtain the total photon self-energy, is the starting point to determine the photon dispersion relations. However, with all its generality, the formula is not particularly useful and therefore we consider below some specific situations of potential interest. ### A Longitudinal dispersion relation As already seen in Section II B, in general the effects of the stream break the degeneracy of the transverse modes and also mixes them with the longitudinal one. When the latter effect is not too large, the longitudinal dispersion relation is obtained approximately by solving the equation $$q^2=\pi _L^{(m)}+\pi _l^{(\nu )},$$ (109) where $$\pi _l^{(\nu )}\frac{\stackrel{~}{u}^\mu \stackrel{~}{u}^\nu }{\stackrel{~}{u}^2}\pi _{\mu \nu }^{(\nu )}.$$ (110) Using the relation $`\stackrel{~}{u}^\mu T_{\mu \alpha }=T_L\stackrel{~}{u}_\alpha `$, together with $`{\displaystyle \frac{\stackrel{~}{u}^\mu \stackrel{~}{u}^\nu }{\stackrel{~}{u}^2}}R_{\mu \nu }(q,u^{})`$ $`=`$ $`{\displaystyle \frac{N_1}{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}`$ (111) $`{\displaystyle \frac{\stackrel{~}{u}^\mu \stackrel{~}{u}^\nu }{\stackrel{~}{u}^2}}Q_{\mu \nu }(q,u^{})`$ $`=`$ $`{\displaystyle \frac{N_1}{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}}+1,`$ (112) we obtain from Eq. (100) $$\text{Re}\pi _l^{(\nu )}=2T_L^2\frac{q^2}{𝒬^{\mathrm{\hspace{0.17em}2}}}\left\{B_\nu N_1\left[\frac{3q^2B_\nu }{𝒬^{\mathrm{\hspace{0.17em}2}}}+\frac{1}{2}A_\nu \right]\right\},$$ (113) where, to simplify the notation, we have omitted the arguments $`\mathrm{\Omega }^{}`$ and $`𝒬^{}`$ in the coefficients $`A_\nu `$ and $`B_\nu `$. The dispersion relation obtained by substituting Eqs. (19) and (113) in Eq. (109) is the same as the corresponding one for a stream of charged particles, with an effective charge $`e_\nu =\frac{1}{\sqrt{2}}T_L`$. #### 1 Long wavelength limit We consider the case analogous to the one discussed in Section II C, namely, a matter background made of a non-relativistic electron gas and a non-relativistic nucleon gas, and the neutrino stream. As in the case mentioned, the photon momentum is assumed to be such that Eq. (49) holds. For the electrons in the matter background we then use the formulas for $`A_e(\mathrm{\Omega },𝒬)`$ and $`B_e(\mathrm{\Omega },𝒬)`$ given in Eq. (52), which are valid for $`\mathrm{\Omega }\overline{v}_e𝒬`$ as indicated in Eq. (54), while the analogous proton terms are negligible. From Eqs. (81) and (91), this yields $$T_L=\frac{q^2}{\mathrm{\Omega }^2}T_0,$$ (114) where $$T_04\sqrt{2}G_Fe\omega _{0e}^2\{\begin{array}{cc}a_e+1\hfill & (\nu _e)\hfill \\ & \\ a_e\hfill & (\nu _{\mu ,\tau })\hfill \end{array}.$$ (115) The neutrinos are, for all practical purposes, ultrarelativistic. For them, we use the formulas analogous to those in Eq. (50), namely $`A_\nu (\mathrm{\Omega }^{},𝒬^{})`$ $`=`$ $`3\omega _{0\nu }^2`$ (116) $`B_\nu (\mathrm{\Omega }^{},𝒬^{})`$ $`=`$ $`3\omega _{0\nu }^2\left(1{\displaystyle \frac{\mathrm{\Omega }^{}}{2𝒬^{}}}\mathrm{ln}\left|{\displaystyle \frac{\mathrm{\Omega }^{}+𝒬^{}}{\mathrm{\Omega }^{}𝒬^{}}}\right|\right),`$ (117) with $$\omega _{0\nu }^2=\frac{1}{6\pi ^2}_0^{\mathrm{}}𝑑𝒫𝒫(f_\nu (𝒫)+f_{\overline{\nu }}(𝒫)),$$ (118) where $$f_{\nu ,\overline{\nu }}(𝒫)=\frac{1}{e^{\beta _\nu 𝒫\alpha _\nu }+1}$$ (119) are the neutrino and antineutrino momentum distributions, in the rest frame of the stream. The longitudinal dispersion relation obtained from Eq. (109) is $$\mathrm{\Omega }^2=4e^2\omega _{0e}^2+2T_0^2\frac{(q^2)^2}{\mathrm{\Omega }^2𝒬^{\mathrm{\hspace{0.17em}2}}}\left\{B_\nu N_1\left(\frac{3q^2}{𝒬^{\mathrm{\hspace{0.17em}2}}}B_\nu +\frac{1}{2}A_\nu \right)\right\}.$$ (120) In this equation, $`𝒬^{}`$ and $`\mathrm{\Omega }^{}`$ are expressed in terms of $`𝒬`$ and $`\mathrm{\Omega }`$ by means of $`𝒬^{}=\sqrt{\mathrm{\Omega }^{\mathrm{\hspace{0.17em}2}}q^2}`$ with $`\mathrm{\Omega }^{}=U^{\mathrm{\hspace{0.17em}0}}\mathrm{\Omega }\stackrel{}{U}^{}\stackrel{}{𝒬}`$, as indicated by Eq. (26). The solutions of Eq. (120) determine the dispersion relation $`\mathrm{\Omega }_L(𝒬)`$ in the long wavelength limit and are valid for $`\mathrm{\Omega }\overline{v}_e𝒬`$. ### B Neutrino driven stream instabilities Besides the usual solution $`\mathrm{\Omega }_L^24e^2\omega _{0e}^2`$, Eq. (120) can have an additional solution if the neutrino term is sufficiently large that it can compete with the electron term. To determine whether this can happen, consider the specific situation in which the velocity of the neutrino stream is not too large, so that the term in Eq. (120) proportional to $`N_1`$ can be neglected \[see Eq. (35)\]. Using the formula for $`B_\nu `$ given in Eq. (116), the longitudinal dispersion relation then becomes $$\mathrm{\Omega }^24e^2\omega _{0e}^2=T_0^2\omega _{0\nu }^2\left(1\frac{𝒬^2}{\mathrm{\Omega }^2}\right)^2f_L,$$ (121) where $`f_L`$ is the same function defined in Eq. (74). For values of $`\mathrm{\Omega }𝒬`$ the function $`f_L`$ becomes large, but with the factor $`(1𝒬^2/\mathrm{\Omega }^2)`$ its contribution in Eq. (121) is negligible in that region. On the other hand, for $`\mathrm{\Omega }0`$, $$B_e(\mathrm{\Omega }0,𝒬)=\beta _em_e\omega _{0e}^2$$ (122) so that $$T_L=T_0\beta _em_e$$ (123) in this limit, instead of Eq. (114). Therefore, the neutrino contribution is not large in the limit $`\mathrm{\Omega }0`$ either. We therefore conclude that the neutrino contribution produces a small correction to the usual dispersion relation but does not introduce any additional branch. In particular, there are no stream-induced instabilities in this system. This conclusion is in sharp contradiction with the result obtained in Ref. , where it was found that, in the same system, the dispersion relation indicates the appearance of neutrino-driven stream instabilities. To understand the origin of this discrepancy it is useful to consider $$f_{\nu ,\overline{\nu }}=(2\pi )^2n_{\nu ,\overline{\nu }}\delta ^{(3)}(\stackrel{}{𝒫}\widehat{U}^{})$$ (124) for the momentum distribution function of the neutrinos, which is of the form used in Ref. . Using it in Eq. (47) to calculate $`B_\nu `$ results in $$B_\nu =\left(\frac{n_\nu +n_{\overline{\nu }}}{2}\right)\frac{𝒬^2(\stackrel{}{𝒬}\widehat{U}^{})^2}{(\mathrm{\Omega }\stackrel{}{𝒬}\widehat{U}^{})^2},$$ (125) which, when substituted in Eq. (120), yields the longitudinal dispersion relation $$\mathrm{\Omega }^2(\mathrm{\Omega }^24e^2\omega _{0e}^2)=\frac{T_0^2(n_\nu +n_{\overline{\nu }})}{}\left(\frac{\mathrm{\Omega }^2𝒬^2}{𝒬^{}}\right)^2\frac{𝒬^2\mathrm{sin}^2\theta }{(\mathrm{\Omega }𝒬\mathrm{cos}\theta )^2},$$ (126) where $`\mathrm{cos}\theta =\stackrel{}{𝒬}\widehat{U}^{}`$ and we have neglected the term proportional to $`N_1`$, as before. Eq. (126) has a solution of the form $$\mathrm{\Omega }_L=𝒬\mathrm{cos}\theta +\delta _L^{(\nu )},$$ (127) with $$\delta _L^{(\nu )\mathrm{\hspace{0.17em}2}}=\frac{T_0^2(n_\nu +n_{\overline{\nu }})}{}\frac{𝒬^2\mathrm{sin}^4\theta }{\mathrm{cos}^2\theta (𝒬^2\mathrm{cos}^2\theta 4e^2\omega _{0e}^2)},$$ (128) which exhibits an instability for $`𝒬^2\mathrm{cos}^2\theta <4e^2\omega _{0e}^2`$. Thus, while we are able to reproduce qualitatively the result of Ref. in this way, we must note that it is based on an inconsistent application of the long wavelength formulas given in Eq. (47). As explained in detail in Ref. , those formulas are obtained by expanding the integrands in powers of $`q/`$ in the one-loop formulas given in Eq. (15), and retaining only those terms that are dominant in the limit $`q/0`$. This requires, among other conditions, that the momentum distribution functions be such that its derivatives do not introduce any singularities in the integrands. The results derived in this way are equivalent to those obtained by semiclasical methods based on kinetic theory or similar approaches. However, the form given in Eq. (124) does not satisfy the required conditions and therefore neither the long wavelength approximation of the one-loop formulas, nor the semiclassical formulas, are applicable in that case. Leaving aside the question of whether or not a distribution function such as that given in Eq. (124) is realistic in any particular physical context, in order to use it the coefficients $`A_\nu ,B_\nu `$ must be calculated with the complete one-loop formulas given in Eq. (15). Following this procedure we obtain $$B_\nu =2(n_\nu +n_{\overline{\nu }})\left[\frac{𝒬^2(\stackrel{}{𝒬}\stackrel{}{U}^{})^2}{4^2(\mathrm{\Omega }\stackrel{}{𝒬}\stackrel{}{U}^{})^2(\mathrm{\Omega }^2𝒬^2)^2}\right]$$ (129) instead of Eq. (125), so that the longitudinal dispersion relation becomes $$\mathrm{\Omega }^2(\mathrm{\Omega }^24e^2\omega _{0e}^2)=4T_0^2(n_\nu +n_{\overline{\nu }})\left(\frac{\mathrm{\Omega }^2𝒬^2}{𝒬^{}}\right)^2\frac{𝒬^2\mathrm{sin}^2\theta }{4^2(\mathrm{\Omega }𝒬\mathrm{cos}\theta )^2(\mathrm{\Omega }^2𝒬^2)^2}.$$ (130) If we neglect here the $`q^2`$ term in the denominator, we recover Eq. (126). However, when that term is taken into account, the neutrino contribution does not become large for $`\mathrm{\Omega }\stackrel{}{𝒬}\stackrel{}{U}^{}`$, and therefore a solution of the form given in Eq. (127) does not exist. ### C Transverse dispersion relation The dispersion relations for the transverse modes are given approximately by solving the equation $$\left[(q^2\pi _T^{(m)})R_{\mu \nu }(q,u)\pi _{t\mu \nu }^{(\nu )}\right]A^\nu =0,$$ (131) where $`\pi _{t\mu \nu }^{(\nu )}`$ $``$ $`R_\mu {}_{}{}^{\alpha }(q,u)R_\nu {}_{}{}^{\beta }(q,u)\pi _{\alpha \beta }^{(\nu )}`$ (132) $`=`$ $`2\left[T_T(q)R_{\mu \lambda }(q,u)+T_P(q)P_{\mu \lambda }(q,u)\right]\left[T_T(q)R_{\nu \rho }(q,u)T_P(q)P_{\nu \rho }(q,u)\right]J^{\lambda \rho }`$ (133) is the transverse projection of the neutrino contribution to the self-energy, and in the second line we have used Eq. (100). For $`J^{\alpha \beta }`$ we will use the formula given in Eq. (105) and consider the case in which the terms with the factor $`N_1`$ can be neglected, as we did in Section III B. Therefore, remembering Eq. (40), we will substitute in Eq. (132) $$J^{\alpha \beta }=\frac{1}{2}\left(A_\nu \frac{B_\nu }{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}\right)R_{\alpha \beta }(q,u)+C_\nu 𝒬\left(\frac{\stackrel{~}{u}u^{}}{\stackrel{~}{u}^2}\right)P_{\alpha \beta }(q,u)$$ (134) It is now useful to introduce the linear combinations $$R_{\alpha \beta }^{(\pm )}\frac{1}{2}\left(R_{\alpha \beta }(q,u)\pm P_{\alpha \beta }(q,u)\right),$$ (135) which satisfy $$R^{(s)\alpha \beta }R_{\beta \gamma }^{(s^{})}=\delta _{ss^{}}\delta _\gamma ^\alpha ,$$ (136) and have the representation $$R_{\alpha \beta }^{(\pm )}=e_\alpha ^{(\pm )}e_\beta ^{(\pm )}$$ (137) where $$e_\alpha ^{(\pm )}=\frac{1}{\sqrt{2}}(e_{1\alpha }\pm ie_{2\alpha }).$$ (138) Writing $`R_{\alpha \beta }`$ and $`P_{\alpha \beta }`$ in terms of $`R_{\alpha \beta }^{(\pm )}`$ and substituting the resulting formulas in Eq. (132), with the help of Eq. (136) we obtain $$\pi _{t\mu \nu }^{(\nu )}=\pi ^{(+)}R_{\mu \nu }^{(+)}+\pi ^{()}R_{\mu \nu }^{()}$$ (139) where $$\pi ^{(\pm )}=2\left(T_T\pm T_P\right)^2\left[\frac{1}{2}\left(A_\nu \frac{B_\nu }{\stackrel{~}{u}^{\mathrm{\hspace{0.17em}2}}}\right)\pm C_\nu 𝒬\left(\frac{\stackrel{~}{u}u^{}}{\stackrel{~}{u}^2}\right)\right].$$ (140) From Eq. (131), the dispersion relations for the transverse modes are then $$q^2=\pi _T^{(m)}+\pi ^{(\pm )}$$ (141) with the corresponding polarization vectors $`e_\alpha ^{(\pm )}`$. ### D Optical Activity of the neutrino gas Eq. (140) exhibits the phenomenon of optical activity of the neutrino gas, in which the two circularly polarized photon modes travel with different dispersion relations as a result of the chiral interactions of the neutrino. Notice that for this occur, $`T_P`$ and/or $`C_\nu `$ must be non-zero. This requires that the chemical potentials in the matter background be such that, for some particle species, $`\alpha _f0`$, or that $`\alpha _\nu 0`$. In the latter case however, there is an additional contribution to the photon self-energy that arises from the set of diagrams that were calculated in detail in Ref. . Those diagrams are not included in Fig. 1 and their result is an additional contribution to the photon self-energy that must be taken into account in Eq. (141). The result of the calculation of Ref. is taken into account by including in the right-hand side of Eq. (100) the term $$\mathrm{\Pi }_P^{(\nu )}P_{\mu \nu }(q,u^{})$$ (142) where, in the notation of the present paper, $$\mathrm{\Pi }_P^{(\nu )}=\frac{\sqrt{2}G_F\alpha }{3\pi }\frac{q^2}{m_e^2}(n_\nu n_{\overline{\nu }})𝒬^{},$$ (143) with $$n_{\nu ,\overline{\nu }}=\frac{d^3𝒫}{(2\pi )^3}f_{\nu ,\overline{\nu }}(𝒫).$$ (144) The result quoted in Eq. (143) is valid for values of $`q<m_e`$. When this term is included in Eq. (132), the net effect is that the right-hand side of Eq. (141) has the additional the term $$\pi _P=\pm \mathrm{\Pi }_P^{(\nu )}\left(\frac{𝒬}{𝒬^{}}\right)\frac{\stackrel{~}{u}u^{}}{\stackrel{~}{u}^2}.$$ (145) If the background contains an equal number of neutrinos and antineutrinos, then $`\mathrm{\Pi }_P^{(\nu )}`$ as well as $`C_\nu `$ are zero. In such a case, the optical activity of the neutrino gas is due to a non-zero value of $`T_P`$ in Eq. (140), which in turn depends on the difference between the particle and antiparticle number densities in the matter background. #### 1 Long wavelength limit $`\pi ^{(\pm )}`$ can be evaluated explicitly by considering specific situations. As an example we consider once more a matter background that consists of non-relativistic electron proton gases, with the photon momentum satisfying Eq. (49) and $`\mathrm{\Omega }\overline{v}_f𝒬`$. The proton contribution to $`T_T`$ is negligible, and using Eq. (47) for $`A_e`$ and $`B_e`$, $$T_T=T_0,$$ (146) with $`T_0`$ given in Eq. (115). On the other hand, $`T_P`$ is given by Eq. (83) where, in the momentum regime that we are considering, $$C_f(\mathrm{\Omega },𝒬0)=\frac{1}{2}\frac{d^3𝒫}{(2\pi )^32}\left(\frac{f_ff_{\overline{f}}}{}\right)\left[1\frac{2𝒫^2}{3^2}\right],$$ (147) as shown in Ref. . For the electron and proton gases we use the non-relativistic limit of this, $`C_f=\omega _{0f}^2/2m_f`$, which implies that the proton term is negligible and therefore $$T_P=\left(\frac{𝒬}{2m_e}\right)T_0^{},$$ (148) where $$T_0^{}=4\sqrt{2}G_Fe\omega _{0e}^2\{\begin{array}{cc}b_e1\hfill & (\nu _e)\hfill \\ & \\ b_e\hfill & (\nu _{\mu ,\tau })\hfill \end{array}$$ (149) For the neutrino gas, the relativistic limit of Eq. (147) yields $$C_\nu =\frac{1}{24\pi ^2}_0^{\mathrm{}}𝑑𝒫(f_\nu (𝒫)f_{\overline{\nu }}(𝒫)),$$ (150) while $`A_\nu `$ and $`B_\nu `$ are given in Eq. (116). With the help of these formulas and remembering that $`\text{Re}\pi _T^{(m)}=4e^2\omega _{0e}^2`$ in the case we are considering, the dispersion relation becomes $$q^2=4e^2\omega _{0e}^2+2T_0^2\omega _{0\nu }^2\left[1\frac{𝒬T_0^{}}{2m_eT_0}\right]^2\left[1\frac{C_\nu 𝒬}{\omega _{0\nu }^2}\left(\frac{\stackrel{~}{u}u^{}}{\stackrel{~}{u}^2}\right)\right]^2\pm \frac{\mathrm{\Pi }_P^{(\nu )}𝒬}{𝒬^{}}\left(\frac{\stackrel{~}{u}u^{}}{\stackrel{~}{u}^2}\right),$$ (151) where we have included the $`\mathrm{\Pi }_P^{(\nu )}`$ term as indicated in Eq. (145). Let us consider first the situation in which $`f_\nu f_{\overline{\nu }}`$, so that $`\mathrm{\Pi }_P^{(\nu )}`$ and $`C_\nu `$ can be neglected in Eq. (151). The solutions for the two modes are then given by $$\mathrm{\Omega }_\pm =\sqrt{𝒬^2+4e^2\omega _{0e}^2}\frac{T_0T_0^{}\omega _{0\nu }^2}{m_e}\frac{𝒬}{\sqrt{𝒬^2+4e^2\omega _{0e}^2}},$$ (152) which is of the same form as that given in Eq. (4.22) of Ref. , if we make the correspondence $$\frac{1}{2}aR_\nu \frac{T_0T_0^{}\omega _{0\nu }^2}{m_e}$$ (153) there. For the situations of potential interest analyzed in Ref. , the effects of the dispersion relations given in Eq. (152) are smaller than those found in that reference by a factor of about $`G_F\omega _{0e}^2G_Fn_e/m_e`$, and whence are unimportant for all practical purposes. Therefore, retaining only the term proportional to $`\mathrm{\Pi }_P^{(\nu )}`$ in Eq. (151), the dispersion relation becomes $$q^2=4e^2\omega _{0e}^2\pm \xi \mathrm{\Pi }_P^{(\nu )},$$ (154) where $$\xi =\frac{𝒬}{𝒬^{}}\left(\frac{\stackrel{~}{u}u^{}}{\stackrel{~}{u}^2}\right).$$ (155) Of course, when the neutrino gas is not moving relative to the matter background ($`\stackrel{}{U}^{}=0`$), $`\xi =1`$ and Eq. (154) reduces to the form given in Ref. . The salient feature here is that, in general, the dispersion relation is not isotropic, so that the splitting between the two circularly polarized modes is different depending on the direction of propagation of the photon relative to the velocity of the neutrino gas. To assess the consequences that this effect can have on the analysis given in Ref. we consider two extreme cases. ##### a $`\stackrel{}{𝒬}`$ perpendicular to $`\stackrel{}{U}^{}`$. Using $`\stackrel{}{𝒬}\stackrel{}{U}^{}=0`$, it is very simple to verify that $$\stackrel{~}{u}u^{}=\frac{U^{\mathrm{\hspace{0.17em}0}}𝒬^2}{q^2},$$ while $`𝒬^{}=\sqrt{\stackrel{}{U}^{\mathrm{\hspace{0.17em}2}}+𝒬^2}`$. Using $`\stackrel{~}{u}^2=𝒬^2/q^2`$, it then follows that $$\xi =\frac{𝒬\sqrt{1+\stackrel{}{U}^{\mathrm{\hspace{0.17em}2}}}}{\sqrt{\mathrm{\Omega }^2\stackrel{}{U}^{\mathrm{\hspace{0.17em}2}}+𝒬^2}}.$$ (156) For small velocities of the neutrino gas this reduces to 1, as it should be, while for large velocities it implies that the effect of the $`\mathrm{\Pi }_P^{(\nu )}`$ term is reduced by the factor $`𝒬/\mathrm{\Omega }`$ for $`\mathrm{\Omega }>𝒬`$. ##### b $`\stackrel{}{𝒬}`$ parallel to $`\stackrel{}{U}^{}`$. We set $$\stackrel{}{Q}=\lambda 𝒬\widehat{U}^{}$$ (157) to include the possibility that the photon propagates antiparallel to the velocity of the neutrino gas. A little algebra shows that in this case $$\stackrel{~}{u}u^{}=\frac{U^{\mathrm{\hspace{0.17em}0}}𝒬^2+\lambda \mathrm{\Omega }𝒬|\stackrel{}{U}^{}|}{q^2}$$ while $`𝒬^{}=|\mathrm{\Omega }|\stackrel{}{U}^{}|\lambda 𝒬U^{\mathrm{\hspace{0.17em}0}}|`$. Therefore $$\xi =\frac{𝒬\lambda \beta ^{}\mathrm{\Omega }}{\left|\mathrm{\Omega }\beta ^{}\lambda 𝒬\right|},$$ (158) where we have defined the speed of the neutrino gas $$\stackrel{}{\beta }^{}=\frac{\stackrel{}{U}^{}}{U^{\mathrm{\hspace{0.17em}0}}}.$$ (159) Eq. (158) reveals in a simple way the anisotropic nature of the dispersion relation. For example, if the velocity of the neutrino stream is such that $$\frac{𝒬}{\mathrm{\Omega }}<\beta ^{},$$ (160) then $`\xi =1`$ or $`+1`$ depending on whether the photon is propagating parallel or antiparallel to $`\stackrel{}{\beta }^{}`$, respectively. In particular, this implies that the frequency difference between the dispersion relations of the two (circularly polarized) transverse modes is the opposite to what it is if the neutrino gas is not moving relative to the matter background. This effect is easy to understand by noticing that, if the condition in Eq. (160) holds, then a photon moving parallel to $`\stackrel{}{\beta }^{}`$ appears to be moving in the opposite direction in the rest frame of the neutrino gas. ## IV Outlook Although our work has been largely motivated by the study of the electromagnetic properties of a neutrino gas that moves, as a whole, relative to a plasma, the approach is applicable to a wider class of problem in similar physical systems, involving relativistic plasmas or high temperature gauge theories. The field theory methods employed here allow us to consider situations for which the semiclassical approaches, such as those based on kinetic theory, are not suitable, and those for which the full power of the techniques and methods that have been developed for high temperature field theory calculations must be employed. Some possible extensions of the present work along those lines would involve the calculation of the imaginary part of the self-energy to determine the damping rates, and the application of the resummation methods to study the dispersion relations in those circumstances in which the perturbative approximations are not reliable. ##### Acknowledgement This work has been partially supported by the U.S. National Science Foundation Grant PHY-9900766.
warning/0001/nucl-th0001001.html
ar5iv
text
# Exploring vector meson masses in nuclear collisions 11footnote 1Talk delivered by B. K. Jain at National Seminar on Nuclear Physics, July 26-29, 1999, Institute of Physics, Bhubaneswar, India ## 1 Introduction The modification of hadron masses in the nuclear medium is an issue of much interest currently . Since in QCD the hadrons are excitations of the vacuum, it is natural that these excitations can get affected by the proximity of other hadrons. Experimentally the masses of unstable hadrons are explored by producing them in nuclear reactions and then measuring the invariant masses of their leptonic decay products . The medium modification of the unstable hadron from these data is, normally, inferred by adding incoherently the decay of the hadron inside and outside the nuclear medium . Recently formalisms have been developed by us and Boreskov et al. for the propagation of resonances produced in proton-nucleus collisions. These formalisms incorporate the interaction of the resonance with the nuclear medium. The invariant mass spectrum of the measured decay products in these formalisms is obtained by adding coherently the contribution of the resonance decay inside and outside the nuclear medium. The formalism of Jain et al. also includes the interaction of the resonance decay products with the nuclear medium if the latter are hadrons. In the first part of the present talk we give the modification of our earlier formalism to heavy-ion collisions, and apply it to the propagation and decay of rho-meson. Our aim is to see as how the medium modification of $`\rho `$-meson shows up in the di-lepton invariant mass distribution using a proper quantum mechanical description for the rho propagation in the nucleus. We do not compare our calculated results with the existing experimental data on rho-meson production in high energy heavy-ion collisions, because these data contain contribution from several other di-lepton processes than that considered in the present paper. In the second part of the talk we write a formalism for coherent rho production in proton nucleus collisions and explore the sensitivity of the (p,p$`{}_{}{}^{}\rho _{}^{0}`$) reaction cross section to medium mass modification of the rho meson. ## 2 Heavy-ion collisions ### 2.1 Formalism Let us suppose that two heavy nuclei, one the projectile $`A`$ and another the target $`B`$, collide at high energies. We assume that one nucleon in the projectile and one in the target collide and a resonance $`R`$ is produced at the collision point. This resonance then moves along the beam direction, which is taken as the z-axis, and decays at some subsequent point. Denoting by $`𝐫`$ the relative coordinate between the target and the projectile, and by $`(𝐫_𝐁,𝐫_𝐀)`$ the intrinsic coordinates of the target and projectile nucleons, respectively, the resonance coordinate is written as, $`𝐫_𝐑=𝐫_𝐀+{\displaystyle \frac{B}{A+B}}𝐫.`$ (1) With this definition the ratio of the cross sections for the resonance production in AB and NN collisions, for an inclusive situation where the state of the (A+B) system is not identified, can be written as, $`{\displaystyle \frac{\mathrm{\Delta }\sigma _R^{AB}}{\mathrm{\Delta }\sigma _R^{NN}}}=[K.F.]{\displaystyle }d𝐛{\displaystyle }dz{\displaystyle }d𝐫_𝐀\rho _A(𝐫_𝐀)\rho _B(𝐫+𝐫_𝐀)|G(𝐫_𝐑;𝐤_𝐑,\mu )|^2,`$ (2) where $`\rho _x`$ are the nuclear densities. \[K.F.\] is the relevant kinematical factor. The function $`G(𝐫_𝐑;𝐤_𝐑,\mu )`$ physically gives the probability amplitude for finding the decay products of the resonance in the detector with the total momentum $`𝐤_𝐑`$ and the invariant mass $`\mu `$, if the resonance is produced at a point $`𝐫_𝐑`$ in the nucleus (for details see Ref. . In terms of the resonance propagator $`G(𝐫_𝐑^{^{}},𝐫_𝐑)`$, the function $`G(𝐫_𝐑;𝐤_𝐑,\mu )`$ is defined as $`G(𝐫_𝐑;𝐤_𝐑,\mu )={\displaystyle }d𝐫_𝐑^{^{}}exp(i𝐤_𝐑.𝐫_𝐑^{^{}})G(𝐫_𝐑^{^{}},𝐫_𝐑),`$ (3) where $`G(𝐫_𝐑^{^{}},𝐫_𝐑)`$ satisfies $`[^2+E^2m_R^2+i\mathrm{\Gamma }_Rm_R\mathrm{\Pi }_R]G(𝐫_𝐑^{^{}},𝐫_𝐑)=\delta (𝐫_𝐑^{^{}}𝐫_𝐑).`$ (4) Here $`\mathrm{\Pi }_R`$ is the self energy of the resonance in the medium and $`\mathrm{\Gamma }_R`$ is its free space decay width. In the eikonal approximation we can write, $`G(𝐫_𝐑;𝐤_𝐑,\mu )=exp(i𝐤_𝐑.𝐫_𝐑)\varphi _R(𝐫_𝐑;𝐤_𝐑,\mu ),`$ (5) where $`\varphi _R`$ is a slowly varying modulating function. With this, and using the Eqs. (3,4), $`\varphi _R`$ approximately works out to $`\varphi (𝐫_𝐑;𝐤_𝐑,\mu )`$ $`=`$ $`{\displaystyle \frac{1}{2ik_R}}{\displaystyle 𝑑z_R^{^{}}exp[\frac{1}{2ik_R}(\mu ^2m_R^2+i\mathrm{\Gamma }_Rm_R)(z_Rz_R^{^{}})]}`$ (6) $`\times exp[{\displaystyle \frac{i}{v_R}}{\displaystyle _{z_R}^{z_R^{}}}V_R(b_R,z{}_{}{}^{\prime \prime })dz_R^{^{\prime \prime }}].`$ Here we have written $`\mathrm{\Pi }_R=2E_RV_R,`$ (7) where $`V_R`$ is the optical potential of the resonance, R, in the nuclear medium. In general, it is complex. Its real part, as we shall see later, is related to the mass shift of the resonance and the imaginary part gives the collision broadening of the resonance in the medium. For a nucleus with a sharp surface, function $`\varphi (𝐫_𝐑;𝐤_𝐑,\mu )`$ splits into a sum of two terms, one corresponding to the decay of the resonance inside the nucleus and another to the decay outside the nucleus, i.e. $`\varphi (𝐫_𝐑;𝐤_𝐑,\mu )=\varphi _{in}(𝐫_𝐑)+\varphi _{out}(𝐫_𝐑)`$ (8) with $`\varphi _{in}(𝐫_𝐑)={\displaystyle \frac{1}{2ik_R}}{\displaystyle _{z_R}^{\sqrt{(}R^2b^2)}}𝑑z_R^{}\varphi _R(𝐛_𝐑;z_R,z_R^{}),`$ (9) and $`\varphi _{out}(𝐫_𝐑)={\displaystyle \frac{1}{2ik_R}}{\displaystyle _{\sqrt{(}R^2b^2)}^{\mathrm{}}}𝑑z_R^{}\varphi _R(𝐛_𝐑;z_R,z_R^{}).`$ (10) Here $`\varphi _R(𝐛_𝐑;z_R,z_R^{})`$ $`=`$ $`exp[{\displaystyle \frac{1}{2ik_R}}(\mu ^2m_R^2+i\mathrm{\Gamma }_Rm_R)(z_Rz_R^{^{}})]`$ (11) $`\times exp[{\displaystyle \frac{i}{v_R}}{\displaystyle _{z_R}^{z_R^{}}}V_R(b_R,z{}_{}{}^{\prime \prime }{}_{R}{}^{})dz_R^{^{\prime \prime }}].`$ After a little bit of manipulations, the final expressions for $`\varphi _{in}`$ and $`\varphi _{out}`$ work out to , $`\varphi _{in}(𝐫_𝐑;k_R,\mu )={\displaystyle \frac{G_0^{}}{2m_R}}[1exp({\displaystyle \frac{i}{v_RG_0^{}}}[L(b_R)z_R])],`$ (12) $`\varphi _{out}(𝐫_𝐑;k_R,\mu )={\displaystyle \frac{G_0}{2m_R}}[exp({\displaystyle \frac{i}{v_RG_0^{}}}[L(b_R)z_R])],`$ (13) where $`v_R`$ is the speed of the resonance and $`L(=\sqrt{(R^2b_R^2)})`$ is the length from the production point to the surface of the nucleus. $`G_0`$ and $`G_0^{}`$ in Eqs. (12) and (13) are the free and the in-medium resonance propagators. Their forms are $`G_0={\displaystyle \frac{2m_R}{\mu ^2m_R^2+i\mathrm{\Gamma }_Rm_R}},`$ (14) $`G_0^{}={\displaystyle \frac{2m_R}{\mu ^2m_R^2+i\mathrm{\Gamma }_R^{}m_R}},`$ (15) with $`m_R^{}`$ $`m_R+{\displaystyle \frac{E_R}{m_R}}U_R.`$ (16) $`\mathrm{\Gamma }_R^{}=\mathrm{\Gamma }_R+{\displaystyle \frac{E_R}{m_R}}|2W_R|.`$ (17) It may be mentioned that, for a nucleus with no sharp surface approximation the expression given in Eq. (6) can be used directly to evaluate the function $`\varphi (𝐫_𝐑;𝐤_𝐑,\mu )`$. In the above we have written, $`V_R=U_R+iW_R.`$ (18) These potentials, as given in Eqs. (16,17) give a measure of the mass and width modification of the resonance in the nuclear medium. Their values are an open question and a subject of much research internationally. In one approach they can be treated as completely unknown quantities and data on appropriate experiments can be used to extract their values. This exercise would be of use if the theoretical formalism used describes the reaction dynamics correctly and the data do not have much uncertainty. Alternatively, they can be estimated in a particular model and the ensuing values can be used to make an estimate of the cross section for the rho production. In literature, various efforts have been made to estimate $`V_R`$ using the high energy ansatz, i.e. $`U_R=\alpha [{\displaystyle \frac{1}{2}}v_R\sigma _T^{RN}\rho _0]`$ (19) and $`W_R=[{\displaystyle \frac{1}{2}}v_R\sigma _T^{RN}\rho _0],`$ (20) where $`\alpha `$ is the ratio of the real to the imaginary part of the elementary RN scattering amplitude and $`\sigma _T^{RN}`$ is the total cross section for it. $`\rho _0`$ is the typical nuclear density. A detailed calculation for the rho-meson has been done on these lines by Kondrayuk et al. which give these potentials as a function of momentum. They use VDM at high enegies and resonance model at low energies to generate the $`\rho `$N scattering parametrs. We have used these values for our calculations in the present paper. Some representative values of the self energies required in our calculations are given in Table 1. ### 2.2 Results and Discussion Examining Eqs. (9-11) $`\varphi (𝐫_𝐑;𝐤_𝐑,\mu )`$ in above we find that the cross sections for the decay of the resonance in the nucleus depends upon the length of the nuclear medium, the speed ($`v_R`$), free decay width and self energy of the resonance. To represent the effect of all these quantities, we present results for the decay of the rho-meson for different values of $`v_R`$ and for two sets of nuclear systems, viz. $`Pb+Pb`$ and $`S+Au`$. The free width of the rho-meson is taken equal to 150 MeV. The optical potentials, which we need at several $`\rho `$ momenta, are taken, as mentioned above, from Kondratyuk et al. . Nuclear densities are taken from Ref. . Denoting the ratio $`\frac{1}{[K.F.]}\frac{\mathrm{\Delta }\sigma _R^{AB}}{\mathrm{\Delta }\sigma _R^{NN}}`$ in Eq. (2) as $`|\mathrm{\Phi }|^2`$, we plot $`|\mathrm{\Phi }|^2`$ as a function of the invariant mass, $`\mu `$, of the decay products of the $`\rho `$ meson. Figures 1-4 show the invariant mass spectra of the $`\rho `$ meson in $`Pb+Pb`$ and $`S+Au`$ collisions at rho velocities of $`0.6c`$ and $`0.9c`$. Here $`c`$ is the speed of light. The solid curve in all the figures gives the coherently summed cross-section from the decay of $`\rho `$-meson inside and outside the nuclear medium. The dashed curve gives the same added incoherently. The individual contributions corresponding to the inside and the outside decay are given by the dash-dot and dash-dot-dot curves respectively. We observe two things. One, the coherent and the incoherent cross-sections are different and second, this difference increases with the increase in the rho-meson speed. At 0.6c speed, while the coherent and incoherent curves differ only in the peak cross sections, at 0.9c speed their shape and peak cross sections both are different. We also observe that the difference is larger for the smaller system like S on Au. If we compare the mass shift seen in our calculations (Figs. 1-4) with those indicated in the high energy heavy ion collisions, our shifts are small and are in the opposite direction. To explore as what kind of optical potentials would produce a shift as large as those indicated experimentally we calculated $`|\mathrm{\Phi }|^2`$ for Pb+Pb at 0.9c for three arbitrarily chosen values of $`U_R`$, viz. -40, -80 and -120 MeV. These results are shown in Fig. 5. We find that only with -120 MeV value the distribution starts having features resembling those indicated in heavy-ion experiments. But this value, compared with the value around +25 MeV coming from the high energy ansatz of Kondratyuk et al. (see Table 1) is very large and is of opposite sign. ## 3 Proton-nucleus collisions The cross section for a coherent rho production reaction, (p,p$`{}_{}{}^{}\rho _{}^{0}`$), is given by $$d\sigma =[PS]S(m^2)<|T_{coh}|^2>,$$ (21) where the phase-space factor, \[$`PS`$\], is written as $$[PS]=\frac{\pi m_p^2m_A}{(2\pi )^6}\frac{k_p^{}k_\rho ^2}{k_p[k_\rho (E_iE_p^{})(𝐤_𝐩𝐤_𝐩^{}).\widehat{k_\rho }E_\rho ]}dm^2dE_p^{}d\mathrm{\Omega }_p^{}d\mathrm{\Omega }_\rho .$$ (22) $`S(m^2)`$ is the free space rho mass distribution function, which is given by $$S(m^2)=\frac{1}{\pi }\frac{m_\rho \mathrm{\Gamma }_\rho }{[(m^2m_\rho ^2)^2+m_\rho ^2\mathrm{\Gamma }_\rho ^2]},$$ (23) with $`m_\rho =770`$ MeV and $`\mathrm{\Gamma }_\rho =150`$ MeV. The T-matrix, $`T_{coh}`$, is given by $$T_{coh}=(\chi _{𝐤_𝐩^{}}^{()},\mathrm{\Psi }_{𝐤_\rho }^{()}<p^{},\rho ^0|_{\rho NN}|p>\chi _{𝐤_𝐩}^{(+)}),$$ (24) where $`\chi `$’s denote the distorted waves for the incoming and outgoing protons. However, in the energy region of interest for rho production the distortion effects are mainly absorptive. Therefore, the proton distorted waves in above can be replaced by plane waves for the present purpose. $`\mathrm{\Psi }_{𝐤_\rho }^{()}`$ is the $`\rho `$-meson scattering wave function with asymptotic momentum $`𝐤_\rho `$. It has the form $$\mathrm{\Psi }_{𝐤_\rho }^{()}=e^{i𝐤_\rho 𝐫}+\mathrm{\Psi }_{scat.}^{}$$ (25) In the absence of any dispersive nuclear distortion of p and p, the first term in this equation does not contribute to $`T_{coh}`$ because $`𝐤_\rho 𝐤_𝐩𝐤_𝐩^{}`$. This in other words means that the rho meson produced at the proton vertex is off-shell. It can not be seen in the detector without incorporating the medium effects on it. $`\mathrm{\Psi }_{scat.}`$ is the part of the wave function which include these effects. If we associate a self energy $`\mathrm{\Pi }(=2\omega V`$, where $`V`$ is the corresponding optical potential) with the $`\rho `$-meson, $`\mathrm{\Psi }_{scat.}`$ is given by $$\mathrm{\Psi }_{scat.}^{}=\chi _{𝐤_\rho }^{()}VG_\rho (t),$$ (26) where $`\chi _{𝐤_\rho }^{()}`$ is the scattering solution of the potential $`V`$. $`G_\rho (t)`$ is the $`\rho `$-meson propagator, and is given by $$G_\rho (t)=\frac{2\omega }{m_\rho ^2ti\omega \mathrm{\Gamma }_\rho }.$$ (27) $`t(=\omega ^2𝐪^2)`$ is the four-momentum transfer squared to $`\rho `$-meson at the production vertex. For the $`\rho `$ production Lagrangian in Eq. (24) we have taken $$_{\rho NN}=\frac{fF(t)}{m_\rho }N^{}(\sigma \mathrm{𝐱𝐪})\tau N\rho ,$$ (28) with $`\rho `$NN coupling constant, $`f`$, equal to 7.81, and the off-shell extrapolation form factor as $$F(t)=\frac{\mathrm{\Lambda }^2m_\rho ^2}{\mathrm{\Lambda }^2t},$$ (29) with $`\mathrm{\Lambda }`$= 2 GeV/c. With the above formalism we calculate the $`\rho `$ production cross section for the <sup>12</sup>C target nucleus. The only quantity required for the calculation is the description of the optical potential, $`V`$, of the $`\rho `$-meson. The values of the optical potential are fixed using the same prescription as given earlier for the heavy ion reactions. Some representative values required by us are given in Table 2. The radial shape of the optical potential is approximated by the radial density distribution of the <sup>12</sup>C nucleus. In Fig. 6 we plot the calculated outgoing proton energy spectrum for p going very near to the forward direction against the energy transfer $`\omega `$(=T<sub>p</sub>-T$`_p^{}`$). This energy transfer and the corresponding momentum transfer q(=$`𝐤_𝐩`$ \- $`𝐤_𝐩^{}`$) are shared between the rho-meson and the recoiling nucleus through the interaction of the rho-meson with the target nucleus. The beam energy is taken equal to 1.5 GeV. We see in the figure that the calculated distribution has a broad peak. The peak cross section is around 0.34 $`\mu `$b/MeV. In Fig. 7 we show the angular distribution of the above rho-mesons at the peak position in Fig 6. It is observed that most of the rho-meson flux gets emitted in the forward direction only. Very little is seen beyond 15<sup>0</sup> or so. Above results are given for a certain choice of the $`\rho `$-meson optical potential. However, they would be sensitive to the change in this potential. In Fig. 8 we have investigated this sensitivity. The optical potential for this purpose has been taken purely real, and different values for it are fixed through different mass-shifts of the $`\rho `$-meson in the medium using the relation given in Eq. (16). In Fig. 8 we show the calculated proton energy spectrum for $`\mathrm{\Delta }\text{m(=m-m}^{})`$ taken equal to 50, 100 and 150 MeV. On x-axis, instead of $`\omega `$, we have $`\frac{\omega }{\mathrm{\Delta }m}`$. This is done because the essential parameter determining the dynamics of the rho-meson in the potential is likely to be the rho energy relative to the depth of the potential. We observe that 1. the magnitude of the cross sections increases with the increase in the strength of the potential. 2. In addition to the broad peak, we see a sharp peak in the small energy region of the rho-meson. The position of this peak on the $`\frac{\omega }{\mathrm{\Delta }m}`$ scale is around 3.7 for $`\mathrm{\Delta }`$m=100 and 150 MeV. For $`\mathrm{\Delta }`$m=50 MeV, this peak is not seen in the results because by then the cross section becomes too small. On examining the phase-shifts of the scattered wave function of rho-meson in the potential, we find that the sharp peak is like a shape elastic resonance seen in the elastic scattering experiments. Of course, when the $`\rho `$-potential is made complex, as is in Fig. 6 the sharp peak disappears. To summarize, we find that in proton scattering on nuclei a measurable cross section exists for $`\rho `$ meson production due to coherent effect of the target nucleus. The actual magnitude of the cross section depends sensitively on the strength of the $`\rho `$-meson optical potential, which is related to the rho-mass modification in the nuclear medium. The cross section increases with the increase in the potential strength. The angular distribution of the emitted rho-meson is such that most of them go in a forward cone of about 15<sup>0</sup>. For a purely real potential a sharp peak appears in the proton energy spectrum in the region of the small rho-meson energy. The authors acknowledge many useful discussions they had with Shashi Phatak and A. B. Santra, and thank them for the same.
warning/0001/hep-th0001171.html
ar5iv
text
# Vacuum Polarization in QED with World-Line Methods ## I Introduction String-inspired methods in QFT were initiated by Bern and Kosower , who applied them to compute one-loop amplitudes in various field theories. These authors and Strassler then recognized that some of the well-known vacuum processes in QED and QCD can be computed rather easily with the aid of one-dimensional path integrals for relativistic point particles. Similar techniques and results are also contained in the monograph by Polyakov . String-inspired methods, particularly in QED, were then extensively studied in a series of papers by Schmidt, Schubert and Reuter, cf.,e.g.,, where the state of the art is reviewed extensively. There are also contributions by McKeon and various co-authors who have proved that world-line methods are extremely useful. In the present article we will try to compactify their work and present some new representations which will finally permit us to write the one-loop diagram tied to an arbitrary number of off-shell photons - together with an applied external constant electromagnetic field of any configuration - in a so far unknown manner . This highly condensed form for the one-loop vacuum process turns out to be directly applicable to various limiting situations, e.g., turning off the off-shell photon lines will bring us back to the effective action of QED which in the low-frequency limit yields the Heisenberg-Euler Lagrangian. Another process should be photon splitting in a prescribed constant magnetic field as calculated by Adler and Schubert . ## II Vacuum Polarization without external Fields - scalar QED. Our starting point is the one loop action, which we write in a path integral representation, $$\mathrm{\Gamma }\left[A\right]=i_0^{\mathrm{}}\frac{dT}{T}𝒩\underset{x(T)=x(0)}{}𝒟xe^{i{\displaystyle _0^T}𝑑\tau \left[{\displaystyle \frac{\dot{x}^2}{4}}eA\dot{x}m^2\right]},$$ (1) where $`𝒩`$ is determined by $$𝒩\underset{x(T)=x(0)}{}𝒟xe^{i{\displaystyle _0^T}𝑑\tau {\displaystyle \frac{\dot{x}^2}{4}}}=\frac{i}{\left(4\pi T\right)^2}V_4.$$ (2) In (1), A denotes a superposition of external plus radiation field. If we then expand the radiation field in plane waves, $`A_\mu (x)=_{i=1}^N\epsilon _i^\mu e^{ik_ix}`$, we obtain $`\mathrm{\Gamma }_N[k_1,\epsilon _1,\mathrm{},k_N,\epsilon _N]`$ $`=`$ $`i(ie)^N{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{T}}𝒩{\displaystyle \underset{x(T)=x(0)}{}}𝒟xe^{i{\displaystyle _0^T}𝑑\tau \left[{\displaystyle \frac{\dot{x}^2}{4}}eA\dot{x}m^2\right]}`$ (4) $`{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^T}dt_i\epsilon _i\dot{x}(t_i)e^{ik_ix\left(t_i\right)}`$ Hence, without external field and keeping only terms linear in $`\epsilon _i`$, (4) becomes $`\mathrm{\Gamma }_N=ie^N{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{T}}𝒩{\displaystyle \underset{x(T)=x(0)}{}}𝒟xe^{i{\displaystyle _0^T}𝑑\tau \left[{\displaystyle \frac{\dot{x}^2}{4}}m^2\right]}\left({\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^T}dt_i\right)`$ (5) $`e^{i{\displaystyle \underset{j=1}{\overset{N}{}}}\left(k_jx\left(t_j\right)i\epsilon _j\dot{x}\left(t_j\right)\right)}|_{linearineach\epsilon _j}`$ $`.`$ (6) Introducing the source for $`x^\mu `$ $$J^\mu \left(\tau \right)=i\underset{j=1}{\overset{N}{}}\left(k_j^\mu \epsilon _j^\mu \frac{}{t_j}\right)\delta \left(\tau t_j\right),$$ (7) we can rewrite (6) as $$\mathrm{\Gamma }_N=ie^N_0^{\mathrm{}}\frac{dT}{T}e^{im^2T}\left(\underset{i=1}{\overset{N}{}}_0^T𝑑t_i\right)𝒩\underset{x(T)=x(0)}{}𝒟xe^{{\displaystyle \frac{i}{4}}{\displaystyle _0^T}𝑑\tau x{\displaystyle \frac{d^2}{d\tau ^2}}x}e^{{\displaystyle _0^T}J\left(\tau \right)x\left(\tau \right)}.$$ (8) The operator $`d^2/d\tau ^2`$, acting on $`x(\tau )`$ with periodical boundary condition $`x(T)=x(0)`$, has zero modes $`x_0`$. These modes will be separated from their orthogonal non-zero mode partners by writing $`x(\tau )=x_0+\xi (\tau )`$ with $`_0^T𝑑\tau \xi ^\mu (\tau )=0`$, i.e.,$`_0^T𝑑\tau x^\mu (\tau )=x_0^\mu `$ and $`𝒟x=d^4x_0𝒟\xi `$. Since $`x(\tau )`$ ( and therefore $`\xi (\tau )`$) is periodic we can write $$x\left(\tau \right)=x_0+\underset{n0}{}c_ne^{{\displaystyle \frac{2\pi in}{T}}\tau }.$$ (9) Now with the use of (2), equation (8) may be written as $`\mathrm{\Gamma }_N`$ $`=`$ $`\left(2\pi \right)^4\delta ^4\left({\displaystyle \underset{i=1}{\overset{N}{}}}k_i\right)e^N{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{\left(4\pi \right)^2T^3}}e^{im^2T}\left({\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^T}𝑑t_i\right){\displaystyle \frac{𝒟\xi e^{{\displaystyle \frac{i}{4}}{\displaystyle _0^T}𝑑\tau \xi {\displaystyle \frac{d^2}{d\tau ^2}}\xi }e^{{\displaystyle _0^T}J\xi }}{𝒟\xi e^{{\displaystyle \frac{i}{4}}{\displaystyle _0^T}𝑑\tau \xi {\displaystyle \frac{d^2}{d\tau ^2}}\xi }}}`$ (10) $`=`$ $`\left(2\pi \right)^4\delta ^4\left({\displaystyle \underset{i=1}{\overset{N}{}}}k_i\right)e^N{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{\left(4\pi \right)^2T^3}}e^{im^2T}{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^T}dt_i`$ (12) $`e^{{\displaystyle \frac{i}{2}}{\displaystyle _0^T}𝑑\tau {\displaystyle _0^T}𝑑\tau ^{}J^\mu \left(\tau \right)G_{\mu \nu }(\tau ,\tau ^{})J^\nu \left(\tau ^{}\right)},`$ where the Green’s function $`G_{\mu \nu }`$ and its properties are given by $`G_{\mu \nu }(\tau ,\tau ^{})=g_{\mu \nu }G(\tau ,\tau ^{}),{\displaystyle \frac{1}{2}}_\tau ^2G(\tau ,\tau ^{})=\delta \left(\tau \tau ^{}\right){\displaystyle \frac{1}{T}}`$ (13) $`G(\tau ,\tau ^{})=\left|\tau \tau ^{}\right|{\displaystyle \frac{\left(\tau \tau ^{}\right)^2}{T}}+const;G(\tau ,\tau ^{})=G\left(\tau \tau ^{}\right)=G(\tau ^{},\tau )`$ (14) $`_\tau G(\tau ,\tau ^{})\dot{G}(\tau ,\tau ^{})=sign\left(\tau \tau ^{}\right){\displaystyle \frac{2\left(\tau \tau ^{}\right)}{T}};\dot{G}(\tau ,\tau ^{})=\dot{G}(\tau ^{},\tau ).`$ (15) Note the generic structure expressed in (12), where particle and off-shell photons are factorized in such a way that the free (i.e., sans external field) scalar particle circulating in the loop becomes multiplied by the exponential term which is solely due to the photons tied to the loop. With $`J^\mu \left(\tau \right)`$ given in (7) we finally obtain $`\mathrm{\Gamma }_N[k_1,\epsilon _1,\mathrm{},k_N,\epsilon _N]=\left(2\pi \right)^4\delta ^4\left({\displaystyle \underset{i=1}{\overset{N}{}}}k_i\right)e^N{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{\left(4\pi \right)^2T^3}}e^{im^2T}{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^T}𝑑t_i`$ (16) $`\mathrm{exp}{\displaystyle \frac{i}{2}}{\displaystyle \underset{i,j=1}{\overset{N}{}}}\left[k_ik_jG(t_i,t_j)k_i\epsilon _j{\displaystyle \frac{}{t_j}}G(t_i,t_j)k_j\epsilon _i{\displaystyle \frac{}{t_i}}G(t_i,t_j)+\epsilon _i\epsilon _j{\displaystyle \frac{^2}{t_it_j}}G(t_i,t_j)\right].`$ (17) Since $`G(\tau ,\tau )=0,\dot{G}(\tau ,\tau )`$, there are no terms with $`k_i^2`$ and $`\epsilon _ik_i`$, i.e., without use of on-shell conditions. As an example we just note that for N=2 we obtain, after keeping only terms linear in $`\epsilon _i`$, $`\mathrm{\Gamma }_2[k_1,\epsilon _1;k_2,\epsilon _2]=\left(2\pi \right)^4\delta ^4\left(k_1+k_2\right)e^2{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{\left(4\pi T\right)^2}}e^{im^2T}{\displaystyle _0^T}𝑑t_1e^{ik_1^2G\left(t_1\right)}`$ (18) $`\{\left(k_1\epsilon _2\right)\left(k_2\epsilon _1\right)\left(\epsilon _1\epsilon _2\right)\left(k_1k_2\right)\}\dot{G}^2\left(t_1\right).`$ (19) This expression is manifestly gauge invariant. Upon using $`G(t_1)=\frac{t_1^2}{T}+t_1`$,$`\dot{G}(t_1)=\frac{2t_1}{T}+1`$ and substituting $`v=\frac{2t_1}{T}1`$, we obtain $`\mathrm{\Gamma }_2[k_1,\epsilon _1;k_2,\epsilon _2]=\left(2\pi \right)^4\delta ^4(k_1+k_2)e^2[(k_1\epsilon _2)(k_2\epsilon _1)(\epsilon _1\epsilon _2)(k_1k_2)]`$ (20) $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{2\left(4\pi \right)^2T}}e^{im^2T}{\displaystyle _1^1}dvv^2e^{ik_1^2{\displaystyle \frac{T}{4}}\left(1v^2\right)}.`$ (21) At this point we can make contact with the vacuum polarization digram in scalar QED: $$\mathrm{\Gamma }_2=\left(2\pi \right)^4\delta ^4\left(k_1+k_2\right)\epsilon _\mu \mathrm{\Pi }^{\mu \nu }\epsilon _\nu ,$$ (22) where, after renormalization, we obtain for $`\mathrm{\Pi }_{\mu \nu }=\left(g_{\mu \nu }k^2k_\mu k_\nu \right)\mathrm{\Pi }\left(k^2\right)`$ $$\mathrm{\Pi }\left(k^2\right)=\frac{\alpha }{4\pi }_0^1𝑑x\left(2x1\right)^2\mathrm{ln}\left[1\frac{k^2}{m^2}x\left(1x\right)\right].$$ (23) All these results have been reproduced here without the use of operator field theory. ## III Vacuum Polarization in scalar QED with external Fields. Following the previous discussion on the free-field case we will briefly describe the effect of an external electromagnetic field on the single-loop process in scalar QED. Our starting point is a short review of the results achieved earlier by one of the present authors . There it was shown that the action in presence of a constant electromagnetic background field is given by $`\mathrm{\Gamma }_N[k_1,\epsilon _1,\mathrm{},k_N,\epsilon _N]=\left(2\pi \right)^4\delta ^4\left({\displaystyle \underset{i=1}{\overset{N}{}}}k_i\right)e^N{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{\left(4\pi \right)^2T^3}}e^{im^2T}`$ (24) $`{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^T}dt_i{\displaystyle \frac{det_P^{1/2}\left[{\displaystyle \frac{d^2}{d\tau ^2}}\right]}{det_P^{1/2}\left[\frac{d^2}{d\tau ^2}2eF\frac{d}{d\tau }\right]}}e^{{\displaystyle \frac{i}{2}}{\displaystyle _0^T}𝑑\tau {\displaystyle _0^T}𝑑\tau ^{}J^\mu \left(\tau \right)G_{\mu \nu }(\tau ,\tau ^{})J^\nu \left(\tau ^{}\right)}`$ (25) where the Green’s function equation is now slightly modified: $$\frac{1}{2}\ddot{G}eF\dot{G}=\delta \left(\tau \tau ^{}\right)\frac{1}{T}.$$ (26) $`det_P^{}`$ stands for non-zero modes in the eigenvalue problem with periodic boundary conditions, just as in the previous chapter. Also note that our former expression (12) has maintained its structure, the only difference being that the former freely circulating particle is now propagating in the external field expressed by the field-dependant determinant in (24). As shown in we can write for the ratio of the two determinants in (24) $$\frac{det_P^{1/2}\left[\frac{d^2}{d\tau ^2}\right]}{det_P^{1/2}\left[\frac{d^2}{d\tau ^2}2eF\frac{d}{d\tau }\right]}=\frac{e^2abT^2}{\mathrm{sin}\left(ebT\right)\mathrm{sinh}\left(eaT\right)}$$ (27) so that the free action of equation (12) is modified according to $`\mathrm{\Gamma }_N[k_1,\epsilon _1,\mathrm{},k_N,\epsilon _N]=\left(2\pi \right)^4\delta ^4\left({\displaystyle \underset{i=1}{\overset{N}{}}}k_i\right)e^N{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{\left(4\pi \right)^2T}}{\displaystyle \frac{e^2abe^{im^2T}}{\mathrm{sin}\left(ebT\right)\mathrm{sinh}\left(eaT\right)}}`$ (28) $`{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^T}dt_ie^{{\displaystyle \frac{i}{2}}{\displaystyle _0^T}𝑑\tau {\displaystyle _0^T}𝑑\tau ^{}J^\mu \left(\tau \right)G_{\mu \nu }(\tau ,\tau ^{})J^\nu \left(\tau ^{}\right)}`$ (29) with $`G_{\mu \nu }(\tau ,\tau ^{})`$ given as in , and a and b are expressed by the invariants $$a^2=\left(^2+𝒥^2\right)^{\frac{1}{2}}+,b^2=\left(^2+𝒥^2\right)^{\frac{1}{2}};=\frac{1}{4}F_{\mu \nu }F^{\mu \nu },𝒥=\frac{1}{4}F_{\mu \nu }^{}F^{\mu \nu }.$$ (30) Having recognized similar structures in the free-field case we now replace (16) by $`\mathrm{\Gamma }_N[k_1,\epsilon _1,\mathrm{},k_N,\epsilon _N]=\left(2\pi \right)^4\delta ^4\left({\displaystyle \underset{i=1}{\overset{N}{}}}k_i\right)e^N{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{\left(4\pi \right)^2T}}{\displaystyle \frac{e^2abe^{im^2T}}{\mathrm{sin}\left(ebT\right)\mathrm{sinh}\left(eaT\right)}}{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^T}dt_i`$ (31) $`\mathrm{exp}{\displaystyle \frac{i}{2}}{\displaystyle \underset{i,j=1}{\overset{N}{}}}\left[k_iG(t_i,t_j)k_jk_i{\displaystyle \frac{}{t_j}}G(t_i,t_j)\epsilon _j\epsilon _i{\displaystyle \frac{}{t_i}}G(t_i,t_j)k_j+\epsilon _i{\displaystyle \frac{^2}{t_it_j}}G(t_i,t_j)\epsilon _j\right].`$ (32) Our next task is to find $`\mathrm{\Gamma }_2`$. However, since the procedure to arrive at a manifestly gauge invariant expression is straightforward, we just report our findings: $`\mathrm{\Gamma }_2[k_1,\epsilon _1;k_2,\epsilon _2]=\left(2\pi \right)^4\delta ^4(k_1+k_2)e^2{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dTT}{32\pi ^2}}{\displaystyle _1^1}dv{\displaystyle \frac{e^2abe^{im^2T}}{\mathrm{sin}\left(ebT\right)\mathrm{sinh}\left(eaT\right)}}e^{i\mathrm{\Psi }}`$ (33) $`\left[\left(\epsilon _1\rho k\right)\left(\epsilon _2\rho k\right)\left(\epsilon _1\lambda k\right)\left(\epsilon _2\lambda k\right)\left(\epsilon _1\rho \epsilon _2\right)\left(k\rho k\right)\right]`$ , (34) where $`k_1kk_2`$ and e.g., $`\left(\epsilon _1\rho k\right)\epsilon _1^\mu \rho _{\mu \nu }k^\nu `$, etc. The explicit expressions for $`\rho `$ and $`\lambda `$ are given by $`\rho =C^2\zeta _3B^2\zeta _4,`$ $`\lambda =C\zeta _1B\zeta _2,`$ (35) $`\mathrm{where}\zeta _1={\displaystyle \frac{\mathrm{cosh}\left(eaT\right)\mathrm{cosh}\left(eaTv\right)}{\mathrm{sinh}\left(eaT\right)}},`$ $`\zeta _2={\displaystyle \frac{\mathrm{cos}\left(ebT\right)\mathrm{cos}\left(ebTv\right)}{\mathrm{sin}\left(ebT\right)}},`$ (37) $`\zeta _3={\displaystyle \frac{\mathrm{sinh}\left(eaTv\right)}{\mathrm{sinh}\left(eaT\right)}},`$ $`\zeta _4={\displaystyle \frac{\mathrm{sin}\left(ebTv\right)}{\mathrm{sin}\left(ebT\right)}}.`$ (38) The matrices $`C_{\mu \nu }`$ and $`B_{\mu \nu }`$ are defined in : $$F_{\mu \nu }=C_{\mu \nu }a+B_{\mu \nu }b,F_{\mu \nu }^{}=C_{\mu \nu }bB_{\mu \nu }a$$ (39) and can be expressed in terms of the field strength tensors $`F_{\mu \nu },F_{\mu \nu }^{}`$ and the invariants a and b. In the special situation of parallel electromagnetic fields E,H along the third direction, $`C_{\mu \nu }`$ and $`B_{\mu \nu }`$ are simply given by $`C_{\mu \nu }=g_\mu ^0g_\nu ^3g_\nu ^0g_\mu ^3`$, $`B_{\mu \nu }=g_\mu ^2g_\nu ^1g_\nu ^2g_\mu ^1`$. However, nowhere in deriving our expression did we employ this special field configuration. Therefore, our result is quite general and was already considered in . The same will be true for spinor QED, which will be our central result presented in the final chapter. Finally we have to define the phase in (34): $$\mathrm{\Psi }=\frac{1}{2}\left[\frac{\left(kB^2k\right)}{eb}\frac{\mathrm{cos}\left(ebT\right)\mathrm{cos}\left(ebTv\right)}{\mathrm{sin}\left(ebT\right)}+\frac{\left(kC^2k\right)}{ea}\frac{\mathrm{cosh}\left(eaT\right)\mathrm{cosh}\left(eaTv\right)}{\mathrm{sinh}\left(eaT\right)}\right].$$ (40) ## IV Vacuum Polarization in spinor QED with external Fields. Here we will start right away with the one-loop action $$\mathrm{\Gamma }\left[A\right]=\frac{1}{8\pi ^2}_0^{\mathrm{}}\frac{dT}{T^3}\frac{𝒟xe^{i{\displaystyle _0^T}𝑑\tau \left[{\displaystyle \frac{\dot{x}^2}{4}}eA\dot{x}m^2\right]}}{𝒟xe^{i{\displaystyle _0^T}𝑑\tau {\displaystyle \frac{\dot{x}^2}{4}}}}\frac{𝒟\psi e^{{\displaystyle _0^T}𝑑\tau \left[{\displaystyle \frac{1}{2}}\psi _\mu \dot{\psi }^\mu +eF_{\mu \nu }\psi ^\mu \psi ^\nu \right]}}{𝒟\psi e^{{\displaystyle \frac{1}{2}}{\displaystyle _0^T}𝑑\tau \psi \dot{\psi }}},$$ (41) where we have introduced the four Grassmann variables $`\psi _\mu (\tau )`$ which anticommute. When we use steps similar to those which took us to equation (29), we arrive again at the background-field one-loop function multiplied by the off-shell photons tied to the spinor particle loop: $`\mathrm{\Gamma }_N[k_1,\epsilon _1,\mathrm{},k_N,\epsilon _N]=\left(2\pi \right)^4\delta ^4\left({\displaystyle \underset{i=1}{\overset{N}{}}}k_i\right)e^N{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{8\pi ^2T^3}}e^{im^2T}`$ (42) $`{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^T}dt_i{\displaystyle }d\theta _id\overline{\theta }_i{\displaystyle \frac{det^{1/2}\left[{\displaystyle \frac{d^2}{d\tau ^2}}\right]}{det^{1/2}\left[{\displaystyle \frac{d^2}{d\tau ^2}}2eF{\displaystyle \frac{d}{d\tau }}\right]}}{\displaystyle \frac{det^{1/2}\left[{\displaystyle \frac{1}{2}}{\displaystyle \frac{d}{d\tau }}eF\right]}{det^{1/2}\left[{\displaystyle \frac{1}{2}}{\displaystyle \frac{d}{d\tau }}\right]}}`$ (43) $`\mathrm{exp}\left\{{\displaystyle \frac{i}{2}}{\displaystyle _0^T}𝑑\tau {\displaystyle _0^T}𝑑\tau ^{}J^\mu \left(\tau \right)G_{\mu \nu }(\tau ,\tau ^{})J^\nu \left(\tau ^{}\right){\displaystyle \frac{1}{4}}{\displaystyle _0^T}𝑑\tau {\displaystyle _0^T}𝑑\tau ^{}\eta ^\mu \left(\tau \right)\overline{G}_{F\mu \nu }(\tau ,\tau ^{})\eta ^\nu \left(\tau ^{}\right)\right\},`$ (44) with $`\left({\displaystyle \frac{1}{2}}{\displaystyle \frac{d}{d\tau }}eF\right)\overline{G}_F(\tau ,\tau ^{})=\delta \left(\tau \tau ^{}\right)`$ (45) $`\mathrm{and}\eta ^\mu \left(\tau \right)=\sqrt{2}{\displaystyle \underset{j=1}{\overset{N}{}}}\left(\theta _j\epsilon _j^\mu i\overline{\theta }_jk_j^\mu \right)\delta \left(\tau t_j\right).`$ (46) As usual, $`\theta _i`$ and $`\overline{\theta }_i`$ denote independent anticommuting Grassmann variables. Calculating the determinants yields for the one-loop action with arbitrary constant electromagnetic field $`\mathrm{\Gamma }_N[k_1,\epsilon _1,\mathrm{},k_N,\epsilon _N]=\left(2\pi \right)^4\delta ^4\left({\displaystyle \underset{i=1}{\overset{N}{}}}k_i\right)e^N{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{8\pi ^2T}}e^{im^2T}`$ (47) $`{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle _0^T}dt_i{\displaystyle }d\theta _id\overline{\theta }_ie^2ab{\displaystyle \frac{\mathrm{cos}\left(ebT\right)\mathrm{cosh}\left(eaT\right)}{\mathrm{sin}\left(ebT\right)\mathrm{sinh}\left(eaT\right)}}`$ (48) $`\mathrm{exp}\left\{{\displaystyle \frac{i}{2}}{\displaystyle _0^T}𝑑\tau {\displaystyle _0^T}𝑑\tau ^{}J^\mu \left(\tau \right)G_{\mu \nu }(\tau ,\tau ^{})J^\nu \left(\tau ^{}\right){\displaystyle \frac{1}{4}}{\displaystyle _0^T}𝑑\tau {\displaystyle _0^T}𝑑\tau ^{}\eta ^\mu \left(\tau \right)\overline{G}_{F\mu \nu }(\tau ,\tau ^{})\eta ^\nu \left(\tau ^{}\right)\right\}.`$ (49) Expressing the sources in the exponential according to equation (7) and (46) we obtain $`\mathrm{exp}\{{\displaystyle \frac{i}{2}}{\displaystyle \underset{i,j=1}{\overset{N}{}}}[k_iG(t_i,t_j)k_jk_i{\displaystyle \frac{}{t_j}}G(t_i,t_j)\epsilon _j\overline{\theta }_j\theta _j\overline{\theta }_i\theta _i\epsilon _i{\displaystyle \frac{}{t_i}}G(t_i,t_j)k_j+`$ (50) $`+\overline{\theta }_i\theta _i\overline{\theta }_j\theta _j\epsilon _i{\displaystyle \frac{^2}{t_it_j}}G(t_i,t_j)\epsilon _j]{\displaystyle \frac{1}{2}}{\displaystyle }^N_{i,j=1}[\theta _i\theta _j\epsilon _i\overline{G}_F(t_i,t_j)\epsilon _j`$ (51) $`i\theta _i\overline{\theta }_j\epsilon _i\overline{G}_F(t_i,t_j)k_ji\overline{\theta }_i\theta _jk_i\overline{G}_F(t_i,t_j)\epsilon _j\overline{\theta }_i\overline{\theta }_jk_i\overline{G}_F(t_i,t_j)k_j]\}.`$ (52) Formula (48) together with (51) is our most general representation for the spin- $`\frac{1}{2}`$ action with external fields. Let us now turn to the special case $`N=2`$. Here we obtained as a check for the free-field case $`\mathrm{\Gamma }_2[k_1,\epsilon _1;k_2,\epsilon _2]=\left(2\pi \right)^4\delta ^4(k_1+k_2)e^2{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{8\pi ^2T^2}}e^{im^2T}{\displaystyle _0^T}dt_1e^{ik_1^2G\left(t_1\right)}`$ (53) $`\left\{[(k_1\epsilon _2)(k_2\epsilon _1)(\epsilon _1\epsilon _2)(k_1k_2)](\dot{G}^21)\right\}=`$ (54) $`=\left(2\pi \right)^4\delta ^4(k_1+k_2)e^2[(k_1\epsilon _2)(k_2\epsilon _1)(\epsilon _1\epsilon _2)(k_1k_2)]`$ (55) $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dT}{16\pi ^2T}}e^{im^2T}{\displaystyle _1^1}dv(1v^2)e^{ik_1^2{\displaystyle \frac{T}{4}}\left(1v^2\right)}.`$ (56) Written in the form $`\mathrm{\Gamma }_2=\left(2\pi \right)^4\delta ^4\left(k_1+k_2\right)\epsilon _\mu \mathrm{\Pi }^{\mu \nu }\epsilon _\nu `$, this yields the well-known result for the spin- $`\frac{1}{2}`$ polarization tensor $`\mathrm{\Pi }^{\mu \nu }=\left(k^2g^{\mu \nu }k^\mu k^\nu \right)\mathrm{\Pi }\left(k^2\right),`$ (57) $`\mathrm{where}\mathrm{\Pi }\left(k^2\right)={\displaystyle \frac{2\alpha }{\pi }}{\displaystyle _0^1}𝑑xx\left(1x\right)\mathrm{ln}\left[1{\displaystyle \frac{k^2}{m^2}}x\left(1x\right)\right].`$ (58) After relatively straightforward calculations which are much easier and shorter than anything published in the literature we finally end up with the following form for the spin- $`\frac{1}{2}`$ polarization tensor with a prescribed constant electromagnetic field of any configuration: $`\mathrm{\Pi }^{\mu \nu }\left(k\right)={\displaystyle \frac{e^2}{16\pi ^2}}{\displaystyle _0^{\mathrm{}}}dTTe^{im^2T}{\displaystyle \frac{e^2ab}{\mathrm{sin}\left(ebT\right)\mathrm{sinh}\left(eaT\right)}}{\displaystyle _1^{+1}}dve^{i\mathrm{\Psi }}`$ (59) $`\{_1(g^{\mu \nu }k^2k^\mu k^\nu )+_2[\left(Ck\right)^\mu \left(Bk\right)^\nu +\left(Ck\right)^\nu \left(Bk\right)^\mu ]+`$ (60) $`+_3\left[\left(C^2k\right)^\mu \left(C^2k\right)^\nu \left(C^2\right)^{\mu \nu }\left(kC^2k\right)\right]+`$ (61) $`+_4[\left(B^2k\right)^\mu \left(B^2k\right)^\nu \left(B^2\right)^{\mu \nu }\left(kB^2k\right)]\}`$ (62) Here we introduced $`N_0_1=\mathrm{cosh}\left(eaTv\right)\mathrm{cos}\left(ebTv\right)+\mathrm{sin}\left(ebTv\right)\mathrm{sinh}\left(eaTv\right)\mathrm{coth}\left(eaT\right)\mathrm{cot}\left(ebT\right)`$ $`N_3_2=\mathrm{sinh}\left(eaTv\right)\mathrm{sin}\left(ebTv\right)+{\displaystyle \frac{\left(1\mathrm{cosh}\left(eaTv\right)\mathrm{cosh}\left(eaT\right)\right)}{\mathrm{sinh}\left(eaT\right)}}{\displaystyle \frac{\mathrm{cos}\left(ebTv\right)\mathrm{cos}\left(ebT\right)1}{\mathrm{sin}\left(ebT\right)}}`$ $`N_1_3=_1+2{\displaystyle \frac{\mathrm{cos}\left(ebT\right)}{\mathrm{sinh}^2\left(eaT\right)}}\left(\mathrm{cosh}\left(eaT\right)\mathrm{cosh}\left(eaTv\right)\right)`$ $`N_2_4=_12{\displaystyle \frac{\mathrm{cosh}\left(eaT\right)}{\mathrm{sin}^2\left(ebT\right)}}\left(\mathrm{cos}\left(ebT\right)\mathrm{cos}\left(ebTv\right)\right).`$ Rewriting the tensor structure of the coefficients $`_3`$ and $`_4`$ we obtain $`\{(\text{60})\}=N_0\left(g^{\mu \nu }k^2k^\mu k^\nu \right)+N_3\left[\left(Ck\right)^\mu \left(Bk\right)^\nu +\left(Ck\right)^\nu \left(Bk\right)^\mu \right]+`$ (63) $`+N_1\left(Ck\right)^\mu \left(Ck\right)^\nu N_2\left(Bk\right)^\mu \left(Bk\right)^\nu .`$ (64) By introducing the quantities $`N_i`$ we have made contact with the work of Urrutia . But while this author treats only parallel E and H fields we allow for any field direction. Urrutias’s result is reproduced by our formula (60) together with (64) by putting $`C_{\mu \nu }=g_\mu ^0g_\nu ^3g_\nu ^0g_\mu ^3`$, $`B_{\mu \nu }=g_\mu ^2g_\nu ^1g_\nu ^2g_\mu ^1`$. We wanted also mention that Gies has given an alternative but much more elaborate, derivation of $`\mathrm{\Pi }^{\mu \nu }`$. Hence we felt that our representation is sufficiently attractive to present it as another example that demonstrates how path integral methods together with world-line techniques can be put to work. ## Acknowledgments The authors profited from discussions with Dr. H. Gies . This work was supported by Deutsche Forschungsgemeinschaft under DFG Di 200/5-1.
warning/0001/cond-mat0001166.html
ar5iv
text
# Microwave Electrodynamics of the Electron-Doped Cuprate Superconductors Pr2-xCexCuO4-y and Nd2-xCexCuO4-y ## 1 Introduction Existing experimental evidence suggests that the pairing state symmetry in the electron-doped cuprate superconductors is s-wave in nature, in contrast to the d-wave symmetry observed in the hole-doped cuprates. There is no compelling theoretical reason for this difference. The strongest evidence for s-wave symmetry comes from measurements of the penetration depth . However, it is possible that the paramagnetism of the Nd rare-earth ion in Nd<sub>2-x</sub>Ce<sub>x</sub>CuO<sub>4-y</sub>(NCCO) may have influenced the determination of the symmetry from temperature-dependent penetration depth measurements . In order to more definitively determine the pairing state symmetry of the electron-doped cuprates, we have again studied the low-temperature behavior of the penetration depth, $`\lambda `$(T). The functional form of $`\lambda `$(T) is determined by the low-energy excitations of the system, and thus indirectly determines the pairing state symmetry of these materials. In addition, we are now able to observe the behavior of the penetration depth to a lower temperature than was done previously . This, along with the possible role of paramagnetism in previous measurements, makes it prudent to study an electron-doped system which is not strongly paramagnetic, such as the cuprate superconductor Pr<sub>2-x</sub>Ce<sub>x</sub>CuO<sub>4-y</sub> (PCCO). The samples studied were crystals grown using a directional solidification technique and have been characterized in previous studies . Typically, these samples exhibited a transition temperature of 19 K for PCCO and 24 K for NCCO with a transition width in the surface resistance of 1.5 K and residual normal state resistivity values of about 60 $`\mu \mathrm{\Omega }`$-cm. A typical sample size was 1 mm x 1 mm x 30 $`\mu `$m. The phase diagram for the electron-doped cuprates shows that a cerium dopant concentration of x=0.15 has the highest T<sub>c</sub>. Based upon the resistivity, T<sub>c</sub>, and transition width, we believe that our samples are at the optimal doping level. ## 2 Experimental Method The in-plane penetration depth, $`\lambda _{ab}`$(T), as well as the surface resistance, R<sub>S</sub>(T), were measured using a superconducting niobium microwave resonant cylindrical cavity operating at 9.6 GHz in which the TE<sub>011</sub> mode is stimulated. In this mode, the magnetic field is maximum and the electric field zero at the center of the cavity. The sample is placed on a hot finger in the center of the cavity with its c-axis aligned parallel to H<sub>rf</sub>. This induces screening currents in the copper-oxide planes of the crystal. The sample temperature was varied from 1.2 K up to above T<sub>c</sub>. As the penetration depth changes with temperature, the resonant frequency, f(T), and quality factor, Q(T), of the cavity change. By measuring these shifts in f(T) and Q(T), quantities such as the temperature dependence of $`\mathrm{\Delta }\lambda `$(T) and R<sub>s</sub>(T) can be deduced. Further details of this technique are given elsewhere . There are some important improvements to this technique that are reported here for the first time. In the past , four issues were of major concern regarding the microwave resonant cavity technique. These involved the base temperature, the reproducibility of the background, the effects of changing liquid helium hydrostatic pressure on the dimensions of the resonator, and the orientation of the sample with respect to H<sub>rf</sub>. In our present design, we are able to reach a base temperature of 1.2 K without the use of an exchange gas, which can corrupt the data. This has been accomplished by pumping on the helium bath and providing a stronger thermal link between the sample and the helium bath. By reaching a lower base temperature, we are able to explore sample properties in a region of temperature in which the distinction between the different possible pairing state symmetries is most apparent. Furthermore, it is at the lower temperatures where any paramagnetic influence would be most evident. A further improvement was realized by permanently mounting the sample rod to the base of the apparatus. To obtain useful data, it is necessary to introduce the sample to the proper position within the resonant cavity. This position has a minimal E<sub>rf</sub>, and the H<sub>rf</sub> is not only maximum, but aligned in a known direction, in our case, along the axis of the cylindrical cavity. Furthermore, by having the sample rod permanently secured, we can ensure that the sample will be introduced to the same location within the cavity every time. This is critical in order for the background measurements to be reproducible. Also, we have addressed the issue of the helium hydrostatic pressure. As the level of the liquid helium drops over the course of an experimental run, the pressure it exerts on the cavity decreases, causing the cavity to expand slightly. This results in a constant drop in resonant frequency with time. In order to minimize this effect, it is necessary to isolate the cavity from the helium bath, while still maintaining a good thermal link between the two. To achieve this, we surround the resonant cavity with a vacuum jacket. Thermal links were maintained by direct surface-to-surface contact between the cavity and this outer cylinder at certain points. The result of this improvement was to reduce any drift in frequency due to changes in the helium hydrostatic pressure to less than the random noise present in the data. Finally, the orientation of the sample in the cavity is different from previous experiments . H<sub>rf</sub> is now applied parallel to the c-axis of the sample, whereas in the past, it was parallel to the a-b plane of the crystal. This change has two effects. First, the rare-earth paramagnetism is stronger when H<sub>rf</sub> is applied in the plane . Thus, by changing the orientation we were able to significantly reduce the paramagnetic influence on our measurements. Secondly, the former orientation induced c-axis currents in the sample. This led to the simultaneous measurement of $`\lambda _{ab}`$(T) and $`\lambda _c`$(T), making it difficult to analyze the in-plane penetration depth without c-axis contributions. However, with H<sub>rf</sub> now applied parallel to the c-axis, only a-b plane currents are induced. This allows for a direct measurement of $`\lambda _{ab}`$(T). ## 3 Results The surface resistance as a function of T/T<sub>c</sub> is presented in figure 1 for one PCCO and one NCCO crystal studied. Note that both samples display a single transition at T<sub>c</sub>. Furthermore, the data have the same behavior until the PCCO crystal surface resistance saturates at the residual resistance value. The homogeneity of the crystals is attested to by the nature of the transition. Furthermore, the width of the transition, about 0.15 T<sub>c</sub>, is typical for our samples. This, along with their reasonable values of T<sub>c</sub>, indicates that the doping is near optimal levels. Therefore, the low temperature behavior of the penetration depth for these samples should be representative of the electron-doped cuprate superconductors, with minimal extrinsic influences. This is in contrast to the behavior of other crystals which displayed much broader transitions as well as step-like features in R<sub>S</sub>(T), indicative of additional superconducting transitions. These crystals also showed multiple transitions in the c-axis resistivity, were judged to be inhomogeneous, and subsequently rejected. The low temperature behavior of the change in penetration depth, $`\mathrm{\Delta }\lambda `$(T), has been used as an indication of the pairing symmetry . In the case of BCS s-wave symmetry, $`\mathrm{\Delta }\lambda `$(T) is expected to behave exponentially, the exact behavior dependent upon the zero-temperature penetration depth, $`\lambda `$(0), and the energy gap, $`\mathrm{\Delta }`$(0). The asymptotic BCS s-wave form is: $`\mathrm{\Delta }\lambda (T)=\lambda (0)(\pi \mathrm{\Delta }(0)/2k_BT)^{1/2}e^{\mathrm{\Delta }(0)/k_BT}`$ for T$`<<`$T<sub>c</sub>/2. In figure 2 we show $`\mathrm{\Delta }\lambda `$(T) for an NCCO crystal and a PCCO crystal up to 0.45T<sub>c</sub>, as well as the asymptotic BCS s-wave model using $`\lambda `$(0) = 1500 Å and 2$`\mathrm{\Delta }(0)/k_BT_c`$ = 3.50. The data for the electron-doped crystals clearly differs from a simple BCS s-wave behavior, with the PCCO data exhibiting clear deviations from an exponential behavior at T/T$`{}_{c}{}^{}<`$0.3. Nevertheless, we can force the data to fit the asymptotic s-wave BCS form, given above, allowing both $`\lambda (0)`$ and $`\mathrm{\Delta }(0)`$ to vary. For the PCCO crystal we found $`\lambda `$(0) = 1800 Å and 2$`\mathrm{\Delta }`$(0)/k<sub>B</sub>T<sub>c</sub> = 2.75. Similarly, the NCCO crystal yields $`\lambda `$(0) = 2725 Å and 2$`\mathrm{\Delta }`$(0)/k<sub>B</sub>T<sub>c</sub> = 2.35. In both cases we find small gap values, much less than the 3.5 expected from BCS theory. The gap values are also much smaller than previously reported in NCCO . Also of note in figure 2 is the upturn in the NCCO $`\mathrm{\Delta }\lambda `$(T) data at low temperatures. We believe that this upturn is due to the paramagnetism of the Nd<sup>3+</sup> ions, and its presence may explain why previous studies, which did not take this into account, found NCCO to be an s-wave superconductor. In the fit considered above, we accounted for the paramagnetism by excluding data below 0.16 T<sub>c</sub> from our analysis. At temperatures above 0.16 T<sub>c</sub>, we believe the paramagnetic influence is less than in previous studies due to the difference in orientation of the sample. It is important to realize that a similar paramagnetic upturn is not observed in the PCCO data due to the much weaker paramagnetism in this compound . Thus, the PCCO data should more closely reflect the intrinsic behavior of the penetration depth. A more detailed analysis of this data, and data on thin films, is forthcoming in which we conclude that our data fits best to a dirty d-wave model . ## 4 Discussion Alff et al. recently obtained $`\mathrm{\Delta }\lambda `$(T)/$`\lambda `$(0) data on NCCO and PCCO grain boundary junctions in thin films . They concluded that their data was consistent with an s-wave scenario. In the case of NCCO, they applied the same paramagnetic correction to their NCCO data which Cooper used to arrive at a d-wave explanation . This is evidence of how sensitive the final determination of the pairing state symmetry is to the choice of parameters for the Curie-Weiss law. Although not shown here, we can also force our $`\mathrm{\Delta }\lambda `$(T) data for a number of PCCO crystals to fit to an s-wave form. In general however, this requires using unphysically large $`\lambda `$(0) and small $`\mathrm{\Delta }`$(0) values. Alff et al. reported a value for 2$`\mathrm{\Delta }(0)/k_BT_c`$ of about 2.9. This is less than the BCS value of 3.5, but compares well with our s-wave fit gap values. However, we interpret this as a reason for rejecting an s-wave behavior. Although both we and Alff et al. obtained very similar data, we believe that the paramater values exclude a standard BCS s-wave explanation. However, there are some s-wave scenarios which might be viable. For instance, gapless superconductivity could lead to T<sup>2</sup> behavior for the temperature dependent penetration depth at low temperatures. Furthermore, the presence of magnetic impurities could potentially reduce the size of the activation gap. This could then explain the small values for $`\mathrm{\Delta }`$(0) we obtain when fitting to an s-wave functional form. Nevertheless, the data is not consistent with a large-gap isotropic BCS s-wave behavior. ## 5 Conclusion In conclusion, we have successfully employed a microwave resonant cavity system to measure the penetration depth and surface resistance of NCCO and PCCO crystals down to 1.2 K. This data supports the conjecture that rare-earth paramagnetism might affect the observed penetration depth temperature dependence in NCCO. By analyzing the functional form of this temperature dependence, we conclude that the pairing symmetry of the electron-doped cuprate superconductors is not a standard s-wave and is more likely d-wave in nature. This work was supported by NSF DMR-9732736, NSF DMR-9624021, and the Maryland Center for Superconductivity Research. We acknowledge discussions with D. H. Wu and R. Prozorov.
warning/0001/cond-mat0001059.html
ar5iv
text
# References
warning/0001/astro-ph0001046.html
ar5iv
text
# NON-THERMAL EMISSION FROM RELATIVISTIC ELECTRONS IN CLUSTERS OF GALAXIES: A MERGER SHOCK ACCELERATION MODEL ## 1 INTRODUCTION Some clusters of galaxies have diffuse non-thermal synchrotron radio halos, which extend in a $``$ Mpc scale (e.g., Giovannini et al. 1993; Röttgering et al. 1997; Deiss et al. 1997). This indicates that there exists a relativistic electron population with energy of a few GeVs (if we assume the magnetic field strength is an order of $`\mu `$G) in intracluster space in addition to the thermal intracluster medium (ICM). Furthermore, it is well known that such clusters of galaxies have evidences of recent major merger in X-ray observations (e.g., Henriksen & Markevitch 1996; Honda et al. 1996; Markevitch, Sarazin, & Vikhlinin 1999; Watanabe et al. 1999). In such clusters of galaxies with radio halos, non-thermal X-ray radiation due to inverse Compton (IC) scattering of cosmic microwave background (CMB) photons by the same electron population is expected (Rephaeli 1979). Indeed, non-thermal X-ray radiation was recently detected in a few rich clusters (e.g., Fusco-Femiano et al. 1999; Rephaeli, Gruber, & Blanco 1999; Kaastra et al. 1999) and several galaxy groups (Fukazawa 1999) although their origins are still controversial. In addition to such relatively high energy non-thermal emission, diffuse extreme ultraviolet (EUV) emission is detected from a number of clusters of galaxies (Lieu et al. 1996; Mittaz, Lieu, & Lockman 1998; Lieu, Bonamente, & Mittaz 1999). Although their origins are also unclear, one hypothesis is that EUV emission is due to IC emission of CMB. If the hypothesis is right, this indicates existence of relativistic electrons with energy of several handled MeVs in intracluster space. The origin of such relativistic electrons is still unclear. Certainly, there are point sources like radio galaxies in clusters of galaxies, which produces such a electron population. However, the electrons can spread only in a few kpc scale by diffusion during the IC cooling time. Clearly, this cannot explain typical spatial size of radio halos. One possible solution of this problem is a secondary electron model, where the electrons are produced through decay of charged pions induced by the interaction between relativistic protons and thermal protons in ICM (Dennison 1980). In the secondary electron model, however, too much gamma-ray emission is produced through neutral pion decay to fit the Coma cluster results (Blasi & Colafrancesco 1999). Moreover, the secondary electron model cannot explicitly explain the association between merger and radio halos. From N-body + hydrodynamical simulations, it is expected that there exist shock waves and strong bulk-flow motion in ICM during merger (e.g, Schindler & Müller 1993; Ishizaka & Mineshige 1996; Takizawa 1999, 2000). This suggests that relativistic electrons are produced at around the shock fronts through 1st order Fermi acceleration and that propagation of the shock waves and bulk-flow of ICM are responsible for extension of radio halos. Obviously, the merger shock acceleration model can explicitly explain the association between merger and radio halos. However, such hydrodynamical effects on time evolution and spatial distribution of relativistic electrons during merger are not properly considered in previous studies. In this paper, we investigate the evolution of a relativistic electron population and non-thermal emission in the framework of the merger shock acceleration model. We perform N-body + hydrodynamical simulations, explicitly considering the evolution of a relativistic electron population produced at around the shock fronts. The rest of this paper is organized as follows. In §2 we show order estimation about the spatial size of radio halos. In §3 we describe the adopted numerical methods and initial conditions for our simulations. In §4 we present the results. In §5 we summarize the results and discuss their implications. ## 2 ORDER ESTIMATION In this section, we estimate several kinds of spatial length scale relevant to the extent of cluster radio halos. ### 2.1 Diffusion Length According to Bohm diffusion approximation, a diffusion coefficient is, $`\kappa ={\displaystyle \frac{\eta E_\mathrm{e}c}{3eB}},`$ (1) where, $`\eta `$ is an enhanced factor from Bohm diffusion limit, $`E_\mathrm{e}`$ is the total energy of an electron, $`c`$ is the velocity of light, $`e`$ is the electron charge, and $`B`$ is the magnetic field strength of intracluster space. Since IC scattering of CMB photons is the dominant cooling process for electrons with energy of $``$ GeVs in typical intracluster conditions (Sarazin 1999), electron cooling time in this energy range is, $`t_{\mathrm{IC}}=1.1\times 10^9\mathrm{yr}({\displaystyle \frac{E_\mathrm{e}}{\mathrm{GeV}}})^1,`$ (2) where we assume that the cluster redsift is much less than unity. Thus, the diffusion length within the cooling time is $`L_{\mathrm{diff}}\sqrt{\kappa t_{\mathrm{IC}}}1.1\times 10^4\mathrm{Mpc}\left({\displaystyle \frac{\eta }{10^2}})^{1/2}\right({\displaystyle \frac{B}{\mu \mathrm{G}}})^{1/2}.`$ (3) This is much less than the spatial size of radio halos. From this reason, electrons which are leaking from point sources such like AGNs cannot be responsible for the radio halos. ### 2.2 Shock Wave Propagation Length From N-body + hydrodynamical simulations of cluster mergers, the propagation speed of shock wave is an order of $`1000`$ km s<sup>-1</sup> (Takizawa 1999). Thus, the length scale where the shock front propagates during the cooling time of equation (2) is, $`L_{\mathrm{shock}.\mathrm{prop}}1.1\mathrm{Mpc}\left({\displaystyle \frac{v_{\mathrm{shock}}}{1000\mathrm{k}\mathrm{m}/\mathrm{s}}}\right)({\displaystyle \frac{E_\mathrm{e}}{\mathrm{GeV}}})^1,`$ (4) where $`v_{\mathrm{shock}}`$ is the propagation speed of the shock front. Roughly speaking, $`L_{\mathrm{shock}.\mathrm{prop}}`$ is related to the extent of radio halos along to the collision axis since the accelerated electrons can emit synchrotron radio radiation only during the cooling time behind the shock front. ### 2.3 Spatial Size of Shock Surfaces In cluster merger, the shock front spread over in a cluster scale ($``$ Mpc). Thus, even if the shock is nearly standing, radio halos spread in a Mpc scale. Roughly speaking, this is related to the extent of radio halos perpendicular to the collision axis. In the merger shock acceleration model, therefore, propagation of shock fronts and spatial size of shock surfaces play more crucial role to the extent of radio halos than diffusion. Furthermore, the model naturally produce 3-dimensionally extended radio halos in a Mpc scale. ## 3 MODELS We consider the merger of two equal mass ($`0.5\times 10^{15}M_{}`$) subclusters. In order to calculate the evolution of ICM, we use the smoothed-particle hydrodynamics (SPH) method. Each subcluster is represented by 5000 N-body particles and 5000 SPH particles. The initial conditions for ICM and N-body components are the same as those of Run A in Takizawa (1999). The numerical methods and initial conditions for N-body and hydrodynamical parts are fully described in §3 of Takizawa (1999). Our code is fully 3-dimensional. To follow the evolution of a relativistic electron population, we should solve the diffusion-loss equation (see Longair 1994) for each SPH particle. Since the diffusion term is negligible as mentioned in §2, the equation is, $`{\displaystyle \frac{dN(E_\mathrm{e},t)}{dt}}={\displaystyle \frac{}{E_\mathrm{e}}}[b(E_\mathrm{e},t)N(E_\mathrm{e},t)]+Q(E_\mathrm{e},t),`$ (5) where $`N(E_\mathrm{e},t)dE_\mathrm{e}`$ is the total number of relativistic electrons per a SPH particle with kinetic energies in the range $`E_\mathrm{e}`$ to $`E_\mathrm{e}+dE_\mathrm{e}`$ (hereafter, we denote kinetic energy of an electron to $`E_\mathrm{e}`$), $`b(E_\mathrm{e},t)`$ is the rate of energy loss for a single electron with an energy of $`E_\mathrm{e}`$, and $`Q(E_\mathrm{e},t)dE_\mathrm{e}`$ gives the rate of production of new relativistic electrons per a SPH particle. According to the standard theory of 1st order Fermi acceleration, we assume that $`Q(E_\mathrm{e},t)E_\mathrm{e}^\alpha `$, where $`\alpha `$ is described as $`(r+2)/(r1)`$ using the compression ratio of the shock front, $`r`$. For the shocks appeared in this simulation, the ratio is roughly $`\sqrt{10}`$ (Takizawa 1999), which provides $`\alpha =2.4`$. Since it is very difficult to monitor the compression ratio at the shock front for each SPH particle per each time step, we neglect the time dependence of $`\alpha `$. The influence of the changes in $`\alpha `$ on the results is discussed in §5. The normalization of $`Q(E_\mathrm{e},t)`$ is proportional to the artificial viscous heating, which is nearly equal to the shock heating. We generate the relativistic electrons everywhere even if explicit shock structures do not appear in the simulation. We assume that sub-shock exists where a fluid element has enough viscous heating. Such sub-shocks are recognized in higher resolution simulations (e.g. Roettiger, Burns, & Stone 1999). We assume that total kinetic energy of accelerated electrons from $`E_\mathrm{e}=0`$ to $`+\mathrm{}`$ is 5% of the viscous energy, which is consistent with the recent TeV gamma ray observational results for the galactic supernova remnant SN 1006 (Tanimori et al. 1998; Naito et al. 1999). Note that equation 5 for the evolution of a relativistic electron population is linear in $`N(E_\mathrm{e},t)`$. Thus, it is easy to rescale our results of $`N(E_\mathrm{e},t)`$ if we choose other parameters for the acceleration efficiency. We neglect energy loss of thermal ICM due to the acceleration. For $`b(E_\mathrm{e},t)`$, we consider IC scattering of the CMB photons, synchrotron losses, and Coulomb losses. We neglect bremsstrahlung losses for simplicity, which is a good approximation in typical intracluster conditions (Sarazin 1999). Then, if we ignore weak energy dependence of coulomb losses, the loss function $`b(E_\mathrm{e},t)`$ becomes $`b(E_\mathrm{e},t)=b_c(t)+b_1(t)E_\mathrm{e}^2,`$ (6) where $`b_c(t)=7.0\times 10^{16}(n_\mathrm{e}(t)/\mathrm{cm}^3)`$ and $`b_1=2.7\times 10^{17}+2.6\times 10^{18}(B(t)/\mu \mathrm{G})^2`$ (if $`b(E_\mathrm{e},t)`$ and $`E_\mathrm{e}`$ are given in units of GeV s<sup>-1</sup> and GeV, respectively.). In the above expressions, $`n_\mathrm{e}`$ is the number density of ICM electrons and $`B`$ is strength of the magnetic field. To integrate equation (5) with the Courant and viscous timestep control (see Monaghan 1992), we use the analytic solution as follows. First, we integrate equation (5) from $`t`$ to $`t+\mathrm{\Delta }t`$, regarding the second term on the right-hand side as being negligible small. Then, $`N(E_\mathrm{e},t+\mathrm{\Delta }t)=\{\begin{array}{cc}N(E_{\mathrm{e},0},t)\frac{b(E_{\mathrm{e},0})}{b(E_\mathrm{e})},\hfill & \text{ (}E_\mathrm{e}<E_{\mathrm{e},\mathrm{max}}\text{), }\hfill \\ 0,\hfill & \text{ (}E_\mathrm{e}>E_{\mathrm{e},\mathrm{max}}\text{), }\hfill \end{array}`$ (9) where $`E_{\mathrm{e},0}`$ $`=`$ $`\sqrt{{\displaystyle \frac{b_c}{b_1}}}\mathrm{tan}(\mathrm{arctan}\sqrt{{\displaystyle \frac{b_1}{b_c}}}E_\mathrm{e}+\sqrt{b_cb_1}\mathrm{\Delta }t),`$ (10) $`E_{\mathrm{e},\mathrm{max}}`$ $`=`$ $`\sqrt{{\displaystyle \frac{b_c}{b_1}}}{\displaystyle \frac{1}{\mathrm{tan}\sqrt{b_1b_c}\mathrm{\Delta }t}}.`$ (11) Next, we add the contribution from the second term to the above $`N(E_\mathrm{e},t+\mathrm{\Delta }t)`$ using the second-order Runge-Kutta method. In the present simulation, $`N(E_\mathrm{e},t)`$ is calculated on logarithmically equally spaced 300 points in the range $`E_\mathrm{e}=0.05`$ to $`50`$ GeVs for each SPH particle. Magnetic field evolution is included by means of the following method. We assume initial magnetic pressure is 0.1 % of ICM thermal pressure. This corresponds to $`B=0.1\mu \mathrm{G}`$ in volume-averaged magnetic field strength. For Lagrangean evolution of $`B`$, due to the frozen-in assumption we apply $`B(t)/B(t_0)=(\rho _{\mathrm{ICM}}(t)/\rho _{\mathrm{ICM}}(t_0))^{2/3}`$. Field changes due to the passage of the shock waves is not considered in this paper. The change may depend on field configuration at the shock front and have value of $`14`$. However, it is difficult to examine it in the present simulation even under high $`\beta `$ condition. We will try this problem in the future paper. Our model implies continuous production of power law distributed relativistic electrons at around the shock fronts. This is valid only when $`\mathrm{\Delta }t_{\mathrm{acc}}`$ is sufficiently shorter than the dynamical timescale of the system ($`10^9`$ yr), where $`\mathrm{\Delta }t_{\mathrm{acc}}`$ denotes acceleration time in which $`Q(E_\mathrm{e},t)`$ becomes power law distribution. It is presented in the framework of the standard shock acceleration theory as $`\mathrm{\Delta }t_{\mathrm{acc}}=3ru^2(r1)^1(\kappa _1+r\kappa _2)`$ , where $`u`$ is the flow velocity of the upstream of the shock front, and $`\kappa _{1,2}`$ are diffusion coefficients of the upstream and downstream, respectively (see e.g. Drury 1983). Assuming $`B_1=B_2`$ and Bohm diffusion approximation of equation (1), $`\mathrm{\Delta }t_{\mathrm{acc}}=1.9\times 10^2\mathrm{yr}\left({\displaystyle \frac{E_\mathrm{e}}{\mathrm{GeV}}}\right)\left({\displaystyle \frac{\eta }{10^2}}\right)\left({\displaystyle \frac{u}{10^3\mathrm{km}\mathrm{s}^1}})^2\right({\displaystyle \frac{B}{\mu \mathrm{G}}})^1.`$ (12) This value is certainly much shorter than the dynamical timescale. ## 4 RESULTS Figure 1 shows the time evolution of non-thermal emission for various energy band: from top to bottom, IC emission of the Extreme Ultraviolet Explorer (EUVE) band (65-245 eV), soft X-ray band (4-10 keV), and hard X-ray band (10-100 keV), and synchrotron radio emission (10 MHz - 10 GHz). The times are relative to the most contracting epoch. The calculation of the luminosity for each band is performed in the simplified assumption that electrons radiate at a monochromatic energy given by $`E_\mathrm{X}=2.5\mathrm{keV}(E_\mathrm{e}/\mathrm{GeV})^2`$ and $`\nu =3.7\mathrm{MHz}(B/\mu \mathrm{G})(E_\mathrm{e}/\mathrm{GeV})^2`$ for IC scattering and synchrotron emission, respectively. Since cooling time is roughly proportional to $`E_\mathrm{e}^1`$ in these energy range, the higher the radiation energy of IC emission is, the shorter duration of luminosity increase is. In other words, luminosity maximum comes later for lower energy band. Hard X-ray and radio emissions come from the electrons with almost the same energy range. The luminosity maximum in the hard X-ray band, however, comes slightly after the most contracting epoch. On the other hand, radio emission becomes maximum at most contracting epoch since the change of magnetic field due to the compression and expression plays an more crucial role than the increase of relativistic electrons. In any cases, radio halos and hard X-ray are well associated to merger phenomena. They are observable only when thermal ICM have definite signatures of mergers such as complex temperature structures, non-spherical and elongated morphology, or substructures. Soft X-ray emission, which is observable only in clusters (or groups) with relatively low temperature ($`1`$ keV) ICM, is still luminous in $``$ 1 Gyr after the merger. Thus, the association of mergers in this band is weaker than in the hard X-ray band. Moreover, EUV emission continues to be luminous after the signatures of the merger have been disappeared in the thermal ICM. Figure 2 shows the IC spectra at $`t=0.0`$ (solid lines) and $`0.25`$ (dotted lines). In lower energies, $`L_\nu \nu ^{0.7}`$, which is originated from the electron source spectrum $`Q(E_\mathrm{e})E_\mathrm{e}^{2.4}`$. On the other hand, in higher energies, the spectrums become close to steady solution, $`L_\nu \nu ^{1.2}`$, owing to the IC and synchrotron losses (Longair 1994). The break point of the spectrum moves toward lower energies as time proceeds. Figure 3 shows the snapshots of synchrotron radio (10MHz-10GHz) surface brightness distribution (solid contours) and X-ray one of thermal ICM (dashed contours) seen from the direction perpendicular to the collision axis. Contours are equally spaced on a logarithmic scale and separated by a factor of 7.4 and 20.1 for radio and X-ray maps, respectively. At $`t=0.25`$, the main shocks are located between the two X-ray peaks and relativistic electrons are abundant there. Thus, the radio emission peak is located between the two X-ray peaks although the magnetic field strength there is weaker. At $`t=0.0`$, relativistic electrons are concentrated around the central region since the main shocks are nearly standing and located near $`X\pm 0.2`$. Furthermore, gas infall compress ICM and the magnetic field. Thus, radio distribution shows rather strong concentration. In these phase (at $`t=0.25`$ and $`0.0`$), the radio halo is located at the high temperature region of ICM. On the other hand, at $`t=0.25`$, relativistic electron distribution becomes rather diffuse since fresh relativistic electrons are producing in the outer regions as the shock waves propagate outwards. At $`t=0.25`$ the main shocks are located at $`X\pm 1`$. Between the shock fronts rather diffuse radio emission is seen. In this phase, the radio halo is located between two high temperature regions of ICM. As described above, the morphology of the radio halo is strongly depending on the phase of the merger when viewed from the direction perpendicular to the collision axis. When viewed nearly along the collision axis, however, this is not the case. Figure 4 shows the same as figure 3, but for seen from the direction tilted at an angle of $`30^{}`$ with respect to the collision axis. Radio and X-ray morphology are similar each other in all phases. When the cluster is viewed along the collision axis, the distribution of relativistic electrons roughly follows that of the thermal ICM since the shock fronts face to the observers and spread over the cluster. The distribution of magnetic field strength also roughly follows that of the thermal ICM. Therefor, the radio morphology follows X-ray one. Figure 5 shows the synchrotron radiation spectra at $`t=0.0`$ (solid lines) and $`0.25`$ (dotted lines). At $`t=0.0`$, the synchrotron spectrum follows that of IC emission. On the other hand, at $`t=0.25`$ there exits a bump at lower energies in the spectrum, which cannot be seen in that of IC emission. Emissivity of synchrotron radiation depends on not only relativistic electron density but also magnetic energy density, which is larger in the central region in this model. On the other hand, IC emissivity depends on the electron density and CMB energy density, which is homogeneous. Thus, the total synchrotron spectrum is more like that in the central region than the total IC spectrum. At $`t=0.0`$, since relativistic electrons are centered, the emission from the outer region is negligible for both spectra. Thus, similar results are given. On the other hand, at $`t=0.25`$, propagation of shocks makes a diffuse distribution of relativistic electrons. Therefor, the contribution from the outer region is not negligible for the total IC spectrum while the total synchrotron spectrum is still biased the central region as seen in figure 3. In the central region, however, electrons produced by the main shocks at $`t0`$ with energies more than $``$ several GeVs have already cooled down. Thus the spectrum in the central region have a bump in lower energies, which is present in the total spectrum. The emission above $`30`$ MHz is mainly due to the electrons produced by the sub-shocks there. If such sub-shocks do not exist, the emission near the main shocks, where the magnetic field strength is rather weaker, can be seen in this energy range. Note that this feature of the synchrotron spectrum is sensitive to the spatial distribution of magnetic field. ## 5 CONCLUSIONS AND DISCUSSION We have investigated evolution of non-thermal emission from relativistic electrons accelerated at around the shock fronts during merger of clusters of galaxies. Hard X-ray and radio radiations are luminous only while merger signatures are left in thermal ICM. Hard X-ray radiation becomes maximum in the later stage of merger. In our simulation, radio emission is the most luminous at the most contracting epoch. This is due to the magnetic field amplification by compression. According to the recent magnetohydrodynamical simulations (Roettiger, Stone, & Burns 1999), however, it is possible that the field amplification occurs as the bulk flow is replaced by turbulent motion in the later stages of merger. If this is effective in real clusters, radio emission can increase by a factor of two or three than our results in the later stages of merger. EUV emission is still luminous after the merger signatures have been disappeared in thermal ICM. This is consistent with the EUVE results. Morphological relation between radio halos and ICM hot regions is described as follows. In the contracting phase, radio halos are located at the hot regions of thermal ICM, or between two substructures (see the left panel of figure 3). This may correspond to A2256 (Röttgering et al. 1994). In the expanding phase, on the other hand, radio halos are located between the two hot regions of ICM and show rather diffuse distribution (see the right panel of figure 3). This may correspond to Coma (Giovannini et al. 1993; Deiss et al. 1997) and A2319 (Feretti, Giovannini, & Böhringer 1997). In the further later phase, the shock fronts reach outer regions and the GeV electrons are already cooled in the central parts. Then, radio halos are located in the cluster outer regions near the shock fronts and we cannot detect radio emission in the central part of the cluster. However, observational correlation between ICM hot regions and radio halos is not clear since the electron temperature there is significantly lower than the plasma mean temperature due to the relatively long relaxation time between ions and electrons (Takizawa 1999, 2000). Note that until now we could only find the electron temperature through X-ray observations. This may correspond to A3667 (Röttgering et al. 1997). It is possible that such radio ‘halos’ located in the outer regions are classified into radio ‘relics’ since their radio powers and spatial scales becomes weaker and smaller than those of typical radio halos, respectively. When the cluster is viewed nearly along the collision axis, however, such morphological relations between radio halo and ICM are unclear and radio and X-ray distributions become similar each other. We neglect the changes of the spectral index in the electron source term. Since the mach number is gradually increasing as merger proceeds (Takizawa 1999), the spectrum of relativistic electrons becomes flatter as time proceeds. We believe that such changes in the spectral index does not influence our results seriously because most of relativistic electrons are produced in the central high density region, where the mach number is almost constant. In the later stage of merger, however, contribution of relativistic electrons produced in the outer region cannot be negligible in higher energy range ($`10`$ GeV) since cooling time is relatively short. Thus, it is probable that the inverse Compton spectrum in the hard X-ray ($`10100`$ keV) becomes flatter in the later stages. The lower energy part of the electron spectrum can emit hard X-ray through bremsstrahlung. Whether IC scattering or bremsstrahlung is dominant in the hard X-ray range is depending on the shape of the electron spectrum. Roughly speaking, when the spectrum of relativistic electrons is flatter than $`E_\mathrm{e}^{2.5}`$, IC scattering dominates the other and vice versa (see Appendix). Furthermore, the shape of the electron spectrum in the lower energy part is flatter than the originally injected form since the cooling time due to the coulomb loss is proportional to $`E_\mathrm{e}`$ and very short (Sarazin 1999). In the present simulation, therefor, it is most likely that the contribution of the bremsstrahlung components in the hard X-ray range is negligible. More detailed calculations, including nonlinear effects for the shock acceleration (Jones & Ellison 1991), by Sarazin & Kempner (1999) show that IC scattering is dominant in the hard X-ray when the accelerated electron momentum spectrum is flatter than $`p_\mathrm{e}^{2.7}`$, which corresponds to the electron energy spectrum of $`E_\mathrm{e}^{2.7}`$ in the fully relativistic range. Thus, the bremsstrahlung contribution in the hard X-ray should be considered in mergers with low mach numbers. A merger shock acceleration model also predicts some gamma-ray emission. Electrons which radiate EUV due to IC scattering also emit $`100`$ MeV gamma-ray through bremsstrahlung. Furthermore, it is most likely that protons as well as electrons are accelerated at around the shock fronts. Such high energy protons also produce gamma-rays peaked at $`100`$ MeV through decay of neutral pions. Although the energy density ratio between electrons and protons in acceleration site is uncertain, the contributions of protons and the bremsstrahlung to the emission become important in the higher energies. We think that it is interesting to investigate from hundreds MeV to multi TeV emissions, which are observable with operating instruments such as EGRET, ground-based air Čerenkov telescopes, and planning projects like GLAST satellites. However, since the diffusion length of protons within the cooling time is much longer than that of electrons, treatment of the diffusion-loss equation for protons are more complex than the model in this paper. We would like to thank Drs. Y. Fukazawa and S. Shibata for helpful comments. MT thanks Drs. S. Mineshige and T. Shigeyama for continuous encouragement. MT is also grateful to S. Tsubaki for fruitful discussion. ## Appendix A ESTIMATION OF THE BREMSSTRAHLUNG CONTRIBUTION IN THE HARD X-RAY RANGE We estimate the contribution of the bremsstrahlung form the lower energy part of the electron spectrum to the hard X-ray emission in our model, which is neglected in this paper. Although the crude estimations discussed here are order-of magnitudes, it is helpful to explain which mechanism, IC scattering and bremsstrahlung radiation, is dominant. The standard theory of 1st order Fermi acceleration provides power law spectrum in momentum distribution. We, therefore, assume the momentum spectrum of accelerated electrons has a form of $`{\displaystyle \frac{dN_\mathrm{e}}{dP_\mathrm{e}}}=cN_0\left({\displaystyle \frac{P_\mathrm{e}}{m_\mathrm{e}c}}\right)^\alpha [\mathrm{cm}^3(\mathrm{eV}c^1)^1],`$ (A1) where $`P_\mathrm{e}`$ is an electron momentum, $`c`$ is the speed of light, and $`m_\mathrm{e}`$ is the electron rest mass. Using the relation between the momentum $`P_\mathrm{e}`$ and the kinetic energy $`E_\mathrm{e}`$, $`E_\mathrm{e}=(P_\mathrm{e}^2c^2+m_\mathrm{e}^2c^4)^{1/2}m_\mathrm{e}c^2`$, the spectrum for kinetic energy is given by $`{\displaystyle \frac{dN_\mathrm{e}}{dE_\mathrm{e}}}=N_0{\displaystyle \frac{E_\mathrm{e}+m_\mathrm{e}c^2}{m_\mathrm{e}c^2}}\left({\displaystyle \frac{P_\mathrm{e}}{m_\mathrm{e}c}}\right)^{\alpha 1}[\mathrm{cm}^3(\mathrm{eV})^1].`$ (A2) For relativistic electrons, $`E_\mathrm{e}m_\mathrm{e}c^2`$, the spectrum becomes $`{\displaystyle \frac{dN_\mathrm{e}}{dE_\mathrm{e}}}=N_0\left({\displaystyle \frac{E_\mathrm{e}}{m_\mathrm{e}c^2}}\right)^\alpha ,`$ (A3) which is consistent with an assumption of $`Q(E_\mathrm{e},t)`$ in section 3. Since the hard X-ray photons are produced from CMB photon field via IC scattering of relativistic electrons in our model, we use this form for the estimation of IC emissivity. For non-relativistic electrons, $`E_\mathrm{e}m_\mathrm{e}c^2`$, the spectrum becomes $`{\displaystyle \frac{dN_\mathrm{e}}{dE_\mathrm{e}}}=2^{\frac{\alpha +1}{2}}N_0{\displaystyle \frac{E_\mathrm{e}+m_\mathrm{e}c^2}{m_\mathrm{e}c^2}}\left({\displaystyle \frac{E_\mathrm{e}}{m_\mathrm{e}c^2}}\right)^{\frac{\alpha +1}{2}},`$ (A4) where we use the relation $`E_\mathrm{e}=P_\mathrm{e}^2/(2m_\mathrm{e})`$. Since the bremsstrahlung radiation at the hard X-ray range is emanated from electrons with almost the same energy range, we use this form for the estimation of bremsstrahlung emissivity. For simplicity, we approximate the emissivity, $`ϵ`$, for both IC and bremsstrahlung processes to be, $`ϵ={\displaystyle \frac{dN_\mathrm{e}}{dE_\mathrm{e}}}{\displaystyle \frac{dE_\mathrm{e}}{d\epsilon _\gamma }}\left|{\displaystyle \frac{dE_\mathrm{e}}{dt}}\right|[\mathrm{erg}\mathrm{s}^1\mathrm{cm}^3\mathrm{eV}^1],`$ (A5) where the emission rate is assumed to be equal to the electron energy loss rate and $`\epsilon _\gamma `$ is the photon energy. To obtain the bremsstrahlung emissivity, we assume that an electron with energy $`E_\mathrm{e}`$ loses its energy to emit photon of energy $`\epsilon _\gamma =E_\mathrm{e}`$ after it has traversed one mean free path $`X_0`$. Hence, $`{\displaystyle \frac{dE_\mathrm{e}}{d\epsilon _\gamma }}=1,`$ (A6) and $`\left|{\displaystyle \frac{dE_\mathrm{e}}{dt}}\right|\epsilon _\gamma {\displaystyle \frac{v_\mathrm{e}}{X_0}}[\mathrm{erg}\mathrm{s}^1],`$ (A7) where $`v_\mathrm{e}`$ is electron velocity. From $`\sigma _\mathrm{T}n_0X_01`$, we approximate $`\left|{\displaystyle \frac{dE_\mathrm{e}}{dt}}\right|\epsilon _\gamma v_\mathrm{e}\sigma _\mathrm{T}n_0,`$ (A8) where $`\sigma _\mathrm{T}`$ denotes the cross section of Thomson scattering and $`n_0`$ denotes the density of ambient matter. Using equations (A4), (A5), (A6), and (A8), we estimate the bremsstrahlung emissivity in the hard X-ray energy range, $`\epsilon _\gamma =\epsilon _{\mathrm{HXR}}`$, as $`ϵ_{\mathrm{brem}}2^{\frac{\alpha }{2}}N_0\sigma _\mathrm{T}cn_0m_\mathrm{e}c^2\left(1+{\displaystyle \frac{\epsilon _{\mathrm{HXR}}}{m_\mathrm{e}c^2}}\right)\left({\displaystyle \frac{\epsilon _{\mathrm{HXR}}}{m_\mathrm{e}c^2}}\right)^{\frac{\alpha }{2}+1},`$ (A9) For IC scattering of CMB photons, the photon energy after the scattering by an electron with energy $`E_\mathrm{e}=m_\mathrm{e}\gamma _\mathrm{e}c^2`$ is approximated by single energy of $`\epsilon _\gamma =\gamma _{\mathrm{e}}^{}{}_{}{}^{2}\overline{\epsilon }_{\mathrm{CMB}}`$, where $`\overline{\epsilon }_{\mathrm{CMB}}`$ is peak energy of CMB spectrum. Thus, $`{\displaystyle \frac{dE_\mathrm{e}}{d\epsilon _\gamma }}={\displaystyle \frac{1}{2}}\left({\displaystyle \frac{\overline{\epsilon }_{\mathrm{CMB}}}{m_\mathrm{e}c^2}}\right)^{\frac{1}{2}}\left({\displaystyle \frac{\epsilon _\gamma }{m_\mathrm{e}c^2}}\right)^{\frac{1}{2}}.`$ (A10) Since the scattering is in Thomson energy range ($`\overline{\epsilon }_{\mathrm{CMB}}\gamma _\mathrm{e}m_\mathrm{e}c^2`$), we set $`\left|{\displaystyle \frac{dE_\mathrm{e}}{dt}}\right|={\displaystyle \frac{4}{3}}\sigma _\mathrm{T}c\gamma _{\mathrm{e}}^{}{}_{}{}^{2}\overline{\epsilon }_{\mathrm{CMB}}n_{\mathrm{CMB}}[\mathrm{erg}\mathrm{s}^1],`$ (A11) where $`n_{\mathrm{CMB}}`$ is photon number density of CMB fields. Using equations (A3), (A5), (A10), and (A11), we estimate the IC emissivity in the hard X-ray energy range, $`\epsilon _\gamma =\epsilon _{\mathrm{HXR}}`$, as $`ϵ_{\mathrm{IC}}={\displaystyle \frac{2}{3}}N_0\sigma _\mathrm{T}c\overline{\epsilon }_{\mathrm{CMB}}n_{\mathrm{CMB}}\left({\displaystyle \frac{\overline{\epsilon }_{\mathrm{CMB}}}{m_\mathrm{e}c^2}}\right)^{\frac{\alpha 3}{2}}\left({\displaystyle \frac{\epsilon _{\mathrm{HXR}}}{m_\mathrm{e}c^2}}\right)^{\frac{\alpha 1}{2}}.`$ (A12) From equations (A9) and (A12), we derive the emissivity ratio as $`{\displaystyle \frac{ϵ_{\mathrm{brem}}}{ϵ_{\mathrm{IC}}}}={\displaystyle \frac{3}{2}}{\displaystyle \frac{n_0}{n_{\mathrm{CMB}}}}\left(1+{\displaystyle \frac{\epsilon _{\mathrm{HXR}}}{m_\mathrm{e}c^2}}\right)\left({\displaystyle \frac{\overline{\epsilon }_{\mathrm{CMB}}}{m_\mathrm{e}c^2}}\right)^{\frac{1}{2}}\left({\displaystyle \frac{\epsilon _{\mathrm{HXR}}}{m_\mathrm{e}c^2}}\right)^{\frac{1}{2}}\left(2{\displaystyle \frac{\overline{\epsilon }_{\mathrm{CMB}}}{m_\mathrm{e}c^2}}\right)^{\frac{\alpha }{2}}.`$ (A13) For $`ϵ_{\mathrm{brem}}<ϵ_{\mathrm{IC}}`$, we get the relation $`\alpha <{\displaystyle \frac{2\mathrm{ln}\left[\frac{3}{2}\frac{n_0}{n_{\mathrm{CMB}}}\left(1+\frac{\epsilon _{\mathrm{HXR}}}{m_\mathrm{e}c^2}\right)\frac{(\overline{\epsilon }_{\mathrm{CMB}}\epsilon _{\mathrm{HXR}})^{\frac{1}{2}}}{m_\mathrm{e}c^2}\right]}{\mathrm{ln}\left(\frac{2\overline{\epsilon }_{\mathrm{CMB}}}{m_\mathrm{e}c^2}\right)}}.`$ (A14) Substituting typical values for ICM at $`z0`$, $`n_0=1\times 10^3\mathrm{cm}^3`$, $`n_{\mathrm{CMB}}=400\mathrm{cm}^3`$, $`\overline{\epsilon }_{\mathrm{CMB}}=6.57\times 10^4\mathrm{eV}`$, and $`\epsilon _{\mathrm{HXR}}=10\mathrm{keV}`$, we obtain $`\alpha <2.5`$. At $`\epsilon _{\mathrm{HXR}}=100\mathrm{keV}`$, the spectrum of equation (A9) becomes steeper as electrons enter the trans-relativistic energy range. As a result, the bremsstrahlung emissivity is reduced so that the larger index of electron spectrum is accepted for $`ϵ_{\mathrm{brem}}<ϵ_{\mathrm{IC}}`$.
warning/0001/nlin0001019.html
ar5iv
text
# Parameter estimation in nonlinear stochastic differential equations ## I Introduction The analysis of complex systems by nonlinear deterministic differential equations has attracted much attention in recent years. Given a parameterized differential equation, best-fit parameters can be obtained by a least-squares minimization, see e.g. . Often, however, the dynamics does not follow a strict deterministic law. In dissipative systems, due to fluctuation-dissipation theorems, thermal noise might have to be taken into account . Furthermore, in open nonequilibrium systems like in biology and economy, the dynamics is often also exposed to highdimensional, effectively random, influences, see for an example. This calls for a description by nonlinear stochastic differential equations (SDE), respectively their corresponding Fokker-Planck equations . While parameter estimation in linear stochastic and nonlinear deterministic differential equations even for data covered by additive observational noise is well known, see for a review, the estimation of parameters in nonlinear SDEs is still under development. In this paper we discuss three approaches for parameter estimation in nonlinear SDEs. The results carry over to modeling by Fokker-Planck equations. The organization of this paper is as follows: In the next section, we briefly summarize the theory of integrating SDEs and exemplify its practice for the van der Pol oscillator undergoing stochastic forcing. In Section III we discuss three approaches for parameter estimation in nonlinear stochastic differential equations. In Section IV we compare two of these approaches in a simulation study using the stochastic van der Pol oscillator. ## II Integrating nonlinear stochastic differential equations ### A Theory A stochastic differential equation (SDE) with parameter vector $`\stackrel{}{\theta }`$ is given by: $$\dot{\stackrel{}{x}}=\stackrel{}{a}(\stackrel{}{\theta },\stackrel{}{x})+\underset{¯}{b}(\stackrel{}{\theta },\stackrel{}{x})\stackrel{}{ϵ},$$ (1) where $`\stackrel{}{ϵ}`$ denotes uncorrelated Gaussian noise or, in mathematical terms, the increment of Brownian motion. This general form includes additive and multiplicative dynamical noise. $`\stackrel{}{a}(\stackrel{}{\theta },\stackrel{}{x})`$ is usually denoted as (deterministic) drift term, $`\underset{¯}{b}(\stackrel{}{\theta },\stackrel{}{x})`$ is called the diffusion term. The integration of a SDE is not straightforward. This is due to the mathematical problems of evaluating integrals which involves the dynamical noise $`\stackrel{}{ϵ}`$, see for a brief introduction and for a detailed discussion. Applying the same ideas underlying higher order integration schemes for deterministic differential equation like Runge-Kutta to SDEs leads to hardly treatable stochastic integrals given in . Thus, only low order integration schemes can be used. The lowest order so-called Euler-scheme for Eq. (1) is given by : $$\stackrel{}{x}(t+\delta t)=\stackrel{}{x}(t)+\delta t\stackrel{}{a}(\stackrel{}{\theta },\stackrel{}{x}(t))+\sqrt{\delta t}\underset{¯}{b}(\stackrel{}{\theta },\stackrel{}{x}(t))\stackrel{}{ϵ}(t),$$ (2) which is of order $`1/2`$ for multiplicative noise and of order $`1`$ for additive noise. The characteristics of integration schemes for SDEs is the $`\sqrt{\delta t}`$ term which results from the integration rules for white noise . For the task of parameter estimation, we assume that the system under observation can be adequately described by Eq. (1) with unknown parameter vector $`\stackrel{}{\theta }`$. The process is observed at discrete time points given by the sampling interval $`\mathrm{\Delta }t`$. For nonlinear deterministic differential equations it is usually possible to chose identical time steps for the integration and the sampling. The necessity of using a low order scheme for SDEs means that the integration step size $`\delta t`$ is usually much smaller than the sampling interval $`\mathrm{\Delta }t`$ by which the time series is recorded. Thus, while Eq. (2) can be written as $$\stackrel{}{x}(t+\delta t)=\stackrel{}{h}(\stackrel{}{x}(t))+\nu (t),$$ (3) with $`\stackrel{}{h}(\stackrel{}{x}(t))`$ a function that can be related back to the parameter vector $`\stackrel{}{\theta }`$ of $`\stackrel{}{a}(\stackrel{}{\theta },\stackrel{}{x}(t))`$ and $`\nu (t)`$ uncorrelated Gaussian noise, on the time scale $`\mathrm{\Delta }t`$ of sampling the relationships are more intricate. While it is still possible to formulate the dependence between present and future states as $$\stackrel{}{x}(t+\mathrm{\Delta }t)=\stackrel{}{g}(\stackrel{}{x}(t))+\eta (t),$$ (4) the parameter vector $`\stackrel{}{\theta }`$ of $`\stackrel{}{a}(\stackrel{}{\theta },\stackrel{}{x}(t))`$ can, in general, not be inferred from $`\stackrel{}{g}(\stackrel{}{x}(t))`$. Furthermore, $`\eta (t)`$, while still uncorrelated with zero mean, does in general not represent Gaussian noise. In other words, in nonlinear SDEs the relation between the mean of the conditional density $`p(\stackrel{}{x}(t+\mathrm{\Delta }t)|\stackrel{}{x}(t))`$ and the drift-term $`\stackrel{}{a}(\stackrel{}{\theta },\stackrel{}{x}(t))`$ is not explicitly known (The analogous problem is given for modeling such systems by their corresponding Fokker-Planck equation). Furthermore, the conditional density $`p(\stackrel{}{x}(t+\mathrm{\Delta }t)|\stackrel{}{x}(t))`$ is, in general, not Gaussian although $`\stackrel{}{ϵ}(t)`$ in Eqs. (1,2) is Gaussian. For these reasons parameter estimation in SDEs is a non-trivial task. ### B An example To exemplify the practical issues of integrating SDEs, we choose the van der Pol oscillator : $$\ddot{x}=\mu (1x^2)\dot{x}x,\mu >0.$$ (5) This system exhibits a limit cycle due to the amplitude dependent change of the sign of the damping term. We expose it to a random force of unit variance, leading to $`\dot{x}_1`$ $`=`$ $`x_2`$ (6) $`\dot{x}_2`$ $`=`$ $`\mu (1x_1^2)x_2x_1+ϵ,`$ (7) where $`x_1`$ denotes the location and $`x_2`$ the velocity. The Euler integration scheme for Eqs. (6,7) is given by $`x_1(t+\delta t)`$ $`=`$ $`x_1(t)+\delta tx_2(t)`$ (8) $`x_2(t+\delta t)`$ $`=`$ $`x_2(t)+\delta t(\mu (1x_1^2(t))x_2(t)x_1(t))+\sqrt{\delta t}ϵ(t).`$ (9) In the following we chose $`\mu =3`$. To obtain a sampled time series of the system, one has to decide on $`\delta t`$ and $`\mathrm{\Delta }t`$. $`\mathrm{\Delta }t`$ should be chosen so that the process is reasonably sampled. Characteristics of the stochastic van der Pol oscillator and related perturbed limit cycles have been investigated in . These authors have shown that the mean period of this system is slightly smaller than the period of the corresponding deterministic system which is approximately 9 s. By choosing $`\mathrm{\Delta }t=0.5`$, we obtain approximately 18 data points per mean period. The choice of the integration step size $`\delta t`$ is more difficult. For deterministic systems adaptive algorithms are well established that guarantee an upper bound for the deviation from the true trajectory . No corresponding straightforward procedure is available for SDEs. For SDEs the characteristic quantities are the conditional densities $`p(\stackrel{}{x}(t+\mathrm{\Delta }t)|\stackrel{}{x}(t))`$. To obtain a sampled trajectory that can be regarded as a realization of the SDE, $`\delta t`$ has to be chosen such small that the conditional densities become independent of $`\delta t`$ for smaller values. Figs. 1 and 2 shows this procedure for the first component of $`\stackrel{}{x}(t+\mathrm{\Delta }t)`$ of the stochastic van der Pol oscillator denoted by $`x_1(t+\mathrm{\Delta }t)`$ for the two state vectors $`\stackrel{}{x}(t)=(0.0935,4.284)`$ and $`\stackrel{}{x}(t)=(1.021,0.9375)`$. The conditional densities $`p(x_1(t+\mathrm{\Delta }t)|\stackrel{}{x}_i(t))`$ were estimated by triangular kernel estimators based on 5000 integrations of the SDE. The quality of these estimated densities with respect to bias and variance depends on the width of the kernel. By visual inspection, the width of the kernel was chosen equal to 0.02 in Fig. 1 and equal to 0.1 in Fig. 2. In Fig. 1, the estimated conditional density changes drastically between $`\delta t=0.1s`$ and $`\delta t=0.01s`$. It becomes independent from $`\delta t`$ for $`\delta t=0.001s`$. In Fig. 2, this already happens for $`\delta t=0.01s`$. The reason for the different behavior is that the time evolution in the first case experiences more of the nonlinearity of the system. By investigating numerous analogous simulations, we regarded $`\delta t=0.001s`$ as an appropriate integration time step. The two state space vectors $`\stackrel{}{x}(t)`$ used for the above simulation were realizations of the trajectory obtained with $`\delta t=0.001s`$. Thus, the procedure for determining $`\delta t`$ is an interactive one that has to be selfconsistent. Fig. 3 shows a realization of the stochastic van der Pol oscillator with $`\mu =3`$, $`\delta t=0.001`$ and $`\mathrm{\Delta }t=0.5`$. The two data points marked by arrows at time 28s, respectively 28.5s were used for the estimation of the conditional densities shown in Figs. 1 and 2. ### C A consequence The above considerations also have consequences for mathematical models used in simulation studies and proposed for analyzing time series in the frame of nonlinear stochastic systems by difference equation of the type : $$x(t+1)=f(\stackrel{}{p},x(t),x(t1),\mathrm{},x(tm))+ϵ(t).$$ (10) Here the sampling interval is set to unity, $`ϵ(t)`$ denotes uncorrelated Gaussian noise, $`\stackrel{}{p}`$ the parameters and $`m`$ the order of the model, see e.g. . For these models the conditional distribution of $`x(t+1)`$ given the history is Gaussian. Due to the Gaussianity, least squares optimization leads to a maximum likelihood estimation of the parameters which is asymptotically unbiased and efficient . Figs. 1 and 2 show that for difference equations that are thought to be discretized versions of SDEs, the dynamical noise should be non-Gaussian, see the skewed distribution in Fig. 2, and state dependent heteroscadistic. It might even be multimodal. ## III Parameter estimation In this section we discuss three methods to estimate parameters in SDEs. Due to its superior statistical properties, the most desirable method would be a maximum likelihood estimation . We argue that this approach is not feasible. Then we discuss recently suggested quasi-maximum likelihood approaches . The third approach applies the integration scheme, Eq. (2), for the whole sampling interval by using $`\mathrm{\Delta }t=\delta t`$. With respect to the identification of the two time scales this last approach is similar to a procedure to estimate parameters in SDEs proposed in . In the simulation studies presented in these publications the time series were sampled at the time step of integration. To simplify the exposition, we consider a scalar dynamics and a single parameter $`\theta `$ in the following. ### A Maximum likelihood estimation Denoting the stationary distribution of the SDE for a given parameter $`\theta `$ by $`\pi (x|\theta )`$, the likelihood for a sampled time series of length $`N`$ reads : $$(x(t_1),x(t_2),\mathrm{},x(t_N);\theta )=\pi (x(t_1)|\theta )\underset{i=1}{\overset{N1}{}}p(x(t_{i+1})|x(t_i),\theta ).$$ (11) Maximizing $`(.;\theta )`$ leads to an estimator $`\widehat{\theta }`$ that is asymptotically unbiased and has least conservative confidence regions. Note that a biased estimator can lead to erroneous interpretations of results, while suboptimal confidence regions ’only’ lowers the power to reject simpler models in favor of more complex ones, but does not lead to statistically false positive results. Usually, the logarithm of the likelihood is considered and the term $`\pi (x(t_1)|\theta )`$ whose influence vanishes asymptotically is neglected: $$L(x(t_2),x(t_3),\mathrm{},x(t_N)|\theta )=\underset{i=1}{\overset{N1}{}}\mathrm{ln}p(x(t_{i+1})|x(t_i),\theta ).$$ (12) The maximum likelihood estimator $`\widehat{\theta }_{mle}`$ is defined by: $$\frac{d}{d\theta }L(x(t_2),x(t_3),\mathrm{},x(t_N);\theta )=0.$$ (13) To obtain $`\widehat{\theta }_{mle}`$ an iterative optimization strategy has to be applied. Starting from an initial guess for the parameter, the conditional densities $`p(x(t_{i+1})|x(t_i),\theta )`$ in Eq. (12) have to be estimated and evaluated. This can be achieved by solving the corresponding Fokker-Planck equation or by applying the integration scheme, Eq. (2), with the actual guess of the parameters several times on the intervals $`[t_i,t_{i+1}]`$ starting from the observed $`x(t_i)`$ to estimate the conditional density, e.g. by kernel estimation . Then, the log-likelihood can be calculated. Based on this procedure, the parameter is changed until the log-likelihood is extremal by applying some optimization algorithm . The performance of this procedure depends heavily on the quality of the density estimation. Thus, for the procedure based on integrating the SDE thousands of trajectories in each interval $`[t_{i_1},t_i]`$ have to be realized. Note that the densities shown in Figs. 1 and 2 that were based on 5000 realizations are not smooth enough to enable a numerically stable estimation of the parameters. Furthermore, this procedure involves the choice of a parameter determining the bandwidth of the density estimator. This parameter has to be chosen data driven and state dependent, e.g. by a computer intensive cross-validation procedure . Therefore, the desirable method of maximum likelihood is hardly applicable. ### B Quasi maximum likelihood estimation Bibby and Sorensen suggested to use a quasi maximum likelihood estimator instead of the infeasible maximum likelihood estimator, Eq. (13). The key idea of this procedure is that Eq. (13) can formally be read as a search for a root : $$G(\theta )=0,$$ (14) defining an estimator $`\widehat{\theta }`$. By virtue of choosing $`G(\theta )`$, the resulting computationally feasible estimator can be forced to be unbiased by paying the price that the resulting confidence region are no longer optimal. Experience shows that the loss in optimality is often rather small . Eq. (14) is called estimation equation, and the resulting estimator is called quasi maximum likelihood estimator. Bibby and Sorensen have shown that in the case of SDEs one possible choice for $`G(\theta )`$ is given by: $$G(\theta )=\underset{i=1}{\overset{N1}{}}g_i(.)(x(t_{i+1})E(x(t_{i+1})|x(t_i),\theta )),$$ (15) where $`g_i(.)`$ is a function that is derived from the SDE and $`E(x(t_{i+1})|x(t_i),\theta )`$ denotes the expected value of $`x(t_{i+1})`$ conditioned on the observed $`x(t_i)`$ for a given $`\theta `$. Different possible procedures for deriving $`g_i(.)`$ given the SDE are discussed in . In opposite to the conditional density necessary for the maximum likelihood case, the conditional expectation in Eq. (15) can the estimated reliably with a comparable small numbers of integrations in the intervals $`[t_i,t_{i+1}]`$ starting from the observed $`x(t_i)`$. ### C $`\mathrm{\Delta }t=\delta t`$” approach In the third approach we use the discretization scheme, Eq. (2), on the interval of the whole sampling time step $`\mathrm{\Delta }t`$ by setting $`\mathrm{\Delta }t=\delta t`$ in Eq. (2). This yields an estimate $`\widehat{x}((t_{i+1})|x(t_i),\theta )`$. The parameters are adjusted until the mean square error $$\underset{i=1}{\overset{N1}{}}(x(t_{i+1})\widehat{x}((t_{i+1})|x(t_i),\theta ))^2$$ (16) is minimized. This method neglects the existence of the two time scales for integrating and sampling SDEs. Furthermore, it implicitly assumes that the variance of the conditional density is state space independent and Gaussian. ## IV Simulation study We investigate the behavior of the quasi maximum likelihood and the ”$`\mathrm{\Delta }t=\delta t`$” approach in a simulation study using the stochastic van der Pol oscillator introduced in Section II B. We integrated the process with $`\delta t=0.001s`$. The smaller the sampling time, the smaller the differences between $`\stackrel{}{h}(.)`$ in Eq. (3) and $`\stackrel{}{g}(.)`$ in Eq. (4) will be. Thus, the less severe the effects of ignoring the two time scales by the ”$`\mathrm{\Delta }t=\delta t`$” approach will be . Therefore, we investigate the behavior of the two approaches described above for different sampling times between $`\mathrm{\Delta }t=0.005s`$ and $`\mathrm{\Delta }t=0.5s`$. As in the simulation study presented in , we assume that the whole state space vector is observed. For the quasi maximum likelihood approach, Section III B, we chose the term $`g_i(.)`$ in Eq. (15) to be: $$g_i(\stackrel{}{x}(t_i))=(1x_1^2)x_2,$$ (17) generalizing the line of argument of . The term ”$`(x(t_{i+1})E(x(t_{i+1})|x(t_i),\theta )`$” in Eq. (15) is read as $`(x_1(t_{i+1})E(x_1(t_{i+1})|\stackrel{}{x}(t_i),\theta )`$. Thus, we predict the observed location $`x_1(t_{i+1})`$ based on the whole state space vector at the present time step $`t_i`$. The expected value of $`x_1(t_{i+1})`$ is estimated based on 50 integrations. Finding the zero of Eq. (15) is performed by routines from . For the ”$`\mathrm{\Delta }t=\delta t`$” approach, Section III C, we use the information from the whole state space from the history and from the present state. The minimization is perform by routines from . Fig. 4 displays the results of the simulation study for sampling time steps $`\mathrm{\Delta }t`$ ranging from $`0.005`$ to $`0.5`$. The 2$`\sigma `$ confidence intervals were calculated from 50 repetitions. The length of the time series for different sampling times were chosen such that 1000 data points enters the estimation. The quasi maximum likelihood approach yields unbiased results independent from the sampling interval. For realistic sampling intervals the ”$`\mathrm{\Delta }t=\delta t`$” approach gives strongly biased results. Only for a sampling time of $`\mathrm{\Delta }t0.025`$ its estimate is consistent with the true parameter. This would require to sample the process with approximately 360 points per period. Note that classical time series like the sunspot data and the Canadian lynx data are sampled with 11, respectively 10 points per period, see e.g. . Fig. 5 shows the results analogous to Fig. 4 for nonlinearity parameter $`\mu =5`$. The relative bias of the ”$`\mathrm{\Delta }t=\delta t`$” approach for sampling intervals $`\delta t=0.05s,0.1s`$ and $`0.5s`$ is larger than for $`\mu =3`$ due to the higher degree of nonlinearity. Thus, in general, a sufficient sampling for the linearly approximating ”$`\mathrm{\Delta }t=\delta t`$” approach to work depends on the degree of nonlinearity. ## V Discussion Modeling time series of open nonequilibrium systems by nonlinear stochastic differential equations (SDE) allows to take into account the effects of the huge number of degrees of freedom that might be active in such systems. A fundamental problem in estimating parameters in SDEs is taught by the theory of integrating nonlinear stochastic differential equations. A central message from this theory is that there are in general two time scales: that of integration and that of sampling. As a consequence, the mean and the distributional characteristics of conditional densities on the time scale of sampling can, in general, not be related back to the parameters of the stochastic differential equation (SDE). We discussed three approaches for parameter estimation in SDEs based on sampled time series. Unfortunately, the desirable maximum likelihood approach is not feasible in the present context due to its extreme computational burden to estimate the conditional distributions. It might become tractable with the advent of more powerful massive parallel computers. In the presented simulation study, the quasi maximum likelihood approach yielded unbiased estimates for the parameter. Disregarding the existence of the two time scales of integration and of sampling in SDEs and using the discretization scheme of the SDE itself on the time scale of the sampling led to results that are unbiased only for the case of heavy oversampling, but biased for conventional sampling rates. Thus, methods that require the sampling time interval to be an admissable integration time step, like suggested in , are applicable to measured time series only for sufficiently sampled data. The same holds for the desirable approach to nonparametrically estimate the complete functional form of the deterministic drift term of a stochastic differential equation based on the discretization of the corresponding Fokker-Planck equation proposed in where formally even the limit ”sampling interval $`0`$” is required. The approaches discussed in this paper require to observe the complete state space vector which is usually not possible. Our future work will concentrate on the generalization to the case of parameter estimation in nonlinear stochastic differential equations based on scalar observations as it is already possible in linear stochastic and nonlinear deterministic systems . ## Figure captions * Estimated conditional densities $`p(x_1(t+\mathrm{\Delta }t)|\stackrel{}{x}(t))`$ for the stochastic van der Pol oscillator for $`\mu =3`$, $`\stackrel{}{x}(t)=(0.0935,4.284)`$, integration time steps $`\delta t=0.1s,0.01s,0.001s,0.0001s`$ and sampling interval $`\mathrm{\Delta }t=0.5s`$. * Estimated conditional densities $`p(x_1(t+\mathrm{\Delta }t)|\stackrel{}{x}(t))`$ for the stochastic van der Pol oscillator for $`\mu =3`$, $`\stackrel{}{x}(t)=(1.021,0.9375)`$, integration time steps $`\delta t=0.1s,0.01s,0.001s,0.0001s`$ and sampling interval $`\mathrm{\Delta }t=0.5s`$. * Realization of the stochastic van der Pol oscillator for $`\mu =3`$, integration time step $`\delta t=0.001s`$ and sampling interval $`\mathrm{\Delta }t=0.5s`$. The two data points marked by arrows at time 28s, respectively 28.5s were used for the estimation of the conditional densities shown in Figs. 1 and 2. * Dependence of the estimated parameter $`\widehat{\mu }`$ of the stochastic van der Pol oscillator on the sampling interval. The integration time step $`\delta t`$ was $`0.001s`$. The 95% confidence intervals were calculated based on 50 repetitions. +: quasi maximum likelihood approach, $``$: ”$`\mathrm{\Delta }t=\delta t`$” approach. For sake of clearity the results are slightly scattered around the applied sampling time steps $`\mathrm{\Delta }t=0.005s,0.01s,0.025s,0.05s,0.1s`$ and $`0.5s`$. The true value of the parameter $`\mu =3`$ is marked by the solid line. * Analogous to Fig. 4 for $`\mu =5`$.
warning/0001/hep-lat0001010.html
ar5iv
text
# Large-𝑞 expansion for the second moment correlation length in the two-dimensional 𝑞-state Potts model ## 1 Introdunction The $`q`$-state Potts model in two dimensions exhibits the first order phase transition for $`q>4`$. Many quantities was solved exactly at the phase transition point, among which the correlation length $`\xi ^{(d)}`$ in the disordered phase is included. On the other hand, no analytic result had been known for the correlation length $`\xi ^{(o)}`$ in the ordered phase at the phase transition point. Here the correlation length means the standard one defined from the ratio of the largest to the second largest eigenvalues of the transfer matrix and it is often called the true correlation length or the exponential correlation length. Janke and Kappler examined the decay rate of the correlation function by the Monte Carlo simulation with large sizes of the lattice for $`q=10,15`$ and $`20`$ giving a result that was consistent with $`\xi ^{(o)}=\xi ^{(d)}`$ and they made a conjecture that this relation would be exact. Iglói and Carlon evaluated the largest and the second largest eigenvalues of the transfer matrix for $`q=9`$ by the density matrix renormalization group technique with a result that supports this conjecture. Recently the author investigated analytically the large-$`q`$ behavior of the eigenvalues of the transfer matrix. He found that from the second largest to the $`L`$-th largest eigenvalues with $`L`$ the size of the lattice make a continuum spectrum for the thermodynamic limit both in the ordered and disordered phases at least in the large-$`q`$ region and that $`\xi ^{(o)}=\xi ^{(d)}`$ at least to order $`z^{3/2}`$ (i.e., in the first 4 terms of the large-$`q`$ expansion). Here we calculate the large-$`q`$ expansion of the second moment correlation length $`\xi _{2nd}`$ at the first order phase transition point both in the ordered and disordered phases of this model. The second moment correlation length also gives important informations on the spectrum of the eigenvalues of the transfer matrix. The large-$`q`$ expansion of the Potts model in two dimensions was calculated to order $`z^{10}`$ ($`z1/\sqrt{q}`$) for the energy cumulants including the specific heat by Bhattacharya, Lacaze and Morel using a graphical method. Tabata and the author extended the large-$`q`$ series to order $`z^{23}`$ using the finite lattice method. They also calculated the large-$`q`$ series for the magnetization cumulants including the magnetic susceptibility to order $`z^{21}`$. These long series enabled them to present the estimate for the numerical values of the cumulants which is orders of magnitude more precise than that of Monte Carlo simulations. The series also served to convince us that the correctness of the conjecture by Bhattacharya, Lacaze and Morel on the behavior of the divergent quantities in the limit of $`q4`$. It is rather straightforward to apply the finite lattice method to the large-$`q`$ expansion of the second moment correlation length, since the algorithm is quite parallel with that of the finite lattice method for the low temperature expansion of the second moment correlation length for the Ising model in three dimensions. The situation is in contrast to the case of the standard correlation length where the continuum spectrum of the eigenvalues of the transfer matrix prevented us from applying the method used to obtain the low temperature series for the Ising model in three dimensions. The results of the large-$`q`$ expansion for the second moment correlation length was reported breafly in reference and here we describe them in detail. ## 2 Expansion series by the finite lattice method The model is defined on the $`L_x\times L_y`$ rectangular lattice by the partition function $$Z=\underset{\{s_i\}}{}\mathrm{exp}\left\{\beta \underset{i,j}{}\delta _{s_i,s_j}+\underset{i}{}(h+\gamma _1x_i+\gamma _2y_i+\eta 𝒓_i^2)\delta _{s_i,1}\right\},$$ (1) where the spin variable $`s_i`$ at each site $`i`$ takes the values $`1,2,\mathrm{},q`$, $`i,j`$ represents the pair of nearest neighbor sites and $`𝒓_i=(x_i,y_i)`$ is the coordinate of the site $`i`$. The phase transition point $`\beta _t`$ for $`h=\gamma _1=\gamma _2=\eta =0`$ is given by $`\mathrm{exp}(\beta _t)1=\sqrt{q}`$. The fixed boundary condition should be taken for the ordered phase in which all the spins outside the $`L_x\times L_y`$ lattice are fixed to be $`\{s_i=+1\}`$, and the free boundary condition should be taken for the disordered phase. The second moment correlation length squared is defined by $$\xi _{2nd}^2=\frac{\mu _2}{2d\mu _0},$$ (2) where $`\mu _2`$ is the second moment of the correlation function $$\mu _2=\underset{L_x,L_y\mathrm{}}{lim}(L_xL_y)^1\underset{i,j}{}(𝒓_i𝒓_j)^2\delta _{s_i,1}\delta _{s_j,1}_c,$$ (3) $`\mu _0`$ is the zeroth moment of the correlation function, i.e., the magnetic susceptibility, and $`d(=2)`$ is the dimensionality of the lattice. The second moment $`\mu _2`$ can be obtained by the derivative of the free energy density as $`\mu _2`$ $`=`$ $`\underset{L_x,L_y\mathrm{}}{lim}(L_xL_y)^1`$ (4) $`\times 2\left({\displaystyle \frac{^2}{h\eta }}{\displaystyle \frac{^2}{\gamma _1^2}}{\displaystyle \frac{^2}{\gamma _2^2}}\right)\mathrm{ln}Z(\beta ,h,\eta ,\gamma _1,\gamma _2)|_{h=\eta =\gamma _1=\gamma _2=0}.`$ The algorithm of the finite lattice method to generate the large-$`q`$ expansion series for the second moment $`\mu _2`$ is the following. We define $`H(l_x,l_y)`$ for each $`l_x\times l_y`$ lattice ($`1l_xL_x,1l_yL_y`$) as $$H(l_x,l_y)=2\left(\frac{^2}{h\eta }\frac{^2}{\gamma _1^2}\frac{^2}{\gamma _2^2}\right)\mathrm{ln}Z(l_x,l_y)|_{h=\eta =\gamma _1=\gamma _2=0},$$ (5) where $`Z(l_x,l_y)`$ is the partition function for the $`l_x\times l_y`$ lattice with the fixed and free boundary condition for the ordered and disordered phase, respectively, and define $`W`$ of each lattice recursively as $`W(l_x,l_y)`$ $`=`$ $`H(l_x,l_y)`$ (6) $`{\displaystyle \underset{\genfrac{}{}{0pt}{}{1l_x^{}l_x,1l_y^{}l_y}{(l_x^{},l_y^{})(l_x,l_y)}}{}}(l_xl_x^{}+1)(l_yl_y^{}+1)W(l_x^{},l_y^{}).`$ Note that $`H(l_x,l_y)`$ and $`W(l_x,l_y)`$ depend on the size $`l_x`$ and $`l_y`$ but not on the position of the origin of the coordinate. Then the second moment of the correlation function in the thermodynamic limit is given by $$\mu _2=\underset{l_x,l_y}{}W(l_x,l_y).$$ (7) We can prove that the Taylor expansion of $`W(l_x,l_y)`$ with respect to $`z1/\sqrt{q}`$ includes the contribution from all the clusters of polymers in the standard cluster expansion that can be embedded into the $`l_x\times l_y`$ lattice but cannot be embedded into any of its rectangular $`l_x^{}\times l_y^{}`$ sub-lattices with $`l_x^{}l_x,l_y^{}l_y`$. It is straightforward to understand from the discussion for the large-$`q`$ expansion of the magnetic susceptibility in reference that the series expansion of $`W(l_x,l_y)`$ starts from the order of $`z^n`$ with $`n=l_x+l_y`$ in the ordered phase and $`n=l_x+l_y2`$ in the disordered phase, respectively. So in order to obtain the expansion series to order $`z^N`$, we should take all the finite-size rectangular lattices for the summation in equation (7) that satisfy $`l_x+l_yN`$ in the ordered phase and $`l_x+l_y2N`$ in the disordered phase, respectively. Using these algorithm of the finite lattice method we have calculated the series for the second moment $`\mu _2`$ at the first order phase transition point to order $`N=21`$ in $`z`$ both in the ordered and disordered phases as $$\mu _2=\underset{n}{}a_nz^n.$$ (8) The coefficients of the series are listed in table 1. We have checked that all of $`W(l_x,l_y)`$ with $`l_x+l_y2N`$ for the ordered phase and with $`l_x+l_yN`$ for the disordered phase have the correct order in $`z`$ as described above. Combining with the large-$`q`$ series of the magnetic susceptibility given in reference, we obtain the series for the second moment correlation length squared $`\xi _{2nd}^2`$ as $$\xi _{2nd}^2=\underset{n}{}b_nz^n,$$ (9) which are listed in table 2. The obtained expansion coefficients for the ordered and disordered phases coincide with each other to order $`z^3`$ and differ a bit from each other in higher orders. ## 3 Analysis of the series It is known that in the limit of the large correlation length $$\frac{\xi _{2nd}^2}{\xi _1^2}\frac{_{j=1}^{\mathrm{}}c_j^2(\xi _j/\xi _1)^3}{_{j=1}^{\mathrm{}}c_j^2(\xi _j/\xi _1)}$$ (10) with $`\xi _j\mathrm{ln}(\mathrm{\Lambda }_j/\mathrm{\Lambda }_0)`$ and $`c_j=<\mathrm{\Lambda }_0|𝒪|\mathrm{\Lambda }_j>`$ where $`\mathrm{\Lambda }_j(j=0,1,2,\mathrm{})`$ are the eigenvalues of the transfer matrix with $`\mathrm{\Lambda }_0`$ the largest one, $`|\mathrm{\Lambda }_j>`$ is the eigenstate of the transfer matrix corresponding to the eigenvalue $`\mathrm{\Lambda }_j`$ and $`𝒪=_i\delta _{s_i,1}`$ with the summation for $`i`$ running over the sites with a fixed coordinate of $`y_i`$. The $`\xi _1\mathrm{ln}(\mathrm{\Lambda }_1/\mathrm{\Lambda }_0)`$ is the standard correlation length. Later on we will call the eigenstate $`\mathrm{\Lambda }_1`$ the first excited state and $`\mathrm{\Lambda }_j`$($`i=2,3,\mathrm{}`$) the higher excited states. The standard correlation length $`\xi _1^{(d)}`$ at the phase transition point in the disordered phase is known exactly with its asymptotic behavior for $`q4`$ as $$\xi _1^{(d)}\frac{1}{8\sqrt{2}}\mathrm{exp}\left(\frac{\pi ^2}{2\theta }\right),$$ (11) where $`2\mathrm{cosh}\theta =\sqrt{q}`$ ($`\theta \sqrt{q4}/2`$ for $`q4`$). It is quite natural to expect that all of the $`\xi _j`$’s behave like $$\xi _j^{(o,d)}C_j^{(o,d)}\mathrm{exp}\left(\frac{\pi ^2}{2\theta }\right)$$ (12) for $`q4`$ both in the ordered and disordered phases, in which case $$\xi _{2nd}^{(o,d)^2}A^{(o,d)}\mathrm{exp}\left(\frac{\pi ^2}{\theta }\right).$$ (13) To check the validity of this conjecture we follow the method in reference adopted to analyze the large-$`q`$ series of the energy and magnetization cumulants including the specific heat and the magnetic susceptibility. The method was used to convince the validity of the Bhattacharya-Lacaze-Morel conjecture on the asymptotic behavior of these quantities for $`q4`$. The latent heat $``$ is known exactly with the asymptotic form for $`q4`$ as $$3\pi \mathrm{exp}\left(\frac{\pi ^2}{4\theta }\right),$$ (14) so, if $`\xi _{2nd}^{(o,d)}^2`$ has the asymptotic form in equation (13), the product $`\xi _{2nd}^2^p`$ is a smooth function of $`z`$ for $`q4`$ when $`p=4`$, and the Padé approximants of the large-$`q`$ series for this product are expected to converge for $`p=4`$. The results are given in figure 1. We can see that the convergence is really good around $`p=4`$ both in the ordered and disordered phases. The estimated values of the second moment correlation length by setting $`p=4`$ are listed in table.3. We note that the ratio of the estimated values of the second moment correlation length in the ordered to disordered phases $`\xi _{2nd}(\text{ordered})/\xi _{2nd}(\text{disordered})`$ is not far from unity. The ratio is 0.935(5) even for $`q4`$. Another interesting and important quantity is the ratio of the second moment correlation length $`\xi _{2nd}`$ to the standard correlation length $`\xi _1`$. From equation (10) we know that the ratio $`\xi _{2nd}/\xi _1`$ should be less than unity in the limit of the large correlation length. If the higher excited states ($`i=2,3,\mathrm{}`$) did not contribute so much, this ratio would be close to unity, as is the case in the Ising model on the simple cubic lattice, where $`\xi _{2nd}/\xi _1=0.970(5)`$ at the critical point in the disordered phase . In figure 2 we plot the ratio $`\xi _{2nd}/\xi _1`$ for the Potts model in the disordered phase. We use the value of $`\xi _{2nd}`$ estimated above from the Padé analysis of $`\xi _{2nd}^2^4`$ and the exact value of $`\xi _1`$. The ratio is much less than unity in the region of $`q`$ where the correlation length is large enough. It approaches $`0.505(6)`$ for $`q4`$. This result implies that the contribution of the higher excited states is important in the Potts model in two dimensions. It is consistent with the fact that the eigenvalue of the transfer matrix for the first excited state locates at the edge of the continuum spectrum of the eigenvalues of the transfer matrix at least in the large-$`q`$ region and it strongly suggests that these excited states ($`i=1,2,\mathrm{},L_x`$) form a continuum spectrum not only in the large-$`q`$ region but also even in the limit of $`q4`$. ## 4 Summary We calculated the large-$`q`$ expansion of the second moment correlation length in the ordered and disordered phases of the $`q`$-state Potts model in two dimensions and found that they coincide with each other to the third term of the expansion but differ a little from each other in higher orders. This suggests that, although the first few terms of the large-$`q`$ expansion for the standard correlation length really coincide in the ordered and disordered phases, their higher order terms might be different from each other. Note, however, that it is not conclusive, since the second moment correlation length depends not only on the spectrum of the eigenvalues of the transfer matrix but also on the overlapping amplitude $`c_i`$ which appeared in equation (10). Numerically the ratio $`\xi _{2nd}^{(o)}/\xi _{2nd}^{(d)}`$ is not far from unity even in the limit of $`q4`$. We also found that $`\xi _{2nd}^{(d)}/\xi _1^{(d)}`$ is far from unity for all region of $`q>4`$. It implies that higher excited states give significant contributions as well as the first excited state and it strongly suggests that these excited states form a continuum spectrum (i.e., there is no particle state in the language of the field theory) not only in the large-$`q`$ region but also in all the region of $`q>4`$. Finally we point out that it is worthwhile to reanalyze the previous Monte Carlo data for the correlation function on the assumption that this would be the case. ## Acknowledgments The author would like to thank W. Janke, A. Sokal and K. Tabata for valuable discussions.
warning/0001/hep-ph0001129.html
ar5iv
text
# hep-ph/0001129FTUV/00-04IFIC/00-04 Neutrino Masses and Mixing one Decade from Now ## 1 Indications for Neutrino Mass: Two–Neutrino Analysis Neutrinos are the only massless fermions predicted by the Standard Model. This seemed to be a reasonable assumption as none of the laboratory experiments designed to measure the neutrino mass have found any positive evidence for a non-zero neutrino mass. However, the confidence on the masslessness of the neutrino is now under question due to the important results of underground experiments, starting by the geochemical experiments of Davis and collaborators till the more recent Gallex, Sage, Kamiokande and Super–Kamiokande (SK) experiments solarexp ; atmexp ; sk99 . Altogether they provide solid evidence for the existence of anomalies in the solar and the atmospheric neutrino fluxes which could be explained by the hypothesis of neutrino oscillations which requires the presence of neutrino masses and mixings. Particularly relevant has been the recent confirmation by the SK collaboration sk99 of the atmospheric neutrino zenith-angle-dependent deficit which strongly indicates towards the existence of $`\nu _\mu `$ conversion. Together with these results there is also the indication for neutrino oscillations in the $`\overline{\nu }_\mu \overline{\nu }_e`$ channel by the LSND experiment lsnd . We first review our present knowledge of the present experimental status for the different evidences and present the results of the different analysis in the framework of two–neutrino oscillations. ### 1.1 Solar Neutrinos At the moment, evidence for a solar neutrino deficit comes from five experiments solarexp : Homestake, Kamiokande, SK and the radiochemical Gallex and Sage experiments. The most recent data on the rates can be summarized as: Clorine $`2.56\pm 0.23\text{SNU}`$ Gallex and Sage $`72.2\pm 5.6\text{SNU}`$ Super–Kamiokande $`(2.45\pm 0.08)\times 10^6\text{cm}\text{-2}\text{s}\text{-1}`$ Super-Kamiokande has also measured the the time dependence of the event rates during the day and night, as well as a measurement of the recoil electron energy spectrum and has also presented preliminary results on the seasonal variation of the neutrino event rates, an issue which will become important in discriminating the MSW scenario from the possibility of neutrino oscillations in vacuum. The different experiments are sensitive to different parts of the energy spectrum of solar neutrinos and putting all these results together seems to indicate that the solution to the problem is not astrophysical but must concern the neutrino properties. Moreover, non-standard astrophysical solutions are strongly constrained by helioseismology studies Bahcall98 . Within the standard solar model approach, the theoretical predictions clearly lie far from the best-fit solution what leads us to conclude that new particle physics is the only way to account for the data. The standard explanation for this deficit is the oscillation of $`\nu _e`$ to another neutrino species either active or sterile. In Fig. 1 we show the allowed two-neutrino oscillation regions obtained in our updated global analysis of the solar neutrino data oursun ; ourfour for both MSW msw as well as vacuum oscillations vacuum into active or sterile neutrinos. These results indicate that for oscillations into active neutrinos there are four possible solutions for the parameters: * vacuum (also called “just so”) oscillations with $`\mathrm{\Delta }m_{ei}^2=(0.5`$$`8)\times 10^{10}`$ eV<sup>2</sup> and $`\mathrm{sin}^2(2\theta )=0.5`$$`1`$ * non-adiabatic-matter-enhanced oscillations (SMA) via the MSW mechanism with $`\mathrm{\Delta }m_{ei}^2=(0.4`$$`1)\times 10^5`$ eV<sup>2</sup> and $`\mathrm{sin}^2(2\theta )=(1`$$`10)\times 10^3`$, and * large mixing (LMA) via the MSW mechanism with $`\mathrm{\Delta }m_{ei}^2=(0.2`$$`5)\times 10^4`$ eV<sup>2</sup> and $`\mathrm{sin}^2(2\theta )=0.6`$$`1`$. * low mass solution (LOW) via the MSW mechanism with $`\mathrm{\Delta }m_{ei}^2=(0.5`$$`2)\times 10^7`$ eV<sup>2</sup> and $`\mathrm{sin}^2(2\theta )=0.8`$$`1`$. For oscillations into an sterile neutrino there are differences partly due to the fact that now the survival probability depends both on the electron and neutron density in the Sun but mainly due to the lack of neutral current contribution to the water cerencov experiments. Unlike active neutrinos which lead to events in the Kamiokande and SK detectors by interacting via neutral current with the electrons, sterile neutrinos do not contribute to the SK event rates. Therefore a larger survival probability for $`{}_{}{}^{8}B`$ neutrinos is needed to accommodate the measured rate. As a consequence a larger contribution from $`{}_{}{}^{8}B`$ neutrinos to the Chlorine and Gallium experiments is expected, so that the small measured rate in Chlorine can only be accommodated if no $`{}_{}{}^{7}Be`$ neutrinos are present in the flux. This is only possible in the SMA solution region, since in the LMA and LOW regions the suppression of $`{}_{}{}^{7}Be`$ neutrinos is not enough. Vacuum oscillations into sterile neutrinos are also ruled out with more than 99% CL. ### 1.2 Atmospheric Neutrinos Atmospheric showers are initiated when primary cosmic rays hit the Earth’s atmosphere. Secondary mesons produced in this collision, mostly pions and kaons, decay and give rise to electron and muon neutrino and anti-neutrinos fluxes. There has been a long-standing anomaly between the predicted and observed $`\nu _\mu `$ $`/\nu _e`$ ratio of the atmospheric neutrino fluxes atmexp . Although the absolute individual $`\nu _\mu `$ or $`\nu _e`$ fluxes are only known to within $`30\%`$ accuracy, different authors agree that the $`\nu _\mu `$ $`/\nu _e`$ ratio is accurate up to a $`5\%`$ precision. In this resides our confidence on the atmospheric neutrino anomaly (ANA), now strengthened by the high statistics sample collected at the SK experiment sk99 . The most important feature of the atmospheric neutrino data is that it exhibits a zenith-angle-dependent deficit of muon neutrinos which is inconsistent with expectations based on calculations of the atmospheric neutrino fluxes. This experiment has marked a turning point in the significance of the ANA. The most likely solution of the ANA involves neutrino oscillations. In principle we can invoke various neutrino oscillation channels, involving the conversion of $`\nu _\mu `$ into either $`\nu _e`$ or $`\nu _\tau `$ (active-active transitions) or the oscillation of $`\nu _\mu `$ into a sterile neutrino $`\nu _s`$ (active-sterile transitions) atmours . Oscillations into electron neutrinos are nowadays ruled out since they cannot describe the measured angular dependence of muon-like contained events atmours . Moreover the most favoured range of masses and mixings for this channel have been excluded by the negative results from the CHOOZ reactor experiment chooz . In Fig. 2 we show the allowed neutrino oscillation parameters obtained in a recent global fit of the full data set of atmospheric neutrino data on vertex contained events at IMB, Nusex, Frejus, Soudan, Kamiokande atmexp and SK experiments sk99 as well as upward going muon data from SK, Macro and Baksan experiments in the different oscillation channels. The two panels corresponding to oscillations into sterile neutrinos in Fig. 2 differ in the sign of the $`\mathrm{\Delta }m^2`$ which was assumed in the analysis of the matter effects in the Earth for the $`\nu _\mu \nu _s`$ oscillations. Notice that in all channels where matter effects play a role the range of acceptable $`\mathrm{\Delta }m^2`$ is shifted towards larger values, when compared with the $`\nu _\mu \nu _\tau `$ case. This follows from looking at the relation between mixing in vacuo and in matter. In fact, away from the resonance region, independently of the sign of the matter potential, there is a suppression of the mixing inside the Earth. As a result, there is a lower cut in the allowed $`\mathrm{\Delta }m^2`$ value, and it lies higher than what is obtained in the data fit for the $`\nu _\mu \nu _\tau `$ channel. Concerning the quality of the fits our results show that the best fit to the full sample is obtained for the $`\nu _\mu \nu _\tau `$ channel although from the global analysis oscillations into sterile neutrinos cannot be ruled out. This arises mainly from the fact that due to matter effects the distribution for upgoing muons in the case of $`\nu _\mu \nu _s`$ are flatter than for $`\nu _\mu \nu _\tau `$ lipari . Data show a somehow steeper angular dependence which can be better described by $`\nu _\mu \nu _\tau `$. This leads to the better quality of the global fit in this channel. Pushing further this feature Super-Kamiokande collaboration has presented a preliminary partial analysis of the angular dependence of the through-going muon data in combination with the up-down asymmetry of partially contained events which seems to exclude the possibility $`\nu _\mu \nu _s`$ at the 2–$`\sigma `$ level sk99 . ### 1.3 LSND Los Alamos Meson Physics Facility (LSND) has searched for $`\overline{\nu }_\mu \overline{\nu }_e`$ oscillations with $`\overline{\nu }_\mu `$ from $`\mu ^+`$ decay at rest lsnd . The $`\overline{\nu }_e`$’s are detected in the quasi elastic process $`\overline{\nu }_epe^+n`$ in correlation with a monochromatic photon of $`2.2`$ MeV arising from the neutron capture reaction $`npd\gamma `$. In Ref. lsnd they report a total of 22 events with $`e^+`$ energy between 36 and $`60`$ MeV while $`4.6\pm 0.6`$ background events are expected. They fit the full $`e^+`$ event sample in the energy range $`20<E_e<60`$ MeV by a $`\chi ^2`$ method and the result yields $`64.3_{16.7}^{+18.5}`$ beam-related events. Subtracting the estimated neutrino background with a correlated gamma of $`12.5\pm 2.9`$ events results into an excess of $`51.8_{16.9}^{+18.7}\pm 8.0`$ events. The interpretation of this anomaly in terms of $`\overline{\nu }_\mu \overline{\nu }_e`$ oscillations leads to an oscillation probability of ($`0.31_{0.10}^{+0.11}\pm 0.05`$)%. Using a likelihood method they obtain a consistent result of ($`0.27_{0.12}^{+0.12}\pm 0.04`$)%. In the two-family formalism this result leads to the oscillation parameters shown in Fig. 3. The shaded regions are the 90 % and 99 % likelihood regions from LSND. Also shown are the limits from BNL776, KARMEN1, Bugey, CCFR, and NOMAD. ## 2 $`\nu `$–Oscillation Searches at Reactor and Accelerator Experiments There are two types of laboratory experiments to search for neutrino oscillations. In a disappearance experiment one looks for the attenuation of a neutrino beam primarily composed of a single flavour due to the mixing with other flavours. On the other hand in an appearance experiment one searches for interactions by neutrinos of a flavour not present in the neutrino beam. Several experiments have been searching for these signatures without any positive observation. Their results are generally presented as exclusion areas in the two-neutrino oscillation approximation. From these figures it is possible to obtain the limits obtained by the experiments on the corresponding transition probabilities. This is the relevant quantity when interpreting the sensitivities in the framework of more–than–two–neutrino mixing. In table 1 we show the limits on the different transition probabilities from the negative results of the most restricting short baseline experiments. Due to the short path length of the neutrino in this experiments they are not sensitive to the low values of $`\mathrm{\Delta }m^2`$ invoked to explain both the solar and the atmospheric neutrino data. For the determination of the neutrino mass matrix structure the most important short baseline experiment will be upcoming MiniBooNE miniboone experiment the first stage of the which is scheduled to start taking data in 2001. It searches for $`\nu _e`$ appearance in the Fermilab $`\nu _\mu `$ beam and it is specially designed to make a conclusive statement about the LSND’s neutrino oscillation evidence. In Fig. 4 we show the 90% CL limits that MiniBooNE can achieve. Should a sign be found then the next step would be the BooNe experiment. Smaller values of $`\mathrm{\Delta }m^2`$ can be accessed at reactor experiments due to the lower neutrino beam energy as well as future long baseline experiments due to the longer distance travelled by the neutrino. In table 2 we show the limits on the different transition probabilities from the negative results of the reactor experiments Bugey and CHOOZ as well as the expected sensitivities at future long baseline experiments both at accelerator and reactors. ## 3 Three–Neutrino Oscillations In the previous section we have discussed the evidences for neutrino masses and mixings as usually formulated in the two neutrino oscillation scenario. We want now to fit all the different evidences in a common framework and see what is our present knowledge of the neutrino mixing and masses and how this may be improved by the upcoming experiments. In doing so it is of crucial relevance the confirmation or reprobation of the LSND result by the MiniBooNE experiment. The three evidences can be interpreted in terms of neutrino oscillations but with the need of three different mass scales. Thus if the LSND is ruled out by the MiniBooNE experiment we could fit both solar and atmospheric data in terms of three–neutrino oscillations. If, on the contrary, LSND result stands the test of time, this would be a puzzling indication for the existence of a light sterile neutrino and the need to work in a four–neutrino framework. In the first case, i.e. three–neutrino framework, the evolution equation for the three neutrino flavours can be written as: $$i\frac{d\nu }{dt}=\left[U\frac{M_\nu }{2E}U^{}+H_{int}\right],$$ (1) where $`M_\nu `$ is the diagonal mass matrix for the three neutrinos and $`U`$ is the unitary matrix relating the flavour and the mass basis. $`H_{int}`$ is the Hamiltonian describing the neutrino interactions. In general $`U`$ contains 3 mixing angles and 1 or 3 CP violating phases depending on whether the neutrinos are Dirac or Majorana (For a detail discussion see Ref.Pilar ) . Here we will neglect the CP violating phases as they are not accessible by the existing experiments. We define the unitary matrix as $$U=R_{23}(\theta _{23})\times R_{13}(\theta _{13})\times R_{12}(\theta _{12}),$$ (2) where $`R_{ij}`$ is a rotation matrix in the plane $`ij`$. In this framework a neutrino of definite flavour $`\nu _\alpha `$, after travelling a distance $`L`$ in vacuum, can be detected in the charged-current (CC) interaction $`\nu N^{}l_\beta N`$ with a probability $$P_{\alpha \beta }=\delta _{\alpha \beta }4\underset{i=1}{\overset{n}{}}\underset{j=i+1}{\overset{n}{}}\text{Re}[U_{\alpha i}U_{\beta i}^{}U_{\alpha j}^{}U_{\beta j}]\mathrm{sin}^2\left(\frac{\mathrm{\Delta }_{ij}}{2}\right).$$ (3) The probability, therefore, oscillates with oscillation lengths $`\mathrm{\Delta }_{ij}`$ given by $$\frac{\mathrm{\Delta }_{ij}}{2}=1.27\frac{|m_i^2m_j^2|}{\text{eV}^2}\frac{L/E}{\text{m/MeV}}=1.27\frac{\mathrm{\Delta }m_{ij}^2}{\text{eV}^2}\frac{L/E}{\text{m/MeV}},$$ (4) where $`E`$ is the neutrino energy. In general the transition probabilities will present an oscillatory behaviour with two oscillation lengths. In order to explain the solar and atmospheric neutrino data we impose the lengths to be in the range such that: $$\mathrm{\Delta }m_{12}^2\mathrm{\Delta }m_{solar}^2<10^4\text{eV}^2,\mathrm{\Delta }m_{23}^2\mathrm{\Delta }m_{atm}^210^3\text{eV}^2.$$ (5) In this way, for instance, the electron and muon neutrino survival probabilities in vacuum are given by $`P_{ee}`$ $`=`$ $`1\mathrm{cos}^4\theta _{13}\mathrm{sin}^2(2\theta _{12})\mathrm{sin}^2\left({\displaystyle \frac{\mathrm{\Delta }_{sol}}{2}}\right)\mathrm{sin}^2(2\theta _{13})\mathrm{sin}^2\left({\displaystyle \frac{\mathrm{\Delta }_{atm}}{2}}\right),`$ (6) $`P_{\mu \mu }`$ $`=`$ $`1\mathrm{cos}^2\theta _{13}\mathrm{sin}^2\theta _{23}(1\mathrm{cos}^2\theta _{13}\mathrm{sin}^2\theta _{23})\mathrm{sin}^2\left({\displaystyle \frac{\mathrm{\Delta }_{atm}}{2}}\right)`$ $`4(\mathrm{sin}\theta _{12}\mathrm{cos}\theta _{23}\mathrm{cos}\theta _{12}\mathrm{sin}\theta _{13}\mathrm{sin}\theta _{23})^2`$ $`(\mathrm{cos}\theta _{12}\mathrm{sin}\theta _{23}+\mathrm{sin}\theta _{12}\mathrm{sin}\theta _{13}\mathrm{sin}\theta _{23})^2\mathrm{sin}^2\left({\displaystyle \frac{\mathrm{\Delta }_{sol}}{2}}\right).`$ For the physically interesting case $`\mathrm{\Delta }m_{solar}^2\mathrm{\Delta }m_{atm}^2`$ we find that the solar and atmospheric neutrino oscillations decouple in the limit $`\theta _{13}=0`$. In this case the values of the mixing angles $`\theta _{12}`$ and $`\theta _{23}`$ can be obtained directly from the results of the analysis in terms of two–neutrino oscillations presented in the first section. Although for simplicity we have restricted here to vacuum oscillations, the decoupling is still valid in the presence of matter. Deviations from the two–neutrino scenario are then determined by the size of the mixing angle $`\theta _{13}`$. The first question to answer is how the presence of this new angle affects our analysis of the solar and atmospheric neutrino data. For vacuum solution to the solar neutrino problem the answer is simply given in Eq. (6). For the case of MSW solutions it has been shown (see Ref.fogli ) that $$P_{ee,MSW}^{3\nu }=\mathrm{sin}^4(\theta _{13})+\mathrm{cos}^4(\theta _{13})P_{ee,MSW}^{2\nu }$$ (8) where $`P_{ee,MSW}^{2\nu }`$ is obtained with the modified sun density $`N_e\mathrm{cos}^2(\theta _{13})N_e`$. In Fig. 5 we show the allowed regions for the oscillation parameters $`\mathrm{\Delta }m_{12}`$ and $`\mathrm{sin}^2(\theta _{12})`$ from the analysis of the solar neutrino experiments event rates in the framework of three–neutrino oscillations for different values of the angle $`\theta _{13}`$. As seen in the figure the effect is small unless very large values of $`\theta _{13}`$ are involved. In particular for $`\mathrm{sin}^2(\theta _{13})<0.2`$ ($`\theta _{13}<25^{}`$) there are still two separated SMA and LMA solutions in the ($`\mathrm{\Delta }m_{12}^2`$,$`\mathrm{sin}^2(\theta _{12})`$ ) plane (see also Fogli’s talk and reference therein). For a detail study of the effect of $`\theta _{13}`$ in the analysis of the atmospheric neutrino data we refer to Fogli’s talk in these proceedings. The conclusion is that for the presently allowed values of $`\theta _{13}`$ by the CHOOZ experiment (see below) the two–flavour analysis of the solar and atmospheric neutrino data are good approximations in the determination of the allowed mass splitings $`\mathrm{\Delta }m_{12}^2`$ and $`\mathrm{\Delta }m_{23}^2`$ and mixing angles $`\theta _{12}`$ and $`\theta _{23}`$. We must now turn to our present knowledge of the value of the mixing angle $`\theta _{13}`$. Short baseline experiments cannot provide any information on the value of this angle, as they are not sensitive to oscillations since for both mass splitings the oscillating phase is too small. On the other hand experiments at reactor and long baseline experiments can be sensitive to oscillations with $`\mathrm{\Delta }m_{23}^2`$ . In Table 3 we show the expression for the transition probabilities relevant for each of the experiments in the three–neutrino framework. In Table 3 $`s_i=\mathrm{sin}(\theta _i)`$ and $`c_i=\mathrm{cos}(\theta _i)`$ and $`S_{atm}^2=\mathrm{sin}^2\left(\frac{\mathrm{\Delta }_{23}}{2}\right)`$. Long baseline experiments at reactors such as Borexino and Kamland due to the long baseline and lower reactor neutrino energy can be sensitive to both oscillation lengths $$P_{ee}^{\text{LBL at reac}}=14s_{12}^2c_{12}^2c_{13}^4S_{sun}^24s_{13}^2c_{13}^2S_{atm}^2$$ (9) In this case if the solution to the solar neutrino deficit is the SMA the contribution from the piece in $`S_{sun}^2`$ is very small and the experiments can be sensitive to $`\theta _{13}`$ by the observation of oscillations with the shorter wavelength. For the LMA solution to the solar neutrino problem, however, both terms can contribute and in consequence the precision attainable on $`\theta _{13}`$ depends on the precise knowledge of the solar neutrino parameters $`\theta _{12}`$ and $`\delta m_{12}^2`$ which would be achieved at future solar neutrino experiments such as SNO presently running at Sudbury and Borexino at Gran Sasso. In Fig. 6 we plot the presently excluded region in the $`\theta _{13}`$ $`\mathrm{\Delta }m_{23}^2`$ plane by the present reactor experiments as well as the attainable sensitivity at the future long baseline experiments listed in Table 2. One must, however, take these value as the “ultimate” sensitivity that could be achieved at these experiments. The results presented in Fig. 6 were obtained by direct translation of the usual two–neutrino exclusion regions to the tree–neutrino scenario. But the original exclusion regions where obtained assuming only two–neutrino oscillations for the corresponding channel while in the case of three–neutrino mixings there may be new sources of backgrounds arising from other channels what can worsen the sensitivity. In order to obtain the definite sensitivity in the angle $`\theta _{13}`$ the experiments should redo their analysis in the framework of three–neutrino oscillations. ## 4 Four-Neutrino Schemes In the previous section we have discussed the neutrino mixing parameters assuming that the LSND result would be not confirmed by the MiniBooNE experiment. If the opposite holds then the simplest way to open the possibility of incorporating the LSND results to the solar and atmospheric neutrino evidences is to invoke a sterile neutrino, i.e. one whose interaction with standard model particles is much weaker than the SM weak interaction so it does not affect the invisible Z decay width, precisely measured at LEP. The sterile neutrino must also be light enough in order to participate in the oscillations involving the three active neutrinos. After imposing the present constrains from the negative searches at accelerator and reactor neutrino oscillation experiments one is left with two possible mass patterns as described in Fig. 7 which we will call scenario I and II. In scenario I there are two lighter neutrinos at the solar neutrino mass scale and two maximally mixed almost degenerate eV-mass neutrinos split by the atmospheric neutrino scale. In scenario II the two lighter neutrinos are maximally mixed and split by the atmospheric neutrino scale while the two heavier neutrinos are almost degenerate separated by the solar neutrino mass difference. In both scenarios solar neutrino data together with reactor neutrino constrains, imply that the electron neutrino must be maximally projected over one of the states belonging to the pair split by the solar neutrino scale: the lighter (heavier) pair for scenario I (II). On the other hand, atmospheric neutrino data together with the bounds from accelerator neutrino oscillation experiments imply that the muon neutrino must be maximally projected over the pair split by the atmospheric neutrino mass difference: the heavier (lighter) pair for scenario I (II). In both scenarios there are two possible assignments for the sterile and tau neutrinos which we denote by .a and .b depending on whether the tau neutrino is maximally projected over the pair responsible for the atmospheric neutrino oscillations and the sterile neutrino is responsible for the solar neutrino deficit ($`\nu _X=\nu _\tau `$ and $`\nu _X^{}=\nu _s`$) or viceversa ($`\nu _X=\nu _s`$ and $`\nu _X^{}=\nu _\tau `$). For a more detail description of these scenarios and general consequences we refer to the talk of C. Giunti in these proceedings four . In the four-neutrino scenarios the evolution equation is given as in Eq. (1) but now $`U`$ is a $`4\times 4`$ unitary matrix which contains in general 6 mixing angles and 3 or 6 CP violating phases four . For the sake of simplicity, in our discussion we will neglect the mixing angles of the sterile neutrino with the heavy states and the CP phases. In this case we are left with four mixing angles which we choose to be $$U=R_{23}(\theta _{24})\times R_{23}(\theta _{23})\times R_{13}(\theta _{34})\times R_{12}(\theta _{12})$$ (10) In general the transition probabilities will present an oscillatory behaviour with three oscillation lengths. In order to explain the solar and atmospheric neutrino data, and the LSND result we impose the osculation lengths to be in the range such that: $`\mathrm{\Delta }m_{12}^2`$ $``$ $`\mathrm{\Delta }m_{solar}^2<10^4\text{eV}^2`$ $`\mathrm{\Delta }m_{34}^2`$ $``$ $`\mathrm{\Delta }m_{atm}^210^3\text{eV}^2`$ (11) $`\mathrm{\Delta }m_{23}^2\mathrm{\Delta }m_{13}^2\mathrm{\Delta }m_{14}^2\mathrm{\Delta }m_{24}^2`$ $``$ $`\mathrm{\Delta }m_{LSND}^20.1\text{eV}^2`$ We can work out all the survival probabilities and find that the solar and atmospheric neutrino oscillations decouple in the limit $`\theta _{24}=\theta _{23}=0`$. In this case the values of the mixing angles $`\theta _{12}`$ and $`\theta _{34}`$ can be obtained directly from the results of the analysis in terms of two–neutrino oscillations presented in the first section. Deviations from the two–neutrino scenario are then determined by the size of the mixing angles $`\theta _{23}`$ and $`\theta _{24}`$. The value of these angles is presently limited by the reactor experiments. For the range of mass differences invoked by the LSND experiment the most constraining experiment is Bugey. The relevant transition probability is the $`\nu _e`$ survival probability. For any value of the atmospheric mass difference this probability always verifies $$P_{lim}P_{ee}^{Bugey}12c_{23}^2c_{24}^2(1c_{23}^2c_{24}^2)c_{23}^2c_{24}^20.992.$$ (12) what implies that $$\mathrm{sin}^2(\theta _{23})<7.5\times 10^3\mathrm{sin}^2(\theta _{24})<7.5\times 10^3$$ (13) or equivalently both angles must be smaller than $`5^{}`$. Further sensitivity on the mixing angle $`\theta _{23}`$ is expected at $`\nu _\mu `$ disappearance experiments at LBL. For instance from the MINOS measurement of $`NC/CC`$ we expect $`P_{lim}P_{\mu \mu }^{Minos}`$ $`=`$ $`1\mathrm{sin}^2(2\theta _{23})S_{LSND}^2\mathrm{sin}^2(2\theta _{34})S_{atm}^2`$ (14) $``$ $`1{\displaystyle \frac{1}{2}}\mathrm{sin}^2(2\theta _{23})\mathrm{sin}^2(\theta _{23})3.7\times 10^3.`$ To improve our knowledge of the mixing $`\theta _{24}`$ one must perform $`\nu _\tau `$ appearance experiments. In these ones the relevant survival probabilities are: $`P_{\mu \tau }`$ $``$ $`\mathrm{sin}^2(2\theta _{23})\mathrm{sin}^2(\theta _{24})S_{LSND}^2+\mathrm{sin}^2(2\theta _{23})\mathrm{cos}^2(\theta _{24})S_{atm}^2`$ (15) $`P_{e\tau }`$ $``$ $`\mathrm{sin}^2(2\theta _{24})S_{LSND}^2+\mathrm{}.`$ (16) The presence of the shorter oscillation wavelength $`S_{LSND}^2`$ suggests that in this four–neutrino scenario the best sensitivity would be achievable at a future high precision short baseline experiment. ## 5 Conclusions At present, indications of non-zero neutrino masses and mixing arise from three different sources: the solar neutrino experiments, atmospheric neutrino data and the LSND result. The analysis of these data in terms of two–neutrino oscillations yield three different oscillation scales which can only be put together in a common framework by invoking the existence of a fourth sterile neutrino. In this respect the most important upcoming experiment is the MiniBooNE experiment which searches for $`\nu _e`$ appearance in the Fermilab $`\nu _\mu `$ beam and it is specially designed to make a conclusive statement about the LSND’s neutrino oscillation evidence. If the LSND is ruled out by the MiniBooNE experiment we can fit both solar and atmospheric data in terms of three–neutrino oscillations. If, on the contrary, LSND result stands the test of time, this would be a puzzling indication for the existence of a light sterile neutrino and the need to work in a four–neutrino framework. We have seen that in both scenarios, the existing limits on neutrino mixings from the negative searches at reactors imply that the two–flavour analysis of the solar and atmospheric neutrino data are good approximations in the determination of the two allowed mass splitings and two mixing angles. In the three–neutrino scenario, due to the long oscillation lengths involved, further improvement in the additional mixing angle $`\theta _{13}`$ can only be achieved at long baseline experiments. In particular we have seen that with the presently designed experiments we can expect to reach at most a sensitivity of about $`\theta _{13}10^{}`$. In the case of four–neutrino oscillations the presence of the shorter oscillation responsible of the LSND observation suggests that the best sensitivity would be achievable at a future high precision short baseline experiment. Acknowledgments We are grateful to F. Didak and J.J. Gomez-Cadenas for their kind hospitality in Lyon. We thank E. Akhmedov for comments. This work was supported by grants DGICYT PB95-1077 and DGICYT PB97-1261, and by the EEC under the TMR contract ERBFMRX-CT96-0090.
warning/0001/nucl-th0001033.html
ar5iv
text
# Dynamics of Hot Bulk QCD Matter: from the Quark-Gluon Plasma to Hadronic Freeze-Out ## I Introduction A major goal of colliding heavy-ions at relativistic energies is to heat up a tiny region of space-time to temperatures as high as are thought to have occured during the early evolution of the universe, a few microseconds after the big bang . In ultra-relativistic heavy-ion collisions the four-volume of hot and dense matter, with temperatures above $`150`$ MeV, is on the order of $`(10`$ fm$`)^4`$. The state of strongly interacting matter at such high temperatures (or density of quanta) is usually called quark-gluon plasma (QGP) . For a discussion of the properties and potential signatures of such a super-dense state see . A particularly interesting aspect of producing such a hot and dense space-time region is that QCD, the fundamental theory of strong interactions, is expected to exhibit a transition to a new thermodynamical phase at a critical temperature $`T_C100300`$ MeV. This phase transition has been observed in numerical studies of the thermodynamics of QCD at vanishing net baryon charge on lattices . It is the only phase transition of a fundamental theory that is accessible to experiments under controlled laboratory conditions. In this paper we shall investigate the dynamics of relativistic heavy ion collisions within a novel transport approach combining a macroscopic and a microscopic model. We shall focus here on collision systems currently under investigation at the CERN Super-Proton-Synchrotron (SPS), the Relativistic Heavy Ion Collider (RHIC) at BNL and the future Large Hadron Collider (LHC) at CERN. We shall work in natural units $`\mathrm{}=c=k=1`$ throughout the paper. ## II General Aspects of Matching Fluid Dynamics to Microscopic Transport In this section we discuss general aspects and assumptions of our model for the space-time evolution of high-energy heavy-ion reactions. In particular, we introduce fluid dynamics for the early, hot stage, and the matching to microscopic transport for the later, more dilute stages of the reaction. Within this section, quantities without subscript refer to the fluid, while properties of the microscopic transport theory carry the subscript $`micro`$. ### A Transport Equation for Incoherent Quanta / Particles The most basic assumption of our model for the evolution of high-energy heavy-ion reactions is that at the initial time<sup>§</sup><sup>§</sup>§Our choice of space-time variables is described in more detail below; for the moment, we assume that suitable variables have been chosen, and that the hypersurfaces of homogeneity are time-orthogonal everywhere. $`t=t_i`$ the highly excited space-time domain produced in the impact can be viewed as being populated by incoherent quanta on the mass-shell. Thus, the system can be described by a distribution function $`f_i(x^\mu ,p^\nu )`$, where $`x^0=t_i`$, $`p^0=\sqrt{\stackrel{}{p}^2+m_i^2}`$, and $`i`$ labels different species of quanta. We will not discuss here how such a state of high entropy density could possibly be reached. That discussion is out of the scope of the present manuscript. Our work addresses the subsequent evolution of that initial state up to the so-called freeze-out of strong interactions in the system. The semi-classical evolution of the distribution function in the forward light-cone is described by means of a so-called transport equation, e.g. the Boltzmann equation $$pf_i(x^\mu ,p^\nu )=𝒞_i.$$ (1) $`𝒞_i`$ is the collision kernel, describing gain or loss of quanta (particles) of species $`i`$ in the phase-space cell around $`(x^\mu ,p^\nu )`$ due to collisions. Note that we have dropped possible classical background fields in eq. (1). ### B Moments of the Transport Equation: Hydrodynamics for the hottest stage For the problem at hand, however, the usefulness of eq. (1) is rather limited. A major difficulty is that to obtain an analytical or numerical solution, in most cases one has to introduce an expansion of the collision kernel in terms of the number of incoming particles per “elementary” collision (In most practical applications that expansion is even truncated at the level of binary collisions, $`2n`$). Obviously, the expansion is ill-defined at very high densities. A second major problem is to describe the hadronization process, i.e. the dynamical conversion of quarks and gluons into hadrons, on the microscopic level. Several interesting approaches to describe the hadronization of a plasma of quarks and gluons microscopically have been proposed in the literature, cf. e.g. and references therein. However, due to the very complicated nature of this process, many of those models have to involve some kind of ad-hoc prescriptions, which have quite significant impact on the results. A first-order QCD phase transition, as assumed in the following, is particularly difficult to model microscopically. At present we are not able to solve these problems in a fully satisfactory way. We can, however, circumvent them to some extent if we are mainly interested in the bulk dynamics of a hot QCD system. In this case we can employ relativistic ideal hydrodynamics for the very dense stage of the reaction up to hadronization. Let us thus assume that it is feasible to employ the continuum limit. The first two moments of eq. (1) yield the continuity equations for the conserved currents and for energy and momentum , $$N_i=0,\mathrm{\Theta }=0.$$ (2) In the following, we will explicitly consider only one conserved current, namely the (net) baryon current. All other currents, as e.g. strangeness, charm, electric charge etc., will be assumed to vanish identically (due to local charge neutrality and an ideal fluid) such that the corresponding continuity equations are trivially satisfied. Ideal fluid dynamics goes even further and assumes that the momentum-space distributions in the local rest frame are given by either Fermi-Dirac or Bose-Einstein distribution functions, respectively. Dissipation and heat conduction which arise from higher moments are neglected. Since we restrict fluid-dynamics to the high-temperature and high-density stage, this approximation is at least logically consistent. In future work, it will be important though to check its quantitative accuracy. The density of secondary partons in the central region of high-energy nuclear collisions is very high. According to present knowledge, it is likely that the central region evolves from a stage of pre-equilibrium towards a QGP in local thermal equilibrium , despite the large expansion rate. On the other hand, the very same calculations do not seem to support rapid chemical equilibration (in particular of the quarks), cf. also . However, in most publications interactions among the secondary partons (and in particular particle production via inelastic processes) were treated perturbatively. Since the running coupling in a thermal plasma with $`T1`$ GeV is not very small, one can not exclude sizeable contributions from processes involving higher powers of $`\alpha _s`$ . Moreover, in addition to the semi-hard partons there might exist a coherent color field in the central region (between the receding nuclear “pancakes”) which produces additional quark-antiquark pairs in its decay . In any case, we will not argue in favor nor against rapid $`q\overline{q}`$ production and chemical equilibration but simply assume that the quark densities at the initial time of the hydrodynamical expansion are close to their chemical equilibrium values. At least for Pb+Pb at CERN-SPS energy, where experimental data already exists, this is basically the only way for our model to account for the fact that measured hadron multiplicity ratios are close to their chemical equilibrium values . Since the expansion rate after hadronization is too large for “chemical cooking” (and in particular for strangeness equilibration), as will be discussed in section IV D, it would be virtually impossible to achieve approximate chemical equilibrium during the later hadronic stages if starting from a QGP far off chemical equilibrium, cf. also . The general picture as described above is summarized in the space-time diagram depicted in Fig. 1. We assume that ideal fluid dynamics is a reasonable approximation between the “initial” time $`t_i`$ and the hadronization hypersurface. After that, we will switch to a microscopic description employing the binary collision approximation for the collision kernel. In particular, we will employ the Ultrarelativistic Quantum Molecular Dynamics (UrQMD) transport model, see below. ### C Microscopic Transport from Hadronization to Freeze-Out One may ask why it is not sufficient to rely on hydrodynamics up to some rather late stage of the reaction, after which one postulates that all particle momenta are “frozen” and thus are equal to those measured in the detector at $`t_f\mathrm{}`$. That approach has been applied to nuclear collisions by many authors, for recent work cf. e.g. , and leads to reasonable results for the single-particle spectra of the most abundant hadron species $`\pi `$, $`K`$, $`p`$, $`\mathrm{\Lambda }`$. However, the following limitations arise: First, the evolution must clearly be non-ideal in the late stages of the reaction , as the system approaches “freeze-out”. This can manifest in decoupling of various components of the fluid (e.g. pions and nucleons), i.e. each component develops an individual collective velocity . Another aspect is that the $`p\mathrm{d}V`$ expansion work performed by the fluid can be partly compensated by entropy production ($`+T\mathrm{d}S`$) such that the expansion may even become isoergic, $`\mathrm{d}E=0`$, instead of isentropic, $`\mathrm{d}S=0`$ . Moreover, since each hadron is propagated individually, and its interactions with other hadrons are described on the basis of elementary processes, microscopic transport models offer the opportunity to calculate the freeze-out conditions instead of just putting them in by hand as is done in the purely fluid-dynamical approaches . There one assumes that freeze-out occurs whenever some criterion is fulfilled, e.g. when the temperature drops below some “guessed” value. In contrast, the non-truncated transport eq. (1) can describe self-consistently the freeze-out of the system: no decoupling hypersurface is imposed by hand, but rather is determined by an interplay between the (local) expansion scalar $`u`$ (where $`u`$ is the four-velocity of the local rest-frame), the relevant elementary cross sections and decay rates, and the equation of state (EoS), which actually changes dynamically as more and more hadron species decouple. This is obviously a key point for being able to study and predict the dependence of the final state on collision energy (i.e. on the initial entropy or energy density), system size etc., instead of just fitting it by an appropriate choice of a freeze-out hypersurface. Note e.g. that the nucleons emerging from the QCD hadronization phase transition in the early universe were able to maintain chemical equilibrium down to temperatures of about $`50`$ MeV . In heavy-ion collisions at CERN-SPS energies, however, one finds chemical freeze-out temperatures on the order of $`140160`$ MeV . The origin of this difference lies in the much smaller expansion rate (Hubble constant) of the early universe as compared to a high-energy heavy ion collision , and can only be explained within kinetic theory but not within pure hydrodynamics. Another complication arises from the fact that close to the freeze-out hypersurface the freeze-out process feeds back on the evolution of the fluid . This will in general deform the freeze-out hypersurface, say an isotherm of given temperature $`T_{fo}`$. It will differ from that found a posteriori from the solution of eqs. (2) in the whole forward light-cone. Furthermore, the idealization that the transition from ideal flow to free-streaming occurs on a sharp hypersurface, i.e. a three-volume in space-time, is rather crude. One instead expects a smooth transition as the temperature (and the density of particles) decreases, cf. e.g. the discussion in . This is supported by studies of the hadron kinetics close to freeze-out with realistic cross-sections , cf. also section IV C. Finally, it is likely that the freeze-out is not universal for all hadron species, simply because their transport cross-sections are very different. One can therefore hardly assume that all hadron species decouple on the same hypersurface . The clearest example for this is the transverse momentum distribution of $`\mathrm{\Omega }`$ baryons obtained by the WA97 collaboration for Pb+Pb collisions at CERN-SPS energy, $`\sqrt{s}=17A`$ GeV. Unlike is the case for pions, nucleons, anti-nucleons, and lambdas, the $`p_T`$ distribution of omegas as calculated within hydrodynamics with freeze-out on the $`T=T_{fo}=130`$ MeV hypersurface is much stiffer than the experimental finding. Indeed, more detailed kinetic treatments which explicitly account for the small transport cross-section of $`\mathrm{\Omega }`$ baryons in a meson-rich hadron gas, emerging either from fragmentation of longitudinally stretched color-strings or an incoherent hot plasma of quarks and gluons , show that these multiple-strange particles freeze out earlier and pick up less collective transverse flow than pions and nucleons, for example. ### D Transition from Fluid Dynamics to Microscopic Transport A few remarks on the transition from hydrodynamics to microscopic transport are in order here. In general, one should introduce source terms $`N_{micro}`$ and $`\mathrm{\Theta }_{micro}`$ on the right-hand-sides of eqs. (2), where $`N_{micro}^\mu (x)`$ $`=`$ $`{\displaystyle \underset{i}{}}{\displaystyle \frac{\mathrm{d}^3k}{k_i^0}k^\mu f_{i,\mathrm{micro}}(x,k)},`$ (3) $`\mathrm{\Theta }_{micro}^{\mu \nu }(x)`$ $`=`$ $`{\displaystyle \underset{i}{}}{\displaystyle \frac{\mathrm{d}^3k}{k_i^0}k^\mu k^\nu f_{i,\mathrm{micro}}(x,k)}`$ (4) denote the net baryon current and the energy-momentum tensor of the microscopic transport model, respectively. Accordingly, external sources of particles have to be introduced in the transport equation, which model the net-baryon charge and energy-momentum transfer from the fluid. This way a self-consistent solution in the whole forward light-cone, starting from the initial hypersurface $`t=t_i`$ could be obtained. However, if a space-time region bounded by a hypersurface $`\sigma _H^\mu `$ exists where fluid-dynamics is an adequate approximation, one can choose an arbitrary hypersurface $`\sigma _{switch}^\mu `$ within this region where to switch from eqs. (2) to (1). One can then simply assume $`N_{micro}0`$ and $`\mathrm{\Theta }_{micro}0`$ in the interior“Interior” meaning towards the origin of our space-time diagram, Fig. 1., and $`N0`$, $`\mathrm{\Theta }0`$ in the exterior. On that hypersurface, one sets $`N_{micro}N`$, $`\mathrm{\Theta }_{micro}\mathrm{\Theta }`$. This is because hydrodynamics is a limiting case of eq. (1), and this more general transport equation will automatically recover the fluid-dynamical solution in the space-time region between $`\sigma _{switch}^\mu `$ and $`\sigma _H^\mu `$. For the particular model discussed here, $`\sigma _{switch}^\mu `$ can not precede the hadronization hypersurface since our microscopic transport model deals with color-singlet states, only. Also, it employs the binary collision approximation of the kernel, which becomes less justified in the hot and dense stage preceding hadronization. Furthermore, as will be discussed in more detail below, in high-energy heavy-ion collisions it turns out that the boundary of validity of (ideal) fluid dynamics, $`\sigma _H^\mu `$, can not extend far into the post-hadronization stage (the less the higher the collision energy). Thus, we conclude that the hadronization hypersurface is the most natural choice for the switch from eqs. (2) to (1). The phase-space distribution of particles of species $`i`$ on $`\sigma _{switch}^\mu `$ is then given by $$E_i\frac{\mathrm{d}N_i}{\mathrm{d}^3p}=d\sigma pf(pu),$$ (5) where $`u^\mu `$ is the four-velocity of the local rest-frame. An explicit expression for the geometry suitable for high-energy collisions will be given below. For time-orthogonal hypersurfaces as depicted in Fig. 1 one has $$\mathrm{d}\sigma _\mu |_{t=\mathrm{const}.}=(\mathrm{d}^3x,\stackrel{}{0}),$$ (6) the left-hand-side of eq. (5) being simply $`E_if_{i,micro}`$. It is clear that by construction the microscopic transport starts from a state of local equilibrium on $`\sigma _{switch}^\mu `$, the hypersurface where the switch is performed. The local energy density, net baryon density, and collective expansion velocity are those obtained from the hydrodynamical solution. Thus, the conserved currents and the energy-momentum tensor of the microscopic transport theory assume the form appropriate for ideal fluids , $`N_{micro}^\mu `$ $`=`$ $`\rho u^\mu ,`$ (7) $`\mathrm{\Theta }_{micro}^{\mu \nu }`$ $`=`$ $`\left(ϵ+p_{micro}\right)u^\mu u^\nu p_{micro}g^{\mu \nu }.`$ (8) Now, in order that $`\mathrm{\Theta }_{micro}=\mathrm{\Theta }`$ on $`\sigma _{switch}^\mu `$, the pressure at given energy and baryon density must equal that of the fluid-dynamical model, i.e., the equations of state in local thermodynamical equilibrium must be the same. In general this requirement is non-trivial. For ideal gases, however, it can be obeyed by simply including the same states $`i`$ in the microscopic transport (1) as in the grand partition function which is used to calculate the equation of state employed in hydrodynamics. We shall discuss this point in more detail when presenting our specific equation of state below. We finally briefly discuss one last aspect of the switch from fluid dynamics to microscopic transport on some hypersurface $`\sigma _{switch}^\mu `$. As already mentioned above, this hypersurface is assumed to be within the region of validity of ideal hydrodynamics, and should be identified with the hadronization hypersurface. However, the latter will in general also exhibit time-like parts (points where the normal vector on $`\sigma _{switch}^\mu `$ is space-like). A schematic example is given in Fig. 2. The initial condition of the microscopic transport on $`\sigma _{switch}^\mu `$ can now not be chosen arbitrarily. This is clear from the fact that the points 1 and 2, for example, are causally connected. The simplest way to prevent violation of the evolution equations is to specify initial conditions on a purely space-like hypersurface (e.g. $`t=t_i`$) and to employ the dynamical equations, in our case the continuity equations (2), to calculate $`\mathrm{\Theta }`$ and $`N`$ on $`\sigma _{switch}^\mu `$. This way the states of the system at 1 and at 2 are consistent. However, one problem with switching on a hypersurface with time-like parts remains. As discussed above, eq. (5) conserves energy-momentum and (net) baryon charge. For this to hold, it actually does not count the flow of the currents from the inside to the outside of $`\sigma _{switch}^\mu `$ but, actually, the net flow. That is, the difference of outflowing and inflowing charge, momentum etc. The inflow is due to those parts of the thermal distribution function $`f(x,p)`$ which move into the opposite direction than the fluid. Due to the exponential tails of $`f`$ such particles clearly always exist, but their number decreases strongly if the collective flow is strong. In this case the locally isotropic momentum-space distribution is strongly boosted. Thus, the in-current is obtained under the assumption that within an infinitesimal region on both sides of the hypersurface there is hydrodynamic flow and local thermodynamical equilibrium. For this reason, $`\sigma _{switch}^\mu `$ must be entirely within the region of validity of fluid dynamics, $`\sigma _H^\mu `$. Again, in this case eq. (5) gives the net flow of all the currents from the fluid-region to the region where we apply the microscopic transport. The problem is, however, that in some part of momentum and coordinate space the left-hand-side of eq. (5) can be negative. This means that the ingoing flow exceeds the outgoing flow. These “negative contributions” were already discussed by several authors . Since we will interpret $`\mathrm{d}^3p\mathrm{d}\sigma pf(pu)/E`$ as a probability distribution we have to require positive definiteness. This can either be achieved by multiplying with a cut-off function $`\mathrm{\Theta }(\mathrm{d}\sigma p)`$, which leads to a slight violation of the conservation laws; or by integration over sufficiently large bins in momentum and coordinate-space, and random redistribution of the particles within the bins, which smears out the distribution over momentum and coordinate space. A rigorous solution of this problem requires to introduce the above-mentioned source-terms in the fluid-dynamical evolution equations as well as in the microscopic transport equation. However, for the cases studied here the negative contributions were not relevant. The main reason is that the collective flow velocity on the time-like parts of the hypersurface is close to one, such that net flow of particles from the microscopic transport to hydrodynamics does not occur. For non-relativistic flow on $`\sigma _{switch}^\mu `$, however, the negative contributions would be more serious. ## III Specific Model for High-Energy Heavy-Ion Collisions In the present paper we shall use hydrodynamics to model a first order phase transition from a QGP to a hadronic fluid, and combine it with a microscopic transport calculation for the later, purely hadronic stages of the reaction. In the following sections we describe the particular hydrodynamical and transport models employed here, cf. also and . ### A Scaling Hydrodynamics As already mentioned above, hydrodynamics for hadronic collisions is defined by (local) energy-momentum and net baryon charge conservation, $$\mathrm{\Theta }=0,N_B=0.$$ (9) $`\mathrm{\Theta }^{\mu \nu }`$ denotes the energy-momentum tensor, and $`N_B^\mu `$ the current of net baryon charge. For ideal fluids, the energy-momentum tensor and the net baryon current assume the simple form $$\mathrm{\Theta }^{\mu \nu }=(ϵ+p)u^\mu u^\nu pg^{\mu \nu },N_B^\mu =\rho _Bu^\mu ,$$ (10) where $`ϵ`$, $`p`$, $`\rho _B`$ are energy density, pressure, and net baryon density in the local rest frame of the fluid, which is defined by $`N_B^\mu =(\rho _B,\stackrel{}{0})`$. Let us, in the following, work in the metric $`g^{\mu \nu }=\mathrm{diag}(+,,,)`$. $`u^\mu =\gamma (1,\stackrel{}{v})`$ is the four-velocity of the fluid ($`\stackrel{}{v}`$ is the three-velocity and $`\gamma =(1\stackrel{}{v}^2)^{1/2}`$ the Lorentz factor). The system of partial differential equations (9) is closed by choosing an equation of state (EoS) in the form $`p=p(ϵ,\rho _B)`$, cf. below. For simplicity, we assume a cylindrically symmetric transverse expansion with a longitudinal scaling flow profile, $`v_z=z/t`$ . At $`z=0`$, equations (9) reduce to $`_tE+_T\left[\left(E+p\right)v_T\right]`$ $`=`$ $`\left({\displaystyle \frac{v_T}{r_T}}+{\displaystyle \frac{1}{t}}\right)\left(E+p\right),`$ (11) $`_tM+_T\left(Mv_T+p\right)`$ $`=`$ $`\left({\displaystyle \frac{v_T}{r_T}}+{\displaystyle \frac{1}{t}}\right)M,`$ (12) $`_tR+_T\left(Rv_T\right)`$ $`=`$ $`\left({\displaystyle \frac{v_T}{r_T}}+{\displaystyle \frac{1}{t}}\right)R,`$ (13) where we defined $`E\mathrm{\Theta }^{00}`$, $`M\mathrm{\Theta }^{0T}`$, and $`RN_B^0`$. In the above expressions, the index $`T`$ refers to the transverse component of the corresponding quantity. The set of equations (11) describes the evolution in the $`z=0`$ plane. Due to the assumption of longitudinal scaling, the solution at any other $`z0`$ can be simply obtained by a Lorentz boost. The above equations also imply $$\frac{p}{\eta }|_{\tau ,r_T}=0,$$ (14) where $`\eta \mathrm{Artanh}v_z`$ and $`\tau \sqrt{t^2z^2}`$. This means that on $`\tau =\mathrm{const}.`$ hypersurfaces pressure gradients in rapidity direction vanish, and there is no flow between adjacent infinitesimal rapidity slices. However, only for net baryon free matter, $`\rho _B0`$, does this automatically also mean that the temperature $`T`$ is independent of the longitudinal fluid rapidity $`\eta `$. In the case $`\rho _B0`$ equation (14) only demands $$s\frac{T}{\eta }|_\tau +\rho _B\frac{\mu _B}{\eta }|_\tau =0.$$ (15) $`s`$ and $`\mu _B`$ denote entropy density and baryon-chemical potential, respectively. If other charges like strangeness or electric charge are locally non-vanishing, additional terms appear. Equation (15) does not imply that the rapidity distribution of produced particles is flat (i.e. independent of rapidity) or that the rapidity distributions of various species of hadrons, e.g. pions, kaons, and nucleons, are similar. Any rapidity-dependent $`T`$ and $`\mu _B`$ that satisfy eq. (15) are in agreement with energy-momentum and net baryon number conservation, as well as with longitudinal scaling flow $`v_z=z/t`$ . Note also that non-trivial solutions of eq. (15) in general also yield $`\mu _S/\eta 0`$ on the hadronization hypersurface (i.e. a rapidity-dependent strangeness-chemical potential), even if the strangeness-density $`\rho _S=0`$ everywhere in the forward light-cone. In this paper, however, we do not explore the rapidity dependence of the particle spectra, and thus simply assume that $`T`$ and $`\mu _B`$ are independent of $`\eta `$. The fluid-dynamical evolution equations can be solved numerically on a discretized space-time grid, cf. e.g. . ### B Equation of State To close the system of coupled equations of hydrodynamics, an equation of state (EoS) has to be specified. From eq. (10) it follows that for an ideal gas the pressure $`p`$ is given by $$p=q\left(q\mathrm{\Theta }\right),$$ (16) where $`q^\mu `$ is orthogonal to $`u^\mu `$ and normalized to $`qq=1`$. In particular, in the local rest-frame $`u^\mu =(1,\stackrel{}{0})`$, we can choose $`q^\mu (0,1,0,0)+(0,0,1,0)+(0,0,0,1)`$. Then, from the definition of the energy-momentum tensor from kinetic theory, eq. (4), we obtain $$p(T,\mu _B,\mu _S)=\underset{i}{}\frac{\mathrm{d}^3k}{k_i^0}\frac{\stackrel{}{k}^2}{3}f_i(k;T,\mu _B,\mu _S).$$ (17) The sum over $`i`$ extends over the various particle species. The grand canonical potential is given by $`\mathrm{\Omega }=pV`$, where $`Vd\sigma u`$ denotes the three-volume of the given hypersurface of homogeneity. All other quantities can be obtained via standard thermodynamical relationships. E.g., the densities of entropy, net baryon charge, and energy are given by $`s(T,\mu _B,\mu _S)`$ $`=`$ $`{\displaystyle \frac{p(T,\mu _B,\mu _S)}{T}},`$ (18) $`\rho _B(T,\mu _B,\mu _S)`$ $`=`$ $`{\displaystyle \frac{p(T,\mu _B,\mu _S)}{\mu _B}},`$ (19) $`\rho _S(T,\mu _B,\mu _S)`$ $`=`$ $`{\displaystyle \frac{p(T,\mu _B,\mu _S)}{\mu _S}}\stackrel{!}{=}0,`$ (20) $`ϵ(T,\mu _B,\mu _S)`$ $`=`$ $`Tsp+\mu _B\rho _B.`$ (21) From $`p(T,\mu _B,\mu _S)`$, $`\rho _B(T,\mu _B,\mu _S)`$ and $`ϵ(T,\mu _B,\mu _S)`$ one can construct the function $`p(ϵ,\rho _B)`$ which is needed to close the system of continuity equations (9). So far, we discussed an ideal gas, only. However, lattice QCD predicts a phase transition from ordinary nuclear matter to a so-called quark-gluon plasma (QGP) at a critical temperature of $`T_C=140160`$ MeV (for $`\rho _B=0`$). We will employ a very simple and intuitive, though not very well justified description of this phase transition. We model the high-temperature phase as an ideal gas of $`u`$, $`d`$, $`s`$ quarks (with masses $`m_u=m_d=0`$, $`m_s=150`$ MeV), and gluons, employing the well-known MIT bag model EoS . In this model the non-perturbative interactions of the “deconfined bag” of quarks and gluons with the true vacuum are parameterized by a bag constant $`B`$. To make this state thermodynamically unfavorable at low temperatures, the bag contribution to the pressure must be negative. Thus, when computing the pressure of the QGP phase we subtract $`B`$ from the right-hand-side of eq. (17). Accordingly, the energy density receives a positive contribution, cf. (21), while $`s`$ and $`\rho _B`$ remain unchanged. This additional “bag term” can also be understood as an additional contribution $`+Bg^{\mu \nu }`$ due to the non-perturbative interactions to the energy-momentum tensor of the QGP-fluid. In the low-temperature region we assume an ideal hadron gas that includes the well-established (strange and non-strange) hadrons up to masses of $`2`$ GeV. They are listed in Tab. I and II. Although heavy states are rare in thermodynamical equilibrium, they have a larger entropy per particle than light states, and therefore have considerable impact on the evolution. In particular, hadronization is significantly faster as compared to the case where the hadron gas consists of light mesons only (see the discussion in ). The actual model used for the hadronic stage of the reaction (UrQMD, see section III D) additionally assumes a continuum of color-singlet states called “strings” above the $`m2`$ GeV threshold to model $`2n`$ processes and inelastic processes at high CM-energy. For example, the annihilation of an $`\overline{p}`$ on an $`\mathrm{\Omega }`$ is described as excitation of two strings with the same quantum numbers as the incoming hadrons, respectively, which are subsequently mapped on known hadronic states according to a fragmentation scheme. Since we shall be interested in the dynamics of the $`\mathrm{\Omega }`$-baryons emerging from the hadronization of the QGP, it is unavoidable to treat string-formation. The fact that string degrees of freedom are not taken into account in the EoS (17) does not represent a problem in our case because we focus on rapidly expanding systems where those degrees of freedom can not equilibrate . The phase coexistence region is constructed employing Gibbs’ conditions of phase equilibrium. The bag parameter of $`B=380`$ MeV/fm<sup>3</sup> is chosen to yield the critical temperature $`T_C160`$ MeV at $`\rho _B=0`$. By construction the EoS exhibits a first order phase transition, as is also expected in QCD for the quark-hadron phase transition in the case of three light quark flavors . The most striking aspect of a first-order phase transition with respect to the dynamical evolution is that the pressure is almost constant within the phase coexistence region (in fact, in a fluid where all conserved currents vanish identically $`p=\mathrm{const}.`$ within the mixed phase). Thus, the isentropic speed of sound, $$c_S^2=\frac{p}{ϵ}|_{s/\rho _B},$$ (22) is very small. This quantity characterizes the pressure gradient caused by a given energy density gradient along an isentrope, i.e. at constant entropy density per net baryon density (recall that all continuous solutions of the relativistic ideal-fluid dynamical equations conserve the entropy). A very small $`c_S`$ means that (isentropic) expansion is inhibited because the fluid does not “respond” to energy density gradients. In heavy-ion collisions this reflects in a particularly “soft” expansion if the mixed phase occupies the largest space-time volume of all three phases . For recent discussions of the consequences of this effect in cosmology (primordial black hole formation, evolution of density perturbations through the QCD phase transition) see e.g. . However, a nearly vanishing isentropic velocity of sound does only occur if the net baryon density is not very large, as e.g. in the cosmological QCD phase transition or in the central region of high-energy collisions studied here. In heavy-ion collisions at much lower energies, where the net baryon density in the central region is rather large, $`c_S`$ is not very small. Despite the first-order phase transition, the isentropic expansion of baryon-dense fluids is not inhibited . One should also be aware of the fact that by constructing the phase coexistence region with Gibbs’ conditions we implicitly assume a “well-mixed” phase, i.e. that the transition from the QGP to the hadronic stage proceeds in equilibrium. This is the common approach widely employed in the literature , and so far it is not in contradiction to existing data. It is based on the picture that the first-order phase transition proceeds via nucleation of hadronic bubbles in the expanding QGP , and that the bubble nucleation and growth is fast as compared to the expansion rate such that the two phases are approximately in pressure equilibrium. However, this scenario is less likely to apply to high-energy heavy-ion collisions than to the cosmological QCD phase transition, because in the former case the expansion rate is many order of magnitude larger . In particular, it has been speculated recently that the time-scale for supercooling down to the spinodal instability is comparable to that for homogeneous bubble nucleation . Thus, it may well be that the phase transition proceeds via spinodal decomposition rather than bubble nucleation. In that case, the “soft” mixed phase with $`c_s^20`$ would be absent and shorter reaction times may be expected. In any case, we postpone a detailed dynamical study of the latter scenario to a future publication, and shall restrict ourselves here to the more conservative picture assuming an adiabatic phase transition. Finally, we have to specify the initial conditions: * For collisions at SPS energy we assume that hydrodynamic flow sets in on the hyperbola $`\tau _i=1`$ fm/c. This is a value conventionally assumed in the literature, cf. e.g. . We further employ a (net) baryon rapidity density (at mid-rapidity) of $`dN_B/dy=80`$, as obtained by the NA49-collaboration for central Pb+Pb reactions . The average specific entropy in these collisions is $`\overline{s}/\overline{\rho }_B=45\pm 5`$ (the bar indicates averaging over the transverse plane). That entropy per net baryon fits most measured hadron multiplicity ratios within $`\pm 20\%`$ . The corresponding initial energy and net baryon densities ($`\overline{ϵ}_i=6.1`$ GeV/fm<sup>3</sup>, $`\overline{\rho }_i=4.5\rho _0`$) are assumed to be distributed in the transverse plane according to a so-called “wounded nucleon” distribution with transverse radius $`R_T=6`$ fm, i.e. $`ϵ(\tau _i,r_T),\rho _B(\tau _i,r_T)f(r_T)`$, with $`f(r_T)=\frac{3}{2}\sqrt{1r_T^2/R_T^2}`$. The initial temperature and quark-chemical potentials are (they are of course not exactly constant over the transverse plane) $`T_i220`$ MeV, $`\mu _q150`$ MeV, $`\mu _s=0`$. The transverse velocity field on the $`\tau =\tau _i`$ hyperbola is assumed to vanish. * Due to the higher parton density at mid-rapidity as compared to collisions at SPS energy, thermalization may be reached earlier at RHIC. According to various studies , thermalization might occur within $`0.31`$ fm. We assume $`\tau _i=R_T/10=0.6`$ fm. The net baryon rapidity density and specific entropy at mid-rapidity in central Au+Au at $`\sqrt{s}=200A`$ GeV is predicted by various models of the initial evolution, e.g. the parton cascade model PCM, RQMD 1.07, FRITIOF 7, and HIJING/B, to be in the range $`\mathrm{d}N_B/\mathrm{d}y2035`$, $`s/\rho _B150250`$ . We will employ $`\mathrm{d}N_B/\mathrm{d}y=25`$ and $`\overline{s}/\overline{\rho }_B=205`$ ($`\overline{ϵ}_i=20`$ GeV/fm<sup>3</sup>, $`\overline{\rho }_i=2.3\rho _0`$). These parameters could of course be fine-tuned once the first experimental data are available. As in the above case, $`ϵ(\tau _i,r_T)`$ and $`\rho _B(\tau _i,r_T)`$ are initially distributed in the transverse plane according to a wounded-nucleon distribution with $`R_T=6`$ fm. The initial temperature and quark-chemical potentials follow as $`T_i300`$ MeV, $`\mu _q45`$ MeV, $`\mu _s=0`$, respectively. This corresponds to a transverse energy on the $`\tau =\tau _i`$ hyperbola of $`\mathrm{d}E_T/\mathrm{d}y1.3`$ TeV, which decreases to $`\mathrm{d}E_T/\mathrm{d}y720`$ GeV on the hadronization hypersurface . * The initial conditions for CERN-LHC energy are, of course, less well known. Qualitatively, and according to present expectations, it appears reasonable to assume that 1. the density of minijets produced at time $`\tau _01/p_0`$, where $`p_02`$ GeV is the minijet cut-off scale, is much larger than at BNL-RHIC energy. The most recent estimates of the energy densities in the central region span the range $`ϵ_0=(0.31.3)`$ TeV/fm<sup>3</sup> . The results to be expected from $`p+p`$, $`p+A`$, and $`A+A`$ at BNL-RHIC will probably not reduce the uncertainties by much because the energy density at $`y0`$ and $`\sqrt{s}=5.5A`$ TeV depends strongly on the model for the nuclear parton distribution functions at very small $`x`$, out of range for RHIC. 2. the higher initial density of partons could also lead to somewhat faster equilibration than at the lower energies. Note that the produced gluons already have the “right” thermal energy per particle, $`ϵ_0/\rho _02.7T_0`$ . The distribution in momentum-space, however, has to become isotropic via rescattering among the partons . 3. the net baryon charge in a rapidity-slice $`\mathrm{\Delta }y=1`$ around $`y=0`$ is even smaller than at RHIC, and can in practice be neglected if one focuses on the bulk dynamics of the central region (in the same way as we neglect net strangeness, charm, etc.). Note however, that smaller rapidity bins may exhibit quite large fluctuations of the initial net baryon charge . Thus, in view of these uncertainties, it is clear that precise quantitative predictions for the CERN-LHC energy are hardly possible at the moment. Our more modest aim will therefore be to discuss a set of even more “extreme” initial conditions than those employed for CERN-SPS and BNL-RHIC energies, to give an idea how the dynamical evolution may continue at even higher energies. Whether or not that set of initial conditions corresponds closely to the LHC case can not be decided presently on solid grounds. Thus, we employ a thermalization time $`\tau _i=0.3`$ fm, an initial energy density (not on the $`\tau _0`$ but on the $`\tau _i`$ hypersurface !) $`\overline{ϵ}_i=230`$ GeV/fm<sup>3</sup>, and a vanishing net baryon charge, $`\mathrm{d}N_B/\mathrm{d}y=\rho _i=0`$. Again, the initial energy density is distributed in the transverse plane according to a wounded nucleon distribution with $`R_T=6`$ fm. The initial temperature is about $`T_i580`$ MeV (it is not exactly constant over the transverse plane), the initial transverse energy is $`\mathrm{d}E_T/\mathrm{d}y7.8`$ TeV. ### C Hadronization and the transition to microscopic dynamics Having specified the initial conditions on the $`\tau =\tau _i`$ hypersurface and the EoS, the hydrodynamical solution in the forward light-cone is determined uniquely. As already mentioned in section II D, we assume that it is not a bad approximation to determine the hadronization hypersurface a posteriori from the solution in the whole forward light-cone. In other words, the hadronization hypersurface is assumed to be within the region of validity of hydrodynamics. In parametric representation, the hypersurface $`\sigma ^\mu `$ is a function of three parameters . In our case, due to the symmetry under rotations around and Lorentz-boosts along the beam axis, two of these parameters can simply be identified with $`\eta `$ and $`\varphi `$, while $`\tau `$ and $`r_T`$ depend only on the third parameter, call it $`\zeta `$. Thus, $`\zeta [0,1]`$ parameterizes the hypersurface in the planes of fixed $`\eta `$ and $`\varphi `$ (in the mathematically positive orientation, i.e. counter clock-wise). The normal is $`\mathrm{d}\sigma _\mu `$ $`=`$ $`ϵ_{\mu \alpha \beta \gamma }{\displaystyle \frac{\sigma ^\alpha }{\zeta }}{\displaystyle \frac{\sigma ^\beta }{\eta }}{\displaystyle \frac{\sigma ^\gamma }{\varphi }}\mathrm{d}\zeta \mathrm{d}\eta \mathrm{d}\varphi `$ (23) $`=`$ $`({\displaystyle \frac{\mathrm{d}r_T}{\mathrm{d}\zeta }}\mathrm{cosh}\eta ,{\displaystyle \frac{\mathrm{d}\tau }{\mathrm{d}\zeta }}\mathrm{cos}\varphi ,{\displaystyle \frac{\mathrm{d}\tau }{\mathrm{d}\zeta }}\mathrm{sin}\varphi ,{\displaystyle \frac{\mathrm{d}r_T}{\mathrm{d}\zeta }}\mathrm{sinh}\eta )r_T\tau \mathrm{d}\zeta \mathrm{d}\eta \mathrm{d}\varphi .`$ (24) This expression naturally looks simpler in the $`(\tau ,\eta ,r_T,\varphi )`$-basis, cf. e.g. , but we will nevertheless write all vectors and tensors in the $`(t,x,y,z)`$-basis throughout the manuscript, even if the components are written in terms of the variables $`\tau `$, $`\eta `$, $`r_T`$, and $`\varphi `$. We can now apply eq. (5) to compute the number of hadrons of species $`i`$ hadronizing at space-time rapidity $`\eta `$, proper time $`\tau `$, and position $`r_T(\mathrm{cos}\left(\chi \varphi \right),\mathrm{sin}\left(\chi \varphi \right))`$, with four-momentum $`p^\mu =(m_T\mathrm{cosh}y,p_T\mathrm{cos}\chi ,p_T\mathrm{sin}\chi ,m_T\mathrm{sinh}y)`$, $$\frac{\mathrm{d}^6N_i}{d^2p_T\mathrm{d}y\mathrm{d}\eta \mathrm{d}\zeta \mathrm{d}\varphi }=r_T\tau \left(p_T\mathrm{cos}\left(\chi \varphi \right)\frac{d\tau }{d\zeta }m_T\mathrm{cosh}(y\eta )\frac{dr_T}{d\zeta }\right)f_i\left(pu\right).$$ (25) $`u^\mu =\gamma _T(\mathrm{cosh}\eta ,v_T\mathrm{cos}(\chi \varphi ),v_T\mathrm{sin}(\chi \varphi ),\mathrm{sinh}\eta )`$ denotes the fluid four-velocity. Thus, the direction of the particle momentum in the transverse plane is determined by the angle $`\chi `$, while the relative angle between $`\stackrel{}{p}_T`$ and the transverse flow velocity, $`\stackrel{}{v}_T`$, is denoted by $`\varphi `$. $`f`$ is either a Bose-Einstein or Fermi-Dirac distribution function, depending on the particle species under consideration. How is the distribution (25) actually passed to the microscopic model ? First, it is integrated over space-time ($`\eta `$, $`\zeta `$, $`\varphi `$) and momentum space ($`\stackrel{}{p}_T`$, $`y`$), rounded to an integer value (the hadronic transport model described in the next section deals with integer number of particles, only), and the distribution (25) divided by $`N_i`$ is used as probability distribution to randomly generate space-time and momentum-space coordinates for $`N_i`$ hadrons of species $`i`$. Of course, due to the fact that our system has a surface and does not extend to infinity in the transverse plane, hadronization does not occur on a $`\tau =\mathrm{const}.`$ hypersurface, cf. Fig. 3. Thus, if we look at our expanding system on $`\tau =\mathrm{const}.`$ surfaces, there exists an interval where the two models, hydrodynamics and the microscopic transport, are applied in parallel. ### D Microscopic dynamics: the UrQMD approach The ensemble of hadrons generated accordingly is then used as initial condition for the microscopic transport model Ultra-relativistic Quantum Molecular Dynamics (UrQMD) . The UrQMD approach is closely related to hadronic cascade , Vlasov–Uehling–Uhlenbeck and (R)QMD transport models . We shall describe here only the part of the model that is important for the application at hand, namely the evolution of an expanding hadron gas in local equilibrium at a temperature of about $`T_C160`$ MeV. The treatment of high-energy hadron-hadron scatterings, as it occurs in the initial stage of ultrarelativistic collisions, is not discussed here. A complete description of the model and detailed comparisons to experimental data can be found in . The basic degrees of freedom are hadrons modeled as Gaussian wave-packets, and strings, which are used to model the fragmentation of high-mass hadronic states via the Lund scheme . The system evolves as a sequence of binary collisions or $`2N`$-body decays of mesons, baryons, and strings. The real part of the nucleon optical potential, i.e. a mean-field, can in principle be included in UrQMD for the dynamics of baryons (using a Skyrme-type interaction with a hard equation of state). However, currently no mean field for mesons (the most abundant hadrons in our investigation) are implemented. Therefore, we have not accounted for mean-fields in the equation of motion of the hadrons. To remain consistent, mean fields were also not taken into account in the EoS on the fluid-dynamical side. Otherwise, pressure equality (at given energy and baryon density) would be destroyed. We do not expect large modifications of the results presented here due to the effects of mean fields, since the “fluid” is not very dense after hadronization and current experiments at SIS and AGS only point to strong medium-dependent properties of mesons (kaons in particular) for relatively low incident beam energies ($`E_{lab}4`$ GeV/nucleon) . Nevertheless, mean fields will have to be included in the future; a fully covariant treatment of baryon and meson dynamics within UrQMD derived from a chiral Lagrangian is currently under development. Binary collisions are performed in a point-particle sense: Two particles collide if their minimum distance $`d`$, i.e. the minimum relative distance of the centroids of the Gaussians during their motion, in their CM frame fulfills the requirement: $$dd_0=\sqrt{\frac{\sigma _{\mathrm{tot}}}{\pi }},\sigma _{\mathrm{tot}}=\sigma (\sqrt{s},\text{ type}).$$ (26) The cross section is assumed to be the free cross section of the regarded collision type ($`NN`$, $`N\mathrm{\Delta }`$, $`\pi N`$ …). The UrQMD collision term contains 53 different baryon species (including nucleon, delta and hyperon resonances with masses up to 2 GeV) and 24 different meson species (including strange meson resonances), which are supplemented by their corresponding anti-particle and all isospin-projected states. The baryons and baryon-resonances which can be populated in UrQMD are listed in table I, the respective mesons in table II – full baryon/antibaryon symmetry is included (not shown in the table), both, with respect to the included hadronic states, as well as with respect to the reaction cross sections. All hadronic states can be produced in string decays, s-channel collisions or resonance decays. Tabulated and parameterized experimental cross sections are used when available. Resonance absorption, decays and scattering are handled via the principle of detailed balance. If no experimental information is available, the cross section is either calculated via an One-Boson-Exchange (OBE) model or via a modified additive quark model which takes basic phase space properties into account. In the baryon-baryon sector, the total and elastic proton-proton and proton-neutron cross sections are well known . Since their functional dependence on $`\sqrt{s}`$ shows a complicated shape at low energies, UrQMD uses a table-lookup for those cross sections. However, many cross sections involving strange baryons and/or resonances are not well known or even experimentally accessible – for these cross sections the additive quark model is widely used. As we shall see later, the most important reaction channels in our investigation are meson-meson and meson-baryon elastic scattering and resonance formation. For example, the total meson-baryon cross section for non-strange particles is given by $`\sigma _{tot}^{MB}(\sqrt{s})`$ $`=`$ $`{\displaystyle \underset{R=\mathrm{\Delta },N^{}}{}}j_B,m_B,j_M,m_MJ_R,M_R{\displaystyle \frac{2S_R+1}{(2S_B+1)(2S_M+1)}}`$ (28) $`\times {\displaystyle \frac{\pi }{p_{CMS}^2}}{\displaystyle \frac{\mathrm{\Gamma }_{RMB}\mathrm{\Gamma }_{tot}}{(M_R\sqrt{s})^2+\frac{\mathrm{\Gamma }_{tot}^2}{4}}}`$ with the total and partial $`\sqrt{s}`$-dependent decay widths $`\mathrm{\Gamma }_{tot}`$ and $`\mathrm{\Gamma }_{RMB}`$. The full decay width $`\mathrm{\Gamma }_{tot}(M)`$ of a resonance is defined as the sum of all partial decay widths and depends on the mass of the excited resonance: $$\mathrm{\Gamma }_{tot}(M)=\underset{br=\{i,j\}}{\overset{N_{br}}{}}\mathrm{\Gamma }_{i,j}(M).$$ (29) The partial decay widths $`\mathrm{\Gamma }_{i,j}(M)`$ for the decay into the final state with particles $`i`$ and $`j`$ is given by $$\mathrm{\Gamma }_{i,j}(M)=\mathrm{\Gamma }_R^{i,j}\frac{M_R}{M}\left(\frac{p_{i,j}(M)}{p_{i,j}(M_R)}\right)^{2l+1}\frac{1.2}{1+0.2\left(\frac{p_{i,j}(M)}{p_{i,j}(M_R)}\right)^{2l}},$$ (30) here $`M_R`$ denotes the pole mass of the resonance, $`\mathrm{\Gamma }_R^{i,j}`$ its partial decay width into the channel $`i`$ and $`j`$ at the pole and $`l`$ the decay angular momentum of the final state. All pole masses and partial decay widths at the pole are taken from the Review of Particle Properties . $`\mathrm{\Gamma }_{i,j}(M)`$ is constructed in such a way that $`\mathrm{\Gamma }_{i,j}(M_R)=\mathrm{\Gamma }_R^{i,j}`$ is fulfilled at the pole. In many cases only crude estimates for $`\mathrm{\Gamma }_R^{i,j}`$ are given in – the partial decay widths must then be fixed by studying exclusive particle production in elementary proton-proton and pion-proton reactions. Therefore, e.g., the total pion-nucleon cross section depends on the pole masses, widths and branching ratios of all $`N^{}`$ and $`\mathrm{\Delta }^{}`$ resonances listed in table I. Resonant meson-meson scattering (e.g. $`\pi +\pi \rho `$ or $`\pi +KK^{}`$) is treated in the same formalism. In order to correctly treat equilibrated matter (we repeat that the hadronic matter with which UrQMD is being initialized in our approach is in local chemical and thermal equilibrium), the principle of detailed balance is of great importance. Detailed balance is based on time-reversal invariance of the matrix element of the reaction. It is most commonly found in textbooks in the form: $$\sigma _{fi}=\frac{\stackrel{}{p}_i^2}{\stackrel{}{p}_f^2}\frac{g_i}{g_f}\sigma _{if},$$ (31) with $`g`$ denoting the spin-isospin degeneracy factors. UrQMD applies the general principle of detailed balance to the following two process classes: 1. Resonant meson-meson and meson-baryon interactions: Each resonance created via a meson-baryon or a meson-meson annihilation may again decay into the two hadron species which originally formed it. This symmetry is only violated in the case of three- or four-body decays and string fragmentations, since N-body collisions with (N$`>2`$) are not implemented in UrQMD. 2. Resonance-nucleon or resonance-resonance interactions: the excitation of baryon-resonances in UrQMD is handled via parameterized cross sections which have been fitted to data. The reverse reactions usually have not been measured - here the principle of detailed balance is applied. Inelastic baryon-resonance de-excitation is the only method in UrQMD to absorb mesons (which are bound in the resonance). Therefore the application of the detailed balance principle is of crucial importance for heavy nucleus-nucleus collisions. Equation (31), however, is only valid in the case of stable particles with well-defined masses. Since in UrQMD detailed balance is applied to reactions involving resonances with finite lifetimes and broad mass distributions, equation (31) has to be modified accordingly. For the case of one incoming resonance the respective modified detailed balance relation has been derived in . Here, we generalize this expression for up to two resonances in both, the incoming and the outgoing channels. The differential cross section for the reaction $`(1,\mathrm{\hspace{0.17em}2})(3,\mathrm{\hspace{0.17em}4})`$ is given by: $$\mathrm{d}\sigma _{12}^{34}=\frac{||^2}{64\pi ^2s}\frac{p_{34}}{p_{12}}\mathrm{d}\mathrm{\Omega }\underset{i=3}{\overset{4}{}}\delta (p_i^2M_i^2)\mathrm{d}p_i^2,$$ (32) here the $`p_i`$ in the $`\delta `$-function denote four-momenta. The $`\delta `$-function ensures that the particles are on mass-shell, i.e. their masses are well-defined. If the particle, however, has a broad mass distribution, then the $`\delta `$-function must be substituted by the respective mass distribution (including an integration over the mass): $$\mathrm{d}\sigma _{12}^{34}=\frac{||^2}{64\pi ^2s}\frac{1}{p_{12}}\mathrm{d}\mathrm{\Omega }\underset{i=3}{\overset{4}{}}p_{34}\frac{\mathrm{\Gamma }}{\left(mM_i\right)^2+\mathrm{\Gamma }^2/4}\frac{\mathrm{d}m}{2\pi }.$$ (33) Incorporating these modifications into equation (31) and neglecting a possible mass-dependence of the matrix element we obtain: $$\frac{\mathrm{d}\sigma _{34}^{12}}{\mathrm{d}\mathrm{\Omega }}=\frac{p_{12}^2}{p_{34}^2}\frac{(2S_1+1)(2S_1+1)}{(2S_3+1)(2S_4+1)}\underset{J=J_{}}{\overset{J_+}{}}j_1m_1j_2m_2JM\frac{\mathrm{d}\sigma _{12}^{34}}{\mathrm{d}\mathrm{\Omega }}.$$ (34) Here, $`S_i`$ indicates the spin of particle $`i`$ and the summation of the Clebsch-Gordan-coefficients is over the isospin of the outgoing channel only. For the incoming channel, isospin is treated explicitly. The summation limits are given by: $`J_{}`$ $`=`$ $`\mathrm{max}(|j_1j_2|,|j_3j_4|)`$ (35) $`J_+`$ $`=`$ $`\mathrm{min}(j_1+j_2,j_3+j_4).`$ (36) The integration over the mass distributions of the resonances in equation (34) has been denoted by the brackets $``$, e.g. $$p_{3,4}^2p_{3,4}^2=p_{CMS}^2(\sqrt{s},m_3,m_4)A_3(m_3)A_4(m_4)dm_3dm_4,$$ with the mass distribution $`A_r(m)`$ given by a free Breit-Wigner distribution with a mass-dependent width according to equation (29): $$A_r(m)=\frac{1}{N}\frac{\mathrm{\Gamma }(m)}{(m_rm)^2+\mathrm{\Gamma }(m)^2/4}\text{with}\underset{\mathrm{\Gamma }0}{lim}A_r(m)=\delta (m_rm),$$ (37) with the normalization constant $$N=\underset{\mathrm{}}{\overset{\mathrm{}}{}}\frac{\mathrm{\Gamma }(m)}{(m_rm)^2+\mathrm{\Gamma }(m)^2/4}dm.$$ (38) Alternatively one can also choose a Breit-Wigner distribution with a fixed width, the normalization constant then has the value $`N=2\pi `$. The most frequent applications of equation (34) in UrQMD are the processes $`\mathrm{\Delta }_{1232}NNN`$ and $`\mathrm{\Delta }_{1232}\mathrm{\Delta }_{1232}NN`$. ## IV Results for heavy-ion collisions at CERN-SPS, BNL-RHIC and CERN-LHC We now present some representative results for central collisions of heavy ions at CERN-SPS, BNL-RHIC and CERN-LHC energies. We will focus on single-inclusive momentum-space distributions and the space-time picture of freeze-out following the hadronization phase transition. As we shall see, one already gains much insight into the dynamics of high-energy heavy-ion collisions from these observables. Many other aspects are thinkable and interesting but have to be postponed to future studies. ### A Hydrodynamical expansion and hadronization We first briefly discuss the evolution and hadronization of the QGP-cylinder present on the $`\tau =\tau _i`$ hypersurface as obtained from the hydrodynamical solution. Similar arguments and results can be found in a variety of papers, see for example . In particular, ref. employed the very same model as here (i.e. longitudinal scaling flow with cylindrically symmetric transverse expansion, the initial conditions and the EoS). However, the evolution at CERN-LHC energy had not been covered, and the hadronization hypersurface was only shown for a step-function like initial transverse energy density distribution, but not for the wounded-nucleon distribution employed here. Therefore, a short discussion of the prehadronic stage may be in order here. Fig. 3 summarizes the space-time picture in the plane $`\eta =\varphi =0`$. We show projections of various hypersurfaces on the $`(\tau ,r_T)`$-plane because their shape in the $`\varphi `$\- and $`\eta `$-directions is trivial: they are simply horizontal lines in the $`(\tau ,\varphi )`$\- $`(\tau ,\eta )`$-planes, extending from $`\pi `$ to $`\pi `$ and $`\mathrm{}`$ to $`\mathrm{}`$, respectively. Thus, no derivatives like $`\tau /\varphi `$ etc. appear in $`\mathrm{d}\sigma _\mu `$, eq. (23). Basically, we start at $`\tau =\tau _i`$ with a pure QGP extending from $`r_T=0`$ up to $`r_T=R_T6`$ fm in the transverse direction. The thickness of the non-QGP region at the surface is very small for the wounded nucleon distribution. Initially, the hot quark-gluon fluid is cooled mainly due to the longitudinal expansion, except close to the surface, where transverse pressure gradients are also large and lead to expansion rates several times larger than the simple $`1/\tau `$ law . The fluid eventually reaches the boundary to the mixed phase, denoted by $`\lambda =1`$. ($`\lambda `$ is the local fraction of quarks and gluons within the mixed phase.) Clearly, the space-time volume of pure QGP increases substantially from SPS to RHIC and then again towards LHC. This leads to stronger transverse flow of matter entering the mixed phase at RHIC and LHC than at SPS. Due to this effect the hadronization hypersurface ($`\lambda =0`$) extends to larger $`r_T`$. At SPS, the hadronization hypersurface $`\lambda =0`$, where the switch to the microscopic model is performed, is almost stationary for some time $`\tau R_T`$, after which the entire fluid hadronizes rapidly. UrQMD is being fed with hadrons from the stationary surface of a “burning log” of mixed phase matter . This is not to be misunderstood as an evaporation process, though. The fluid is moving with substantial velocity through the hadronization hypersurface, in particular near the point where the $`\lambda =0`$ and the $`\tau =\tau _i`$ hypersurfaces meet (there, the very dilute fluid comes close to the light-cone). Thus, the momenta of the emitted hadrons, which are purely thermal in the local rest frame, are boosted in transverse direction. At RHIC, and of course even more so at LHC, that pre-acceleration by the QGP “explosion” is so strong that the hadronization hypersurface is initially even driven outwards, before the mixed-phase cylinder finally collapses (when it can not balance the vacuum pressure $`B`$ any more) and emits hadrons from all over the transverse area. Thus, it is clear from Fig. 3 that the dynamics at SPS is characterized by the large space-time volume occupied by the mixed phase, while the stiffer and more “explosive” QGP gains importance at higher energies. ### B Post-hadronization kinetics: evolution of $`p_T`$ The choice of the hypersurface at which to perform the transition from the macroscopic hydrodynamical calculation to the microscopic transport model may affect the reaction dynamics and the results of our calculation. However, concerning the variation of the hypersurface for that transition one has to note that the hadronic part of the EoS used in the hydrodynamic solution contains the same states as UrQMD, and the energy-momentum tensors on both sides of the hadronization hypersurface match. If the assumption of local equilibrium is indeed fulfilled, UrQMD will simply continue the hydrodynamic flow since it reduces to hydrodynamics in the equilibrium-limit. However, as we shall see later, for some hadron species with small interaction cross sections deviations from ideal hydrodynamic flow can be observed immediately after complete hadronization (see also refs. ). It is found that the expansion of the hadronic fluid is dissipative rather than ideal; due to the fast local expansion generated by the QGP before hadronization the ideal flow is disturbed. Therefore it does not make much sense to choose a later hypersurface for the matching because one would precisely assume that ideal flow persists even after hadronization. Nevertheless, it is interesting to study how the choice of a later hypersurface for the transition from the macroscopical to the microscopical part of the calculation affects the results. This reveals “how wrong” the assumption of an ideal evolution of the state at hadronization is. Figure 4 shows the final mean transverse momentum $`p_T`$ for various hadron species as a function of the temperature on the hydro$``$micro transition isotherm, $`T_{sw}`$. The grey lines in the upper frame denote the $`p_T`$ of the hadrons at hadronization, i.e. at $`T_C=160`$ MeV. As shall be discussed in greater detail in section IV E, the change in $`p_t`$ in the hadronic phase (for our “default” choice $`T_{sw}=T_C=160`$ MeV) depends strongly on the individual hadron species. Protons and hyperons gain most, the $`\mathrm{\Omega }^{}`$ does not acquire any additional $`p_T`$ at all, and pions even loose some $`p_T`$ due to rescattering and additional soft pion production. The results change only marginally when decreasing $`T_{sw}`$ to 150 MeV. Simply speaking, UrQMD reproduces the fluid-dynamical solution down to about $`T150`$ MeV, for central Au+Au at RHIC energy. At this stage, fluid-dynamics predicts that the transverse rarefaction in the hadron fluid reaches the center. Consequently, the expansion becomes rather spherical and transverse flow increases strongly in this “hadronic explosion”. The lower frame in Fig. 4 shows the kink in $`p_T(T_{sw})`$ of heavy hadrons at $`T_{sw}150`$ MeV predicted by ideal hydrodynamics. Remarkably, however, the system apparently is already in a state of too rapid expansion for this “hadronic explosion” to happen. Given the state at hadronization, UrQMD (applying realistic cross-sections) predicts that the hadronic fluid basically freezes out right at the point where the hadronic rarefaction is about to make the expansion more spherical and to increase the expansion rate, see e.g. Fig. 2 in . Any later transition from hydrodynamics to the microscopic transport model leads to a strong increase of $`p_T`$ at freeze-out, which depends only on the mass of the hadron, but not on its flavor (resp. its quark content). The lines in the lower frame of figure 4 show the $`p_T`$ of the respective hadron species at the transition hypersurface (i.e. at $`T_{sw}`$). By comparing the $`p_T`$ value indicated by the line to that given by the plot symbol for each $`T_{sw}`$ one can determine the amount of $`p_T`$ gained or lost during the microscopic evolution of the reaction. Again, protons acquire the most $`p_T`$ during the microscopic evolution (even though the amount of $`p_T`$ gained decreases the lower $`T_{sw}`$ is and the closer the system comes to freeze-out), whereas $`\mathrm{\Xi }`$’s and $`\mathrm{\Omega }`$’s do not experience any $`p_T`$ increase at all. It is obvious from this analysis that the conditions of applicability for hydrodynamics in the hadronic phase deteriorate rapidly. A general freeze-out criterion can not be given since the freeze-out depends on the system size and the centrality, the energy etc. However, our transport calculation with realistic cross-sections in the hadron gas, starting in the wake of a hadronizing QGP, shows that the expansion is too rapid to allow cooling of the strong interactions much below $`T_C`$. In particular, adiabatic expansion breaks down once the expansion of the hadron fluid effectively becomes (3+1)-dimensional. ### C Space-time distributions of hadronic freeze-out Let us now turn to the freeze-out “hypersurfaces” of pions and nucleons in central (impact parameter $`b=0`$ fm) collisions of gold or lead nuclei at SPS ($`\sqrt{s}=17`$ GeV per incident colliding nucleon-pair), RHIC ($`\sqrt{s}=200`$ GeV per incident colliding nucleon-pair) and LHC ($`\sqrt{s}=5500`$ GeV per incident colliding nucleon-pair). We start with the nucleons, the most abundant baryon species in the system, restricting ourselves to the central rapidity region. Figure 5 shows the freeze-outFreeze-out meaning the space-time point of last interaction, irrespective of how “soft” that last interaction might be. We remind you also that mean-fields are not taken into account. They could even prolong the freeze-out due to very soft interactions of the hadrons with the mean field. time and transverse radius distributions $`1/r_T`$ d$`{}_{}{}^{3}N/`$d$`r_T`$d$`\tau _{fr}`$d$`y`$ for LHC (top), RHIC (middle) and SPS (bottom). The right column shows the result of the pure hydrodynamical calculation up to complete hadronization, with subsequent hadronic resonance decays, but without hadronic reinteraction. The left column shows the same calculation including full microscopic hadronic dynamics. The freeze-out characteristics of the nucleons are significantly modified due to the hadronic interaction phase. The average transverse freeze-out radius doubles at SPS and RHIC and increases by a factor of 2.5 at LHC (see also table III). The respective average freeze-out times increase by similar factors (see table IV). E.g., at RHIC the average freeze-out time for protons changes from 11.3 to 25.8 fm/c due to hadronic rescattering. As the meson multiplicity in the system at RHIC is fifty times larger than the baryon multiplicity, baryons propagate through a relativistic meson gas, acting as probes of this highly excited meson medium. Thus, we use the proton and hyperon freeze-out values listed in table IV for a first rough estimate of the duration of the hadronic phase via $`\mathrm{\Delta }\tau _{had}=\tau _{fr}^{\mathrm{Hydro}+\mathrm{UrQMD}}\tau _{fr}^{\mathrm{Hydro}+\mathrm{had}.\mathrm{decays}}`$. At the SPS $`\mathrm{\Delta }\tau _{had}`$ is found to be $`13.5`$ fm/c, very similar to the value at RHIC ($`15`$ fm/c) and at the LHC we obtain $`\mathrm{\Delta }\tau _{had}23`$ fm/c. The transverse spatial extent of the hadronic phase can be estimated in a similar way, using table III and defining the thickness $`\mathrm{\Delta }r_{had}`$ of the hadronic phase as: $`\mathrm{\Delta }r_{had}=r_{t,fr}^{\mathrm{Hydro}+\mathrm{UrQMD}}r_{t,fr}^{\mathrm{Hydro}+\mathrm{had}.\mathrm{decays}}`$. Here we find values of $`4.4`$ fm at the SPS, $`5.8`$ fm at RHIC and $`13.3`$ fm at the LHC. The Hydro+UrQMD model predicts a space-time freeze-out picture which is very different from that usually employed in the hydrodynamical model, e.g. in refs. : Here , freeze-out is found to occur in a four-dimensional region within the forward light-cone rather than on a three-dimensional “hypersurface” . Similar results have also been obtained within other microscopic transport models when the initial state was not a quark-gluon plasma. This finding seems to be a generic feature of such models: the elementary binary hadron-hadron interactions smear out the sharp signals to be expected from simple hydro. This predicted additional fourth dimension of the freeze-out domain could affect the HBT parameters considerably. This does not mean that the momentum-distributions alone can not be calculated assuming freeze-out on some effective three-dimensional hypersurface. For example, if interactions on the outer side of that hypersurface are very “soft”, the single-particle momentum distributions at not too small $`p_T`$ will not change anymore. The two-particle correlator does change, however, since it probes rather small relative momenta. Thus, the freeze-out condition, e.g. the temperature, as measured by single-particle spectra and two-particle correlations needs not be the same. The shapes of the freeze-out hypersurfaces (FOHS) show broad radial maxima for intermediate freeze-out times. Thus, transverse expansion has not developed scaling-flow (in that case the FOHS would be hyperbolas in the $`\tau r_T`$ plane). This agrees with the discussion of the evolution of the $`p_T`$ after hadronization in section IV B, which already indicated the transition to free-streaming once the transverse expansion rate becomes comparable to the longitudinal expansion rate. Furthermore, the hypersurfaces of pions and nucleons, and their shapes, are distinct from each other (as also found in at the lower BNL-AGS and CERN-SPS energies). Thus, the ansatz of a unique freeze-out hypersurface for all hadrons appears to be a very rough approximation, cf. also refs. . Figure 6 shows the transverse freeze-out radius distributions for $`\pi `$, $`K`$, $`p`$, $`\mathrm{\Lambda }+\mathrm{\Sigma }^0`$, $`\mathrm{\Xi }`$ and $`\mathrm{\Omega }^{}`$ at LHC (top), RHIC (middle) and SPS (bottom). They are rather broad and similar to each other, though the $`\mathrm{\Omega }^{}`$ shows a somewhat narrower freeze-out distribution. The average transverse freeze-out radii are listed in table III; e.g. at RHIC we find 9.5 fm for pions, 10.2 fm for kaons, 11.3 fm for protons, 11.6 fm for Lambda- and Sigma-Hyperons, 14.2 fm for Cascades, but only 7.3 fm for the $`\mathrm{\Omega }^{}`$. The freeze-out of the $`\mathrm{\Omega }^{}`$ occurs rather close to the phase-boundary , due to its very small hadronic interaction cross section. This observation holds true for all three studied beam energies. The respective thickness $`\mathrm{\Delta }r_{had}`$ of the hadronic phase is reduced by a factor of 2 for the $`\mathrm{\Omega }^{}`$, compared to that of the other baryon species. This behavior could be responsible for the experimentally observed hadron-mass dependence of the inverse slopes of the $`m_T`$-spectra at SPS energies . For the $`\mathrm{\Omega }^{}`$, the inverse slope remains practically unaffected by the purely hadronic stage of the reaction, due to its small interaction cross section, while the flow of $`p`$’s and $`\mathrm{\Lambda }`$’s increases (see also section IV E). Figure 7 shows the freeze-out time distributions d$`{}_{}{}^{2}N/`$d$`\tau _{fr}`$d$`y`$ for $`\pi `$, $`p`$ and $`\mathrm{\Omega }^{}`$ at LHC (top), RHIC (middle) and SPS (bottom). Open symbols denote the distributions for a pure hydrodynamical calculation up to hadronization with subsequent hadron resonance decays (but without hadronic reinteraction), whereas the full symbols show the full calculation with hadronic rescattering. As we have already seen previously in the transverse freeze-out radii, hadronic rescattering strongly modifies the shape of the distributions and significantly increases the lifetime of the system. Table IV lists the average freeze-out times for $`\pi `$, $`K`$, $`p`$, $`Y(=\mathrm{\Lambda }+\mathrm{\Sigma }^0)`$, $`\mathrm{\Xi }`$ and $`\mathrm{\Omega }^{}`$ with and without hadronic rescattering. One issue of great interest is the predicted significant increase of the lifetime of the system from SPS to RHIC energies , being due to the time-delay caused by a first-order phase transition . However, our model calculation (which does exhibit a first order phase transition) shows no huge difference in the freeze-out time distributions of $`\pi `$, $`p`$, and $`\mathrm{\Omega }^{}`$ from SPS to RHIC energies (note, however, the logarithmic scale). Origin of this prediction is that we include many more states in the hadronic EoS, which speeds up hadronization considerably . Furthermore, decays of resonances partly hide the remaining small increase of the hadronization time. Thus, the “time-delay signal” can not be expected to be well above $`2030\%`$, and must be approached by a detailed excitation function. Note that the multi-strange $`\mathrm{\Omega }^{}`$ baryons freeze out far earlier than all other baryons, as discussed already previously in the context of figure 6. The duration of the hadronic reinteraction phase, $`\mathrm{\Delta }\tau _{had}=\tau _{fr}^{\mathrm{Hydro}+\mathrm{UrQMD}}\tau _{fr}^{\mathrm{Hydro}+\mathrm{had}.\mathrm{decays}}`$ remains nearly unchanged, e.g. at 5.9 fm/c for pions, 8.0 fm/c for kaons, 14.5 fm/c for protons, 15.4 fm/c for hyperons and 8.0 fm/c for the $`\mathrm{\Omega }^{}`$ between RHIC and SPS. Note that the lifetime of the pre-hadronic stage in this approach is a factor of $`23`$ longer than when employing the parton cascade model (PCM) for the initial reaction stage. It will be interesting to check whether this is related to the first-order phase transition built into the EoS which is used here. The final transverse freeze-out radii and times (after hadronic rescattering), however, are very similar in both approaches . Figure 8 shows the estimated freeze-out volume $`V^{}=\pi r_{T,fr}^2\tau _{fr}`$ as a function of the pion rapidity density d$`N_\pi `$/d$`y`$ for four different bins in transverse momentum. For all $`p_T`$-bins $`V^{}`$ exhibits a nearly linear increase with d$`N_\pi `$/d$`y`$. Thus, the freeze-out density of the pions remains virtually constant over a large range of multiplicities (or energies). We will see in the next section that this is due to the fact that the chemical freeze-out of pions occurs rather shortly after hadronization of the QGP, at all energies studied here. Since the local density of pions on the hadronization hypersurface is similar in all cases (because the temperature is almost the same), the density at chemical freeze-out is, too. We also observe that low-$`p_T`$ pions are basically emitted from the entire volume, while at higher $`p_T`$ the pions seem only to be emitted from an outer shell, the radius of the hollow core increasing with $`p_T`$. The inset of figure 8 shows the dependence of the non-emitting core-volume $`V_0`$ on the transverse momentum of the pions. $`V_0`$ has been calculated by a linear fit of $`V^{}`$ to d$`N_\pi `$/d$`y`$: $`V^{}=V_0+c`$(d$`N_\pi `$/d$`y`$). The increase of $`V_0`$ with $`p_T`$ is a manifestation of the collective flow effect; high-$`p_T`$ pions can not be emitted from the center, $`r_T0`$, since the collective velocity field vanishes there. ### D Chemical freeze-out So far, we have only discussed the kinetic freeze-out of individual hadron species. However, apart from the kinetic freeze-out, the chemical freeze-out of the system, which fixes the chemical composition, is of interest, too. The chemical freeze-out hypersurface of hadron species $`i`$ is in principle defined as the surface $`\sigma _{chem}^\mu `$ separating the space-time region where $`N_i=0`$ from that where the number-current $`N_i`$ is not conserved. Usually, the chemical freeze-out is defined modulo hadronic resonance decays which are performed on $`\sigma _{chem}^\mu `$, even for short-lived resonances like the $`\rho `$-meson or $`\mathrm{\Delta }`$-baryon. However, that definition is not very useful in the present case, since most inelastic processes are actually modeled via resonance excitation and subsequent decay, cf. section III D. Furthermore, as in the case of kinetic freeze-out studied above, the microscopic transport model does not yield sharp hypersurfaces (three-dimensional volumes) but rather freeze-out domains (four-dimensional volumes). We shall therefore mainly discuss the evolution of hadron multiplicities after hadronization, and their time-dependence. Figure 9 shows the time evolution of on-shell hadron multiplicities for LHC (top), RHIC (middle) and SPS (bottom). The dark grey shaded area indicates the duration of the QGP phase, whereas the light grey shaded area depicts the mixed phase (both averaged over $`r_T`$; only hadrons that have already “escaped” from the mixed phase into the purely hadronic phase are shown). Hadronic resonances are formed and are populated for a long time. One can rather nicely observe the stronger transverse expansion as beam energy increases: on $`\tau =\mathrm{const}.`$ hypersurfaces the resonance-decay “tails” get boosted to larger $`\tau `$. Due to those transversely boosted resonances the hadron yields saturate only at rather large $`\tau `$, approximately 25 fm/c at SPS and RHIC and about 40 fm/c at LHC. By comparing the final hadron yields resulting from the hydrodynamical calculation (up to hadronization, including subsequent hadronic decays, but no hadronic reinteractions) to that of the full calculation, which includes microscopic hadronic dynamics, we can quantify the changes of the hadrochemical content due to hadronic rescattering. Figure 10 shows the relative change (in %) of the multiplicity for various hadron species for SPS (bottom), RHIC (middle) and LHC (top). As to be expected, the state of rapid expansion prevailing at hadronization does not allow chemical equilibrium to hold down to much lower temperatures. The hadronic rescattering changes the multiplicities by less than a factor of two, cf. also . Thus, we have first evidence that a QGP expanding and hadronizing as an ideal fluid produces a too rapidly expanding background for a hadron-fluid with known elementary cross-sections to maintain chemical equilibrium down to much lower temperatures than $`T_C`$. However, a closer look provides more insight into the chemical composition. The changes are most pronounced at the SPS, were the baryon-antibaryon asymmetry is highest (since the net-baryon density at mid-rapidity is highest). This manifests e.g. in a reduction of the antiproton multiplicity by 40-50% due to baryon-antibaryon annihilation. $`\overline{\mathrm{\Lambda }}`$ and $`\overline{\mathrm{\Xi }}`$ are affected in similar fashion. The baryon-antibaryon asymmetry decreases at higher beam energy, and at LHC particle-antiparticle symmetry is restored for our initial conditions. The remaining small asymmetries (compare e.g. the $`p`$-$`\overline{p}`$, $`K`$-$`\overline{K}`$, and $`Y`$-$`\overline{Y}`$ evolutions in Fig. 10) are due to fluctuations triggered by the finite number of particles, which distort the ideal longitudinal boost-invariance present (by construction) at hadronization. Interestingly, the $`\mathrm{\Omega }^{}`$ multiplicity decreases stronger towards higher beam energy. This is due to the higher antibaryon density in the system, leading to more $`\mathrm{\Omega }^{}`$ annihilations on antibaryons with subsequent redistribution of the three strange quarks. (This process is modeled in UrQMD as string excitation and subsequent fragmentation, cf. .) Thus the hadronic phase becomes slightly more opaque for the $`\mathrm{\Omega }^{}`$ with increasing beam-energy. Collision rates offer another approach to determine the duration of the hadronic phase, in particular $`B`$-$`\overline{B}`$ collisions which almost always lead to annihilation. Fig. 11 shows the time-evolution of the rates for hadron-hadron collisions at LHC (top), RHIC (middle) and SPS (bottom). Meson-meson (MM) and – to a lesser extent – meson-baryon (MB) interactions dominate the dynamics in the hadronic phase at RHIC and LHC. At the SPS meson-baryon and meson-meson interactions are equally frequent. Note that while at the SPS baryon-baryon (BB) collisions significantly outnumber baryon-antibaryon annihilations, the situation at RHIC and LHC is reversed, where $`B`$-$`\overline{B}`$ annihilation is far more frequent than BB collisions. This is a consequence of the fact that the $`B`$-$`\overline{B}`$ annihilation cross sections at small relative momenta increase faster then the total $`B`$-$`B`$ cross sections . In the case of (approximate) baryon-antibaryon symmetry, one therefore expects more $`B`$-$`\overline{B}`$ than $`B`$-$`B`$ interactions, as seen for RHIC and LHC energies. Of course, all collision rates reach their maxima at the end of the mixed phase, then decreasing roughly according to a power-law. After $`35`$ fm/c, less than one hadron-hadron collision occurs per unit of time and rapidity at SPS and RHIC energies; due to the higher transverse $`\gamma `$-factor the time is $`60`$ fm/c at the LHC. At this stage the system is certainly kinetically and chemically frozen-out. ### E Transverse flow: Emission of multi-strange baryons from the phase-boundary In this section we analyze the transverse mass spectra at freeze-out, and discuss their evolution from the hadronization hypersurface. The results obtained for Pb+Pb collisions at CERN-SPS energy are in reasonable agreement with the data obtained by the NA49-collaboration and by the WA97-collaboration . For a comparison to that data we refer to ; here, we focus on the model-results. Fig. 12 compares the $`m_T`$-spectra on the hadronization hypersurface (open symbols), obtained from Eq. (25) (plus strong resonance decays), with those at freeze-out (full symbols). One observes that the transverse flow of $`p`$’s and $`\mathrm{\Lambda }`$’s increases during the hadronic stage, since those spectra flatten. On the other hand, the spectra of $`\mathrm{\Omega }`$’s and of $`\mathrm{\Xi }`$’s with $`m_T\stackrel{>}{}1.6`$ GeV are practically unaffected by the hadronic stage and closely resemble those on the phase boundary. This is due to the fact that the scattering rates of $`\mathrm{\Xi }`$ and $`\mathrm{\Omega }`$ in a pion-rich hadron gas are significantly smaller than those of $`N`$’s and $`\mathrm{\Lambda }`$’s . As shown in Fig. 13, on average the baryons which finally emerge as $`\mathrm{\Xi }`$’s and $`\mathrm{\Omega }`$’s suffer far less interactions than the final-state $`p`$’s and $`\mathrm{\Lambda }`$’s. Thus, within the model presented here, these particles are basically emitted directly from the phase boundary with very little further rescattering in the hadronic stage. The hadron gas emerging from the hadronization of the QGP (in these high-energy reactions) is almost “transparent” for the multiple strange baryons. On the other hand, $`p`$’s and $`\mathrm{\Lambda }`$’s on average suffer several collisions with other hadrons before they freeze-out. This behavior holds generally true for all three studied energy domains, at the SPS, RHIC and LHC. These findings manifest themselves most strikingly in the mass-dependence of the inverse slopes of the $`m_T`$ spectra. A simple isentropic hydrodynamical expansion leads to broader $`m_T`$ spectra of heavier states, i.e. $`p_T`$ or the inverse slope $`T^{}`$ increase with mass . This observation agrees with the inverse slopes of $`\pi `$, $`K`$, and $`p`$ measured for central collisions of $`Pb`$ nuclei at a CM-energy of $`17A`$ GeV . However, it has also been found that the $`\mathrm{\Xi }`$ and $`\mathrm{\Omega }`$ baryons do not follow this general trend . Fig. 14 depicts the inverse slopes $`T^{}`$ obtained from our model by a fit of $`\mathrm{d}^3N_i/\mathrm{d}^2m_T\mathrm{d}y`$ to $`\mathrm{exp}(m_T/T^{})`$ in the range $`m_Tm_i<1`$ GeV. The statistical error of this fit is $`10\%`$. Open symbols denote the SPS calculation and data, whereas full symbols show the RHIC prediction. The lines show a purely hydrodynamical calculation with a freeze-out temperature of $`T_{fo}=130`$ MeV for SPS (dotted line) and RHIC (full line), respectively. The trend of the SPS data (open circles), namely the “softer” spectra of $`\mathrm{\Xi }`$’s and $`\mathrm{\Omega }`$’s as compared to a linear $`T^{}(m)`$ relation, is reproduced reasonably well. As already mentioned, this is not the case for “pure” hydrodynamics with kinetic freeze-out on a common hypersurface (e.g. the $`T=130`$ MeV isotherm), where the stiffness of the spectra increases monotonically with mass, cf. Fig. 14 and also refs. . Resonance decays are not included in the hydrodynamic spectra on the $`T=130`$ MeV isotherm. When going from SPS to RHIC energy, the model discussed here generally yields only a slight increase of the inverse slopes, although the specific entropy is larger by a factor of 4-5 ! The reason for this behavior is the first-order phase transition that softens the transverse expansion considerably . For our set of initial conditions, the average collective transverse flow velocity (at mid-rapidity) on the hadronization hypersurface increases only from $`0.3`$ (for Pb+Pb at SPS) to $`0.35`$ (for Au+Au at RHIC) . (However, there are high-$`v_T`$ tails on the hadronization hypersurface which get more pronounced at RHIC than at SPS.) As can be seen from the present calculation, this is not counterbalanced by increased rescattering in the purely hadronic stage – compare to the inverse slopes obtained from “pure” hydrodynamics with freeze-out on the $`T=130`$ MeV isotherm! The transverse flow at LHC beam energy is so strong that the $`m_T`$-spectra can not be fitted any more by an exponential distribution. We have therefore refrained to extract the slopes for the LHC calculation. Instead, in figure 15 we show the mean transverse momenta of the different hadron species as a function of their mass. As in figure 12 we compare the $`p_T`$ on the hadronization hypersurface (open symbols), obtained from Eq. (25) (plus strong resonance decays), with that at freeze-out (full symbols). Hadronic rescattering leads to a transfer of transverse energy/momentum from pions to heavier hadrons (the pions actually suffer a reduction of $`p_T`$ in the hadronic phase) . This phenomenon has also been termed the pion-wind , pushing heavier hadrons to higher $`p_T`$. Nucleons gain most transverse momentum, while the $`\mathrm{\Omega }^{}`$ remains nearly unchanged due to its small interaction cross section in the meson dominated hadronic medium, as discussed earlier in this section. Those hadrons are the best “messengers” of the early pre-hadronization evolution. Furthermore, one clearly observes the rather moderate increase of $`p_T`$ from SPS to RHIC energy, as discussed already in Fig. 14. In contrast, in our model the collective dynamics at the much higher CERN-LHC energy is dominated by the stiff QGP, cf. also Fig. 3, and the average transverse momenta increase appreciably. ## V Summary and outlook In summary, we have introduced a combined macroscopic/microscopic transport approach, combining relativistic hydrodynamics for the early deconfined stage of the reaction and the hadronization process with a microscopic non-equilibrium model for the later hadronic stage at which the hydrodynamic equilibrium assumptions are not valid anymore. Within this approach we have self-consistently calculated the freeze-out of the hadronic system, accounting for the collective flow on the hadronization hypersurface generated by the QGP expansion. The reaction dynamics, hadronic freeze-out and transverse flow in ultra-relativistic heavy ion collisions at SPS, RHIC and LHC have been discussed in detail. We find that the space-time domains of the freeze-out for the investigated hadron species are actually four-dimensional, and differ drastically between the individual hadrons species. The thickness of the hadronic phase is found to be between 2 fm and 6 fm (at RHIC), depending on the respective hadron species. Its lifetime is between 5 fm/c and 13 fm/c, respectively. Freeze-out radii distributions have similar widths for most hadron species, though the $`\mathrm{\Omega }^{}`$ is found to be emitted rather close to the phase boundary and shows the smallest freeze-out radii and times among all baryon species. The total lifetime of the system does not increase drastically when going from SPS to RHIC energies. Our model-calculation shows that in high-energy nuclear collisions the hadron multiplicities at midrapidity change by less than $`40\%`$ after hadronization, unlike e.g. in the early universe. However, a closer look is warranted and reveals interesting information. For example, more strange baryons ($`\mathrm{\Lambda }`$, $`\mathrm{\Sigma }`$, $`\mathrm{\Xi }`$, $`\mathrm{\Omega }`$) are annihilated as the energy increases because the anti-baryon density at hadronization increases. Interactions within the hadron gas increase the collective flow beyond that present at hadronization, and reduce the temperature below the QCD phase transition temperature (we assume $`T_C=160`$ MeV). As an exception, we find that multiple strange baryons practically do not rescatter within the hadron gas. Their $`m_T`$-spectra are therefore determined by the conditions on the hadronization hypersurface, i.e. $`T_C`$ and the collective flow created by the expansion preceding hadronization. Their spectra therefore are less sensitive to the confined phase, $`T<T_C`$, but are closely related to the EoS of the QGP and the phase transition temperature $`T_C`$. Average transverse momenta and inverse slopes are predicted to increase only moderately from SPS to RHIC, despite the significant increase of the entropy to net baryon ratio. In this sense, the collective evolution at RHIC energy is strongly characterized by the presence of a well-mixed coexistence phase with small isentropic speed of sound. It will be very interesting to see if this picture of hadronization of bulk QCD matter, which is based on similar models for the QCD phase transition in the much slower expanding early universe, agrees with the data to be taken by the various experiments at BNL-RHIC. Towards the much higher CERN-LHC energy, the evolution changes appreciably. The pure QGP occupies a larger space-time volume than the mixed phase. If the QGP-EoS at high energy density is anywhere close to an ultrarelativistic ideal gas with $`pϵ/3`$, transverse expansion should be much stronger than at RHIC and SPS. Consequently, the average transverse momenta of the heavier hadrons increase by $`6080\%`$ as compared to RHIC energy. We believe that the model presented here and in refs. represents a step forward towards the understanding and the description of the evolution of a quark-gluon plasma, its hadronization, and the subsequent freeze-out of the strong interactions. Nevertheless, it is clear that many improvements are thinkable and necessary before a really detailed comparison to experimental data can be attempted. For example, corrections to ideal fluid dynamics (before hadronization) should be studied, at least within the Navier-Stokes approximation. In the present approach dissipative effects are only taken into account after hadronization, where we expect them to be most significant, particularly as freeze-out is approached. The widely used Bag-model EoS can certainly be improved as well. It is well known that it yields a substantially higher latent heat than extracted from present lattice-QCD results. Thus, it may over-pronounce the effects of a first-order QCD phase transition. One may even try a cross-over transition to see whether that is ruled out by experimental data or not. Also, we have already commented on the fact that due to the huge expansion rate in high-energy collisions (which is not much smaller than strong interaction rates) more radical scenarios like spinodal decomposition rather than an adiabatic phase transition should be examined as well. To simplify the switch from hydrodynamics to the microscopic transport model we assumed longitudinal boost-invariance and azimuthal symmetry. The latter approximation, in particular, disables us to study many up-to-date topics that will be addressed by the experimental data, as e.g. anisotropies in the hadron spectra (in non head-on collisions) that may be sensitive to the QGP-EoS. A fully 3+1 dimensional solution without symmetry assumptions is thus highly desirable. One application not covered at all in the present studies is correlated particle emission. Two-particle correlations may for example allow to extract the volume of the emission region in space-time. The two-particle correlator probes rather soft interactions and therefore contains information about the freeze-out process (e.g. the thickness of the emission region, background mean-fields etc.) . Fluctuations in the rapidity and transverse momentum spectra induced (or suppressed) by the QCD phase transition are another highly interesting topic. So far, we have studied only the evolution on average, and have allowed only for fluctuations to develop in the post-hadronization stage (which occur naturally in the microscopic transport model). However, the hadronization process could in principle induce larger or other types of fluctuations, e.g. due to droplet formation , spinodal decomposition with possible DCC formation , or due to the change of the order of the phase transition in the vicinity of a second order critical point . It would obviously be highly desirable to know if they can survive the hadronic interaction stage. Other fluctuations may be there from the very beginning, e.g. those arising in the production process of secondary hadrons, and one could study the evolution of strangeness-rich rapidity bins , or of rapidity bins with negative baryon-number, through the hadronization phase transition until freeze-out. There are many other questions that can not be listed here but can be addressed within this model. Work along those lines is in progress and will be reported in forthcoming publications. ###### Acknowledgements. S.A.B. has been supported in part by the Alexander von Humboldt Foundation through a Feodor Lynen Fellowship, and in part by DOE grant DE-FG02-96ER40945. A.D. acknowledges support from the DOE Research Grant, Contract No. De-FG-02-93ER-40764. S.A.B. thanks Berndt Müller for many helpful and inspiring discussions. A.D. thanks M. Bleicher, K.A. Bugaev, W. Greiner, M. Gyulassy, D.H. Rischke, and H. Stöcker for numerous inspiring discussions. The computer-programs (implemented in FORTRAN 77) with which the numerical calculations described in this paper have been performed can be obtained free of charge (but copyright protected) from the “Open Standard Codes and Routines” working group homepage, http://rhic.phys.columbia.edu/oscar. The authors thank the UrQMD collaboration for the permission to use the UrQMD transport model for the actual computations performed in this manuscript; and Dirk Rischke for the permission to use parts of his RHLLE implementation of the fluid-dynamical equations, to solve the fluid-dynamical continuity equations.
warning/0001/hep-th0001213.html
ar5iv
text
# 1 Introduction ## 1 Introduction Recently there has been much interest in studying the properties of the (D$`(p2)`$, D$`p`$) bound states and their consequences (see for example - and references therein). Such bound states can be viewed as D$`p`$-branes with a nonzero (rank two) Neveu-Schwarz (NS) $`B`$ field. At present, it is known that the worldvolume coordinates will become noncommutative if a D$`p`$-brane carries a nonvanishing NS $`B`$ field on its worldvolume -, and the field theories on the worldvolume of such D-branes are called noncommutative field theories in order to distinguish the ordinary field theories on the worldvolume of D-branes without NS $`B`$ field. The gauge theories on the noncommutative spacetimes can naturally be realized in string theories. According to the Maldacena conjecture -, the string theories can be used to study the large $`N`$ noncommutative field theories in the strong ’t Hooft coupling limit. On the basis of consideration of planar diagrams, Bigatti and Susskind argued that the large $`N`$ noncommutative and ordinary gauge field theories are equivalent in the weak coupling limit and noncommutative effects can be seen only in the nonplanar diagrams. Explicit perturbative calculations render evidence to this assertion. On the supergravity side, it has also been found that the thermodynamics of near-extremal D$`p`$-branes with a nonvanishing NS $`B`$ field coincides exactly with that of the corresponding D$`p`$-branes without $`B`$ field , which means that in the large $`N`$ and strong coupling limit, the number of the degrees of freedom of the noncommutative gauge fields remains unchanged despite the noncommutativity of space. More recently, Lu and Roy have found that in the system of (D$`(p2)`$, D$`p`$) bound states ($`2p6`$), the noncommutative effects of gauge fields are actually due to the presence of infinitely many D$`(p2)`$-branes which play the dominant role over the D$`p`$-branes in the large $`B`$ field limit. The D$`p`$-branes with a constant $`B`$ field represent dynamically the system of infinitely many D$`(p2)`$-branes with two smeared transverse coordinates (additional isometrics) and no $`B`$ field in the decoupling limit. With this observation, Lu and Roy further argued that there is an equivalence between the noncommutative super Yang-Mills theory in $`(p+1)`$ dimensions and an ordinary one with gauge group $`U(\mathrm{})`$ in $`(p1)`$ dimensions. For related discussions, see also . In the present paper, we would like to discuss this equivalence from the viewpoint of the thermodynamics of the (D$`(p2)`$, D$`p`$) bound states in the dual gravity description with two dimensions compactified on a torus. The case discussed by Lu and Roy corresponds to the infinite volume limit of the torus, and we would like to clarify some subtle questions in this analysis. In the next section, we study the black (D$`(p2)`$, D$`p`$) configuration and some of its basic thermodynamic properties. As mentioned above, it has been noticed that the thermodynamics of the black D$`p`$-branes with $`B`$ field is the same as that of D$`p`$-branes without $`B`$ field. We show here that the thermodynamics of the D$`p`$-branes with nonzero $`B`$ field is also completely the same as that of D$`(p2)`$-branes with two smeared coordinates and zero $`B`$ field. We obtain a relation \[eq. (2.21) below\] between the numbers of D$`p`$-branes with $`B`$ field and the D$`(p2)`$-branes without $`B`$ field when this equivalence is valid. Since the worldvolume theory of D$`p`$-branes with $`B`$ field is the noncommutative super Yang-Mills with gauge group $`U(N_p)`$ in ($`p+1`$) dimensions with two dimensions compactified on a torus and that of D$`(p2)`$-branes with two smeared coordinates and zero $`B`$ field is an ordinary super Yang-Mills with gauge group $`U(N_{p2})`$ in ($`p+1`$) dimensions on the dual torus, as we will see, this implies that these theories are equivalent in the large $`N`$ limit. When the volume of the torus is sent to infinity, the latter theory reduces to $`U(\mathrm{})`$ gauge theory in ($`p1`$) dimensions for fixed $`N_p`$, in agreement with . This is also in accordance with the proposal that D2-brane is a condensate of D0-branes . Furthermore, we give the relation of Yang-Mills coupling constants between the theories in different dimensions. Investigating the interactions between a probe and a source is a useful method to see some of the properties of the source. For instance, a scalar field has been used as a probe to find out the noncommutative effects on the absorption by the (D1, D3) bound state. In section 3, we study the descriptions of the (D$`(p2)`$, D$`p`$) bound state in terms of the D$`p`$-branes with $`B`$ field and in terms of the D$`(p2)`$-branes without $`B`$ field by analyzing the thermodynamics of two probes, one of which is a bound state of D$`(p2)`$\- and D$`p`$-branes and the other is a D$`(p2)`$-brane. In the decoupling limit, we find a relation (3.19) similar to (2.21). In section 4, we further reveal the equivalence of the descriptions by examining the dynamics of the probes in the bound state backgrounds. We find that non-extremal D-branes can be located at the horizon from the viewpoint of probes. We summarize our results in section 5. ## 2 The (D$`(p2)`$, D$`p`$) bound states and implications of their thermodynamics The supergravity solutions of the (D$`(p2)`$, D$`p`$) bound states ($`2p6`$ ) in type II superstring theories have been constructed by many authors in . The supergravity solutions of the (D$`(p2)`$, D$`p`$) bound states can be acquired by taking the extremal limit of the corresponding black configurations. We start with the general solution $`ds^2=H^{1/2}[fdt^2+dx_1^2+\mathrm{}+dx_{p2}^2+h(dx_{p1}^2+dx_p^2)]+H^{1/2}(f^1dr^2+r^2d\mathrm{\Omega }_{8p}^2),`$ $`e^{2\varphi }=g^2H^{\frac{3p}{2}}h,B_{p1,p}=\mathrm{tan}\theta H^1h,`$ $`A_{012\mathrm{}p}^p=g^1(H^11)h\mathrm{cos}\theta \mathrm{coth}\alpha ,A_{012\mathrm{}(p2)}^{p2}=g^1(H^11)\mathrm{sin}\theta \mathrm{coth}\alpha ,`$ (2.1) and $$H=1+\frac{r_0^{7p}\mathrm{sinh}^2\alpha }{r^{7p}},f=1\left(\frac{r_0}{r}\right)^{7p},h^1=\mathrm{cos}^2\theta +H^1\mathrm{sin}^2\theta .$$ (2.2) Here $`g`$ is the string coupling constant, $`r_0`$ is the non-extremal Schwarzschild mass parameter and $`\alpha `$ is the boost parameter. The solution (2.1) interpolates between the black D$`(p2)`$-brane solution with two smeared coordinates $`x_{p1}`$ and $`x_p`$ ($`\theta =\pi /2`$), and the black D$`p`$-brane with zero $`B`$ field ($`\theta =0`$). Note that a constant part of the $`B`$ field can be gauged away so that the constant value for $`B_{p1,p}`$ can be changed. The parameter $`\theta `$ characterizes the interpolation. The coordinates $`x_{p1}`$ and $`x_p`$ are relative transverse directions for the D$`(p2)`$-branes and parametrize a rectangular 2-torus. Denote the area of the 2-torus spanned by $`x_{p1}`$ and $`x_p`$ by $`V_2`$ and the spatial volume of the D$`(p2)`$-brane with worldvolume coordinates $`(t,x_1,\mathrm{},x_{p2})`$ by $`V_{p2}`$. The spatial volume of the D$`p`$-brane with worldvolume coordinates $`(t,x_1,\mathrm{},x_p)`$ is then $`V_p=V_{p2}V_2`$. The charge density of the D$`p`$-brane in the bound state system is given by $$Q_p=\frac{1}{2\kappa ^2}_{\mathrm{\Omega }_{8p}}F_{p+2}=\frac{(7p)\mathrm{\Omega }_{8p}\mathrm{cos}\theta }{2\kappa ^2g}r_0^{7p}\mathrm{sinh}\alpha \mathrm{cosh}\alpha ,$$ (2.3) where $`2\kappa ^2=(2\pi )^7\alpha ^4`$ is the gravity constant in ten dimensions and $`\mathrm{\Omega }_{8p}`$ is the volume of a unit $`(8p)`$-sphere: $$\mathrm{\Omega }_{8p}=\frac{2\pi ^{(9p)/2}}{\mathrm{\Gamma }[(9p)/2]}=\frac{4\pi \pi ^{(7p)/2}}{(7p)\mathrm{\Gamma }[(7p)/2]}.$$ (2.4) The D$`(p2)`$-brane charge density on its worldvolume is $$Q_{p2}=\frac{1}{2\kappa ^2}_{V_2\times \mathrm{\Omega }_{8p}}F_p=\frac{(7p)\mathrm{\Omega }_{8p}V_2\mathrm{sin}\theta }{2\kappa ^2g}r_0^{7p}\mathrm{sinh}\alpha \mathrm{cosh}\alpha .$$ (2.5) In fact the D$`(p2)`$-brane charge density on the worldvolume of D$`p`$-brane is $$\stackrel{~}{Q}_{p2}=\frac{Q_{p2}}{V_2}.$$ (2.6) According to the charge quantization rule $`Q_p=T_pN_p`$ in terms of the tension of the D$`p`$-branes, we can obtain the number $`N_p`$ of the D$`p`$-branes and $`N_{p2}`$ of the D$`(p2)`$-branes in the bound states. Defining $`\stackrel{~}{R}^{7p}=r_0^{7p}\mathrm{sinh}\alpha \mathrm{cosh}\alpha `$, we have the relation between the number $`N_p`$ of D$`p`$-branes and $`N_{p2}`$ of D$`(p2)`$-branes: $$\stackrel{~}{R}^{7p}=N_p\frac{2\kappa ^2gT_p}{(7p)\mathrm{\Omega }_{8p}\mathrm{cos}\theta }=N_{p2}\frac{2\kappa ^2gT_{p2}}{(7p)\mathrm{\Omega }_{8p}V_2\mathrm{sin}\theta },$$ (2.7) where the tensions $`T_p`$ and $`T_{p2}`$ have a unified expression as $`T_p=(2\pi )^p(\alpha ^{})^{(p+1)/2}`$. From (2.3) and (2.5), we can see that the asymptotic value $`\mathrm{tan}\theta `$ of the $`B`$ field has the following relation to the charges of the D$`p`$\- and D$`(p2)`$-branes: $$\mathrm{tan}\theta =\frac{\stackrel{~}{Q}_{p2}}{Q_p}=\frac{1}{V_2}\frac{Q_{p2}}{Q_p}=\frac{1}{V_2}\frac{T_{p2}}{T_p}\frac{N_{p2}}{N_p}.$$ (2.8) The solution (2.1) has the event horizon at $`r=r_0`$ and hence has the associated thermodynamics. A standard calculation gives us the ADM mass $`M`$, Hawking temperature $`T`$ and entropy $`S`$ of the black configuration: $`M={\displaystyle \frac{(8p)\mathrm{\Omega }_{8p}V_pr_0^{7p}}{2\kappa ^2g^2}}\left(1+{\displaystyle \frac{7p}{8p}}\mathrm{sinh}^2\alpha \right),`$ $`T={\displaystyle \frac{7p}{4\pi r_0\mathrm{cosh}\alpha }},`$ $`S={\displaystyle \frac{4\pi \mathrm{\Omega }_{8p}V_p}{2\kappa ^2g^2}}r_0^{8p}\mathrm{cosh}\alpha .`$ (2.9) It is somewhat surprising that these thermodynamic quantities are completely the same as those for the D$`p`$-branes without $`B`$ field, just as noticed in . The conclusion remains valid even if the angular rotation is introduced . Here we focus on another aspect of the thermodynamics of this black configuration: These thermodynamic quantities are all independent of the parameter $`\theta `$. As pointed out above, the parameter $`\theta `$ characterizes the interpolation of the solution between the black D$`p`$-brane without $`B`$ field and the black D$`(p2)`$-brane with two smeared coordinates and zero $`B`$ field (a nonvanishing constant $`B`$ field along the directions transverse to the D$`(p2)`$-branes can be gauged away). Thus these thermodynamic quantities are also those of the black D$`(p2)`$-branes with two smeared coordinates and no $`B`$ field. Therefore there is the thermodynamic equivalence not only between black D$`p`$-branes with $`B`$ field and those without $`B`$ field, but also between the black D$`p`$-branes with $`B`$ field and black D$`(p2)`$-branes with two smeared coordinates and no $`B`$ field. This fact is important in the following discussions. The thermodynamic quantities (2.9) with the charges (2.3) and (2.5) satisfy the first law of black hole thermodynamics as expected: $`dM`$ $`=`$ $`TdS+\mu _pdq_p+\mu _{p2}dq_{p2}`$ (2.10) $`=`$ $`TdS+\mu _pV_pT_pdN_p+\mu _{p2}V_{p2}T_{p2}dN_{p2},`$ where $`\mu _p`$ and $`\mu _{p2}`$ are the chemical potentials corresponding to the total charges $`q_p=Q_pV_p`$ and $`q_{p2}=V_{p2}Q_{p2}`$, respectively, $$\mu _p=\mathrm{cos}\theta \mathrm{tanh}\alpha /g,\mu _{p2}=\mathrm{sin}\theta \mathrm{tanh}\alpha /g.$$ (2.11) In addition, we notice that in the extremal limit (by taking $`r_00`$ and $`\alpha \mathrm{}`$, but keeping $`\stackrel{~}{R}^{7p}`$ constant), $$M_{\mathrm{ext}.}^2=q_p^2+q_{p2}^2,$$ (2.12) which indicates that the bound state (D$`(p2)`$, D$`p`$) is a nonthreshold one. Now we turn to the field theory limit (decoupling limit) of the bound state solution (2.1), in which the gravity decouples from the field theory on the worldvolume of D$`p`$-branes. Following , in the decoupling limit $`\alpha ^{}0:`$ $`\mathrm{tan}\theta ={\displaystyle \frac{\stackrel{~}{b}}{\alpha ^{}}},r=\alpha ^{}u,r_0=\alpha ^{}u_0,`$ $`g=\stackrel{~}{g}\alpha ^{(5p)/2},x_{0,1,\mathrm{},p2}=\stackrel{~}{x}_{0,1,\mathrm{},p2},x_{p1,p}={\displaystyle \frac{\alpha ^{}}{\stackrel{~}{b}}}\stackrel{~}{x}_{p1,p},`$ (2.13) and $`\stackrel{~}{b}`$, $`\stackrel{~}{g}`$, $`u`$, $`u_0`$, and $`\stackrel{~}{x}_\mu `$ held fixed, one has the following decoupling limit solution: $`ds^2=\alpha ^{}[\left({\displaystyle \frac{u}{R}}\right)^{(7p)/2}(\stackrel{~}{f}dt^2+d\stackrel{~}{x}_1^2+\mathrm{}+d\stackrel{~}{x}_{p2}^2+\stackrel{~}{h}(d\stackrel{~}{x}_{p1}^2+d\stackrel{~}{x}_p^2))`$ $`+\left({\displaystyle \frac{R}{u}}\right)^{(7p)/2}(\stackrel{~}{f}^1du^2+u^2d\mathrm{\Omega }_{8p}^2)],`$ $`e^{2\varphi }=\stackrel{~}{g}^2\stackrel{~}{b}^2\stackrel{~}{h}\left({\displaystyle \frac{R}{u}}\right)^{(7p)(3p)/2},B_{p1,p}={\displaystyle \frac{\alpha ^{}}{\stackrel{~}{b}}}{\displaystyle \frac{(au)^{7p}}{1+(au)^{7p}}},`$ (2.14) where the corresponding RR fields are not exposed explicitly, $$\stackrel{~}{f}=1\left(\frac{u_0}{u}\right)^{7p},\stackrel{~}{h}=\frac{1}{1+(au)^{7p}},a^{7p}=\stackrel{~}{b}^2/R^{7p},$$ (2.15) and $$R^{7p}=\frac{1}{2}(2\pi )^{6p}\pi ^{(7p)/2}\mathrm{\Gamma }[(7p)/2]\stackrel{~}{g}\stackrel{~}{b}N_p.$$ (2.16) According to the generalized AdS/CFT correspondence, the solution (2.14) is the dual gravity description of the noncommutative gauge field theory with gauge group $`U(N_p)`$ in ($`p+1)`$ dimensions . When $`a=0`$, the solution (2.14) reduces to the usual decoupling limit solution of D$`p`$-branes without $`B`$ field . This implies that the noncommutativity effect is weak in field theories for $`au<<1`$ or at long distance. In the decoupling limit, the thermal excitations above the extremality have the energy $`E`$, temperature $`T`$ and entropy $`S`$: $`E={\displaystyle \frac{(9p)\mathrm{\Omega }_{8p}\stackrel{~}{V}_p}{2(2\pi )^7(\stackrel{~}{g}\stackrel{~}{b})^2}}u_0^{7p},`$ $`T={\displaystyle \frac{7p}{4\pi }}R^{\frac{7p}{2}}u_0^{\frac{5p}{2}},`$ $`S={\displaystyle \frac{2\mathrm{\Omega }_{8p}\stackrel{~}{V}_p}{(2\pi )^6(\stackrel{~}{g}\stackrel{~}{b})^2}}R^{\frac{7p}{2}}u_0^{(9p)/2}.`$ (2.17) Here $`\stackrel{~}{V}_p=V_{p2}\stackrel{~}{V_2}`$ is the spatial volume of the D$`p`$-brane after taking the decoupling limit (2.13), and $`\stackrel{~}{V}_2=V_2\stackrel{~}{b}^2/\alpha ^2`$ is the area of the torus. Using (2.17), one finds the free energy, defined as $`F=ETS`$, of the thermal excitations: $`F`$ $`=`$ $`{\displaystyle \frac{(5p)\mathrm{\Omega }_{8p}\stackrel{~}{V}_p}{2(2\pi )^7(\stackrel{~}{g}\stackrel{~}{b})^2}}u_0^{7p}`$ (2.18) $`=`$ $`{\displaystyle \frac{\mathrm{\Omega }_{8p}V_{p2}\stackrel{~}{V}_2}{(2\pi )^7\stackrel{~}{g}^2\stackrel{~}{b}^2}}{\displaystyle \frac{5p}{2}}\left({\displaystyle \frac{4\pi }{7p}}\right)^{\frac{2(7p)}{5p}}R^{\frac{(7p)^2}{5p}}T^{\frac{2(7p)}{5p}},`$ in terms of the temperature. In the decoupling limit, the numbers of two kinds of branes are constants. Hence the first law of thermodynamics becomes $$dE=TdS,\mathrm{and}dF=SdT.$$ (2.19) In the spirit of the generalized AdS/CFT correspondence, the thermodynamics of the thermal excitations should be equivalent to that of the corresponding noncommutative gauge fields at finite temperature $`T`$ in the large $`N`$ and strong ’t Hooft coupling limit. We notice that these thermodynamic quantities, after rescaling the string coupling constant as $`\stackrel{~}{g}\stackrel{~}{b}=\widehat{g}`$, are exactly the same as those of the black D$`p`$-branes without $`B`$ field in the decoupling limit. This means that in this supergravity approximation, the thermodynamics of the large $`N`$ noncommutative and ordinary gauge field theories both in $`(p+1)`$ dimensions are equivalent to each other. This also implies that in the planar limit, the number of the degrees of freedom in the noncommutative gauge theories coincides with that in the ordinary field theories not only in the weak coupling limit , but also in the strong coupling limit . Now let us recall the fact that the thermodynamics (2.9) of the black D$`p`$-branes with nonzero $`B`$ field is the same as that of the black D$`(p2)`$-branes with two smeared coordinates and zero $`B`$ field. In the decoupling limit (2.13), the solution (2.14) is described by the quantities of D$`p`$-branes. For instance, $`R^{7p}`$ is proportional to the number $`N_p`$ of the coinciding D$`p`$-branes \[see (2.16)\]. In fact the “radius” $`R`$ can also be expressed by quantities of D$`(p2)`$-branes. Using (2.8), we obtain $$R^{7p}=\frac{1}{2}(2\pi )^{6p}\pi ^{(7p)/2}\mathrm{\Gamma }[(7p)/2]\stackrel{~}{g}\stackrel{~}{b}N_{p2}\times \frac{(2\pi )^2\stackrel{~}{b}}{\stackrel{~}{V_2}}.$$ (2.20) In the decoupling limit, we are thus led to the relation between the numbers of D$`p`$\- and D$`(p2)`$-branes: $$\mathrm{tan}\theta =\frac{\stackrel{~}{b}}{\alpha ^{}}=\frac{(2\pi )^2\stackrel{~}{b}^2}{\alpha ^{}\stackrel{~}{V}_2}\frac{N_{p2}}{N_p}\frac{N_{p2}}{N_p}=\frac{\stackrel{~}{V}_2}{(2\pi )^2\stackrel{~}{b}}.$$ (2.21) Because $`\stackrel{~}{V}_2`$ and $`\stackrel{~}{b}`$ can be kept finite, we can thus conclude that in the decoupling limit of the D$`p`$-branes with NS $`B`$ field, the number of the D$`(p2)`$-branes can be kept finite. This looks different from the claim by Lu and Roy where they conclude that the number of the D$`(p2)`$-branes becomes infinity in the decoupling limit. This is so because they take a little different decoupling limit and there $`\stackrel{~}{x}_{p1}`$ and $`\stackrel{~}{x}_p`$ are infinitely extended. Mathematically our decoupling limit becomes the same as theirs by taking $`\stackrel{~}{V}_2\mathrm{}`$ but keeping $`N_{p2}/\stackrel{~}{V}_2=N_p/(2\pi )^2\stackrel{~}{b}`$ finite. In the decoupling limit (2.13), $`\mathrm{tan}\theta \mathrm{}`$ as $`\alpha ^{}0`$. If we set $`\theta =\pi /2`$, the solution (2.1) reduces to the black D$`(p2)`$-brane with two smeared coordinates and zero $`B`$ field after gauging away the constant value. This shift of the constant $`B`$ is allowed in the large $`N`$ limit .<sup>1</sup><sup>1</sup>1Alternatively, if one simply takes the usual D$`(p2)`$-brane solution, there is no $`B`$ field. The decoupling limit solution (2.14) for the black D$`p`$-brane with NS $`B`$ field is thus expected to be related with the solution of black D$`(p2)`$-brane with two smeared coordinates and no $`B`$ field in the same decoupling limit. For our convenience, we rewrite the black D$`(p2)`$-brane with two smeared coordinates: $`ds^2=H^{1/2}[fdt^2+dx_1^2+\mathrm{}+dx_{p2}^2+H(dx_{p1}^2+dx_p^2)]+H^{1/2}(f^1dr^2+r^2d\mathrm{\Omega }_{8p}),`$ $`e^{2\varphi }=g^2H^{(5p)/2},A_{01\mathrm{}(p2)}^{p2}=g^1(H^11)\mathrm{coth}\alpha ,B_{p1,p}=0,`$ (2.22) where $`H`$ and $`f`$ are the same as those in (2.2). Note that here $`x_{p1}`$ and $`x_p`$ are two smeared transverse coordinates for the D$`(p2)`$-branes. In the decoupling limit (2.13), we reach $`ds^2=\alpha ^{}[\left({\displaystyle \frac{u}{R}}\right)^{(7p)/2}(\stackrel{~}{f}dt^2+d\stackrel{~}{x}_1^2+\mathrm{}+d\stackrel{~}{x}_{p2}^2+{\displaystyle \frac{1}{(au)^{7p}}}(d\stackrel{~}{x}_{p1}^2+d\stackrel{~}{x}_p^2))`$ $`+\left({\displaystyle \frac{R}{u}}\right)^{(7p)/2}(\stackrel{~}{f}^1du^2+u^2d\mathrm{\Omega }_{8p}^2)],`$ $`e^{2\varphi }=\stackrel{~}{g}^2\stackrel{~}{b}^{5p}(au)^{(7p)(p5)/2},B_{p1,p}=0,`$ (2.23) where $`\stackrel{~}{f}`$ and $`R^{7p}`$ are given in (2.15) and (2.20), respectively. When $`\stackrel{~}{f}=1`$, the solution reduces to that given in . Obviously for $`au>>1`$, the decoupling solution (2.14) of the D$`p`$-brane with NS $`B`$ field is indeed equivalent to the decoupling limit solution (2.23) of the black D$`(p2)`$-branes with two smeared coordinates and no NS $`B`$ field, as noticed in . (We will also discuss the equivalence from the thermodynamics point of view shortly.) Note that the coordinate $`u`$ corresponds to an energy scale of worldvolume gauge field theories, and in (2.14) $`au`$ reflects the noncommutative effect of gauge fields. It has been found that in order for the dual gravity description (2.14) of noncommutative gauge fields to be valid, $`au>>1`$ should be satisfied , in which case $`N_p`$ can be small and the noncommutativity effect is strong in the corresponding field theories. We know that the solution (2.14) is a dual gravity description of a noncommutative super Yang-Mills theory with gauge group $`U(N_p)`$ in ($`p+1)`$ dimensions. What is the field theory corresponding to the supergravity solution (2.23) for large $`au`$? To see this, let us note that the supergravity description (2.23) breaks down for large $`au`$ since the effective size of the torus shrinks. Nevertheless, we can make a T-duality along the directions $`\stackrel{~}{x}_{p1}`$ and $`\stackrel{~}{x}_p`$. We then obtain a usual decoupling limit solution of $`N_{p2}`$ coincident D$`p`$-branes without $`B`$ field : $`ds^2`$ $`=`$ $`\alpha ^{}[\left({\displaystyle \frac{u}{R}}\right)^{(7p)/2}(\stackrel{~}{f}dt^2+d\stackrel{~}{x}_1^2+\mathrm{}+d\stackrel{~}{x}_{p2}^2+dx_{p1}^2+dx_p^2)`$ $`+\left({\displaystyle \frac{R}{u}}\right)^{(7p)/2}(\stackrel{~}{f}^1du^2+u^2d\mathrm{\Omega }_{8p}^2)],`$ $`e^{2\varphi }`$ $`=`$ $`{\displaystyle \frac{(2\pi )^4\stackrel{~}{g}^2\stackrel{~}{b}^4}{\stackrel{~}{V}_2^2}}\left({\displaystyle \frac{u}{R}}\right)^{(7p)(p3)/2},\stackrel{~}{B}_{p1,p}=0.`$ (2.24) This solution describes a ($`p+1`$)-dimensional ordinary super Yang-Mills theory with gauge group $`U(N_{p2})`$ on the dual torus with area $`\widehat{V}_2=(2\pi )^4\stackrel{~}{b}^2/\stackrel{~}{V}_2`$. Since the dual torus is characterized by the periodicity $`x_{p1,p}x_{p1,p}+\sqrt{\widehat{V}_2}`$, the radii of the dual torus go to zero for $`\stackrel{~}{V}_2\mathrm{}`$ and the ($`p+1`$)-dimensional ordinary super Yang-Mills theory then reduces to a ($`p1`$)-dimensional one. This means that if the torus in (2.23) is very large ($`\stackrel{~}{V}_2\mathrm{}`$), the solution is a dual gravity description of a ($`p1)`$-dimensional ordinary super Yang-Mills theory with gauge group $`U(\mathrm{})`$. The ($`p1`$)-dimensional theory has the Yang-Mills coupling constant $$g_{\mathrm{YM}}^2=(2\pi )^{p4}\stackrel{~}{g},$$ (2.25) while the coupling constant of the $`(p+1)`$-dimensional noncommutative gauge field is $`g_{\mathrm{YM}}^2=(2\pi )^{p2}\stackrel{~}{g}\stackrel{~}{b}`$. Thus we reach the equivalence argued by Lu and Roy between the noncommutative super Yang-Mills theory in ($`p+1)`$ dimensions and the ordinary one with gauge group $`U(\mathrm{})`$ in ($`p1`$) dimensions. This equivalence can also be understood from a T-duality of the decoupling limit solution (2.14) for the D$`p`$-branes with $`B`$ field. A usual T-duality transformation is, however, not enough for this purpose since the resulting radius for the torus is not large after the usual T-duality in the presence of $`B`$ field. This can be remedied if we use more general T-duality transformation $`SL(2,𝐙)`$ as described in refs. . The duality transformation $$\rho \frac{a\rho +b}{c\rho +d},\rho \frac{\stackrel{~}{V}_2}{(2\pi )^2\alpha ^{}}\left(B_{p1,p}+i\sqrt{G_{(p1)(p1)}G_{pp}}\right),$$ (2.26) gives a dual solution $`ds^2=\alpha ^{}[\left({\displaystyle \frac{u}{R}}\right)^{(7p)/2}(\stackrel{~}{f}dt^2+d\stackrel{~}{x}_1^2+\mathrm{}+d\stackrel{~}{x}_{p2}^2+dx_{p1}^2+dx_p^2)`$ $`+\left({\displaystyle \frac{R}{u}}\right)^{(7p)/2}(\stackrel{~}{f}^1du^2+u^2d\mathrm{\Omega }_{8p}^2)],`$ $`e^{2\varphi }={\displaystyle \frac{(2\pi )^4\stackrel{~}{g}^2\stackrel{~}{b}^4}{\stackrel{~}{V}_2^2}}\left({\displaystyle \frac{u}{R}}\right)^{(7p)(p3)/2},\stackrel{~}{B}_{p1,p}={\displaystyle \frac{\alpha ^{}}{\stackrel{~}{b}}},`$ (2.27) by choosing $`c=1`$ and $`d=\stackrel{~}{V}_2/(2\pi )^2\stackrel{~}{b}`$ when the latter is an integer. Note that $`d=N_{p2}/N_p`$ must be a rational number. If this is not an integer, after some steps of Morita equivalence transformation following , one can reach a solution like (2.27). For $`p=3`$ and $`\stackrel{~}{f}=1`$, the solution (2.27) reduces to the case discussed in . Note that the solution (2.27) is the same as (2.24) except that the former has a nonvanishing constant $`B`$ field while the latter has zero $`B`$ field. The solution (2.27) describes a twisted ordinary super Yang-Mills theory with gauge group $`U(N_{p2})`$ in $`(p+1)`$ dimensions, living on the dual torus with area $`\widehat{V}_2=(2\pi )^4\stackrel{~}{b}^2/\stackrel{~}{V}_2`$. For $`\stackrel{~}{V}_2\mathrm{}`$, however, the theory reduces to a ($`p1`$)-dimensional ordinary super Yang-Mills theory. Therefore we again arrive at the conclusion that the ($`p+1`$)-dimensional noncommutative gauge field is equivalent to an ordinary gauge field with gauge group $`U(\mathrm{})`$ in ($`p1`$) dimensions for $`\stackrel{~}{V}_2\mathrm{}`$, from the point of view of dual gravity description. Next let us address another evidence to render support of the above equivalence from the viewpoint of thermodynamics. We find that the thermodynamics of decoupling limit solution (2.23) of the D$`(p2)`$-branes is completely the same as those in (2.17). The worldvolume theory of the (D$`(p2)`$, D$`p`$) bound states (2.1) (or $`N_p`$ coinciding D$`p`$-branes with $`B`$ field) is a noncommutative gauge field theory with gauge group $`U(N_p)`$ in $`(p+1)`$ dimensions with two dimensions compactified on a torus, while the worldvolume theory is an ordinary gauge field theory with the same gauge group $`U(N_p)`$ if the NS $`B`$ field is absent. The above equivalence of the descriptions of the (D$`(p2)`$, D$`p`$) bound states implies that the bound states can also be described by ordinary gauge field theories in $`(p+1)`$ dimensions. Moreover, the equivalence is valid also between the $`(p+1)`$-dimensional noncommutative $`U(N_p)`$ theory and the $`(p1)`$-dimensional $`U(\mathrm{})`$ ordinary theory in the limit $`\stackrel{~}{V}_2\mathrm{}`$ for the reason described above. For finite volume, the equivalence is between the $`(p+1)`$-dimensional noncommutative $`U(N_p)`$ gauge field and a (twisted) ordinary $`U(N_{p2})`$ gauge field with the relation (2.21), the latter living on a dual torus. In this case, the Yang-Mills coupling constant is $$g_{\mathrm{YM}}^2=\frac{(2\pi )^p\stackrel{~}{g}\stackrel{~}{b}^2}{\stackrel{~}{V}_2},$$ (2.28) for the $`(p+1)`$-dimensional ordinary gauge field theory. As a self-consistency check, one may find that the coupling constant (2.25) can also be obtained from (2.28) after a trivial dimensional reduction. In the following sections we will further discuss the equivalence of the descriptions and the relation (2.21) from the point of view of probe branes. ## 3 The static probes: thermodynamics The D-brane probe is a useful tool to explore the structure of D-brane bound states (see for example - and references therein). Recently the D-brane probes have been used to check a certain aspect of AdS/CFT correspondence -. In this section we consider the static interaction potentials (thermodynamics) of two kinds of probes in the background of (D$`(p2)`$, D$`p`$) bound states. One of them is a bound state probe consisting of D$`(p2)`$\- and D$`p`$-branes, or D$`p`$-brane probe with $`B`$ fields, or noncommutative D$`p`$-brane probe; the other is a D$`(p2)`$-brane probe. Let us first discuss the noncommutative D$`p`$-brane probe. ### 3.1 A noncommutative D$`p`$-brane probe Due to the presence of the nonvanishing NS $`B`$ field in the noncommutative D$`p`$-brane probe, the probe should have the following action: $$S_p=T_pd^{p+1}xe^\varphi \sqrt{det(G_{ab}+_{ab})}+T_pA^p+T_pA^{p2},$$ (3.1) where $`_{ab}=(2\pi \alpha ^{})F_{ab}+B_{ab}`$, $`F_{ab}`$ is the gauge field strength on the worldvolume of the D$`p`$-brane and $`B_{ab}`$ is the NS $`B`$ field. Here we set $`F_{ab}=0`$. As demonstrated in (2.1), the occurrence of the NS $`B`$ field in the D$`p`$-branes is always accompanied by the appearance of D$`(p2)`$-branes in the system, forming (D$`(p2)`$, D$`p`$) bound states. The probe can be regarded as a bound state probe consisting of D$`(p2)`$\- and D$`p`$-branes for the following reasons: (1) The probe has the tension $`T_p\sqrt{1+\mathrm{tan}^2\theta }`$, the same as the (D$`(p2)`$, D$`p`$) bound states; (2) We will see shortly that the static interaction potential of the probe vanishes in the background of the nonthreshold (D$`(p2)`$, D$`p`$) bound states; (3) Except for the source $`A^p`$ of the D$`p`$-branes, the source $`A^{p2}`$ of the D$`(p2)`$-branes also occurs in the action (3.1); (4) From its thermodynamics we will obtain further evidence of this interpretation. Because of the presence of the NS $`B`$ field, we can also view the probe as a noncommutative D$`p`$-brane probe. Substituting the solution (2.1) into the probe action (3.1), one has $$S_p=\frac{T_pV_p}{g\mathrm{cos}\theta }𝑑\tau H^1[\sqrt{f}1+H_0H],$$ (3.2) where we have subtracted a constant potential at spatial infinity and $$H_0=1+\left(\frac{\stackrel{~}{R}}{r}\right)^{7p}=1+\frac{r_0^{7p}\mathrm{sinh}\alpha \mathrm{cosh}\alpha }{r^{7p}}.$$ (3.3) In the extremal limit ($`f=1`$), the static interaction potential vanishes, which verifies that the probe is a bound state of D$`(p2)`$-branes and D$`p`$-branes because the source is nonthreshold bound states of D$`(p2)`$\- and D$`p`$-branes. Unless the probe is such a kind of bound state, the static potential will no longer vanish. In the non-extremal background, the interaction potential always exists. Now suppose the probe is moved from spatial infinity to the horizon of the source . From (3.2) we can obtain the potential difference (which is just the potential at the horizon because we have set the potential zero at spatial infinity) $$U_p|_{r=r_0}=\frac{T_pV_p}{g\mathrm{cos}\theta }\left(1\mathrm{tanh}\alpha \right).$$ (3.4) Since the tension of the probe is $`T_p\sqrt{1+\mathrm{tan}^2\theta }`$, it is easy to show that the first term is just the mass of the probe because $$m_p=\frac{T_pV_p}{g}\sqrt{1+\mathrm{tan}^2\theta }=\frac{T_pV_p}{g\mathrm{cos}\theta }.$$ (3.5) The second term in (3.4) has the following interpretation. Let us denote the numbers of D$`p`$-branes and D$`(p2)`$-branes in the probe by $`\delta N_p`$ and $`\delta N_{p2}`$, respectively. We then have $`\mu _pV_pT_p\delta N_p`$ $`+`$ $`\mu _{p2}V_{p2}T_{p2}\delta N_{p2}`$ (3.6) $`={\displaystyle \frac{T_pV_p}{g\mathrm{cos}\theta }}\left[\mathrm{cos}^2\theta +\mathrm{sin}\theta \mathrm{cos}\theta {\displaystyle \frac{V_{p2}T_{p2}}{V_pT_p}}{\displaystyle \frac{\delta N_{p2}}{\delta N_p}}\right]\delta N_p,`$ $`={\displaystyle \frac{T_pV_p}{g\mathrm{cos}\theta }}\mathrm{tanh}\alpha \delta N_p,`$ where in obtaining the third line we have used the fact that the form of the tension of the probe implies that the number of the branes obey $`\delta N_{p2}/\delta N_p=\mathrm{tan}\theta V_pT_p/(V_{p2}T_{p2})`$. When $`\delta N_p=1`$, this quantity (3.6) gives the second term in (3.4). This process satisfies the first law of thermodynamics (2.10). In fact eq. (3.4) reduces to (2.10) with $`dM=m_p`$, $`U_p|_{r=r_0}=TdS`$ and (3.6). It follows that the potential of the probe is converted into heat energy and is absorbed by the source when the probe moves to the horizon from spatial infinity. Our calculation also shows that the probe is a bound state of D$`(p2)`$\- and D$`p`$-branes. Now we consider the decoupling limit of the static probe action. In this limit, we obtain $$S_p=\frac{V_{p2}\stackrel{~}{V}_2}{(2\pi )^p\stackrel{~}{g}\stackrel{~}{b}}𝑑\tau \left(\frac{u}{R}\right)^{7p}\left[\sqrt{\stackrel{~}{f}}1+\frac{u_0^{7p}}{2u^{7p}}\right].$$ (3.7) From the action we can also obtain the free energy of the probe at the temperature $`T`$, which is just the Hawking temperature of the source given in (2.17): $$F_p=\frac{V_{p2}\stackrel{~}{V}_2}{(2\pi )^p\stackrel{~}{g}\stackrel{~}{b}}\left(\frac{u}{R}\right)^{7p}\left[\sqrt{\stackrel{~}{f}}1+\frac{u_0^{7p}}{2u^{7p}}\right].$$ (3.8) In the generalized AdS/CFT correspondence, the thermodynamics (2.17) is equivalent to that of noncommutative supersymmetric gauge fields with gauge group $`U(N_p)`$ in the large $`N`$ and strong coupling limit (within the valid regime of dual gravity description). From the viewpoint of field theory, the thermodynamics corresponds to that of the gauge field in the Higgs branch, where the gauge group is not broken and hence the vacuum expectation values of scalars vanish. According to the interpretation of a D-brane probe action , the thermodynamics of a D-brane probe can be regarded as the thermodynamics of the supersymmetric gauge field in the Coulomb branch (Higgs phase) , in which the original gauge group is broken; some vacuum expectation values of scalar fields do not vanish; and the distance $`u`$ between the probe and the source can be viewed as a energy scale in the gauge fields. This interpretation of thermodynamics of D-brane probe turns out to be consistent with the expectation on the field theory side . Since the rescaled string coupling is $`\widehat{g}=\stackrel{~}{g}\stackrel{~}{b}`$, we find from (3.8) that in the decoupling limit, the free energy of a noncommutative D$`p`$-brane probe in the noncommutative D$`p`$-brane background ((D$`(p2)`$, D$`p`$) bound states) is exactly the same as that of an ordinary D$`p`$-brane probe in the D$`p`$-brane background without NS $`B`$ field (for the interaction potential of the latter see ). Thus, in the supergravity approximation, the thermodynamics of noncommutative gauge fields remains the same as the ordinary case both in the Higgs and Coulomb branches. We thus conclude that in the large $`N`$ limit, the number of the degrees of freedom of noncommutative gauge fields coincides with the ordinary case, not only in the weak coupling limit, but also in the strong coupling limit. Now we consider the difference between the interaction potentials (free energy) of the probe at the infinity and at the horizon in the decoupling limit, and compare it with the asymptotically flat case already discussed before. Note that the free energy still vanishes at the infinity ($`u\mathrm{}`$) as can be seen in (3.8). Thus the difference in the free energies is just the free energy of the probe at the horizon $`u_0`$: $`F_p|_{u=u_0}`$ $`=`$ $`{\displaystyle \frac{V_{p2}\stackrel{~}{V}_2}{2(2\pi )^p\stackrel{~}{g}\stackrel{~}{b}}}\left({\displaystyle \frac{u_0}{R}}\right)^{7p}`$ (3.9) $`=`$ $`{\displaystyle \frac{V_{p2}\stackrel{~}{V}_2}{2(2\pi )^p\stackrel{~}{g}\stackrel{~}{b}}}\left({\displaystyle \frac{4\pi RT}{7p}}\right)^{\frac{2(7p)}{5p}}.`$ We find that the free energy of the probe at the horizon (3.9) has the relation with that of the source (2.18) as $$F_p|_{u=u_0}=\frac{dF}{dN_p}\delta N_p,$$ (3.10) with $`\delta N_p=1`$. Note that we are considering a D$`p`$-brane with $`B`$ field in the background of $`N_p`$ D$`p`$-branes. Therefore in the large $`N_p`$ limit (that is $`N_p>>1`$), the probe free energy is expected to be $$F_p|_{u=u_0}F(N_p+1)F(N_p),$$ (3.11) consistent with (3.10). This relation supports the argument that the non-extremal D$`p`$-branes is located at the horizon. (In the next section we will further show that indeed D$`p`$-branes can be located at the horizon). This also supports the interpretation of thermodynamics of probe branes given in , because the probe brane at the horizon of the source can be considered to coincide with source branes and the gauge symmetry is restored and the probe brane can be seen as a part of the source branes in this case. ### 3.2 A D$`(p2)`$-brane probe Let us next consider a D$`(p2)`$-brane probe. The action of a D$`(p2)`$-brane in the background of the (D$`(p2)`$, D$`p`$) bound states (2.1) is $$S_{p2}=T_{p2}d^{p1}xe^\varphi \sqrt{detG_{ab}}+T_{p2}A^{p2}.$$ (3.12) Substituting the background solution (2.1) into the action yields, for a static probe, $$S_{p2}=\frac{T_{p2}V_{p2}}{g}𝑑\tau H^1\left[H^{1/2}h^{1/2}\sqrt{f}(1H_0)\mathrm{sin}\theta H\right],$$ (3.13) where we have also subtracted a constant potential so that the interaction potential vanishes at spatial infinity. Note that the static interaction potential is quite different from that of the noncommutative D$`p`$-brane probe in the same background (2.1). Indeed the potential (3.13) does not vanish even when the background is sent to the extremal limit of the solution. This is consistent with the fact that the source is a nonthreshold bound state consisting of D$`(p2)`$\- and D$`p`$-branes. As in the noncommutative D$`p`$-brane probe, let us first consider the asymptotically flat background. In this case, we find that the deference between the potentials at the spatial infinity and at the horizon is $$U_{p2}|_{r=r_0}=\frac{V_{p2}T_{p2}}{g}\left(1\mathrm{sin}\theta \mathrm{tanh}\alpha \right).$$ (3.14) The first term is the mass of the probe $`m_{p2}=T_{p2}V_{p2}/g`$, while the second term is equal to $`\mu _{p2}V_{p2}T_{p2}\delta N_{p2}`$ with $`\delta N_{p2}=1`$ since we are considering a probe D$`(p2)`$-brane. Therefore the D$`(p2)`$-brane probe falling to the horizon from the spatial infinity satisfies the first law (2.10) again with $`dM=m_{p2}`$, $`\delta N_{p2}=1`$, and $`\delta N_p=0`$. The potential difference of the D$`(p2)`$-brane probe is converted into heat energy at the horizon and thereby is absorbed by the source. Comparing (3.2) and (3.13), a priori one may think that they are quite different and there seems to be no relation between them. Actually once the decoupling limit (2.13) is taken, one may find that there is a close relation between (3.2) and (3.13). In the decoupling limit, the action of the D$`(p2)`$-brane probe becomes $$S_{p2}=\frac{V_{p2}}{(2\pi )^{p2}\stackrel{~}{g}}𝑑\tau \left(\frac{u}{R}\right)^{7p}\left[\sqrt{\frac{1+(au)^{7p}}{(au)^{7p}}}\sqrt{\stackrel{~}{f}}1+\frac{u_0^{7p}}{2u^{7p}}\right].$$ (3.15) As mentioned above, the validity of the dual gravity description of gauge field theories requires $`au>>1`$. The above action then reduces to $$S_{p2}=\frac{V_{p2}}{(2\pi )^{p2}\stackrel{~}{g}}𝑑\tau \left(\frac{u}{R}\right)^{7p}\left[\sqrt{\stackrel{~}{f}}1+\frac{u_0^{7p}}{2u^{7p}}\right].$$ (3.16) The corresponding free energy of the probe at the distance $`u`$ is $$F_{p2}=\frac{V_{p2}}{(2\pi )^{p2}\stackrel{~}{g}}\left(\frac{u}{R}\right)^{7p}\left[\sqrt{\stackrel{~}{f}}1+\frac{u_0^{7p}}{2u^{7p}}\right].$$ (3.17) At the horizon $`u_0`$ it is $`F_{p2}|_{u=u_0}`$ $`=`$ $`{\displaystyle \frac{V_{p2}}{2(2\pi )^{p2}\stackrel{~}{g}}}\left({\displaystyle \frac{u_0}{R}}\right)^{7p}`$ (3.18) $`=`$ $`{\displaystyle \frac{V_{p2}}{2(2\pi )^{p2}\stackrel{~}{g}}}\left({\displaystyle \frac{4\pi RT}{7p}}\right)^{\frac{2(7p)}{5p}}.`$ Comparing the free energy (3.18) with the one (3.9) of the noncommutative D$`p`$-brane probe, one may find that they are the same up to a different prefactor. Consequently the free energy of $`\delta N_{p2}`$ D$`(p2)`$-branes is the same as that of $`\delta N_p`$ noncommutative D$`p`$-branes if the relation $$\frac{\delta N_{p2}}{\delta N_p}=\frac{\stackrel{~}{V}_2}{(2\pi )^2\stackrel{~}{b}},$$ (3.19) is obeyed. We see that this relation coincides with eq. (2.21). Note that the relation (2.21) is derived from the two equivalent descriptions of the bound state source, while (3.19) is obtained from the equivalence of probes in the same background. In other words, the probe consisting of $`\delta N_p`$ noncommutative D$`p`$-branes is equivalent to the probe consisting of $`\delta N_{p2}`$ D$`(p2)`$-branes since they get the same response in the same background. Furthermore, we find $$F_{p2}|_{u=u_0}=\frac{dF}{dN_{p2}}\delta N_{p2},$$ (3.20) with $`\delta N_{p2}=1`$. When $`N_{p2}>>1`$, once again, we have $$F_{p2}|_{u=u_0}F(N_{p2}+1)F(N_{p2}).$$ (3.21) This implies that from the point of view of the D$`(p2)`$-brane probe, the bound state source (D$`(p2)`$, D$`p`$) can be viewed as $`N_{p2}`$ coincident D$`(p2)`$-branes with two smeared coordinates and zero NS $`B`$ field, while from the noncommutative D$`p`$-brane probe, the source is $`N_p`$ coincident D$`p`$-branes with nonzero NS $`B`$ field. Therefore from the thermodynamics of probe branes, we again find that the bound states (D$`(p2)`$, D$`p`$) have two equivalent descriptions. ## 4 The dynamical probes: absorbing or scattering In this section we will consider the dynamical aspect of the two kinds of probes discussed in the previous section. ### 4.1 The noncommutative D$`p`$-brane probe To investigate the dynamics of the probe, it is convenient to take static gauge: $`\tau =t`$, $`x_i`$ act just as the worldvolume coordinates and other transverse coordinates depend only on $`\tau `$. In the background (2.1), the action (3.1) of the noncommutative D$`p`$-brane probe reduces to $$S_p=\frac{T_pV_p}{g\mathrm{cos}\theta }𝑑\tau H^1[\sqrt{fH(f^1\dot{r}^2+r^2\dot{\mathrm{\Omega }}_{8p}^2)}1+H_0H],$$ (4.1) where an overdot denotes the derivative with respect to $`\tau `$. In the decoupling limit, the action becomes $`S_p=m_p{\displaystyle 𝑑\tau \left(\frac{u}{R}\right)^{7p}\left[\sqrt{\stackrel{~}{f}\left(\frac{R}{u}\right)^{7p}\left(\stackrel{~}{f}^1\dot{u}^2+u^2\dot{\mathrm{\Omega }}_{8p}^2\right)}1+\frac{u_0^{7p}}{2u^{7p}}\right]},`$ (4.2) where $`m_p=V_{p2}\stackrel{~}{V}_2/[(2\pi )^p\stackrel{~}{g}\stackrel{~}{b}]`$ is the mass of the probe. With the definition of the isotropic coordinates $$\stackrel{~}{f}^1du^2+u^2d\mathrm{\Omega }_{8p}^2=u^2\rho ^2(d\rho ^2+\rho ^2d\mathrm{\Omega }_{8p}^2),$$ (4.3) where $$u^{7p}=\rho ^{7p}\left(1+\frac{u_0^{7p}}{4\rho ^{7p}}\right)^2,$$ (4.4) we can define the velocity of the probe as $$\stackrel{~}{f}^1\dot{u}^2+u^2\dot{\mathrm{\Omega }}_{8p}^2u^2\rho ^2v^2.$$ (4.5) In the low velocity and long distance approximation, expanding (4.2) yields $$S_p=𝑑\tau [\frac{1}{2}m_pv^2𝒱(\rho ,v)+𝒪(1/\rho ^{2(7p)})],$$ (4.6) where the interaction potential $`𝒱`$ is $$𝒱(\rho ,v)=m_p\frac{u_0^{7p}}{\rho ^{7p}}\left\{\frac{9p}{4(7p)}v^2+\frac{1}{8}\left[\left(\frac{u_0}{R}\right)^{7p}+\left(\frac{R}{u_0}\right)^{7p}v^4\right]\right\}$$ (4.7) When $`\theta =0`$, the background (2.1) reduces to the black D$`p`$-brane without $`B`$ and the probe action (4.1) to the one for a D$`p`$-brane without $`B`$ field. The interaction potential (4.7) therefore is also the one for a D$`p`$-brane probe in other D$`p`$-brane background. Up to order $`v^4`$, it can be seen that (4.7) reduces to the result in for a D$`p`$-brane in other D$`p`$-brane background. This interaction potential can be reproduced in the one-loop calculation of the noncommutative field theory. Using the relation between the phase shift of scattering and the potential $$\delta (\rho ,v)=_0^{\mathrm{}}𝑑\tau 𝒱[\rho (\tau ),v],\rho ^2(\tau )=\rho ^2+v^2\tau ^2,$$ (4.8) we obtain the phase shift of the probe. Writing $`𝒱(\rho ,v)=\lambda (v)\rho ^{(7p)}`$, we find $$\delta (\rho ,v)=\frac{B(\frac{1}{2},\frac{6p}{2})}{2v\rho ^{6p}}\lambda (v),$$ (4.9) where $`B`$ is the beta function. Next we discuss the classical motion of the noncommutative D$`p`$-brane probe near the horizon of the source (that is, in the decoupling limit). For simplicity, let us consider the case of the probe with angular momentum only in a single direction (say, $`\varphi `$-direction). Following refs. , from (4.2) we obtain its angular momentum $$L=\frac{m_pu^2\dot{\varphi }}{\sqrt{\stackrel{~}{f}\left(\frac{R}{u}\right)^{7p}\left(\stackrel{~}{f}^1\dot{u}^2+u^2\dot{\varphi }^2\right)}}.$$ (4.10) The energy of the probe is $$E=\frac{m_p\left(\frac{u}{R}\right)^{7p}\stackrel{~}{f}}{\sqrt{\stackrel{~}{f}\left(\frac{R}{u}\right)^{7p}\left(\stackrel{~}{f}^1\dot{u}^2+u^2\dot{\varphi }^2\right)}}m_p\left(\frac{u}{R}\right)^{7p}\left(1\frac{u_0^{7p}}{2u^{7p}}\right).$$ (4.11) With the relation $$E=\frac{1}{2}m_p\dot{u}^2+V(u),$$ (4.12) one may obtain an effective central potential of the radial motion of the probe $$V(u)=E\left[1\frac{m_p\stackrel{~}{f}^2}{2E}\left(\frac{u}{R}\right)^{7p}\left(1\frac{\stackrel{~}{f}}{𝒜^2}\right)\right]+\frac{L^2\stackrel{~}{f}^3}{2m_pu^2𝒜^2}.$$ (4.13) where $$𝒜=1\frac{u_0^{7p}}{2u^{7p}}+\frac{E}{m_p}\left(\frac{R}{u}\right)^{7p}.$$ (4.14) The qualitative features of the motion of the probe can be understood by finding out the turning points, at which $`\dot{u}=0`$. Let us first discuss the case of extremal background (or in the background of the nonthreshold bound state (D$`(p2)`$, D$`p`$). One has $`\stackrel{~}{f}=1`$. From the effective central potential one can see that if the angular momentum vanishes, there is no turning point for the probe. It follows that the probe will be captured by the source. When the angular momentum does not vanish, the potential (4.13) reduces to $$V(u)=E\left\{1\frac{1}{2}\left(\frac{u}{u_{}}\right)^{7p}\left[1\frac{1}{\left(1+(u_{}/u)^{7p}\right)^2}\right]\right\}+\frac{Eu_{}^2}{2u^2\left(1+(u_{}/u)^{7p}\right)^2},$$ (4.15) where we have introduced two characterizing lengths $$u_{}=R\left(\frac{E}{m_p}\right)^{1/(7p)},u_{}=L\left(\frac{1}{m_pE}\right)^{1/2}.$$ (4.16) The turning point satisfies the following equation: $$2+\left(\frac{u_{}}{u_c}\right)^{7p}=\left(\frac{u_{}}{u_c}\right)^2.$$ (4.17) If $`u_{}/u>>1`$, namely, very near the source branes, the turning point is $$u_c=\left(\frac{u_{}^{7p}}{u_{}^2}\right)^{1/(5p)}.$$ (4.18) In the non-extremal background, there may exist some points satisfying the turning-point condition $`\dot{u}=0`$. From the effective potential (4.13) we find that the horizon, where $`\stackrel{~}{f}=0`$, must be one of those points, regardless of the angular momentum. In particular, we notice that the central force exerted on the probe, defined as $`F(u)=dV(u)/du`$, vanishes at the horizon. It means that once the probe reaches the horizon, it can stay at the horizon since the horizon is the turning point and the central force is zero there. In the previous section we have shown that the non-extremal branes are “located” at the horizon from the point of view of thermodynamics of a probe brane. Here we provide another evidence to support the argument from the dynamical aspect of a probe brane. ### 4.2 The D$`(p2)`$-brane probe In this subsection we consider the dynamics of a D$`(p2)`$-brane in the background of (D$`(p2)`$, D$`p`$) bound states in the decoupling limit. In this case, the worldvolume is $`(t,x_1,\mathrm{},x_{p2})`$, in the static gauge, one has the action of the probe $`S_{p2}`$ $`=`$ $`{\displaystyle \frac{T_{p2}V_{p2}}{g}}{\displaystyle }d\tau H^1[(Hh^1)^{1/2}\sqrt{fh(\dot{x}_{p1}^2+\dot{x}_p^2)H(f^1\dot{r}^2+r^2\dot{\mathrm{\Omega }}_{8p}^2)}`$ (4.19) $`(1H_0)\mathrm{sin}\theta H],`$ In the decoupling limit, it reduces to $`S_{p2}`$ $`=`$ $`m_{p2}{\displaystyle }d\tau \left({\displaystyle \frac{u}{R}}\right)^{7p}[\sqrt{{\displaystyle \frac{1}{(au)^{7p}\stackrel{~}{h}}}}\sqrt{\stackrel{~}{f}\stackrel{~}{h}\left(\dot{\stackrel{~}{x}}_{p1}^2+\dot{\stackrel{~}{x}}_p^2\right)\left({\displaystyle \frac{R}{u}}\right)^{7p}\left(\stackrel{~}{f}^1\dot{u}^2+u^2\dot{\mathrm{\Omega }}_{8p}^2\right)}`$ (4.20) $`1+{\displaystyle \frac{u_0^{7p}}{2u^{7p}}}].`$ Here $`m_{p2}=V_{p2}/[(2\pi )^{p2}\stackrel{~}{g}]`$ is the mass of the probe. When $`au>>1`$, this action approximates to $`S_{p2}`$ $`=`$ $`m_{p2}{\displaystyle }d\tau \left({\displaystyle \frac{u}{R}}\right)^{7p}[\sqrt{\stackrel{~}{f}{\displaystyle \frac{1}{(au)^{7p}}}\left(\dot{\stackrel{~}{x}}_{p1}^2+\dot{\stackrel{~}{x}}_p^2\right)\left({\displaystyle \frac{R}{u}}\right)^{7p}\left(\stackrel{~}{f}^1\dot{u}^2+u^2\dot{\mathrm{\Omega }}_{8p}^2\right)}`$ (4.21) $`1+{\displaystyle \frac{u_0^{7p}}{2u^{7p}}}].`$ In fact, this is also the action of a D$`(p2)`$-brane probe in the background produced by the source D$`(p2)`$-branes (2.23). In the extremal limit, up to $`𝒪(v^4)`$, its motion is a geodesic of the following moduli space: $$ds_m^2=du^2+u^2d\mathrm{\Omega }_{8p}^2+\frac{1}{\stackrel{~}{b}^2}\left(d\stackrel{~}{x}_{p1}^2+d\stackrel{~}{x}_p^2\right).$$ (4.22) Therefore, in this approximation, the D$`(p2)`$-brane probe moves as in a flat space. Namely, in the large noncommutative effect limit, the effect of the $`N_p`$ coincident D$`p`$-branes in the source on the motion of a D$`(p2)`$-brane probe vanishes. The source looks like the one consisting of only D$`(p2)`$-branes with two smeared coordinates and no $`B`$ field. In the large $`au>>1`$ limit, if one does not consider the motion of the probe along the relative transverse directions (that is, setting $`\dot{\stackrel{~}{x}}_{p1}=\dot{\stackrel{~}{x}}_p=0`$ in the the action (4.20)), the action of the probe then reduces to that (4.2) of a noncommutative D$`p`$-brane probe, except a difference in the mass of probes. If the mass of both probes is equal to each other, then they have the same interaction potential and the same phase shift depending on the distance and the velocity. Because we are considering only a single D-brane probe, to match the mass of probes, we should have $`\delta N_pm_p=\delta N_{p2}m_{p2}`$, from which one gets the relation (3.19) again. Thus, from the point of view of the dynamics of probes, we see again the equivalence between $`\delta N_p`$ noncommutative D$`p`$-branes and $`\delta N_{p2}`$ D$`(p2)`$-branes with two smeared coordinates and no $`B`$ field. Without the motion along the relative transverse directions, similarly to the case of the D$`p`$-brane probe, we also obtain an effective central potential for the D$`(p2)`$-brane probe as $$V(u)=E\left\{1\frac{m_{p2}\stackrel{~}{f}^2}{2E}\left(\frac{u}{R}\right)^{7p}\left[1\frac{𝒞\stackrel{~}{f}}{^2}\right]\right\}+\frac{L^2\stackrel{~}{f}^3}{2m_{p2}u^2^2},$$ (4.23) where we have not taken the large $`au`$ limit and $$𝒞=\frac{1+(au)^{7p}}{(au)^{7p}},=1\frac{u_0^{7p}}{2u^{7p}}+\frac{E}{m_{p2}}\left(\frac{R}{u}\right)^{7p}.$$ (4.24) Due to the appearance of $`𝒞`$, the motion of the probe D$`(p2)`$-brane is a little different from that of the D$`p`$-brane probe. However, we find that the horizon of the background is still the turning point of the probe and the central force on the D$`(p2)`$-brane probe vanishes. This implies that the D$`(p2)`$-brane probe can also stay at the horizon. It is also consistent with the result from the analysis of thermodynamics of the probe. Indeed, as an ingredient of the bound state (D$`(p2)`$, D$`p`$), non-extremal D$`(p2)`$-branes should also be located at the horizon of the background. ## 5 Conclusions As is well known by now, the worldvolume coordinates of D$`p`$-branes will become noncommutative if a nonvanishing constant NS $`B`$ field is present on the worldvolume of the D$`p`$-branes. The worldvolume theory is then the super Yang-Mills theory in a noncommutative space (noncommutative gauge field theory). Each of the nonthreshold (D$`(p2)`$, D$`p`$) bound states ($`2p6`$) can be viewed as a D$`p`$-brane bound state with a nonvanishing NS $`B`$ field of rank two. In this paper we have investigated two equivalent descriptions of the nonthreshold (D$`(p2)`$, D$`p`$) bound states in the dual gravity description. In the decoupling limit, the bound states can be described as D$`p`$-branes with nonvanishing NS $`B`$ field, and then the worldvolume theory is a noncommutative gauge field with gauge group $`U(N_p)`$ in ($`p+1`$) dimensions (with two dimensions compactified on a torus in our case) if the number of the coincident D$`p`$-branes is $`N_p`$. On the other hand, the nonthreshold (D$`(p2)`$, D$`p`$) bound state will reduce to the solution of D$`(p2)`$-branes with two smeared coordinates and zero $`B`$ field in the decoupling limit and the large $`au>>1`$ limit. The latter condition is necessary for the validity of the dual gravity description. The worldvolume theory of the D$`(p2)`$-branes should be an ordinary gauge field theory with gauge group $`U(N_{p2})`$ in ($`p+1)`$ dimensions if the number of the coincident D$`(p2)`$-branes is $`N_{p2}`$. From the viewpoint of the thermodynamics of dual gravity solutions for the bound states (D$`(p2)`$, D$`p`$), we have found that $`N_p`$ coincident D$`p`$-branes with NS $`B`$ field is equivalent to $`N_{p2}`$ coincident D$`(p2)`$-branes with two smeared coordinates and no $`B`$ field. In the equivalence, $`N_p`$ and $`N_{p2}`$ must obey the relation (2.21), where $`\stackrel{~}{V}_2`$ is the area of the two additional dimensions and $`\stackrel{~}{b}`$ is a noncommutativity parameter. When the volume of the torus is sent to infinity keeping this relation, the ordinary super Yang-Mills theory reduces to the one with gauge group $`U(\mathrm{})`$ in $`(p1)`$ dimensions. We have identified the Yang-Mills coupling constant for the ($`p1)`$-dimensional ordinary Yang-Mills theory. We have also shown the equivalence from the thermodynamics and dynamics of two probes in the background of the bound states (D$`(p2)`$, D$`p`$). One of the probes is a bound state of D$`p`$\- and D$`(p2)`$-branes, which we called a noncommutative D$`p`$-brane probe. The other is a D$`(p2)`$-brane probe. In the asymptotically flat limit, when the two probes fall into the horizon of the source from spatial infinity, their static interaction potentials at the horizon are converted into heat and thereby are absorbed by the source. In this process, the first law of black hole thermodynamics is obeyed. In the decoupling limit, we have found that the thermodynamics and dynamics of the two probes are identical if the numbers of probe branes satisfy the relation (3.19), completely the same relation as (2.21). As a byproduct, we have found that the free energy of the noncommutative D$`p`$-brane probe in the D$`p`$-brane background with a nonvanishing NS $`B`$ field is the same as that of a D$`p`$-brane probe in the D$`p`$-brane background without $`B`$ field. It shows that the thermodynamics of the noncommutative super Yang-Mills coincides with the ordinary case in the large $`N`$ limit, not only in the Higgs branch, but also in the Coulomb branch. In addition, from the analysis of dynamics of probes, we have derived that the non-extremal D$`p`$-branes can be located at the horizon. Our discussions support the argument by Lu and Roy that there is an equivalence between the noncommutative super Yang-Mills with gauge group $`U(N_p)`$ in ($`p+1)`$ dimensions with two dimensions compactified on a torus and the ordinary one with gauge group $`U(N_{p2})`$ in ($`p1`$) dimensions when the area of the torus $`\stackrel{~}{V}_2\mathrm{}`$, with the relation (2.21) between $`N_p`$ and $`N_{p2}`$. This result is also consistent with the Morita equivalence . ## Acknowledgments We would like to thank J.X. Lu and E. Kiritsis for helpful correspondences. This work was supported in part by the Japan Society for the Promotion of Science and by grant-in-aid from the Ministry of Education, Science, Sports and Culture No. 99020.
warning/0001/nlin0001060.html
ar5iv
text
# Rate increase in chemical reaction and its variance under turbulent equilibrium ## 1 Introduction — A perspective on the classical and turbulence- corrected Boltzmann formalism It is well-known that each of the transport processes in fluids, namely, diffusion, heat transfer, momentum transfer due to shearing motion, and chemical reaction is enhanced by the presense of turbulence. The physical origin that causes such appreciable hike in those rates is the fractal nature of turbulence: it is ascribed to drastic increase in net contact area of adjacent bulk of fluid, taking place through the bounding surface penetrating into each other area, thereby making it easier to transport mass, momentum and energy. It results in drag rise of a flat plate several times the laminar counterpart, also in flame propagation velocity elevated by a few ten times under internal combustion engine environment(Fig.1). Obviously the latter is a favorable facet of turbulence characteristies, whereas the former is detrimental from the viewpoint of transport vehicle technologies. In fact, the evidence that the rate increase due to turbulence is much higher for mechanism generating thrust rather than drag tempts us to imagine a midget planet where airplanes would have to install disproportionally huge engines, under the environment of no turbulence. Thus chemical reaction rates, among other transport processes, must be influenced by turbulence most sensitively, so is worth looking into most detail. In general there are two aspects of approach to this problem: The one is to pursue the instantaneous value of a physical quantity $`\underset{¯}{W}`$ which is stochastic and fractal, using direct numerical simulation (DNS) for example. In principle, any of these quantities are described in terms of microscopic density advocated by Klimontovich, which we will call hereafter the K-formalism. The DNS currently being used in numerical simulation of fluid flows is the macroscopic version of the K-formalism. It has been applied to problems of combustion science, enabling to compute wiggling premixed flame front with sufficient resolvability. The state of the art, however, is yet to go far to be able to provide information such as turbulent flame speed for engine designers. Another approach to turbulent combustion is based on an axiom of statistical mechanics that claims equivalence of solving for instantaneous value $`\underset{¯}{W}`$ of a stochastic quantity and for a set of averaged quantities $$\{W,\overline{W^{}\widehat{W}^{}},\overline{W^{}\widehat{W}^{}\stackrel{~}{W}^{}},\mathrm{}\}$$ (1) In the above $`W=\overline{\underset{¯}{W}}`$ is the average value, $$W^{}=\underset{¯}{W}W$$ (2) is the instantaneous fluctuation subject to $`\overline{W^{}}=0`$, and fluctuation correlations $`\overline{W^{}\widehat{W}^{}}`$ etc., refer to those between different points in the independent variable $`z,\mathrm{e}.\mathrm{g}.,\mathrm{W}=\mathrm{W}(\mathrm{z}),\widehat{\mathrm{W}}=\mathrm{W}(\widehat{\mathrm{z}})`$, etc. The second formalism is the basis on which to found statistical mechanics for example; in fact, if one identifies $`\underset{¯}{W}`$ with the microscopic density, then its average $`W`$ represents Boltzmann’s function. So the latter may well be called B-formalism as contrast to the K-formalism as defined above. A flow of turbulent gas is dually stochastic; the one is microscopic (molecular) level with the average taken with respect to the molecular velocity, whereas the other is macroscopic (fluid-dynamic) level where the average is taken over a volumelet whose size is greater than the Kolmogorov scale yet smaller than fluid-dynamic characteristic length. If one identifies $`W`$ with chemical reaction rate, there are three levels of descriptions: Following the K-formalism $`\underset{¯}{W}`$ is expressed as a sum of on-off step functions depending on whether each molecular collision clears the potential barrier to effect inelastic collisions. Next level is the classical B-formalism where the molecular average alone is employed. Then $`W`$ is shown to be expressed as Arrhenius’ law. It is coupled with fractal analysis to estimate turbulent flame surface area to evaluate turbulent burning rate. The third stage, namely, the B-formalism to be discussed here is intended to encompass both molecular and turbulent averaging procedures in its framework. A series of earlier works along this line has shed some light on the conflicts between the classical Arrhenius kinetics and experimental data attributable to temperature turbulence using the exact solution that represents simplified reality. In Fig.2 is shown shock-tube data of low-temperature (spotty) ignition of hydrogen oxygen premixture as compared with the classical theory ($`\delta T=0`$), also with turbulence-corrected one, a primitive form of the present theory. Also is shown in Fig.3 flame velocity of turbulent premixed gas as compared with existing experiment and with renormalization group theory. Although the earlier theory has thus enabled to predict some features of turbulent combustion as to their dependence on turbulent intensity without relying on any empirical parameters, no information has been provided regarding their dependence on the scale of turbulence. Revised version in the basic formalism has been proposed in ref. , with applications to nonreactive turbulence, where full informations on the scale of turbulence have been made available. The objective of this paper is three-fold: The one is to show the exact solution for turbulent equilibrium still valid even without the condition of only temperature turbulence prevalent. The other is to remodel the previous theory so as to include scale of turbulence into formalism for reactive flows now at issue. The third is to show that the B-formalism, an average-oriented formalism, provides a sound description of turbulent combustion if based on turbulent equilibrium, which otherwise has been considered as out-of-date. ## 2 Turbulent equilibrium In paralled with the classical kinetic theory providing molecular basis of nonturbulent fluid dynamics, molecular description of gases based on the microscopic density $$\underset{¯}{f}(z)=\underset{s=1}{\overset{N}{}}\delta (zz_s(t))$$ (3) has given the ground for the direct numerical simulation of turbulent gases. In the above expression, $`z=(𝒙,𝒗)`$ is a point in the phase space, namely, physical ($`𝒙`$) plus molecular velocity ($`𝒗`$) space, and $`\delta `$ denotes the (6D) delta function. This function gives, by definition, the instantaneous number density in the phase space, so is identified with unaveraged Boltzmann function. It was first applied to the plasma kinetic theory where the equation governing $`\underset{¯}{f}`$ is Vlasov’s equation. For monatomic (neutral) ideal gases, the governing equation, derived as an equation of continuity in the phase space with precise spatial and temporal resolvability of Hamiltonian mechanics level, proves to be an unaveraged Boltzmann equation $$B(\underset{¯}{f})\left(\frac{}{t}+𝒗\frac{}{𝒙}\right)\underset{¯}{f}J(z;\stackrel{~}{z})[\underset{¯}{f}\underset{¯}{\overset{~}{f}}]=0$$ (4) where $`J`$ is the classical collision integral operator $$J(z;\stackrel{~}{z})[\underset{¯}{f}\underset{¯}{\overset{~}{f}}]=\left[(f^{}\stackrel{~}{f}^{}f\stackrel{~}{f})V𝑑\mathrm{\Omega }𝑑\stackrel{~}{𝒗}\right]_{\stackrel{~}{𝒙}=𝒙}$$ (5) In the above $`\underset{¯}{\overset{~}{f}}=\underset{¯}{f}(\stackrel{~}{z})`$ is the microscopic density of the collision partner molecule $`\stackrel{~}{z}`$, $`\stackrel{~}{f}^{}=f(\stackrel{~}{z}^{})`$ likewise with $`\widehat{z}^{}`$ specifying the pre-collision state of the collision partner leading to $`\stackrel{~}{z}`$ after the collision, $`V=|\stackrel{~}{𝒗}𝒗|`$, and $`d\mathrm{\Omega }`$ is the differential collision-cross section of the molecular encounter. The microscopic density, being averaged over a space with a number of unidentifiable molecules included, defines the Boltzmann function $`f`$; $$f=\overline{\underset{¯}{f}}.$$ (6) This is where the statistical concept is introduced for a quantity of deterministic mechanics. Similarly, two-point correlation function $`\psi (z,\stackrel{~}{z})`$ is defined using the same averaging concept as $$\psi (z,\stackrel{~}{z})=\overline{\underset{¯}{f}\underset{¯}{\overset{~}{f}}}f\stackrel{~}{f}$$ (7) Higher order correlations can be defined likewise. The classical kinetic theory is founded on Eq.(4) averaged over the same space as (6) together with the Boltzmann’s (binary) molecular chaos hypothesis $$\psi =0$$ (8) that is the Boltzmann equation in the classical sense; $$\left(\frac{}{t}+𝒗\frac{}{𝒙}\right)fJ(z;\stackrel{~}{z})[f\stackrel{~}{f}]=0$$ (9) Under the equilibrium condition ($`/t=0,/𝒙=0`$) the equation reduces to $$J(z;\stackrel{~}{z})[f\stackrel{~}{f}]=0$$ (10) whose solution is known as the Maxwellian distribution $$\begin{array}{c}f=f_0=\frac{n}{(2\pi c^2)^{3/2}}\mathrm{exp}\left(\frac{𝒘^2}{2c^2}\right)\hfill \\ 𝒘𝒗𝒖\hfill \\ c=(T/)^{1/2}\hfill \end{array}\}$$ (11) where $`n,𝒖,T`$ are the number density, fluid velocity and temperature, $``$ and $``$ denote the universal gas constant and molar weight, respectively. If a gas is turbulent, fluid variables exhibit (long-range) correlations, so the classical binary chaos (8) ceases to hold. Then Eq.(9) is to be replaced with $$\overline{B(\underset{¯}{f})}=\left(\frac{}{t}+𝒗\frac{}{𝒙}\right)fJ(z;\stackrel{~}{z})[f\stackrel{~}{f}+\psi (z,\stackrel{~}{z})]=0$$ (12) Thus the equation is not closed at this level: The equation of next step to govern $`\psi `$ is $$\overline{f^{}B(\underset{¯}{\overset{^}{f}})+B(\underset{¯}{f})\widehat{f}^{}}=0(f^{}\underset{¯}{f}f)$$ (13) which is shown to be identical with the two-particle hierarchy equation of the BBGKY theory. The actual form of (13) reads $$\begin{array}{c}(\frac{}{t}+𝒗\frac{}{𝒙}+\widehat{𝒗}\frac{}{\widehat{𝒙}})\psi (z,\widehat{z})=J(z;\stackrel{~}{z})[f\psi (\widehat{z},\stackrel{~}{z})+\stackrel{~}{f}\psi (z,\widehat{z})\hfill \\ +\psi _{\text{III}}(z,\widehat{z},\stackrel{~}{z})]+J(\widehat{z};\stackrel{~}{z})[\widehat{f}\psi (z,\stackrel{~}{z})+\stackrel{~}{f}\psi (z,\stackrel{~}{z})+\psi _{\text{III}}(z,\widehat{z},\stackrel{~}{z})]\hfill \end{array}$$ (14) The simplest possible truncation of the chain of equations at this level is to put $$\psi _{\text{III}}=0$$ (15) namely, to invoke tertiary molecular chaos to replace the classical binary chaos (8). Eq.(14) subject to condition (15) turns out to be separable into those for respective variables in terms of the wave number $`𝒌`$, $$\psi (z,\widehat{z})=\mathrm{R}.\mathrm{P}.l^3\varphi (z,𝒌)\varphi ^{}(\widehat{z},𝒌)𝑑𝒌$$ (16) where symbol $``$ denotes complex conjugate, $`l`$ denotes a length characteristic of the flow geometry and R.P. stands for the real part. Upon its substitution into (14) we have the equation governing $`\varphi `$ as $$\begin{array}{c}i\omega (𝒌)\varphi =\mathrm{\Omega }_0(\varphi )\hfill \\ \mathrm{\Omega }_0(\varphi )\left(\frac{}{t}+𝒗\frac{}{𝒙}\right)\varphi J(z;\stackrel{~}{z})[f\stackrel{~}{\varphi }+\varphi \stackrel{~}{f}]\hfill \end{array}\}$$ (17) where $`\omega (𝒌)`$ is the separation constant having dimension of the frequency, and connected with the wave number through phase velocity $`𝑽_p`$ as $$\omega =𝑽_p𝒌$$ (18) ¿From Eqs.(17) and (18) the following relationship is seen to hold; $$\varphi ^{}(𝒌)=\varphi (𝒌)$$ (19) A remark should be mentioned on a renovation that has brought the scale of turbulence into formalism: In the older theory variable separation (16) is effected in the domain of frequency $`\omega `$ in place of wave number $`𝒌`$. Information on the scale of turbulence is missed in this form, which may be retrieved in the following way: Here turbulence is characterized as a wave with frequency $`\omega `$ and wave number $`𝒌`$ obeying dispersion relationship (18). We may assume a plane wave $$\varphi =e^{i𝒌𝒙}\mathrm{\Phi }(z,𝒌)$$ (20) and discuss its amplitude $`\mathrm{\Phi }`$. If further we introduce Fourier transform $$\mathrm{\Phi }(z,𝒌)=\frac{1}{(2\pi )^3}_{\mathrm{}}^{\mathrm{}}e^{i𝒌𝒔}F(z,𝒔)𝑑𝒔$$ (21) to work with ‘eddy’ space $`𝒔`$, and employ separation rule written in terms of wave number $`𝒌`$, namely after (16), a simple formula follows: $$\psi _{\text{II}}(z,\widehat{z})=\frac{1}{(2\pi l)^3}_{\mathrm{}}^{\mathrm{}}F(z,𝒔)F(\widehat{z},𝒔+𝒓)𝑑𝒔$$ (22) where $`𝒓=\widehat{𝒙}𝒙`$ is the distance between the two space points. Any informations on scale of turbulence can be obtained from this correlation formula by taking fluid moment of (22), for example, the velocity-velocity fluctuation correlation as $$\overline{u_j^{}(𝒙)𝒖_l^{}(𝒙+𝒓)}=(\rho \widehat{\rho })^1w_j\widehat{w}_l\psi _{\text{II}}𝑑𝒗𝑑\widehat{𝒗}$$ (23) where $`𝒘`$ has been defined by (11). It is a well-posed problem to seek the equilibrium solution of the new set of equations (12) and (17). It stands for a ‘turbulent equilibrium’ solution to substitute the local Maxwellian for nonturbulent gases. This is to solve simultaneous integral equations $$J(z;\stackrel{~}{z})[f\stackrel{~}{f}+\varphi \stackrel{~}{\varphi }^{}𝑑𝒌]=0$$ (24) $$J(z;\stackrel{~}{z})[f\stackrel{~}{\varphi }+\varphi \stackrel{~}{f}]=0$$ (25) The exact solution under the assumption that turbulence prevails only in the temperature has been obtained in a closed form by Sagara by summing an infinite series in the temperature fluctuation $`\delta T=(\overline{T^{}_{}{}^{}2})^{1/2}`$. The same result has been reached using much simpler method. We can show that the exact solution also exists for realistic turbulence where the fluctuations in number density $`\delta n=(\overline{n^{}_{}{}^{}2})^{1/2}`$ and fluid velocity $`\delta u_j=(\overline{u_j^{}_{}{}^{}2})^{1/2}`$ (no summation convention here) are also prevalent. Put $$f=f^0=\frac{1}{2}(f_0^++f_0^{})$$ (26) $$\varphi =\varphi ^0=\frac{1}{2}K(𝒌)(f_0^+f_0^{})$$ (27) where $`f_0`$ is the local Maxwellian defined by (11), and $`f_0^\pm `$ is given as follows; $$\begin{array}{ccc}f_0^\pm & =& \frac{n^\pm }{(2\pi c^{\pm 2})^{3/2}}\mathrm{exp}\left(\frac{w_k^{\pm 2}}{2c^{\pm 2}}\right)\hfill \\ n^\pm & =& n\pm \delta n\hfill \\ w_j^\pm & =& v_ju_j^\pm \hfill \\ u_j^\pm & =& u_j\pm \delta u_j\hfill \\ c^{\pm 2}& =& R(T\pm \delta T)\hfill \\ R& =& /\hfill \end{array}\}$$ (28) It is easily confirmed that the form of (26) is designed to be consistent with definitions of the average quantity $`(n,𝒖,T)`$, and the form of (27) with that of the variance $`(\delta n,\delta 𝒖,\delta T)`$ of each quantity provided that the following relationship holds; $$_{\mathrm{}}^{\mathrm{}}KK^{}𝑑𝒌=1$$ (29) Direct substitution of (26) and (27) into (24) and (25) reads, respectively, $$\begin{array}{c}\frac{1}{2}J(z;\widehat{z})[f_0^+\widehat{f}_0^++f_0^{}\widehat{f}_0^{}]=\frac{1}{2}(J^++J^{})\\ \frac{K}{2}J(z;\widehat{z})[f_0^+\widehat{f}_0^+f_0^{}\widehat{f}_0^{}]=\frac{K}{2}(J^+J^{})\end{array}\}$$ (30) where $`J^\pm `$ has been defined as $$J^\pm =J(z;\widehat{z})[f_0^\pm \widehat{f}_0^\pm ]$$ (31) This integral vanishes in view of the fact that $`J(z;\widehat{z})[f_0\widehat{f}_0]=0`$, also that the bilinear Maxwellians in (31) consist of the same family in fluid parameters (Q.E.D.). The deduction above claims that the state of turbulent equilibrium is represented by the Boltzmann function of the form (26), along with correlation function $$\psi (z,\widehat{z})=\frac{1}{4}(f_0^+f_0^{})(\widehat{f}_0^+\widehat{f}_0^{})$$ (32) or the two-point distribution function $$\begin{array}{ccc}\hfill f_{\text{II}}(z,\widehat{z})& & f\widehat{f}+\psi (z,\widehat{z})\hfill \\ & =& \frac{1}{2}(f_0^+\widehat{f}_0^++f_0^{}\widehat{f}_0^{})\hfill \end{array}$$ (33) At a glance difference between the two Boltzmann functions $`f_0`$ of (3) and $`f^0`$ of (26) looks trivial, because the difference is of the first order smallness in $`\delta `$. The point at issue, however, is that they differ appreciably in the population of high energy molecules that play a crucial role in chemical reactions, particularly when the activation energy is high. In Fig.4 is shown the population density of molecules having absolute molecular velocity $`v`$ with turbulence only in the temperature. The Maxwellian distribution $`(\delta T=0)`$ is seen to have lean high energy tail, namely, underestimates population of high energy molecules compared to reality of turbulence. ## 3 Turbulent chemical reactions Chemical reaction rate of elementary reaction of the form $$[\mathrm{a}]+[\mathrm{b}][\mathrm{c}]+[\mathrm{d}]([];\mathrm{chemical}\mathrm{symbol})$$ usually obeys Arrhenius’ law; $$W=An_\mathrm{a}n_\mathrm{b}\mathrm{exp}(E/T)$$ (34) where $`n_\mathrm{a}`$ and $`n_\mathrm{b}`$ denote partial number densities of reactants $`[\mathrm{a}]`$ and $`[\mathrm{b}]`$, $`A`$ is the frequency factor and $`E`$ is the activation energy per mol. Rate law (34) is warranted on the classical kinetic theory using a simple collision model. A most clearcut model among various ones proposed in late 40’s through 50’s is due to Present : Let the intermolecular potential of a molecule \[a\] relative to a molecule \[b\] be repulsive, of a volcano shape having a crater with radius $`r_0`$ at altitude $`ϵ`$. Let also the locus of the \[b\] molecule relative to the \[a\] molecule written in axisymmetric polar coordinate be $`[r(t),\chi (t)]`$ where $`\chi `$ is the angle between the symmetry axis parallel to $`𝑽_{\mathrm{ba}}=𝒗_\mathrm{b}𝒗_\mathrm{a}`$ and the position vector $`𝒓=𝒙_\mathrm{b}𝒙_\mathrm{a}`$ then we have $$\begin{array}{c}r^2\dot{\chi }=bV\hfill \\ (m^{}/2)(\dot{r}^2+r^2\dot{\chi }^2)+ϵ(r)=(m^{}/2)V^2\hfill \end{array}\}$$ (35) relationships representing conservation of angular momentum and of energy, where $`\dot{\chi }=dx/dt`$ etc., $`m^{}=(m_\mathrm{a}^1+m_\mathrm{b}^1)^1`$ is the reduced mass, $`b`$ is the impact parameter, and $`ϵ(r)`$ denotes the intermolecular potential. Since the radius of the closest approach $`r=r_c`$ is given by the condition $`dr/d\chi =0`$, the critical collision that separates elastic from inelastic collisions is such that the closest approach occurs on the ridge line of the crater ; $`r_c=r_0`$. This condition determines the critical impact parameter $`b_c(V)`$ from (35) as $$\frac{b_c(V)}{r_0}=\left(1\frac{2ϵ}{m^{}V^2}\right)^{1/2}$$ (36) Instantaneous reaction rates for respective species are expressed in the following form, $$\begin{array}{c}\underset{¯}{W}_\mathrm{a}=m_\mathrm{a}\underset{¯}{W},\underset{¯}{W}_\mathrm{b}=m_\mathrm{b}\underset{¯}{W},\underset{¯}{W}_\mathrm{c}=m_\mathrm{c}\underset{¯}{W},\underset{¯}{W}_\mathrm{d}=m_\mathrm{d}\underset{¯}{W}\hfill \\ \underset{¯}{W}=_{\mathrm{inel}}\underset{¯}{f}(z_\mathrm{a})\underset{¯}{f}(z_\mathrm{b})V𝑑\mathrm{\Omega }_{\mathrm{ab}}𝑑𝒗_\mathrm{a}𝑑𝒗_\mathrm{b}\hfill \end{array}$$ (37) Utilizing the following formula $$_{\mathrm{inel}}𝑑\mathrm{\Omega }_{\mathrm{ab}}=\pi b_c^2=\pi r_0^2\left(1\frac{2ϵ}{m^{}V^2}\right)$$ and transforming integral variables from $`(𝒗_\mathrm{a},𝒗_\mathrm{b})`$ to $`(𝑼,𝑽)`$ defined by $$\begin{array}{ccc}\hfill 𝑼& =& (m_\mathrm{a}𝒗_\mathrm{a}+m_\mathrm{b}𝒗_\mathrm{b})/(m_\mathrm{a}+m_\mathrm{b})\hfill \\ \hfill 𝑽& =& 𝒗_\mathrm{b}𝒗_\mathrm{a}\hfill \end{array}\}$$ (38) we have for (33) $$\underset{¯}{W}=\pi r_0^2_{V_c}^{\mathrm{}}𝑑𝑽V\left(1\frac{V_c^2}{V^2}\right)_{\mathrm{}}^{\mathrm{}}\underset{¯}{f}(z_\mathrm{a})\underset{¯}{f}(z_\mathrm{b})𝑑𝑼$$ (39) where $$V_c=(2ϵ/m^{})^{1/2}$$ (40) denotes the critical relative velocity for a head-on collision. Chemical reaction rate in classical sense is obtained from (39), upon averaging and employing classical chaos (8), as $$W=\pi r_0^2_{V_c}^{\mathrm{}}𝑑𝑽V\left(1\frac{V_c^2}{V^2}\right)_{\mathrm{}}^{\mathrm{}}f(z_\mathrm{a})f(z_\mathrm{b})𝑑𝑼$$ (41) If the Maxwellian equilibrium (11) is employed Eq.(41) is integrated out to give $$\begin{array}{c}W_0(T,n_\mathrm{a},n_\mathrm{b})=An_\mathrm{a}n_\mathrm{b}e^{ϵ/kT}\hfill \\ A=(8kT/\pi m^{})^{1/2}\pi r_0^2\hfill \end{array}\}$$ (42) that is nothing but Arrhenius law (34) with activation energy $`E`$ and the frequency factor $`A`$ written in microscopic parameters. It should be noted that turbulence counterpart of the above deduction faces with difficulty if conducted phenomenologically, whereas it is simply straightforward if one follows the molecular approach as employed here. In fact, if one assumes that the instantaneous reaction rate is Arrhenius form (34) $$\underset{¯}{W}=\underset{¯}{A}\underset{¯}{n}_\mathrm{a}\underset{¯}{n}_\mathrm{b}\mathrm{exp}(E/\underset{¯}{T})$$ and its average given by $$W=\overline{\underset{¯}{A}\underset{¯}{n}_\mathrm{a}\underset{¯}{n}_\mathrm{b}\mathrm{exp}(E/\underset{¯}{T})}$$ (43) Decomposing the r.h.s. term under average into means and fluctuations we have $$W=An_\mathrm{a}n_\mathrm{b}\overline{(1+O(n_\mathrm{a}^{},n_\mathrm{b}^{}))e^{\beta +\beta ^2\tau \beta ^3\tau ^2+\mathrm{}}}$$ where $`\beta =E/T`$ is the Zel’dovich number and $`\tau =T^{}/E`$ stands for the temperature fluctuation. Leading term in the Taylor expansion of the expression under average sign gives for $`\beta 1`$, $$\overline{\frac{d^nW}{d\tau ^n}\frac{\tau ^n}{n!}}\frac{\beta ^n}{n!}\overline{(T^{})^n}$$ which implies that convergence of the series is extremely slow since the Zel’dovich number for combustion reaction is of $`O(10^1)`$. For example we need to take 23 terms before $`\beta ^n/n!`$ diminishes to unity for $`\beta =10`$. It also means that we need to know 22 consecutive self correlations in the temperature fluctuation. Needless to say, therefore, that the first term (Arrhenius) or a few-term approximation is far from satisfactory. On the other hand, if one adopts microscopic approach, namely, starts with the averaged version of (39) insted of (43); $$W=\pi r_0^2_{V_c}^{\mathrm{}}𝑑𝑽V\left(1\frac{V_c^2}{V^2}\right)_{\mathrm{}}^{\mathrm{}}\overline{\underset{¯}{f}_\mathrm{a}\underset{¯}{f}_\mathrm{b}}𝑑𝑼$$ (44) and employs turbulent equilibrium (33) for $`\overline{\underset{¯}{f}_\mathrm{a}\underset{¯}{f}_\mathrm{b}}f_{\text{II}}`$, a simple calculation leads immediately to the final expression $$W^0=\frac{1}{2}(W_0^++W_0^{})$$ (45) with $$W_0^\pm W_0(T\pm \delta T,n_\mathrm{a}\pm \delta n_\mathrm{a},n_\mathrm{b}\pm \delta n_\mathrm{b})$$ (46) where $`W_0`$ has been defined by (42). Formula (45) for the turbulent reaction rate turns out to be a bimodal Arrhenius: It allows us to draw a intuitive picture that a half of the chemical reaction takes place under temperature elevated and population of reactant molecules denser by their root-mean-square turbulent intensity, and another half under those quantities lowered by the same amount. The bimodal Arrhenius low (45) is the full turbulence version derived on the basis of exact solution of Eqs. (24) and (25), supplementing the previous one, where turbulence prevails only in the temperature. As already sketched it is obvious that anomalous increase in the reaction rate is overwhelmingly due to the temperature turbulence that feeds extra high energy molecules capable of reaction (see Fig.4). The first term of (45) looks as if the potential barrier is lowered by the factor of $`(1+\delta T/T)^1`$; a macroscopic equivalent of tunnelling effect of quantum mechanics where the potential barrier is lowered by a factor of Planck’s constant due to the uncertainty principle. On the other hand, fluctuation in the reactant number density does not have such effects on the reaction rate beyond the first order smallness. However, when coupled with the temperature effect, it leaves with a possible account for below-critical lean combustion that actually occurs in the operation of the direct injection engines (GDI). Rate law (45) also tells us that turbulence in the fluid velocity has no influence on the reaction rate (as it should not). In fact, the collision integral for a pair of molecules about to collide are not influenced by the fluctuation in the fluid velocity, because they are aboard on the ‘same boat’ in the turbulent ocean. Its effects on turbulent flame velocity are exclusively via other transport processes, namely, turbulent diffusion and turbulent heat transfer. (See Table 1, next section). ## 4 Reaction rate variance It is an experimental evidence that instantaneous chemical reaction rate $`\underset{¯}{W}=W+W^{}`$ at a fixed point undergoes fluctuation over a wide range, therefore, quality of a statistical theory employing any average concepts hinges crucially on reliable estimate for its variance around the average $`W`$. This issue again is beyond the reach of phenomenologies and we have to address to statistical approach as developed in the preceding section. To be remarked at this point is that what appear in the governing equations are not the variance itself, but the fluctuation correlations $`\overline{Z^{}W^{}}`$ where $`Z`$ stands for fluid variables such as velocity, temperature, etc.. The equation governing $`W^{}`$ is obtained from (39) by separating it into mean and fluctuating parts. We have, then, $$W^{}=I_{\mathrm{inel}}\left[f^{}(z_\mathrm{a})f(z_\mathrm{b})+f(z_\mathrm{a})f^{}(z_\mathrm{b})+f^{}(z_\mathrm{a})f^{}(z_\mathrm{b})\overline{f^{}(z_\mathrm{a})f^{}(z_\mathrm{b})}\right]$$ (47) where we have defined $$I_{\mathrm{inel}}[\mathrm{\Phi }(z_\mathrm{a},z_\mathrm{b})]\pi r_0^2_{V_c}^{\mathrm{}}𝑑𝑽V\left(1\frac{V_c^2}{V^2}\right)_{\mathrm{}}^{\mathrm{}}\mathrm{\Phi }(z_\mathrm{a},z_\mathrm{b})𝑑𝑼$$ Since, in general, fluid variable fluctuation is expressed as a moment of fluctuation of the Boltzmann function in the velocity space as $$Z^{}=_{\mathrm{}}^{\mathrm{}}\zeta (z_\alpha )f^{}(z_\alpha )𝑑𝒗_\alpha $$ (48) we have, from (43) and (44), $$_{\mathrm{}}^{\mathrm{}}\zeta 𝑑𝒗_\alpha \left\{\overline{W^{}f_\alpha ^{}}I_{\mathrm{inel}}\left[\overline{f_\mathrm{a}^{}f_\alpha ^{}}f_\mathrm{b}+f_\mathrm{a}\overline{f_\alpha ^{}f_\mathrm{b}^{}}+\overline{f_\mathrm{a}^{}f_\alpha ^{}f_\mathrm{b}^{}}\right]\right\}=0$$ (49) Let the variable separation be invoked after (16) as $$\overline{W^{}f_\alpha ^{}}=\mathrm{R}.\mathrm{P}.l^3_{\mathrm{}}^{\mathrm{}}g_W(𝒌)\varphi _\alpha ^{}(𝒌)𝑑𝒌$$ (50) and $$\begin{array}{ccc}\hfill \overline{f_\mathrm{a}^{}f_\alpha ^{}f_\mathrm{b}^{}}& =& \mathrm{R}.\mathrm{P}.l^9\varphi _\mathrm{a}(𝒌_\mathrm{a})\varphi _\alpha (𝒌_\alpha )\varphi _\mathrm{b}(𝒌_\mathrm{b})\delta (𝒌_\mathrm{a}+𝒌_\alpha +𝒌_\mathrm{b})𝑑𝒌_\mathrm{a}𝑑𝒌_\alpha 𝑑𝒌_\mathrm{b}\hfill \\ & =& \mathrm{R}.\mathrm{P}.l^6\varphi _\alpha ^{}(𝒌)\varphi _\mathrm{a}(\stackrel{~}{𝒌})\varphi _\mathrm{b}(𝒌\stackrel{~}{𝒌})𝑑𝒌𝑑\stackrel{~}{𝒌}\hfill \end{array}$$ where, in the last row, $`𝒌_\alpha `$ is replaced with $`𝒌`$, also $`𝒌_\mathrm{a}`$ with $`\stackrel{~}{𝒌}`$, and condition (19) has been made use of. Then Eq.(49) leads to $$\begin{array}{c}_{\mathrm{}}^{\mathrm{}}d𝒌\varphi _\alpha ^{}(𝒌)\{g_W(𝒌)I_{\mathrm{inel}}[\varphi _\mathrm{a}(𝒌)f_\mathrm{b}+f_\mathrm{a}\varphi _\mathrm{b}(𝒌)\hfill \\ +\mathrm{R}.\mathrm{P}.l^3_{\mathrm{}}^{\mathrm{}}d\stackrel{~}{𝒌}\varphi _\mathrm{a}(𝒌\stackrel{~}{𝒌})\varphi _\mathrm{b}(\stackrel{~}{𝒌})d\stackrel{~}{𝒌}]\}=0\hfill \end{array}$$ (51) If one decomposes $`g_W`$ and $`\varphi `$ into a plane wave after (20) $$\begin{array}{ccc}\hfill g_W& =& e^{i𝒌𝒙}G_W\hfill \\ \hfill \varphi & =& e^{i𝒌𝒙}\mathrm{\Phi }\hfill \end{array}\}$$ (52) and deals with its amplitude $`G_W`$ and $`\mathrm{\Phi }`$, or its Fourier-transform version $$\begin{array}{ccc}\hfill G_W(𝒙,𝒌)& =& \frac{1}{(2\pi l)^3}_{\mathrm{}}^{\mathrm{}}𝑑𝒔e^{i𝒌𝒔}q_W(𝒙,𝒔)\hfill \\ \hfill \mathrm{\Phi }(𝒙,𝒌)& =& \frac{1}{(2\pi l)^3}_{\mathrm{}}^{\mathrm{}}𝑑𝒔e^{i𝒌𝒔}F(𝒙,𝒔)\hfill \end{array}\}$$ (53) we have two alternatives for the reaction rate fluctuation from (51), $$G_W(𝒙,𝒌)=I_{\mathrm{inel}}[\mathrm{\Phi }_\mathrm{a}f_\mathrm{b}+f_\mathrm{a}\mathrm{\Phi }_\mathrm{b}+\mathrm{R}.\mathrm{P}.l^3_{\mathrm{}}^{\mathrm{}}d\stackrel{~}{𝒌}\mathrm{\Phi }_\mathrm{a}(𝒌\stackrel{~}{𝒌})\mathrm{\Phi }_\mathrm{b}(\stackrel{~}{𝒌})d\stackrel{~}{𝒌}]$$ (54) $$q_W(𝒙,𝒔)=I_{\mathrm{inel}}[F_\mathrm{a}f_\mathrm{b}+f_\mathrm{a}F_\mathrm{b}+F_\mathrm{a}F_\mathrm{b}]$$ (55) depending on which of the wave number $`(𝒌)`$ or the eddy $`(𝒔)`$ spaces is more convenient to work with. The actual form of $`q_W`$ associated with turbulent equilibrium (26) and (27) is straightforward. In fact, in view of (52) and (53), we have $$F_\alpha ^0(z,𝒔)=S(𝒙,𝒔)\frac{f_{0\alpha }^+f_{0\alpha }^{}}{2}$$ (56) with $$S(𝒙,𝒔)=(2\pi l)^{3/2}\frac{q_0}{\left({\displaystyle _{\mathrm{}}^{\mathrm{}}}q_0^2𝑑𝒔\right)^{1/2}}$$ (57) where $`q_0`$ is related to variance of the density fluctuation by $$(\delta \rho )^2=\frac{1}{(2\pi l)^3}_{\mathrm{}}^{\mathrm{}}q_0^2𝑑𝒔$$ (58) Utilizing (56) in (55) we have $$q_W(𝒙,𝒌)=\frac{S}{2}\left[W_0^+W_0^{}+\frac{S}{2}(W_0^++W_0^{}W_0^+)\right]$$ (59) where $$\begin{array}{c}W_0^+=A\{(n_\mathrm{a}+\delta n_\mathrm{a})(n_\mathrm{b}\delta n_\mathrm{b})\mathrm{exp}[E/(T+\frac{m_\mathrm{a}m_\mathrm{b}}{m_\mathrm{a}+m_\mathrm{b}}\delta T)]\hfill \\ +(n_\mathrm{a}\delta n_\mathrm{a})(n_\mathrm{b}+\delta n_\mathrm{b})\mathrm{exp}[E/(T\frac{m_\mathrm{a}m_\mathrm{b}}{m_\mathrm{a}+m_\mathrm{b}}\delta T)]\}\hfill \end{array}$$ (60) Once we have obtained explicit expression for $`q_W`$ variance of the reaction rate can be calculated from (see Eq.(50)) $$\begin{array}{ccc}\hfill \overline{W^{}_{}{}^{}2}& =& \mathrm{R}.\mathrm{P}.l^3_{\mathrm{}}^{\mathrm{}}g_Wg_W^{}𝑑𝒌\hfill \\ & =& \frac{1}{(2\pi l)^3}_{\mathrm{}}^{\mathrm{}}q_W^2𝑑𝒔\hfill \end{array}$$ (61) where (52), (53) have been taken into account. In Fig.5a is shown the relative variance $`(\overline{W^{}_{}{}^{}2})^{1/2}/W_0`$ of reaction rate around classical Arrhenius, which amounts to quantities exceeding $`O(1)`$ by far, indicating extraordinary fluctuation intensity to be observed by a fixed-point measurement. Also is shown in Fig.5b the same ratio for turbulent reaction rate (41). In this case the relative variance stays within sound realm, and approaches unity with increase in turbulence in the temperature and with Zel’dovich number. This is a sign of stochastic signals characteristic of on-off type. ## 5 Phenomenology-coupled Boltzmann formalism on turbulent equilibrium Statistical description of stochastic phenomena where relative variance far exceeds unity is pathological. It is for this reason that a formalism using the averaged Arrhenius law fails in describing turbulent combustion. However, as asserted in the preceding section, such formalism using the bimodal law as the averaged reaction rate deserves reconsideration. For this purpose it is advisable to follow the recipe prepared for nonreactive turbulence: We start with phenomenological equations generalized to reactive gases written in instantaneous quantities: $$\underset{¯}{\mathrm{\Lambda }}_0\frac{\underset{¯}{\rho }}{t}+\frac{\underset{¯}{m}_r}{x_r}=0$$ (62) $$\underset{¯}{m}_r\underset{¯}{\rho }\underset{¯}{u}_r$$ (63) $$\underset{¯}{\mathrm{\Lambda }}_j\frac{m_j}{t}+\frac{}{x_r}\left(\frac{\underset{¯}{m}_j\underset{¯}{m}_r}{\underset{¯}{\rho }}+\underset{¯}{p}\delta _{jr}+\underset{¯}{p}_{jr}\right)=0$$ (64) $$\underset{¯}{p}_{jr}\underset{¯}{\mu }\left[\frac{}{x_j}\left(\frac{\underset{¯}{m}_r}{\underset{¯}{\rho }}\right)+\frac{}{x_r}\left(\frac{\underset{¯}{m}_j}{\underset{¯}{\rho }}\right)\frac{2}{3}\delta _{jr}\frac{}{x_k}\left(\frac{\underset{¯}{m}_k}{\underset{¯}{\rho }}\right)\right]$$ (65) $$\begin{array}{c}\underset{¯}{\mathrm{\Lambda }}_4\frac{\underset{¯}{E}}{t}+\frac{}{x_r}\left(\underset{¯}{m}_r\underset{¯}{H}+\underset{¯}{Q}_r\right)+e_\alpha ^0\underset{¯}{W}_\alpha =0\hfill \\ \underset{¯}{E}\underset{¯}{\rho }\underset{¯}{e}+\frac{\underset{¯}{m}_k^2}{2\underset{¯}{\rho }},\underset{¯}{H}\underset{¯}{E}+\underset{¯}{p}\hfill \end{array}\}$$ (66) $$\underset{¯}{Q}_r\underset{¯}{\lambda }\frac{}{x_r}\left(\frac{\underset{¯}{p}}{\underset{¯}{\rho }R}\right)$$ (67) $$\begin{array}{c}\underset{¯}{\mathrm{\Lambda }}_\mathrm{a}\frac{\underset{¯}{\rho }_\mathrm{a}}{t}+\frac{}{x_r}\left(\underset{¯}{\rho }_\mathrm{a}\underset{¯}{u}_r+\underset{¯}{M}_{\mathrm{a},r}\right)\underset{¯}{W}_\mathrm{a}=0\hfill \\ \underset{¯}{Y}_\mathrm{a}\underset{¯}{\rho }_\mathrm{a}/\underset{¯}{\rho }\hfill \end{array}\}$$ (68) $$\underset{¯}{M}_{\mathrm{a},r}\underset{¯}{\rho }\underset{¯}{D}_\mathrm{a}\frac{\underset{¯}{Y}_\mathrm{a}}{x_r}$$ (69) In the above, $`\underset{¯}{\rho },\underset{¯}{m}_j,\underset{¯}{p}`$ and $`\underset{¯}{e}`$ denote density, massflux density, pressure and specific internal energy (thermal part only), $`\underset{¯}{p}_{jr}`$, $`\underset{¯}{Q}_r`$ and $`\underset{¯}{M}_{\mathrm{a}r}`$ are viscous stress, heat flux density and diffusion mass flux density for species a, and $`\underset{¯}{\rho }_\mathrm{a}`$, $`\underset{¯}{Y}_\mathrm{a}`$ and $`e_\mathrm{a}^0`$ are partial density, mass fraction and zero-point enthalpy of species a, respectively. Also, $`\underset{¯}{\mu }`$, $`\underset{¯}{\lambda }`$ and $`\underset{¯}{D}_\mathrm{a}`$ are coefficients of viscosity, thermal conductivity and diffusion for species a, respectively. Note that all the equations have the form in accordance with a rule (rule I) that they are written in terms of quantities proportional to the density. This is to use mass-flux density instead of fluid velocity, pressure instead of temperature, and internal energy per unit of volume instead of specific internal energy. A few words must be mentioned about modified Fick’s law (69) that is not exact consequence of the kinetic theory of gases as distinct from other molecular transport relationships. The ‘effective’ diffusion coefficient $`D_\mathrm{a}`$ is expressed in terms of binary diffusion coefficient for each pair of constituents as $$\begin{array}{ccc}\hfill D_\mathrm{a}& =& (1Y_\mathrm{a})/\underset{\mathrm{b}}{}D_{\mathrm{ab}}^1Y_\mathrm{b}\hfill \\ & =& 1/\underset{\mathrm{b}}{}D_{\mathrm{ab}}^1Y_\mathrm{b}^{}\hfill \end{array}$$ (70) where $`Y_\mathrm{b}^{}=Y_\mathrm{b}/(1Y_\mathrm{a})`$ is mass fraction of species b of the mixture with species a excepted. The approximate formula (69) with (70) warrants validity for practical use except a minor flaw that condition $`{\displaystyle \underset{\mathrm{a}}{}}\underset{¯}{M}_{\mathrm{a},\mathrm{r}}=0`$ is not exactly met. Equations for the average are provided by $$\mathrm{\Lambda }_\alpha \overline{\underset{¯}{\mathrm{\Lambda }}}_\alpha =0(\alpha ;\mathrm{\hspace{0.17em}0},j,\mathrm{\hspace{0.17em}4},\mathrm{a},\mathrm{b},\mathrm{})$$ (71) whose actual forms are given as follows: $$\mathrm{\Lambda }_0=\frac{\rho }{t}+\frac{m_r}{x_r}=0$$ (72) $$\mathrm{\Lambda }_j=\frac{m_j}{t}+\frac{}{x_r}\left(\frac{m_jm_r}{\rho }+p\delta _{jr}+p_{jr}\right)=0$$ (73) $$p_{jr}(\overline{\underset{¯}{p}}_{jr})_{\mathrm{NS}}+\rho \overline{u_j^{}u_r^{}}$$ (74) $$\mathrm{\Lambda }_4=\frac{E}{t}+\frac{}{x_r}\left(\frac{m_r}{\rho }H+Q_r\right)+\underset{\mathrm{a},\mathrm{b},\mathrm{}}{}e_\mathrm{a}^0W_\mathrm{a}=0$$ (75) $$Q_r(\overline{\underset{¯}{Q}}_r)_{\mathrm{Fourier}}+\rho \overline{u_r^{}h^{}}$$ (76) $$\mathrm{\Lambda }_\mathrm{a}=\frac{\rho _\mathrm{a}}{t}+\frac{}{x_r}\left(\frac{m_r}{\rho }\rho _\mathrm{a}+M_{\mathrm{a},r}\right)W_\mathrm{a}=0$$ (77) $$M_{\mathrm{a},r}(\overline{\underset{¯}{M}}_{\mathrm{a},r})_{\mathrm{Fick}}+\rho \overline{u_r^{}Y_\mathrm{a}^{}}$$ (78) $$W_\mathrm{a}=\frac{1}{2}(W_{0\mathrm{a}}^++W_{0\mathrm{a}}^{})$$ (79) where $`p_{jr}`$, $`Q_r`$, $`M_{\mathrm{a},r}`$ and $`W_\mathrm{a}`$ are average transports of momentum, thermal energy, species mass due to diffusion and chemical reactions, viewed from the center of gravity frame of reference, respectively. Of the four transport processes three but chemical reactions have a structure such that molecular and turbulent transports are additive, whereas these are tightly coupled and built in a single exponential form for chemical reactions (Table 1). A few remarks are in order in deriving Eqs.(72) through (78) from (71): They are exact to $`O(Z^{}_{}{}^{}2)`$ if Mach number is small enough. Single term turbulence corrections as derived here would not be available unless one followed rule I plus the following one (rule II) that fluctuation of each density-proportional quantity be written in density-independent quantities except fluctuation in density itself, for example, $`\overline{\underset{¯}{m}}_j=m_j+\rho u_j^{}+\rho ^{}u_j+\rho ^{}u_j^{}\overline{\rho ^{}u_j^{}}`$. It is remarkable that turbulence correction to each of Navier-Stokes, Fourier and Fick laws is represented by such a simple expression without employing the so-called mass average concept. Note also that artifice of this kind is not necessary to reach the same result if phenomenologies are discarded and the whole deduction to fluid equations starts from the B-formalism in the phase space, which, unfortunately, is practicable only to nonreactive monatomic gases. Equations governing turbulent transports to close the system emerge from the following equations $$\overline{Z_\alpha ^{}\widehat{\mathrm{\Lambda }}_\beta ^{}+\mathrm{\Lambda }_\alpha ^{}\widehat{Z}_\beta ^{}}=0,((\alpha ,\beta );(0,j,\mathrm{\hspace{0.17em}4},\mathrm{a},\mathrm{b},\mathrm{}))$$ (80) where $`\mathrm{\Lambda }_\alpha ^{}=\underset{¯}{\mathrm{\Lambda }}_\alpha \mathrm{\Lambda }_\alpha `$, and $`Z_\alpha ^{}`$ denotes fluctuation of fluid variable $`Z_\alpha `$. Regardless of the fact that those equations are nonlinear, they are again separable exactly in the form of (22) and (23), resulting in the following set of equations (see Appendix for its derivation); $$Dq_0+_rq_r=0$$ (81) $$\begin{array}{c}Dq_j+_r(u_r^{}q_j+q_{40}\delta _{rj}+q_{rj})+q_ru_j^{}/x_r\rho ^1(p/x_j)q_0=0\hfill \\ q_{jr}=\mu [_j(\rho ^1q_r)+_r(\rho ^1q_j)(2/3)\delta _{jr}_k(\rho ^1q_k)]\hfill \\ [u_r^{}/x_j+u_j^{}/x_r(2/3)\delta _{jr}u_k^{}/x_k]\frac{q_4}{\rho R}\frac{d\mu }{dT}\hfill \\ +\rho ^1[q_jq_r(\delta _{jr}/3)q_k^2]\hfill \end{array}\}$$ (82) $$\begin{array}{c}\frac{1}{\gamma 1}Dq_{40}+_r\left(\frac{\gamma }{\gamma 1}q_{40}u_r^{}+q_{rkk}\right)+\underset{\mathrm{a}}{}e_\mathrm{a}^0q_{\mathrm{a}W}=0\hfill \\ q_{rkk}=(\lambda /R)[_r(\rho ^1q_4)+p^1q_4\frac{d\lambda }{dT}RT/x_r]\hfill \\ +[\gamma /(\gamma 1)](RTq_r+\rho ^1q_rq_4)\hfill \end{array}\}$$ (83) $$\begin{array}{c}Dq_\mathrm{a}+_r(u_r^{}q_\mathrm{a}+q_{\mathrm{a},r})+q_rY_\mathrm{a}/x_rq_{W\mathrm{a}}+\rho ^1q_0W_\mathrm{a}=0\hfill \\ q_{\mathrm{a},r}=\rho D_\mathrm{a}_r(\rho ^1q_\mathrm{a})\frac{d(\rho D_\mathrm{a})}{dT}\frac{q_4}{\rho R}\frac{Y_\mathrm{a}}{x_r}+\rho ^1q_\mathrm{a}q_r\hfill \\ q_{W\mathrm{a}}=\pm m_\mathrm{a}q_W(\pm ;\mathrm{species}[\mathrm{a}]\mathrm{produced}/\mathrm{disappeared})\hfill \\ q_W=\frac{S}{2}\left[W_0^+W_0^{}+\frac{S}{2}(W_0^++W_0^{}W_0^+)\right]\hfill \end{array}\}$$ (84) $$q_{40}=q_4+\rho Rq_0$$ (85) where $`_j/x_j+/s_j`$ $`Dq_\alpha q_\alpha /tV_{pr}q_\alpha /s_r`$ Of these equations terms of fluctuations in transport processes are listed in Table 2. Turbulent transports having appeared in equations for the average are written in terms of $`q`$’s as $$\begin{array}{c}\rho \overline{u_j^{}u_r^{}}=\frac{1}{\rho (2\pi )^3}_{\mathrm{}}^{\mathrm{}}q_jq_r𝑑𝒔\hfill \\ \rho \overline{h^{}u_r^{}}=\frac{1}{\rho (2\pi )^3}\frac{\gamma }{\gamma 1}_{\mathrm{}}^{\mathrm{}}q_4q_r𝑑𝒔\hfill \\ \rho \overline{Y_\mathrm{a}^{}u_r^{}}=\frac{1}{\rho (2\pi )^3}_{\mathrm{}}^{\mathrm{}}q_\mathrm{a}q_r𝑑𝒔\hfill \\ (\overline{T^{}_{}{}^{}2})^{1/2}=\frac{1}{\rho R}\left(_{\mathrm{}}^{\mathrm{}}q_4^2𝑑𝒔\right)^{1/2}\hfill \end{array}\}$$ (86) Thus the two sets of equations for the average (Eqs.(71)) and separated equations for fluctuation-correlations (Eqs.(80)) are coupled through those quantities. They have to be solved simultaneously subject to homogeneous boundary conditions for $`q`$’s as $$\begin{array}{c}q_\alpha (𝒙_\mathrm{b},𝒔)=0(\alpha 40)\hfill \\ (q_{40}/x_n)_{𝒙=𝒙_\mathrm{b}}=0\hfill \end{array}\}$$ (87) at solid boundaries and at laminar-turbulent flow boundaries $`𝒙=𝒙_\mathrm{b}`$, also $$q_\alpha (𝒙,\pm \mathrm{})=0$$ (88) a necessary condition to secure convergence of the integral in the $`𝒔`$-space. ## 6 Concluding remarks So far the ‘Reynolds average’ concept is considered not to be applicable to turbulent combustion gasdynamics. This negative view has its origin in poor predictivity for one of the four average turbulent transports, namely, the chemical reaction rate. It is concluded that if one employs bimodal Arrhenius law to replace the classical one, sound ‘average’ description of turbulent combustion may well be expected despite its on-off structure of signals to be observed at the reaction front, a sign of marginal variance amplitude allowed for statistical description. The bimodal law as the average turbulent reaction rate and its variance are derived on the basis of the turbulent equilibrium, an exact solution of non-equilibrium statistical mechanics. It is only through the microscopic approach that the correct estimate for population of high energy (reactive) molecules is possible, and thereby high reactivity under turbulent environment is elucidated. Also proposed are Reynolds averaged gasdynamic equations for turbulent combustion with those renovations incorporated. They are coupled with another set governing turbulent fluctuations through the turbulent transport processes to constitute a closed system. ## Appendix. Equations governing turbulent fluctuation-correlations The actual expression for $`\mathrm{\Lambda }_\alpha ^{}`$ of Eq.(80) subject to rules I and II are $$\begin{array}{c}\mathrm{\Lambda }_0^{}\rho ^{}/t+(/x_r)(\rho u_r^{}+\rho ^{}u_r^{}+\rho ^{}u_r^{}\overline{\rho ^{}u_r^{}})=0\hfill \\ u_r^{}=m_r/\rho =(\rho u_r+\overline{\rho ^{}u_r^{}})/\rho \hfill \end{array}\}$$ (89) $$\begin{array}{c}\mathrm{\Lambda }_j^{}\frac{}{t}(\rho ^{}u_j^{}+\rho u_j^{})+\frac{}{x_r}[\rho u_r^{}u_j^{}+\rho u_j^{}u_r^{}\hfill \\ +\rho ^{}u_j^{}u_r^{}+\rho u_j^{}u_r^{}\rho \overline{u_j^{}u_r^{}}+p^{}\delta _{jr}+(p_{jr}^{})_{\mathrm{NS}}]=0\hfill \\ (p_{jr}^{})_{\mathrm{NS}}=\mu [u_r^{}/x_j+u_j^{}/x_r(2\delta _{jr}/3)u_k^{}/x_k]\hfill \\ (d\mu /dT)T^{}[u_r^{}/x_j+u_j^{}/x_r(2\delta _{jr}/3)u_k^{}/x_k]\hfill \end{array}\}$$ (90) $$\begin{array}{c}\mathrm{\Lambda }_4^{}=\frac{1}{\gamma 1}\frac{p^{}}{t}+\frac{}{x_r}[\frac{\gamma }{\gamma 1}(p^{}u_r^{}+pu_r^{}+p^{}u_r^{}\overline{p^{}u_r^{}})\hfill \\ \frac{}{}+(Q_r^{})_{\mathrm{Fourier}}]+_\mathrm{a}e_\mathrm{a}^0W_\mathrm{a}^{}=0\hfill \\ (Q_r^{})_{\mathrm{Fourier}}=\lambda T^{}/x_r(d\lambda /dT)T^{}T/x_r\hfill \end{array}\}$$ (91) $$\begin{array}{c}\mathrm{\Lambda }_\mathrm{a}^{}=\frac{\rho Y_\mathrm{a}^{}}{t}+\frac{}{x_r}\left[\rho u_r^{}Y_\mathrm{a}^{}+\rho Y_\mathrm{a}^{}u_r^{}\rho \overline{Y_\mathrm{a}^{}u_r^{}}+(M_{\mathrm{a}r}^{})_{\mathrm{Fick}}\right]\hfill \\ +\rho u_r^{}\frac{Y_\mathrm{a}}{x_r}W_\mathrm{a}^{}+\frac{\rho ^{}}{\rho }W_\mathrm{a}=0\hfill \\ (M_{\mathrm{a}r}^{})_{\mathrm{Fick}}=\rho D_\mathrm{a}\frac{Y_\mathrm{a}^{}}{x_r}\frac{d(\rho D_\mathrm{a})}{dT}T^{}\frac{Y_\mathrm{a}}{x_r}\hfill \end{array}\}$$ (92) Eqs.(80) are again separable into respective independent variables $`𝒙`$ and $`\widehat{𝒙}`$ with full nonlinear terms retained when subjected to the following separation rules: $$\begin{array}{c}\overline{Z_\alpha ^{}\widehat{Z}_\beta ^{}}=\mathrm{R}.\mathrm{P}.l^3_{\mathrm{}}^{\mathrm{}}𝑑𝒌g_\alpha (𝒌)\widehat{g}_\beta ^{}(𝒌)\hfill \\ \overline{Z_\alpha ^{}\widehat{Z}_\beta ^{}Z_\gamma ^{}}=\mathrm{R}.\mathrm{P}.l^6_{\mathrm{}}^{\mathrm{}}𝑑𝒌\widehat{g}_\beta ^{}(𝒌)_{\mathrm{}}^{\mathrm{}}𝑑\widehat{𝒌}g_\alpha (𝒌\widehat{𝒌})g_\gamma (\widehat{𝒌})\hfill \\ \overline{Z_\alpha ^{}\widehat{Z}_\beta ^{}\widehat{Z}_\gamma ^{}}=\mathrm{R}.\mathrm{P}.l^6_{\mathrm{}}^{\mathrm{}}𝑑𝒌g_\alpha (𝒌)_{\mathrm{}}^{\mathrm{}}𝑑\widehat{𝒌}\widehat{g}_\beta ^{}(𝒌\widehat{𝒌})\widehat{g}_\gamma ^{}(\widehat{𝒌})\hfill \end{array}\}$$ (93) This separation rule preserves the property that periodic part in $`g`$’s be separated out after (52), namely, $$g_\alpha (𝒙,𝒌)=e^{i𝒌𝒙}G_\alpha (𝒙,𝒌)$$ which converts Eqs.(78) in 6D space $`(𝒙,\widehat{𝒙})`$ into those in another 6D space $`(𝒙,𝒌)`$ as $$D(𝒌)G_0+_r(𝒌)G_r=0$$ (94) $$\begin{array}{c}D(𝒌)G_j+_r(𝒌)(u_r^{}G_j+G_{40}\delta _{rj}+G_{rj})\hfill \\ +(u_j^{}/x_r)G_r\rho ^1(p/x_j)G_0=0\hfill \\ G_{rj}=\mu [_j(𝒌)(\rho ^1G_r)+_r(𝒌)(\rho ^1G_j)(2/3)\delta _{jr}_k(𝒌)(\rho ^1G_k)]\hfill \\ [u_r^{}/x_j+u_j^{}/x_r(2/3)\delta _{jr}u_k^{}/x_k]\frac{G_4}{\rho R}\frac{d\mu }{dT}\hfill \\ +(1/\rho )[\mathrm{\Gamma }(G_jG_r)(\delta _{jr}/3)\mathrm{\Gamma }(G_kG_k)]\hfill \end{array}\}$$ (95) $$\begin{array}{c}\frac{1}{\gamma 1}D(𝒌)G_{40}+_r(𝒌)\left(\frac{\gamma }{\gamma 1}G_{40}u_r^{}+G_{rkk}\right)+\underset{\mathrm{a}}{}e_\mathrm{a}^0G_{\mathrm{a}W}=0\hfill \\ G_{rkk}=(\lambda /R)[_r(𝒌)(\rho ^1G_4)+p^1G_4\frac{d\lambda }{dT}RT/x_r]\hfill \\ +[\gamma /(\gamma 1)][RTG_r+\rho ^1\mathrm{\Gamma }(G_rG_4)]\hfill \end{array}\}$$ (96) $$\begin{array}{c}D(𝒌)G_\mathrm{a}+_r(𝒌)(u_r^{}G_\mathrm{a}+G_{\mathrm{a},r})+G_rY_\mathrm{a}/x_rG_{W\mathrm{a}}+\frac{G_0}{\rho }W_\mathrm{a}=0\hfill \\ G_{\mathrm{a},r}=\rho D_\mathrm{a}_r(𝒌)(\rho ^1G_\mathrm{a})\frac{d(\rho D_\mathrm{a})}{dT}\frac{G_4}{\rho R}\frac{Y_\mathrm{a}}{x_r}+\rho ^1\mathrm{\Gamma }(G_\mathrm{a}G_r)\hfill \end{array}\}$$ (97) where we have defined the following quantities, $$D(𝒌)GG/ti𝒌𝑽_pG$$ (98) $$_r(𝒌)/x_r+ik_r$$ (99) $$\mathrm{\Gamma }(G_\alpha G_\beta )_{\mathrm{}}^{\mathrm{}}G_\alpha (𝒌𝒌^{})G_\beta (𝒌^{})𝑑𝒌^{}$$ (100) $$G_{40}=G_4+RTG_0$$ (101) Eqs.(A4) through (A7) govern the ‘wave’ functions $`G_0,G_j,G_4(G_{40})`$ and $`G_\mathrm{a},G_\mathrm{b},\mathrm{}`$ from which turbulent correlations $`\overline{Z_\alpha ^{}\widehat{Z}_\beta ^{}}`$ are calculated. Relationships between the two groups of variables are the following; $$Z_\alpha ^{}=\left(\begin{array}{c}\rho ^{}\\ \rho u_j^{}\\ \rho RT^{}\\ p^{}\\ \rho Y_\mathrm{a}^{}\\ W_\mathrm{a}^{}\end{array}\right)G_\alpha =\left(\begin{array}{c}G_0\\ G_j\\ G_4\\ G_{40}\\ G_\mathrm{a}\\ G_{W\mathrm{a}}\end{array}\right)$$ (102) Turbulent transports that appear in the equations for the average are $$\begin{array}{c}\rho \overline{u_j^{}u_r^{}}=\frac{1}{\rho }\mathrm{R}.\mathrm{P}.l^3_{\mathrm{}}^{\mathrm{}}G_jG_r^{}𝑑𝒌,(\mathrm{Reynolds}\mathrm{stress})\hfill \\ \rho \overline{h^{}u_r^{}}=\frac{1}{\rho }\frac{\gamma }{\gamma 1}\mathrm{R}.\mathrm{P}.l^3_{\mathrm{}}^{\mathrm{}}G_4G_r^{}𝑑𝒌,(\mathrm{turbulent}\mathrm{heat}\mathrm{flux}\mathrm{density})\hfill \\ \rho \overline{Y_\mathrm{a}^{}u_r^{}}=\frac{1}{\rho }\mathrm{R}.\mathrm{P}.l^3_{\mathrm{}}^{\mathrm{}}G_\mathrm{a}G_r^{}𝑑𝒌,(\mathrm{turbulent}\mathrm{diffusion})\hfill \\ (\overline{T^{}_{}{}^{}2})^{1/2}=\frac{1}{\rho R}\left(l^3_{\mathrm{}}^{\mathrm{}}G_4G_4^{}𝑑𝒌\right)^{1/2},(\mathrm{turbulent}\mathrm{chemical}\mathrm{reaction})\hfill \end{array}\}$$ (103) These equations as derived above in the $`𝒌`$-space are integro-differential equations having nonlinear integrals $`\mathrm{\Gamma }`$. These integral expressions are of convolution type, so are eliminated by Fourier transform $$G_\alpha (𝒙,𝒌)=\frac{1}{(2\pi l)^3}_{\mathrm{}}^{\mathrm{}}𝑑𝒔e^{i𝒌𝒔}q_\alpha (𝒙,𝒔)$$ (104) leading to an equivalent set of equations in eddy ($`𝒔`$) space simply through a transformation rule posted in Table A. Their actual forms are: $$Dq_0+_rq_r=0$$ (105) $$\begin{array}{c}Dq_j+_r(u_r^{}q_j+q_{40}\delta _{rj}+q_{rj})+q_ru_j^{}/x_r\rho ^1(p/x_j)q_0=0\hfill \\ q_{jr}=\mu [_j(\rho ^1q_r)+_r(\rho ^1q_j)(2/3)\delta _{jr}_k(\rho ^1q_k)]\hfill \\ [u_r^{}/x_j+u_j^{}/x_r(2/3)\delta _{jr}u_k^{}/x_k]\frac{q_4}{\rho R}\frac{d\mu }{dT}\hfill \\ +\rho ^1[q_jq_r(\delta _{jr}/3)q_k^2]\hfill \end{array}\}$$ (106) $$\begin{array}{c}\frac{1}{\gamma 1}Dq_{40}+_r\left(\frac{\gamma }{\gamma 1}q_{40}u_r^{}+q_{rkk}\right)+\underset{\mathrm{a}}{}e_\mathrm{a}^0q_{\mathrm{a}W}=0\hfill \\ q_{rkk}=(\lambda /R)[_r(\rho ^1q_4)+p^1q_4\frac{d\lambda }{dT}RT/x_r]\hfill \\ +[\gamma /(\gamma 1)](RTq_r+\rho ^1q_rq_4)\hfill \end{array}\}$$ (107) $$\begin{array}{c}Dq_\mathrm{a}+_r(u_r^{}q_\mathrm{a}+q_{\mathrm{a},r})+q_rY_\mathrm{a}/x_rq_{W\mathrm{a}}+\rho ^1q_0W_\mathrm{a}=0\hfill \\ q_{\mathrm{a},r}=\rho D_\mathrm{a}_r(\rho ^1q_\mathrm{a})\frac{d(\rho D_\mathrm{a})}{dT}\frac{q_4}{\rho R}\frac{Y_\mathrm{a}}{x_r}+\rho ^1q_\mathrm{a}q_r\hfill \\ q_{W\mathrm{a}}=\pm m_\mathrm{a}q_W(\pm ;\mathrm{species}[\mathrm{a}]\mathrm{produced}/\mathrm{disappeared})\hfill \\ q_W=\frac{S}{2}\left[W_0^+W_0^{}+\frac{S}{2}(W_0^++W_0^{}W_0^+)\right]\hfill \end{array}\}$$ (108) $$q_{40}=q_4+\rho Rq_0$$ (109) ## References ## Figure captions Fig.1 Dependence of flame velocity $`u_T/u_L`$ on turbulent intensity $`u^{}/u_L`$ of methane-air premixture in elevated pressure environments. Kobayashi et al (1996). Fig.2 Failure of the classical theory ($`\delta T=0`$) in predicting ignition time $`\tau `$ at low-temperatures ($`T<1000^{}K`$), as corrected by turbulent reaction rate formula ($`\delta T>0`$). Fig.3 Turbulent flame velocity $`u_T/u_L`$ as dependent on turbulent intensity $`u^{}/u_L`$ of 9.0 percent $`H_2`$-air premixture due to a predecessor theory of the current one as compared with existing experiments, also with Arrhenius law equivalent and renormalization group theory. Fig.4 Maxwellian equilibrium $`f_0(\delta T=0)`$ and Turbulent equilibrium $`f^0(\delta T>0)`$ distribution function plotted against absolute molecular velocity, revealing appreciable difference in the population of high-energy molecules eligible for chemical reactions. Fig.5 Relative variance of reaction rates under turbulence around (a) averaged Arrhenius reaction rate and (b) turbulent (bimodal) reaction rate. ## Table captions Table 1 Molecular and turbulent transport processes Table 2 Fluctuations in transport processes Table A Conversion rules from wave-number space $`(𝒌)`$ to eddy space $`(𝒔)`$ | momentum (shear) flux density | $`p_{jl}=\mu \left({\displaystyle \frac{u_j^{}}{x_l}}+{\displaystyle \frac{u_l^{}}{x_j}}{\displaystyle \frac{2}{3}}\delta _{jl}{\displaystyle \frac{u_k^{}}{x_k}}\right)+\rho \overline{u_j^{}u_l^{}}`$ | | --- | --- | | | (Navier-Stokes’ law) (Reynolds stress) | | | $`(u_j^{}=m_j/\rho )`$ | | heatflux density | $`q_j=\lambda {\displaystyle \frac{T^{}}{x_j}}+\rho \overline{u_j^{}h^{}}`$ | | | (Fourier’s law) (turbulent heatflux density) | | | $`(T^{}=p/\rho R)`$ | | partial mass flux density | $`=\rho D{\displaystyle \frac{Y_\mathrm{a}^{}}{x_j}}+\rho \overline{u_j^{}Y_\mathrm{a}^{}}`$ | | | (Fick’s law) (turbulent diffusion) | | | $`(Y_\mathrm{a}^{}=\rho _\mathrm{a}/\rho )`$ | | chemical reaction rate | $`\begin{array}{c}W={\displaystyle \frac{1}{2}}[A(n_\mathrm{a}+\delta n_\mathrm{a})(n_\mathrm{b}+\delta n_\mathrm{b})\mathrm{exp}({\displaystyle \frac{E}{(T+\delta T)}})\hfill \\ +A(n_\mathrm{a}\delta n_\mathrm{a})(n_\mathrm{b}\delta n_\mathrm{b})\mathrm{exp}({\displaystyle \frac{E}{(T\delta T)}})]\hfill \end{array}`$ | | | (turbulence-corrected Arrhenius law) | Table 1 | fluctuations in | molecular | turbulent | | --- | --- | --- | | momentum (shear) flux density | $`\begin{array}{c}\mu [_j(\rho ^1q_r)+_r(\rho ^1q_j)\hfill \\ (2/3)\delta _{jr}_k(\rho ^1q_k)]\hfill \\ [u_r^{}/x_j+u_j^{}/x_r\hfill \\ (2/3)\delta _{jr}u_k^{}/x_k]{\displaystyle \frac{q_4}{\rho R}}{\displaystyle \frac{d\mu }{dT}}\hfill \end{array}`$ | $`+\rho ^1[q_jq_r(\delta _{jr}/3)q_k^2]`$ | | | $`\begin{array}{c}(\lambda /R)[_r(\rho ^1q_4)\hfill \\ +p^1q_4{\displaystyle \frac{d\lambda }{dT}}RT/x_r]\hfill \end{array}`$ | $`+[\gamma /(\gamma 1)](RTq_j+\rho ^1q_jq_4)`$ | | partial mass flux density | $`\begin{array}{c}\rho D_\mathrm{a}_r(\rho ^1q_\mathrm{a})\hfill \\ {\displaystyle \frac{d(\rho D_\mathrm{a})}{dT}}{\displaystyle \frac{q_4}{\rho R}}{\displaystyle \frac{Y_\mathrm{a}}{x_r}}\hfill \end{array}`$ | $`+\rho ^1q_\mathrm{a}q_r`$ | | chemical reaction rate | $`{\displaystyle \frac{S}{2}}\left[W_0^+W_0^{}+{\displaystyle \frac{S}{2}}(W_0^++W_0^{}W_0^+)\right]`$ | | Table 2 | wave number ($`𝒌`$) space | $``$ | eddy $`(𝒔)`$ space | | --- | --- | --- | | $`G_\alpha (𝒙,𝒌)`$ | $``$ | $`q_\alpha (𝒙,𝒔)`$ | | $`ik_j`$ | $``$ | $`/s_j`$ | | $`_j(𝒌)=/x_j+ik_j`$ | $``$ | $`_j=/x_j+/s_j`$ | | $`D(𝒌)G_\alpha =G_\alpha /ti𝒌𝑽_pG_\alpha `$ | $``$ | $`Dq_\alpha =q_\alpha /tV_{pr}q_\alpha /s_r`$ | | $`l^3{\displaystyle _{\mathrm{}}^{\mathrm{}}}G_\alpha (𝒌^{})G_\beta (𝒌𝒌^{})𝑑𝒌^{}`$ | $``$ | $`q_\alpha q_\beta `$ | Table A
warning/0001/cond-mat0001089.html
ar5iv
text
# Field Dependent Specific-Heat of Rare Earth Manganites ## Abstract The low temperature specific heat C(H) of several rare-earth manganites (La<sub>0.7</sub>Sr<sub>0.3</sub>MnO<sub>3</sub>, Nd<sub>0.5</sub>Sr<sub>0.5</sub>MnO<sub>3</sub>, Pr<sub>0.5</sub>Sr<sub>0.5</sub>MnO<sub>3</sub>, La<sub>0.67</sub>Ca<sub>0.33</sub>MnO<sub>3</sub>, La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub>, La<sub>0.45</sub>Ca<sub>0.55</sub>MnO<sub>3</sub> and La<sub>0.33</sub>Ca<sub>0.67</sub>MnO<sub>3</sub>) was measured as a function of magnetic field. We observed behaviour consistent with thermodynamic expectations, i.e., C(H) decreases with field for ferromagnetic metallic compounds by an amount which is in quantitative agreement with spin wave theory. We also find that C(H) increases with field in most compounds with a charge-ordered antiferromagnetic ground state. In compounds which show evidence of a coexistence of ferromagnetic metallic and antiferromagnetic charge-ordered states, C(H) displays some unusual non-equilibrium effects presumably associated with the phase-separation of the two states. We also observe a large anomalous low temperature specific heat at the doping induced metal-insulator transition (at x $`=`$ 0.50) in La<sub>1-x</sub>Ca<sub>x</sub>MnO<sub>3</sub>. The rare-earth perovskite manganites, R<sub>1-x</sub>A<sub>x</sub>MnO<sub>3</sub> (R and A are trivalent and divalent ions respectively) exhibit some intriguing features originating from the strong interplay between the electronic, magnetic and structural degrees of freedom. Recent studies have demonstrated that the ground state of these materials can be changed by varying either the band-filling or the band-width , and that the behaviour is extremely sensitive to the application of an external magnetic field leading to the so-called $`\mathrm{`}\mathrm{`}`$colossal magnetoresistance$`\mathrm{"}`$. The stoichiometric manganites (RMnO<sub>3</sub>) are antiferromagnetic insulators , but the ground state of these materials can be tuned to a metallic ferromagnetic state by increasing the band filling . At larger values of x, the ferromagnetic conducting state becomes unstable to a charge-ordered antiferromagnetic ground state with large resistivity. For some values of x, the charge-ordered state can be dissociated by an external magnetic field, leading to a large magnetoresistance of many orders of magnitude. It has been recently demonstrated that there is a coexistence of phase-separated ferromagnetic metallic (FMM) and antiferromagnetic charge-ordered (ACO) states at both values of x near the cross-over between the two ground states, and also in the lower doping regime . Measurements of specific heat provide insight into the nature of the excitations in these various phases and the phenomena responsible for the metal-insulator transitions. There have been a number of studies of the specific heat of the manganites, but these studies primarily concentrated on the temperature dependence of the specific heat, and the contributions of different excitations to the specific heat were derived from analytical fits to the data. Since the field-dependent properties of these materials are responsible for much of the recent interest, we have performed a study of the field dependence of the low temperature C(H) on a range of manganite samples to elucidate how the thermodynamic properties of rare-earth manganites evolve as a function of magnetic field. We observe that the magnetic specific heat dominates the behaviour of C(H) and that the behaviour is typically consistent with simple thermodynamic expectations. Specific heat was measured on both polycrystalline (pc) and single crystal (sc) manganite samples. We studied two samples, La<sub>0.67</sub>Ca<sub>0.33</sub>MnO<sub>3</sub> (pc) and La<sub>0.7</sub>Sr<sub>0.3</sub>MnO<sub>3</sub> (sc) with FMM ground states, three samples, Nd<sub>0.5</sub>Sr<sub>0.5</sub>MnO<sub>3</sub> (sc), Pr<sub>0.5</sub>Sr<sub>0.5</sub>MnO<sub>3</sub> (sc) and La<sub>0.33</sub>Ca<sub>0.67</sub>MnO<sub>3</sub> (pc) with ACO ground states and two samples, La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub> (pc) and La<sub>0.45</sub>Ca<sub>0.55</sub>MnO<sub>3</sub> (pc) which display strong evidence of a coexistence of FMM and ACO states. The single crystal samples were grown by floating-zone method, and the polycrystalline samples were synthesized by the solid-state reaction method. All our samples were single phase as adjudged by x-ray diffraction. Specific heat was measured by semi-adiabatic heat pulse method. All the samples were zero-field cooled to the prescribed temperature and the specific heat was measured as a function of field at a constant temperature in discrete steps of 0.25 T, as the field was swept from 0 $``$ 9 T, 9 T $``$ 0 $``$ -9 T and -9 T $``$ 0 $``$ 9 T. Each step in field took a total of $``$ 10-30 minutes, depending on the thermal mass of the samples, resulting in some non-equilibrium effects as discussed below. Moreover, the measured values of C(T) are in good agreement with those of previous studies . From simple thermodynamics, one expects the field dependence of the magnetic contribution to the specific heat to be given by: $$\left(\frac{C}{H}\right)_TT\left(\frac{^2M}{T^2}\right)_H$$ (1) For a ferromagnet in a positive field $`\frac{^2M}{T^2}`$ $`<`$ 0, thus one expects $`\frac{C}{H}`$ $`<`$ 0, i.e., C(H) decreases in a field. This can also be interpreted as the suppression of spin waves with increasing field. By contrast, for an antiferromagnet in a positive field, $`\frac{^2M}{T^2}`$ $`>`$ 0 $``$ $`\frac{C}{H}`$ $`>`$ 0, i.e., C(H) increases with increasing field, which can alternately be viewed as the softening of the AFM order parameter in a field. Figure 1 shows C(H) of La<sub>0.7</sub>Sr<sub>0.3</sub>MnO<sub>3</sub> (sc) and La<sub>0.67</sub>Ca<sub>0.33</sub>MnO<sub>3</sub> (pc) at T = 5.45 K. The specific heat of both samples decreases monotonically with increasing field, dropping by $``$ 10 $`\%`$ as the field is changed from 0 to 9T, although C(H) of La<sub>0.67</sub>Ca<sub>0.33</sub>MnO<sub>3</sub> is $``$ 18$`\%`$ higher, attributable to the larger lattice contribution due to the smaller size of Ca ion. The reduction in C(H) with increasing field can be explained as the suppression of thermal excitations of the spin waves as described above. Similar observations were also drawn from the recent temperature dependent specific-heat experiments on La<sub>1-x</sub>Sr<sub>x</sub>MnO<sub>3</sub>. The magnon contributions to the specific heat of a ferromagnet in an external field neglecting any demagnetization effects, and assuming no spin gap at H = 0, so that the dispersion relation at an external field H is $`\omega =g\mu _BH+Dk^2`$, can be written as , $$C_{magnon}(H)=\frac{k_B^{5/2}T^{3/2}}{4\pi ^2D^{3/2}}\underset{g\mu _BH/k_BT}{\overset{\mathrm{}}{}}\frac{x^2e^x}{(e^x1)^2}\sqrt{x\frac{g\mu _BH}{k_BT}}𝑑x$$ (2) where $`x=(g\mu _BH+Dk^2)/k_BT`$, and D = 2JS$`a^2`$ is the spin stiffness constant. We can thus calculate the spin-wave contributions to the specific heat of these two samples from calculations based on Eq. 2, and using stiffness constant and unit cell length from previous neutron scattering and x-ray scattering data respectively. As shown by the solid lines in figure 1, the spin wave calculation fits the C(H) data extremely well with only a single free parameter – a constant offset which accounts for the non-magnetic contributions to the heat capacity. Figure 2 and 3 show the low temperature C(H) of single crystals of Pr<sub>0.5</sub>Sr<sub>0.5</sub>MnO<sub>3</sub> (PSMO) and Nd<sub>0.5</sub>Sr<sub>0.5</sub>MnO<sub>3</sub> (NSMO). Recent neutron and x-ray scattering experiments have suggested that at low temperatures while the orbital ordering of d$`{}_{}{}^{2}{}_{x^2}{}^{}`$-d$`{}_{}{}^{2}{}_{y^2}{}^{}`$ leads to A-type AFM state in PSMO with no clear evidence for a charge ordering, d$`_{z^2}`$ orbitals order in NSMO resulting in a CE-type ($`\pi `$, 0, $`\pi `$) type charge ordered AFM state. However, PSMO undergoes a sharp rise in resistivity with an accompanying drop in magnetization, which is usually associated with charge-ordering. At moderate fields the low temperature C(H) of both PSMO and NSMO increases monotonically with increasing field in contrast to the FM samples, but consistent with the expectations for the AFM materials. The low temperature C(H) of PSMO (figure 2) increases monotonically with field at low fields due to the softening of the stiffness constant in the AFM state. However, in the region (4 T $``$ H $``$ 8 T), C(H) has a reduced slope, suggesting that the AFM $``$ FM transition which occurs at $``$ 5 T is not completed until H $``$ 8 T. At higher fields (H $``$ 8 T), in contrast to NSMO (see below) we believe that the change in C(H) with field is electronic in nature, and the steep rise in C(H) with increasing field, which coincides with the sharp drop in $`\rho (H)`$ is due to the enhancement of the free carriers as the FM state is percolated throughout the sample. This sample also shows some non-equilibrium effects, e.g., C(H) increases by 8 $`\%`$ in $``$ 10 mins immediately after the field is raised to 9 T, even when the external parameters, such as temperature and the field remained unchanged. This kind of non-equilibrium behaviour is much more pronounced in the sample which shows strong evidence of a coexistence of FM and charge-ordered state, as discussed below. The gradual relaxation of the FM state in this sample is also possibly responsible for the smaller C(H) (by $``$ 10$`\%`$) at low fields during subsequent field sweeps. The low temperature specific heat of NSMO is an order of magnitude higher than the rest of the measured manganites, this is possibly associated with Schottky-like anomaly connected with the large magnetic moment (J $`=`$ 9/2) of the Nd<sup>3+</sup> ion . The low temperature C(H) of NSMO increases monotonically with field at H $``$ 7 T, demonstrating a stable AFM state. At H $``$ 7 T, the slope of C(H) changes, and C(H) displays a sharp downward turn (see figure 3), suggesting the suppression of thermal excitation of the spin-waves as the sample approaches an AFM $``$ FM transition. During subsequent field sweeps M(H) shows no hysteresis, although $`\rho (H)`$ remains order of magnitude smaller, indicating that the carriers remain delocalized even when the field is removed. Since C(H) shows no hysteresis despite the increase in the population of free carriers , this further suggests that the magnetic-contribution dominates the low temperature heat capacity. Figure 4 shows the low temperature C(H) of La<sub>0.33</sub>Ca<sub>0.67</sub>MnO<sub>3</sub>. This sample exhibits all the features of an ACO ground state, with magnetization reaching only $``$ 2 $`\%`$ of the total saturation magnetization even at H $`=`$ 7 T. However, the behaviour of C(H) is quite different from the other AFM CO samples we studied. The low temperature C(H) of this sample exhibits features which are rather inconsistent with the thermodynamical expectations for an AFM state, i.e, C(H) decreases monotonically with increasing field, although the slope of C(H) is not as sharp as that of the FMM samples. However, when the field is increased from 0 to 9 T, C(H) drops by only 3.5 mJ. The absolute magnitude of this change is at least a factor of 3 smaller than that of the other measured AFM samples, but quantitatively and qualitatively similar to the equilibrium C(H) of the phase-separated samples, as discussed below. This uncharacteristic behaviour of C(H) in the AFM state can not be attributed to the magnetic excitations, and is perhaps associated with charge/orbital ordering of the La<sub>1-x</sub>Ca<sub>x</sub>MnO<sub>3</sub> compounds. Figure 5 and 6 show low temperature C(H) of polycrystalline sample of La<sub>0.50</sub>Ca<sub>0.5</sub>MnO<sub>3</sub> (Mn$`{}_{}{}^{4+}\%`$ = 53.8) and La<sub>0.45</sub>Ca<sub>0.55</sub>MnO<sub>3</sub> (Mn$`{}_{}{}^{4+}\%`$ = 58.2) with well characterized Mn<sup>4+</sup> content. These samples show strong evidence of phase separation into FMM and insulating ACO states . The charge-ordered state of the sample with x $`=`$ 0.50 can be partially dissociated by a moderate external magnetic field, greatly reducing the resistivity, but the sample with x $`=`$ 0.55 remains largely charge-ordered showing very little magnetoresistance even at H $`=`$ 9 T. We find that the behaviour of C(H) in the phase-separated samples is more complex than the others we have studied, displaying significant non-equilibrium effects. The low temperature zero-field cooled C(H) of La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub> decreases with increasing field, when the field is swept for the first time from H $`=`$ 0 $``$ 9 T. After reaching 9 T, however, C(H) increases by $``$ 19$`\%`$ in $``$ 10 mins even when field is kept constant at 9 T. On decreasing the field, C(H) not only remains 17$`\%`$ higher than the initial sweep but is almost constant for H $``$ 3 T. However, at lower fields, C(H) drops sharply at H $``$ 3 T with a minimum at H $``$ 1 T, and then recovers to near its maximum value at H $``$ -1 T. Although the low temperature C(H) changes by only $``$ 3$`\%`$ on increasing the field in the reverse direction for H $``$ -6 T, C(H) drops sharply at H $``$ -6 T, and similar features to those of the initial sweep are observed for H $``$ -6 T and also during subsequent field sweep from H = -9 T $``$ 9 T. To characterize equilibrium behaviour of C(H), we also performed the same measurements in an equilibrium mode, i.e., data were taken after waiting for 1 hour at every field, ensuring that the system had reached equilibrium at that field by monitoring C as a function of time at every field. We observe that the data taken in this equilibrium mode displays only $``$ 2$`\%`$ rise in C(H) at H $``$ 3 T, but at higher fields remains almost flat up to 9 T. The sharp rise in C after the field is first swept to 9 T suggests that the FM regimes are growing with time at the expense of the ACO phase. This increase in C could then be attributable to an increased population of free carriers or as the enhancement of long wavelength low energy spin-wave excitations with the percolation of the small FM regimes. The minima we observe in C(H) at $``$ 1 T when sweeping the field could similarly be explained by the gradual domain formation or a time-dependent increase in the fraction of the sample which is phase separated into the ACO phase. That C(H) measured in equilibrium changes very little with field is consistent with behaviour of the La<sub>0.33</sub>Ca<sub>0.67</sub>MnO<sub>3</sub> sample. This suggests that the magnetic heat capacity of the CO state in La<sub>1-x</sub>Ca<sub>x</sub>MnO<sub>3</sub> is much smaller than that of the other ACO samples studied. This is perhaps due to the microscopic phase-separation which inhibits the formation of long wavelength spin modes in La<sub>1-x</sub>Ca<sub>x</sub>MnO<sub>3</sub> systems . The charge-lattice of La<sub>0.45</sub>Ca<sub>0.55</sub>MnO<sub>3</sub> remains primarily intact even at H $``$ 9 T. Although magnetization reveals the presence of small clusters of ferromagnetism, which undergo a first order transition to charge-ordering at a lower temperature than the bulk of the sample, a FM state is not established throughout the sample even at H $`=`$ 9 T . On increasing the field, the low temperature C(H) increases by 7$`\%`$ but C(H) displays a small downward turn at H $``$ 4 T before finally dropping sharply at H $``$ 6 T. After reaching 9 T, C(H) rises by 26$`\%`$ in $``$ 10 minutes even when the external parameters remained unchanged. Similar features are also observed during subsequent field sweeps though C(H) shows no non-equilibrium effect. The non-equilibrium effects appear to be intrinsic to the samples which display a coexistence of small clusters of ferromagnetism in the primarily charge-ordered regime, although the extent of this effect decreases for samples far from x $`=`$ 0.50. One other anomalous feature of the heat capacity of the La<sub>0.5</sub>Ca<sub>0.5</sub> MnO<sub>3</sub> sample is that the magnitude of C(H=0) is much larger than that of La<sub>0.67</sub>Ca<sub>0.33</sub>MnO<sub>3</sub> or La<sub>0.33</sub>Ca<sub>0.67</sub>MnO<sub>3</sub>. In fact the enhancement of heat capacity at x $`=`$ 0.50 has been previously noted, and was suggested to be due to an extra contribution to C by orbital excitations of the d$`_{z^2}`$ orbital ordering . Since we have available samples of La<sub>1-x</sub>Ca<sub>x</sub>MnO<sub>3</sub> for a wide range of x , we also studied this phenomenon briefly with the results plotted in figure 7. We see that there is a sharp rise in C(x) for x $`=`$ 0.5 with some scatter around the maximum. This scatter is greatly reduced, however when the data are plotted as a function of the actual Mn<sup>4+</sup> content of the samples (as determined by redox titration). This maximum appears to be associated with the doping-induced metal-insulator transition at x $``$ 0.50, and since it is not reflected in the field-dependence of the heat capacity, and since the lattice properties of these materials are quite similar. We hypothesize that the enhancement of C(x) near x $`=`$ 0.50 is associated with an enhancement of the electron mass near the MIT, but there is no detailed theory to support this possibility. In conclusion, we have measured low temperature field dependent specific heat of a range of rare earth manganites. Our data suggest that the behaviour of C(H) is dominated by the magnetic contributions to the specific heat regardless of the nature of the magnetic or electronic state of these materials. Moreover, we observed that the slope of C(H) with respect to the field for both ferromagnetic and antiferromagnetic samples is typically consistent with the thermodynamic expectations. We also observed that C(H) of compounds which exhibit evidence of a coexistence of ferromagnetic metallic and antiferromagnetic charge-ordered states show some unusual non-equilibrium effects presumably associated with the phase-separation of the two states. We are grateful to Dr. A. P. Ramirez for many enlightening discussions. This research has been supported by NSF grant DMR 97-01548, the Alfred P. Sloan Foundation and the Dept. of Energy, Basic Energy Sciences-Materials Sciences under contract $`\mathrm{\#}`$W-31-109-ENG-38.
warning/0001/physics0001014.html
ar5iv
text
# 1 Low-energy bands of the porphyrin spectra for 𝐶⁢𝐻₂⁢𝐶⁢𝑙₂ as solvent. PHOTO-INDUCED INTERMOLECULAR CHARGE TRANSFER IN PORPHYRIN COMPLEXES Michael Schreiber, Dmitry Kilin, and Ulrich Kleinekathöfer Institut für Physik, Technische Universität, D-09107 Chemnitz, Germany Optical excitation of the sequential supermolecule $`H_2PZnPQ`$ induces an electron transfer from the free-base porphyrin ($`H_2P`$) to the quinone ($`Q`$) via the zinc porphyrin ($`ZnP`$). This process is modeled by equations of motion for the reduced density matrix which are solved numerically and approximately analytically. These two solutions agree very well in a great region of parameter space. It is shown that for the majority of solvents the electron transfer occurs with the superexchange mechanism. I. INTRODUCTION The investigation of photoinduced charge transfer is important both for the description of natural photosynthesis \[?\] and for the creation of artificial photoenergy-converting devices \[?\]. For experimental realizations of such artificial devices porphyrin complexes are good candidates \[?,?,?\]. Of major interest are those complexes with an additional bridging block between donor and acceptor \[?,?,?,?,?\]. Electron transfer reactions can occur through different mechanism \[?,?,?\]: sequential transfer (ST) or superexchange (SE). Changing a building block of the complex \[?,?\] or changing the environment \[?\] can modify which mechanism is most significant. To clarify which mechanism is present one sequentially varies the energetics of the complex \[?,?,?,?\]. This is done by radical substituting the porphyrin complexes \[?,?,?\] or by changing the polarity of the solvent \[?,?\]. Also the geometry and size of a bridging block can be varied and in this way the length of the subsystem through which the electron has to be transfered \[?,?,?,?\]. SE \[?\] occurs due to coherent mixing of the levels \[?,?,?,?\] and plays a role for any detuning of the energy levels \[?,?,?,?\]. The transfer rate in this channel decreases exponentially with increasing length of the bridge \[?,?\]. When incoherent effects such as dissipation and dephasing dominate \[?,?\], the transfer is mainly sequential \[?,?\], i. e., the levels are occupied mainly in sequential order \[?,?\]. An increase in the bridge length induces only a small reduction in the transfer rate \[?,?\]. In the case of coherent SE the dynamics is mainly Hamiltonian and can be described on the basis of the Schrödinger equation. The physically important results can be obtained by perturbation theory \[?\], most successfully by the Marcus theory \[?\]. In case of ST the environmental influence has to be taken into account. The more natural description of the relaxation process is based on the density matrix (DM) formalism \[?,?,?,?,?,?,?\]. The master equation that governs the DM evolution as well as the appropriate relaxation coefficients can be derived from such basic information as system-environment coupling strength and spectral density of the environment \[?,?,?\]. The main physics of the system can be described by a DM equation which accounts for relaxation effects phenomenologically \[?,?\]. The master equation is analytically solvable only for the simplest models \[?,?\]. Most investigations are based on the numerical solution of this equation \[?,?,?\]. However, an estimations can be obtained within the steady-state approximation \[?\]. Here we perform numerical as well as approximate analytical calculations. II. MODEL We investigate the photoinduced electron transfer in supermolecules that consist of sequentially connected molecular blocks, namely donor, bridge, and acceptor. The donor (D) is not able to transfer its charge directly to the acceptor (A) because of their spatial separation. D and A can exchange their charges only through B (Fig. 1). In the present investigation the supermolecule consists of free-base porphyrin($`H_2P`$) as donor, zinc substituted porphyrin($`ZnP`$) as bridge, and benzoquinone as acceptor \[?\]. In each of those molecular blocks we consider only two molecular orbitals, the LUMO and the HOMO. Each of those orbitals can be occupied by an electron ($`|1`$) or not ($`|0`$). This model allows us to describe the neutral nonexcited molecule $`|1_{HOMO}|0_{LUMO}`$ and the following three states of the molecule: neutral excited molecule $`|0_{HOMO}|1_{LUMO}`$, positive ion $`|0_{HOMO}|0_{LUMO},`$ and negative ion $`|1_{HOMO}|1_{LUMO}`$. Below Roman indices indicate molecular orbitals ($`m=0`$ \- HOMO, $`m=1`$ \- LUMO), while Greek indices indicate molecular blocks ($`\mu =1`$ \- donor, $`\mu =2`$ \- bridge, $`\mu =3`$ \- acceptor). Each of the electronic states has its own vibrational substructure. However the time of vibrational relaxation \[?\] is two orders of magnitude faster than the characteristic time of the electron transfer \[?\]. Because of this we assume that only the vibrational ground states play a dominant role in electron transfer. One can describe the occupation of an orbital by an electron with the appropriate creation operator $`c_{\mu m}^+=|1_{\mu m}0|_{\mu m}`$ as well as its annihilation $`c_{\mu m}=|0_{\mu m}1|_{\mu m}`$. Then $`\widehat{n}_\mu =_mc_{\mu m}^+c_{\mu m}`$ gives the number of electrons in the molecular block $`\mu `$. For the description of charge transfer and other dynamical processes in the system we introduce the Hamiltonian $`\widehat{H}=\widehat{H}_S+\widehat{H}_B+\widehat{H}_{SB},`$ (1) where $`H_S`$ characterizes the supermolecule, $`H_B`$ the dissipative bath, and $`H_{SB}`$ the interaction between the two. $`H_S`$, however, includes the static influence of the environment, namely of the solvent dipoles, which gives rise to a reduction of the energy levels, $`\widehat{H}_S={\displaystyle \underset{\mu m}{}}E_{\mu m}\widehat{n}_{\mu m}+{\displaystyle \frac{3}{ϵ_s+2}}\left(\widehat{E}_{el}+\widehat{E}_{ion}\right)+\widehat{V},.`$ (2) The energies $`E_{\mu m}`$ are calculated in the independent particle approximation \[?\]. $`ϵ_s`$ denotes the static dielectric constant of the solvent. $`\widehat{E}_{el}=_\mu (\widehat{n}_\mu 1)e^2/(4ϵ_0r_\mu )`$ describes the energy to create an isolated ion. This term depends on the characteristic radius $`r_\mu `$ of the molecular blocks. $`\widehat{E}_{ion}=_\mu _\nu (\widehat{n}_\mu 1)(\widehat{n}_\nu 1)e^2/(4\pi ϵ_0r_{\mu \nu })`$ includes the interaction between the already created ions. It depends on the distance between the molecular blocks $`r_{\mu \nu }`$. The last contribution to the system Hamiltonian is the hopping term $`\widehat{V}=_{\mu \nu }v_{\mu \nu }(\widehat{V}_{\mu \nu }^++\widehat{V}_{\mu \nu }^{})((\widehat{n}_\mu 1)^2+(\widehat{n}_\nu 1)^2),`$ which includes the coherent hopping between each pair of LUMO $`\widehat{V}_{\mu \nu }^{}=c_{\nu 1}^+c_{\mu 1}`$, $`\widehat{V}^+=(\widehat{V}^{})^+`$ as well as the corresponding intensities $`v_{\mu \nu }`$. The matrix elements of this operator give nonzero contribution only if one of the states has a charge separation. Because there is no direct connection between donor and acceptor we assume $`v_{13}=0`$. As usual the bath is given by harmonic oscillators with creation and anhilation operators $`a_\lambda ^+`$ and $`a_\lambda `$. The system bath interaction comprises both irradiative and radiative transitions. For $`t110ns`$ the latter one can be neglected . The irradiative contribution corresponds to energy transfer to the solvent and spreading of energy over vibrational modes of the supermolecule $`\widehat{H}_{SB}={\displaystyle \underset{\lambda }{}}{\displaystyle \underset{\mu \nu }{}}K_{\lambda ,\mu \nu }v_{\mu \nu }\left(a_\lambda ^++a_\lambda \right)\left(\widehat{V}_{\mu \nu }^++\widehat{V}_{\mu \nu }^{}\right),`$ (3) where $`K_{\lambda ,\mu \nu }`$ reflects the interaction strength between bath mode $`\lambda `$ and quantum transition between LUMO levels of molecules $`\mu `$ and $`\nu `$. Initially we use the whole density matrix of system and bath for the description of the dynamics. After applying the Markov and rotating wave approximations and tracing out the bath modes \[?\] we obtain the equation of motion for the reduced density matrix (RDM) $`\dot{\sigma }=i/\mathrm{}[\widehat{H}_S,\sigma ]`$ $`+`$ $`{\displaystyle \underset{\mu \nu }{}}\mathrm{\Gamma }_{\mu \nu }\{(n\left(\omega _{\mu 1\nu 1}\right)+1)([\widehat{V}_{\mu \nu }^{}\sigma ,\widehat{V}_{\mu \nu }^+]+[\widehat{V}_{\mu \nu }^{},\sigma \widehat{V}_{\mu \nu }^+])`$ (4) $`+n\left(\omega _{\mu 1\nu 1}\right)([\widehat{V}_{\mu \nu }^+\sigma ,\widehat{V}_{\mu \nu }^{}]+[\widehat{V}_{\mu \nu }^+,\sigma \widehat{V}_{\mu \nu }^{}])\},`$ where the dissipation intensity $`\mathrm{\Gamma }_{\mu \nu }=\pi K_{\mu \nu }^2\rho (\omega _{\mu 1\nu 1})v_{\mu \nu }^2`$ depends on the coupling $`K_{\mu \nu }`$ of the transition $`\mu 1\nu 1`$ and on the bath mode of the same frequency. $`\mathrm{\Gamma }_{\mu \nu }`$ depends also on the density $`\rho `$ of bath modes at the transition frequency $`\omega _{\mu 1\nu 1}`$ and on the corresponding coherent coupling $`v_{\mu \nu }`$ between the system states. $`n(\omega )`$ denotes Bose-Einstein distribution. For simplicity we introduce a superindex $`i=\{\mu m\}`$, the intensities of the dissipative transitions $`d_{ij}=\mathrm{\Gamma }_{ij}n(\omega _{ij})`$ between each pair of states, as well as the corresponding dephasing intensities $`\gamma _{ij}=1/2_k(d_{ik}+d_{kj})`$. Taking these simplifications into account one gets $`\dot{\sigma }_{ii}`$ $`=`$ $`i/\mathrm{}{\displaystyle \underset{j}{}}\left(V_{ij}\sigma _{ji}\sigma _{ij}V_{ji}\right){\displaystyle \underset{i}{}}d_{ij}\sigma _{ii}+{\displaystyle \underset{j}{}}d_{ji}\sigma _{jj},`$ (5) $`\dot{\sigma }_{ij}`$ $`=`$ $`\left(i\omega _{ij}\gamma _{ij}\right)\sigma _{ij}i/\mathrm{}V_{ij}\left(\sigma _{jj}\sigma _{ii}\right).`$ (6) The simplification is that we do not calculate the system parameters, rather we extract them from experimental data. III. EXTRACTION OF SYSTEM PARAMETERS The porphyrin absorption spectra \[?\] consist of high frequency Soret bands and low frequency $`Q`$ bands. In case of $`ZnP`$ the $`Q`$ band has two subbands, $`Q(0,0)`$ and $`Q(1,0)`$. In the free-base porphyrin $`H_2P`$ the reduction of symmetry induces a splitting of each subband into two, namely $`Q^x(0,0)`$, $`Q^y(0,0)`$ and $`Q^x(1,0)`$, $`Q^y(1,0)`$. So the emission spectra of $`ZnP`$ and $`H_2P`$ consist of two and four bands, respectively. Each of the abovementioned spectra can be represented as a sum of Lorentzians with good precision. It is important to note that the spectra of porphyrin complexes contain all bands of the isolated porphyrins without essential changes. We use the lowest band of each spectrum. The corresponding frequencies and widths are shown in table 1. On the basis of the experimental spectra we determine $`E_{D^{}BA}=1.82eV`$ and $`E_{DB^{}A}=2.03eV`$ (in $`CH_2Cl_2`$). The authors of Ref. ? give the energies of two other levels, $`E_{D^+B^{}A}=2.44eV`$ and $`E_{D^+BA^{}}=1.42eV`$. This allows to calculate $`E_{DB^+A^{}}=1.21eV`$. The hopping intensity $`v_{23}=v=2.2meV`$ is calculated in Ref. ?. On the other hand Rempel et al. \[?\] estimate the electron coupling of the initially excited and charged bridge states $`v_{12}=V=65meV`$. We take the intensity of the intermolecular conversions $`\mathrm{\Gamma }_{21}`$, $`\mathrm{\Gamma }_{23}`$ in range $`110\times 10^{11}s^1`$ \[?\]. The main parameter which controls the electron transfer in a triad is the relative energy of the state $`D^+B^{}A`$. This state has a strong coupling to the solvent that changes the energy of the state. The values of the energy $`E_{D^+B^{}A}`$ calculated in the present model are shown in table 2 for some solvents. In table 2 $`ϵ_s`$ denotes the static dielectric permittivity, $`ϵ_{\mathrm{}}`$ the optic dielectric permittivity, MTHF 2-methil-tetrahydrofuran, and CYCLO denotes cyclohexane. The calculated value $`E_{D^+B^{}A}=2.86eV`$ deviates $`15\%`$ from the data of Ref. ?. IV. RESULTS The time evolution of charge transfer within the supermolecule is described by Eqs. (5) and (6). At initial time only the donor state is occupied. The calculations were performed with two methods, direct numerical integration and analytic approximation. For the numerical simulation the eigenvalues and -vectors of the system are calculated and with these the time evolution of the system is known. The simulation of the system dynamics with the parameters determined in the previous section shows exponential growth of the acceptor population. Such a behavior can be accurately fitted to the formula $`P_3(t)=P_3(\mathrm{})[1\mathrm{exp}(k_{ET}t)],`$ where $`k_{ET}5\times 10^9s^1`$ and $`P_3(\mathrm{})0.95`$ for $`CH_2Cl_2`$ as solvent. The population of the bridge state does not exceed $`0.005`$. This shows that the SE mechanism dominates over the ST for the chosen set of parameters. In this case the system dynamics can be described by two values: the acceptor population at infinite time $`P_3(\mathrm{})`$ and the reaction rate $`k_{ET}`$ that we deduce from the dynamics via the following formula $`k_{ET}=P_3(\mathrm{})/\{_0^{\mathrm{}}[1P_3(t)]𝑑t\}`$. The analytical approach is valid for the kinetic limit $`t1/\gamma _{ij}`$. In Laplace-space we can replace $`1/(i\omega _{ij}+\gamma _{ij}+s)`$ by $`1/(i\omega _{ij}+\gamma _{ij})`$, where $`s`$ denotes the Laplace variable. This allows to simplify Eqs. (5) and (6) and we define a new relaxation operator $`(L\sigma )_{ii}^{new}=_ig_{ij}\sigma _{ii}+_jg_{ji}\sigma _{jj}.`$ In this expression the transition coefficients $`g_{ij}`$ contain both, dissipative and coherent contributions $$g_{ij}=d_{ij}+v_{ij}v_{ji}\gamma _{ij}/\left[\mathrm{}^2\left(\omega _{ij}^2+\gamma _{ij}^2\right)\right].$$ (7) Assuming the bridge population to be zero allows us to find the dynamics of the acceptor state in the form $`P_3(t)=P_3(\mathrm{})[1\mathrm{exp}(k_{ET}t)],`$ where the final population $`P_3(\mathrm{})`$ and the reaction rate $`k_{ET}`$ are expressed in terms of the coefficients $`g_{ij}`$ $`k_{ET}=g_{23}+{\displaystyle \frac{g_{23}\left(g_{12}g_{32}\right)}{g_{21}+g_{23}}},P_3\left(\mathrm{}\right)={\displaystyle \frac{g_{12}g_{23}}{g_{21}+g_{23}}}\left(k_{ET}\right)^1.`$ (8) V. DISCUSSION The following question will now be discussed: How does the mechanism and speed of the reaction depend on a deviation of the parameters from the determined values? Namely which parameters have to be changed in order to change not only the reaction rate quantitatively, but the dominant mechanism of reaction and the qualitative behavior of dynamics at all. To answer these questions we calculate the system dynamics while varying one parameter at a time and keeping the other parameters unchanged. The dependencies of transfer rate $`k_{ET}`$ and final population $`P_3(\mathrm{})`$ on coherent couplings $`V=v_{12}`$, $`v=v_{23}`$ and dissipation intensities $`\mathrm{\Gamma }=\mathrm{\Gamma }_{21}`$, $`\gamma =\mathrm{\Gamma }_{23}`$ are shown in Fig. 2. In particular, the decrease of the coherent coupling $`V`$ induces a quadratic decrease of the reaction rate $`k_{ET}`$ until saturation $`V10^{10}ps^1`$. Then $`k_{ET}`$ reaches its lower bound and does not depend on $`V`$ anymore. This corresponds to the crossover of the reaction mechanism from SE mechanism to ST. But, due to the big energy difference between donor and bridge state the efficiency of this ST is extremely low, i. e., $`P_30`$. The considered variation of the coherent coupling can be experimentally performed by exchanging building blocks in the supermolecule. The most crucial change in the reaction dynamics can be induced by changing the energies of the system levels. As discussed above this can be done by altering the solvent. Most important is the relative energy of the bridge state $`|D^+B^{}A`$. The results of the corresponding calculations are presented in Fig. 3. For high energies of the bridge state $`E_{D^+B^{}A}E_{D^{}BA}`$ the numerical and analytical results do not differ from each other. The reaction occurs with the SE mechanism that coincides with the conclusion of Ref. ?. This is the case for the most of solvents (see table 2). The smooth decrease of energy induces an increase of the reaction rate up to the maximal value near $`1ps^1`$. While the bridge energy approaches the energy of the donor state the ST mechanism starts to contribute to the process. As can be seen in table 2 this regime can be reached by the use strong polar solvents. The analytical solution does not coincide with the numerical one anymore because the used approximations are no more valid in this region. In the case $`E_{D^+B^{}A}<E_{D^+BA^{}}`$ one cannot approximate the dynamics of the acceptor population in the form $`P_3[1\mathrm{exp}(k_{ET}t)]`$. A high value of the bridge energy ensures the transition of the whole population to the acceptor state $`|D^+BA^{}`$. In the intermediate case, when the bridge state has the same energy as the acceptor state, the final population spreads itself over these two states $`P_3(\mathrm{})=0.5`$. At even lower bridge energies the population gets trapped at the bridge state. We performed calculations for the electron transfer in the supermolecular complex $`H_2PZnPQ`$ within the RDM formalism. The resulting analytical and numerical reaction rates are in good agreement with each other and in qualitative correspondence with experimental data \[?,?,?\]. The SE mechanism of electron transfer dominates over the sequential one. The qualitative character of the transfer reaction is stable with respect to a small variation of the system parameter. The crossover between the reaction mechanisms can be forced by lowering the bridge state energy to the energy of the donor state. REFERENCES
warning/0001/cond-mat0001104.html
ar5iv
text
# Decay on several sorts of heterogeneous centers: Special monodisperse approximation in the situation of strong unsymmetry. 1. General results ## 1 Introduction Metastable phase decay on the several types of heterogeneous centers remains a rather actual problem for theoretical investigation. For the first time the theory for the kinetics description was constructed in . This approach decomposes the general situation into characteristic situations which are rather simple. All limit situations characterized by small values of characteristic parameters can be solved by the slightly modified versions of the iteration method initially proposed in . Only the intermediate situation requires a special method of consideration which is based on the special monodisperse approximation. When the total number of the heterogeneous centers is one and the same for different types of centers then as it is shown in there are only two characteristic situations: the intermediate situation and the situation of the strong unsymmetry. The special monodisperse approximation is well based and can be spread on a more general situation. This generalization is important not only to present the description in a more compact form. The special monodisperse approximation allows to reduce the error appeared in the limit situations. Really, in the situation of the strong unsymmetry (this is a standard limit situation) one has to use the monodisperse approximation. If we use the special monodisperse approximation instead of the already used one we shall reduce the error. The universal character of obtained solution was shown in . The general recipe to use the special monodisperse approximation was suggested in in the abstract manner. So, now it rather natural to show how the special monodisperse approximation works in the limit situations concretely. This publication is intended to show how to use the special monodisperse approximation in the limit situations. We shall use for example the situation of the strong unsymmetry which appeared when the number of the first type centers equals to the number of the second type centers (see ). This situation together with the intermediate situation completes the consideration of the general case . ## 2 Formulation of the problem Consider the system with two sorts of heterogeneous centers (they are marked by subscripts $`1`$ and $`2`$). Suppose that the total numbers of heterogeneous centers $`\eta _{tot1}`$ and $`\eta _{tot2}`$ are equal $$\eta _{tot1}=\eta _{tot2}$$ (1) At the initial moment of time which is denoted by the subscript $``$ there exists only heterogeneous centers. We shall call the period of intensive formation of the droplets on the heterogeneous centers of given type as the nucleation on the centers of given type. We shall use in the estimates some characteristic values. Denote by $`\mathrm{\Delta }_1t`$ the duration of the nucleation on the first type centers and by $`\mathrm{\Delta }_2t`$ the duration of the nucleation on the second type centers. Consider the situation when the rate of nucleation on the first type of heterogeneous centers strongly exceeds the rate of nucleation on the second type of heterogeneous centers. We shall choose the sorts of heterogeneous centers to have $$f_1f_2$$ where $`f_i`$ is the amplitude value of the stationary distribution function. Due to (1) the last inequality is practically equivalent to $$\mathrm{exp}(\mathrm{\Delta }F_1)\mathrm{exp}(\mathrm{\Delta }F_2)$$ where $`\mathrm{\Delta }F`$ is the height of the activation barrier. The power of metastability is characterized by the supersaturation $$\zeta =(nn_{\mathrm{}})/n_{\mathrm{}}$$ where $`n`$ is the molecules number density of the vapor, $`n_{\mathrm{}}`$ is the molecule number density of the saturated vapor. Then the balance of the substance in the closed substance requires $$\zeta _{}=\zeta +G_1+G_2$$ where $`G_1`$ is the number of the molecules in the liquid phase around the first type centers (taken in units of $`n_{\mathrm{}}`$), $`G_2`$ is the number of molecules in the droplets formed on the second type centers. We shall describe the droplet by the value of dimensionless radius $`\rho `$ which equals to the cube root of the number $`\nu `$ of the molecules inside the droplet $$\rho =\nu ^{1/3}$$ This value is convenient because the rate of its growth is one and the same for all sizes $$\frac{d\rho }{dt}=\zeta /\tau $$ where the constant $`\tau `$ is the characteristic time. The last equation is valid for the supercritical droplets under the free molecular regime of the substance exchange. Then automatically one can see that the distribution $`f_i(\rho ,t)`$ of the droplets over $`\rho `$ in the moment $`t`$ depends only on the intensity of the droplets formation in the time when the droplet of given size were formed. The application of the ordinary monodisperse approximation (”total monodisperse approximation”) is based on the following explanation: \- Suppose we can suggest an approximation $$G_1N_{1tot}\rho _0^3/n_{\mathrm{}}$$ where $`N_{1tot}`$ is the total number of the droplets formed on the first type centers during the condensation process. The last approximation is good when $`\rho _0`$ is many times grater than the value of $`\rho _0`$ at the end of the period of intensive formation of the droplets on the first type centers, i.e. $`\rho _0(\mathrm{\Delta }_1t)`$: $$\rho _0(t)\rho _0(\mathrm{\Delta }_1t)$$ This approximation is important when $`G_1`$ stops the formation of the droplets on the second type centers, i.e. at $`\mathrm{\Delta }_2t`$. So, it is necessary to be $$\rho _0(\mathrm{\Delta }_1t)\rho _0(\mathrm{\Delta }_2t)$$ As far as $`d\rho /dt`$ is rather smooth function of time one can rewrite the last estimate as $$\mathrm{\Delta }_1t\mathrm{\Delta }_2t$$ (2) One can estimate the number $`N_{itot}`$ of the droplets formed on the centers of given sort as $$N_{itot}J_i\mathrm{\Delta }_it$$ where $$J_if_in_{\mathrm{}}\tau /\zeta $$ is the initial rate of nucleation<sup>1</sup><sup>1</sup>1$`f_i`$ is expressed in units of $`n_{\mathrm{}}`$. Then the violation of (2) means that $$N_{1tot}N_{2tot}$$ As the result the number of the droplets formed on the second type centers is negligible. The negligible value of $`N_{2tot}`$ in the unique situation when the ordinary monodisperse approximation fails was the ground to apply this approximation in the situation of the strong unsymmetry . But we see that if we are interested in the value of $`N_{2tot}`$ without any respect to $`N_{1tot}`$ then the question is still open. Below we shall resolve this problem. ## 3 Special monodisperse approximation For $`G_i`$ one can write the following relation $$G_i=_0^{\mathrm{}}\rho ^3f_i(\rho ,t)𝑑\rho $$ Now one can analyze the subintegral function $$g_i(\rho ,t)=\rho ^3f_i(\rho ,t)$$ which has the sense only for positive $`\rho `$. One can see that $$g_i(\rho ,t)=0$$ for $$\rho >\rho _0(t)_0^t\frac{\zeta (t^{})}{\tau }𝑑t^{}$$ One can see that $$g_i(\rho ,t)<\rho ^3f_ig_{iappr}$$ Consider now the pseudo homogeneous situation. It means that we neglect the exhaustion of the heterogeneous centers by the droplets. Here $$g_i(\rho ,t)\rho ^3f_i$$ for all $`\rho `$ from $$\rho _0\rho <(0.7÷0.8)\zeta _{}\mathrm{\Delta }_it/\tau $$ As far as $`\rho ^3f_i`$ is the sharp function of $`\rho `$ we see that $`g_i`$ is even more sharp function of $`\rho `$. Then it is quite reasonable to suggest the monodisperse approximation for $`g_i`$. To construct the monodisperse approximation for $`g_i`$ one has to solve how to cut the tail at small $`\rho `$. Despite the rapid decrease at small $`\rho `$ the tail can not be integrated at least on the base of approximation $`g_{iappr}`$ (if we forget about the restriction $`\rho >0`$). There are two ways to do it. The first way is to cut off the spectrum on the halfwidth of $`g_{iappr}`$. It gives the halfwidht $$\mathrm{\Delta }\rho =(12^{1/3})\rho _0(t)=0.21\rho _0(t)\mathrm{\Delta }_{diff}\rho $$ which is small in comparison with $`\rho _0(g)`$. So, really, the approximation $`g_{iappr}`$ can be used here. The second way is more close to the iteration procedure from . One can define $`\mathrm{\Delta }\rho `$ by the integral way. One can integrate the approximation $`g_{iappr}`$ from $`\rho _0`$ up to $`0`$ taking into account that $`g_i=0`$ for $`\rho <0`$. Then $$\mathrm{\Delta }\rho =\rho _0(t)/4\mathrm{\Delta }_{int}\rho $$ As far as $`\mathrm{\Delta }_{diff}\rho \mathrm{\Delta }_{int}\rho `$ one can use both these two ways. The integral way is more convenient as far as it gives precise asymptotes. Now we can suggest approximation $$G_1=f_1\mathrm{\Delta }\rho \rho _0(t)^3$$ and rewrite it as $$G_1=N_1(t/4)\rho _0(t)^3$$ where $`N_1(t/4)`$ is the number of the droplets formed until the moment $`t/4`$ and the behavior of $`G`$ is analyzed in the current moment $`t`$. When we consider the heterogeneous condensation with essential exhausting of centers one can easily note that the function $`g_i`$ becomes even more sharp. So, the previous derivation is suitable here. Certainly, the value of $`N_1(t/4)`$ has to be calculated with account of exhaustion of the heterogeneous centers as it was described in , . Here we can present the last approximation as $$G_1(\eta _{1tot}\eta _1(t/4))\rho _0(t)^3$$ where $`\eta _1`$ is the number of the free heterogeneous centers of the first sort. ## 4 Floating monodisperse approximation In we were interested in the final parameters of the whole nucleation periods and used the monodisperse approximation at $`t=\mathrm{\Delta }_2t`$. It allowed to use for $`N_1(\mathrm{\Delta }_2t/4)`$ the following approximation $$N_1(\mathrm{\Delta }_2t/4)=\eta _{1tot}(1\mathrm{exp}(B\mathrm{\Delta }_2t/4))$$ where $$B=f_1n_{\mathrm{}}\zeta _{}/\tau \eta _{1tot}$$ Now we don’t want to use the monodisperse approximation only at $`t=\mathrm{\Delta }_2t`$. So, we shall act without the last approximation. But now we have to use the monodisperse approximation at the arbitrary moment of time $`t`$. We shall use the variables $`x,z`$ (see , ) and shall investigate the system of equations $$\zeta _{}=\zeta +G_1+G_2$$ $$G_1=f_1_0^z(zx)^3\mathrm{exp}(\mathrm{\Gamma }_1(\zeta \zeta _{})/\zeta _{})\theta _1(x)𝑑x$$ $$G_2=f_2_0^z(zx)^3\mathrm{exp}(\mathrm{\Gamma }_2(\zeta \zeta _{})/\zeta _{})\theta _2(x)𝑑x$$ $$\theta _1=\mathrm{exp}(\frac{f_1n_{\mathrm{}}}{\eta _{1tot}}_0^z\mathrm{exp}(\mathrm{\Gamma }_1(\zeta \zeta _{})/\zeta _{})𝑑x)$$ $$\theta _2=\mathrm{exp}(\frac{f_2n_{\mathrm{}}}{\eta _{2tot}}_0^z\mathrm{exp}(\mathrm{\Gamma }_2(\zeta \zeta _{})/\zeta _{})𝑑x)$$ where $`\mathrm{\Gamma }_i`$ are some parameters (see ), $`\theta _i`$ are the relative numbers of the free heterogeneous centers of the given sort. In the situation of the strong unsymmetry we can rewrite this system in the following manner $$G_1=f_1_0^z(zx)^3\mathrm{exp}(\mathrm{\Gamma }_1G_1(x)/\zeta _{})\theta _1(x)𝑑x$$ $$G_2=f_2_0^z(zx)^3\mathrm{exp}(\mathrm{\Gamma }_2(G_1+G_2)/\zeta _{})\theta _2(x)𝑑x$$ $$\theta _1=\mathrm{exp}(\frac{f_1n_{\mathrm{}}}{\eta _{1tot}}_0^z\mathrm{exp}(\mathrm{\Gamma }_1G_1(x)/\zeta _{})𝑑x)$$ $$\theta _2=\mathrm{exp}(\frac{f_2n_{\mathrm{}}}{\eta _{2tot}}_0^z\mathrm{exp}(\mathrm{\Gamma }_2(G_1+G_2)/\zeta _{})𝑑x)$$ The first and the third equations of the previous system form the closed system which allows to consider $`G_1`$ in the second and the forth equations as some known value. According to the monodisperse approximation it can be presented as $$G_1(z)=\frac{f_1}{E}(1\theta _1(z/4))z^3$$ where $$\theta _1(z/4)=\mathrm{exp}(E_0^{z/4}\mathrm{exp}(\mathrm{\Gamma }_1f_1x^4/4\zeta _{})𝑑x)$$ $$E=\frac{f_1n_{\mathrm{}}}{\eta _{1tot}}$$ One can simplify the last expression as $$\eta _1(z/4)=\eta _{tot}\mathrm{exp}(Ez/4)$$ (3) for $$z/4<z_m$$ and $$\eta _1(z/4)=\eta _{tot}\mathrm{exp}(E(\frac{4\zeta _{}}{\mathrm{\Gamma }_1f_1})^{1/4}A)$$ (4) for $$z/4>z_m$$ where $$A=_0^{\mathrm{}}\mathrm{exp}(x^4)𝑑x=0.905$$ and $$z_m=(\frac{4\zeta _{}}{\mathrm{\Gamma }_1f_1})^{1/4}A$$ Then the nucleation on the second sort centers can be described by the following equations $$G_2=f_2_0^z(zx)^3\mathrm{exp}(\mathrm{\Gamma }_2(\frac{f_1}{E}(1\theta _1(z/4))z^3+G_2)/\zeta _{})\theta _2(x)𝑑x$$ $$\theta _2=\mathrm{exp}(\frac{f_2n_{\mathrm{}}}{\eta _{2tot}}_0^z\mathrm{exp}(\mathrm{\Gamma }_2(\frac{f_1}{E}(1\theta _1(x/4))x^3+G_2)/\zeta _{})𝑑x)$$ We can adopt with a rather high accuracy the following expression for $`\theta _2`$ (the reasons are the same as in the section ”Final iterations” in ) $$\theta _2=\mathrm{exp}(\frac{f_2n_{\mathrm{}}}{\eta _{2tot}}_0^z\mathrm{exp}(\mathrm{\Gamma }_2(\frac{f_1}{E}(1\theta _1(x/4))x^3+f_2x^4/4)/\zeta _{})𝑑x)$$ or with the help of approximation (3), (4) it can be presented in more simple form. If $`z/4<z_m`$ then $$\theta _2=\mathrm{exp}(\frac{f_2n_{\mathrm{}}}{\eta _{2tot}}_0^z\mathrm{exp}(\mathrm{\Gamma }_2(\frac{f_1}{E}(1\mathrm{exp}(Ex/4))x^3+f_2x^4/4)/\zeta _{})𝑑x)$$ (5) If $`z/4>z_m`$ then $`\theta _2=\mathrm{exp}({\displaystyle \frac{f_2n_{\mathrm{}}}{\eta _{2tot}}}({\displaystyle _0^{4z_m}}\mathrm{exp}(\mathrm{\Gamma }_2({\displaystyle \frac{f_1}{E}}(1\mathrm{exp}(Ex/4))x^3+f_2x^4/4)/\zeta _{})dx+`$ (6) $`{\displaystyle _{4z_m}^z}\mathrm{exp}(\mathrm{\Gamma }_2({\displaystyle \frac{f_1}{E}}(1\mathrm{exp}(E({\displaystyle \frac{4\zeta _{}}{\mathrm{\Gamma }_1f_1}})^{1/4}A))x^3+f_2x^4/4)/\zeta _{})dx))`$ Now we have to calculate the integrals appeared in the last two expressions. We shall start from the first one. Consider (5). One can see that function $$\varphi \mathrm{\Gamma }_2(\frac{f_1}{E}(1\mathrm{exp}(Bx/4))x^3+f_2x^4/4)/\zeta _{}$$ is very sharp function. It is more sharp than $$\varphi _0constx^3+constx^4$$ The function $`(1\mathrm{exp}(Bx/4))`$ is rather smooth in comparison with $`\varphi `$ and $`\varphi _0`$. One can note that integrals $`_0^{\mathrm{}}\mathrm{exp}(x^3)𝑑x=0.89`$ and $`_0^{\mathrm{}}\mathrm{exp}(x^4)𝑑x=0.90`$ are approximately equal. Both subintegral functions have very sharp back front and one can speak about the cut-off in both cases. The approximate equality of these integrals means that these cut-off have approximately same values. We shall define the characteristic parameter $`z_q`$ by equality $$\mathrm{\Gamma }_2(\frac{f_1}{E}(1\mathrm{exp}(Ez_q/4))z_q^3+f_2z_q^4/4)/\zeta _{}=1$$ (7) Then the integral in (5) can be rewritten as $`{\displaystyle _0^z}\mathrm{exp}(\mathrm{\Gamma }_2({\displaystyle \frac{f_1}{E}}(1\mathrm{exp}(Ex/4))x^3+`$ (8) $`f_2x^4/4)/\zeta _{})dx=\mathrm{\Theta }(zCz_q)Cz_q+\mathrm{\Theta }(Cz_qz)z`$ where $$C=\frac{1}{2}(_0^{\mathrm{}}\mathrm{exp}(x^4)𝑑x+_0^{\mathrm{}}\mathrm{exp}(x^3)𝑑x)$$ This representation of the integral transfers (5) into $$\theta _2=\mathrm{exp}(\frac{f_2n_{\mathrm{}}}{\eta _{2tot}}(\mathrm{\Theta }(zCz_q)Cz_q+\mathrm{\Theta }(Cz_qz)z))$$ (9) Now we shall analyze the integral in (4). The reasons are the same. We shall introduce parameter $`z_l`$ by the following relation $$\mathrm{\Gamma }_2(\frac{f_1}{E}(1\mathrm{exp}(E(\frac{4\zeta _{}}{\mathrm{\Gamma }_1f_1})^{1/4}A))z_l^3+f_2z_l^4/4)/\zeta _{}=1$$ Note that we need only one parameter as far as $$1\mathrm{exp}(Ex/4)1\mathrm{exp}(E(\frac{4\zeta _{}}{\mathrm{\Gamma }_1f_1})^{1/4}A)$$ for $`\frac{x}{4}<z_m`$ and $$1\mathrm{exp}(Ex/4)|_{x/4=z_m}1\mathrm{exp}(E(\frac{4\zeta _{}}{\mathrm{\Gamma }_1f_1})^{1/4}A)$$ If $`z_l<4z_m`$ then $`{\displaystyle _0^{4z_m}}\mathrm{exp}(\mathrm{\Gamma }_2({\displaystyle \frac{f_1}{E}}(1\mathrm{exp}(Ex/4))x^3+f_2x^4/4)/\zeta _{})𝑑x`$ (10) $`{\displaystyle _{4z_m}^z}\mathrm{exp}(\mathrm{\Gamma }_2({\displaystyle \frac{f_1}{E}}(1\mathrm{exp}(E({\displaystyle \frac{4\zeta _{}}{\mathrm{\Gamma }_1f_1}})^{1/4}A))x^3+f_2x^4/4)/\zeta _{})𝑑x`$ and one can analyze only $$I_1=_0^{4z_m}\mathrm{exp}(\mathrm{\Gamma }_2(\frac{F_1}{E}(1\mathrm{exp}(Ex/4))x^3+f_2x^4/4)/\zeta _{})𝑑x$$ It was already done in consideration of (5). If $`z_l>4z_m`$ then both $$I_1=_0^{4z_m}\mathrm{exp}(\mathrm{\Gamma }_2(\frac{f_1}{E}(1\mathrm{exp}(Ex/4))x^3+f_2x^4/4)/\zeta _{})𝑑x$$ and $$I_2=_{4z_m}^z\mathrm{exp}(\mathrm{\Gamma }_2(\frac{f_1}{E}(1\mathrm{exp}(E(\frac{4\zeta _{}}{\mathrm{\Gamma }_1f_1})^{1/4}A))x^3+f_2x^4/4)/\zeta _{})𝑑x$$ are essential. Then $$I_1=4z_m$$ and $`I_2`$ can be analyzed quite analogously. Namely, we shall introduce $`z_t`$ from equality<sup>2</sup><sup>2</sup>2Certainly $`z_t=z_l`$ $$\mathrm{\Gamma }_2(\frac{f_1}{E}(1\mathrm{exp}(E(\frac{4\zeta _{}}{\mathrm{\Gamma }_1f_1})^{1/4}A))z_t^3+f_2z_t^4/4)/\zeta _{}=1$$ If $`z_t`$ is near $`4z_m`$ then $`I_2`$ is small in comparison with $`I_1`$ and there is no need to analyze $`I_2`$. If $`I_2`$ is essential in comparison with $`I_1`$ one can use the following approximation: $$I_2=(z4z_m)$$ for $`z<Cz_t`$ $$I_2=Cz_t4z_m$$ for $`z>Cz_t`$. This completes the approximate analysis of the expression for $`\theta _2`$. Some parameters $`z_l`$, $`z_m`$, $`z_q`$, $`z_t`$ may coincide but they are conserved in order to avoid misunderstanding. The main interesting value is $`\theta _2(\mathrm{})`$. The final expressions for this value are more simple. They can be directly obtained from the already presented ones. On the base of $`\theta _i(z)`$ one can easily find the number of droplets $`N_i`$ as $$N_i=\eta _{toti}(1\theta _i)$$ To find the total number of the droplets one has to put the arguments to $`\mathrm{}`$.
warning/0001/hep-th0001193.html
ar5iv
text
# Untitled Document hep-th/0001193 A Note On Relation Between Holographic RG Equation And Polchinski’s RG Equation Miao Li Institute of Theoretical Physics Academia Sinica Beijing 100080 and Department of Physics National Taiwan University Taipei 106, Taiwan mli@phys.ntu.edu.tw We clarify the relation between the recently formulated holographic renormalization group equation and Polchinski’s exact renormalization group equation. Jan. 2000 The holographic renormalization group flow has been clarified in the context of AdS/CFT correspondence as well as in the context of open string versus closed string . (For earlier attempts in this, see \[3----9\], also see .) It was noticed in these papers that the holographic RG equation arising either from the Hamilton-Jacobi theory of supergravity or from world-sheet considerations has a strong resemblance to Polchinski’s exact RG equation . Apparently the two equations are different. We aim in this short note to clarify the relation between them. The effective action in the AdS/CFT correspondence is defined as a functional of coupling constants, while the effective action of Polchinski is a functional of the fundamental fields. Thus the holographic RG equation is naturally a differential equation in coupling coupling constants, and the Polchinski RG equation is one in fundamental fields. For simplicity and without loss of generality, we will consider the field theory of a single Hermitian matrix. We start with a single matrix model in 0 dimension to illustrate some of our ideas. Although in the 0-dimensional matrix model there is no infinity to remove, one still can design an artificial RG flow by introducing a “cut-off” in the quadratic term in the action $$S_0=\frac{1}{2}NK(t)\mathrm{tr}\mathrm{\Phi }^2,$$ where $`\mathrm{\Phi }`$ is a Hermitian matrix, $`K(t)`$ is the cut-off propagator, depending on $`t=\mathrm{ln}a`$, $`a`$ is the running cut-off. Unlike in a genuine field theory, where once an interaction term is introduced in the action, many other terms will be generated with a nontrivial $`K(t)`$ in order to keep the physics invariant under changing $`t`$. In our case, there is much freedom in satisfying the RG flow, as we shall explain later. To mimic the $`𝒩=4`$ super Yang Mills theory, we introduce the interaction part as a sum of single trace operators $$S_1=N\underset{n3}{}\varphi _n(t)\mathrm{tr}\mathrm{\Phi }^n,$$ where again a factor N is introduced to follow the usual large N field theory convention. With this convention, the effective action as a functional of $`\varphi _n`$ defined by $$e^{S(\varphi (t),t)}=[d\mathrm{\Phi }]e^{S_0+S_1},$$ has the usual genus expansion $$S(\varphi _n)=N^{22h}F_h,$$ and the connected two point functions of operators $`\mathrm{tr}\mathrm{\Phi }^n`$ in the leading order is proportional to $`N^0`$. We pause to emphasize that it is crucial to introduce only single trace operators. In the AdS/CFT correspondence, a single trace operator is related to a field in SUGRA or string theory, the role of which is played by $`\varphi _n`$ in our toy model. Multi-trace operators are related to multi-particle states. The RG equation as formulated in has to do with only single trace operators. To ensure the effective action $`S`$ be independent of $`t`$, $`S_1`$ must satisfy a differential equation . As we shall see, this differential equation can not be satisfied by our model in which $`S_1`$ contains only single trace operators. Thus we need to relax this equation to be the one valid only when taken average in the path integral, namely $$_tS_1+K_1^{ij,lk}\frac{S_1}{\mathrm{\Phi }_{ij}}\frac{S_1}{\mathrm{\Phi }_{lk}}+K_2^{ij,lk}\frac{^2S_1}{\mathrm{\Phi }_{ij}\mathrm{\Phi }_{lk}}+K_3=0,$$ where $$𝒪=\frac{[d\mathrm{\Phi }]𝒪e^{S_0+S_1}}{[d\mathrm{\Phi }]e^{S_0+S_1}}.$$ we will determine $`K_1,K_2,K_3`$ momentarily. Note that the constant term $`K_3`$ can be removed by a shift of $`S_1`$, and this shift can be absorbed into the definition for the measure of $`\mathrm{\Phi }`$. Thus in the following we will ignore this term. For completeness, we will derive eq.(1). We start with $$_te^S=[d\mathrm{\Phi }](\frac{1}{2}_tK\mathrm{tr}\mathrm{\Phi }^2+_tS_1)e^{S_0+S_1}=0.$$ Use (1) to replace $`_tS_1`$ in (1). Next, use the fact $$\begin{array}{cc}& [d\mathrm{\Phi }]K_1^{ij,lk}\frac{S_1}{\mathrm{\Phi }_{ij}}\frac{S_1}{\mathrm{\Phi }_{lk}}e^{S_0+S_1}=[d\mathrm{\Phi }]K_1^{ij,lk}\frac{S_1}{\mathrm{\Phi }_{ij}}(NK\mathrm{\Phi }_{kl}+\frac{}{\mathrm{\Phi }_{lk}})e^{S_0+S_1}\hfill \\ & =[d\mathrm{\Phi }]\left(NKK_1^{ij,lk}\frac{S_1}{\mathrm{\Phi }_{ij}}\mathrm{\Phi }_{kl}K_1^{ij,lk}\frac{^2S_1}{\mathrm{\Phi }_{ij}\mathrm{\Phi }_{lk}}\right)e^{S_0+S_1},\hfill \end{array}$$ we see that if we choose $`K_2^{ij,lk}=K_1^{ij,lk}`$, then the second derivatives of $`S_1`$ cancel. Applying the same trick to the first term in the last line of the above equation $$\begin{array}{cc}& [d\mathrm{\Phi }]NKK_1^{ij,lk}\mathrm{\Phi }_{kl}\left(NK\mathrm{\Phi }_{ji}+\frac{}{\mathrm{\Phi }_{ij}}\right)e^{S_0+S_1}\hfill \\ & =[d\mathrm{\Phi }](N^2K^2K_1^{ij,lk}\mathrm{\Phi }_{ji}\mathrm{\Phi }_{kl}NK_1^{ij,ji})e^{S_0+S_1}.\hfill \end{array}$$ Now the first term in the above can be used to cancel the first term in (1) if $$K_1^{ij,lk}=\frac{1}{2}N^1_tK^1\delta _{ik}\delta _{jl}.$$ And the inhomogeneous term is removed by choosing $`K_3=\frac{1}{2}N^2_t\mathrm{ln}K`$. However, as we remarked before, this term can be absorbed into a redefinition of the measure and henceforth we will ignore it. To summarize, we have derived the following equation $$_tS_1+\frac{1}{2}N^1_tK^1\left(\frac{S_1}{\mathrm{\Phi }_{ij}}\frac{S_1}{\mathrm{\Phi }_{ji}}+\frac{^2S_1}{\mathrm{\Phi }_{ij}\mathrm{\Phi }_{ji}}\right)=0.$$ As we advertised, this is the weak form of Polchinski’s equation. The above equation is not valid if the average symbol is removed. (We call this equation the strong form of Polchinski equation) To see this, we compute $$\frac{S_1}{\mathrm{\Phi }_{ij}}\frac{S_1}{\mathrm{\Phi }_{ji}}==N^2\underset{n4}{}g_n\mathrm{tr}\mathrm{\Phi }^n,$$ where $$g_n=\underset{l}{}l(n+2l)\varphi _l\varphi _{n+2l}.$$ And $$\frac{^2S_1}{\mathrm{\Phi }_{ij}\mathrm{\Phi }_{ji}}=NG_{mn}\mathrm{tr}\mathrm{\Phi }^m\mathrm{tr}\mathrm{\Phi }^n$$ with $$G_{mn}=(m+n+2)\varphi _{m+n+2}.$$ If the original Polchinski equation applies, then $`_tS_1`$ contains only single trace operators, and can be used to balance the single trace operators in (1). However, (1) contains double trace operators, and can not be balanced in Polchinski equation, in the large N limit, since these operators are new independent operators. In order to solve Polchinski equation, we need to introduce in $`S_1`$ double trace operators. This in turn generates triple trace operators in $`^2S_1`$, etc. Thus in order for the Polchinski equation to hold, all multiple trace operators must be introduced. With a little thought, it is easy to realize that this conclusion holds for any matrix model, including $`𝒩=4`$ SYM. We thus learn that it is impossible satisfy the strong form of Polchinski equation without introducing multiple trace operators. On the other hand, there is no problem to satisfy the weak form of Polchinski’s equation, eq.(1). It simply generates a first order differential equations for $`\varphi _n(t)`$. Also, as we shall see shortly, the term $`^2S_1`$ in (1) is the same order as the term $`S_1S_1`$, in the large N limit. This is quite different from the speculation of , where it is conjectured that the holographic RG equation is just the strong form of Polchinski equation, if so, then $`^2S_1`$ is suppressed by $`1/N^2`$. Define the beta function $$\beta _n(\varphi )=\frac{d\varphi _n}{dt},$$ then $$_tS_1=\beta _nN\mathrm{tr}\mathrm{\Phi }^n=\beta _n\frac{S}{\varphi _n},$$ where we suppressed summation over $`n`$. Use (1), $$\frac{S_1}{\mathrm{\Phi }_{ij}}\frac{S_1}{\mathrm{\Phi }_{ji}}=Ng_n\frac{S}{\varphi _n}.$$ Use (1), $$\frac{^2S_1}{\mathrm{\Phi }_{ij}\mathrm{\Phi }_{ji}}=N^1G_{mn}e^S\frac{^2}{\varphi _m\varphi _n}e^S=N^1G_{mn}(\frac{S}{\varphi _m}\frac{S}{\varphi _n}+\frac{^2S}{\varphi _m\varphi _n}).$$ Substituting (1), (1) and (1) into (1), we find $$(\beta _n+\frac{1}{2}_tK^1g_n)\frac{S}{\varphi _n}+\frac{1}{2}N^2_tK^1G_{mn}(\frac{S}{\varphi _m}\frac{S}{\varphi _n}+\frac{^2S}{\varphi _m\varphi _n})=0.$$ This equation is almost the same as the holographic RG equation, say as presented in . The RG equation in is derived for the leading order in the large N limit. To compare with that, we use the genus expansion of (1) to derive in the leading order $$(\beta _n+\frac{1}{2}_tK^1g_n)\frac{F_0}{\varphi _n}+\frac{1}{2}_tK^1G_{mn}\frac{F_0}{\varphi _m}\frac{F_0}{\varphi _n}=0.$$ Two crucial points deserve mentioning explicitly. unlike one would naively think, the term $`S_1S_1`$ in the weak form of Polchinski equation is not identified with $`SS`$ in the holographic RG equation. Rather, it generates only the form $`S`$, a correction to the beta function in (1). On the other hand the term $`^2S_1`$ in the weak form of Polchinski equation generates both the terms $`SS`$ and $`^2S`$ in the holographic RG equation, with the term $`^2S`$ subleading to $`SS`$ in the large N limit. Clearly, the former is identified with the disconnected part of two point functions, as also used in , while the latter is identified with connected part of two point functions. Clearly, both of these terms appear only in a large N theory, since they come from the double traced operators in the Polchinski equation. With a single scalar field, one has only the $`S`$ term. The resulting RG equation can not be interpreted as coming from the Hamilton-Jacobi equation of a gravity theory. In the AdS/CFT correspondence, the beta functions are determined by the local part of the effective action $`S`$. Denote this local part by $`S_{loc}`$. It contains kinetic term $`\frac{1}{2}G^{mn}_t\varphi _m_t\varphi _n`$, so the beta function is given by $$\beta _m=G_{mn}\frac{S_{loc}}{\varphi _n},$$ where $`S_{loc}`$ is a functional of $`\varphi _m(t)`$ which is determined by the initial value problem. However, in our toy model there is no such a kinetic term in the extra dimension $`t`$, the reason is quite simple: the effective action (1) is defined already as a functional of initial values $`\varphi _m(t)`$. We do not know how to define an “off-shell” action which can be expressed as an integral over the whole range of $`t`$. In fact, the beta functions are not completely determined by demanding RG invariance of the effective action only. The 0-dimensional one-matrix model can be solved completely in the large N limit . For instance, one can derive the Schwinger-Dyson equation in this limit. Apparently given a finite set of functions $`\{\varphi _n(t)\}`$, one of them is determined by the rest by requiring RG invariance. Our above discussions serve only for the purpose of deriving the holographic RG equation from the weak Polchinski equation. Eq.(1) has a flavor of the string equation and Virasoro constraints in one-matrix model. Incidentally these equations can be derived by an action principle . It may be worthwhile to pursue along this direction. The beta-functions are completely determined in a field theory. The new ingredient here is the requirement of the cut-off independent correlation functions. It is straightforward to generalize the above consideration to one matrix model in D dimensional spacetime. We use Euclidean signature. Now the regularized kinetic action is $$S_0=\frac{1}{2}Nd^Dxd^DyK(xy,t)\mathrm{tr}_i\mathrm{\Phi }(x)_i\mathrm{\Phi }(y),$$ where $`K(xy,t)`$ is the cut-off inverse propagator. When $`t\mathrm{}`$, it tends to a delta function. Since we are dealing with a field theory, in order to have a RG invariant partition function, it is not enough to have a local interaction action $`S_1`$. Assume the inverse of $`K(xy,t)`$ exist (so that the kinetic term is not degenerate), denote this inverse by $`K^1(xy)`$: $$d^DzK^1(xz)K(zy)=\delta ^D(xy).$$ The weak form of Polchinski equation reads $$_tS_1+N^1d^Dxd^DyK_1(xy)\left(\frac{S_1}{\mathrm{\Phi }(x)_{ij}}\frac{S_1}{\mathrm{\Phi }(y)_{ji}}+\frac{^2S_1}{\mathrm{\Phi }(x)_{ij}\mathrm{\Phi }(y)_{ji}}\right)=0,$$ where $$K_1(xy)=\frac{1}{2}\mathrm{\Delta }^1d^Dx^{}d^Dy^{}K^1(xx^{})_tK(x^{}y^{})K^1(y^{}y).$$ In order to satisfy the above equation, $`S_1`$ must contain all the nonlocal terms. If we start with a local action $$S_1=Nd^Dx\varphi _n(x)\mathrm{tr}\mathrm{\Phi }^n(x),$$ then $$\frac{S_1}{\mathrm{\Phi }(x)_{ij}}\frac{S_1}{\mathrm{\Phi }(y)_{ji}}=N^2\varphi _m(x)\varphi _n(y)\mathrm{tr}\mathrm{\Phi }^m(x)\mathrm{\Phi }^n(y),$$ infinitely many nonlocal terms are generated. The above can be expanded in derivatives of $`\mathrm{\Phi }`$. On the other hand, $$\frac{^2S_1}{\mathrm{\Phi }(x)_{ij}\mathrm{\Phi }(y)_{ji}}=N\delta ^D(xy)(m+n+2)\mathrm{tr}\mathrm{\Phi }^m(x)\mathrm{tr}\mathrm{\Phi }^n(y),$$ yielding a contact term. This contributes to the holographic RG equation a term $$N^2K_1(0)(m+n+2)\varphi _{m+n+2}(x)\left(\frac{S}{\varphi _m(x)}\frac{S}{\varphi _n(x)}+\frac{^2S}{\varphi _m(x)\varphi _n(x)}\right).$$ This is good news, since in the holographic RG equation, this is indeed a contact term. It originates from the fact that $`\varphi _m(x)`$ is a field in the AdS space, thus has a local quadratic term in the effective action. Denote the Fourier transform of $`K(xy)`$ by $`K(p)`$, then $`K^1(p)0`$ when $`|p|\mathrm{}`$, and the coefficient $`K_1(0)`$ in (1) is given by $$K_1(0)=\frac{1}{2}d^Dp\frac{1}{p^2}_tK^1(p),$$ which is certainly convergent. To have a closed form of RG equation, we thus introduce all possible single trace operators into the interaction part $`S_1`$. A generic operator is $$\mathrm{tr}_{i_1}\mathrm{}_{i_n}\mathrm{\Phi }\mathrm{}_{j_1}\mathrm{}_{j_m}\mathrm{\Phi }.$$ Denote such a generic operator by $`𝒪_I(x)`$, and the corresponding coupling by $`\varphi ^I(x)`$. Now go through the above steps, we will arrive at the following holographic RG equation $$d^Dx\left[(\beta ^I+g^I)\frac{S}{\varphi ^I(x)}+G_{IJ}\left(\frac{S}{\varphi ^I(x)}\frac{S}{\varphi ^J(x)}+\frac{^2S}{\varphi ^I(x)\varphi ^J(x)}\right)\right]=0,$$ where all the components of the metric $`G_{IJ}`$ are proportional to an integral of $`p^2_tK^1(p)`$ weighted by a polynomial of $`p`$. So for each component of $`G_{IJ}`$ to be well-defined, the cut-off propagator $`K^1(p)`$ must fall off more rapidly than any negative power of $`p`$ for large $`|p|`$. Note again that the correction to the beta function, $`g^I`$, comes from the part $`S_1S_1`$ in the weak Polchinski equation. To compare with , we can identify $`\beta ^I+g^I`$ with the beta function defined on the world-sheet, where the cut-off is defined on the world-sheet. As already pointed out in , there is a UV/UV relation between string world-sheet physics and spacetime physics. The two cut-offs are not identical, thus the two definitions of the beta functions are not the same. Once again, our RG equation (1) is valid for any $`N`$. Use the genus expansion (1), one recovers the RG equation of \[1,,2\] in the large N limit. The subleading term $`^2S`$ in (1) is to be interpreted as coming from quantum corrections in the AdS/CFT context. We also want to emphasize the fact that our RG equation (1) involves a single integral. This is also true for the equation derived in . In order to remove this integral to obtain a local form, we need to introduce position-dependent cut-off. This is also related to general covariance in AdS/CFT. We leave a detailed discussion of this to another work. More equations can be derived by demanding the renormalized correlation functions to be independent of the cut-off. These equations are just Callan-Symanzik equations. They can be derived from the local form of the RG equation , but not from (1). We can derive them in the matrix field theory by generalizing the steps leading to (1). These equations put together will give a closed system of equations for $`S`$ and $`\beta ^I`$. We are not sure whether these beta functions can be written in a form (1). Note that the metric $`G_{IJ}`$ in (1) on the moduli space is the same as in (1), and is already determined in deriving (1). It would be highly nontrivial if all these beta functions are given by a single functional $`S_{loc}`$. Maybe this is the most crucial criteria for a holographic theory, and is generically violated by an arbitrary matrix field theory such as the single scalar field theory. It remains to generalize our construction to $`𝒩=4`$ super Yang-Mills theory. Although we do not see essential difficulty in doing this, we need to resolve the problem of introducing a gauge invariant cut-off. A naive cut-off will not work, since this is not compatible with local gauge transformation which mixes all energy scales. Another way to see this is through the naive cut-off Yang-Mills action $$d^Dx\mathrm{tr}F_{\mu \nu }(x)F_{\mu \nu }(y)K(xy,t),$$ it is certainly not gauge invariant. It has been suggested to use Wilson loop as gauge invariant variable to overcome this difficulty , and this line of approach was followed up in . However, the existence of AdS/CFT correspondence indicates that a gauge invariant cut-off exists for local gauge invariant variables. We suspect that stochastic quantization may be one way to gauge invariantly regulate Yang-Mills theory. And the stochastic time has been interpreted as the RG scale recently in . The discussion presented here may be viewed as a zero-slope limit of approach of . However, the relation between string world-sheet and large N diagrams need to be further clarified. Also, our approach does not has the drawback of assuming perturbation theory as in . The open/closed string duality need to be understood. One particularly nice example was discussed in . Acknowledgments. This work was supported by a grant of NSC and by a “Hundred People Project” grant of Academia Sinica. I thank M. Yu for several useful conversations, and T. Yoneya for comments on the manuscript. References relax J. de Boer, E. Verlinde and H. Verlinde, On the Holographic Renormalization Group, hep-th/9912012; E. Verlinde and H. Verlinde, Gravity and the Cosmological Constant, hep-th/9912018. relax J. Khoury and H. Verlinde, On Open/Closed String Duality, hep-th/0001056. relax E.T. Akhmedov, A Remark on the AdS/CFT Correspondence and the Renormalization Group Flow, Phys. Lett. B442 (1998) 152, hep-th/9806217. relax E. Alvarez and C. Gómez Geometric Holography, the Renormalization Group and the c-Theorem, Nucl. Phys. B541 (1999) 441, hep-th/9807226. relax D. Z. Freedman, S. S. Gubser, K. Pilch, N. P. Warner, Renormalization Group Flows from Holography--Supersymmetry and a c-Theorem, hep-th/9906194. relax L. Girardello, M. Petrini, M. Porrati, A. Zaffaroni, The Supergravity Dual of $`N=1`$ Super Yang-Mills Theory, hep-th/9909047; Novel Local CFT and Exact Results on Perturbations of N=4 Super Yang Mills from AdS Dynamics, hep-th/9810126; M. Porrati, A. Starinets, RG Fixed Points in Supergravity Duals of 4-d Field Theory and Asymptotically AdS Spaces, Phys.Lett. B454 (1999) 77, hep-th/9903241. relax V. Balasubramanian, P. Kraus, Space-time and the Holographic Renormalization Group, Phys.Rev.Lett 83 (1999) 3605, hep-th/9903190. relax K. Skenderis, P.K. Townsend, Gravitational Stability and Renormalization-Group Flow, hep-th/9909070. relax O. DeWolfe, D.Z. Freedman, S.S. Gubser, A. Karch, Modelling t he fifth dimension with scalars and gravity, hep-th/9909134. relax C. Schmidhuber, AdS-Flows and Weyl Gravity, hep-th/9912155. relax J. Polchinski, Renormalization and Effective Lagrangians, Nucl. Phys. B231 (1984) 269. relax J. Maldacena, The Large N Limit of Superconformal Field Theories and Supergravity, Adv. Theor. Math. Phys. 2 (1998) 231, hep-th/9711200; E. Witten, Anti-de Sitter Space and Holography, Adv. Theor. Math. Phys. 2 (1998) 253, hep-th/9802150; S. Gubser, I. Klebanov and A. Polyakov, Gauge Theory Correlators from Noncritical String Theory, Phys. Lett. B428 (1998) 105, hep-th/9802109. relax E. Brezin and S. Wadia, The large N expansion in quantum field theory and statistical physics: from spin systems to two-dimensional gravity, Singapore, World Scientific (1993). relax T. Yoneya, Action Principle, Virasoro Structure and Analyticity in Nonperturbative Two-dimensional Gravity, Int. J. Mod. Phys. A7 (1992) 4015. relax T. R. Morris, A Manifestly Gauge Invariant Exact Renormalization Group, hep-th/9810104. relax S. Hirano, Exact Renormalization Group and Loop Equation, hep-th/9910256. relax G. Parisi and Y.S. Wu, Sci. Sin. 24 (1981) 484. relax V. Periwal, String field theory Hamiltonian from Yang-Mills theories, hep-th/9906052; G. Lifschytz and V. Periwal, Dynamical Truncation of the String Spectrum at Finite N, hep-th/9909152. relax M. Li and Y.S. Wu, Holography and Noncommutative Yang-Mills, hep-th/9909085.
warning/0001/cond-mat0001210.html
ar5iv
text
# Anisotropy of the Energy Gap in the Insulating Phase of the 𝑼-𝒕-𝒕' Hubbard Model During the last decade of research work on high temperature superconductors angular resolved photoemission spectroscopy (ARPES) has played an important role in elucidating their electronic properties . An understanding of the single particle properties in the normal state is a prerequisite for a theory of a mechanism for superconductivity as well as for transport properties. Over the years ARPES data have continuously provided surprises and new insights and thereby served as a guidance to theoretical developments. The observation of the $`d_{x^2y^2}`$-shape of the energy gap, pseudogap structures in the metallic phase , strong anisotropies in the quasiparticle (qp) peak lineshapes and a possible partial destruction of the Fermi surface , or the unusual frequency and temperature dependence of the qp peakwidth are examples for intriguing information obtained from ARPES experiments. For the normal state an important goal is the description for the evolution with doping from the antiferromagnetic (AF) and insulating parent compounds to the overdoped superconductors. This demands control over the spectral features of the Mott-Hubbard insulating state as a starting point. Yet, it proved to be difficult to reproduce ARPES data for the single hole dispersion in the insulating cuprate Sr<sub>2</sub>CuO<sub>2</sub>Cl<sub>2</sub> within $`t`$-$`J`$ or Hubbard models. In particular, for momenta along the Brillouin zone (BZ) axis next-nearest neighbor (nnn) or even longer range hopping amplitudes had to be introduced to achieve a reasonable comparison to the measured spectra . Even more striking are recent results on AF Ca<sub>2</sub>CuO<sub>2</sub>Cl<sub>2</sub> that demonstrated an anisotropy of the insulating energy gap which was claimed to follow closely a $`d_{x^2y^2}`$-wave modulation along a remnant Fermi surface with a modulation amplitude of $``$ 300meV comparable in magnitude to the AF exchange interaction . It has been pointed out that a $`d_{x^2y^2}`$ gap modulation might follow naturally in the context of a projected SO(5) theory unifying antiferromagnetism and d-wave superconductivity by a symmetry principle . In this paper we show that a $`\gamma _d^2(\stackrel{}{k})=(\mathrm{cos}k_x\mathrm{cos}k_y)^2`$ modulation of the AF energy gap is realized in the half-filled Hubbard model on a square lattice with nearest (nn) and nnn hopping amplitudes. This result is obtained in an analytic strong coupling expansion around the atomic limit following a recently proposed strategy with a mapping to a Jacobi continued fraction representation for the propagator . Given that the Mott-Hubbard insulator is the appropriate starting point for studying the hole doping evolution in cuprates this intrinsic anisotropy may bear an important preformed structure for the doped metallic phase. The single particle Matsubara Green’s function for the Hubbard model is represented in terms of Grassmann fields $`\gamma ^{}`$ and $`\gamma `$ in the Feynman path integral representation by $`G_{ij}(\tau \sigma |\tau ^{}\sigma ^{})=T_\tau ^{}c_{i\sigma }^{}(\tau )`$ $`c_{j\sigma ^{}}^{}(\tau ^{})`$ (1) $`={\displaystyle \frac{1}{Z}}{\displaystyle [d\gamma ^{}d\gamma ]\gamma _{i\sigma \tau }^{}\gamma _{j\sigma ^{}\tau ^{}}^{}}`$ $`\mathrm{exp}(S[\gamma ^{},\gamma ])`$ (2) with the partition function $`Z`$ and the action $`S=S_{kin}+S_{atom}`$ where $`S_{kin}[\gamma ^{},\gamma ]={\displaystyle _0^\beta }d\tau {\displaystyle \underset{i,j,\sigma }{}}t_{ij}(\gamma _{i\sigma \tau }^{}\gamma _{j\sigma \tau }^{}+h.c.),`$ (3) $`S_{atom}[\gamma ^{},\gamma ]={\displaystyle _0^\beta }d\tau [`$ $`{\displaystyle \underset{i,\sigma }{}}\gamma _{i\sigma \tau }^{}({\displaystyle \frac{}{\tau }}\mu )\gamma _{i\sigma \tau }^{}`$ (4) $`+U`$ $`{\displaystyle \underset{i}{}}\gamma _{i\tau }^{}\gamma _{i\tau }^{}\gamma _{i\tau }^{}\gamma _{i\tau }^{}].`$ (5) Here, $`c_{i\sigma }^{}`$ creates an electron at site $`i`$ with spin $`\sigma `$, and we restrict the hopping amplitudes $`t_{ij}`$ to nn ($`t`$) and nnn ($`t^{}`$) sites. $`U`$ is the on-site Coulomb repulsion, $`\mu `$ the chemical potential, and $`\beta =1/T`$ the inverse temperature. $`G_{ij}(\tau \sigma |\tau ^{}\sigma ^{})`$ is evaluated diagrammatically in terms of a cumulant expansion around the atomic limit, i.e. an expansion in powers of $`t/U`$ . As early on suggested by Sarker it is convenient to introduce auxiliary Grassmann fields $`\{\psi _{i\sigma \tau }^{},\psi _{i\sigma \tau }^{}\}`$ by performing a Hubbard-Stratonovich transformation with respect to the kinetic energy term Eq. 3. From the Green’s function for the auxiliary fields $`𝒱_{ij}(\tau \sigma |\tau ^{}\sigma ^{})=T_\tau \psi _{i\sigma \tau }^{}\psi _{j\sigma ^{}\tau ^{}}^{}`$ the propagator $`G_{ij}(\tau \sigma |\tau ^{}\sigma ^{})`$ in Eq. 2 is obtained via the relation $`G_{ij}(\tau \sigma |\tau ^{}\sigma ^{})=(\mathrm{\Gamma }^1V)_{ij}^1(\tau \sigma |\tau ^{}\sigma ^{})`$ . Here $`\mathrm{\Gamma }`$ denotes the self energy and $`V`$ is the non-interacting Green’s function for the auxiliary fields, i.e. $`𝒱=V+V\mathrm{\Gamma }𝒱`$. As is known from previous work on strong-coupling expansions for the Hubbard model the Green’s function calculated this way does not have the correct analytic properties, because higher order cumulants contain also higher order poles. This problem was circumvented in Ref. by mapping the Green’s function as calculated by the diagrammatic expansion to a Green’s function $`G_J`$ in a finite Jacobi continued fraction representation $$G_J(z)=\frac{a_0}{z+b_1{\displaystyle \frac{a_1}{z+b_2{\displaystyle \frac{a_2}{z+\mathrm{}{\displaystyle \frac{a_{N1}}{z+b_N}}}}}}}$$ (6) such that $`G_J`$ has the same series expansion as $`G`$ to the same order in $`t/U`$. Specifically, we have calculated the Fourier transformed Green’s function $`G(\stackrel{}{k},i\omega _n)`$ at half filling $`\mu =\frac{U}{2}`$ up to order $`(t/U)^4`$. While our independently obtained results agree with Ref. for nn hopping only, in the case of a finite nnn hopping $`t^{}`$ an additional $`tt^2/U^3`$ diagram contributes to $`G(\stackrel{}{k},i\omega _n)`$ as shown in Fig. 1. In this type of diagrams the vertices represent local n-particle cumulants $`G_C^{(n)}`$ (connected Green’s functions). Each line between two vertices represents a hopping process between two sites. In the absence of a Wick theorem for the local Green’s functions the higher order cumulants must be calculated separately. The resulting algebraic expressions become very involved for higher order cumulants; for the performance of this diagrammatic expansion we have therefore developed a special purpose computer algebra code. As a result we obtain a Jacobi continued fraction expression for $`G_J`$ that has eight fraction levels, i.e. the coefficients $`a_N`$ in Eq. 6 vanish for $`N>8`$. This termination for $`G_J`$ translates into eight simple poles. Each of the continued fraction coefficients $`a_i`$, $`b_i`$ $`(i=1,\mathrm{},8)`$ is given by a fourth order polynomial in $`t/U`$ or $`t^{}/U`$, respectively, with coefficients that depend on $`T`$, $`\mu `$, $`U`$, and $`\stackrel{}{k}`$. The explicit result for $`G_J`$ is accessible electronically . In Fig. 2 we show the corresponding spectral function $$A(\stackrel{}{k},\omega )=\frac{1}{\pi }\mathrm{I}mG(\stackrel{}{k},\omega +i0^+)$$ (7) for different $`\stackrel{}{k}=(k_x,k_y)`$ along the path $`(0,0)(\pi ,\pi )(\pi ,0)(0,0)`$ in the first BZ. On the scale of this plot only six out of eight poles of $`G_J(\stackrel{}{k},\omega )`$ carry significant and visible spectral weight. In analyzing the spectrum we map out the dispersion of the lowest energy hole excitation, i.e. the dispersion of the first peak below the gap (marked by an arrow in Fig. 2). Its spectral weight is largest at the BZ center and drops monotonically towards $`(\pi ,\pi )`$. Fig. 3 shows the peak dispersion for different temperatures. Upon cooling the system approaches the AF ordered ground state with a doubled unit cell and a reduced magnetic BZ (determined by $`\mathrm{cos}k_x+\mathrm{cos}k_y0`$). The dispersion along the zone diagonal approaches a perfectly symmetric shape with respect to the point $`\stackrel{}{k}_d=(\frac{\pi }{2},\frac{\pi }{2})`$ on the magnetic BZ boundary reflecting the growing AF spin correlations. Remarkably, along the BZ axis the dispersion remains very flat at all temperatures. In fact, the dispersion along the axis is obtained flatter than in previous numerical studies of Hubbard or $`t`$-$`J`$ models . In units of the exchange coupling $`J=4t^2/U`$ the total bandwidth of the energy dispersion is $`1.58`$ for $`T=0.2t`$, which is roughly consistent with experiment and previous calculations. In Fig. 4 we compare our result for the single hole dispersion with the ARPES data on Sr<sub>2</sub>CuO<sub>2</sub>Cl<sub>2</sub> from Ref. . For $`U=10t`$ the parameters $`t`$ and $`t^{}`$ were chosen to obtain a best fit to the data. The overall agreement, in particular along the BZ axis, is quite satisfactory. Hopping amplitudes beyond nnn hopping are found unnecessary for reproducing the flat dispersion along the BZ axis.The deviation of the theoretical result from the experimental data points along the BZ diagonal, i.e. the lack of symmetry with respect to $`\stackrel{}{k}_d`$ – as realized in the AF state – is due to the finite temperature. In fact, the range of applicability of the $`t/U`$ expansion has a lower bound in temperature. (Following the arguments of Ref. we estimate that our results are valid for $`T>0.16t`$.) When the magnetic correlation length is much larger than the hopping range as given by the order of the $`t/U`$ expansion, an accurate description of the single particle Green’s function is no longer expected. The data were taken 100K above the Néel temperature $`T_N=256`$K of Sr<sub>2</sub>CuO<sub>2</sub>Cl<sub>2</sub>; at this temperature the magnetic correlation length is already as large as 250Å as measured by neutron scattering . Nevertheless, the flatness of the dispersion along the BZ axis in the $`t/U`$ expansion is a robust and temperature insensitive feature. In Fig. 5 the calculated energy dispersion $`E(\stackrel{}{k}_d)E(\stackrel{}{k})`$ of the lowest energy peak and the experimental data are plotted along the BZ path from $`\stackrel{}{k}_d`$ to $`(\pi ,0)`$ which is parametrized in the form $`|\gamma _d(\stackrel{}{k})|/2`$. Originally, the data of Ref. (black squares) were anticipated to imply a $`d_{x^2y^2}`$-modulation of the energy gap which in the parametrization of Fig. 5 would translate into a straight line. The perfectly linear relation between $`E(\stackrel{}{k}_d)E(\stackrel{}{k})`$ and $`|\gamma _d(\stackrel{}{k})|`$ along the chosen path would furthermore imply an unreasonable cusp-like feature of the dispersion on the magnetic BZ boundary. Indeed, the more recent data by Ronning et al. (white circles in Fig. 5) resolve this problem and rather provide evidence for a quadratic dependence of the energy gap on $`|\gamma _d(\stackrel{}{k})|`$ in the vicinity of $`\stackrel{}{k}_d`$. Given the experimental error bars of Ref. included in Fig. 5 and in the absence of an error estimate for the recent data by Ronning et al. our results are clearly compatible with experiment. With the same parameter set used for the fit of the ARPES dispersion in Fig. 4 the gap modulation amplitude $``$ 300meV as measured in Ca<sub>2</sub>CuO<sub>2</sub>Cl<sub>2</sub> is reproduced as well. Fig. 6 shows the temperature dependence of the gap modulation. In fact, when plotted versus $`\gamma _d^2(\stackrel{}{k})`$ it becomes evident that the gap approaches a perfect $`\gamma _d^2(\stackrel{}{k})`$ momentum dependence at low temperatures. For $`t^{}=0`$ the modulation vanishes in the expansion up to the order $`(t/U)^4`$ because the $`\stackrel{}{k}`$ dependence of all diagrams to this order arises in the form $`\mathrm{cos}k_x+\mathrm{cos}k_y`$ for $`t^{}=0`$ which vanishes on the BZ boundary for $`t^{}=0`$. If diagrams of order $`(t/U)^6`$ or higher were taken into account, a small modulation is expected to appear also for $`t^{}=0`$. In summary, we have found in a strong coupling expansion to order $`(t/U)^4`$ that the energy gap in the insulating phase of the half-filled $`U`$-$`t`$-$`t^{}`$ Hubbard model develops a $`(\mathrm{cos}k_x\mathrm{cos}k_y)^2`$ modulation at low temperatures. Without the need to include hopping amplitudes beyond nnn an excellent fit is achieved for the single hole dispersion in Sr<sub>2</sub>CuO<sub>2</sub>Cl<sub>2</sub>. With the same dispersion fit parameters also the calculated gap modulation amplitude compares well with the ARPES data. It is natural to expect that these spectral features in the insulator carry over to the metallic, doped case; the consequences and the connection to anisotropic pseudogap structures or even $`d`$-wave superconductivity remain yet to be understood. We thank D. Duffy and F. Ronning for discussions and S. Pairault for sharing his insight into the efficient calculation of high order diagrams. This work was supported by the Deutsche Forschungsgemeinschaft through SFB 484.
warning/0001/astro-ph0001207.html
ar5iv
text
# The redshift evolution of bias and baryonic matter distribution. ## 1 Introduction In the standard cosmological scenario large-scale structures observed in the present universe were formed by the amplification through gravitational instability of small primordial density fluctuations (usually assumed to be gaussian). Within hierarchical models (as in the CDM case: Peebles 1982) small scales collapse first to form bound objects which later merge to build more massive halos as larger scales become non-linear. This describes the evolution under the action of gravity of the dark matter component. However, observed objects (e.g., quasars, galaxies, Lyman-$`\alpha `$ clouds…) are formed by baryonic matter (gas, stars) which does not necessarily follow the behaviour of the dark matter component since it undergoes additional processes (e.g., radiative cooling). Moreover, by looking at specific astrophysical objects (e.g. bright galaxies) one selects peculiar dark matter environments where the former are most likely to arise. Thus, to relate observations to the underlying dark matter component (which is usually gravitationally dominant) and to test scenarios for structure formation one needs to study the link between the dark and baryonic components. This can also constrain the astrophysical models used to build galaxies and other objects from the density field. One tool to describe the distribution of objects in the universe, in addition to their abundance, is their two-point correlation function which measures their clustering. Then one often defines the bias $`b^2`$ of these halos as the ratio of this quantity to the dark matter correlation function, which shows in a clear quantitative way whether these objects are more or less clustered than the dark matter component. In this article, we present the bias (and the correlation length) associated with various objects (galaxies, quasars, Lyman-$`\alpha `$ clouds, clusters). The formulation of this bias is obtained (Schaeffer 1985,1987; Bernardeau & Schaeffer 1992,1999) within the framework of a general description of structure formation in the universe provided by a scaling model (Schaeffer 1984; Balian & Schaeffer 1989) which describes the density field in the non-linear regime. These predictions specifically take into account the deep non-linearity of the density field and are conceptually different from those obtained using linear theory (e.g., Mo & White 1996). Indeed, the latter authors evaluate the bias of primordial overdensities that will collapse later, using linear theory, with the implicit assumption that this bias is conserved during the course of evolution. They find the bias depends separately on the mass, radius and overdensity. Here, we use the bias calculated directly using non-linear theory. This bias is seen to depend on very specific internal properties of the considered non-linear objects, namely on a unique parameter $`x`$, quite close to the internal velocity dispersion within the objects, thus exhibiting a scaling law. Such a specific dependence has indeed been found to hold in numerical simulations (Munshi et al. 1999b), and we consider this approach to provide a more trustworthy description of the present clustering properties as well as their evolution at higher redshifts. In the present paper, we specifically study the dependence of the bias on the observable properties of the considered objects (luminosity, column density) and its evolution with redshift. In contrast with most previous works, our analytical model relating dark to observable mass is not an ad-hoc parameterization. Indeed, it is simply a consequence of a general model (Valageas & Schaeffer 1997) which has already been applied to galaxies (Valageas & Schaeffer 1999), Lyman-$`\alpha `$ clouds (Valageas et al. 1999), clusters (Valageas & Schaeffer 2000) and reionization studies (Valageas & Silk 1999a,b). Thus, its predictions have already been compared with many observations (e.g., galaxy luminosity function, column density distribution of Lyman-$`\alpha `$ clouds, amplitude of the UV background radiation field, cluster X-ray luminosity function,…). Hence there is no free parameter chosen in this article. Next, in addition to the bias we also present the distribution of baryonic matter over the various objects we describe. This article is organized as follows. In Sect.2 we describe our prescription for mass functions, while in Sect.3 we recall how the bias is obtained in this model and we compare our formalism with other prescriptions. In Sect.4 we present our results for the case of a critical density universe. We first describe the clustering of galaxies, quasars, Lyman-$`\alpha `$ clouds and clusters and then we study the repartition of baryonic matter within various classes of objects. Finally, in Sect.5 we present the case of an open universe. ## 2 Multiplicity functions Since we wish to describe the properties of various classes of objects like Lyman-$`\alpha `$ clouds and galaxies which are defined by specific constraints we need a formalism which can handle more general multiplicity functions than the usual mass function of “just-virialized halos”. To this order, we shall assume that the non-linear density field obeys the scaling model detailed in Balian & Schaeffer (1989). Thus, we define each class of objects (Lyman-$`\alpha `$ clouds or galaxies) by a constraint $`\mathrm{\Delta }(M,z)`$ on the density contrast of the underlying dark-matter halo of mass $`M`$ at redshift $`z`$. The “just-virialized halos” correspond to the special case $`\mathrm{\Delta }(M,z)=\mathrm{\Delta }_c(z)`$ with $`\mathrm{\Delta }_c177`$. For our purposes we consider the cases of a constant density contrast, arising from a virialization constraint, and of a constant radius (i.e. $`(1+\mathrm{\Delta })M`$), arising from a fixed cooling radius or Jeans length. In any case, we write the multiplicity function of these objects as (see Valageas & Schaeffer 1997 for details): $$\eta (M,z)\frac{dM}{M}=\frac{\overline{\rho }}{M}x^2H(x)\frac{dx}{x}.$$ (1) Here, the parameter $`x`$ associated with a halo of mass $`M`$ at redshift $`z`$ is defined by: $$x(M,z)=\frac{1+\mathrm{\Delta }(M,z)}{\overline{\xi }[R(M,z),z]},$$ (2) where $$\overline{\xi }(R)=_V\frac{d^3r_1d^3r_2}{V^2}\xi _2(𝐫_1,𝐫_2)\text{with}V=\frac{4}{3}\pi R^3$$ is the average of the two-body correlation function $`\xi _2(𝐫_1,𝐫_2)`$ over a spherical cell of radius $`R`$ and provides the measure of the density fluctuations in such a cell (thus typical objects have $`x1`$). The scaling function $`H(x)`$ depends on the initial power-spectrum of the density fluctuations and must be taken from numerical simulations. However, from theoretical considerations (Balian & Schaeffer 1989) it is expected to satisfy the asymptotic behaviour: $$x1:H(x)x^{\omega 2},x1:H(x)x^{\omega _s1}e^{x/x_{}}$$ with $`\omega 0.5`$, $`\omega _s3/2`$, $`x_{}10`$ to 20, and by definition it must satisfy $$_0^{\mathrm{}}xH(x)𝑑x=1.$$ (3) These properties have been checked for various $`P(k)`$ by Bouchet et al (1991), Colombi et al. (1992,1994,1995,1997), Munshi et al. (1999a), while the mass functions obtained from (1) for various constraints $`\mathrm{\Delta }(M)`$ (for a constant $`\mathrm{\Delta }`$ which was taken from $`0.5`$ to $`5000`$ or for $`(1+\mathrm{\Delta })M`$) have been shown to provide reasonable approximations to numerical results in Valageas et al. (2000). ## 3 Bias ### 3.1 Our formulation Within the framework of the scale-invariance of the many-body matter correlation functions $`\xi _p(𝐫_1,\mathrm{},𝐫_p)`$ which led to the mass function (1), one can show that in the highly non-linear regime the bias characteristic of two objects factorizes and is a function of the sole parameter $`x`$ introduced in Sect.2, see Bernardeau & Schaeffer (1992,1999). Thus we write the correlation function of objects of type (1) with objects of type (2) as: $$\xi _{1,2}(r)=b(x_1)b(x_2)\xi (r)$$ (4) with $$x1:b(x)x^{(1\omega )/2}\text{and}x1:b(x)x.$$ (5) This behaviour has been sucessfully checked in numerical simulations by Munshi et al. (1999c). This allows us to obtain the redshift evolution of the bias associated with galaxies, Lyman-$`\alpha `$ clouds or quasars, once we define the constraints $`\mathrm{\Delta }(M,z)`$ which characterize these various astrophysical objects (that in turn define the proper value of $`x`$ to be used). ### 3.2 Comparison with some other models We can note that alternative models have been proposed in the litterature to describe the clustering of galaxies. Here we briefly compare our prescription to such models, in order to clarify their main differences. Thus, semi-analytic models of galaxy and dark matter clustering based on Press-Schechter-like ideas have recently been proposed (Seljak 2000; Peacock & Smith 2000). They describe the dark matter density field as a collection of smooth halos with a mean density profile. The mass function of these “just-virialized” halos (defined by an overall density contrast $`\mathrm{\Delta }=\mathrm{\Delta }_c177`$) is obtained from a slightly modified Press-Schechter mass function (Press & Schechter 1974). By construction, such a model neglects the substructures of virialized halos since they are described by an average density profile. This is not a problem at large scales where the correlations among galaxies are governed by the distribution of their host virialized halos (note that since the Press-Schechter prescription is based on the linear density field these correlations should actually apply to very large scales which are still in the linear regime). However, at scales of the order of the size of large “just-virialized” halos the clustering of galaxies is set by the distribution of these galaxies within those larger objects (which actually correspond to clusters of galaxies at large masses). Hence one needs to add a specific prescription to cope with this regime. Thus, one introduces the mean number $`N(M)`$ of galaxies contained in halos of mass $`M`$. This function is derived from N-body simulations or observations. Then, one puts a galaxy at the center of the halo and the $`(N1)`$ remaining galaxies at random within the halo, following the mean dark matter density profile. In order to include the fluctuations of $`N(M)`$ one can even reproduce the second moment $`N(N1)(M)`$ obtained from simulations (Seljak 2000). However, it is clear that the predictive power of such a model is quite limited. Indeed, the output of the model (i.e. the two-point correlation function of galaxies) is almost directly related to the input (the first two moments of the galaxy distribution within virialized halos). Thus, this can at best serve as a test of N-body simulations if this data is taken from simulations (while it would not make much sense to obtain this data from the observations one would later on try to recover). In particular, in order to derive the $`p`$point correlation function of galaxies in this line one would need the moment $`N^p`$ of order $`p`$… On the other hand, if one only includes the mean $`N(M)`$ (taken from observations for instance) as input, one can argue that the model tests whether the spatial galaxy distribution is well described by the mean dark matter density profile. However, as shown in Valageas (1999) such a model contradicts the results of numerical simulations for the dark matter distribution itself when one takes into account the constraints provided by the first five order moments of counts-in-cells statistics. This implies that one could not get higher-order moments (e.g., the three-point galaxy correlation function) from this model. Moreover, it is not clear how much information one can draw from the mere comparison of the galaxy power-spectrum predicted by the model with observations. Indeed, any sensible model should provide a reasonable agreement with the data since it should more or less follow the dark matter power-spectrum which is itself normalized to observations of large-scale clustering (through $`\sigma _8`$). Then, it seems difficult to obtain narrow constraints on the input parameters from only one curve, in view of the number of these free parameters and the approximations involved in such models (e.g., importance of the scatter of galaxy properties, shape of the distribution $`P(N)`$ within virialized halos, role of dark matter substructures,…). Such approaches differ from ours in many respects. First, we use a mass function derived from a scaling ansatz for the dark matter density field (Balian & Schaeffer 1989) which applies to the non-linear regime. It is fully specified by the sole function $`H(x)`$, obtained from counts-in-cells statistics, and by the behaviour of the non-linear power-spectrum. Most importantly, it allows us to get various multiplicity functions for different classes of objects in addition to “just-virialized” halos, defined by a relation $`\mathrm{\Delta }(M)`$ or $`R(M)`$ which specifies their overall density contrast or radius as a function of their mass (Valageas & Schaeffer 1997; Valageas et al. 2000). This enables us to bypass the link between galaxies and “just-virialized” halos which hinders the approaches described above. Indeed, we directly recognize the galaxies themselves, defined by a relation $`R(M)`$ obtained from cooling constraints (Valageas & Schaeffer 1999, 2000). Then, as explained above the clustering of these objects is obtained (Bernardeau & Schaeffer 1992,1999) from the behaviour of the $`p`$point correlation functions $`\xi _p`$ which led to the mass function (1). Hence it does not require any additional information. Moreover, although we restrict ourselves to the two-point correlation functions, our formulation also predicts the $`p`$point correlation functions of any order $`p`$ for all the objects we study here (galaxies, QSOs,…). This clearly illustrates the predictive power of our model. Besides, although we do not take into account the scatter of galaxy properties with respect to their mass, our procedure automatically includes the part of the scatter of the galaxy distribution with respect to the distribution of “just-virialized” halos which is due to the statistics of the dark matter itself. Finally, the spirit of our present study is rather different from the works which are mainly intended to describe galaxy clustering (Seljak 2000; Peacock & Smith 2000) as we merely compare with observations the predictions of a general model of structure formation which was fully constrained in previous papers which dealt with its consequences for galaxies, QSOs, Lyman-$`\alpha `$ clouds and clusters. In particular, we have no free parameters left to improve the agreement with the observed galaxy clustering and we check our model against a vast array of observations, ranging from Lyman-$`\alpha `$ clouds up to clusters, through galaxies and QSOs. Thus, our goal is not to use observations of galaxy clustering to constrain some free parameters, but to check whether a sensible scenario of structure formation can match this data. Besides, since we use an analytic model where the various physical processes can be clearly identified, we can get some valuable insight into the clustering properties of the range of objects we study here. For instance, we shall see that a model like ours can explain in simple physical terms the qualitative and quantitative differences between the redshift evolution of the clustering of bright galaxies and faint galaxies, small Lyman$`\alpha `$ forest clouds and damped systems, or QSOs and galaxies. Another method to study the clustering of galaxies is to use N-body simulations, which may be supplemented by semi-analytic modelling of star-formation or galaxy formation (e.g., Benson et al. 2000a). In principle, this allows one to avoid the assumptions and approximations involved in an analytic model like ours to describe the dark matter density field. However, this technique may be limited by numerical resolution and computer time constraints. In any case, this is certainly a promising method but it does not provide yet the clustering properties of the full range of objects we study here in a unified fashion. ## 4 Critical density universe We first consider the case of a critical density universe $`\mathrm{\Omega }_\mathrm{m}=1`$ with a CDM power-spectrum (Davis et al. 1985), normalized to $`\sigma _8=0.5`$. We choose a baryonic density parameter $`\mathrm{\Omega }_b=0.04`$ and $`H_0=60`$ km/s/Mpc. We use the scaling function obtained by Bouchet et al. (1991) for $`H(x)`$. ### 4.1 Galaxies #### 4.1.1 Present universe We use the model of galaxy formation and evolution described in detail in Valageas & Schaeffer (1999), where it was checked against many observations (luminosity function, Tully-Fisher relation…). We define galaxies by two constraints: 1) a virialization condition $`\mathrm{\Delta }>\mathrm{\Delta }_c`$ where $`\mathrm{\Delta }_c177`$ is given by the usual spherical model, and 2) a cooling constraint $`t_{cool}<t_H`$ which requires the gas to cool within a few Hubble times at formation. At high virial temperature and low redshift, the condition 2) is the most stringent and can be approximated by a constant radius constraint $`R=R_{cool}`$. At high redshift it becomes irrelevant since any halo which satisfies 1) also satisfies 2). These two conditions define for galaxies the constraint $`\mathrm{\Delta }(M,z)`$ we introduced in Sect.2. Once we have identified in this manner the dark-matter halos which correspond to galaxies we still need to set up a model for star formation to obtain the stellar properties of these objects. As detailed in Valageas & Schaeffer (1999) our model involves 4 components: (1) short-lived stars which are recycled, (2) long-lived stars which are not recycled, (3) a central gaseous component which is deplenished by star formation and ejection by supernovae winds, replenished by infall from (4), a diffuse gaseous component. This allows us to obtain the luminosity of these galaxies. Since observations usually refer to the mean bias (or correlation function) of galaxies above a given luminosity, we define: $$b(>L)=\frac{_L^{\mathrm{}}b(L)\eta _g(L)\frac{dL}{L}}{_L^{\mathrm{}}\eta _g(L)\frac{dL}{L}},$$ (6) where we used the factorization of the bias, see (4). We present in Fig.1 the dependence of the bias on the B-band luminosity. We see that we recover the trend of larger bias for brighter galaxies, although our dependence is somewhat steeper than the observed relation. However, we note that in our analytical model each luminosity is associated with a specific object of a given mass, radius,… In a more realistic model one should include some scatter in this relation which would lessen the final luminosity dependence of the bias (for a large scatter such that the luminosity is uncorrelated with the properties of the underlying dark matter halo we would have no B-magnitude dependence left). From the relation (4) we can also obtain the cross-correlation of galaxies of luminosity $`L_B`$ with all galaxies, in a fashion similar to (6). We can convert the bias $`b`$ evaluated in this manner into a correlation length $`r_0(b)`$ defined by: $$r_0(b)=b^{2/\gamma }r_0,$$ (7) where $`\gamma =1.8`$ is the approximate slope of the correlation function and $`r_0`$ is the correlation length of the dark matter density field. Indeed in our model the correlation functions of various objects and of the matter distribution all have the same slope in the non-linear regime (note that Dave et al. 1999 find from numerical simulations that the bias shows very little scale dependence, if any, in the non-linear regime). This prescription allows us to compare in Fig.2 our results to observations performed by Loveday et al. (1995). We see that our results are compatible with the observations. We find an increase of the bias (hence of $`r_0`$) for brighter galaxies in accordance with the data, except for the last point at $`22.8<M_B<20.8`$ where the bias is measured to be slightly smaller than for $`20.8<M_B<19.8`$. Nevertheless, since this contradicts the trend observed by Benoist et al. (1996) this disagreement is probably not conclusive. We present in Fig.3 the correlation function of galaxies brighter than $`M_B=18.6`$ at $`z=0`$. Our results agree reasonably well with observations, although the slope we obtain may be a bit too steep at large scales. This is a well-known discrepancy for CDM initial conditions with $`\mathrm{\Omega }_\mathrm{m}=1`$ and could be improved with a slightly different initial power-spectrum. The dashed line is the correlation function given by the linear theory multiplied by the relevant bias factor. Hence it coincides with the actual correlation function at large scales. At small scales the non-linear dynamics increases the clustering as compared to the linear predictions, in agreement with observation. Note that in principle the non-linear predictions are valid for $`\overline{\xi }1`$, but are in practice seen to hold as soon as $`\overline{\xi }1`$. #### 4.1.2 Redshift evolution We now turn to our predictions for the redshift evolution of the bias associated with galaxies recognized by their B-band luminosity, shown in Fig.4. At “low” luminosities $`M_B>21`$, the bias grows with redshift since galactic halos correspond to increasingly rare overdense regions (because structure formation was not as developed as it is today). For very bright galaxies $`M_B<21`$ we recover the same trend except at low redshifts where the bias suddenly increases with time. This evolution is due to a luminosity effect. Indeed, in our model at low $`z<1`$ very massive galaxies see their luminosity decrease with time as their initial gas content gets exhausted by star formation (which is very efficient in this high-density environment). As a consequence, when selecting galaxies by a fixed B-band luminosity, one looks at more massive, higher virial temperature halos at $`z=0`$ than at $`z=1`$. This shift to higher density fluctuations acts opposite to the usual trend seen for lower luminosity galaxies and leads to the increase with time of the bias at low $`z`$ for these very bright magnitudes. We present directly in Fig.5 the redshift evolution of the bias associated with galaxies brighter than three B-band magnitude thresholds. This clearly shows the dependence on luminosity of the redshift evolution itself. We show in Fig.6 the redshift evolution of the comoving correlation length $`r_0`$ associated with galaxies brighter than various magnitude thresholds. The two data points at $`z<0.1`$ correspond to $`z=0`$ but they have been slightly shifted to be more clearly seen. We can see that our results agree reasonably well with observations, except for the data points from Le Fevre et al. (1996) which however are inconsistent with data from Carlberg et al. (1998) and Postman et al. (1998). The comoving number density of Lyman-break galaxies at $`z=3`$ is $`n2.7\times 10^3`$ Mpc<sup>-3</sup> (see Adelberger et al. 1998) which is also the abundance of galaxies brighter than $`M_B=18.7`$ (and $`M5\times 10^{11}M_{}`$) in our model. The square at $`z=3`$ shows our prediction for the clustering strength of galaxies with this abundance. We see that this simple identification leads to good agreement with observations (filled triangle). The comoving correlation length of galaxies first decreases at higher $`z`$, following the decline of the dark-matter two-point correlation function. This trend is even larger for very bright galaxies since it is amplified by the decline of their bias, as shown in Fig.5. However, at high redshifts $`z>2`$, $`r_0`$ increases with $`z`$ because of the strong growth of the bias. Of course, this trend is stronger for brighter galaxies. We note that contrary to Kauffmann et al. (1999) we obtain a “dip” in the evolution of $`r_0`$, while these authors only found this feature for a low-density flat cosmology. This is explained by the slower growth with $`z`$ of the bias associated with galaxies in our model, as compared to theirs, which is insufficient to override at low $`z`$ the decrease of clustering measured by the decline of the dark-matter correlation function. Thus, such a feature depends on the astrophysical model used to build galaxies. However, we note that observations show that $`r_0`$ decreases with larger $`z`$ in the range $`0z0.5`$, in agreement with our model. We recall here that our results are direct consequences of a detailed model for structure formation which has already been compared with many observations, so that no attempt was made to “improve” the agreement of our predictions with the data. In fact, as shown by the discrepancy between various observations, the latter still have a significant uncertainty. However, we recover the general trend of a decrease of $`r_0`$ up to $`z2`$ and a subsequent growth at higher redshifts. This behaviour also agrees with the results obtained by Baugh et al. (1999) from a semi-analytic model. ### 4.2 Quasars From the description of galaxies presented above we have also developed a model for the quasar luminosity function, as detailed in Valageas & Silk (1999a). We assume that a fraction $`\lambda _Q1`$ of galactic halos host a quasar with a mass $`M_Q`$ proportional to the mass of gas $`M_{gc}`$ available in the inner parts of the galaxy: $`M_Q=FM_{gc}`$. For most galaxies this also implies $`M_QFM_s`$ where $`M_s`$ is the stellar mass and we use $`F=0.01`$. We also assume that quasars shine at the Eddington limit with a radiative efficiency $`ϵ=0.1`$, which means that the quasar life-time is $`t_Q=4.4ϵ\mathrm{\hspace{0.33em}10}^8`$ yr. This determines the quasar bolometric luminosity: $$L_Q=\frac{ϵM_Qc^2}{t_Q}.$$ (8) We present in Fig.7 the redshift evolution of the quasar luminosity - bias relation. Here $`b(>L_B,z)`$ is the bias of quasars brighter than $`L_B`$ at redshift $`z`$. We see that the bias increases with redshift since at higher $`z`$ quasars of a given luminosity correspond to rarer density fluctuations. The high luminosity cutoff seen in the figure comes from the fact that in our model very massive galaxies have already consumed most of their gas content at low $`z`$, as explained in the previous section, so that the maximum quasar luminosity declines with time because of fuel exhaustion. Of course, at higher $`z`$ the quasar luminosity function cutoff also declines because of the knee of the mass function of collapsed objects (Valageas & Silk 1999a). From the bias $`b(>L_B,z)`$ we can derive the comoving correlation length of quasars. This is shown in Fig.8. We can check that our results agree reasonably well with observations. In particular, we recover the increase with redshift of the correlation length. La Franca et al. (1998) noticed that this behaviour differs from the decrease of $`r_0`$ observed for galaxies from $`z=0`$ to $`z=1`$ which we also obtain, as shown in Fig.6. Thus a simple hierarchical model like ours is able to simultaneously explain both of these redshift dependences. The stronger increase with redshift of the correlation length associated with quasars, as compared with galaxies, also translates into the faster growth of their bias factor, see Fig.4 and Fig.7. This difference between galaxies and quasars is due to the redshift evolution of their respective ratios of (mass)/(luminosity). As described above, in our model the luminosity of a quasar is proportional to its mass of gas, see (8), and to the mass of stars in the host galaxy. Hence the bias of quasars of a given luminosity grows strongly with redshift since objects with a fixed mass of central gas are increasingly rare at earlier times when structure formation is less advanced. This overrides the decrease of the dark matter correlation length with $`z`$ and it leads to a rise of $`r_0(>L_B)`$, as seen in Fig.8. On the other hand, galaxies with a constant mass of star-forming gas had a higher luminosity in the past because the ratio $`M_s/L`$ decreases with redshift (see for instance Fig.17 in Valageas & Schaeffer 1999). This is due to the fact that the smaller age of the universe $`t_H`$ selects more massive and brighter stars with a life-time $`t_{}t_H`$. In other words, at later times the contribution from faint long-lived stars which progressively accumulated in the galaxy ever since it was born becomes increasingly important as compared with the contribution of bright newly-born short-lived stars which quickly disappear as supernovae and thus cannot accumulate. This implies that by looking at a fixed luminosity one selects a smaller mass of gas and stars at higher redshift. Thus, if we start at $`z=0`$ with galaxy and quasar luminosities which correspond to the same host halos, at large redshift $`z>0`$ the galaxies with the same luminosity correspond to smaller halos than the quasar hosts. Hence galaxies correspond to less extreme objects and have a smaller bias. Thus at low $`z`$ their correlation length follows the decreases of the dark matter $`r_0`$, until $`z1`$, beyond which the rise of their bias becomes sufficiently strong to lead to a growth of $`r_0(>L_B)`$, see Fig.6. The precise amplitude of these redshift evolutions depends somewhat on the astrophysical models used for quasars and stars (IMF,…). However, the agreement we obtain with these observations is remarkable since our model was built independently of these considerations, with no free parameter left for this present discussion. Thus this behaviour appears as a natural outcome of such models which suggests they provide a realistic description of the global properties of star and quasar formation, even though the details of these processes are still poorly known. ### 4.3 Lyman$`\alpha `$ absorbers As shown in Valageas et al. (1999), the formalism introduced in Sect.2 also allows us to build a model for Lyman-$`\alpha `$ clouds, some of which are objects with a lower density than the mean density of the universe. We describe these absorbers as three different populations of objects. Low density halos with a small virial temperature see their baryonic density fluctuations erased over scales of the order of the Jeans length $`R_J`$. These mass condensations form our first class of objects, defined by the scale $`R_J`$, which can be identified with the Lyman-$`\alpha `$ forest. They correspond to low density clouds which can even be underdense (but stick out within very low-density regions or voids). On the other hand, potential wells with a large virial temperature maintain their density profile. Thus, they constitute a second population of absorbers where the column density observed along a line of sight depends strongly on the impact parameter. They can be identified with the Lyman-limit systems. Finally, the deep neutral cores of these halos (because of self-shielding) form our third class of objects which correspond to the damped systems. We present in Fig.9 the redshift evolution of the bias associated with Lyman-$`\alpha `$ absorbers above various column density thresholds. We note that large column density absorbers have a bias greater than unity since they correspond to high density fluctuations (massive halos) while forest clouds which are low density fluctuations located in underdense areas have a small bias $`b<1`$. For $`N_{HI}<10^{14}`$ cm<sup>-2</sup> the bias decreases at larger redshift because the higher mean density of the universe implies that to observe a constant small column density one must look at increasingly underdense objects. In a similar fashion, for Lyman-limit or damped systems the bias first declines from $`z=0`$ to $`z=2.5`$. However, since these objects correspond to virialized halos the bias eventually grows with $`z`$ as the structure formation process is less advanced. Of course there is also an intermediate range of $`N_{HI}`$ where the behaviour is more intricate. We display the redshift evolution of the bias associated with clouds above three different column densities thresholds in Fig.10. We see that the redshift evolution of the bias of Lyman-$`\alpha `$ absorbers can display a non-trivial behaviour, since many processes play a role in the properties of these clouds, including the amplitude of the UV background radiation field. Thus, a phenomenological parameterisation of the form $`bN_{HI}^\alpha (1+z)^\beta `$ is unlikely to provide an accurate description. Note however that intermediate clouds in the range $`10^{16}<N_{HI}<10^{21}`$ cm<sup>-2</sup> have a bias which remains reasonably close to unity at all redshifts $`z<5`$ and thus can be used as tracers of the matter density field. A comparison of our results with observations for the two-point correlation function in velocity space is described in Valageas et al. (1999). ### 4.4 Clusters Finally, we show in Fig.11 the redshift evolution of the correlation length associated with clusters above various virial temperature thresholds. Here we simply defined clusters as “just-virialized” halos characterized by the usual density contrast $`\mathrm{\Delta }_c177`$. A detailed description of our model for clusters and the evolution of their X-ray luminosity function is presented in Valageas & Schaeffer (2000). We see that our results agree reasonably well with observations. As was the case for quasars, the rise with redshift of the bias associated to these halos leads to an increase of their correlation length. However, this increase is slower than for quasars, see Fig.8, because here we select these halos by a constant temperature threshold. This implies that their mass evolves as $`M(1+z)^{3/2}`$ hence we consider smaller halos at high $`z`$ which are more common (and have a smaller bias) than objects of constant mass. Thus, we have obtained a simple model of structure formation which is consistent with observations from Lyman-$`\alpha `$ clouds ($`M10^8M_{}`$) up to clusters ($`M10^{15}M_{}`$), i.e. over 7 orders of magnitude. ### 4.5 Distribution of matter The models we have built in previous studies for galaxies (Valageas & Schaeffer 1999) and Lyman-$`\alpha `$ clouds (Valageas et al. 1999) allow us to obtain the redshift evolution of the fraction of baryonic matter associated with various components. We present in Fig.12 the contributions of stars, galactic gas and Lyman-$`\alpha `$ forest clouds. In our model the sum of these three constituents is unity. Indeed, all the matter is embedded within mass condensations which we identify with galaxies (deep potential wells where the gas can cool) or with Lyman-$`\alpha `$ forest clouds (small density fluctuations heated by the UV background radiation). We see that the mass fraction within stars is always small ($`10\%`$ at $`z=0`$) while the contribution of Lyman-$`\alpha `$ forest clouds shows a fast increase with redshift to constitute $`70\%`$ of the matter at $`z=5`$ when there are few deep potential wells. Bright galaxies ($`M_B<19.5`$) contain most of the mass embedded within stars but they only constitute a small part of the gas which was able to cool. Moreover, their total mass of gas starts declining at low redshifts $`z<1`$ as it is converted into stars. This trend does not appear for faint galaxies because their star formation process is less efficient while they accrete gas which was previously part of the Lyman-$`\alpha `$ forest clouds. Moreover, some of these small galaxies have just formed as new potential wells have built up. Note that if the IGM is reheated by supernovae or quasars up to $`T\mathrm{5\hspace{0.33em}10}^5`$ K, which is suggested by the observed relation temperature - X-ray luminosity of clusters (e.g., Ponman et al. 1999), the fraction of matter which has been able to cool and appears as galactic gas in Fig.12 would be smaller. This was studied in details in Valageas & Silk (1999b). We show in Fig.13 the mass fraction we associate with groups and clusters of galaxies, as well as with stars and galaxies. We define groups as virialized halos (i.e. with a density contrast $`\mathrm{\Delta }_c`$) which do not satisfy the constraint $`t_{cool}<t_H`$. They correspond to massive potential wells with a large virial temperature (at high $`T`$ we have $`t_{cool}\sqrt{T}`$). At high redshifts this contribution declines since an increasing proportion of virialized halos can cool (indeed $`t_{cool}1/\rho `$ while $`t_H1/\sqrt{\rho }`$), see also Valageas & Schaeffer (1999). We identify clusters as groups with a virial temperature $`T1`$ keV. At low $`z`$ they only form a small sub-class of galaxy groups while at high $`z`$ all groups have large virial temperatures $`T1`$ keV as explained above. Note that in our model the mass associated with groups is also associated to the member galaxies, so that it is a part of the fraction of matter recognized as embedded within galaxies, shown in Fig.12 and in Fig.13 (upper solid line). A detailed study of groups and clusters in presented in Valageas & Schaeffer (2000). From the fraction of matter $`F_i`$ associated with various objects and the baryonic density parameter $`\mathrm{\Omega }_b=0.04`$ we used in our model, we also obtain the baryonic density parameters $`\mathrm{\Omega }_i=F_i\mathrm{\Omega }_b`$ for various classes of objects. We display in Tab.1 a comparison of our results with observational estimates by Fukugita et al. (1998). We note that our predictions are consistent with the data. In Tab.1 we removed from the mass embedded within clusters and groups shown in Fig.13 the fraction of baryons in stars or galactic cores which they contain. ## 5 Open universe Using the same formalism we can also study the case of an open universe $`\mathrm{\Omega }_\mathrm{m}=0.3`$, $`\mathrm{\Omega }_\mathrm{\Lambda }=0`$, with a CDM power spectrum normalized to $`\sigma _8=0.77`$. We use $`\mathrm{\Omega }_b=0.03`$ and $`H_0=60`$ km/s/Mpc. ### 5.1 Galaxies We show in Fig.14 the correlation function of galaxies brighter than $`M_B=18.6`$ at $`z=0`$ (upper solid line). Our predictions are similar to the results obtained in Fig.3. However, a bump appears at intermediate scales ($`R2`$ Mpc) where $`\overline{\xi }10`$. This corresponds to the sharp increase of $`\overline{\xi }(R)`$ from the linear regime ($`\overline{\xi }1`$) to the non-linear regime ($`\overline{\xi }100`$). This leads to a “step-like” profile which is also seen in numerical simulations (e.g., Valageas et al. 2000) and which gets larger for low density universes (Peacock & Dodds 1996). This is due to the slow-down of the linear growth factor while the two-point correlation function keeps increasing as $`\overline{\xi }(R,a)a^3`$ in the highly non-linear regime (Valageas & Schaeffer 1997). The observed two-point correlation function on the other hand is surprisingly close to a power-law as soon as $`\overline{\xi }1`$. Such a behaviour requires a precise balance between the curvature of the initial power-spectrum, the distortions induced by the non-linear dynamics and the bias of galaxies (which also depends on non-gravitational processes which affect galaxy formation). Note however that using numerical simulations Pearce et al. (1999) and Benson et al. (2000a) obtain a good match to observations as the galaxies they get are antibiased relative to the dark matter on small scales which “conspires” to provide a power-law galaxy correlation function. This behaviour is partly due to the spatial exclusion of halos. Moreover, the expression (5) we use for the bias is only valid at large separations $`rR`$ between objects of size $`R`$. Thus, the discrepancy at small scales might be cured by a more rigorous calculation which would not use the approximation $`rR`$. On the other hand, note that in the case of a critical density universe this “bump” in the mildly non-linear regime is much smaller, which leads to good agreement with observations (see Fig.3). In particular, we also display in Fig.14 the dark matter correlation function we would obtain for a low-density flat universe (upper dashed line) with $`\mathrm{\Omega }_\mathrm{m}=0.3`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }=0.7`$, with a slightly different power-spectrum from Bardeen et al. (1989) (and $`\mathrm{\Gamma }=0.18`$) using the non-linear fit from Peacock & Dodds (1996). Thus this corresponds to a bias of unity, which is a good approximation for galaxies with $`M_B18.6`$ at $`z=0`$, see Fig.1 and Fig.15. The main point is that the “bump” is much weaker for a flat cosmology, as can also be seen in Peacock & Dodds (1996). The intermediate lines show the dark matter correlation functions obtained for $`\mathrm{\Omega }_\mathrm{m}=0.5`$ (with $`\mathrm{\Omega }_\mathrm{\Lambda }=0`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }=0.5`$) and normalized to $`\sigma _8=0.6`$ (and $`\mathrm{\Gamma }=0.3`$). Thus, the shape itself of the galaxy correlation function does not allow one to strongly discriminate between these various cosmological scenarios. Although the “bump” which appears for the open universe with $`\mathrm{\Omega }_\mathrm{m}=0.3`$ seems to disfavour this model, the inaccuracy involved in the modelling of the galaxy clustering properties is certainly too large to draw meaningful conclusions, especially since numerical simulations suggest that simple physical processes could remove this feature (Benson et al. 2000a). We present in Fig.15 the redshift evolution of the luminosity-bias relation. We can see that our results are again similar to those obtained in Fig.4, since the astrophysical model is the same and it must satisfy the same observational constraints at low $`z`$. However, the bias of bright galaxies increases more slowly with redshift than for a critical density universe (compare the curve at $`z=4`$ with Fig.4). This is due to the higher normalization $`\sigma _8`$ of the power-spectrum and to the slower redshift evolution of the linear growth factor which means that massive galactic halos correspond to less extreme density fluctuations at $`z=4`$ than for the case $`\mathrm{\Omega }_\mathrm{m}=1`$ when they have a similar clustering pattern at $`z=0`$. We show in Fig.16 the redshift evolution of the comoving correlation length $`r_0`$ associated with galaxies below various B-band magnitude thresholds. The galaxy correlation length is larger than for the previous case of a critical density universe because the normalization of the power-spectrum ($`\sigma _8`$) is slightly higher. This also implies a slower redshift evolution. We note that the clustering pattern we predict for Lyman break galaxies is somewhat too strong as compared to observations. However, this could be due to some extinction by dust of the luminosity of these LBG. Indeed, in our model the number of galaxies brighter than $`M_B=18`$ is $`n=3\times 10^3`$ Mpc<sup>-3</sup> while the abundance of LBG is $`n8.5\times 10^4`$ Mpc<sup>-3</sup> (Adelberger at al. 1998). This means that we would recover the right correlation length if we assumed that LBG correspond to $`30\%`$ of these galaxies with $`M_B18`$. An alternative would be that LBG correspond to galaxies which are in the midst of a starburst phase (e.g., Somerville et al. 2000) which increases their luminosity (note that the absolute LBG magnitude is rather large: $`M_R21.8`$ for an apparent magnitude $`25.5`$ without K-correction). In this case, Lyman-break galaxies would also be drawn from a larger parent population which could again lead to a smaller correlation length. Note on the other hand that it would be difficult to reconcile the model with observations if we had predicted a smaller value of $`r_0`$ than the one given by the data. ### 5.2 Distribution of matter We show in Fig.17 the contributions of stars, galactic gas and Lyman-$`\alpha `$ clouds to the baryonic density in the universe as a function of redshift. We obtain a behaviour similar to Fig.12. We display in Fig.18 the redshift evolution of the fraction of matter embedded within galaxies, groups, clusters and stars. We see that the mass fraction formed by groups and clusters declines more slowly than in Fig.13. Again this is due to the higher normalization $`\sigma _8`$ and to the slower redshift evolution of the linear growth factor. We show in Tab.2 the baryonic density parameters $`\mathrm{\Omega }_i`$ associated with various objects. We note that our results still agree reasonably well with observations. ## 6 Conclusion In this article we have described the redshift evolution of the clustering of various objects (galaxies, quasars, Lyman-$`\alpha `$ clouds, clusters) within the framework of a unified model of structure formation which has already been checked against many observations. Hence there are no free parameters chosen to fit the observations considered in this article, since the model was already overconstrained by previous studies. We have shown that the bias of these astrophysical objects displays an intricate behaviour which cannot be described by simple power-law parameterizations. Indeed, the link between these objects and the underlying dark matter component reflects the influence of many factors (cooling processes, star formation efficiency, evolution of the UV background radiation field…) which can have competing effects. We have shown that in our model the bias associated with galaxies strongly increases with their luminosity. This was already emphasized in the earlier calculations (Schaeffer 1987; Bernardeau & Schaeffer 1992) as being a genuine outcome of the hierarchical clustering hypothesis. We recover the trend seen in observations, although our dependence may be slightly too strong. However, this could be due to the simplicity of our model which does not include any scatter in the properties of the objects we model. The redshift evolution we obtain also agrees with data, in particular the correlation length of Lyman-break galaxies observed at $`z=3`$ is consistent with our results. We have also described the bias associated with Lyman-$`\alpha `$ clouds. Thus, it appears that absorbers in the range $`10^{16}<N_{HI}<10^{21}`$ cm<sup>-2</sup> have a bias close to unity (hence they trace the dark matter density field) in contrast to other clouds and galaxies which show a very different behaviour of their clustering properties. Our results also agree with observations for quasars and clusters. In particular, they explain in a natural fashion the qualitative and quantitative differences between the redshift evolution of the clustering of QSOs and of galaxies. Thus, our model provides a good description of the intrinsic properties and the spatial clustering of a wide variety of objects, from small underdense Lyman-$`\alpha `$ absorbers up to massive clusters, which covers six orders of magnitude in mass. Finally, we have described the redshift evolution of the distribution of matter over the various objects observed in the universe. Our results show a reasonable agreement with observations. This is not surprising since our model has already been checked against the galaxy luminosity function, the number density of Lyman-$`\alpha `$ clouds, which constrain the amount of matter embedded within these objects. In particular, it appears that light does not trace baryonic mass as bright galaxies ($`M_B<19.5`$) which contain most of the stars only provide a small fraction of the mass associated with virialized and cooled halos. We have also performed the same analysis for the case of an open universe. In both cosmologies our results agree reasonably well with observations but the details of astrophysical models are still too uncertain to discriminate between these two cases. However, for the open universe the galaxy correlation function we obtain shows a shoulder which does not appear in observations. This could be due to the approximation involved in our calculation which is only valid on scales much larger than the size of the objects. In particular, some numerical simulations do not get this “bump” (Benson et al. 2000a). On the other hand, they do not recover the luminosity dependence of the galaxy bias (Benson et al. 2000b) and simulations cannot simultaneously describe the clustering of the whole range of objects we studied in this paper. Nevertheless, the fact that our predictions are nearly consistent with the data over such a wide range of objects shows that hierarchical scenarios provide a very good description of structure formation. On the other hand, more detailed observations of the amplitude of the clustering of galaxies selected by their luminosity will constrain the astrophysical models. ###### Acknowledgements. Much of this work was performed at the Center for Particle Astrophysics, University of California, Berkeley, whose hospitality we gratefully acknowledge.
warning/0001/astro-ph0001097.html
ar5iv
text
# HST NICMOS Observation of Proto-Planetary Nebulae ## 1. Introduction Although the circumstellar envelopes of asymptotic giant branch (AGB) stars are mostly spherical, many planetary nebulae (PN) have bipolar or butterfly morphologies. When and how this morphological transformation occurs are two of the major unsolved problems in PN research today. Recent imaging observations of PPN have found that several PPN have already developed reflection nebulosities of bipolar shape, suggesting that shaping occurs soon after the end of the AGB phase (Kwok et al. 1996, Kwok et al. 1998, Su et al. 1998, Hrivnak et al. 1999). In this paper, we present HST NICMOS observations of four of these bipolar PPN. ## 2. Observations & Data Reduction The NICMOS observations (ID No. 7840; S. Kwok, PI) were obtained in 1998 with the NICMOS camera 2 (NIC2), which has the highest resolution ($`0{}_{}{}^{\prime \prime }.076pixel^1`$) and a 19<sup>′′</sup>$`\times `$ 19<sup>′′</sup> field of view. The observations were taken with one broad-band (F160W) and one medium-band (F222M) filters, three polarizers POL0, POL120, POL240, and the narrow-band filter F212N (H<sub>2</sub>) and F215N (continuum). Each target was imaged three times using a predefined spiral dither pattern, which allowed us to compensate for bad pixel/columns and the area blocked by the coronographic mask. ## 3. H & K Wide-band Images The reduced images of IRAS 17150-3224 (the Cotton Candy Nebula) are shown in Fig. 1. The two bipolar reflection lobes are found to have similar sizes in the V, H, and K bands. Superimposed on the two lobes and the surrounding halo is a series of concentric, circular arcs or rings. There are a total 8 rings (A to H, Fig. 1d) in the V-band image (Kwok et al. 1998), 6 of which are also detected in the H-band. The average separation of the 6 arcs in H is 0.51<sup>′′</sup>, which agrees with the results measured in the V-band. While the two lobes are separted by a dark lane in the V image, the central star can be seen in the H and K images. The position of the central star agrees well with the position of the crossing point of the searchlight beams in the V-band image. A linear plot (from 10 $`\sigma _{sky}`$ to 100 $`\sigma _{sky}`$) of the continuum-subtracted H<sub>2</sub> image is shown in Fig. 1 (e). The clumpy H<sub>2</sub> emission regions at the end of the lobes could be the result of shock excitation as a fast, collimated outflow runs into the remnant AGB envelope. The V, H, and K images of IRAS 17441-2441 (the Silkworm Nebula) are shown in Fig. 1 (f), (g) and (h) respectively. The arcs in the V-band image can also be seen in the H-band. The S-shaped morphology is more prominent in the K-band. A bright central star is resolved in the K-band image at the sky position which agrees with the center of the circular arcs as determined from the V-band image. Figure 2 shows the results of IRAS 17245-3951 (the Walnut Nebula) and IRAS 16594-4656 (the Water Lily Nebula). The central star in the Walnut Nebula is obscured by a dark lane in the V-band image but is visible in the H- and K-band images. For the Water Lily Nebula, very little nebulosity is detected in the K-band image. However, the nebulosity in the H-band image bears some resemblance to that seen in V-band image. ## 4. 2 $`\mu `$m Polarized Images We used the algorithm developed by Hines et al. (1997) to derive the Stokes images $`(I,Q,U)`$, the percentages of polarization, and the position angles in the NICMOS polarization images. In order to minimize the noise contributions to the calculation, a threshold value of 10 $`\sigma _{sky}`$ was used for each pixel. The data were binned by 2$`\times `$2 pixels before calculating polarization percentage and position angle. All of the polarized vector maps (Fig. 3) show centrosymmetric patterns, similar to those observed in the equatorial regions of bipolar reflection nebulae (Kastner et al. 1995) and the Egg Nebula (Sahai et al. 1998). The presence of such patterns provides strong, independent evidence for the presence of a circumstellar disk in these PPN. The higher polarization regions are found at the edges of two reflection lobes far from the central stars, suggesting that the reflection nebulae are the result of single scattering processes. The polariztion is low in the equatorial regions, implying that the light is either unpolarized or the result of multiple scattering. The centroids of the polarimetric vectors of all four PPN coincide with the positions of the central stars in the K-band images, confirming that the central stars are the source of the scattered light. ## References Hines, D.C. 1998, NICMOS and the VLT Workshop, Freudling and Hook eds., p62 Hrivnak, B. J, Kwok, S., & Su, K. Y. L. 1999, ApJ, 524, in press. Kastner, J.H., & Weintraub, D.A. 1995, AJ, 109, 1211 Kwok, S., Hrivnak, B. J., Zhang, C. Y., & Langill, P. P. 1996, ApJ, 472, 287 Kwok, S., Su, K. Y. L., & Hrivnak, B. J. 1998, ApJ, 501, L117 Sahai, R. et al. 1998, ApJ, 492, L163 Su, K.Y.L., Volk, K., Kwok, S. and Hrivnak, B.J. 1998, ApJ, 508, 744
warning/0001/cond-mat0001168.html
ar5iv
text
# Raman Scattering from Magnetic Excitations in the Spin Ladder Compounds CaV2O5 and MgV2O5 ## I Introduction The spin-1/2 ladder systems, which are the intermediate situation between one- dimensional (1D) and two-dimensional Heisenberg antiferromagnet (HA), show fascinating quantum effects such as dramatic dependence of the ground state on the width of the ladder (the width of the ladder is defined as a number of coupled chains) . Namely, ladders made of even number of legs have spin-liquid ground state with an exponential decay of the spin-spin correlations produced by the finite spin gap. Among them, the two-leg ladder is the first member of the family, with a largest spin gap of about 0.5 J (assuming the same coupling constant along the legs and the rungs). Realizations of the ladders are recognized in Sr-Cu-O structures , and recently in AV<sub>2</sub>O<sub>5</sub>, A=Ca, Mg (two-leg ladder) . According to their crystal structures, vanadium oxides are quasi-2D layered materials, with spin-1/2 vanadium ions which form two-leg ladders coupled in the trellis lattice. The spin gap is observed in CaV<sub>2</sub>O<sub>5</sub> to be around 600 K , and the magnetic susceptibility measurements demonstrated almost perfect agreement with spin-1/2 two-leg ladder HA model. On the other hand, there is only contradicting evidence for the existence of the spin- gap in MgV<sub>2</sub>O<sub>5</sub> . The excitation spectrum of the ladder has been analyzed by a variety of the mathematical techniques. However, it was recently shown that the two-magnon bound state (we will use word magnon although classical magnons do not exist in quantum systems), besides usual triplet (gap mode) and two-magnon continuum, may appear in the excitation spectra . Such a bound state is also predicted for the dimerized quantum spin chain and indeed observed in CuGeO<sub>3</sub> , using a Raman-scattering technique. These singlet-singlet transitions (the 30 cm<sup>-1</sup> mode in the Raman spectra of CuGeO<sub>3</sub> is believed to be the transition between ground state and the two-magnon bound state which are both singlets) are allowed and take place via exchange scattering mechanism . On the contrary, the Raman scattering process on spin-gap excitations (one-magnon scattering) takes place via spin- orbit interaction , and it is usually negligible due to quenched orbital momentum in the transition metal ions. It is therefore, desirable to study Raman scattering in magnetic ladder materials such as CaV<sub>2</sub>O<sub>5</sub> and MgV<sub>2</sub>O<sub>5</sub>, since these experiments provide information about the spin-dynamics and the type of magnetic order. In this paper, we present a study of the spin dynamics in CaV<sub>2</sub>O<sub>5</sub> and MgV<sub>2</sub>O<sub>5</sub> using Raman scattering technique. We analyze the origin of the magnetic modes in the spectra and discuss the difference between magnetic ordering in these two compounds. ## II Experiment The measurements were performed on polycrystalline samples, prepared as described in Ref. . As an excitation source we used the 514.5 nm line of an Ar<sup>+</sup> ion laser. An average power of 5 mW was focused with microscope optics on the surfaces of the pellets. The temperature was controlled with a liquid helium micro-Raman cryostat (CryoVac, Vacuubrand GMBH). The spectra were measured in back-scattering geometry using a DILOR triple monochromator equipped with a CCD camera. ## III Raman spectra of CaV<sub>2</sub>O<sub>5</sub> CaV<sub>2</sub>O<sub>5</sub> has an orthorhombic unit cell, space group Pmmn , with a crystalline structure formed by layers of VO<sub>5</sub> square pyramids. The Ca ions are situated between these layers, see Fig.1. The pyramids are mutually connected via common edges and corners making characteristic V zigzag chains along the b axis. All vanadium atoms are in 4+ valence state thus carrying the spin=1/2. Effectively, this structure can be regarded as composed from spin-1/2 two-leg ladders connected with each other in the trellis lattice. CaV<sub>2</sub>O<sub>5</sub> is isostructural with NaV<sub>2</sub>O<sub>5</sub>. Therefore, the same irreducible representations correspond to the vibrational modes of CaV<sub>2</sub>O<sub>5</sub> and NaV<sub>2</sub>O<sub>5</sub> : $`\mathrm{\Gamma }_{opt}`$=8A<sub>g</sub>(aa,bb,cc)+3B<sub>1g</sub>(ab)+8B<sub>2g</sub>(ac)+5B<sub>3g</sub>(bc)+ 7B<sub>1u</sub>(E$`||`$c)+4B<sub>2u</sub>(E$`||`$b)+7B<sub>3u</sub>(E$`||`$a) The room- temperature Raman spectra of CaV<sub>2</sub>O<sub>5</sub> are presented in Fig. 2. Comparing the phonon averaged intensities, of HH, HV and VV scattering configurations, with the polarized Raman scattering spectra of NaV<sub>2</sub>O<sub>5</sub> single crystals we assign the phonon modes of CaV<sub>2</sub>O<sub>5</sub>. The phonon frequencies of CaV<sub>2</sub>O<sub>5</sub> are found to be close to those in NaV<sub>2</sub>O<sub>5</sub>, as expected from similarity between lattice constants and interatomic distances in these two oxides. The low-temperature spectra of CaV<sub>2</sub>O<sub>5</sub> are shown in Fig. 3. In addition to phonons we found a strong asymmetric line at 795 cm<sup>-1</sup> (1137 K) with a tail towards higher energies. This line decreases in intensity and broadens by increasing the temperature, together with the intensity decrease of the associated continuum, see the left inset in Fig. 3. Moreover, its energy is in the range of 2$`\mathrm{\Delta }_s`$, estimated from the spin-gap measurements, thus suggesting its magnetic origin. Below, we will show by detailed discussion of the experimental results that this is indeed the case, i.e. that 795 m<sup>-1</sup> feature corresponds to the onset of the two-magnon continuum at $`2\mathrm{\Delta }_s`$. The magnetic susceptibility measurements show a broad maximum around 400 K, with a decrease towards zero at low temperatures, thus reflecting the low dimensionality of the magnetic interaction, the singlet ground state and the finite spin gap. The spin gap is found to be around 660 K using electron spin resonance and at around 500 K using nuclear magnetic resonance experiments. Since the magnon dispersion is not reported in the literature, the exchange couplings are estimated by comparison of the calculated and measured temperature dependencies of the susceptibility. The Monte Carlo simulations showed that the magnetic properties of CaV<sub>2</sub>O<sub>5</sub> can be indeed described using a coupled ladder spin-1/2 HA model (trellis lattice) with in-rung exchange (J$`{}_{}{}^{}`$ 600 K) five times or even an order of magnitude larger then the exchange along the legs. The interladder exchange is found to be negligible. Thus we start analyzing our Raman spectra using Heisenberg Hamiltonian for the ladder model: $$H=J_{}\underset{i,jrungs}{}𝐒_i𝐒_j+J_{}\underset{i,jlegs}{}𝐒_i𝐒_j$$ (1) where the $`𝐒_i`$ represents a spin-1/2 operator at site i of the ladder. The properties of the magnetic excitations in the Heisenberg ladder have been obtained using perturbation method in the strong coupling limit . In this limit $`J_{}>J_{}`$ the lowest spin-triplet dispersion is : $`\omega (k)=J_{}+J_{}cos(k)+\frac{3}{4}\frac{J_{}^2}{J_{}}+\mathrm{}`$ The minimum and the maximum energies of the triplet are at k=$`\pi `$ ($`\omega J_{}J_{}`$) and at k=$`0`$ ($`\omega J_{}+J_{}`$). Thus the gap is $`\mathrm{\Delta }_sJ_{}J_{}`$, and the total width of the branch $`2J_{}`$. Therefore, from the measurements of the energy onset and the width of the two-magnon continuum we could uniquely obtain the values of the exchange integrals $`J_{}`$ and $`J_{}`$. As we have already mentioned, the two-magnon excitations in the Raman spectra can be observed through the exchange-scattering process. Then, the well defined peak around 2$`\mathrm{\Delta }_s`$ is expected in the Raman spectra of the ladder, due to the singularities of the one-dimensional density of (two-magnon) states (DOS) , or due to the appearance of the magnon bound state . In the first case the peak position is exactly 2$`\mathrm{\Delta }_s`$ while in the second case the peak can be found at 2$`\mathrm{\Delta }_s\delta `$, $`\delta `$ being the magnon binding energy. We believe that the 795 cm<sup>-1</sup> line in CaV<sub>2</sub>O<sub>5</sub> Raman spectra does not correspond to a magnon bound state due to the following reasons: (1) The frustration in this material is expected to be small , and the binding energy is then estimated to be small . (2) The intensity of the 795 cm<sup>-1</sup> line saturates at temperatures below 80 K. In the case of the bound state the linear dependence of its intensity as a function of the temperature is believed to be the fingerprint of it, without saturation up to very low temperatures . Thus, we assign the onset of the magnon continuum to 2$`\mathrm{\Delta }_s`$=795 cm<sup>-1</sup> and obtain the spin-gap in CaV<sub>2</sub>O<sub>5</sub> to be $`\mathrm{\Delta }_s`$=398 cm<sup>-1</sup> (570 K). Moreover, the width of the continuum is estimated to be around 200 cm<sup>-1</sup>, see left inset of Fig. 3. The shape of the magnon continuum in the spectra is compared with classical noninteracting two-magnon spin-wave calculation (DOS), $`I_k\delta [E2\omega (k)]`$, and with numerical simulations of the spin-ladder model using exact diagonalization technique. These spectra are presented in the right inset of Fig. 3. As expected, the DOS (full curve), obtained using $`J_{}`$ 640 K and $`J_{}`$ 70 K, failed to explain the nonobservation of the van Hove singularity, associated with the top of the two-magnon branch. It is well known, that in the ideal one-dimensional systems magnons do not behave classically but form a continuum of excitations . This effect is clearly seen in exact diagonalization result (dashed curve, in the right inset of Fig.3) obtained with $`J_{}/J_{}`$ 0.1 in Ref. . However, a discrepancy between the theory and the experiment is still found around 1000 cm<sup>-1</sup>. The possible explanation may lay in the fact that the exact diagonalization result, has been obtained strictly for the two-magnon scattering process, thus neglecting higher-order magnon contributions. Also, we have to note, that the inclusion of the higher-order magnon processes would produce the scattering intensity above the two-magnon high-energy cutoff but it will still fail to explain the weak broad peak observed in the spectra around 980 cm<sup>-1</sup>. We suggest that such a shape of the two-magnon continuum may come from the next-nearest-neighbor interactions, which were not taken into consideration in the numerical simulations. ## IV Raman spectra of MgV<sub>2</sub>O<sub>5</sub> MgV<sub>2</sub>O<sub>5</sub> has an orthorhombic unit cell, space group Cmcm , with a similar V-O layered structure as in CaV<sub>2</sub>O<sub>5</sub>. The structure can be described as a linkage of VO<sub>5</sub> pyramids, with apical oxygens along the b axis. The V zigzag chains run along the a axis. The factor group analysis (FGA) gives $`\mathrm{\Gamma }_{opt}`$=8A<sub>g</sub>(aa,bb,cc)+5B<sub>1g</sub>(ab)+5B<sub>2g</sub>(ac)+6B<sub>3g</sub>(bc)+ 7B<sub>1u</sub>(E$`||`$c)+7B<sub>2u</sub>(E$`||`$b)+4B<sub>3u</sub>(E$`||`$a) The room temperature and the T=10 K Raman spectra of MgV<sub>2</sub>O<sub>5</sub> are shown in Fig.4. In addition to phonon excitations, which we will not discuss here, we observe the continuumlike asymmetric mode around 340 cm<sup>-1</sup>. This structure shows large intensity and frequency dependence as a function of the temperature. The temperature dependence of the spectra in the frequency range from 50 to 450 cm<sup>-1</sup> is shown in the inset of Fig. 4. The continuum clearly extends from 150 to 350 cm<sup>-1</sup> with a spectral weight cut off around 350 cm<sup>-1</sup>. This mode shifts to lower energies, as well, and decreases in intensity by increasing temperature. Such a form of the spectra and its temperature dependence suggests the two-magnon origin of the continuum, with a peak at 340 cm<sup>-1</sup> corresponding presumably to zone boundary magnons. On the other hand, no feature is found in the vicinity of 2$`\mathrm{\Delta }_s`$ 21 cm<sup>-1</sup> (30 K) - 32 cm<sup>-1</sup> (45 K), so the existence of the spin gap is not confirmed by our results. However, we have to note that this energy range is close to the frequency limit (10 cm<sup>-1</sup>) of our experiment. However, the existence of the large zone boundary magnon spectral weight in the Raman spectra, reflects strong deviation of the magnetic ordering in MgV<sub>2</sub>O<sub>5</sub> from the 1D two-leg-ladder type observed in CaV<sub>2</sub>O<sub>5</sub>. In principle, such a deviation may be a consequence of the two effects (1) There is a spin-gap in MgV<sub>2</sub>O<sub>5</sub> (according to our measurements it must be less than 10 cm<sup>-1</sup>) and the values of the exchange constants along legs and rungs are either very close to each other or the rung exchange is much smaller that the leg exchange (the latter case correspond to homogenous Heisenberg antiferromagnetic chain with a large frustration). (2) There is no spin-gap and magnetic ordering is quasi-2D. For the first case, since the existence (and the value) of the gap is still an open question and since all measurements are performed on polycrystalline samples, it is hard to make any estimation of the exchange integrals. On the other side, if we assume 2D magnetic ordering, for example, the effective exchange constant can be estimated from the position of the two-magnon peak using relation $`E_{twomagnon}=2.7J`$. In this case we obtain J=125 cm<sup>-1</sup> (180 K), which is close to the energy $`J_{}=`$ 144 K obtained in Ref. . However, this formula is strictly valid only for 2D square lattice with the isotropic exchange, which is clearly not the case in MgV<sub>2</sub>O<sub>5</sub>, even though the exchange couplings (along legs and rungs) are found to be all of the same order . Because of that, we believe that neither one of the above mentioned models are applicable to MgV<sub>2</sub>O<sub>5</sub>, since this compound belongs to the class of strongly frustrated trellis lattices. Unfortunately, the existence of the spin-gap and the type of magnetic ordering in this type of lattice are not yet known . ## V Conclusion In conclusion, we present the Raman-scattering spectra of CaV<sub>2</sub>O<sub>5</sub> and MgV<sub>2</sub>O<sub>5</sub>. In CaV<sub>2</sub>O<sub>5</sub> we found the onset of the two-magnon continuum at the energy 2$`\mathrm{\Delta }_s`$ $``$ 795 cm<sup>-1</sup> in the form of the strong asymmetric line with a tail on the high-energy side. From the analysis of its spectral shape, we argue that the magnetic ordering in CaV<sub>2</sub>O<sub>5</sub> can be described using a S=1/2 two-leg ladder Heisenberg antiferromagnetic model with $`J_{}/J_{}`$ 0.1, and a small interladder exchange. The spin gap and exchange constant are estimated to be $`\mathrm{\Delta }_s`$ $``$ 400 cm<sup>-1</sup> (570 K) and $`J_{}`$ 640 K. No magnon bound states are found. In contrast to CaV<sub>2</sub>O<sub>5</sub> the the existence of the spin gap is not confirmed in MgV<sub>2</sub>O<sub>5</sub>. Instead, we observe a strong two-magnon excitation at about 340 cm<sup>-1</sup> corresponding presumably to zone boundary magnons. These results reflects strong deviation of the magnetic ordering in MgV<sub>2</sub>O<sub>5</sub> from the 1D two-leg-ladder type observed in CaV<sub>2</sub>O<sub>5</sub>. Acknowledgment Z.V.P. acknowledges support from the Research Council of the K.U. Leuven and DWTC. The work at the K.U. Leuven is supported by Belgian IUAP and Flemish FWO and GOA Programs. MJK thanks Roman Herzog - AvH for financial support. $``$ present address: Physics Department, Simon Fraser University, 8888 University Drive, Burnaby B.C. Canada V5A1S6.
warning/0001/hep-th0001221.html
ar5iv
text
# A Renormalization Group Study of the (ϕ^∗⁢ϕ)³ Model coupled to a Chern-Simons Field*footnote **footnote *Copyright 1999 by The American Physical Society ## I Introduction Self interacting scalar fields are the simplest nontrivial field theories. Nevertheless they have found large application in many different phenomena. Renormalization group analyses of the model of scalar fields in $`(2+1)`$ dimensions, with a self interaction of the form $`\lambda \varphi ^6`$ have appeared in the literature in conjunction with other self interactions, and also in interaction with other fields. On the other side, the Chern-Simons (CS) field theory is known to cause some strange effects in matter fields, the most known being the transmutation of their spins and statistics . Bosons (fermions) interacting with a CS field get an extra contribution to their spins and statistical phases, changing to anyons and even to fermions (bosons). Studies of the change in the scale behavior of matter fields due to their interaction with the CS field have also been considered . In this paper we study the model of a massless charged scalar field with a self interaction of the form $`\lambda (\varphi ^{}\varphi )^3`$ and interacting with an Abelian CS field. Classically it only involves dimensionless parameters and is scale invariant. It is also strictly renormalizable: no induction of terms of the forms $`m^2(\varphi ^{}\varphi )`$ or $`g(\varphi ^{}\varphi )^2`$ occurs. Besides the calculation of the anomalous dimensions of $`\varphi `$ and $`A^\mu `$ and the $`\beta `$ functions related to their coupling constants, we also calculate the anomalous dimensions of composite operators of the form $`(\varphi ^{}\varphi )^n`$. Some of our conclusions agree and others disagree with the previous literature. This will be discussed in section III and in the Conclusions. To regulate the ultraviolet (UV) behavior we use a simplified version of dimensional regularization, the so called “Dimensional Reduction” method. It consists of contracting and simplifying the Lorentz tensors, before extending the Feynman integrals out of 3 dimensions. This procedure, previously used by several authors , greatly simplifies calculations involving the CS field, because it does not require the extension of the Levi-Civita tensor $`\epsilon ^{\mu \nu \rho }`$ out of $`3`$ dimensions. In Feynman integrals only involving scalar vertices and propagators, no difference appear between the results gotten by using one or the other method. In graphs involving the CS field and the $`\epsilon ^{\mu \nu \rho }`$, the differences of this method to a “full” dimensional regularization would only show up in sub-leading contributions to the Feynman integrals; that is, if $`D`$ stands for the extended dimension of the space time when the Feynman integrals are expanded in Laurent series in $`ϵ(D3)`$, no difference in the leading divergent term in $`1/ϵ`$ will appear. It is, on the other side, a characteristic of dimensional regularization in $`(2+1)`$ dimensions, that one loop graphs are finite, and 2 loops graphs have at most a single pole divergence in $`ϵ`$. As the calculation of the renormalization group parameters only involve the use of the divergent parts of the graphs, no differences to the full dimensional regularization is expected up to 2 loops in graphs that involve the CS propagator, and any number of loops in graphs only involving the scalar propagator. In this paper we will restrict the calculations to up 2 loops in all graphs involving propagators of the CS field, and 4 loops in graphs involving only the propagator of the scalar field. As we will explicitly show that ( at least ) to that orders, dimensional reduction is enough to regularize the model and to preserve the gauge symmetry, as expressed by the Ward Identities (WI). We will work in the Landau gauge and, avoiding exceptional momenta, no infrared (IR) divergences appear. The plan of the paper is as follows. In section II the model is presented, and the divergent UV counterterms, necessary for the renormalization group study, are obtained by calculating the CS 2 point function and the scalar field 2 and 6 point functions. In section III the renormalization group $`\beta `$ functions and anomalous dimensions of the fields are obtained, and compared with other calculations. The change in the dynamical behavior of the $`\varphi `$ field due to the interaction of the CS field is discussed. The influence of the CS in the dimension and renormalizability of operators of the form $`(\varphi ^{}\varphi )^n`$ is also studied. A summary of the results are presented in the Conclusions. In Appendix A the explicit verification of the WI is given, and in Appendix B some Feynman integrals are calculated as examples. ## II The Model The model is constituted by a massless charged boson in $`2+1`$ dimensions represented by a field $`\varphi `$, with a self interaction of the form $`(\varphi ^{}\varphi )^3`$ and minimally interacting with a Chern-Simons (CS) field $`A_\mu `$. Its Lagrangian density is given by $``$ $`=`$ $`_\mu \varphi _0^{}^\mu \varphi _0ie_0A_0^\mu (\varphi _0^{}_\mu \varphi _0_\mu \varphi _0^{}\varphi _0)+e_0^2A_0^\mu A_{0\mu }(\varphi _0^{}\varphi _0)`$ (1) $``$ $`{\displaystyle \frac{\lambda _0}{6^2}}(\varphi _0^{}\varphi _0)^3+{\displaystyle \frac{1}{2}}\epsilon _{\mu \nu \rho }A_0^\mu ^\nu A_0^\rho .`$ (2) The metric is $`g_{\mu \nu }=(1,1,1)`$, $`_\mu `$ stands for $`\frac{}{x^\mu }`$, $`\epsilon _{\mu \nu \rho }`$ is the antisymmetric Levi-Civita tensor with $`\epsilon ^{012}=1`$, and $`e_0`$ and $`\lambda _0`$ are dimensionless coupling constants. The subscript “$`0`$” means that the corresponding quantity is “unrenormalized”. The model is renormalizable: all the UV infinities of the perturbative series can be absorbed in a redefinition of the unrenormalized quantities. It also has a gauge symmetry what suggests the use of dimensional regularization . However, the presence of the Levi-Civita tensor in the CS term makes dimensional regularization cumbersome and the calculations become awkward in more than one loop. We will so, take advantage of some characteristics of $`(2+1)`$ dimensions and use a simplified version of Dimensional Regularization, the so called Dimensional Reduction. In this procedure, the Lorentz tensor algebra is considered in $`(2+1)`$ dimensions and only the remaining scalar Feynman integrals are extended out of $`(2+1)`$ dimensions. It was verified in , up to 2 loops, that for the non-Abelian Chern-Simons theory, this procedure, in fact preserves the Slavnov-Taylor identities. As we will also show below up to 2 loops, it also preserves the Ward identities in our model, and no inconsistencies appear. To get information on the asymptotic behavior of the model, we need to calculate the renormalization group parameters: $`\beta `$ functions and anomalous dimensions of the fields. For this task, adopting the Renormalization Group approach of t’Hooft based on minimal subtraction, we only need to calculate the divergent parts of some vertex functions, more precisely the residues of the poles in $`1/ϵ`$, where $`ϵ=3D`$ and $`D`$ is the “extended”dimension of the space time. In $`(2+1)`$ dimension this means that we must go to at least 2 loops calculations, because as a characteristic of dimensional regularization, one loop integrals are finite. Introducing the renormalized fields $`\varphi `$ and $`A_\mu `$ and the renormalized coupling constants $`e`$ and $`\lambda `$ through the definitions $$\varphi _0=Z_\varphi ^{\frac{1}{2}}\varphi =(1+A)^{\frac{1}{2}}\varphi $$ (3) $$A_0^\mu =Z_A^{\frac{1}{2}}A^\mu =(1+B)^{\frac{1}{2}}A^\mu $$ (4) $$e_0=e\mu ^{\frac{ϵ}{2}}(1+D)/Z_\varphi Z_A^{\frac{1}{2}}$$ (5) $$e_0^2=e^2\mu ^ϵ(1+E)/Z_\varphi Z_A$$ (6) $$\lambda _0=\mu ^{2ϵ}(\lambda +C)/Z_\varphi ^3$$ (7) where $`\mu `$ is a mass parameter introduced to keep $`e`$ and $`\lambda `$ dimensionless quantities, and $`A`$ to $`E`$ are the counterterms to be chosen so as to make the renormalized quantities finite, in each order of perturbation. As will be seen in the calculations, the renormalization of $`\lambda `$ in presence of the CS field is not multiplicative. By substituting these definitions in (2) we get for $``$: $``$ $`=`$ $`_\mu \varphi ^{}^\mu \varphi +{\displaystyle \frac{1}{2}}\epsilon _{\mu \nu \rho }A^\mu ^\nu A^\rho ie\mu ^{\frac{ϵ}{2}}A^\mu (\varphi ^{}_\mu \varphi _\mu \varphi ^{}\varphi )+e^2\mu ^ϵA^\mu A_\mu (\varphi ^{}\varphi )`$ (10) $`{\displaystyle \frac{\lambda \mu ^{2ϵ}}{6^2}}(\varphi ^{}\varphi )^3+A_\mu \varphi ^{}^\mu \varphi +{\displaystyle \frac{B}{2}}\epsilon _{\mu \nu \rho }A^\mu ^\nu A^\rho ie\mu ^{\frac{ϵ}{2}}DA^\mu (\varphi ^{}_\mu \varphi _\mu \varphi ^{}\varphi )`$ $`+e^2\mu ^ϵEA^\mu A_\mu (\varphi ^{}\varphi ){\displaystyle \frac{\mu ^{2ϵ}C}{6^2}}(\varphi ^{}\varphi )^3.`$ The Feynman rules for this Lagrangian in the Landau gauge are depicted in figure 1. This gauge can be implemented by adding to the Lagrangian a gauge fixing term: $`\frac{1}{2\xi }(^\mu A_\mu )^2`$, inverting the free quadratic part of the $`A^\mu `$ to get the CS propagator and, then letting $`\xi \mathrm{}`$. The would be Faddeev-Popov ghost field is completely decoupled of the other fields and does not have any effect. Calling $`\mathrm{\Gamma }(p)`$ the scalar field 1PI two point function, and $`\mathrm{\Gamma }^\mu (q;p,p^{})`$ and $`\mathrm{\Gamma }^{\mu \nu }(q,k;p,p^{})`$, respectively the trilinear and quadrilinear CS scalar field vertices, where $`q`$ and $`k`$ represent “photon” momenta and $`p`$ and $`p^{}`$ scalar field momenta, we have the WI $$q^\mu \mathrm{\Gamma }_\mu (q;p,p^{})=e[\mathrm{\Gamma }(p^{})\mathrm{\Gamma }(p)]$$ (11) $$q^\mu \mathrm{\Gamma }_{\mu \nu }(q,k;p,p^{})=e[\mathrm{\Gamma }_\nu (k;p+q,p^{})\mathrm{\Gamma }_\mu (k;p,p^{}q)],$$ (12) which require that $`E=D=A`$, leaving us with only three (we choose $`A`$, $`B`$, $`C`$) counterterms to be fixed. An explicit proof of these WI in two loops is given in the Appendix A. To determine $`A`$, $`B`$, and $`C`$ we need to calculate the simple pole part of the 2 point function of the CS field, $`\mathrm{\Pi }_{\mu \nu }(q)`$), and the scalar field 2 and 6 point functions, respectively $`\mathrm{\Gamma }_2`$ and $`\mathrm{\Gamma }_6`$. In graphs involving the CS field, we will extend the calculations up to two loops getting at most a simple pole in $`1/ϵ`$; in graphs only involving the scalar field we will go up to four loops. So, in the tensorial Feynman integrals, in which dimensional reduction could possibly differ from dimensional regularization ( in the sub leading terms in $`1/ϵ`$ ) no difference between the two methods are expected in the calculation of the counterterms and in the renormalization group parameters. Let us start with $`\mathrm{\Pi }_{\mu \nu }`$. The only divergent diagrams, up to 2 loops, are those shown in figure 2 ( the possible counterterm is also drawn in the figure ). Their contributions are given by $$(2a)=4e^4𝒟q𝒟k\frac{\epsilon _{\mu \nu \rho }k^\rho }{k^2(qp)^2(q+k)^2}$$ (13) $$(2b)=e^4𝒟q𝒟k\frac{(2k+q)^\alpha \epsilon _{\alpha \beta \gamma }q^\gamma (2k+q2p)^\beta (2kp)_\mu (2k+2qp)_\nu }{k^2(k+q)^2q^2(kp)^2(q+kp)^2},$$ (14) where $`𝒟q\mu ^ϵd^{3ϵ}q/(2\pi )^{3ϵ}`$ and an infinitesimal imaginary part is supposed in every propagator denominator ($`p^2p^2+i\eta `$, $`\eta 1`$). Both integrals are logarithmically divergent. The divergent parts are of the form $`\epsilon _{\mu \nu \rho }p^\rho I`$ where $`I`$ is a scalar integral, that can be calculated by the usual dimensional continuation, after reducing the denominator to a single monomial through the use of Feynman parameters. The results are $$(2a)=4\epsilon _{\mu \nu \rho }p^\rho \left(\frac{e^4}{96\pi ^2}\frac{1}{ϵ}\right)+\text{finite part}$$ (15) $$(2b)=\epsilon _{\mu \nu \rho }p^\rho \left(\frac{e^4}{24\pi ^2}\frac{1}{ϵ}\right)+\text{finite part}.$$ (16) As can be seen, the divergent parts of the two integrals cancel each other and we are left with only finite contributions to $`\mathrm{\Pi }_{\mu \nu }`$. So, the counterterm $`B`$ can be chosen as $`B=0`$, and no infinite wave function renormalization of the CS field interacting with massless scalar field is needed. This result extends for massless matter, the result of the Coleman-Hill theorem . Let us now look at the scalar two point function, $`\mathrm{\Gamma }_2(p)`$. The divergent graphs up to second order in $`\alpha `$ and $`\lambda `$ are shown in figure 3, together with the counterterm. Their contributions are given by $$(3a)=2e^4i𝒟q\frac{1}{(p+q)^2}𝒟k\epsilon _{\mu \nu \rho }\frac{k^\rho }{k^2}\epsilon ^{\nu \mu \gamma }\frac{(k+q)_\gamma }{(k+q)^2}$$ (17) $$(3b)=e^4i^3𝒟q\epsilon _{\alpha \beta \gamma }\frac{q^\gamma }{q^2}\frac{(2p+q)^\beta }{(p+q)^2}𝒟k\frac{(2p+k)^\mu (2p+2q+k)^\nu (2k+2p+q)^\alpha }{(p+k)^2(p+q+k)^2}\epsilon _{\mu \nu \rho }\frac{k^\rho }{k^2}$$ (18) $$(3c)=e^4i^3𝒟q\frac{1}{(p+q)^2}\epsilon _{\mu \alpha \lambda }\frac{q^\lambda }{q^2}\epsilon _{\beta \nu \rho }\frac{q^\rho }{q^2}𝒟k\frac{(2k+q)^\alpha (2k+q)^\beta (2p+q)^\mu (2p+q)^\nu }{k^2(k+q)^2}$$ (19) $$(3d)=\frac{\lambda ^2}{2^2.3}i^5𝒟k_1𝒟k_2𝒟k_3𝒟k_4\frac{1}{k_1^2k_2^2k_3^2k_4^2(p+k_1+k_2k_3k_4)^2}.$$ (20) After the simplification of the tensor algebra in $`(2+1)`$ dimensions we are left with multiple scalar integrals that can be made, one loop at time, through the reduction of the denominators by successive use of Feynman parameters. The results are $$(3a)=2ie^4\left(\frac{p^2}{96\pi ^2}\frac{1}{ϵ}+\mathrm{}\right)$$ (21) $$(3b)=ie^4\left(\frac{p^2}{12\pi ^2}\frac{1}{ϵ}+\mathrm{}\right)$$ (22) $$(3c)=ie^4\left(\frac{p^2}{24\pi ^2}\frac{1}{ϵ}+\mathrm{}\right)$$ (23) $$(3d)=i\frac{\lambda ^2}{2^2.3}\left(\frac{p^2}{3.2^{11}\pi ^4}\frac{1}{ϵ}+\mathrm{}\right).$$ (24) For the contribution $`(iAp^2)`$ of the counterterm to cancel these divergences we must choose: $$A=\left(\frac{7}{48\pi ^2}\alpha ^2\frac{1}{3^2.2^{13}\pi ^4}\lambda ^2\right)\frac{1}{ϵ}.$$ (25) Let us now proceed to the calculation of $`C`$, the counterterm of the coupling constant $`\lambda `$. For this task we need to get the divergent parts of $`\mathrm{\Gamma }_6(p_1,\mathrm{},p_6)`$. After a lengthly analyses of the many graphs involved, we are left with the divergent contributions drawn in figure 4. The bullets on the diagrams $`4p,4q,4r`$ $`4s`$ and $`4t`$ mean the insertion of the counterterm in the corresponding vertex. The calculation of all these diagrams can be reduced to the calculation of the nine integrals represented in figure 5. In the Appendix B we show, as examples, the calculation (of the divergent parts) of $`5a,5b,5d`$ and $`5f`$. Here we present only the results: $$𝒢(p,q)=\frac{1}{2^5\pi ^2}\frac{1}{ϵ}+\text{finite part},$$ (26) $$(p)=\frac{i}{16\pi ^2}\frac{1}{ϵ}+\text{finite part},$$ (27) $$\mathrm{\Delta }_3(p)=\frac{i}{2^5\pi ^2}[\frac{1}{ϵ}++(\mathrm{log}\frac{4\pi \mu ^2}{p^2}32\gamma 2\mathrm{log}2)+𝒪(ϵ)],$$ (28) $$𝒴(p)=\frac{1}{2^{12}\pi ^4}\frac{1}{ϵ}+\text{finite part},$$ (29) $$𝒵(p,q)=\frac{1}{2^{11}\pi ^4}\frac{1}{ϵ}+\text{finite part},$$ (30) $$𝒲(q,p)=\frac{1}{2^{11}\pi ^4}\left[\frac{1}{ϵ^2}+\frac{1}{ϵ}\left(2\mathrm{log}\frac{4\pi \mu ^2}{(p+q)^2}+8\frac{11}{2}\gamma \right)+\text{finite part}\right],$$ (31) $$(a,c,d)=\frac{3i}{2^6\pi ^2}\frac{1}{ϵ}+\text{finite part},$$ (32) $$𝒩(a,c,d)=\frac{i}{2^5\pi ^2}\frac{1}{ϵ}+\text{finite part},$$ (33) and $$𝒬(a,b,c)=\frac{1}{2^5\pi ^2}\frac{1}{ϵ}+\text{finite part}$$ (34) where $`\gamma `$ is the Euler constant. In some graphs we will need the result of $`\mathrm{\Delta }_{3}^{}{}_{}{}^{2}(p)`$: $$\mathrm{\Delta }_{3}^{}{}_{}{}^{2}(p)=\frac{1}{2^{10}\pi ^4}\left[\frac{1}{ϵ^2}+\frac{1}{ϵ}\left(2\mathrm{log}\frac{4\pi \mu ^2}{p^2}2(3+2\gamma +2\mathrm{log}2)\right)+\text{finite part}\right].$$ (35) By collecting all contributions of figure 4 we can write $`\mathrm{\Gamma }_6(p_1,`$ $`p_2`$ $`,p_3,p_1^{},p_2^{},p_3^{})\mu ^{2ϵ}`$ (36) $`=`$ $`{\displaystyle \frac{\lambda ^2}{6}}\mathrm{\Delta }_3(p_1+p_2+p_3){\displaystyle \frac{\lambda ^2}{2}}[\mathrm{\Delta }_3(p_1+p_2p_2^{})+8\text{terms}]`$ (54) $`+2i\lambda \alpha ^2[𝒢(p_1,p_1^{})+8\text{terms}]+2i\lambda \alpha ^2[𝒢(p_1,p_2)+2\text{terms}]`$ $`+2i\lambda \alpha ^2[𝒢(p_1^{},p_2^{})+2\text{terms}]2\lambda \alpha ^2[(p_1p_1^{})+8\text{terms}]`$ $`+i{\displaystyle \frac{5}{4}}\lambda ^3[𝒴(p_1p_1^{},p_2p_2^{})+5\text{terms}]`$ $`+i{\displaystyle \frac{3}{4}}\lambda ^3[𝒴(p_1+p_2,(p_1+p_2^{}))+8\text{terms}]`$ $`+i{\displaystyle \frac{1}{4}}\lambda ^3[𝒵(p_1,p_2)+2\text{terms}]{\displaystyle \frac{5}{12}}\lambda ^2[𝒵(p_1,p_1^{})+8\text{terms}]`$ $`+i{\displaystyle \frac{1}{4}}\lambda ^3[𝒵(p_1^{},p_2^{})+2\text{terms}]`$ $`+i{\displaystyle \frac{1}{36}}\lambda ^3[\mathrm{\Delta }_{3}^{}{}_{}{}^{2}(p_1+p_2+p_3)]+i{\displaystyle \frac{1}{4}}\lambda ^3[\mathrm{\Delta }_{3}^{}{}_{}{}^{2}(p_1+p_2p_1^{})+8\text{terms}]`$ $`+i{\displaystyle \frac{1}{4}}\lambda ^3[𝒲(p_1,p_2+p_3)+2\text{terms}]`$ $`+i{\displaystyle \frac{1}{4}}\lambda ^3[𝒲(p_1^{},(p_2^{}+p_3^{}))+2\text{terms}]`$ $`+i{\displaystyle \frac{3}{4}}\lambda ^3[𝒲(p_1,p_2p_1^{})+17\text{terms}]`$ $`+i{\displaystyle \frac{3}{4}}\lambda ^3[𝒲(p_1^{},p_1p_2^{})+17\text{terms}]`$ $`+i{\displaystyle \frac{7}{12}}\lambda ^3[𝒲(p_1,(p_1^{}+p_2^{}))+8\text{terms}]`$ $`+i{\displaystyle \frac{7}{12}}\lambda ^3[𝒲(p_1^{},p_1+p_2)+8\text{terms}]{\displaystyle \frac{\lambda C}{3}}\mathrm{\Delta }_3(p_1+p_2+p_3)`$ $`\lambda C[\mathrm{\Delta }_3(p_1+p_2p_1^{})+8\text{terms}]iC`$ $`+\mathrm{\hspace{0.17em}2}^4\alpha ^4[(p_1,p_2p_2^{},p_3p_3^{})+17\text{terms}]`$ $`+\mathrm{\hspace{0.17em}2}^5\alpha ^4[𝒩(p_1,p_2p_2^{},p_3p_3^{})+17\text{terms}]`$ $`+i2^2\alpha ^4[𝒬(p_1p_1^{},p_2p_2^{},p_3p_3^{})+35\text{terms}]`$ from which, after imposing that the result be finite, we get $`C=`$ $`\lambda ^2{\displaystyle \frac{7}{48\pi ^2}}{\displaystyle \frac{1}{ϵ}}\lambda \alpha ^2\left[{\displaystyle \frac{33}{16\pi ^2}}\right]{\displaystyle \frac{1}{ϵ}}+\alpha ^4{\displaystyle \frac{72}{2\pi ^2}}{\displaystyle \frac{1}{ϵ}}.`$ (56) $`\lambda ^3\left[{\displaystyle \frac{582+57\pi ^21092\gamma }{2^{14}\pi ^4}}\right]{\displaystyle \frac{1}{ϵ}}+\lambda ^3\left[{\displaystyle \frac{49}{2^83^2\pi ^4}}\right]{\displaystyle \frac{1}{ϵ^2}},`$ The term proportional to $`\alpha ^4`$ in the above expression shows that the renormalization of $`\lambda `$ is not multiplicative, a fact that will lead to an interesting effect in the renormalization group equations. In the next section, results (25), (56) and $`B=0`$ will be used to determine the renormalization group parameters. ## III Renormalization Group Analyses Let us start by verifying the that the CS coupling does not run. Equation (6) is $$\alpha _0=\alpha \mu ^ϵ\frac{(1+E)}{(1+A)(1+B)}.$$ (57) As we have seen in the last section, $`B=0`$ and, as consequence of the Ward Identities, we also have $`E=A`$. Thus (57) reduces to $$\alpha _0=\alpha \mu ^ϵ,$$ (58) from which, in the way of we get $$0\mu ^{1ϵ}\frac{d\alpha _0}{d\mu }=ϵ\alpha +\mu \frac{d\alpha }{d\mu },$$ (59) and therefore $$\beta _\alpha =\mu \frac{d\alpha }{d\mu }|_{ϵ0}0,$$ (60) showing that $`\alpha `$ does not run under a rescaling of $`\mu `$ or the momenta of the Green function. A similar result was get in for a model of a scalar field interacting with a non Abelian CS field. These results extend to massless matter, the result of the theorem of Coleman-Hill . For calculating $`\beta _\lambda `$ we start with equation (7): $$\lambda _0=\mu ^{2ϵ}\frac{\lambda +C}{(1+A)^3}=\mu ^{2ϵ}(\lambda +C3A+\mathrm{})$$ (61) By substituting (25) and (56) in this equation we get $$\lambda _0=\mu ^{2ϵ}\left(\lambda +\frac{\lambda _1(\alpha ,\lambda )}{ϵ}+\mathrm{}\right),$$ (62) where $$\lambda _1(\alpha ,\lambda )=a(\lambda ^2c\alpha ^2\lambda +d\alpha ^4b\lambda ^3)$$ (63) with $$a=\frac{7}{48\pi ^2}=0.01478,$$ (64) $$b=\frac{1}{2^{10}7\pi ^2}(1744+171\pi ^23276\gamma )=0.0218,$$ (65) $$c=\frac{120}{7}=17.1429,$$ (66) and $$d=\frac{1728}{7}=246.86.$$ (67) From (62) we have $`0`$ $`=`$ $`\mu ^{12ϵ}{\displaystyle \frac{d\lambda _0}{d\mu }}`$ (68) $`=`$ $`2ϵ\left(\lambda +{\displaystyle \frac{\lambda _1}{ϵ}}+\mathrm{}\right)+\left(\mu {\displaystyle \frac{\lambda }{\mu }}+\mu {\displaystyle \frac{\lambda }{\mu }}{\displaystyle \frac{\lambda _1}{\lambda }}{\displaystyle \frac{1}{ϵ}}+\mu {\displaystyle \frac{\alpha }{\mu }}{\displaystyle \frac{\lambda _1}{\alpha }}{\displaystyle \frac{1}{ϵ}}+\mathrm{}\right),`$ (69) and using (59) we get $`\beta _\lambda `$ $`=`$ $`\mu {\displaystyle \frac{\lambda }{\mu }}`$ (70) $`=`$ $`\left(\alpha {\displaystyle \frac{}{\alpha }}+2\lambda {\displaystyle \frac{}{\lambda }}2\right)\lambda _1(\alpha ,\lambda )2\lambda ϵ`$ (71) $`=`$ $`2a(\lambda ^2c\lambda \alpha ^2+d\alpha ^42b\lambda ^3)(\text{for}ϵ0)`$ (72) Up to 2 loops ( terms in $`\lambda ^2`$, $`\lambda \alpha ^2`$ and $`\alpha ^4`$ ) this result qualitatively coincides with that of for this same model. It does not, however, coincide with the result of ( we will discuss this fact in the conclusions ). As can be seen from (66), the contribution of the 4 loops graphs ( term in $`\lambda ^3`$ ) is small and will not qualitatively change the results for $`\beta _\lambda `$. Making $`\alpha =0`$ we go to the pure $`(\varphi ^{}\varphi )^3`$ model. In this case $`\beta `$ starts at zero for $`\lambda =0`$ and increases monotonically with $`\lambda `$ . The model presents an infrared (IR) fix point at $`\lambda =0`$. For $`\alpha 0`$ a drastic change occurs. In this case $`\beta `$ starts at $`(4ad\alpha ^4)>0`$ for $`\lambda =0`$ and never vanishes in the perturbative range of the two coupling constants. A similar behavior of the $`\beta `$ function, already in one loop order, is shown in the Coleman-Weinberg model (CW) in (3+1) dimensions. There, a dynamical symmetry breakdown occurs and masses are generated for the two fields. In the effective potential was calculated in two loops and a breakdown of symmetry was also shown to appear. We would like to stress that our results for $`\mathrm{\Gamma }^2`$ and $`\mathrm{\Gamma }^6`$ are compatibles with that conclusion. The $`\mathrm{\Gamma }^2(v)`$ for the displaced field $`\psi =\varphi v`$, were $`v`$ is a constant with dimension $`(mass)^{1/2}`$, would be written, in terms of the functions that we calculated for $`\varphi `$, as a series of the form $`\mathrm{\Gamma }^2(v)=\mathrm{\Gamma }^2+(v^2/2)\mathrm{\Gamma }^4+(v^4/4!)\mathrm{\Gamma }^6+\mathrm{}`$ As can be seen from the graphs proportional to $`\alpha ^4`$ in figure 5, $`\mathrm{\Gamma }^6`$ receives a constant ( independent of p ) finite contribution. As consequence, $`\mathrm{\Gamma }^2(v)`$ will have a singularity displaced to some non null value of $`p^2`$, compatible with a non null dynamically generated mass for $`\varphi `$. The anomalous dimensions of the fields $`A_\mu `$ and $`\varphi `$ are given by $$\gamma _A=\frac{1}{2}\frac{\mu }{Z_A}\frac{dZ_A}{d\mu }$$ (73) $$\gamma _\varphi =\frac{1}{2}\frac{\mu }{Z_\varphi }\frac{dZ_\varphi }{d\mu }.$$ (74) As shown in section II, $`Z_A=1+B=1`$ and so $`\gamma _A=0`$. By writing $$Z_\varphi =1+A=1+\frac{a_1(\alpha ,\lambda )}{ϵ}+\mathrm{}$$ (75) where $`a_1`$ is given in (25) we can write (74) in the form $$2\left(1+\frac{a_1}{ϵ}+\mathrm{}\right)\gamma _\varphi =\mu \frac{\lambda }{\mu }\frac{a_1}{\lambda }\frac{1}{ϵ}+\mu \frac{\alpha }{\mu }\frac{a_1}{\alpha }\frac{1}{ϵ}+\mathrm{},$$ (76) and using (59) and (72) we get $$\gamma _\varphi =\lambda \frac{a_1}{\lambda }\frac{\alpha }{2}\frac{a_1}{\alpha }.$$ (77) By substituting $`a_1`$, from (25), in (77) we have $$\gamma _\varphi =\frac{7}{48\pi ^2}\alpha ^2+\frac{1}{3^2.2^{12}\pi ^4}\lambda ^2.$$ (78) The contribution in $`\alpha ^2`$ qualitatively agrees with the result of . The term in $`\lambda ^2`$ comes from 4 loops graphs ( not calculated in ) and is very small compared to the term in $`\alpha ^2`$. It can be seen from (77), that the scalar field dimension, $`D_\varphi =\frac{1}{2}+\gamma _\varphi `$, decreases with the coupling to the CS field. As it is well known, in non perturbative approach in quantum mechanics, the coupling of matter fields to a CS field, changes the spin and statistics of the matter fields, driving bosons into anyons and also, for strong enough coupling, into fermions. Based on these results, there is a conjecture in the literature that, even in perturbative quantum field approach (in which the strength $`\alpha 1`$) the dimension of a boson coupled to a CS should receive an increase in the direction of the fermion dimension $`d_\psi =1`$ ( for the corresponding problem of fermions a decrease in the direction of the boson dimension should be expected ). As shown in (78) this conjecture is not realized: the coupling to the CS field works in the direction of decreasing the dimension of $`\varphi `$. To get a bit farther in testing this conjecture, we have also looked at the anomalous dimensions of the composite operators $`[(\varphi ^{}\varphi )^n]`$, where n is an integer number. As we are mainly interested in the effect of the coupling of the boson to the CS field, to simplify the calculations, we will restrict the analyze to $`\lambda =0`$. In terms of monomials of $`\varphi `$ this composite operator can be written $$[(\varphi ^{}\varphi )^n]=Z_n(\varphi ^{}\varphi )^n+Z_{n1}^0(\varphi ^{}\varphi )^{n1}+Z_{n2}^2(\varphi ^{}\varphi )^{n2}(\varphi ^{}^2\varphi )+\mathrm{}.$$ (79) Determination of the $`Z_m^i`$ ($`mn`$) require the calculation of the divergent parts of the 2m scalar field 1PI vertex functions with the insertion of one integrated composite operator $$\mathrm{\Gamma }_{[(\varphi ^{}\varphi )^n]}(x_1,\mathrm{},y_m)=d^3z<T[(\varphi ^{}\varphi )^n](z)\varphi (x_1)\mathrm{}\varphi (x_m)\varphi ^{}(y_1)\mathrm{}\varphi ^{}(y_m)>,$$ (80) or, in momentum space, the $`\mathrm{\Gamma }_{[(\varphi ^{}\varphi )^n]}(p_1,\mathrm{},p_{2m})`$ function with zero momentum $`q`$ entering through the special vertex $`[(\varphi ^{}\varphi )^n]`$. Up to order $`\alpha ^2`$, the divergent graphs contributing to $`\mathrm{\Gamma }_{[(\varphi ^{}\varphi )^n]}(p_1,\mathrm{},p_{2n})`$ are shown in figure 6. In figure 7 we draw some of the graphs that could contribute to $`\mathrm{\Gamma }_{[(\varphi ^{}\varphi )^n]}(p_1,\mathrm{}p_{2(n1)})`$. Diagrams in figure 7, are in fact all nulls, what imply that the renormalization parameters $`Z_{n1}^i`$ also vanish. The same can be shown to be true for all $`Z_m^i`$ which any $`m<n`$. So, the right side of (80) reduces to only the first monomial and $`[(\varphi ^{}\varphi )^n]`$ does not mix with other operators ( mixing will however appear if we consider $`\lambda 0`$ ). Its renormalization only requires the calculation of $`Z_n`$, what means to calculate the divergent parts of the graphs in figure 6. The involved Feynman integrals are the $`𝒢(p,q)`$ and $`(p)`$ from figure 5. By writing $`Z_n=1+A_n`$ we have $$\mathrm{\Gamma }_{[(\varphi ^{}\varphi )^n]}(p_1,\mathrm{},p_{2n})=(n!)^2[A_n(4n^22n)\alpha ^2𝒢2in^2\alpha ^2]+\text{finite graphs},$$ (81) and we have for $`A_n`$ $`A_n`$ $`=`$ $`\text{DivPart}\{(4n^22n)\alpha ^2𝒢+2in^2\alpha ^2\}`$ (82) $`=`$ $`{\displaystyle \frac{4n^2n}{16\pi ^2}}{\displaystyle \frac{\alpha ^2}{ϵ}},`$ (83) where ”DivPart” stands for keeping only the divergent part of the following expression. With these results for $`Z_m`$ and (25) for $`Z_\varphi `$, the equation (79) rewritten in terms of the unrenormalized (see also (3)) field $`\varphi _0`$, becomes $$[(\varphi ^{}\varphi )^n]=Z_{cn}^1(\varphi _0^{}\varphi _0)^n,$$ (84) where $$Z_{cn}=(Z_n)^1(Z_\varphi )^n=1+\frac{a_{cn}(\alpha )}{ϵ}+\mathrm{},$$ (85) and $$a_{cn}(\alpha )=\frac{\alpha ^2}{4\pi ^2}\left(n^2+\frac{n}{3}\right).$$ (86) By deriving the two sides of (84) with respect to $`\mu `$ and remembering that $`\varphi _0`$ is independent of $`\mu `$ we have $$\mu \frac{d}{d\mu }[(\varphi ^{}\varphi )^n]=\gamma _{cn}[(\varphi ^{}\varphi )^n],$$ (87) where $$\gamma _{cn}=\frac{\mu }{Z_{cn}}\frac{dZ_{cn}}{d\mu },$$ (88) is the anomalous dimension of the composite operator. Going through the same steps that leads (75) to (77) we get $$\gamma _{cn}=\alpha \frac{da_{cn}}{d\alpha }=\frac{\alpha ^2}{2\pi ^2}\left(n^2+\frac{n}{3}\right).$$ (89) The dimension of the composite operator $`[(\varphi ^{}\varphi )^n]`$ becomes $$D_{[(\varphi ^{}\varphi )^n]}=n\frac{\alpha ^2}{2\pi ^2}\left(n^2+\frac{n}{3}\right).$$ (90) This result is in disagreement with . Their calculation seems to miss the contribution of the second graph in our figure 6. But it is not this fact what makes the major difference. Our counting of the combinatorial factors of the graphs in figure 6, gives a term proportional to $`n^2`$ ( besides the term in $`n`$ ), different from theirs which is only proportional to $`n`$. No matter if the composite operator is super-renormalizable ($`n<3`$), renormalizable ($`n=3`$) or non-renormalizable ($`n>3`$), the effect of the coupling to the CS field is to lower its dimension. Nevertheless, the lowest non-renormalizable operator, $`(\varphi ^{}\varphi )^4`$, with effective dimension: $`D_4=4\frac{52}{6\pi ^2}\alpha ^2`$ will never, in the perturbative regime, be driven to be renormalizable. Yet, due to the quadratic dependence of the anomalous dimension on $`n`$, given any $`\alpha 1`$, the operators $`[(\varphi ^{}\varphi )^n]`$ with $`n`$ bigger than $`n_c\frac{2\pi ^2}{\alpha ^2}\frac{10}{3}1`$ have their operator dimensions driven to values smaller than 3. To finish this section, let us look at the renormalization group equations for the $`\mathrm{\Gamma }_{(2n)}(p,\lambda ,\alpha ,\mu )`$ functions ( p is a short for the 2n external momenta ). As the 4 loops contributions are very small we will restrict the analyses to 2 loops. As $`\beta _\alpha `$ and $`\gamma _A`$ are null we have the renormalization group equation $$(\mu \frac{}{\mu }+\beta _\lambda \frac{}{\lambda }2n\gamma _\varphi )\mathrm{\Gamma }_{(2n)}(p,\lambda ,\alpha ,\mu )=0.$$ (91) The solution of this equation can be written as $$\mathrm{\Gamma }_{(2n)}(p,\lambda ,\alpha ,\mu )=\mathrm{\Gamma }_{(2n)}(p,\overline{\lambda },\alpha ,\mu s_{\overline{\lambda },\lambda })s_{\overline{\lambda },\lambda }^{\mathrm{\hspace{0.17em}\hspace{0.17em}\hspace{0.17em}\hspace{0.17em}2}n\gamma _\varphi }$$ (92) where we used the fact that up to two loops, $`\gamma _\varphi =\frac{7}{48\pi ^2}\alpha ^2`$ does not change with s. In the above equation, $`s_{\overline{\lambda }\lambda }`$ stands for the solution of $`s{\displaystyle \frac{d}{ds}}\overline{\lambda }`$ $`=`$ $`\beta _\lambda (\overline{\lambda })`$ (93) $`=`$ $`2a(\overline{\lambda }^2c\alpha ^2\overline{\lambda }+d\alpha ^4),`$ (94) with the condition $`\overline{\lambda }(s=1)=\lambda `$, that is : $`s_{\overline{\lambda }\lambda }`$ $`=`$ $`exp\left({\displaystyle \frac{1}{af\alpha ^2}}\left[tan^1\left({\displaystyle \frac{2\overline{\lambda }}{f\alpha ^2}}{\displaystyle \frac{c}{f}}\right)tan^1\left({\displaystyle \frac{2\lambda }{f\alpha ^2}}{\displaystyle \frac{c}{f}}\right)\right]\right)`$ (95) $``$ $`exp\left({\displaystyle \frac{2.86}{\alpha ^2}}\left[tan^1\left({\displaystyle \frac{\overline{\lambda }}{12\alpha ^2}}0.71\right)tan^1\left({\displaystyle \frac{\lambda }{12\alpha ^2}}0.71\right)\right]\right),`$ (96) where $`f=(4dc^2)^{1/2}`$. As $`\beta _\lambda `$ is non null for $`\lambda =0`$ ( for $`\alpha 0`$ ) this equation is well defined if we choose $`\lambda =0`$. With this choice in (96) we can write $$\mathrm{\Gamma }_{(2n)}(p,\overline{\lambda },\alpha ,\mu )=\mathrm{\Gamma }_{(2n)}(p,0,\alpha ,\mu s_{\overline{\lambda }\mathrm{\hspace{0.17em}0}}^1)s_{\overline{\lambda }\mathrm{\hspace{0.17em}0}}^{2n\gamma _\varphi }.$$ (97) This equation shows that, up to two loops, the $`\mathrm{\Gamma }_{(2n)}`$ functions of the model defined by Lagrangian (10), can be get from the corresponding $`\mathrm{\Gamma }_{(2n)}`$ for the model where only the interaction term with the $`A_\mu `$ field is present, or what is equivalent, from the calculation of the sub set of diagrams contributing to $`\mathrm{\Gamma }_{(2n)}`$, which only involves the interaction vertex with the $`A_\mu `$ field. A short inspection of the CW results, shows that a similar fact is also true for that model ( at least in one loop ). ## IV Conclusions The coupling to the CS field lowers the dimension of $`\varphi `$ and of $`(\varphi ^{}\varphi )^n`$ . This goes in the opposite direction of the conjecture that the transmutation of the boson into anyon ( due to the coupling to the CS field ) should be signaled by the dimension of these operators to increase in the direction of the canonical dimension of a fermion field $`\psi `$ and their composite operators $`(\psi ^{}\psi )^n`$, respectively. In the present paper, as in previous calculations in the literature, the function $`\beta _\alpha `$ and the anomalous dimension of the CS field are shown to vanish; the CS coupling constant $`\alpha `$ does not run with the change of the energy scale. The function $`\beta _\lambda `$ instead, shows a drastic change in the presence of the CS field. From an IR trivial fix point for the pure $`\lambda (\varphi ^{}\varphi )^3`$ interaction, the model is driven, to a phase in which no fix point appears for $`\beta _\lambda `$, in a behavior similar to that of $`\beta _\lambda `$ for the model of Coleman-Weinberg . In , the renormalization group functions were calculated up to 2 loops, although their main aim was to study the effective potential and dynamical symmetry breakdown. The model of , defined by their Lagrangian (2.1) can be made to coincide with ours by deleting their $`\lambda (\varphi ^{}\varphi )^2`$ interaction and the $`m^2(\varphi ^{}\varphi )`$ mass term, that is, by making their $`\lambda `$ and $`m`$ zero. Considering also, that their coefficient, $`\nu `$, of the $`(\varphi ^{}\varphi )^3`$ interaction, differs from our $`\lambda `$ by a factor of 2/5, what also implies in a 2/5 factor of difference in the corresponding $`\beta `$ functions, their results ( equations (10.7-9) and (11.8) ), after translated to our notation, can be summarized as: $`\mathrm{𝟏}.`$ $`\beta _\alpha =0`$ and $`\gamma _A=0`$. These results are in agreement with our equation (60) and the observation below equation (74). $`\mathrm{𝟐}.`$ $`\gamma _\varphi =𝒪(\lambda ^2)`$, and $`\beta _\lambda =2a\lambda ^2+𝒪(\lambda ^3)`$, both independent of $`\alpha `$. Our results (72) and (78) differ from these last ones by terms dependents on the CS coupling $`\alpha `$. Their conclusion is that the model has an IR trivial fix point in $`\lambda `$. Ours instead, is that $`\beta _\lambda `$ never vanishes, a result similar to that of CW in a model in which a dynamical symmetry breakdown occurs. A dynamical symmetry breakdown was also seen in for the present model. Our result for $`\beta `$ looks so, in accordance with their result on symmetry breakdown. The discrepancies between ours and the $`\beta `$ function of can be attributed to the different regularization schemes we are using. In , the model is regularized through a full dimensional regularization, by extending out of 3D, all the tensor structures ( including the definition of the $`ϵ_{\mu \nu \rho }`$ ) that appear in the Feynman graphs. As they conclude, in that method, the renormalizability of the model is only achieved, if an extra regularization, represented by a Maxwell term for the $`A^\mu `$ field ( besides the CS one ), is introduced. Their method requires, that this extra regularization be dismissed ( their parameter “a” taken to zero ), only after the continuation back to 3D is made. As can be seen from their results (11.8), some of their $`\beta `$ functions become singular, when $`a0`$, showing that a better understanding of the structure of the renormalization group equation is still lacking in that method. Also, as discussed in their Section 10, if a regularization directly in 3D ( exists and ) were used, $`\gamma _\varphi `$ and $`\beta _\lambda `$ would be expected to depend also on $`\alpha `$. In this paper we used the Dimensional Reduction regularization scheme, in which all the tensor contractions are first made in 3D and only the remaining scalar Feynman integrals are extended out of 3D. We explicitly verified that this method controls all the UV infinities and preserves the Ward identities ( and so, the gauge covariance ) up to the order of approximation in which we are working ( 2 loops in graphs involving the CS propagator and 4 loops in graphs only involving the scalar propagator ). Although we can not say that it is a regularization directly in 3D, our results is consistent with the above mentioned discussion in . As a definitive answer to this problem is desirable, we are presently working in a related model, using a direct in 3D version of the BPHZ renormalization method. The preliminary results confirm those of the present paper for the renormalization group functions, together with the dynamical symmetry breakdown got in . To finalize we would like to summarize the results of two previous papers , in which we studied the scale behavior of fermions interacting with a CS field. In the first one, a single fermion with its most general 4-fermion ( non renormalizable ) self interaction $`g(\overline{\psi }\psi )^2`$ was considered. We saw that, although $`\psi `$ gets a negative anomalous dimension, making its operator dimension to approach that of a boson, no definite pattern of approach to a bosonic scale behavior due to the interaction with the CS field is seen for composite operators: the super-renormalizable operator $`\overline{\psi }\psi `$ gets a negative anomalous dimension, but the non-renormalizable operator $`(\overline{\psi }\psi )^2`$ gets a positive one. In the second paper an extended version of this model with N ( small ) fermion fields, with their most general 4-fermion interaction: $`g(\overline{\psi }\psi )^2+h(\overline{\psi }\gamma ^\mu \psi )^2`$ was considered. We studied operators of canonical dimension four. We showed that one of them has positive anomalous dimension, other has a very small negative anomalous dimension and the third one, more interesting from the renormalization view point, has a negative anomalous dimension, making, through a fine tuning of the coupling constants, its operator dimension as close to 3 as wanted . Nevertheless, no general pattern of approach to a bosonic like behavior ( negative anomalous dimension ), as advanced by the conjecture in the literature, was seen. ## A The Ward Identities The two relations among the counterterms: A to E can be get from the WI among the 1PI 4-linear photon-scalar vertex, $`\mathrm{\Gamma }_{\mu \nu }`$, the trilinear photon-scalar vertex, $`\mathrm{\Gamma }_\mu `$, and the scalar self energy $`\mathrm{\Gamma }_2`$. In tree approximation they are given by ( see figure 1 ): $`\mathrm{\Gamma }_{\mu \nu }=2ie^2\mu ^ϵg_{\mu \nu }`$, $`\mathrm{\Gamma }_\mu =ie\mu ^{ϵ/2}(p+p)_\mu `$ and $`\mathrm{\Gamma }_2=iAp^2`$. It is ease to use that they satisfy the relations $$q^\mu \mathrm{\Gamma }_\mu (q;p,p^{})=e\mu ^{\frac{ϵ}{2}}\left\{\mathrm{\Gamma }_2(p^{})\mathrm{\Gamma }_2(p)\right\},$$ (A1) $$q^\mu \mathrm{\Gamma }_{\mu \nu }(q,q^{};p,p^{})=e\mu ^{\frac{ϵ}{2}}\left\{\mathrm{\Gamma }_\nu (q^{};p^{}q^{},p^{})\mathrm{\Gamma }_\nu (q^{};p,p+q^{})\right\}.$$ (A2) As we explicitly verified these relations are, in fact, valid up to 2-loop order. Instead of considering the WI among the sum of all graphs up to 2-loops contributing to each of the 3 vertex functions above, we can take advantage of the fact that they can be separated in sub classes to be seen to be separately related through the the WI (A1) and (A2). As an example consider the graphs (8-a) to (8-h) contributing to $`\mathrm{\Gamma }_\mu `$ and (9-a) and (9-b) contributing to $`\mathrm{\Gamma }_2`$. Let us call $`\stackrel{~}{\mathrm{\Gamma }}_\mu `$ the sum of contributions of diagrams (8-a) to (8-q) and $`\stackrel{~}{\mathrm{\Gamma }_2}`$ the graph (9-a). Let also $`\stackrel{~}{D}`$ and $`\stackrel{~}{A}`$ be the possible divergent contributions to the counterterms $`D`$ and $`A`$, chosen so as to make finite the sums of graphs in figure 8 and 9, respectively. By using dimensional reduction regularization, and explicitly writing all the Feynman integrands involved, we can verify that $$q^\mu \left\{\stackrel{~}{\mathrm{\Gamma }}_\mu (q;p,p^{})ie\mu ^{\frac{ϵ}{2}}(p^{}+p)_\mu \stackrel{~}{D}\right\}=e\mu ^{\frac{ϵ}{2}}\left\{(\stackrel{~}{\mathrm{\Gamma }_2}(p^{})+ip^2\stackrel{~}{D})(\stackrel{~}{\mathrm{\Gamma }_2}(p)+ip^2\stackrel{~}{D})\right\}.$$ (A3) As $`\stackrel{~}{D}`$ is chosen so as to make the bracket in the left side of these equations finite, the right side is also finite, what implies that: $`ip^2\stackrel{~}{D}=\text{DivPart}\{\stackrel{~}{\mathrm{\Gamma }_2}(p)\}ip^2\stackrel{~}{A}`$, that is $`\stackrel{~}{D}=\stackrel{~}{A}`$. A more direct verification is obtained by explicitly calculating: $$ie\mu ^{\frac{ϵ}{2}}(p^{}+p)_\mu \stackrel{~}{D}=\text{DivPar}\{\stackrel{~}{\mathrm{\Gamma }}_\mu (q;q,p^{})\}$$ (A4) and $$ip^2\stackrel{~}{A}=\text{DivPart}\{\stackrel{~}{\mathrm{\Gamma }_2}(p)\}.$$ (A5) The only really divergent graphs contributing to $`\stackrel{~}{\mathrm{\Gamma }}_\mu `$ are (8-a) and (8-g). By going through the calculation of the divergent parts of (8-a) plus (8-g) as exemplified in Appendix B we get $`ie(p^{}+p)_\mu \stackrel{~}{D}`$ $`=`$ $`\text{DivPart}\{(ie)^3(ie^2)(i)^3{\displaystyle \frac{12}{3!}}{\displaystyle }𝒟k{\displaystyle }𝒟k^{}\epsilon _{\beta \nu \rho }{\displaystyle \frac{k^\rho }{k^2}}\epsilon _{\alpha \mu \gamma }{\displaystyle \frac{k^\gamma }{k^2}}`$ (A7) $`\times {\displaystyle \frac{(2p+k)_\beta (2p+2k+k^{})_\alpha (2p+2k^{}+k)_\nu }{(p+k)^2(p+k^{})^2(p+k+k^{})^2}}+pp^{}\}`$ $`=`$ $`i{\displaystyle \frac{e^5}{12\pi ^2}}{\displaystyle \frac{1}{ϵ}}p_\mu +i{\displaystyle \frac{e^5}{12\pi ^2}}{\displaystyle \frac{1}{ϵ}}p_\mu ^{}`$ (A8) that is: $`\stackrel{~}{D}=\frac{\alpha ^2}{12\pi ^2}\frac{1}{ϵ}`$. For $`\stackrel{~}{A}`$ we have $`ip^2\stackrel{~}{A}`$ $`=`$ $`\text{DivPart}\{\stackrel{~}{\mathrm{\Gamma }}(p)\}`$ (A9) $`=`$ $`\text{DivPart}\left\{\text{Graph (3-b)}\right\}`$ (A10) $`=`$ $`i{\displaystyle \frac{e^4}{12\pi ^2}}{\displaystyle \frac{1}{ϵ}}p^2`$ (A11) as given by (2.8). So, we have $`\stackrel{~}{D}=\stackrel{~}{A}=\frac{\alpha ^2}{12\pi ^2}\frac{1}{ϵ}`$. An example of subset of graphs that match through the $`2^d`$ WI are depicted in Fig. 10 and Fig. 11. The identification of $`\stackrel{~}{E}^{}=\stackrel{~}{D}^{}`$ follows through the same steps as in the example above. ## B Feynman Integrals To illustrate the method adopted to get the divergent parts of the Feynman integrals that appear in the paper, we will explicitly show as examples the calculation of the diagrams 5.a, 5.b, 5.d and 5.f. Let us start with (5.a). In the figure, $`\mathrm{\Delta }(k)=i/(k^2+i\eta )`$ as usual, and $`\mathrm{\Delta }_2(k)`$ and $`\mathrm{\Delta }_3(k)`$ stand for the subgraphs formed respectively by 2 and 3 scalar propagators connecting 2 vertices, with total momentum $`k`$ passing through. Its integration can be done successively one loop at time, first getting $`\mathrm{\Delta }_2`$ and then $`\mathrm{\Delta }_3`$. $`\mathrm{\Delta }_2(p)`$ is given by $$\mathrm{\Delta }_2(p)=𝒟k\frac{i}{k^2+i\eta }\frac{i}{(k+p)^2+i\eta },$$ (B1) where $`𝒟k=\mu ^ϵd^{3ϵ}k/(2\pi )^{3ϵ}`$. By introducing a Feynman parameter through the use of the identity $$\frac{1}{A^\alpha B^\beta }=\frac{\mathrm{\Gamma }(\alpha +\beta )}{\mathrm{\Gamma }(\alpha )\mathrm{\Gamma }(\beta )}_0^1𝑑x\frac{x^{\alpha 1}(1x)^{\beta 1}}{[Ax+B(1x)]^{\alpha +\beta }},$$ (B2) the $`k`$ integration can be done and then the parametric integration to give $$\mathrm{\Delta }_2(p)=i\frac{(4\pi \mu ^2)^{\frac{ϵ}{2}}}{(4\pi )^{\frac{3}{2}}}\frac{\mathrm{\Gamma }^2\left(\frac{1}{2}\frac{ϵ}{2}\right)\mathrm{\Gamma }\left(\frac{1}{2}+\frac{ϵ}{2}\right)}{\mathrm{\Gamma }(1ϵ)}(p^2i\eta )^{\left(\frac{1}{2}\frac{ϵ}{2}\right)}.$$ (B3) $`\mathrm{\Delta }_3(q)`$ can be written as $$\mathrm{\Delta }_3(q)=𝒟p\frac{i}{(p+q)^2+i\eta }\mathrm{\Delta }_2(p).$$ (B4) This integration can also be done following the same steps as for $`\mathrm{\Delta }_2(p)`$, after explicitly substituting in this last equation, the expression (B3) for $`\mathrm{\Delta }_2(p)`$. The result is $$\mathrm{\Delta }_3(q)=\frac{i}{(4\pi )^3}\frac{\mathrm{\Gamma }^3\left(\frac{1}{2}\frac{ϵ}{2}\right)\mathrm{\Gamma }\left(ϵ\right)}{\mathrm{\Gamma }\left(\frac{3}{2}\frac{3ϵ}{2}\right)}\left(\frac{4\pi \mu ^2}{q^2+i\eta }\right)^ϵ.$$ (B5) For 5.d we have $$𝒲(q,p)=𝒟k\frac{i}{(k+q)^2+i\eta }\mathrm{\Delta }_2(k+p)\mathrm{\Delta }_3(k),$$ (B6) This integral can be done by first reducing the 3 denominators to a single one, by twice using (B2) to get a single denominator and then doing the $`k`$ integration. In terms of the 2 remaining Feynman parameters it has the form $$𝒲(q,p)=\frac{1}{(4\pi )^6}\frac{\mathrm{\Gamma }\left(2ϵ\right)\mathrm{\Gamma }^5\left(\frac{1}{2}\frac{ϵ}{2}\right)}{\mathrm{\Gamma }\left(1ϵ\right)\mathrm{\Gamma }\left(\frac{3}{2}\frac{3ϵ}{2}\right)}\left(\frac{4\pi \mu ^2}{p^2+i\eta }\right)^{2ϵ}I_ϵ(q,p),$$ (B7) where $`I_ϵ(q,p)`$ is given by $$I_ϵ(q,p)=_0^1𝑑y(1y)^{ϵ1}f_ϵ(y),$$ (B8) and $$f_ϵ(y)=_0^1𝑑xx^{\frac{1}{2}+\frac{ϵ}{2}}y^{\frac{1}{2}+\frac{ϵ}{2}}\left\{\frac{q^2}{p^2}[y^2(1x)^2y(1x)]+2\frac{p.q}{p^2}y^2x(1x)+yx(yx1)\right\}^{2ϵ}.$$ (B9) $`I_ϵ(q,p)`$ has a single pole in $`ϵ`$ coming from the integration region in the vicinity of $`y=1`$. As (B7) already has a factor $`\mathrm{\Gamma }(2ϵ)`$ the integral (B6) will present both a single and a double pole in $`ϵ`$. To separate their contributions we must calculate the first 2 terms (single pole and the $`ϵ`$ independent term) of the Laurent expansion of $`I_ϵ(q,p)`$. We have $`I_ϵ(q,p)`$ $`=`$ $`I_{1ϵ}(q,p)+I_{2ϵ}(q,p)`$ (B10) $`=`$ $`{\displaystyle \frac{A_1}{ϵ}}+(B_1+B_2)+(C_1+C_2)ϵ+\mathrm{},`$ (B11) where $`I_{1ϵ}(q,p)`$ $`=`$ $`{\displaystyle _0^1}𝑑y(1y)^{ϵ1}f_ϵ(1)`$ (B12) $`=`$ $`{\displaystyle \frac{A_1}{ϵ}}+B_1+C_1ϵ+\mathrm{},`$ (B13) $`I_{2ϵ}(q,p)`$ $`=`$ $`{\displaystyle _0^1}dy(1y)^{ϵ1}(f_ϵ(y)f_ϵ(1)`$ (B14) $`=`$ $`B_2+C_2ϵ+\mathrm{},`$ (B15) where $`A_1`$, $`B_1`$ and $`B_2`$ are still to be determined. $`B_2`$ is given by $$B_2=I_{20}(q,p)=_0^1𝑑y(1y)^1_0^1𝑑x(y^{\frac{1}{2}}1)x^{\frac{1}{2}}=4(1+\mathrm{log}2).$$ (B16) $`A_1`$ and $`B_1`$ come from $`I_{1ϵ}(p,q)`$ $`=`$ $`{\displaystyle _0^1}𝑑y(1y)^{ϵ1}{\displaystyle _0^1}𝑑xx^{\frac{ϵ}{2}\frac{1}{2}}(x^2x)^{2ϵ}\left\{{\displaystyle \frac{(qp)^2}{(p^2)}}\right\}^{2ϵ}`$ (B17) $`=`$ $`(1)^{2ϵ}\left({\displaystyle \frac{p^2}{(qp)^2}}\right)^{2ϵ}{\displaystyle \frac{\mathrm{\Gamma }\left(\frac{1}{2}\frac{3ϵ}{2}\right)\mathrm{\Gamma }\left(12ϵ\right)\mathrm{\Gamma }\left(ϵ\right)}{\mathrm{\Gamma }\left(\frac{3}{2}\frac{7ϵ}{2}\right)\mathrm{\Gamma }\left(1+ϵ\right)}}.`$ (B18) The results are $`A_1=2`$ and $`B_1=2\left\{52\mathrm{log}2\frac{7\gamma }{2}+2\mathrm{log}\left(\frac{p^2}{(pq)^2}\right)\right\}`$. Multiplying the Laurent expansion of $`I_ϵ(p,q)`$ by the Laurent expansion of the multiplying factor in (B7) we get $$𝒲(q,p)=\frac{1}{2^{11}\pi ^4}\frac{1}{ϵ^2}\frac{1}{2^{10}\pi ^4}\left\{4\frac{11}{4}\gamma +\mathrm{log}\left(\frac{4\pi \mu ^2}{(pq)^2}\right)\right\}\frac{1}{ϵ}+\text{finite part}.$$ (B19) Let us go to (5.b). The sub diagram $`D_2(k)`$ is given by $$D_2(k)=𝒟q\epsilon ^{\mu \nu \lambda }\frac{q_\lambda }{q^2+i\eta }\epsilon _{\nu \mu \rho }\frac{(q+k)^\rho }{(q+k)^2+i\eta }.$$ (B20) After contracting the tensors in ($`2+1`$) dimension we are left with the (finite) integral $`D_2(k)`$ $`=`$ $`2{\displaystyle 𝒟q\frac{k.q}{[q^2+i\eta ][(q+k)^2+i\eta ]}}`$ (B21) $`=`$ $`{\displaystyle \frac{i}{8}}{\displaystyle \frac{(4\pi \mu ^2)^{\frac{ϵ}{2}}}{(k^2i\eta )^{\frac{1}{2}+\frac{ϵ}{2}}}},`$ (B22) where $`ϵ`$ was made zero whenever possible. Graph (5.b) is given by $$𝒢(p_1,p_2)=𝒟k\frac{i}{(p_1+k)^2+i\eta }\frac{i}{(p_2+k)^2+i\eta }D_2(k).$$ (B23) This integral is logarithmically divergent and their residue is independent of $`p_1`$ and $`p_2`$. To get this residue it is sufficient to calculate it for $`p_2=p_1`$ $$𝒢\frac{}{}_{p_2=p_1}=𝒟k\frac{i}{[(p+k)^2i\eta ]^2}\frac{1}{(k^2i\eta )^{\frac{1}{2}+\frac{ϵ}{2}}},$$ (B24) where whenever possible we have put $`ϵ=0`$. After introducing a Feynman parameter through (B2) and integrating in $`k`$ we get $$𝒢\frac{}{}_{p_2=p_1}=\frac{1}{2^5\pi ^2}\mathrm{\Gamma }(ϵ)(p_1^2)^ϵ=\frac{1}{32\pi ^2}\frac{1}{ϵ}+\mathrm{}.$$ (B25) Contribution of diagram (5.f) is given by $$(p)=i𝒟q𝒟k\epsilon ^{\mu \nu \rho }\frac{(p+q)_\rho }{(p+q)^2}\epsilon ^{\alpha \beta \lambda }\frac{q_\lambda }{q^2}\frac{g_{\nu \alpha }(2k+pq)_\mu (2kq)_\beta }{(k+p)^2(kq)^2k^2}.$$ (B26) or $$(p)=2i\epsilon ^{\mu \nu \rho }\epsilon ^{\alpha \beta \lambda }g_{\nu \alpha }\frac{\mu ^ϵ}{(2\pi )^d}𝒟q\frac{1}{(p+q)^2q^2}I_{\beta \mu \lambda \rho }(q),$$ (B27) where $$I_{\beta \mu \lambda \rho }(q)=d^dk\frac{2k_\beta k_\mu (q_\nu p_\rho +q_\nu q_\rho )+k_\beta (q_\nu q_\rho p_\mu q_\nu q_\mu p_\rho )}{(k+p)^2(kq)^2k^2}.$$ (B28) Using the identity $$\frac{1}{ABC}=2_0^1y𝑑y_0^1𝑑x\frac{1}{[C(1y)+y(Ax+B(1x))]^3}$$ (B29) and doing the $`k`$ integration we get $`(p)`$ $`=`$ $`{\displaystyle \frac{2\mathrm{\Gamma }\left(2\frac{d}{2}\right)\mu ^ϵ}{2^d\pi ^{\frac{d}{2}}}}\epsilon ^{\mu \nu \rho }\epsilon ^{\alpha \beta \lambda }g_{\nu \alpha }{\displaystyle _0^1}y𝑑y{\displaystyle _0^1}𝑑x{\displaystyle \frac{1}{[a^{}]^{3\frac{d}{2}}}}{\displaystyle 𝒟q\frac{1}{(p+q)^2q^2}}`$ (B30) $`\times `$ $`[{\displaystyle \frac{q_\nu q_\mu p_\rho p_\beta xy[(4d)y(x1)+1]+q_\nu q_\rho p_\mu p_\beta xy[(4d)xy1]}{[q^22q.p\frac{b^{}}{a^{}}p^2\frac{c^{}}{a^{}}]^{3\frac{d}{2}}}}`$ (B32) $`+{\displaystyle \frac{a^{}g_{\beta \mu }(q_\nu p_\rho +q_\nu q_\rho )}{[q^22q.p\frac{b^{}}{a^{}}p^2\frac{c^{}}{a^{}}]^{2\frac{d}{2}}}}]`$ with $`a^{}=y(x1)[y(x1)+1]`$, $`b^{}=xy^2(x1)`$ and $`c^{}=xy(xy1)`$. The only divergent term is the last monomial in the square bracket of (B32). We can so, write $$(p)=\frac{\mathrm{\Gamma }\left(2\frac{d}{2}\right)\mu ^ϵ}{2^{d+1}\pi ^{\frac{d}{2}}}\epsilon ^{\mu \nu \rho }\epsilon ^{\alpha \beta \lambda }g_{\nu \alpha }_0^1𝑑yy_0^1𝑑x\frac{1}{[a^{}]^{2\frac{d}{2}}}\frac{\mu ^ϵ}{(2\pi )^d}I_{DivPar}(p)+\text{fin parts}$$ (B33) where $$I(p)=d^dq\frac{q_\nu q_\rho }{(p+q)^2q^2[q^22q.p\frac{b^{}}{a^{}}p^2\frac{c^{}}{a^{}}]^{2\frac{d}{2}}}.$$ (B34) By reducing the denominators through the use of the identity $$\frac{1}{A^\alpha B^\beta C^\gamma }=\frac{\mathrm{\Gamma }(\alpha +\beta +\gamma )}{\mathrm{\Gamma }(\alpha )\mathrm{\Gamma }(\beta )\mathrm{\Gamma }(\gamma )}_0^1𝑑z_0^z𝑑t\frac{t^{\gamma 1}(zt)^{\beta \alpha }}{[A+(BA)z+(CB)t]^{\alpha +\beta +\gamma }}.$$ (B35) and doing the integration in $`q`$ we get for the divergent part $$I_{DivPart}(p)=i\frac{g_{\lambda \rho }}{2^{d+1}\pi ^{\frac{d}{2}}[p^2]^ϵ}\frac{\mathrm{\Gamma }(ϵ)}{\mathrm{\Gamma }\left(2\frac{d}{2}\right)}_0^1𝑑z_0^z𝑑tt^{1\frac{d}{2}}[a^{\prime \prime }b^{\prime \prime }]^{d3},$$ (B36) where $`a^{\prime \prime }=1z+\frac{b^{}}{a^{}}t`$ and $`b^{\prime \prime }=1z+\frac{c^{}}{a^{}}t`$. By inserting (B36) in (B34), expanding in $`ϵ`$ and doing the parametric integrations we get $$=\frac{i}{16\pi ^2}\frac{1}{ϵ}+\text{finite parts}.$$ (B37) ACKNOWLEDGEMENTS This work was partially supported by the Brazilian agencies: Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), Coordenadoria de Aperfeiçoamento de Pessoal de Ensino Superior (CAPES) and by Fundacão de Amparo a Pesquisa do Estado de São Paulo (FAPESP). Figure Captions * Figure 1. Feynman rules in the Landau gauge. * Figure 2. Divergent diagrams contributing to the CS 2 point function. * Figure 3. Divergent diagrams contributing to the scalar field 2 point function. * Figure 4. Divergent contributions to the scalar 6 point function. Three others, not drawn, diagrams similar to 4.n, 4.o and 4.p, but with the sense of all external lines reversed must also be considered. * Figure 5. Representation of the divergent integrals that appear in the diagrams of Figure 4. * Figure 6. Divergent contributions to $`\mathrm{\Gamma }_{[(\varphi ^{}\varphi )^n]}(2n)`$, that is, the 2n point function with one insertion of the composite operator $`[(\varphi ^{}\varphi )^n]`$. * Figure 7. Some possible contributions to $`\mathrm{\Gamma }_{[(\varphi ^{}\varphi )^n]}(2(n1))`$ ,the 2(n-1) point function with one insertion of the composite operator $`[(\varphi ^{}\varphi )^n]`$. * Figure 8. An example of a family of diagrams contributing to $`\mathrm{\Gamma }_\mu (q;p,p^{})`$. * Figure 9. The diagrams contributing to $`\mathrm{\Gamma }(p)`$, related by the Ward identity, to the family of diagrams in Figure 8. * Figure 10. An example of family of diagrams contributing to $`\mathrm{\Gamma }_{\mu \nu }(q,q^{};p,p^{})`$. * Figure 11. The diagrams contributing to $`\mathrm{\Gamma }_\mu `$, related by the Ward identity to the family of diagrams in Figure 10.
warning/0001/hep-ph0001233.html
ar5iv
text
# Particle Interferometry from 40 MeV to 40 TeV ## 1 Introduction Although the concept of Bose–Einstein or intensity interferometry was discovered in particle and nuclear physics more than 30 years ago , some basic questions in the field are still unanswered, namely, what the form of the Bose–Einstein correlation functions is, and what this form means. However, even if the ultimate understanding of the effect is still lacking, the level of sophistication in the theoretical descriptions and the level of sophistication in the experimental studies of Bose–Einstein correlations and particle interferometry has increased drastically, particularly in the field of heavy ion physics . ### 1.1 W-mass determination and particle interferometry The study of Bose–Einstein correlations is interesting in its own right, but it should be noted that consequences may spill over into other fields of research, that are seemingly unrelated. Such is the topic of the W-mass determination at LEP2, a top priority research in high energy physics. It turned out that the non-perturbative Bose–Einstein correlations between the pions from decaying W<sup>+</sup>W<sup>-</sup> pairs could be responsible for the presently largest systematic errors in W-mass determination at LEP2 . Hence, the theoretical understanding and the experimental control of Bose–Einstein correlations at LEP2 is essential to make a precision measurement of the W mass, which in turn may carry information via radiative corrections about the value of the Higgs mass or signals of new physics beyond the Standard Model. ### 1.2 Quark–gluon plasma and particle interferometry Heavy ion physics is the physics of colliding atomic nuclei. At the presently largest energies, the aim of heavy ion physics is to study the sub-nuclear degrees of freedom by successfully creating and identifying the quark–gluon plasma (QGP). This presently only hypothetical phase of matter would consist of freely moving quarks and gluons, over a volume which is macroscopical relative to the characteristic 1 fm size of hadrons. Theoretically proposed signals of the expected phase transition from hot hadronic matter to QGP were tested till now by fixed target experiments. At AGS, Brookhaven, collisions were made with nuclei as big as <sup>197</sup>Au accelerated to14.5 AGeV bombarding energy. At CERN SPS, collisions were made with 60 and 200 AGeV beams of <sup>16</sup>O nuclei, 200 AGeV beams of <sup>32</sup>S nuclei, 40 and 158 AGeV beams of <sup>208</sup>Pb nuclei . The really heavy projectile runs were made relatively recently, the data are being published and the implications of the new measurements are explored theoretically, with claims of a possible QGP production at CERN SPS Pb + Pb reactions, however, without a clear-cut experimental proof of the identification of the new phase . Both at CERN and at BNL, new collider experiments are planned and being constructed. The Relativistic Heavy Ion Collider (RHIC) at Brookhaven will collide $`100+100`$ AGeV <sup>197</sup>Au nuclei, which yields about 40 TeV total energy in the center of mass frame. RHIC started to deliver its first results in 2000. The construction stage of the RHIC accelerator rings was declared to be complete by the US Department of Energy on August 14, 1999, during a NATO Advanced Study Institute in Nijmegen, The Netherlands, where the material of this review paper has been presented. The forthcoming Large Hadron Collider (LHC) at CERN is scheduled to start in 2005. LHC will collide nuclei up to <sup>208</sup>Pb with 2.76 + 2.76 ATeV bombarding energy, yielding a total energy of 1150 TeV in the center of mass frame. The status quo has been summarized recently in Refs . At such large bombarding energies, the sub-nuclear structure of matter is expected to determine the outcome of the experiments. However, the observed single particle spectra and two-particle correlations indicated rather simple dependences on the transverse mass of the produced particles , that had a natural explanation in terms of hydrodynamical parameterizations. Although hydrodynamical type of models are also able to fit the final hadronic abundances, spectra and correlations, these models are not able to describe the ignition part of the process, thus their predictions are dependent on the assumed initial state. The hydro models come in two classes: i) hydro parameterizations, that attempt to parameterize the flow, temperature and density distributions on or around the freeze-out hypersurface by fitting the observed particle spectra and correlations, for example , but without solving the time-dependent (relativistic) hydrodynamical equations. The class ii) comes in the form of hydrodynamical solutions, that assume an equation of state and an initial condition, and follow the time evolution of the hydrodynamical system until a freeze-out hypersurface. These are better substantiated but more difficult to fit calculations, than class i) type of parameterizations. The exact hydro solutions are obtained either in analytical forms, , or from numerical solutions, see for example Refs . An even more substantiated approach is hydrodynamical approach with continuous emission of particles, which takes into account the small sizes of heavy ion reactions as compared to the mean free path of the particles . Such a continuous emission of hadrons during the time evolution of the hot and dense hadronic matter is supported by microscopic simulations . In principle, the exact hydrodynamical solutions can be utilized in a time-reversed form: after fixing the parameters to describe the measured particle spectra and correlations at the time when the particles are produced, the hydro code can be followed backwards in time, and one may learn about the initial condition in a given reaction: was it a QGP or a conventional hadron gas initial state? ### 1.3 Basics of quantum statistical correlations Essentially, intensity correlations appear due to the Bose–Einstein or Fermi–Dirac symmetrization of the two-particle final states of identical bosons or fermions, in short, due to quantum statistics. The simplest derivation is as follows: suppose that a particle pair is observed, one with momentum $`k_1`$ the other with momentum $`k_2`$. The amplitude has to be symmetrized over the unobservable variables, in particular over the points of emissions $`x_1`$ and $`x_2`$. If Coulomb, strong or other final-state interactions can be neglected, the amplitude of such a final state is proportional to $$A_{12}\frac{1}{\sqrt{2}}[\mathrm{e}^{ik_1x_1+ik_2x_2}\pm \mathrm{e}^{ik_1x_2+ik_2x_1}],$$ (1) where $`+`$ sign stands for bosons, $``$ for fermions. If the particles are emitted in an incoherent manner, the observable two-particle spectrum is proportional to $$N_2(k_1,k_2)𝑑x_1\rho (x_1)𝑑x_2\rho (x_2)|A_{12}|^2$$ (2) and the resulting two-particle intensity correlation function is $$C_2(k_1,k_2)=\frac{N_2(k_1,k_2)}{N_1(k_1)N_2(k_2)}=1\pm |\stackrel{~}{\rho }(k_1k_2)|^2$$ (3) that carries information about the Fourier-transformed space-time distribution of the particle emission $$\stackrel{~}{\rho }(q)=𝑑x\mathrm{e}^{iqx}\rho (x)$$ (4) as a function of the relative momentum $`q=k_1k_2`$. As compared to the idealized case when quantum-statistical correlations are negligible (or neglected), Bose–Einstein or Fermi–Dirac correlations modify the momentum distribution of the hadron pairs in the final state by a weight factor $`1\pm \mathrm{cos}[(\mathrm{k}_1\mathrm{k}_2)(\mathrm{x}_1\mathrm{x}_2)]`$. ### 1.4 Correlations between particle and heavy ion physics In case of pions, that are produced abundantly in relativistic heavy ion experiments, Bose–Einstein symmetrization results in an enhancement of correlations of pion pairs with small relative momentum, and the correlation function carries information about the space-time distribution of pion production points. This in turn is expected to be sensitive to the formation of a transient quark–gluon plasma stage . In particle physics, reshuffling or modification of the momentum of pions in the fully hadronic decays of the W<sup>+</sup>W<sup>-</sup> pairs happens due to the Bose–Einstein symmetrization of the full final stage, that includes symmetrization of pions with similar momentum from different W-s. As a consequence of this quantum interference of pions, a systematic error as big as 100 MeV may be introduced to the W-mass determination from reconstruction of the invariant masses of $`(\mathrm{q}\overline{\mathrm{q}})`$ systems in 4-jet events . It is very difficult to handle the quantum interference of pions from the W<sup>+</sup> and W<sup>-</sup> jets with Monte-Carlo simulations, perturbative calculations and other conventional methods of high energy physics. Unexpectedly, a number of recent experimental results arose suggesting that the Bose–Einstein correlations and the soft components of the single-particle spectra in high energy collisions of elementary particles show similar features to the same observables in high energy heavy ion physics . These striking similarities of multi-dimensional Bose–Einstein correlations and particle spectra in high energy particle and heavy ion physics have no fully explored dynamical explanation yet. This review intends to give a brief introduction to various sub-fields of particle interferometry, highlighting those phenomena that may have applications or analogies in various different type of reactions. The search for such analogies inspired a study of non-relativistic heavy ion reactions in the 30 – 80 AMeV energy domain and a search for new exact analytic solutions of fireball hydrodynamics, reviewed briefly for a comparison. As some of the sections are mathematically more advanced, and other sections deal directly with data analysis, I attempted to formulate the various sections so that they be self-standing as much as possible, and be of interest for both the experimentally and the theoretically motivated readers. ## 2 Formalism The basic properties of the Bose–Einstein $`n`$-particle correlation functions (BECF-s) can be summarized as follows, using only the generic aspects of their derivation. The $`n`$-particle Bose–Einstein correlation function is defined as $$C_n(𝐤_1,\mathrm{},𝐤_n)=\frac{N_n(𝐤_1,\mathrm{},𝐤_n)}{N_1(𝐤_1)\mathrm{}N_1(𝐤_n)},$$ (5) where $`N_n(𝐤_1,\mathrm{},𝐤_n)`$ is the $`n`$-particle inclusive invariant momentum distribution, while $$N_n(𝐤_1,\mathrm{},𝐤_n)=\frac{1}{\sigma }E_{𝐤_1}\mathrm{}E_{𝐤_n}\frac{d^{3n}\sigma }{d𝐤_1\mathrm{}d𝐤_n}$$ (6) is the invariant $`n`$-particle inclusive momentum distribution. It is quite remarkable that the complicated object of Eq. (5) carries quantum mechanical information on the phase-space distribution of particle production as well as on possible partial coherence of the source, can be expressed in a relatively simple, straight-forward manner both in the analytically solvable pion-laser model of Refs as well as in the generic boosted-current formalism of Gyulassy, Padula and collaborators as $$C_n(𝐤_1,\mathrm{},𝐤_n)=\frac{{\displaystyle \underset{\sigma ^{(n)}}{}}{\displaystyle \underset{i=1}{\overset{n}{}}}G(𝐤_i,𝐤_{\sigma _i})}{{\displaystyle \underset{i=1}{\overset{n}{}}}G(𝐤_i,𝐤_i)},$$ (7) where $`\sigma ^{(n)}`$ stands for the set of permutations of indices $`(1,2,\mathrm{},n)`$ and $`\sigma _i`$ denotes the element replacing element $`i`$ in a given permutation from the set of $`\sigma ^{(n)}`$, and, regardless of the details of the two different derivations, $$G(𝐤_i,𝐤_j)=\sqrt{E_{𝐤_i}E_{𝐤_j}}a^{}(𝐤_i)a(𝐤_j)$$ (8) stands for the expectation value of $`a^{}(𝐤_i)a(𝐤_j)`$. The operator $`a^{}(𝐤)`$ creates while operator $`a(𝐤)`$ annihilates a boson with momentum $`𝐤`$. The quantity $`G(𝐤_i,𝐤_j)`$ corresponds to the first order correlation function in the terminology of quantum optics. In the boosted-current formalism, the derivation of Eq. (7) is based on the assumptions that i) the bosons are emitted from a semi-classical source, where currents are strong enough so that the recoils due to radiation can be neglected, ii) the source corresponds to an incoherent random ensemble of such currents, as given in a boost-invariant formulation in Ref. , and iii) that the particles propagate as free plane waves after their production. Possible correlated production of pairs of particles is neglected here. Note also the recent clarification of the proper normalization of the two-particle Bose–Einstein correlations . A formally similar result is obtained when particle production happens in a correlated manner, generalizing the results of Refs . Namely, the $`n`$-particle exclusive invariant momentum distributions of the pion-laser model read as $$N_n^{(n)}(𝐤_1,\mathrm{},𝐤_n)=\underset{\sigma ^{(n)}}{}\underset{i=1}{\overset{n}{}}G_1(𝐤_i,𝐤_{\sigma _i}),$$ (9) with $$G_1(𝐤_i𝐤_j)=\sqrt{E_{𝐤_i}E_{𝐤_j}}\text{Tr}\{\widehat{\rho }_1a^{}(𝐤_i)a(𝐤_j)\},$$ (10) where $`\widehat{\rho }_1`$ is the single-particle density matrix in the limit when higher-order Bose–Einstein correlations are negligible. Q.H. Zhang has shown , that the $`n`$-particle inclusive spectrum has a similar structure: $`N_n(𝐤_1,\mathrm{},𝐤_n)`$ $`=`$ $`{\displaystyle \underset{\sigma ^{(n)}}{}}{\displaystyle \underset{i=1}{\overset{n}{}}}G(𝐤_i,𝐤_{\sigma _i}),`$ (11) $`G(𝐤_i,𝐤_j)`$ $`=`$ $`{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}G_n(𝐤_i,𝐤_j).`$ (12) This result, valid only if the density of pions is below a critical value , was obtained if the multiplicity distribution was assumed to be a Poissonian one in the rare gas limit. The formula of Eq. (12) has been generalized by Q.H. Zhang in Ref. to the case when the multiplicity distribution in the rare gas limit is arbitrary. The functions $`G_n(𝐤_i,𝐤_j)`$ can be considered as representatives of order $`n`$ symmetrization effects in exclusive events where the multiplicity is fixed to $`n`$, see Refs for more detailed definitions. The function $`G(𝐤_i,𝐤_j)`$ can be considered as the expectation value of $`a^{}(𝐤_i)a(𝐤_j)`$ in an inclusive sample of events, and this building block includes all the higher-order symmetrization effects. In the relativistic Wigner function formalism, in the plane wave approximation $`G(𝐤_1,𝐤_2)`$ can be rewritten as $`G(𝐤_1,𝐤_2)`$ $``$ $`\stackrel{~}{S}(q_{12},K_{12})={\displaystyle d^4xS(x,K_{12})\mathrm{exp}(iq_{12}x)},`$ (13) $`K_{12}`$ $`=`$ $`0.5(k_1+k_2),`$ (14) $`q_{12}`$ $`=`$ $`k_1k_2,`$ (15) where a four-vector notation is introduced, $`k=(E_𝐤,𝐤)`$, and the energy of quanta with mass $`m`$ is given by $`E_𝐤=\sqrt{m^2+𝐤^2}`$, the mass-shell constraint. Notation $`ab`$ stands for the inner product of four-vectors. In the following, the relative momentum four-vector shall be denoted also as $`\mathrm{\Delta }k=q=(q_0,q_x,q_y,q_z)=(q_0,𝐪)`$, the invariant relative momentum is $`Q=\sqrt{qq}`$. The covariant Wigner transform of the source density matrix, $`S(x,𝐤)`$ is a quantum-mechanical analogue of the classical probability that a boson is produced at a given $`(x,k)`$ point in the phase-space, where $`x=(t,𝐫)=(t,r_x,r_y,r_z)`$. The quantity $`S(x,K_{12})`$ corresponds to the off-shell extrapolation of $`S(x,𝐤)`$, as $`K_{12}^0\sqrt{m^2+𝐊_{12}^2}`$. Fortunately, Bose–Einstein correlations are non-vanishing at small values of the relative momentum $`q`$, where $`K_{12}^0E_{𝐊_{12}}`$. Due to the mass-shell constraints, $`G`$ depends only on 6 independent momentum components. For the two-particle Bose–Einstein correlation function, Eqs (7,8,13) yield the following representation: $$C_2(𝐤_1,𝐤_2)=1+\frac{|\stackrel{~}{S}(q_{12},K_{12})|^2}{\stackrel{~}{S}(0,𝐤_1)\stackrel{~}{S}(0,𝐤_2)}.$$ (16) Due to the unknown off-shell behavior of the Wigner functions, it is rather difficult to evaluate this quantity from first principles, in a general case. When comparing model results to data, two kind of simplifying approximations are frequently made: i) The on-shell approximation can be used for developing Bose–Einstein afterburners to Monte-Carlo event generators, where only the on-shell part of the phase-space is modeled. In this approximation, Eq. (16) is evaluated with the on-shell mean momentum, $`\stackrel{~}{K}=(\sqrt{m^2+𝐊_{12}^2},𝐊_{12})`$. This on-shell approximation was used e.g. in Ref. to sample $`S(x,\stackrel{~}{K})`$ from the single-particle phase-space distribution given by Monte-Carlo event generators, and to calculate the corresponding Bose–Einstein correlation functions in a numerically efficient manner. The method yields a straightforward technique for the inclusion of Coulomb and strong final-state interactions as well, see e.g. Ref. . ii) The smoothness approximation can be used when describing Bose–Einstein correlations from a theoretically parameterized model, e.g. from a hydrodynamical calculation. In this case, the analytic continuation of $`S(x,𝐤)`$ to the off-shell values of $`K`$ is providing a value for the off-shell Wigner function $`S(x,K_{12})`$. However, in the normalization of Eq. (16), the product of two on-shell Wigner functions appear. In the smoothness approximation, one evaluates this product as a leading order Taylor series in $`q`$ of the exact expression $`\stackrel{~}{S}(0,𝐊𝐪/2)S(0,𝐊+𝐪/2)`$. The resulting formula, $$C_2(𝐤_1,𝐤_2)=1+\frac{|\stackrel{~}{S}(q_{12},K_{12})|^2}{|\stackrel{~}{S}(0,K_{12})|^2},$$ (17) relates the two-particle Bose–Einstein correlation function to the Fourier-transformed off-shell Wigner function $`S(x,K)`$. This provides an efficient analytic or numeric method to calculate the BECF from sources with known functional forms. The correction terms to the smoothness approximation of Eq. (17) are given in Ref. . These corrections are generally on the 5% level for thermal like momentum distributions. ## 3 Model-Independent Analysis of Short-Range Correlations Can one model-independently characterize the shape of two-particle correlation functions? Let us attempt to answer this question on the level of statistical analysis, without theoretical assumptions on the thermal or non-thermal nature of the particle emitting source. In this approach, the usual theoretical assumptions are not made, neither on the presence or the negligibility of Coulomb and other final-state interactions, nor on the presence or the negligibility of a coherent component in the source, nor on the presence or the negligibility of higher-order quantum statistical symmetrization effects, nor on the presence or the negligibility of dynamical effects (e.g. fractal structure of gluon-jets) on the short-range part of the correlation functions. The presentation follows the lines of Ref. . The reviewed method is really model-independent, and it can be applied not only to Bose–Einstein correlation functions but to every experimentally determined function, which features the properties i) and ii) listed below. The following experimental properties are assumed: i) The measured function tends to a constant for large values of the relative momentum. ii) The measured function has a non-trivial structure at a certain value of its argument. The location of the non-trivial structure in the correlation function is assumed for simplicity to be close to $`Q=0`$. The properties i) and ii) are well satisfied by e.g. the conventionally used two-particle Bose–Einstein correlation functions. For a critical review on the non-ideal features of short-range correlations, (e.g. non-Gaussian shapes in multi-dimensional Bose–Einstein correlation studies), we recommend Ref. . The core/halo intercept parameter $`\lambda _{}`$ is defined as the extrapolated value of the two-particle correlation function at $`Q=0`$, see Section 5 for greater details. It turns out that $`\lambda _{}`$ is an important physical observable, related to the degree of partial restoration of $`U_A(1)`$ symmetry in hot and dense hadronic matter , as reviewed in Section 15. Various non-ideal effects due to detector resolution, binning, particle mis-identification, resonance decays, details of the Coulomb and strong final-state interactions etc. may influence this parameter of the fit. One should also mention, that if all of these difficulties are corrected for by the experiment, the extrapolated intercept parameter $`\lambda _{}`$ for like-sign charged bosons is (usually) not larger, than unity as a consequence of quantum statistics for chaotic sources, even with a possible admixture of a coherent component. However, final-state interactions, fractal branching processes of gluon jets, or the appearance of one-mode or two-mode squeezed states in the particle emitting source might provide arbitrarily large values for the intercept parameter. A really model-independent approach is to expand the measured correlation functions in an abstract Hilbert space of functions. It is reasonable to formulate such an expansion so that already the first term in the series be as close to the measured data points as possible. This can be achieved if one identifies the approximate shape (e.g. the approximate Gaussian or the exponential shape) of the correlation function with the abstract measure $`\mu (t)dt`$ in the abstract Hilbert-space $``$. The orthonormality of the basis functions $`\varphi _n(t)`$ in $``$ can be utilized to guarantee the convergence of these kind of expansions, see Refs for greater details. ### 3.1 Laguerre expansion and exponential shapes If in a zeroth order approximation the correlation function has an exponential shape, then it is an efficient method to apply the Laguerre expansion, as a special case of the general formulation of Refs : $`C_2(Q)`$ $`=`$ $`𝒩\left\{1+\lambda _L\mathrm{exp}(QR_L)\left[1+c_1L_1(QR_L)+{\displaystyle \frac{c_2}{2!}}L_2(QR_L)+\mathrm{}\right]\right\}.`$ In this and the next subsection, $`Q`$ stands symbolically for any, experimentally chosen, one dimensional relative momentum variable. The fit parameters are the scale parameters $`𝒩`$, $`\lambda _L`$, $`R_L`$ and the expansion coefficients $`c_1`$, $`c_2`$, … . The $`n`$th order Laguerre polynomials are defined as $`L_n(t)`$ $`=`$ $`\mathrm{exp}(t){\displaystyle \frac{d^n}{dt^n}}t^n\mathrm{exp}(t),`$ (19) they form a complete orthogonal basis for an exponential measure as $`\delta _{n,m}`$ $``$ $`{\displaystyle _0^{\mathrm{}}}𝑑t\mathrm{exp}(t)L_n(t)L_m(t).`$ (20) The first few Laguerre polynomials are explicitly given as $`L_0(t)`$ $`=`$ $`1,`$ (21) $`L_1(t)`$ $`=`$ $`t1,`$ (22) $`L_2(t)`$ $`=`$ $`t^24t+2,\mathrm{}.`$ (23) As the Laguerre polynomials are non-vanishing at the origin, $`C(Q=0)1+\lambda _L`$. The physically significant core/halo intercept parameter $`\lambda _{}`$ can be obtained from the parameter $`\lambda _L`$ of the Laguerre expansion as $`\lambda _{}`$ $`=`$ $`\lambda _L[1c_1+c_2\mathrm{}].`$ (24) ### 3.2 Edgeworth expansion and Gaussian shapes If, in a zeroth-order approximation, the correlation function has a Gaussian shape, then the general form given in Ref. takes the particular form of the Edgeworth expansion as: $`C(Q)`$ $`=`$ $`𝒩\{1+\lambda _E\mathrm{exp}(Q^2R_E^2)\times `$ (25) $`[1+{\displaystyle \frac{\kappa _3}{3!}}H_3(\sqrt{2}QR_E)+{\displaystyle \frac{\kappa _4}{4!}}H_4(\sqrt{2}QR_E)+\mathrm{}]\}.`$ The fit parameters are the scale parameters $`𝒩`$, $`\lambda _E`$, $`R_E`$, and the expansion coefficients $`\kappa _3`$, $`\kappa _4`$, … that coincide with the cumulants of rank 3, 4, … of the correlation function. The Hermite polynomials are defined as $`H_n(t)`$ $`=`$ $`\mathrm{exp}(t^2/2)\left({\displaystyle \frac{d}{dt}}\right)^n\mathrm{exp}(t^2/2),`$ (26) they form a complete orthogonal basis for a Gaussian measure as $`\delta _{n,m}`$ $``$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑t\mathrm{exp}(t^2/2)H_n(t)H_m(t).`$ (27) The first few Hermite polynomials are listed as $`H_1(t)`$ $`=`$ $`t,`$ (28) $`H_2(t)`$ $`=`$ $`t^21,`$ (29) $`H_3(t)`$ $`=`$ $`t^33t,`$ (30) $`H_4(t)`$ $`=`$ $`t^46t^2+3,\mathrm{}`$ (31) The physically significant core/halo intercept parameter $`\lambda _{}`$ can be obtained from the Edgeworth fit of Eq. (25) as $`\lambda _{}`$ $`=`$ $`\lambda _E\left[1+{\displaystyle \frac{\kappa _4}{8}}+\mathrm{}\right].`$ (32) This expansion technique was applied in the conference contributions to the AFS minimum bias and 2-jet events to characterize successfully the deviation of data from a Gaussian shape. It was also successfully applied to characterize the non-Gaussian nature of the correlation function in two-dimensions in case of the preliminary E802 data in Ref. , and it was recently applied to characterize the non-Gaussian nature of the three-dimensional two-pion BECF in e$`{}_{}{}^{+}+`$ e<sup>-</sup> reactions at LEP1 . Figure 1 indicates the ability of the Laguerre expansions to characterize two well-known, non-Gaussian correlation functions : the second-order short-range correlation function $`D_2^s(Q)`$ as determined by the UA1 and the NA22 experiments . The convergence criteria of the Laguerre and the Edgeworth expansions is given in Ref. . From Table 1 the core/halo model intercept parameter is obtained as $`\lambda _{}=1.14\pm 0.10`$ (UA1) and $`\lambda _{}=1.11\pm 0.17`$ (NA22). As both of these values are within errors equal to unity, the maximum of the possible value of the intercept parameter $`\lambda _{}`$ in a fully chaotic source, we conclude that either there are other than Bose–Einstein short-range correlations observed by both collaborations, or the full halo of long lived resonances is resolved in case of this measurement . If the two-particle BECF can be factorized as a product of (two or more) functions of one variable each, then the Laguerre and the Edgeworth expansions can be applied to the multiplicative factors — functions of one variable, each. This method was applied recently to study the non-Gaussian features of multi-dimensional Bose–Einstein correlation functions e.g. in Refs . The full, non-factorized form of two-dimensional Edgeworth expansion and the interpretation of its parameters is described in the handbook on mathematical statistics by Kendall and Stuart . ## 4 Coulomb Wave Corrections for Higher-Order Correlations The short-range part of the two- and multi-particle correlation function of charged particles is strongly effected by Coulomb interactions. Even in the non-relativistic case, the $`n`$-body Coulomb scattering problem is solvable exactly only for the $`n=2`$ case, the full 3-body Coulomb wave-function is unknown. However, when studying higher-order Bose–Einstein correlations and e.g. searching for the onset of (partial) coherence in the source, it is desired that the Coulomb-induced correlations be removed from the data. In any given frame, the boost-invariant decomposition of Eq. (13) can be rewritten into the following, seemingly not invariant form: $`G(𝐤_1,𝐤_2)`$ $`=`$ $`{\displaystyle d^3𝐱S_{𝐊_{12}}(𝐱)\mathrm{exp}(i𝐪_{12}𝐱)},`$ (33) $`S_{𝐊_{12}}(𝐱)`$ $`=`$ $`{\displaystyle 𝑑t\mathrm{exp}(i𝜷_{K_{12}}𝐪_{12}t)S(𝐱,t,K_{12})},`$ (34) $`𝜷_{K_{12}}`$ $`=`$ $`(𝐤_1+𝐤_2)/(E_1+E_2).`$ (35) Note that the relative source function $`S_{𝐊_{12}}(𝐱)`$ reduces to a simple time integral over the source function $`S(x,K)`$ in the frame where the mean momentum of the pair (hence the pair velocity $`𝜷_{K_{12}}`$) vanishes. Based on a Poisson cluster picture, the effect of multi-particle Coulomb final-state interactions on higher-order intensity correlations is determined in general in Ref. , with the help of a scattering wave function which is a solution of the $`n`$-body Coulomb Schrödinger equation in (a large part of) the asymptotic region of the $`n`$-body configuration space. If $`n`$ particles are emitted with similar momenta, so that their $`n`$-particle Bose–Einstein correlation functions may be non-trivial, Eqs (3335) form the basis for evaluation of the Coulomb and strong final-state interaction effects on the observables for any given cluster of particles, assuming that the relative motion of the particles is non-relativistic within the cluster, see Ref. . The Coulomb correction factor $`K^1`$ can be integrated for arbitrary large number of particles and for any kind of model source, by replacing the plane wave approximation with the approximate $`n`$-body Coulomb wave-function. In the limit of vanishing source sizes, the generalization of the Gamow penetration factor was obtained to the correlation function of arbitrary large number of particles . In particular, Coulomb effects on the $`n`$-particle Bose–Einstein correlation functions of similarly charged particles were studied for Gaussian effective sources, for $`n=3`$ in Ref. and for $`n=4`$ and 5 in Ref. . For the typical $`R=1`$ fm effective source sizes of the elementary particle reactions, the generalized $`n`$-body Gamow penetration factor gave rather precise estimates of the Coulomb correction (within 5% from the Coulomb-wave correction). In contrast, for typical effective source sizes observed in high energy heavy ion reactions, Fig. 2 indicates that the new Coulomb wave-function integration method allows for a removal of a systematic error as big as 100% from higher-order multi-particle Bose–Einstein correlation functions. See Ref. for greater details. ## 5 Core/Halo Picture of Bose–Einstein Correlations The core/halo model deals with the consequences of a phenomenological situation, when the boson source can be considered to be a superposition of a central core surrounded by an extended halo. In the forthcoming sections, final-state interactions are neglected, we assume that the data are corrected for final-state Coulomb (and possibly strong) interactions. Bose–Einstein correlations are measured at small relative momenta of particle pairs. In order to reliably separate the near-by tracks of particle pairs in the region of the Bose enhancement, each experiment imposes a cut-off $`Q_{\mathrm{min}}`$, the minimum value of the resolvable relative momentum. The value of this cut-off may vary slightly from experiment to experiment, but such a cut-off exists in each measurement. In the core/halo model, the following assumptions are made: Assumption 0: The emission function does not have a no-scale, power-law-like structure. This possibility was discussed and related to intermittency and effective power-law shapes of the two-particle Bose–Einstein correlation functions in Ref. . Assumption 1: The bosons are emitted either from a central part or from the surrounding halo. Their emission functions are indicated by $`S_c(x,𝐤)`$ and $`S_h(x,𝐤)`$, respectively. According to this assumption, the complete emission function can be written as $`S(x,𝐤)=S_c(x,𝐤)+S_h(x,𝐤),`$ (36) and $`S(x,𝐤)`$ is normalized to the mean multiplicity, $`d^4x(d𝐤/E)S(x,𝐤)=n`$. Assumption 2: The emission function that characterizes the halo is assumed to change on a scale $`R_h`$ that is larger than $`R_{\mathrm{max}}\mathrm{}/Q_{\mathrm{min}}`$, the maximum length-scale resolvable by the intensity interferometry microscope. The smaller central core of size $`R_c`$ is assumed to be resolvable, $`R_h>R_{\mathrm{max}}>R_c.`$ (37) This inequality is assumed to be satisfied by all characteristic scales in the halo and in the central part, e.g. in case the side, out or longitudinal components of the correlation function are not identical. Assumption 3: The core fraction $`f_c(𝐤)=N_c(𝐤)/N_1(𝐤)`$ varies slowly on the relative momentum scale given by the correlator of the core . The emission function of the core and the halo are normalized as $`{\displaystyle d^4x\frac{d𝐤}{E}S_c(x,𝐤)}=n_c`$ and $`{\displaystyle d^4x\frac{d𝐤}{E}S_h(x,𝐤)}=n_h.`$ (38) One finds that $`N_1(𝐤)=N_c(𝐤)+N_h(𝐤)`$ and $`n=n_c+n_h.`$ (39) Note that in principle the core as well as the halo part of the emission function could be decomposed into more detailed contributions, e.g. $`S_h(x,𝐤)`$ $`=`$ $`{\displaystyle \underset{r=\omega ,\eta ,\eta ^{},K_S^0}{}}S_{\mathrm{halo}}^{(r)}(x,𝐤).`$ (40) In case of pions and NA44 acceptance, the $`\omega `$ mesons were shown to contribute to the halo, Ref. . For the present considerations, this separation is indifferent, as the halo is defined with respect to $`Q_{\mathrm{min}}`$, the experimental two-track resolution. For example, if $`Q_{\mathrm{min}}=1015`$ MeV, the decay products of the $`\omega `$ resonances can be taken as parts of the halo . If future experimental resolution decreases below $`5`$ MeV and the error bars on the measurable part of the correlation function decrease significantly in the $`Q<\mathrm{}/\mathrm{\Gamma }_\omega =8`$ MeV region, the decay products of the $`\omega `$ resonances will contribute to the resolvable core, see Refs for greater details. If Assumption 3 is also satisfied by some experimental data set, then Eq. (16) yields a particularly simple form of the two-particle Bose–Einstein correlation function: $`C_2(𝐤_1,𝐤_2)`$ $`=`$ $`1+\lambda _{}(𝐊){\displaystyle \frac{\stackrel{~}{S}_c(\mathrm{\Delta }k,𝐊)^2}{\stackrel{~}{S}_c(0,𝐤_1)\stackrel{~}{S}_c(0,𝐤_2)}},`$ (41) $``$ $`\mathrm{\hspace{0.17em}1}+\lambda _{}(𝐊){\displaystyle \frac{\stackrel{~}{S}_c(\mathrm{\Delta }k,K)^2}{|\stackrel{~}{S}_c(0,K)|^2}},`$ (42) where mean and the relative momentum four-vectors are defined as $$K=0.5(k_1+k_2),\mathrm{\Delta }k=k_1k_2,$$ (43) with $`K=(K^0,𝐊)`$ and $`\mathrm{\Delta }k=(\mathrm{\Delta }k^0,𝚫𝐤)`$, and the effective intercept parameter $`\lambda _{}(𝐊)`$ is given as $$\lambda _{}(𝐊)=\left[N_c(𝐊)/N_1(𝐊)\right]^2.$$ (44) As emphasized in Ref. , this effective intercept parameter $`\lambda _{}`$ shall in general depend on the mean momentum of the observed boson pair, which within the errors of $`Q_{\mathrm{min}}`$ coincides with any of the on-shell four-momentum $`k_1`$ or $`k_2`$. Note that $`\lambda _{}\lambda _{xct}=1`$, the latter being the exact intercept parameter at $`Q=0`$ MeV. The core/halo model is summarized in Fig. 3, see Ref. for further details. The core/halo model correlation function is compared to the so-called “model-independent”, Gaussian approximation of Refs and to the full correlation function in Fig. 4, see appendix of Ref. and that of Ref. for further details. The measured two-particle BECF is determined for $`𝚫𝐤>Q_{\mathrm{min}}10`$ MeV/c, and any structure within the $`𝚫𝐤<Q_{\mathrm{min}}`$ region is not resolved. However, the $`(c,h)`$ and $`(h,h)`$ type boson pairs create a narrow peak in the BECF exactly in this $`\mathrm{\Delta }k`$ region according to Eq. (36), which cannot be resolved according to Assumption 2. The general form of the BECF of systems with large halo, Eq. (42), coincides with the most frequently applied phenomenological parameterizations of the BECF in high energy heavy ion as well as in high energy particle reactions . Previously, this form has received a lot of criticism from the theoretical side, claiming that it is in disagreement with quantum statistics or that the $`\lambda `$ parameter is just a kind of fudge parameter, “a measure of our ignorance”. In the core/halo picture, Eq. (42) is derived with a standard inclusion of quantum statistical effects. Reactions including e$`{}_{}{}^{+}+`$ e<sup>-</sup> annihilations, lepton–hadron and hadron–hadron reactions, nucleon–nucleus and nucleus–nucleus collisions are phenomenologically well described by a core/halo picture. ### 5.1 Partial coherence and higher-order correlations In earlier studies of the core/halo model it was assumed that $`S_c(x,p)`$ describes a fully incoherent (thermal) source. In Ref. an additional assumption was also made: Assumption 4: A part of the core may emit bosons in a coherent manner: $$S_c(x,𝐤)=S_c^p(x,𝐤)+S_c^i(x,𝐤),$$ (45) where upper index $`p`$ stands for coherent component (which leads to partial coherence), upper index $`i`$ stands for incoherent component of the source. The invariant spectrum is given by $$N(𝐤)=d^4xS(x,𝐤)=N_c(𝐤)+N_h(𝐤)$$ (46) and the core contribution is a sum: $$N_c(𝐤)=d^4xS_c(x,𝐤)=N_c^p(𝐤)+N_c^i(𝐤).$$ (47) One can introduce the momentum dependent core fractions $`f_c(𝐤)`$ and partially coherent fractions $`p(𝐤)`$ as $`f_c(𝐤)`$ $`=`$ $`N_c(𝐤)/N(𝐤),`$ (48) $`p_c(𝐤)`$ $`=`$ $`N_c^p(𝐤)/N_c(𝐤).`$ (49) Hence the halo and the incoherent fractions $`f_h,p_i`$ are $`f_h(𝐤)`$ $`=`$ $`N_h(𝐤)/N(𝐤)=1f_c(𝐤),`$ (50) $`f_i(𝐤)`$ $`=`$ $`N_c^i(𝐤)/N_c(𝐤)=1p_c(𝐤).`$ (51) ### 5.2 Strength of the $`n`$-particle correlations We denote the $`n`$-particle correlation function of Eq. (5) as $`C_n(1,2,\mathrm{},n)`$ $`=`$ $`C_n(𝐤_1,𝐤_2,\mathrm{},𝐤_n)={\displaystyle \frac{N_n(1,2,\mathrm{},n)}{N_1(1)N_1(2)\mathrm{}N_1(n)}},`$ (52) where a symbolic notation for $`𝐤_i`$ is introduced, only the index of $`𝐤`$ is written out in the argument. In the forthcoming, we shall apply this notation consistently for the arguments of various functions of the momenta, i.e. $`f(𝐤_i,𝐤_j,\mathrm{},𝐤_m)`$ is symbolically denoted by $`f(i,j,\mathrm{},m)`$. The strength of the $`n`$-particle correlation function (extrapolated from a finite resolution measurement to zero relative momentum for each pair) is denoted by $`C_n(0)`$, given by the following simple formula, $$C_n(0)=1+\underset{j=2}{\overset{n}{}}({}_{j}{}^{n})\alpha _jf_c^j[(1p_c)^j+jp_c(1p_c)^{j1}].$$ (53) Here, $`\alpha _j`$ indicates the number of permutations, that completely mix exactly $`j`$ non-identical elements. There are $`({}_{j}{}^{n})`$ different ways to choose $`j`$ different elements from among $`n`$ different elements. Since all the $`n!`$ permutations can be written as a sum over the fully mixing permutations, the counting rule yields a recurrence relation for $`\alpha _j`$, Refs : $`\alpha _n`$ $`=`$ $`n!{\displaystyle \underset{j=0}{\overset{n1}{}}}({}_{j}{}^{n})\alpha _j,`$ (54) $`\alpha _0`$ $`=`$ $`1.`$ (55) The first few values of $`\alpha _j`$ are given as $$\alpha _1=0,\alpha _2=1,\alpha _3=2,\alpha _4=9,\alpha _5=44,\alpha _6=265,$$ (56) the first few intercept parameters, $`\lambda _{,n}=C_n(0)1`$, are given as $`\lambda _{,2}`$ $`=`$ $`f_c^2[(1p_c)^2+2p_c(1p_c)],`$ (57) $`\lambda _{,3}`$ $`=`$ $`3f_c^2[(1p_c)^2+2p_c(1p_c)]`$ (58) $`+2f_c^3[(1p_c)^3+3p_c(1p_c)^2],`$ $`\lambda _{,4}`$ $`=`$ $`6f_c^2[(1p_c)^2+2p_c(1p_c)]`$ (59) $`+8f_c^3[(1p_c)^3+3p_c(1p_c)^2]`$ $`+9f_c^4[(1p_c)^4+4p_c(1p_c)^3],`$ $`\lambda _{,5}`$ $`=`$ $`10f_c^2[(1p_c)^2+2p_c(1p_c)]`$ (60) $`+20f_c^3[(1p_c)^3+3p_c(1p_c)^2]`$ $`+45f_c^4[(1p_c)^4+4p_c(1p_c)^3]`$ $`+44f_c^5[(1p_c)^5+5p_c(1p_c)^4].`$ In general, terms proportional to $`f_c^j`$ in the incoherent case shall pick up an additional factor $`[(1p_c)^j+jp_c(1p_c)^{j1}]`$ in case the core has a coherent component . This extra factor means that either all $`j`$ particles must come from the incoherent part of the core, or one of them must come from the coherent, the remaining $`j1`$ particles from the incoherent part. If two or more particles come from the coherent component of the core, the contribution to intensity correlations vanishes as the intensity correlator for two coherent particles is zero . If the coherent component is present, one can introduce the normalized incoherent and partially coherent core fractions as $`\stackrel{~}{s}_c^i(j,k)`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{S}_c^i(j,k)}{\stackrel{~}{S}_c^i(j,j)}},`$ (61) $`\stackrel{~}{s}_c^p(j,k)`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{S}_c^p(j,k)}{\stackrel{~}{S}_c^p(j,j)}}.`$ (62) In the partially coherent core/halo picture, one obtains the following closed form for the order $`n`$ Bose–Einstein correlation functions : $`C_n(1,\mathrm{},n)`$ $`=`$ $`1+{\displaystyle \underset{j=2}{\overset{n}{}}}{\displaystyle \underset{m_1\mathrm{}m_j=1}{\overset{n_{}}{}}}{\displaystyle \underset{\rho ^{(j)}}{}}\{{\displaystyle \underset{k=1}{\overset{j}{}}}f_c(m_k)[1p_c(m_k)]\stackrel{~}{s}_c^i(m_k,m_{\rho _k})`$ (63) $`+{\displaystyle \underset{l=1}{\overset{j}{}}}f_c(m_l)p_c(m_l)\stackrel{~}{s}_c^p(m_l,m_{\rho _l}){\displaystyle \underset{k=1,kl}{\overset{j}{}}}f_c(m_k)[1p_c(m_k)]\stackrel{~}{s}_c^i(m_k,m_{\rho _k})\}.`$ Here, $`\rho ^{(j)}`$ stands for the set of permutations that completely mix exactly $`j`$ elements, $`\rho _i`$ stands for the permuted value of index $`i`$ in one of these permutations. By definition, $`\rho _ii`$ for all $`i=1,2,\mathrm{},j`$. The notation $`\mathrm{\Sigma }^{}`$ indicates summation for different values of indexes, $`m_im_l`$ for all $`i,l`$ pairs. The expression Eq. (63) contains two (momentum dependent) phases in the Fourier-transformed, normalized source distributions: one denoted by $`\varphi ^i(𝐤_m,𝐤_n)`$ in the Fourier-transformed normalized incoherent core emission function, $`\stackrel{~}{s}_c^i(𝐤_m,𝐤_n)`$ and another independent phase denoted by $`\varphi ^c(𝐤_m,𝐤_n)`$ is present in the the Fourier-transformed normalized coherent core emission function, $`\stackrel{~}{s}_c^p(𝐤_m,𝐤_n)`$. One can write $`\stackrel{~}{s}_c^i(𝐤_m,𝐤_n)`$ $`=`$ $`|\stackrel{~}{s}_c^i(𝐤_m,𝐤_n)|\mathrm{exp}[i\varphi ^i(𝐤_m,𝐤_n)],`$ (64) $`\stackrel{~}{s}_c^p(𝐤_m,𝐤_n)`$ $`=`$ $`|\stackrel{~}{s}_c^p(𝐤_m,𝐤_n)|\mathrm{exp}[i\varphi ^p(𝐤_m,𝐤_n)].`$ (65) The shape of both the coherent and the incoherent components is arbitrary, but corresponds to the space-time distribution of particle production. If the variances of the core are finite, the emission functions can be parameterized by Gaussians, for the sake of simplicity . If the core distributions have power-law-like tails, like in case of the Lorentzian distribution , then the Fourier-transformed emission functions correspond to exponentials or to power-law structures. For completeness, we list these possibilities below: $`|\stackrel{~}{s}_c^i(𝐤_m,𝐤_n)|^2`$ $`=`$ $`\mathrm{exp}(R_i^2Q_{mn}^2)\text{or}`$ (66) $`|\stackrel{~}{s}_c^i(𝐤_m,𝐤_n)|^2`$ $`=`$ $`\mathrm{exp}(R_iQ_{mn})\text{or}`$ (67) $`|\stackrel{~}{s}_c^i(𝐤_m,𝐤_n)|^2`$ $`=`$ $`a_i(R_iQ_{mn})^{b_i}\text{etc … ,}`$ (68) $`|\stackrel{~}{s}_c^p(𝐤_m,𝐤_n)|^2`$ $`=`$ $`\mathrm{exp}(R_p^2Q_{ij}^2)\text{or}`$ (69) $`|\stackrel{~}{s}_c^p(𝐤_m,𝐤_n)|^2`$ $`=`$ $`\mathrm{exp}(R_pQ_{mn})\text{or}`$ (70) $`|\stackrel{~}{s}_c^p(𝐤_m,𝐤_n)|^2`$ $`=`$ $`a_p(R_pQ_{mn})^{b_p}\text{etc … }.`$ (71) In the above equations, subscripts $`i`$ and $`p`$ index the parameters belonging to the incoherent or to the partially coherent components of the core, and $`Q_{mn}`$ stands for certain experimentally defined relative momentum component determined from $`𝐤_m`$ and $`𝐤_n`$. A straightforward counting yields that in the limiting case when all momenta are equal, the simple formula of Eq. (53) follows from the shape of the $`n`$-particle Bose–Einstein correlation functions of Eq. (63), as $`\stackrel{~}{s}_c^i(i,i)=\stackrel{~}{s}_c^p(i,i)=1`$. ### 5.3 Graph rules Graph rules were derived for the evaluation of the $`n`$-particle correlation function $`C_n(𝐤_1,\mathrm{},𝐤_n)`$ in Ref. . Graphs contributing to the $`n`$ = 2 and 3 case are shown in Fig. 5, the case of $`n=4`$ is shown in Fig. 6. Circles can be either open or full. Each circle carries one label (e.g. $`j`$) standing for a particle with momentum $`𝐤_j`$. Full circles represent the incoherent core component by a factor $`f_c(j)[1p_c(j)]`$, whereas open circles correspond to the coherent component of the core, a factor of $`f_c(j)p_c(j)`$. For the $`n`$-particle correlation function, all possible $`j`$-tuples of particles have to be found. Such $`j`$-tuples can be chosen in $`({}_{j}{}^{n})`$ different manner. In a $`j`$-tuple, either each circle is filled, or the circle with index $`k`$ is open and the other $`j1`$ circle is filled, which gives $`j+1`$ different possibilities. All the permutations that fully mix either $`j=2`$ or 3, …, or $`n`$ different elements have to be taken into account for each choice of filling the circles. The number of different fully mixing permutations that permute the elements $`i_1,\mathrm{},i_j`$ is given by $`\alpha _j`$ and can be determined from the recurrence of Eq. (54). Lines, that connect a pair of circles (or vertexes) $`(i,j)`$ stand for factors that depend both on $`𝐤_i`$ and $`𝐤_j`$. Full lines represent incoherent–incoherent particle pairs, and corresponds to a factor of $`\stackrel{~}{s}_c^i(i,j)`$. Dashed lines correspond to incoherent-coherent pairs, and carry a factor of $`\stackrel{~}{s}_c^p(i,j)`$. The lines are oriented, they point from circle $`i`$ to circle $`j`$, corresponding to the given permutation, that replaces element $`j`$ by element $`i`$. Dashed lines start from an open circle and point to a full circle. All graphs contribute to the order $`n`$ correlation function, that are in agreement with the above rules. The result corresponds to the fully mixing permutations of all possible $`j`$-tuples $`(j=2,\mathrm{},n)`$ chosen in all possible manner from elements $`(1,2,\mathrm{},n)`$. Each graph adds one term to the correlation function, given by the product of all the factors represented by the circles and lines of the graph. Note that the directions of the arrows matter. The correlation function $`C(1,\mathrm{},n)`$ is given by 1 plus the sum of all the graphs. Note that for the $`n`$-particle cumulant correlation function, $`n`$ circles, representing the $`n`$ particles, should be connected in all possible manner corresponding only to the fully mixing permutations of elements $`(1,\mathrm{},n)`$. Disconnected graphs do not contribute to the cumulant correlation functions, as they correspond to permutations, that either do not mix all of the $`n`$ elements or can be built up from two or more independent permutations of certain sub-samples of elements $`(1,2,\mathrm{},n)`$. ### 5.4 Application to three-particle correlation data In the CERN SPS S + Pb reactions, the strength of the two- and three-particle correlation functions was determined experimentally by the NA44 Collaboration as $`\lambda _{,2}=0.44\pm 0.04`$ in Ref. and by $`\lambda _{,3}=1.35\pm 0.12\pm 0.09`$, Ref. . Note that the value of $`\lambda _{,3}`$ was determined with the help of the Coulomb 3-particle wave-function integration method of Ref. , reviewed in Section 4, because the estimatebased only on the 3-body Gamow penetration factor introduced unacceptably large systematic errors to the three-particle Bose–Einstein correlation function. The two experimental values, $`\lambda _{,2}`$ and $`\lambda _{,3}`$ can be fitted with the two theoretical parameters $`f_c`$ and $`p_c`$, as done in Ref. . Figure 7 illustrates the 2 $`\sigma `$ contour plots in the $`(f_c,p_c)`$ plane, obtained using the published value of $`\lambda _{,2}=0.44\pm 0.04`$ and the preliminary value of $`\lambda _{,3}=1.35\pm 0.12`$. A range of $`(f_c,p_c)`$ values is found to describe simultaneously the strength of the two-particle and the three-particle correlation functions within two standard deviations from these values. Thus neither the fully chaotic, nor the partially coherent source picture can be excluded at this level of precision. Cramer and Kadija pointed out, that for higher values of $`n`$ the difference between a partially coherent source and between the fully incoherent particle source with an unresolvable component (halo or mis-identified particles) will become larger and larger . Indeed, similar values can be obtained for the strength of the second and third order correlation function, if the source is assumed to be fully incoherent $`(f_c=0.6,p_c=0)`$ or if the source has no halo but a partially coherent component $`(f_c=1,p_c=0.75)`$, but the strength of the 5th order correlation function is almost a factor of 2 larger in the former case, as can be seen from Table 2. Precision measurements of 4th and 5th order correlations are necessary to determine the value of the degree of partial coherence in the pion source. ## 6 Particle Interferometry in e$`{}_{}{}^{+}+`$ e<sup>-</sup> Reactions The hadronic production in e<sup>+</sup>e<sup>-</sup> annihilations is usually considered to be a basically coherent process and therefore no Bose–Einstein effect was expected, whereas hadronic reactions should be of a more chaotic nature giving rise to a sizable effect. It was even argued that the strong ordering in rapidity, preventing neighboring $`\pi ^{}\pi ^{}`$ or $`\pi ^+\pi ^+`$ pairs, would drastically reduce the effect . Therefore it was a surprise when G. Goldhaber at the Lisbon Conference in 1981 presented data which showed that correlations between identical particles in e<sup>+</sup>e<sup>-</sup> annihilations were very similar in size and shape to those seen in hadronic reactions, see the review paper Ref. for further details. ### 6.1 The Andersson–Hofmann model The Bose–Einstein correlation effect, a priori unexpected for a coherent process, has been given an explanation within the Lund string model by B. Andersson and W. Hofmann . The space-time structure of an e<sup>+</sup>e<sup>-</sup> annihilation is shown for the Lund string model in Fig. 8. The probability for a particular final state is given by the expression $$\mathrm{Prob}.\mathrm{phasespace}\mathrm{exp}(bA),$$ (72) where $`A`$ is the space-time area spanned by the string before it breaks and $`b`$ is a parameter. The classical string action is given by $`S=\kappa A`$, where $`\kappa `$ is the string tension. It is natural to interpret the result in Eq. (72) as resulting from an imaginary part of the action such that $$S=(\kappa +ib/2)A,$$ (73) and an amplitude $`M`$ given by $$M\mathrm{exp}(iS),$$ (74) which implies $$\mathrm{Prob}.M^2\mathrm{exp}(bA).$$ (75) Final states with two identical particles are indistinguishable and can be obtained in different ways. Suppose that the two particles indicated as 1 and 2 in Fig. 8 are identical, then the hadron state in the left panel can be considered as being the same as that in the right panel (where 1 and 2 are interchanged). The amplitude should, for bosons, be the sum of two terms $$M\mathrm{exp}[i(\kappa +ib/2)A_1]+\mathrm{exp}[i(\kappa +ib/2)A_2],$$ (76) where $`A_1`$ and $`A_2`$ are the two string areas, giving a probability proportional to $$M^2[\mathrm{exp}(bA_1)+\mathrm{exp}(bA_2)]\left[1+\frac{\mathrm{cos}(\kappa \mathrm{\Delta }A)}{\mathrm{cosh}(b\mathrm{\Delta }A/2)}\right]$$ (77) with $`\mathrm{\Delta }AA_1A_2`$. The magnitudes of $`\kappa `$ and $`b`$ are known from phenomenological studies. The energy per unit length of the string is given by $`\kappa 1`$ GeV/fm, and $`b`$ describes the breaking of the string at a constant rate per unit area, $`b/\kappa ^20.7`$ GeV<sup>-2</sup> . The difference in space-time area $`\mathrm{\Delta }A`$ is marked as the hatched area in Fig. 8. It can be expressed by the $`(t,r_z)`$ components $`(E,k)`$ of the four-momenta of the two identical particles 1 and 2, and the intermediate system $`I`$: $$\mathrm{\Delta }A=[E_2k_1E_1k_2+E_I(k_1k_2)k_I(E_1E_2)]/\kappa ^2.$$ (78) To take into account also the component transverse to the string a small additional term is needed. The change in area $`\mathrm{\Delta }A`$ is Lorentz invariant to boosts along the string direction and is furthermore approximately proportional to $`Q=\sqrt{(k_1k_2)^2}`$. The interference pattern between the amplitudes will be dominated by the phase change of $`\mathrm{\Delta }\mathrm{\Phi }=\kappa \mathrm{\Delta }A`$. It leads to a Bose–Einstein correlation which, as a function of the four-momentum transfer, reproduces the data well but shows a steeper dependence at small $`Q`$ than a Gaussian function. A comparison to TPC data confirmed the existence of such a steeper than Gaussian dependence on $`Q`$, although the statistics at the small $`Q`$-values did not allow a firm conclusion . Recently, the interest for multi-dimensional analysis of Bose–Einstein correlations increased also in the particle physics community, see Ref. for a critical review of the present status. I would like to highlight three interesting features: i) The effect seems to depend on the transverse momentum of the produced pion pairs, decreasing effective radii were observed for increasing transverse mass . This effect is also seen in the LUBOEI algorithm of JETSET, although no intrinsic momentum-dependent scale is plugged into the algorithm . ii) The three-dimensional Bose–Einstein correlations of L3 indicate a non-Gaussian structure . iii) The effective source sizes of heavier particles (K, $`\mathrm{\Lambda }`$) were measured recently , based on spin statistics developed by Alexander and Lipkin . The measured source sizes show a clear decrease with increasing particle masses. The latter effect was explained by Alexander, Cohen and Levin by arguments based on the Heisenberg uncertainty relation, and independently with the help of virial theorem applied for a QCD motivated confining potential. See Fig. 9, reproduced from Ref. . Note that a similar decrease was predicted in Ref. , which would depend not on the mass, but on the transverse mass of the particles, if the particle production happens so that the position of the emission is very strongly correlated with the momentum of the emitted particle . So, it would be timely to check whether the effect depends on the particle mass, or on the transverse mass. Although the side radius components indicate such a decrease in case of pions, similar measurements for kaons and $`\mathrm{\Lambda }`$-s would be indispensable to clarify the origin of the observed behavior. The question arises: can the effects i)–iii) be explained in a unified framework, that characterizes the hadronization process in e<sup>+</sup>e<sup>-</sup> annihilation? An explanation of the rather small effective size of the source of the $`\mathrm{\Lambda }`$-s seems to be a challenge for the Lund string model. The three-dimensional analysis of the NA22 data on h + p reactions indicated a strong decrease of all the characteristic radii with increasing values of transverse momenta of the pair in the NA22 experiment . A decrease of the effective source sizes with increasing values of the transverse mass for a given kind of particle is seen in heavy ion collisions, similarly to effect i) in particle physics. The property iii), the decrease of the effective source size with the increase of the mass of the particle is seen in heavy ion physics and is explained in terms of hydrodynamical expansion, similarly to the explanation of effect i), see Figs 15 and 16 in Section 12. Can one give a unified explanation of these similarities between results of particle interferometry in e$`{}_{}{}^{+}+`$ e<sup>-</sup>, h + p and heavy ion physics? We do not yet know the answer to this question. ## 7 Invariant Buda–Lund Particle Interferometry The $`n`$-particle Bose–Einstein correlation function of Eq. (5) is defined as the ratio of the $`n`$-particle invariant momentum momentum distribution divided by an $`n`$-fold product of the single-particle invariant momentum distributions. Hence these correlation functions are boost-invariant. The invariant Buda–Lund parameterization (or BL in short) deals with a boost-invariant, multi-dimensional characterization of the building blocks $`a_{𝐤_1}^{}a_{𝐤_2}`$ of arbitrary high order Bose–Einstein correlation functions, based on Eqs (8,13). The BL parameterization was developed by the Buda–Lund Collaboration in Refs . The essential part of the BL is an invariant decomposition of the relative momentum $`q`$ in the $`\mathrm{exp}(iq\mathrm{\Delta }x)`$ factor into a temporal, a longitudinal and two transverse relative momentum components. This decomposition is obtained with the help of a time-like vector in the coordinate space, that characterizes the center of particle emission in space-time, see Fig. 10. Although the BL parameterization was introduced in Ref. for high energy heavy ion reactions, it can be used for other physical situations as well, where a dominant direction of an approximate boost-invariant expansion of the particle emitting source can be identified and taken as the longitudinal direction $`r_z`$. For example, such a direction is the thrust axis of single jets or of back-to-back two-jet events in case of high energy particle physics. For longitudinally almost boost-invariant systems, it is advantageous to introduce the boost invariant variable $`\tau `$ and the space-time rapidity $`\eta `$, $`\tau `$ $`=`$ $`\sqrt{t^2r_z^2},`$ (79) $`\eta `$ $`=`$ $`0.5\mathrm{log}\left[(t+r_z)/(tr_z)\right].`$ (80) Similarly, in momentum space one introduces the transverse mass $`m_t`$ and the rapidity $`y`$ as $`m_t`$ $`=`$ $`\sqrt{E^2p_z^2},`$ (81) $`y`$ $`=`$ $`0.5\mathrm{log}\left[(E+p_z)/(Ep_z)\right].`$ (82) The source of particles is characterized in the boost invariant variables $`\tau `$, $`m_t`$ and $`\eta y`$. For systems that are only approximately boost-invariant, the emission function may also depend on the deviation from mid-rapidity, $`y_0`$. The scale on which the approximate boost-invariance breaks down is denoted by $`\mathrm{\Delta }\eta `$, a parameter that is related to the width of the rapidity distribution. The correlation function is defined with the help of the Wigner function formalism, Eq. (13), the intercept parameter $`\lambda _{}`$ is introduced in the core/halo picture of Eq. (42). The case of $`n=2`$ particles and a chaotic core with $`p_c=0`$ was discussed in Ref. . In the following, we evaluate the building block for arbitrary high order Bose–Einstein correlation functions. We assume for simplicity that the core is fully incoherent, $`p_c(j)=0`$ in Eq. (63). A further simplification is obtained if we assume that the emission function of Eqs (13,42) factorizes as a product of an effective proper-time distribution, a space-time rapidity distribution and a transverse coordinate distribution : $`S_c(x,K)d^4x`$ $`=`$ $`H_{}(\tau )G_{}(\eta )I_{}(r_x,r_y)d\tau \overline{\tau }d\eta dr_xdr_y.`$ (83) The subscript $``$ stands for a dependence on the mean momentum $`K`$, the mid-rapidity $`y_0`$ and the scale of violation of boost-invariance $`\mathrm{\Delta }\eta `$, using the symbolic notation $`f_{}f[K,y_0,\mathrm{\Delta }\eta ]`$. The function $`H_{}(\tau )`$ stands for such an effective proper-time distribution (that includes, by definition, an extra factor $`\tau `$ from the Jacobian $`d^4x=d\tau \tau d\eta ,dr_xdr_y`$, in order to relate the two-particle Bose–Einstein correlation function to a Fourier transformation of a distribution function in $`\tau `$). The effective space-time rapidity distribution is denoted by $`G_{}(\eta )`$, while the effective transverse distribution is denoted by $`I_{}(r_x,r_y)`$. In Eq. (83), the mean value of the proper-time $`\overline{\tau }`$ is factored out, to keep the distribution functions dimensionless. Such a pattern of particle production is visualized in Fig. 10. In case of hydrodynamical models, as well as in case of a decaying Lund strings , production of particles with a given momentum rapidity $`y`$ is limited to a narrow region in space-time around $`\overline{\eta }`$ and $`\overline{\tau }`$. If the sizes of the effective source are sufficiently small (if the Bose–Einstein correlation function is sufficiently broad), the $`\mathrm{exp}(iq\mathrm{\Delta }x)`$ factor of the Fourier transformation is decomposed in the shaded region in Fig. 10 as $`\mathrm{exp}[i(q^0\mathrm{\Delta }tq_z\mathrm{\Delta }r_z)]`$ $``$ $`\mathrm{exp}[i(Q_=\mathrm{\Delta }\tau Q_{}\overline{\tau }\mathrm{\Delta }\eta )],`$ (84) $`\mathrm{exp}[i(q_x\mathrm{\Delta }r_x+iq_y\mathrm{\Delta }r_y)]`$ $``$ $`\mathrm{exp}[i(Q_:\mathrm{\Delta }r_:+Q_{..}\mathrm{\Delta }r_{..})].`$ (85) The invariant temporal, parallel, sideward, outward (and perpendicular) relative momentum components are defined, respectively, as $`Q_=`$ $`=`$ $`q_0\mathrm{cosh}[\overline{\eta }]q_z\mathrm{sinh}[\overline{\eta }],`$ (86) $`Q_{}`$ $`=`$ $`q_z\mathrm{cosh}[\overline{\eta }]q_0\mathrm{sinh}[\overline{\eta }],`$ (87) $`Q_{..}`$ $`=`$ $`(q_xK_yq_yK_x)/\sqrt{K_x^2+K_y^2},`$ (88) $`Q_:`$ $`=`$ $`(q_xK_x+q_yK_y)/\sqrt{K_x^2+K_y^2},`$ (89) $`Q_{}`$ $`=`$ $`\sqrt{q_x^2+q_y^2}=\sqrt{Q_:^2+Q_{..}^2}.`$ (90) The time-like normal-vector $`\overline{n}`$ indicates an invariant direction of the source in coordinate space . It is parameterized as $`\overline{n}^\mu =(\mathrm{cosh}[\overline{\eta }],0,0,\mathrm{sinh}[\overline{\eta }])`$, where $`\overline{\eta }`$ is a mean space-time rapidity . The parameter $`\overline{\eta }`$ is one of the fitted parameters in the BL type of decomposition of the relative momenta. The above equations are invariant, they can be evaluated in any frame. To simplify the presentation, in the following we evaluate $`q`$ and $`\overline{\eta }`$ in the LCMS. The acronym LCMS stands for the Longitudinal Center of Mass System, where the mean momentum of a particle pair has vanishing longitudinal component, $`K_z=0.5(k_{1,z}+k_{2,z})=0.`$ In this frame, introduced in Ref. , $`𝐊`$ is orthogonal to the beam axis, and the time-like information on the duration of the particle emission couples to the out direction. The rapidity of the LCMS frame can be easily found from the measurement of the momentum vectors of the particles. As $`\overline{\eta }`$ is from now on a space-time rapidity measured in the LCMS frame, it is invariant to longitudinal boosts: $`\overline{\eta }^{}=(\overline{\eta }y)(0y)=\overline{\eta }`$. The symbolic notation for the side direction is two dots side by side as in $`Q_{..}`$. The remaining transverse direction, the out direction was indexed as in $`Q_:`$, in an attempt to help to distinguish the zeroth component of the relative momentum $`Q_0`$ from the out component of the relative momentum $`Q_:Q_o=Q_{\mathrm{out}}`$, $`Q_0Q_o`$. Hence $`K_:=|𝐊_{}|`$ and $`K_{..}=0`$. The geometrical idea behind this notation is explained in details in Ref. . The perpendicular (or transverse) component of the relative momentum is denoted by $`Q_{}`$. By definition, $`Q_{..}`$, $`Q_:`$ and $`Q_{}`$ are invariants to longitudinal boosts, and $`Q^2=qq=Q_{..}^2+Q_:^2+Q_{||}^2Q_=^2`$. With the help of the small source size (or large relative momentum) expansion of Eq. (84), the amplitude $`\stackrel{~}{s}_c(1,2)=\stackrel{~}{s}_c^i(1,2)`$ that determines the arbitrary order Bose–Einstein correlation functions in Eq. (63) can be written as follows: $$\stackrel{~}{s}_c^i(1,2)=\frac{\stackrel{~}{H}_{}(Q_=)\stackrel{~}{G}_{}(Q_{})\stackrel{~}{I}_{}(Q_:,Q_{..})}{\stackrel{~}{H}_{}(0)\stackrel{~}{G}_{}(0)\stackrel{~}{I}_{}(0,0)}.$$ (91) This expression and Eq. (63) yield a general, invariant, multi-dimensional Buda–Lund parameterization of order $`n`$ Bose–Einstein correlation functions, valid for all $`n`$. The Fourier-transformed distributions are defined as $`\stackrel{~}{H}_{}(Q_=)`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}𝑑\tau \mathrm{exp}(iQ_=\tau )H_{}(\tau ),`$ (92) $`\stackrel{~}{G}_{}(Q_{})`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\eta \mathrm{exp}(iQ_{}\overline{\tau }\eta )G_{}(\eta ),`$ (93) $`\stackrel{~}{I}_{}(Q_:,Q_{..})`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑r_:{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑r_{..}\mathrm{exp}(iQ_:r_:iQ_{..}r_{..})I_{}(r_:,r_{..}).`$ (94) As a particular case of Eqs (91,63) for $`n=2`$ and $`p_c(j)=0`$, the two-particle BECF can be written into a factorized Buda–Lund form as $`C(𝐤_1,𝐤_2)`$ $`=`$ $`1+\lambda _{}(K){\displaystyle \frac{|\stackrel{~}{H}_{}(Q_=)|^2}{|\stackrel{~}{H}_{}(0)|^2}}{\displaystyle \frac{|\stackrel{~}{G}_{}(Q_{})|^2}{|\stackrel{~}{G}_{}(0)|^2}}{\displaystyle \frac{|\stackrel{~}{I}_{}(Q_:,Q_{..})|^2}{|\stackrel{~}{I}_{}(0,0)|^2}}.`$ (95) Thus, the BL results are rather generic. For example, BL parameterization may in particular limiting cases yield the power-law, the exponential, the double-Gaussian, the Gaussian, or the less familiar oscillating forms of Eq. (128), see also Ref. . The Edgeworth, the Laguerre or other similarly constructed low-momentum expansions can be applied to any of the factors of one variable in Eq. (95) to characterize these unknown shapes in a really model-independent manner, relying only on the convergence properties of expansions in terms of complete orthonormal sets of functions . In a Gaussian approximation and assuming that $`R_:=R_{..}=R_{}`$, the Buda–Lund form of the Bose–Einstein correlation function reads as follows: $$C_2(𝐤_1,𝐤_2)=1+\lambda _{}\mathrm{exp}\left(R_=^2Q_=^2R_{}^2Q_{}^2R_{}^2Q_{}^2\right),$$ (96) where the 5 fit parameters are $`\lambda _{}`$, $`R_=`$, $`R_{}`$, $`R_{}`$ and the value of $`\overline{\eta }`$ that enters the definitions of $`Q_=`$ and $`Q_{}`$ in Eqs (86,87). The fit parameter $`R_=`$ reads as $`R`$-time-like, and this variable measures a width of the proper-time distribution $`H_{}`$. The fit parameter $`R_{}`$ reads as $`R`$-parallel, it measures an invariant length parallel to the direction of the expansion. The fit parameter $`R_{}`$ reads as $`R`$-perpendicular or $`R`$-perp. For cylindrically symmetric sources, $`R_{}`$ measures a transversal rms radius of the particle emitting source. The BL radius parameters characterize the lengths of homogeneity in a longitudinally boost-invariant manner. The lengths of homogeneity are generally smaller than the momentum-integrated, total extension of the source, they measure a region in space and time, where particle pairs with a given mean momentum $`𝐊`$ are emitted from. The following Edgeworth expansion can be utilized to characterize non-Gaussian multidimensional Bose–Einstein correlation functions, in a longitudinally boost-invariant manner: $`C_2(𝐤_1,𝐤_2)`$ $`=`$ $`1+\lambda _E\mathrm{exp}(Q_=^2R_=^2Q_{||}^2R_{||}^2Q_{}^2R_{}^2)\times `$ (97) $`[1+{\displaystyle \frac{\kappa _{3=}}{3!}}H_3(\sqrt{2}Q_=R_=)+{\displaystyle \frac{\kappa _{4=}}{4!}}H_4(\sqrt{2}Q_=R_=)+\mathrm{}]\times `$ $`[1+{\displaystyle \frac{\kappa _{3||}}{3!}}H_3(\sqrt{2}Q_{||}R_{||})+{\displaystyle \frac{\kappa _{4||}}{4!}}H_4(\sqrt{2}Q_{||}R_{||})+\mathrm{}]\times `$ $`\left[1+{\displaystyle \frac{\kappa _3}{3!}}H_3(\sqrt{2}Q_{}R_{})+{\displaystyle \frac{\kappa _4}{4!}}H_4(\sqrt{2}Q_{}R_{})+\mathrm{}\right].`$ This yields $`5`$ free scale parameters for cylindrically symmetric, longitudinally expanding sources, and three series of shape parameters. The scale parameters are $`\lambda _E`$, $`R_=`$, $`R_{}`$, $`R_{}`$ and $`\overline{\eta }`$, that characterize the effective source at a given mean momentum, by giving the vertical scale of the correlations, the invariant temporal, longitudinal and transverse extensions of the source and its invariant direction, which is the space-time rapidity of the effective source in the LCMS frame (the frame where $`k_{1,z}+k_{2,z}=0`$, ). The three series of shape parameters are $`\kappa _{3=}`$,$`\kappa _{4=}`$, … , $`\kappa _{3||}`$, $`\kappa _{4||}`$, … , $`\kappa _3`$, $`\kappa _4`$, … . Each of these parameters may depend on the mean momentum $`𝐊`$. A multi-dimensional Laguerre or a mixed Edgeworth–Laguerre expansion can be introduced in a similar manner by replacing the Edgeworth expansion in Eq. (97) by a Laguerre one in any of the principal directions. In Eqs (97,96), the spatial information about the source distribution in $`(r_x,r_y)`$ was combined to a single perp radius parameter $`R_{}`$. In a more general Gaussian form, suitable for studying rings of fire and opacity effects, the Buda–Lund invariant BECF can be denoted as $$C_2(𝐤_1,𝐤_2)=1+\lambda _{}\mathrm{exp}\left(R_=^2Q_=^2R_{}^2Q_{}^2R_{..}^2Q_{..}^2R_:^2Q_:^2\right).$$ (98) The 6 fit parameters are $`\lambda _{}`$, $`R_=`$, $`R_{}`$, $`R_{..}`$, $`R_:`$ and $`\overline{\eta }`$, all are in principle functions of $`(𝐊,y_0,\mathrm{\Delta }\eta )`$. Note that this equation is identical to Eq. (44) of Ref. , rewritten into the new, symbolic notation of the Lorentz-invariant directional decomposition. The above equation may be relevant for a study of expanding shells, or rings of fire, as discussed first in Ref. . We shall argue, based on a simultaneous analysis of particle spectra and correlations, and on recently found exact solutions of non-relativistic fireball hydrodynamics that an expanding, spherical shell of fire is formed protons in 30 AMeV <sup>40</sup>Ar + <sup>197</sup>Au reactions, and that a two-dimensional, expanding ring of fire is formed in the transverse plane in NA22 h + p reactions at CERN SPS. The experimental signatures for the formation of these patterns will be discussed in Section 11. Opacity effects, as suggested recently by H. Heiselberg , also require the distinction between $`R_{..}`$ and $`R_:`$. The lack of transparency in the source may result in an effective source function, that looks like a crescent in the side-out reference frame . When integrated over the direction of the mean momentum, the effective source looks like a ring of fire in the $`(r_x,r_y)`$ frame. The price of the invariant decomposition of the basic building blocks of any order Bose–Einstein correlation functions in the BL parameterization is that the correlation functions cannot be directly binned in the BL variables, as these can determined after the parameter $`\overline{\eta }`$ is fitted to the data — so the correlation function has to be binned first in some directly measurable relative momentum components, e.g. the (side, out, long) relative momenta in the LCMS frame, as discussed in the next subsection. After fitting $`\overline{\eta }`$ in an arbitrary frame, the BECF can be re-binned into the BL form. ### 7.1 Gaussian parameterizations of BE Correlations We briefly summarize here the Bertsch–Pratt and the Yano–Koonin parameterization of the Bose–Einstein correlation functions, to point out some of their advantages as well as drawbacks and to form a basis for comparison. #### 7.1.1 The Bertsch–Pratt parameterization The Bertsch–Pratt (BP) parameterization of Bose–Einstein correlation functions is one of the oldest, widely used multi-dimensional decomposition, called also as the side–out–longitudinal decomposition . This directional decomposition was devised to extract the contribution of a long duration of particle emission from an evaporating Quark–Gluon Plasma, as expected in the mixture of a hadronic and a QGP phase if the re-hadronization phase transition is a strong first order transition. The BP parameterization in a compact form reads as $`C_2(𝐤_1,𝐤_2)`$ $`=`$ $`1+\lambda \mathrm{exp}\left[R_s^2Q_s^2R_o^2Q_o^2R_l^2Q_l^22R_{ol}^2Q_lQ_o\right].`$ (99) Here index $`o`$ stands for out (and not the temporal direction), $`s`$ for side and $`l`$ for longitudinal. The out–longitudinal cross-term was introduced by Chapman, Scotto and Heinz in Refs — this term is non-vanishing for axially symmetric systems, if the source is not fully boost-invariant, or if the measurement is made not at mid-rapidity. In a more detailed form, the mean momentum dependence of the various components is shown as $`C_2(𝐤_1,𝐤_2)`$ $`=`$ $`1+\lambda (𝐊)\mathrm{exp}[R_s^2(𝐊)Q_s^2(𝐊)R_o^2(𝐊)Q_o^2(𝐊)`$ (100) $`R_l^2(𝐊)Q_l^22R_{ol}^2(𝐊)Q_lQ_o(𝐊)],`$ where the mean and the relative momenta are defined as $`𝐊`$ $`=`$ $`0.5(𝐤_1+𝐤_2),`$ (101) $`𝚫𝐤`$ $`=`$ $`𝐤_1𝐤_2,`$ (102) $`Q_l`$ $`=`$ $`k_{z,1}k_{z,2},`$ (103) $`Q_o`$ $`=`$ $`Q_o(𝐊)=𝚫𝐤𝐊/|𝐊|,`$ (104) $`Q_s`$ $`=`$ $`Q_s(𝐊)=|𝚫𝐤\times 𝐊|/|𝐊|.`$ (105) It is emphasized that the BP radius parameters are also measuring lengths of homogeneity . Not only the radius parameters but also the decomposition of the relative momentum to the side and the out components depends on the (direction of) mean momentum $`𝐊`$. In an arbitrary frame, Gaussian radius parameters can be defined, and sometimes they are also referred to as BP radii, when the spatial components of the relative momentum vector are taken as independent variables. The BP radii reflect space-time variances of the core of the particle emission, if a Gaussian approximation to the core is warranted: $`C_2(𝐤_1,𝐤_2)`$ $`=`$ $`1+\lambda _{}(𝐊)\mathrm{exp}\left(R_{i,j}^2(𝐊)𝚫𝐤_i𝚫𝐤_j\right),`$ (106) $`\lambda _{}(𝐊)`$ $`=`$ $`[N_𝐜(𝐊)/N(𝐊)]^2,`$ (107) $`R_{i,j}^2(𝐊)`$ $`=`$ $`\underset{¯}{x}_i\underset{¯}{x}_j_𝐜\underset{¯}{x}_i_𝐜\underset{¯}{x}_j_𝐜,`$ (108) $`\underset{¯}{x}_i`$ $`=`$ $`x_i\beta _it,`$ (109) $`f(x,𝐤)_𝐜`$ $`=`$ $`{\displaystyle d^4xf(x,𝐤)S_𝐜(x,𝐤)/d^4xS_𝐜(x,𝐤)},`$ (110) where $`S_𝐜(x,𝐤)`$ is the emission function that characterizes the central core and subscripts $`i`$ or $`j`$ stand for $`x`$, $`y`$ or $`z`$, i.e. any of the spatial directions in the frame of the analysis. This method is frequently called as “model-independent” formulation, because the applied Gaussian approximation is independent of the functional form of the emission function $`S_(x,𝐤)`$ . In the literature, this result is often over-stated, it is claimed that such a Taylor expansion would provide a general “proof” that multi-dimensional Bose–Einstein correlation functions must be Gaussians. Although the “proof” is indeed not depending on the exact shape of $`S(x,K)`$, it relies on a second order Taylor expansion of the shape of the correlation function around its exact value at $`Q=0`$. At this point not only the derivatives of the correlation function are unmeasurable, but the very value of the correlation function $`C_2(0)`$ is unmeasurable as well, see Figs 4 and 3 for graphical illustration. For exponential or for power-law type correlations, the building block $`\stackrel{~}{S}_c(q,K)`$ of the correlation function is not analytic at $`Q=0`$, so a Taylor expansion cannot be applied in their case. For the oscillatory type of correlation functions, the Gaussian provides a good approximation in the experimentally unresolvable low $`Q`$ domain, but it misses the structure of oscillations at large values of $`Q`$, which appear because $`S(x,K)`$ has more than one maxima, like a source distribution of a binary star. Thus, the exact shapes of multi-dimensional BECF-s cannot be determined a priori and in case of non-Gaussian correlators one has to evaluate more (but still not fully) model-independent relationships, for example Eqs (13,63,91), which are valid for broader than Gaussian classes of correlation functions. Note that the tails of the emission function are typically dominated by the halo of long-lived resonances $`S_h(x,𝐤)`$ and even a small admixture of e.g. $`\eta `$ and $`\eta ^{}`$ mesons increases drastically the space-time variances of particle production, and makes the interpretation of the BP radii in terms of space-time variances of the total emission function $`S=S_c+S_h`$ unreliable both qualitatively and quantitatively, as pointed out already in Ref. . In the Longitudinal Center of Mass System (LCMS, Ref. ), the BP radii have a particularly simple form , if the coupling between the $`r_x`$ and the $`t`$ coordinates is also negligible, $`\stackrel{~}{r}_x\stackrel{~}{t}=\stackrel{~}{r}_x\stackrel{~}{t}`$: $`R_s^2(𝐊)`$ $`=`$ $`\stackrel{~}{r}_y^2_𝐜,`$ (111) $`R_o^2(𝐊)`$ $`=`$ $`\stackrel{~}{r}_x^2_𝐜+\beta _t^2\stackrel{~}{t}^2_𝐜,`$ (112) $`R_l^2(𝐊)`$ $`=`$ $`\stackrel{~}{r}_z^2_𝐜,`$ (113) $`R_{ol}^2(𝐊)`$ $`=`$ $`\stackrel{~}{r}_z(\stackrel{~}{r}_x\beta _t\stackrel{~}{t})_𝐜,`$ (114) where $`\stackrel{~}{x}=xx`$. Although this method cannot be applied to characterize non-Gaussian correlation functions, the the above form has a number of advantages:it is straightforward to obtain and it is easy to implement for a numerical evaluation of the BP radii of Gaussian correlation functions . In the LCMS frame, information on the duration of the particle emission couples only to the out direction. This is one of the advantages of the LCMS frame. Using the BP, the time distribution enters the out radius component as well as the out-long cross-term. Other possible cross-terms were shown to vanish for cylindrically symmetric sources . For completeness, we give the relationship between the invariant BL radii and the BP radii measured in the LCMS, if the BL forms are given in the Gaussian approximation of Eq. (98): $`R_s^2`$ $`=`$ $`R_{..}^2,`$ (115) $`R_o^2`$ $`=`$ $`R_:^2+\beta _t^2[\mathrm{cosh}^2(\overline{\eta })R_=^2+\mathrm{sinh}^2(\overline{\eta })R_{}^2],`$ (116) $`R_{ol}^2`$ $`=`$ $`\beta _t\mathrm{sinh}(\overline{\eta })\mathrm{cosh}(\overline{\eta })(R_=^2+R_{}^2),`$ (117) $`R_l^2`$ $`=`$ $`\mathrm{cosh}^2(\overline{\eta })R_{}^2+\mathrm{sinh}^2(\overline{\eta })R_=^2,`$ (118) where the dependence of the fit parameters on the value of the mean momentum, $`𝐊`$ is suppressed. The advantage of the BP parameterization is that there are no kinematic constraints between the side, out and long components of the relative momenta, hence the BP radii are not too difficult to determine experimentally. A drawback is that the BP radii are not invariant, they depend on the frame where they are evaluated. The BP radii transform as a well-defined mixture of the invariant temporal, longitudinal and transverse BL radii, given e.g. in Ref. . #### 7.1.2 The Yano–Koonin–Podgoretskii parameterization A covariant parameterization of two-particle correlations has been worked out for non-expanding sources by Yano, Koonin and Podgoretskii (YKP) . This parameterization was recently applied to expanding sources by the Regensburg group , by allowing the YKP radius and velocity parameters be momentum dependent: $`C_2(𝐤_1,𝐤_2)`$ $`=`$ $`1+\mathrm{exp}[R_{}^2(𝐊)q_{}^2R_{}^2(𝐊)(q_z^2q_0^2)`$ (119) $`(R_0^2(𝐊)+R_{}^2(𝐊))(qU(𝐊))^2],`$ where the fit parameter $`U(𝐊)`$ is interpreted as a four-velocity of a fluid-element . (Note that in YKP index $`0`$ refers to the time-like components). This generalized YKP parameterization was introduced to create a diagonal Gaussian form in the “rest frame of a fluid-element”. This form has an advantage as compared to the BP parameterization: the three extracted YKP radius parameters, $`R_{}`$, $`R_{}`$ and $`R_0`$ are invariant, independent of the frame where the analysis is performed, while $`U^\mu `$ transforms as a four-vector. The price one has to pay for this advantage is that the kinematic region may become rather small in the $`q_0`$, $`q_l`$, $`q_{}`$ space, where the parameters are to be fitted, as follows from the inequalities $`Q^2=qq0`$ and $`q_0^20`$: $$0q_0^2q_z^2+q_{}^2,$$ (120) and the narrowing of the regions in $`q_0^2q_z^2`$ with decreasing $`q_{}`$ makes the experimental determination of the YKP parameters difficult, especially when the analysis is performed far from the LCMS rapidities \[or more precisely from the frame where $`U^\mu =(1,0,0,0)`$ \]. Theoretical problems with the YKP parameterization are explained as follows. a) The YKP radii contain components proportional to $`1/\beta _t`$, which lead to divergent terms for particles with very low $`p_t`$ . b) The YKP fit parameters are not even defined for all Gaussian sources . Especially, for opaque sources, for expanding shells, or for rings of fire with $`\stackrel{~}{r}_x^2<\stackrel{~}{r}_y^2`$ the algebraic relations defining the YKP “velocity” parameter become ill-defined and result in imaginary values of the YKP “velocity”, . c) The YKP “flow velocity” $`U^\mu (𝐊)`$ is defined in terms of space-time variances at fixed mean momentum of the particle pairs , corresponding to a weighted average of particle coordinates. In contrast, the local flow velocity $`u^\mu (x)`$ is defined as a local average of particle momenta. Hence, in general $`U^\mu (𝐊)u^\mu (x)`$, and the interpretation of the YKP parameter $`U^\mu (𝐊)`$ as a local flow velocity of a fluid does not correspond to the principles of kinetic theory. ## 8 Hydrodynamical Parameterization à la Buda–Lund (BL-H) The Buda–Lund hydro parameterization (BL-H) was invented in the same paper as the BL parameterization of the Bose–Einstein correlation functions , but in principle the general BL forms of the correlation function do not depend on the hydrodynamical ansatz (BL-H). The BL form of the correlation function can be evaluated for any, non-thermalized expanding sources, e.g. for the Lund string model also. The BL-H assumes, that the core emission function is characterized with a locally thermalized, volume-emitting source: $`S_c(x,𝐤)d^4x`$ $`=`$ $`{\displaystyle \frac{g}{(2\pi )^3}}{\displaystyle \frac{k^\mu d^4\mathrm{\Sigma }_\mu (x)}{\mathrm{exp}\left({\displaystyle \frac{u^\mu (x)k_\mu }{T(x)}}{\displaystyle \frac{\mu (x)}{T(x)}}\right)+s}}.`$ (121) The degeneracy factor is denoted by $`g`$, the four-velocity field is denoted by $`u^\mu (x)`$, the temperature field is denoted by $`T(x)`$, the chemical potential distribution by $`\mu (x)`$ and $`s=0`$, $`1`$ or $`1`$ for Boltzmann, Bose–Einstein or Fermi–Dirac statistics. The particle flux over the freeze-out layers is given by a generalized Cooper–Frye factor, assuming that the freeze-out hypersurface depends parametrically on the freeze-out time $`\tau `$ and that the probability to freeze-out at a certain value is proportional to $`H(\tau )`$, $`k^\mu d^4\mathrm{\Sigma }_\mu (x)`$ $`=`$ $`m_t\mathrm{cosh}[\eta y]H(\tau )d\tau \overline{\tau }d\eta dr_xdr_y.`$ (122) The four-velocity $`u^\mu (x)`$ of the expanding matter is assumed to be a scaling longitudinal Bjorken flow appended with a linear transverse flow, characterized by its mean value $`u_t`$, see Refs : $`u^\mu (x)`$ $`=`$ $`(\mathrm{cosh}[\eta ]\mathrm{cosh}[\eta _t],\mathrm{sinh}[\eta _t]{\displaystyle \frac{r_x}{r_t}},\mathrm{sinh}[\eta _t]{\displaystyle \frac{r_y}{r_t}},\mathrm{sinh}[\eta ]\mathrm{cosh}[\eta _t]),`$ $`\mathrm{sinh}[\eta _t]`$ $`=`$ $`u_tr_t/R_G,`$ (123) with $`r_t=[r_x^2+r_y^2]^{1/2}`$. Such a flow profile, with a time-dependent radius parameter $`R_G`$, was recently shown to be an exact solution of the equations of relativistic hydrodynamics of a perfect fluid at a vanishing speed of sound, Ref. . Instead of applying an exact hydrodynamical solution with evaporation terms, the BL-H characterizes the local temperature, flow and chemical potential distributions of a cylindrically symmetric, finite hydrodynamically expanding system with the means and the variances of these distributions. The hydrodynamical variables $`1/T(x)`$, $`\mu (x)/T(x)`$, are parameterized as $`{\displaystyle \frac{\mu (x)}{T(x)}}`$ $`=`$ $`{\displaystyle \frac{\mu _0}{T_0}}{\displaystyle \frac{r_x^2+r_y^2}{2R_G^2}}{\displaystyle \frac{(\eta y_0)^2}{2\mathrm{\Delta }\eta ^2}},`$ (124) $`{\displaystyle \frac{1}{T(x)}}`$ $`=`$ $`{\displaystyle \frac{1}{T_0}}\left(1+{\displaystyle \frac{\mathrm{\Delta }T}{T}}_r{\displaystyle \frac{r_t^2}{2R_G^2}}\right)\left(1+{\displaystyle \frac{\mathrm{\Delta }T}{T}}_t{\displaystyle \frac{(\tau \overline{\tau })^2}{2\mathrm{\Delta }\tau ^2}}\right),`$ (125) the temporal distribution of particle evaporation $`H(\tau )`$ is assumed to have the form of $$H(\tau )=\frac{1}{(2\pi \mathrm{\Delta }\tau ^2)^{3/2}}\mathrm{exp}\left[\frac{(\tau \overline{\tau })^2}{2\mathrm{\Delta }\tau ^2}\right],$$ (126) and it is assumed that the widths of the particle emitting sources, e.g. $`R_G`$ and $`\mathrm{\Delta }\eta `$ do not change significantly during the course of the emission of the observable particles. The parameters $`\mathrm{\Delta }T/T_r`$ and $`\mathrm{\Delta }T/T_t`$ control the transversal and the temporal changes of the local temperature profile, see Refs for further details. This formulation of the BL hydro source includes a competition between the transversal flow and the transverse temperature gradient, in an analytically tractable form. In the analytic evaluation of this model, it is assumed that the transverse flow is non-relativistic at the point of maximum emissivity , the temperature gradients were introduced following the suggestion of Akkelin and Sinyukov . Note that the shape of the profile function in $`\eta `$ is assumed to be a Gaussian in Eq. (124) in the spirit of introducing only means and variances. However, in Ref. a formula was given, that allows the reconstruction of this part of the emission function from the measured double-differential invariant momentum distribution in a general manner, for arbitrary sources with scaling longitudinal expansions. ### 8.1 Correlations and spectra for the BL-Hydro Using the binary source formulation, reviewed in the next section, the invariant single particle spectrum is obtained as $`N_1(𝐤)`$ $`=`$ $`{\displaystyle \frac{g}{(2\pi )^3}}\overline{E}\overline{V}\overline{C}{\displaystyle \frac{1}{\mathrm{exp}\left({\displaystyle \frac{u^\mu (\overline{x})k_\mu }{T(\overline{x})}}{\displaystyle \frac{\mu (\overline{x})}{T(\overline{x})}}\right)+s}}.`$ (127) The two-particle Bose–Einstein correlation function was evaluated in the binary source formalism in Ref. : $`C_2(𝐤_1,𝐤_2)`$ $`=`$ $`1+\lambda _{}\mathrm{\Omega }(Q_{})\mathrm{exp}\left(Q_{}^2\overline{R}_{}^2Q_=^2\overline{R}_=^2Q_{}^2\overline{R}_{}^2\right),`$ (128) where the pre-factor $`\mathrm{\Omega }(Q_{})`$ induces oscillations within the Gaussian envelope as a function of $`Q_{}`$. This oscillating pre-factor satisfies $`0\mathrm{\Omega }(Q_{})1`$ and $`\mathrm{\Omega }(0)=1`$. This factor is given as $`\mathrm{\Omega }(Q_{})`$ $`=`$ $`\mathrm{cos}^2(Q_{}\overline{R}_{}\mathrm{\Delta }\overline{\eta })+\mathrm{sin}^2(Q_{}\overline{R}_{}\mathrm{\Delta }\overline{\eta })\mathrm{tanh}^2(\overline{\eta }).`$ (129) The invariant BL decomposition of the relative momentum is utilized to present the correlation function in the simplest possible form. Although the shape of the BECF is non-Gaussian, because the factor $`\mathrm{\Omega }(Q_{})`$ results in oscillations of the correlator, the result is still explicitly boost-invariant. Although the source is assumed to be cylindrically symmetric, we have 6 free fit parameters in this BL form of the correlation function: $`\lambda _{}`$, $`R_=`$, $`R_{}`$, $`R_{}`$, $`\overline{\eta }`$ and $`\mathrm{\Delta }\overline{\eta }`$. The latter controls the period of the oscillations in the correlation function, which in turn carries information on the separation of the effective binary sources. This emphasizes the importance of the oscillating factor in the BL Bose–Einstein correlation function. The parameters of the spectrum and the correlation function are the same, defined as follows. In the above equations, $`\overline{a}`$ means a momentum-dependent average of the quantity $`a`$. The average value of the space-time four-vector $`\overline{x}`$ is parameterized by $`(\overline{\tau },\overline{\eta },\overline{r_x},\overline{r_y})`$, denoting longitudinal proper-time, space-time rapidity and transverse directions. These values are obtained in terms of the BL-H parameters in a linearized solution of the saddle-point equations as $`\overline{\tau }`$ $`=`$ $`\tau _0,`$ (130) $`\overline{\eta }`$ $`=`$ $`(y_0y)/\left[1+\mathrm{\Delta }\eta ^2m_t/T_0\right],`$ (131) $`\overline{r_x}`$ $`=`$ $`u_tR_G{\displaystyle \frac{p_t}{T_0+\overline{E}\left(u_t+\mathrm{\Delta }T/T_r\right)}},`$ (132) $`\overline{r_y}`$ $`=0.`$ (133) In Eq. (127), $`\overline{E}`$ stands for an average energy, $`\overline{V}`$ for an average volume of the effective source of particles with a given momentum $`k`$ and $`\overline{C}`$ for a correction factor, each defined in the LCMS frame: $`\overline{E}`$ $`=`$ $`m_t\mathrm{cosh}(\overline{\eta }),`$ (134) $`\overline{V}`$ $`=`$ $`(2\pi )^{\frac{3}{2}}\overline{R}_{}\overline{R}_{}^2{\displaystyle \frac{\mathrm{\Delta }\overline{\tau }}{\mathrm{\Delta }\tau }},`$ (135) $`\overline{C}`$ $`=`$ $`\mathrm{exp}\left(\mathrm{\Delta }\overline{\eta }^2/2\right)/\sqrt{\lambda _{}}.`$ (136) The average invariant volume $`\overline{V}`$ is given as a time-averaged product of the transverse area $`\overline{R}_{}`$ and the invariant longitudinal source size $`\overline{R}_{}`$, given as $`\overline{R}_{}^2`$ $`=`$ $`R_{..}^2=R_:^2=R_G^2/\left[1+\left(u_t^2+\mathrm{\Delta }T/T_r\right)\overline{E}/T_0\right],`$ (137) $`\overline{R}_{}^2`$ $`=`$ $`\overline{\tau }^2\mathrm{\Delta }\overline{\eta }^2,`$ (138) $`\mathrm{\Delta }\overline{\eta }^2`$ $`=`$ $`\mathrm{\Delta }\eta ^2/\left(1+\mathrm{\Delta }\eta ^2\overline{E}/T_0\right),`$ (139) $`\overline{R}_=^2`$ $`=`$ $`\mathrm{\Delta }\overline{\tau }^2=\mathrm{\Delta }\tau ^2/\left(1+\mathrm{\Delta }T/T_t\overline{E}/T_0\right).`$ (140) This completes the specification of the shape of particle spectrum and that of the two-particle Bose–Einstein correlation function. These results for the spectrum correspond to the equations given in Ref. although they are expressed here using an improved notation. In a generalized form, the thermal scales are defined as the $`\overline{E}/T_0`$ $`\mathrm{}`$ limit of Eqs (137140), while the geometrical scales correspond to dominant terms in the $`\overline{E}/T_00`$ limit of these equations. In all directions, including the temporal one, the length-scales measured by the Bose–Einstein correlation function are dominated by the smaller of the thermal and the geometrical length-scales. As shown in Sections 11 and 12, the width of the rapidity distribution and the slope of the transverse-mass distribution is dominated by the bigger of the geometrical and the thermal length-scales. This is the analytic reason, why the geometrical source sizes, the flow and temperature profiles of the source can only be reconstructed with the help of a simultaneous analysis of the two-particle Bose–Einstein correlation functions and the single-particle momentum distribution . If the geometrical contributions to the HBT radii are sufficiently large as compared to the thermal scales, they cancel from the measured HBT radius parameters. In this case, even if the geometrical source distribution for different particles (pions, kaons, protons) were different, the HBT radii (lengths of homogeneity) approach a scaling function in the large $`\overline{E}/T_0`$ limit. Up to the leading order calculation in the transverse coordinate of the saddle-point, this model predicts a scaling in terms of $`\overline{E}`$, which variable coincides with the transverse mass $`m_t`$ at mid-rapidity.Phenomenologically, the scaling law can be summarized as $`R_im_t^{\alpha _i}`$, where $`i`$ indexes the directional dependence, and the exponent $`\alpha _i`$ may be slightly rapidity dependent, due to the difference between $`\overline{E}`$ and $`m_t`$, and it may phenomenologically reflect the effects of finite size corrections as well. Note also that such a scaling limiting case is only a possibility in the BL-H, valid in certain domain of parameter space, but it is not a necessity. The analysis of Pb + Pb collisions at 158 AGeV indicates that BL-H describes the data fairly well, but the longitudinal radius component exhibits different scaling behavior from the transverse radii, see Section 12 for more details. ## 9 Binary Source Formalism Let us first consider the binary source representation of the BL-H model. The two-particle Bose–Einstein correlation function was evaluated in Ref. only in a Gaussian approximation, without applying the binary source formulation. An improved calculation was recently presented in Ref. , where the correlation function was evaluated using in the binary source formulation, and the corresponding oscillations were found. Using the exponential form of the $`\mathrm{cosh}[\eta y]`$ factor, the BL-H emission function $`S_c(x,𝐤)`$ can be written as a sum of two terms: $`S_c(x,𝐤)`$ $`=`$ $`0.5[S_+(x,𝐤)+S_{}(x,𝐤)],`$ (141) $`S_\pm (x,𝐤)`$ $`=`$ $`{\displaystyle \frac{g}{(2\pi )^3}}m_t\mathrm{exp}[\pm \eta y]H_{}(\tau ){\displaystyle \frac{1}{[f_B(x,𝐤)+s]}},`$ (142) $`f_B(x,𝐤)`$ $`=`$ $`\mathrm{exp}\left[{\displaystyle \frac{k^\mu u_\mu (x)\mu (x)}{T(x)}}\right].`$ (143) Let us call this splitting as the binary source formulation of the BL-H parameterization. The effective emission function components are both subject to Fourier transformation in the BL approach. In an improved saddle-point approximation, the two components $`S_+(x,k)`$ and $`S_{}(x,k)`$ can be Fourier-transformed independently, finding the separate maxima (saddle point) $`\overline{x}_+`$ and $`\overline{x}_{}`$ of $`S_+(x,k)`$ and $`S_{}(x,k)`$, and performing the analytic calculation for the two components separately. The oscillations in the correlation function are due to this effective separation of the pion source to two components, a splitting caused by the Cooper–Frye flux term. These oscillations in the intensity correlation function are similar to the oscillations in the intensity correlations of photons from binary stars in stellar astronomy . Due to the analytically found oscillations, the presented form of the BECF goes beyond the single Gaussian version of the saddle-point calculations of Refs . This result goes also beyond the results obtainable in the YKP or the BP parameterizations. In principle, the binary-source saddle-point calculation gives more accurate analytic results than the numerical evaluation of space-time variances, as the binary-source calculation keeps non-Gaussian information on the detailed shape of the Bose–Einstein correlation function. Note that the oscillations are expected to be small in the BL-H picture, and the Gaussian remains a good approximation to Eq. (128), but with modified radius parameters. ### 9.1 The general binary source formalism In the previous subsection, we have seen how effective binary sources appear in the BL-H model in high energy physics. However, binary sources appear generally: in astrophysics, in form of binary stars, in particle physics, in form of W<sup>+</sup>W<sup>-</sup> pairs, that separate before they decay to hadrons. Let us consider first the simplest possible example, to see how the binary sources result in oscillations in the Bose–Einstein or Fermi–Dirac correlation function. Suppose a source distribution $`s(xx_+)`$ describes e.g. a Gaussian source, centered on $`x_+`$. Consider a binary system, where the emission happens from $`s_+=s(xx_+)`$ with fraction $`f_+`$, or from a displaced source, $`s_{}=s(xx_{})`$, centered on $`x_{}`$, with a fraction $`f_{}`$. For such a binary source, the amplitude of the emission is $$\rho (x)=f_+s(xx_+)+f_{}s(xx_{}),$$ (144) and the normalization requires $$f_++f_{}=1.$$ (145) The two-particle Bose–Einstein or Fermi–Dirac correlation function is $`C(q)`$ $`=`$ $`1\pm |\stackrel{~}{\rho }(q)|^2=1\pm \mathrm{\Omega }(q)|\stackrel{~}{s}(q)|^2,`$ (146) where $`+`$ is for bosons, and $``$ for fermions. The oscillating pre-factor $`\mathrm{\Omega }(q)`$ satisfies $`0\mathrm{\Omega }(q)1`$ and $`\mathrm{\Omega }(0)=1`$. This factor is given as $`\mathrm{\Omega }(q)`$ $`=`$ $`\left[(f_+^2+f_{}^2)+2f_+f_{}\mathrm{cos}[q(x_+x_{})]\right].`$ (147) The strength of the oscillations is controlled by the relative strength of emission from the displaced sources and the period of the oscillations can be used to learn about the distance of the emitters. In the limit of one emitter ($`f_+=1`$ and $`f_{}=0`$, or vice versa), the oscillations disappear. The oscillating part of the correlation function in high energy physics is expected to be much smaller, than that of binary stars in stellar astronomy. In particle physics, the effective separation between the sources can be estimated from the uncertainty relation to be $`x_\pm =|x_+x_{}|2\mathrm{}/M_\mathrm{W}0.005`$ fm. Although this is much smaller, the effective size of the pion source, 1 fm, one has to keep in mind that the back-to-back momenta of the W<sup>+</sup>W<sup>-</sup> pairs can be large, as compared to the pion mass. Due to this boost, pions with similar momentum may be emitted from different W-s with a separation which is already comparable to the 1 fm hadronization scale, and the resulting oscillations may become observable. In stellar astronomy, the separation between the binary stars is typically much larger than the diameter of the stars, hence the oscillations are well measurable. In principle, similar oscillations may provide a tool to measure the separation of the W<sup>+</sup> from W<sup>-</sup> in 4-jet events at LEP2. The scale of separation of W<sup>+</sup>W<sup>-</sup> pairs is a key observable to estimate in a quantum-mechanically correct manner the influence of the Bose–Einstein correlations on the reconstruction of the W mass. In heavy ion physics, oscillations are seen in the long-range part of the p + p Fermi–Dirac correlation function , with a half-period of $`Q_h=30`$ MeV. This implies a separation of $`x_\pm =\pi \mathrm{}/Q_h20`$ fm, which can be attributed to interference between the the two peaks of the NA49 proton $`dn/dy`$ distribution , separated by $`\mathrm{\Delta }y=2.5`$. As for the protons we have $`mT_0=140`$ MeV, we can identify this rapidity difference with the space-time rapidity difference between the two peaks of the rapidity distribution. The longitudinal scale of the separation is then given by $`x_\pm =2\overline{\tau }\mathrm{sinh}(\mathrm{\Delta }\eta _p/2)`$, which can be used to estimate the mean freeze-out time of protons, $`\overline{\tau }=\pi \mathrm{}/[2Q_h\mathrm{sinh}(\mathrm{\Delta }\eta _p/2)]6.4`$ fm/c, in a good agreement with the average value of $`\overline{\tau }=5.9\pm 0.6`$ as extracted from the simultaneous analysis of the single-particle spectra and HBT radii in NA44, NA49 and WA98 experiments in the Buda–Lund picture, as summarized in Section 12. ## 10 Particle Correlations and Spectra at 30 – 160 AMeV There are important qualitative differences between relativistic heavy ion collisions at CERN SPS and those at non-relativistic energies from the point of view of particle sources. Low and intermediate energy reactions may create a very long-lived, evaporative source, with characteristic lifetimes of a few 100 fm/c, in contrast to the relatively short-lived systems of lifetimes of the order of 10 fm/c at CERN SPS. During such long evaporation times, cooling of the source is unavoidable and has to be included into the model. Furthermore, in the non-relativistic heavy ion collisions mostly protons and neutrons are emitted and they have much stronger final-state interactions than the pions dominating the final state at ultra-relativistic energies, see Refs for recent reviews. The evolution of the particle emission in a heavy-ion collision at intermediate energies may roughly be described as: production of pre-equilibrium particles; expansion and possible freeze-out of a compound source; possible evaporation from an excited residue of the source. Note though that this separation is not very distinct and there is an overlap between the different stages. The importance of the various stages above also depends on the beam energy and the impact parameter of the collision. See the review paper of Ref. for greater details. Sophisticated microscopical transport descriptions , such as the BUU(Boltzmann–Uehling–Uhlenbeck) and the QMD (Quantum Molecular Dynamics) models are well-known and believed to provide a reasonable picture of proton emission in central heavy ion collisions from a few tenths up to hundreds of MeV per nucleon. However, the BUU model predicts too large correlations and underpredicts the number of protons emitted with low energies, for the reaction <sup>36</sup>Ar + <sup>45</sup>Sc at $`E=120`$ and 160 MeV/nucleon, see Ref. . This indicates that the simultaneous description of two-particle correlations and single-particle spectra is a rather difficult task. For energies below a few tens of MeV per nucleon, where long-lived evaporative particle emission is expected to dominate, the measured two-protoncorrelation functions were found to be consistent with compound-nucleus model predictions . A simultaneous analysis of proton and neutron single particle spectra and two-particle correlation was presented in Ref. . This model calculation described the second stage above and, for long emission times, also part of the third stage. In Ref. , the competition among particle evaporation, temperature gradient and flow was investigated in a phenomenological manner, based on a simultaneous analysis of quantum statistical correlations and momentum distributions for a non-relativistic, spherically symmetric, three-dimensionally expanding, finite source. The model used can be considered as a non-relativistic, spherically symmetric version of the BL-H hydro parameterization . The non-relativistic kinetic energy is denoted by $`E_k(𝐤)=𝐤^{\mathrm{\hspace{0.17em}2}}/(2m)`$. The following result is obtained for the effective source size $`R_{}`$: $`R_{}^2(𝐤)`$ $`=`$ $`{\displaystyle \frac{R_G^2}{1+\left[\mathrm{\Delta }T/T_rE_k(𝐤)+mu_t^2\right]/T_0}}.`$ (148) The analytic results for the momentum distribution and the quantum statistical correlation function are given in the Boltzmann approximation as $`N_1(𝐤)`$ $`=`$ $`{\displaystyle \frac{g}{(2\pi )^3}}E_k(𝐤)V_{}(𝐤)\mathrm{exp}\left[{\displaystyle \frac{(𝐤m𝐮(𝐫_s(𝐤)))^2}{2mT(𝐫_s(𝐤))}}+{\displaystyle \frac{\mu (𝐫_s(𝐤))}{T(𝐫_s(𝐤))}}\right],`$ $`V_{}(𝐤)`$ $`=`$ $`\left[2\pi R_{}^2(𝐤)\right]^{3/2},`$ (150) $`C(K,\mathrm{\Delta }k)`$ $`=`$ $`1\pm \mathrm{exp}(R_{}^2(𝐊)𝚫𝐤^2\mathrm{\Delta }t^2\mathrm{\Delta }E^2).`$ (151) The effects of final-state Coulomb and Yukawa interactions on the two-particle relative wave-functions are neglected in these analytic expressions. When comparing to data, the final-state interactions were taken into account, see Ref. for further details. These general results for the correlation function indicate structural similarity between the non-relativistic flows in low/intermediate energy heavy ion collisions and the transverse flow effects in relativistic high energy heavy ion and elementary particle induced reactions . The radius parameters of the correlation function and the slopes of the single-particle spectra are momentum dependent both for the non-relativistic versions of the model, presented in Refs and for the model-class with scaling relativistic longitudinal flows, discussed in Refs . Such a momentum-dependent effective source size has been seen in the proton–proton correlation functions in the <sup>27</sup>Al (<sup>14</sup>N, pp) reactions at $`E=75`$ MeV/nucleon : the larger the momentum of the protons the smaller the effective source size , in qualitative agreement with Eq. (148). This model was applied in Ref. to the reaction <sup>40</sup>Ar + <sup>197</sup>Au at 30 MeV/ nucleon. With the parameter set presented in Table 3, we have obtained a simultaneous description of the n and p single particle spectra as well as the nn and pp correlation functions as given by Refs . See Ref. for further details and discussions. The main effects of the temperature gradient are that it introduces i) a momentum-dependent effective temperature which is decreasing for increasing momentum, resulting in a suppression at high momentum as compared to the Boltzmann distribution; ii) a momentum-dependent effective source size which decreases with increasing total momentum. Agreement with the experimental data is obtained only if the time of duration of the particle emission was rather long, $`t520`$ fm with a variance of $`320`$ fm/c. The obtained parameter set reflects a moderately large system (Gaussian radius parameter $`R_G`$ = 4.0 fm) at a moderate temperature ($`T_0(n)`$ = 3 MeV and $`T_0(p)`$ = 5 MeV) and small flow. The neutrons and the protons seem to have different local temperature distributions: the neutron temperature distribution is homogeneous, while the temperature of the proton source decreases to $`T_s(p)=4.3`$ MeV at the Gaussian radius, a difference that could be attributed to the difference between their Coulomb interactions . An agreement between the model and the data was obtained only if some amount of flow was included . After the completion of the data analysis, a new family of exact solutions of fireball hydrodynamics was found in Ref. , which features scaling radial Hubble flow, and an initial inhomogeneous, arbitrary temperature profile. The competition of the temperature gradients and flow effects were shown to lead to the formation of spherical shells of fire in this class of exact hydrodynamical solutions , if the temperature gradient was stronger than the flow, $`\mathrm{\Delta }T/T_r>mu_t^2/T_0`$. This is the case found from the analysis of proton spectra and correlations in Ref. , while the neutron data do not satisfy this condition. Assuming the validity of non-relativistic hydrodynamics to characterize this reaction, one finds that a slowly expanding, spherical shell of fire is formed by the protons, while the neutrons remain in a central, slightly colder and even slower expanding, normal fireball in 30 AMeV <sup>40</sup>Ar + <sup>197</sup>Au heavy ion reactions. ## 11 Description of h + p Correlations and Spectra at CERN SPS The invariant spectra of $`\pi ^{}`$ mesons produced in $`(\pi ^+/\mathrm{K}^+)\mathrm{p}`$ interactions at 250 GeV/c are analysed in this section in the framework of the BL-H model of three-dimensionally expanding cylindrically symmetric finite systems, following the lines of Ref. . The EHS/NA22 Collaboration has been the first to perform a detailed and combined analysis of single-particle spectra and two-particle Bose–Einstein correlations in high energy physics . NA22 reported a detailed study of multi-dimensional Bose–Einstein correlations, by determining the side, out and the longitudinal radius components at two different values of the mean transversal momenta in $`(\pi ^+/\mathrm{K}^+)\mathrm{p}`$ at CERN SPS energies . It turned out, however, that the experimental two-particle correlation data were equally well described by a static Kopylov–Podgoretskii parameterization as well as by the predictions of hydrodynamical parameterizations for longitudinally expanding, finite systems. In Refs we have shown, that the combined analysis of two-particle correlations and single-particle spectra may result in a dramatic enhancement of the selective power of data analysis. The double-differential invariant momentum distribution of Eq. (127) can be substantially simplified for one-dimensional slices . i) At fixed $`m_t`$, the rapidity distribution reduces to $`N_1(𝐤)`$ $`=`$ $`C_m\mathrm{exp}\left[{\displaystyle \frac{(yy_0)^2}{2\mathrm{\Delta }y^2}}\right],`$ (152) $`\mathrm{\Delta }y^2`$ $`=`$ $`\mathrm{\Delta }\eta ^2+T_0/m_t,`$ (153) where $`C_m`$ is an $`m_t`$-dependent normalization coefficient and $`y_0`$ is defined above. The width parameter $`\mathrm{\Delta }y^2`$ extracted for different $`m_t`$-slices is predicted to depend linearly on $`1/m_t`$, with slope $`T_0`$ and intercept $`\mathrm{\Delta }\eta ^2`$ . Observe, that this width is dominated by the bigger of the geometrical scale ($`\mathrm{\Delta }\eta `$) and the thermal scale $`T_0/m_t`$. Note that for static fireballs or spherically expanding shells (152) and (153) are satisfied with $`\mathrm{\Delta }\eta =0`$ . Hence the experimental determination of the $`1/m_t`$ dependence of the $`\mathrm{\Delta }y`$ parameter can be utilized to distinguish between longitudinally expanding finite systems versus static fireballs or spherically expanding shells. ii) At fixed $`y`$, the $`m_t^2`$-distribution reduces to $$N_1(𝐤)=C_ym_t^\alpha \mathrm{exp}\left(\frac{m_t}{T_{\mathrm{eff}}}\right),$$ (154) where $`C_y`$ is a $`y`$-dependent normalization coefficient and $`\alpha `$ is related to the effective dimensions of inhomogeneity in the source as $`\alpha =1d_{\mathrm{eff}}/2`$ . The $`y`$-dependent “effective temperature” $`T_{\mathrm{eff}}(y)`$ reads as $$T_{\mathrm{eff}}(y)=\frac{T_{}}{1+a(yy_0)^2},$$ (155) where $`T_{}`$ is the maximum of $`T_{\mathrm{eff}}(y)`$ achieved at $`y=y_0`$, and parameter $`a`$ can be expressed with the help of the other fit parameters, see Refs . The slope parameter at mid-rapidity, $`T_{}`$ is also determined by an interplay of the central temperature $`T_0`$ the flow effects modeled by $`u_t^2`$ and the temperature difference between the surface and the center, as characterized by $`\mathrm{\Delta }T/T_r`$ . Eq. (66) of Ref. can be rewritten as $$T_{}=T_0+mu_t^2\frac{T_0}{T_0+m\frac{\mathrm{\Delta }T}{T}_r}.$$ (156) The approximations of Eqs (152) and (154) explicitly predict a specific narrowing of the rapidity and transverse-mass spectra with increasing $`m_t`$ and $`y`$, respectively (cf. (153) and (155)). The character of these variations is expected to be different for the various scenarios of hadron matter evolution. These features of the spectra were found to be in agreement with the NA22 data , and were utilized to reconstruct the particle source of h + p reactions in the $`(t,r_z)`$ plane. ### 11.1 Combination with two-particle correlations As already mentioned in the introduction, more comprehensive information on geometrical and dynamical properties of the hadron matter evolution are expected from a combined consideration of two-particle correlations and single-particle inclusive spectra . At mid-rapidity, $`y=y_0`$ and in the LCMS where $`k_{1,z}=k_{2,z}`$, the effective BP radii can be approximately expressed form the BL-H parameterization as : $`R_l^2`$ $`=`$ $`\overline{\tau }^2\mathrm{\Delta }\overline{\eta }^2,`$ (157) $`R_o^2`$ $`=`$ $`\overline{R}_{}^2+\beta _t^2\mathrm{\Delta }\overline{\tau }^2,`$ (158) $`R_s^2`$ $`=`$ $`\overline{R}_{}^2`$ (159) with $`{\displaystyle \frac{1}{\mathrm{\Delta }\overline{\eta }^2}}`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{\Delta }\eta ^2}}+{\displaystyle \frac{M_t}{T_0}},`$ (160) $`\overline{R}_{}^2`$ $`=`$ $`{\displaystyle \frac{R_\mathrm{G}^2}{1+\frac{M_t}{T_0}(u_t^2+\frac{\mathrm{\Delta }T}{T}_r)}},`$ (161) where parameters $`\mathrm{\Delta }\eta ^2,T_0,u_t`$ and $`\mathrm{\Delta }T/T_r`$ are defined and estimated from the invariant spectra; $`R_\mathrm{G}`$ is related to the transverse geometrical rms radius of the source as $`R_\mathrm{G}(\mathrm{rms})=\sqrt{2}R_\mathrm{G}`$; $`\overline{\tau }`$ is the mean freeze-out (hadronization) time; $`\mathrm{\Delta }\overline{\tau }`$ is related to the duration time $`\mathrm{\Delta }\overline{\tau }`$ of pion emission and to the temporal inhomogeneity of the local temperature, as the relation $`\mathrm{\Delta }\tau \mathrm{\Delta }\overline{\tau }`$ holds; the variable $`\beta _t`$ is the transverse velocity of the pion pair. The effective longitudinal radius $`R_l`$, extracted for two different mass ranges, $`M_t=0.26\pm 0.05`$ and $`0.45\pm 0.09`$ GeV/c<sup>2</sup> are found to be $`R_l=0.93\pm 0.04`$ and $`0.70\pm 0.09`$ fm, respectively. This dependence on $`M_t`$ matches well the predicted one. Using Eq. (159) with $`T_0=140\pm 3`$ MeV and $`\mathrm{\Delta }\eta ^2=1.85\pm 0.04`$ (Table 4), one finds that the values of $`\overline{\tau }`$ extracted for the two different $`M_t`$ regions are similar to each other: $`\overline{\tau }=1.44\pm 0.12`$ and $`1.36\pm 0.23`$ fm/c. The averaged value of the mean freeze-out time is $`\overline{\tau }=1.4\pm 0.1`$ fm/c. The width of the (longitudinal) space-time rapidity distribution of the pion source was found to be $`\mathrm{\Delta }\eta =1.36\pm 0.02`$. Since this value of $`\mathrm{\Delta }\eta `$ is significantly bigger than 0, the static fireballs or the spherically expanding shells fail to reproduce the NA22 single-particle spectra , although each of these models was able to describe the NA22 two-particle correlation data in Ref. . The transverse-plane radii $`R_o`$ and $`R_s`$ were reported in Ref. for the whole $`M_t`$ range are: $`R_o=0.91\pm 0.08`$ fm and $`R_s=0.54\pm 0.07`$ fm. Substituting in (157) and (158), one obtains (at $`\beta _t=0.484`$ c ): $`\mathrm{\Delta }\overline{\tau }=1.3\pm 0.3`$ fm/c. The mean duration time of pion emission can be estimated as $`\mathrm{\Delta }\tau \mathrm{\Delta }\overline{\tau }=1.3\pm 0.3`$ fm/c. A possible interpretation of $`\mathrm{\Delta }\tau \overline{\tau }`$ might be that the radiation process occurs during almost all the hydrodynamical evolution of the hadronic matter produced in meson–proton collisions. An estimation for the parameter $`R_\mathrm{G}`$ can be obtained from (158) and (160) using the quoted values of $`R_s`$, $`T_0`$, $`u_t`$ and $`\mathrm{\Delta }T/T`$ at the mean value of $`M_t=0.31\pm 0.04`$ GeV/c (averaged over the whole $`M_t`$ range): $`R_\mathrm{G}=0.88\pm 0.13`$ fm. The geometrical rms transverse radius of the hydrodynamical tube, $`R_\mathrm{G}(\mathrm{rms})=\sqrt{2}R_\mathrm{G}=1.2\pm 0.2`$ fm, turns out to be larger than the proton rms transverse radius. The data favour the pattern according to which the hadron matter undergoes predominantly longitudinal expansion and non-relativistic transverse expansion with mean transverse velocity $`u_t=0.20\pm 0.07`$, and is characterized by a large temperature inhomogeneity in the transverse direction: the extracted freeze-out temperature at the center of the tube and at the transverse rms radius are $`140\pm 3`$ MeV and $`82\pm 7`$ MeV, respectively. ### 11.2 The space-time distribution of $`\pi `$ emission A reconstruction of the space-time distribution of pion emission points is shown in Fig. 13, expressed as a function of the cms time variable $`t`$ and the cms longitudinal coordinate $`zr_z`$. The momentum-integrated emission function along the $`z`$ axis, i.e. at $`\text{r}_t=(r_x,r_y)=(0,0)`$ is given by $$S(t,z)\mathrm{exp}\left(\frac{(\tau \overline{\tau })^2}{2\mathrm{\Delta }\tau ^2}\right)\mathrm{exp}\left(\frac{(\eta y_0)^2}{2\mathrm{\Delta }\eta ^2}\right).$$ (162) It relates the parameters fitted to the NA22 single-particle spectrum and HBT radii to the particle production in space-time. The coordinates $`(t,z)`$ are expressed with the help of the longitudinal proper-time $`\tau `$ and space-time rapidity as $`\eta `$ as $`(\tau \mathrm{cosh}(\eta ),\tau \mathrm{sinh}(\eta ))`$. We find a structure looking like a boomerang, i.e. particle production takes place close to the regions of $`z=t`$ and $`z=t`$, with gradually decreasing probability for ever larger values of space-time rapidity. Although the mean proper-time for particle production is $`\overline{\tau }=1.4`$ fm/c, and the dispersion of particle production in space-time rapidity is rather small, $`\mathrm{\Delta }\eta =1.35`$ fm, we still see a characteristic long tail of particle emission on both sides of the light-cone, giving a total of 40 fm maximal longitudinal extension in $`z`$ and a maximum of about 20 fm/c duration of particle production in the time variable $`t`$. In the transverse direction, only the rms width of the source can be directly inferred from the BP radii. However, the additional information from the analysis of the transverse momentum distribution on the values of $`u_t`$ and on the values of $`\mathrm{\Delta }T/T_r`$ can be used to reconstruct the details of the transverse density profile, as an exact, non-relativistic hydro solution was found in Ref. , given in terms of the parameters $`u_t`$ and $`\mathrm{\Delta }T/T_r`$ and using an ideal gas equation of state. Assuming the validity of this non-relativistic solution in the transverse direction, in the mid-rapidity range, one can reconstruct the detailed shape of the transverse densityprofile. The result looks like a ring of fire in the $`(r_x,r_y)`$ plane, see Fig. 14. In this hydro solution, $`\mathrm{\Delta }T/T_r<mu_t^2/T_0`$ corresponds to self-similar, expanding fireballs, while $`\mathrm{\Delta }T/T_r>mu_t^2/T_0`$ corresponds to self-similar, expanding shells or rings of fire. Due to the strong surface cooling and the small amount of the transverse flow, one finds that the particle emission in the transverse plane of h + p reactions at CERN SPS corresponds to a ring of fire. This transverse distribution, together with the scaling longitudinal expansion, creates an elongated, tube-like source in three dimensions, with the density of particle production being maximal on the surface of the tube. ## 12 Pb + Pb Correlations and Spectra at CERN SPS In Ref. , an analysis similar to that of the NA22 Collaboration has been performed, with improved analytic approximations, using Fermi–Dirac or Bose–Einstein statistics $`(s=\pm 1)`$ in the analytic expressions fitted to single particle spectra. The spectra were evaluated with the binary source method, the Bose–Einstein correlation functions were calculated with the saddle-point method without invoking the binary source picture. The analytical formulas for the BECF and IMD, as were used in the fits, were summarized in their presently most advanced form in Section 8, their development was described in Refs . In case of homogeneous freeze-out temperatures, or particles with small masses, Eq. (156) implies a linear rise of the slope with $`m`$ as $$T_{}(m)=T_0+mu_t^2,\text{if }\frac{\mathrm{\Delta }T}{T}_rT_0/m.$$ (163) For heavy particles, or for large, non-vanishing temperature gradients, a flattening of the initial linear rise is obtained as $$T_{}(m)=T_0\left[1+\frac{u_t^2}{\frac{\mathrm{\Delta }T}{T}_r}\right],\text{if }\frac{\mathrm{\Delta }T}{T}_rT_0/m.$$ (164) This means that very heavy particles resolve the temperature inhomogeneities of the source, and they are produced with a mass-independent effective slope parameter in the BL-H parameterization, if $`T_0/m`$ becomes smaller than the temperature inhomogeneity. In a general case, the $`T_{}(m)`$ function starts with an initial linear $`m`$ dependence, with a slope given by the transverse flow $`u_t`$, then $`T_{}(m)`$ flattenes out to a mass-independent value if the source has temperature inhomogeneities in the transverse direction. Such a behavior was reported by Pb + Pb heavy ion experiments at CERN SPS . The central temperature is $`T_0140`$ MeV, the flattening of the slopes sets in at about $`m=1400`$ MeV , which then leads to about 10% temperature inhomogeneity in the transverse direction of the Pb + Pb source. This estimate is in a good agreement with the results of the combined analysis of the single-particle spectra and the two-particle Bose–Einstein correlation functions, see Table 5. The NA49, NA44 and WA98 data on single particle spectra of h<sup>-</sup>, identified $`\pi `$, K and p as well as detailed rapidity and $`m_t`$ dependent HBT radius parameters are found to be consistent with each other as well as with BL-H. The BL-H fit results to these data sets is summarized in Table 5, Ref. . ## 13 Comparison of h + p and Pb + Pb Final States at CERN SPS with Heavy Ion Reactions at Low and Intermediate Energies The final state of central Pb + Pb collisions at CERN SPS corresponds to a cylindrically symmetric, large ($`R_G=7.1\pm 0.2`$ fm) and transversally homogenous ($`T_0=139\pm 6`$ MeV) fireball, expanding three-dimensionally with $`u_t=0.55\pm 0.06`$. A large mean freeze-out time, $`\overline{\tau }=5.9\pm 0.6`$ is found with a relatively short duration of emission, $`\mathrm{\Delta }\overline{\tau }=1.6\pm 1.5`$ fm, which is similar to the time-scale of emission in the h + p reaction. Note that the temporal cooling in Pb + Pb reactions seems to be stronger than in h + p, which can be expained by the faster, three-dimensional expansion in the former case, as compared to the essentially one-dimensional expansion in the case of h + p reactions. By the time the particle production is over, the surface of Pb + Pb collisions cools down from 139 MeV to $`T_0/(1+\mathrm{\Delta }T/T_0_r)/(1+\mathrm{\Delta }T/T_0_t)83`$ MeV. It is very interesting to note that this value is similar to the surface temperature of $`T_s=82\pm 7`$ MeV, found in h + p reactions as a consequence of the transverse temperature inhomogeneities, as described in Section 11, Ref. . Such snowballs with relatively low values of surface temperature $`T_s`$ and a possible hotter core were reported first in 200 AGeV S + Pb reactions in Ref. . Other hydro parameterizations, as reviewed in Ref. , frequently neglect the effects of temperature inhomogeneities during the expansion and particle production stage. Energy conservation implies that the temperature cannot be exactly constant when particles are freezing out in a non-vanishing period of time from a three-dimensionally expanding source. The exact solution of non-relativistic, spherically symmetric fireball hydrodynamics implies that Gaussian fireballs with spatially uniform temperature profiles satisfy the collisionless Boltzmann equation. Fixing the temperature to a constant in the fits yields an average freeze-out temperature in the range of $`T_f=110\pm 30`$ MeV . Based on the recently found new family of non-relativistic hydrodynamics and on the analysis of h + p single particle spectra and two-particle Bose–Einstein correlation function , we concluded that the pion emission function $`S(r_x,r_y)`$ in h + p reactions corresponds to the formation of a ring of fire in the transverse plane, because the transverse flow is rather small and because the sudden drop of the temperature in the transverse direction leads to large pressure gradients in the center and small pressure gradients and a density built-up at the expanding radius of the fire-ring. We presented arguments for a similar formation of a spherical shell of fire in the proton distributions at 30 AMeV <sup>40</sup>Ar + <sup>197</sup>Au reactions. The formation of shells of fire seems thus to be of a rather generic nature, related to the initial conditions of self-similar radial flows. It is natural to ask the question: can we learn more about this phenomena in other physical systems? Radial expansion is a well established phenomena in heavy ion collisions from low energy to high energy reactions. See Refs for recent reviews and for example see Refs for the evidence of collective flow in central heavy ion collisions from 100 AMeV to 2 AGeV as measured by the FOPI Collaboration at GSI SIS. The FOPI Collaboration measured recently the proton–proton correlation functions at 1.93 AGeV Ni + Ni collisions . To interpret their data, they utilized a version of the hydrodynamical solution, found in Ref. . They assumed a linear flow profile, a Gaussian density distribution and a constant temperature. Such a solution of fireball hydrodynamics exists, but it corresponds to a collisionless Knudsen gas . A collisionless approximation has to break down. Indeed, only the peak of the FOPI proton–proton correlation function was reproduced by the collisionless model, however, the tails had to be excluded from the FOPI analysis. Perhaps it is worthwhile to search for a possible formation of shells of fire at the SIS energy domain, by re-analyzing the FOPI data . ## 14 Shells of Fire and Planetary Nebulae In transport calculations based on the Boltzmann–Uehling–Uhlenbeck equation, a formation of toroidal density distributions was predicted for central <sup>36</sup>Ar + <sup>45</sup>Sc collisions at $`E=80`$ AMeV in Ref. , which leads to ring-like configurations for $`S(r_x,r_y)`$. However, the clearest experimental observation of the development of expanding shell like structures in the time evolution of exploding fireballs comes from stellar astronomy. Stars with initial masses of less than about eight solar masses end their lives by ejecting planetary nebulae, stellar remnants turning to white dwarfs. After the star has completed its core hydrogen burning, it becomes a red giant. In the core of the star, helium burns while hydrogen continues to burn in a thin shell surrounding the core. This hydrogen rich shell swells to enormous size, and the surface temperature drops to a rather low value for stars. A solar wind develops that carries away most of the hydrogen envelope surrounding the star’s central core. The envelope material ejected by the star forms an expanding shell of gas that is known as a planetary nebula. Planetary nebulae are illuminated by their central stars and display a variety of often beautiful structures. Some are spherical or helical, others have bipolar shapes, and still others are rather irregularly shaped. In a matter of a few tens of thousands of years, they intermingle with the interstellar medium and disperse. The space-time evolution of planetary nebulae is in many aspects similar to the solution of non-relativistic hydrodynamics given in Ref. . We argued, that this solution seems to describe also low and intermediate heavy ion collisions in the 30 – 80 AMeV energy domain. A similar hydro solution may also describe the non-relativistic transverse dynamics at mid-rapidity in hadron + proton collisions in the CERN SPS energy domain, compare Figs 14 and 20, the latter from Ref. . In all of these physical systems, expansion competes with the drop of the pressure gradients, which in turn is induced by the drop of the temperature on the surface. If the flow is small enough, the drop of the temperature on the surface results in a drop of the pressure gradients on the surface, which implies density pile-up. On the other hand, if the flow is strong enough, it blows away the material from the surface, preventing the formation of shells of fire, and an ordinary expanding fireball is obtained. Finally I note that this situation is just a special class of the more general solutions given in Ref. . Arbitrary number of self-similarly expanding, simultaneously existing shells of fire can be described by the general form of new class of exact solutions of fireball hydrodynamics . ## 15 Signal of Partial $`U_A(1)`$ Symmetry Restoration from Two-Pion Bose–Einstein Correlations In this section let me summarize Ref. , where the effective intercept parameter of the two-pion Bose–Einstein Correlation function, $`\lambda _{}`$ was shown to carry a sensitive and measurable signal of partial restoration of the axial $`U_A(1)`$ symmetry and the related increase of the $`\eta ^{}`$ production in ultra-relativistic nuclear collisions: An increase in the yield of the $`\eta ^{}`$ meson, proposed earlier as a signal of partial $`U_A(1)`$ restoration, was shown to create a “hole” in the low $`p_t`$ region of $`\lambda _{}`$. In the chiral limit ($`m_u=m_d=m_s=0`$), QCD possesses a $`U(3)`$ chiral symmetry. When broken spontaneously, $`U(3)`$ implies the existence of nine massless Goldstone bosons. In Nature, there are only eight light pseudoscalar mesons, a discrepancy which is resolved by the Adler–Bell–Jackiw $`U_A(1)`$ anomaly; the ninth would-be Goldstone boson gets a mass as a consequence of the non-zero density of topological charges in the QCD vacuum . In Refs , it is argued that the ninth (“prodigal” ) Goldstone boson, the $`\eta ^{}`$, would be abundantly produced if sufficiently hot and dense hadronic matter is formed in nucleus–nucleus collisions. Estimates of Ref. show that the corresponding production cross section of the $`\eta ^{}`$ should be enhanced by a factor of 3 up to 50 relative to that for p + p collisions. If the $`\eta ^{}`$ mass is decreased, a large fraction of the $`\eta ^{}`$s will not be able to leave the hot and dense region through thermal fluctuation since they need to compensate for the missing mass by large momentum . These $`\eta ^{}`$s will thus be trapped in the hot and dense region until it disappears, after which their mass becomes normal again; as a consequence, the $`\eta ^{}`$-s will have small transverse momenta $`p_t`$. Then they decay to pions via $$\eta ^{}\eta +\pi ^++\pi ^{}(\pi ^0+\pi ^++\pi ^{})+\pi ^++\pi ^{}.$$ (165) It is important to observe that the $`p_t`$ of pions produced in this decay chain is small since many of the $`\eta ^{}`$ appear at $`p_t0`$ and also since the rest mass of the decay products from the $`\eta ^{},\eta `$ decays use up most of the remaining energy. Based on the kinematics of the $`\eta ^{},\eta `$ decay chain to pions, an enhanced production of $`\pi `$ mesons was estimated to happen dominantly in the $`p_t150`$ MeV region, extending to a maximum $`p_t407`$ MeV . In the core/halo picture the $`\eta ^{},\eta `$ decays contribute to the halo due to their large decay time ($`1/\mathrm{\Gamma }_{\eta ^{},\eta }20`$ fm/c). Thus, we expect a hole in the $`0p_t150`$ MeV region of the effective intercept parameter, $`\lambda _{}=[N_{\mathrm{core}}(𝐩)/N_{\mathrm{total}}(𝐩)]^2`$. To calculate the $`\pi ^+`$ contribution from the halo region, the bosons ($`\omega `$, $`\eta ^{}`$, $`\eta `$ and $`\mathrm{K}_S^0`$) are given both a rapidity $`(1.0<y<1.0)`$ and an $`m_t`$, then are decayed using Jetset 7.4 . The $`m_t`$ distribution of the bosons is given by $$N(m_t)=Cm_t^\alpha e^{m_t/T_{\mathrm{eff}}},$$ (166) where $`C`$ is a normalization constant, $`\alpha =1d/2`$ and where $$T_{\mathrm{eff}}=T_{\mathrm{fo}}+mu_t^2.$$ (167) In the above expression, $`d=3`$ is the dimension of expansion, $`T_{\mathrm{fo}}=140`$ MeV is the freeze-out temperature and $`u_t`$ is the average transverse flow velocity. It should be noted that the $`m_t`$ distribution for the core pions is also obtained from Eq. (166). The contributions from the decay products of the different regions (halo and core) are then added together according to their respective fractions, allowing for the determination of $`\lambda _{}(m_t)`$. The respective fractions of pions are estimated separately by Fritiof and by RQMD as summarized in Ref. . Simulating the presence of the hot and dense region involves increasing the relative abundance of the $`\eta ^{}`$ and also changing their $`p_t`$ spectrum. The $`p_t`$ spectrum of the $`\eta ^{}`$ is obtained by assuming energy conservation and zero longitudinal motion at the boundary between the two phases. This conservation of transverse mass at the boundary implies $$m_\eta ^{}^2+p_{t}^{}{}_{\eta ^{}}{}^{2}=m_\eta ^{}^2+p_{t}^{}{}_{\eta ^{}}{}^{2},$$ (168) where the $``$ denotes the $`\eta ^{}`$ in the hot dense region. The $`p_t`$ distribution then becomes a two-fold distribution. The first part of the distribution is from the $`\eta ^{}`$ which have $`p_t^{}[m_\eta ^{}^2m_\eta ^{}^2]^{1/2}`$. These particles are given a $`p_t=0`$. The second part of the distribution comes from the rest of the $`\eta ^{}`$’s which have big enough $`p_t`$ to leave the hot and dense region. These have the same, flow motivated $`p_t`$ distribution as the other produced resonances and are given a $`p_t`$ according to the $`m_t`$ distribution $$N_\eta ^{}(m_t^{})=Cm_{t}^{}{}_{}{}^{0.5}e^{m_t^{}/T^{}},$$ (169) where $`C`$ is a normalization constant and where $`T^{}=200`$ MeV and $`m_\eta ^{}^{}`$ is the effective temperature and mass, respectively, of the hot and dense region. Using the value given above for the effective temperature and letting $`m_\eta ^{}^{}=500`$ MeV implies an increase in the production cross section of the $`\eta ^{}`$ in the hot and dense region by a factor of 10. Using three different effective masses for the $`\eta ^{}`$ in the hot and dense region, calculations of $`\lambda _{}(m_t)`$ including the hot and dense regions are compared to those assuming the standard abundances in Fig. 21. A similar $`m_t`$ dependence but with slightly higher values of $`\lambda _{}(m_t)`$ is obtained when using RQMD abundances. The effective mass of 738 MeV corresponds to an enhancement of the production cross section of the $`\eta ^{}`$ by a factor of 3, while the effective mass of 403 MeV and 140 MeV correspond to factors of 16 and 50 respectively. The two data points shown are taken from NA44 data on central S + Pb reactions at the CERN SPS with incident beam energy of 200 AGeV . The lowering of the $`\eta ^{}`$ mass and the partial chiral restoration result in a hole in the effective intercept parameter at low $`m_t`$. This happens even for a modest enhancement of a factor of 3 in the $`\eta ^{}`$ production. Similar results are obtained when using RQMD abundances. See Ref. for further details of the simulation. In addition, $`\lambda _{}(m_t)`$ is calculated using Fritiof abundances with different average flow velocities in Fig. 22. Here it is shown that $`\lambda _{}(m_t)`$ can also be a measure of the average collective flow. In Ref. , an average flow velocity of $`u_t=0.50`$ resulted in an approximately flat, $`m_t`$ independent shape for the effective intercept parameter $`\lambda _{}(m_t)`$ distribution . Calculations using RQMD abudances result in a similar dependence on $`u_t`$, but with slightly higher values of $`\lambda _{}(m_t)`$. This analysis of NA44 S + Pb data indicated no visible sign of $`U_A(1)`$ restoration at SPS energies. In addition, a mean transverse flow of $`u_t0.50`$ in S + Pb reactions was deduced . The suggested $`\lambda _{}`$-hole signal of partial $`U_A(1)`$ restoration cannot be faked in a conventional thermalized hadron gas scenario, as it is not possible to create significant fraction of the $`\eta `$ and $`\eta ^{}`$ mesons with $`p_t0`$ in such a case, Ref. . ## 16 Squeezed Correlations and Spectra for Mass-Shifted Bosons In this section, let me follow the lines of Refs to show that novel back-to-back correlations (BBC) arise for thermal ensembles of squeezed bosonic states associated with medium-modified mass shifts. It was observed in Ref. , that the strength of the BBC could become unexpectedly large in heavy ion collisions, and may thus provide an experimentally observable signal of boson modification in hot and dense matter. Consider, in the rest frame of matter, the following model Hamiltonian, $$H=H_0\frac{1}{2}d^3𝐱d^3𝐲\varphi (𝐱)\delta M^2(𝐱𝐲)\varphi (𝐲),$$ (170) where $`H_0`$ is the asymptotic Hamiltonian, $$H_0=\frac{1}{2}d^3𝐱\left(\dot{\varphi }^2+|\varphi |^2+m_0^2\varphi ^2\right).$$ (171) The scalar field $`\varphi (𝐱)`$ in this Hamiltonian, $`H`$, corresponds to quasi-particles that propagate with a momentum-dependent medium-modified effective mass, which is related to the vacuum mass, $`m_0`$, via $$m_{}^2(|𝐤|)=m_0^2\delta M^2(|𝐤|).$$ The mass shift is assumed to be limited to long wavelength collective modes: $$\delta M^2(|𝐤|)m_0^2\text{if}|𝐤|>\mathrm{\Lambda }_s.$$ The invariant single-particle and two-particle momentum distributions are given as: $`N_1(𝐤_1)`$ $`=`$ $`\omega _{𝐤_1}{\displaystyle \frac{d^3N}{d𝐤_1}}=\omega _{𝐤_1}a_{𝐤_1}^{}a_{𝐤_1}^{},`$ (172) $`N_2(𝐤_1,𝐤_2)`$ $`=`$ $`\omega _{𝐤_1}\omega _{𝐤_2}a_{𝐤_1}^{}a_{𝐤_2}^{}a_{𝐤_2}^{}a_{𝐤_1}^{},`$ (173) $`a_{𝐤_1}^{}a_{𝐤_2}^{}a_{𝐤_2}^{}a_{𝐤_1}^{}`$ $`=`$ $`a_{𝐤_1}^{}a_{𝐤_1}^{}a_{𝐤_2}^{}a_{𝐤_2}^{}+a_{𝐤_1}^{}a_{𝐤_2}^{}a_{𝐤_2}^{}a_{𝐤_1}^{}`$ (174) $`+a_{𝐤_1}^{}a_{𝐤_2}^{}a_{𝐤_2}^{}a_{𝐤_1}^{},`$ where $`a_𝐤`$ is the annihilation operator for asymptotic quanta with four-momentum $`k^\mu =(\omega _𝐤,𝐤)`$, $`\omega _𝐤=\sqrt{m^2+𝐤^2}`$ and the expectation value of an operator $`\widehat{O}`$ is given by the density matrix $`\widehat{\rho }`$ as $`\widehat{O}=\mathrm{Tr}\widehat{\rho }\widehat{O}`$. Eq.(174) has been derived as a generalization of Wick’s theorem for locally equilibriated (chaotic) systems in Ref. . The chaotic and squeezed amplitudes were introduced as $`G_c(1,2)`$ $`=`$ $`\sqrt{\omega _{𝐤_1}\omega _{𝐤_2}}a_{𝐤_1}^{}a_{𝐤_2}^{},`$ (175) $`G_s(1,2)`$ $`=`$ $`\sqrt{\omega _{𝐤_1}\omega _{𝐤_2}}a_{𝐤_1}^{}a_{𝐤_2}^{}.`$ (176) In most situations, the chaotic amplitude, $`G_c(1,2)G(1,2)`$ is dominant, and carries the Bose–Einstein correlations, while the squeezed amplitude, $`G_s(1,2)`$ vanishes. ### 16.1 Mass modification in a homogenous heat bath The terms involving $`G_s(1,2)`$ become non-negligible when mass shift becomes non-vanishing, i.e. $`\delta M^2(𝐤)0`$. Given such a mass shift, the dispersion relation is modified to $`\mathrm{\Omega }_𝐤^2=\omega _𝐤^2\delta M^2(𝐤)`$, where $`\mathrm{\Omega }_𝐤`$ is the frequency of the in-medium mode with momentum $`𝐤`$. The annihilation operator for the in-medium quasi-particle $`b_𝐤`$, and that of the asymptotic field, $`a_𝐤`$, are related by a Bogolyubov transformation : $$a_{𝐤_1}^{}=c_{𝐤_1}^{}b_{𝐤_1}^{}+s_{𝐤_1}^{}b_{𝐤_1}^{}C_1^{}+S_1^{},$$ (177) where $`c_𝐤=\mathrm{cosh}[r_𝐤]`$, $`s_𝐤=\mathrm{sinh}[r_𝐤]`$ and $`r_𝐤`$ reads as $$r_𝐤=\frac{1}{2}\mathrm{log}(\omega _𝐤/\mathrm{\Omega }_𝐤).$$ (178) We introduce the shorthand, $`C_1^{}`$ and $`S_1^{}`$, to simplify later notation. As the Bogolyubov is a squeezing transformation, let us call $`r_𝐤`$ mode dependent squeeze parameter. While it is the $`a`$-quanta that are observed, it is the $`b`$-quanta that are thermalized in medium . Let us consider the average for a globally thermalized gas of the $`b`$-quanta, that is homogenous in volume $`V`$: $$\widehat{\rho }=\frac{1}{Z}\mathrm{exp}\left(\frac{1}{T}\frac{V}{(2\pi )^3}d^3𝐤\mathrm{\Omega }_𝐤b_𝐤^{}b_𝐤^{}\right).$$ (179) When this thermal average is applied, $`G_c(1,2)`$ $`=`$ $`\sqrt{\omega _{𝐤_1}\omega _{𝐤_2}}\left[C_1^{}C_2^{}+S_1^{}S_2^{}\right],`$ (180) $`G_s(1,2)`$ $`=`$ $`\sqrt{\omega _{𝐤_1}\omega _{𝐤_2}}\left[S_1^{}C_2^{}+C_1^{}S_2^{}\right].`$ (181) If this thermal $`b`$ gas freezes out suddenly at some time at temperature $`T`$, the observed single $`a`$-particle distribution takes the following form: $`N_1(𝐤)`$ $`=`$ $`{\displaystyle \frac{V}{(2\pi )^3}}\omega _𝐤n_1(𝐤),`$ (182) $`n_1(𝐤)`$ $`=`$ $`|c_𝐤^{}|^2n_𝐤^{}+|s_𝐤|^2(n_𝐤+1),`$ (183) $`n_𝐤`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{exp}(\mathrm{\Omega }_𝐤/T)1}}.`$ (184) This spectrum includes a squeezed vacuum contribution in addition to the mass-modified thermal spectrum. In this homogeneous limiting case, the two particle correlation function is unity except for the parallel (HBT) and antiparallel (BBC) cases: $`C_2(𝐤,𝐤)`$ $`=`$ $`2,`$ (185) $`C_2(𝐤,𝐤)`$ $`=`$ $`1+{\displaystyle \frac{|c_𝐤^{}s_𝐤^{}n_𝐤^{}+c_𝐤^{}s_𝐤^{}(n_𝐤^{}+1)|^2}{n_1(𝐤)n_1(𝐤)}}.`$ (186) The dynamical correlation due to the two mode squeezing associated with mass shifts is therefore back-to-back, as first pointed out in Ref. . The strength of the HBT correlations remains 2 for identical momenta. It follows from Eq. (186) that the intercept of the BBC is unlimited from above: $`1C_2(𝐤,𝐤)<\mathrm{}`$. As $`|𝐤|\mathrm{}`$, $`C_2(𝐤,𝐤)1+1/|s_𝐤|^21+1/n_1(𝐤)\mathrm{}`$. Hence, at large values of $`|𝐤|`$, the particle production is dominated by that of back-to-back correlated pairs for any non-vanishing value of the in-medium mass shifts . ### 16.2 Suppression by finite duration of emission To describe a more gradual freeze-out, the probability distribution $`F(t_i)`$ of the decay times $`t_i`$ is introduced. The sudden approximation is recovered in the $`F(t_i)=\delta (t_it_0)`$ limiting case. The time evolution of the operators is given by $`a_𝐤(t)=a_𝐤(t_i)\mathrm{exp}[i\omega _𝐤(tt_i)]`$. This leads to a suppression of BBC as $$C_2(𝐤,𝐤)=1+|\stackrel{~}{F}(\omega _𝐤+\omega _𝐤)|^2\frac{|c_𝐤^{}s_𝐤^{}n_𝐤^{}+c_𝐤^{}s_𝐤^{}(n_𝐤^{}+1)|^2}{n_1(𝐤)n_1(𝐤)}.$$ (187) Here $`\stackrel{~}{F}(\omega )=𝑑tF(t)\mathrm{exp}(i\omega t)`$, so for an exponential decay, $`F(t)=\mathrm{\Theta }(tt_0)\times `$ $`\mathrm{\Gamma }\mathrm{exp}[\mathrm{\Gamma }(tt_0)]`$ the suppression factor is $$|\stackrel{~}{F}(\omega _𝐤+\omega _𝐤)|^2=1/[1+(\omega _𝐤+\omega _𝐤)^2/\mathrm{\Gamma }^2].$$ (188) In the adiabatic limit, $`\mathrm{\Gamma }0`$, this factor suppresses completely the BBC, while in the sudden approximation, $`\mathrm{\Gamma }\mathrm{}`$, the full strenght of the BBC is preserved. For a typical $`\delta t=\mathrm{}/\mathrm{\Gamma }=2`$ fm/c decay time, and for BBC of $`\varphi `$ mesons with $`m_{}=0.61.4`$ GeV, this suppression factor is about 0.001, which decreases the BBC of $`\varphi `$ mesons from the scale of 2000 to 2, the scale of the HBT correlations. This emphasizes the enormous strength of the BBC . The formalism to evaluate the BBC for locally thermalized, expanding sources was also developed, see Ref. for greater details. As the Bogolyubov transformation always mixes particles with anti-particles, the above considerations hold only for particles that are their own anti-particles, e.g. the $`\varphi `$ meson and $`\pi ^0`$. The extension to particle–anti-particle correlations is straightforward. Let $`+`$ label particles, $``$ antiparticles if antiparticle is different from particle, let $`0`$ label both particle and antiparticle if they are identical. The non-trivial correlations from mass modification for pairs of $`(++)`$, $`(+)`$ and $`(00)`$ type read as follows: $`C_2^{++}(𝐤_1,𝐤_2)`$ $`=`$ $`1+{\displaystyle \frac{|G_c(1,2)|^2}{G_c(1,1)G_c(2,2)}},`$ (189) $`C_2^+(𝐤_1,𝐤_2)`$ $`=`$ $`1+{\displaystyle \frac{|G_s(1,2)|^2}{G_c(1,1)G_c(2,2)}},`$ (190) $`C_2^{00}(𝐤_1,𝐤_2)`$ $`=`$ $`1+{\displaystyle \frac{|G_c(1,2)|^2}{G_c(1,1)G_c(2,2)}}+{\displaystyle \frac{|G_s(1,2)|^2}{G_c(1,1)G_c(2,2)}},`$ (191) where we assume that mass modifications of particles and anti-particles are the same as happens at vanishing baryon density. This theory of particle correlations and spectra for bosons with in-medium mass shifts predicts huge back-to-back correlations of $`\varphi ^0,\varphi ^0`$ and $`\mathrm{K}^+,\mathrm{K}^{}`$ meson pairs . These BBC could become observable at the STAR and PHENIX heavy ion experiments at RHIC , and could be looked for in present CERN SPS experiments. Further model calculations are required to study the mass-shift effects on realistic source models. ## 17 A Pion-Laser Model and Its Solution In high energy heavy ion collisions hundreds of bosons are created in the present CERN SPS reactions when Pb + Pb reactions are measured at 160 AGeV laboratory bombarding energy. At the RHIC accelerator, thousands of pions could be produced in a unit rapidity interval . If the number of pions in a unit value of phase-space is large enough these bosons may condense into the same quantum state and a pion laser could be created . In this section a consequent quantum mechanical description of multi-boson systems is reviewed, based on properly normalized projector operators for overlapping multi-particle wave-packet states and a model of stimulated emission, following the lines of Refs . One of the new analytic results is that multi-boson correlations generate momentum-dependent radius and intercept parameters even for static sources, as well as induce a special directional dependence of the correlation function. This is to be contrasted to the simplistic but very frequently invoked picture of Eq. (3), where sources without expansion correspond to a correlation function that depends only on the relative momentum, but not on the mean momentum of the particle pairs. A solvable density matrix of a generic quantum mechanical system is $`\widehat{\rho }`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{p_n}{𝒩(n)}}{\displaystyle \underset{i=1}{\overset{n}{}}d\alpha _i\rho _1(\alpha _i)\left(\underset{\sigma ^{(n)}}{}\underset{k=1}{\overset{n}{}}\alpha _k|\alpha _{\sigma _k}\right)|\alpha _1,\mathrm{},\alpha _n\alpha _1,\mathrm{},\alpha _n|}.`$ Here the index $`n`$ characterizes sub-systems with particle number fixed to $`n`$, the multiplicity distribution is prescribed by the set of $`\left\{p_n\right\}_{n=0}^{\mathrm{}}`$, normalized as $`_{n=0}^{\mathrm{}}p_n=1`$. The density matrixes are normalized as $`\text{Tr}\widehat{\rho }=1`$ and $`\text{Tr}\widehat{\rho }_n=1`$. The states $`|\alpha _1,\mathrm{},\alpha _n`$ denote properly normalized $`n`$-particle wave-packet boson states: $$|\alpha _1,\mathrm{},\alpha _n=\left(\underset{\sigma ^{(n)}}{}\underset{i=1}{\overset{n}{}}\alpha _i|\alpha _{\sigma _i}\right)^{\frac{1}{2}}\alpha _n^{}\mathrm{}\alpha _1^{}|0.$$ (193) Here $`\sigma ^{(n)}`$ denotes the set of all the permutations of the indexes $`\{1,2,\mathrm{},n\}`$ and the subscript sized $`\sigma _i`$ denotes the index that replaces the index $`i`$ in a given permutation from $`\sigma ^{(n)}`$. The wave-packet creation operators, $`\alpha _i^{}`$ create the normalized single-particle states $`|\alpha _i=\alpha _i^{}|0`$, with $`\alpha _i|\alpha _i=1`$. The $`\alpha _i=(\xi _i,\pi _i,\sigma _i,t_i)`$ stands for a given value of the parameters of a single-particle wave-packet: the mean coordinate, the mean momentum, the width of the wave-packet in coordinate space and the time of the production. The distribution function $`\rho _1(\alpha _i)`$ provides the probability distribution for a given value of the wave-packet parameters. For simplicity, we assume a static source at rest, uniform wave-packet widths and simultaneous production, $`\sigma _i=\sigma `$ and $`t_i=t_0`$. A Gaussian distribution of the centers of wave-packets is also assumed: in the coordinate space, the distribution of $`\xi _i`$ is a characterized with a radius $`R`$, while in the momentum-space, the centers of wave-packets $`\pi _i`$ are assumed to have a non-relativistic Boltzmann distribution corresponding to a temperature $`T`$ and mass $`m`$. The coefficient of proportionality, $`𝒩(n)`$, can be determined from the normalization condition. The density matrix given in Eq. (LABEL:e:dtrick) describes a quantum-mechanical wave-packet system with induced emission, and the amount of the induced emission is controlled by the overlap of the $`n`$ wave-packets , yielding a weight in the range of $`[1,n!]`$. Although it is very difficult numerically to operate with such a wildly fluctuating weight, the problem of overlapping multi-boson wave-packets with stimulated emission was reduced in Refs to an already discovered “ring”-algebra of permanents for plane-wave outgoing states , with modified source parameters . Assuming a non-relativistic, non-expanding Gaussian source at rest, and a Poisson multiplicity distribution $`p_n^{(0)}`$ in the rare gas limiting case: $$p_n^{(0)}=\frac{n_0^n}{n!}\mathrm{exp}(n_0),$$ (194) the ring-algebra was reduced in Ref. to a set of recurrences, which reduced the complexity of the problem from the numerically impossible $`n!`$ to the numerically easy $`n^2`$. These recurrences were solved analytically in Refs , further reducing the complexity of the problem to $`n^0`$, and yielding analytic insight to the behavior of the multi-boson symmetrization effects. The probability of events with fixed multiplicity $`n`$, the single-particle and the two-particle momentum distribution in such events are given as $`p_n`$ $`=`$ $`\omega _n\left({\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}\omega _k\right)^1,`$ (195) $`N_1^{(n)}(𝐤_1)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \frac{\omega _{ni}}{\omega _n}}G_i(1,1),`$ (196) $`N_2^{(n)}(𝐤_1,𝐤_2)`$ $`=`$ $`{\displaystyle \underset{l=2}{\overset{n}{}}}{\displaystyle \underset{m=1}{\overset{l1}{}}}{\displaystyle \frac{\omega _{nl}}{\omega _n}}\left[G_m(1,1)G_{lm}(2,2)+G_m(1,2)G_{lm}(2,1)\right],`$ where $`\omega _n=p_n/p_0`$. Averaging over the multiplicity distribution $`p_n`$ yields the inclusive spectra as $`G(1,2)`$ $`=`$ $`{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}G_n(1,2),`$ (198) $`N_1(𝐤_1)`$ $`=`$ $`{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}p_nN_1^{(n)}(𝐤_1)=G(1,1),`$ (199) $`N_2(𝐤_1,𝐤_2)`$ $`=`$ $`G(1,1)G(2,2)+G(1,2)G(2,1).`$ (200) Let us introduce the following auxiliary quantities: $`\gamma _\pm ={\displaystyle \frac{1}{2}}\left(1+x\pm \sqrt{1+2x}\right),`$ $`x=R_e^2\sigma _T^2,`$ (201) $`\sigma _T^2=\sigma ^2+2mT,`$ $`R_e^2=R^2+{\displaystyle \frac{mT}{\sigma ^2\sigma _T^2}},`$ (202) The general analytical solution of the model is given through the generating function of the multiplicity distribution $`p_n`$ $`G(z)`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}p_nz^n=\mathrm{exp}\left({\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}C_n(z^n1)\right),`$ (203) where $`C_n`$ is introduced as $$C_n=\frac{1}{n}d^3𝐤_1G_n(1,1)=\frac{n_0^n}{n}\left[\gamma _+^{\frac{n}{2}}\gamma _{}^{\frac{n}{2}}\right]^3.$$ (204) The general analytic solution for the functions $`G_n(1,2)`$ is given as: $`G_n(1,2)`$ $`=`$ $`j_n\mathrm{exp}\left\{{\displaystyle \frac{b_n}{2}}\left[\left(\gamma _+^{\frac{n}{2}}𝐤_1\gamma _{}^{\frac{n}{2}}𝐤_2\right)^2+\left(\gamma _+^{\frac{n}{2}}𝐤_2\gamma _{}^{\frac{n}{2}}𝐤_1\right)^2\right]\right\},`$ (205) $`j_n`$ $`=`$ $`n_0^n\left[{\displaystyle \frac{b_n}{\pi }}\right]^{\frac{3}{2}},b_n={\displaystyle \frac{1}{\sigma _T^2}}{\displaystyle \frac{\gamma _+\gamma _{}}{\gamma _+^n\gamma _{}^n}}.`$ (206) The detailed proof that the analytic solution to the multi-particle wave-packet model is indeed given by the above equations is described in Ref. . The representation of Eq. (203) indicates that the quantities $`C_n`$-s are the so called combinants of the probability distribution of $`p_n`$ and in this case their explicit form is known for any set of model parameters. In the generator functional formalism of multi-particle production, the combinants can be introduced in general as the integrals of the exclusive correlation functions . The form of the multiplicity distribution, given by Eqs (203,204) does not correspond to the multiplicity distributions described in standard textbooks of mathematical statistics, e.g. Ref. . It has the very interesting property, that the probability distribution simultaneously corresponds to an infinite convolution of independently distributed clusters of particle singlets, pairs, triplets and higher-order $`n`$-tuples, as well as to an infinite convolution of strongly correlated Bose–Einstein distribution of particle singlets, pairs, triplets etc. As far as I know, this is a new type of physically motivated discrete distribution in the theory of probability and statistics. The large $`n`$ behavior of $`p_n`$ depends on the ratio of $`n_0/n_c`$, where the critical value of $`n_0`$ is $`n_c=\gamma _+^{3/2}`$, . If $`n_0<n_c`$, one finds $`n(n1)>n^2`$, a super-Poissonian multiplicity distribution, and a chaotic or thermal behavior of the inclusive correlations, $`C_2(𝐤,𝐤)=2`$. If $`n_0n_c`$, the multiplicity distribution, the inclusive spectra and the inclusive correlations become mathematically undefined, but the exlusive quantities remain finite for any fixed value of $`n`$. To calculate inclusive observables, a regularization has to be introduced similarly to the description of Bose–Einstein condensation of massive quanta in the limit of $`\mu m`$ in standard statistical mechanics. Highly condensed limiting case. In Refs we have related the divergence for $`n_0n_c`$ of the mean multiplicity $`n`$ to the onset of a generalized type of Bose–Einstein condensation of the wave-packets to the wave-packet state with the smallest energy. Note that the onset of Bose–Einstein condensation happens in the limit when $`p_n/p_{n+1}1`$, which happens if $`n_0n_c`$ from below , and this limiting case formally corresponds to an “infinite temperature” case — if the finite slope parameters of the $`N_1^{(n)}(𝐤)`$ single-particle distributions in exclusive events are not taken into account and the concept of the temperature is inferred only from the number distribution. In a physical situation, the total number of pions is limited: $`nn_E=E_{\mathrm{tot}}/E_0`$, where $`E_0`$ is the energy of the wave-packet with the smallest energy (including the mass $`m`$). Thus, energy conservation induces a cut-off in the number of pions, that has to be taken into account explicitly . Such a cut in the multiplicity distribution can be straightforwardly implemented, as the basic building block, the fully symmetrized $`n`$-particle invariant momentum distribution in events with exactly $`n`$ particles is always finite for every fixed values of $`n`$, similarly to the bosonic enhancement factor $`\omega _n`$. At $`n_0=n_c`$, the series $`S_n=_{j=0}^n\omega _j`$ changes from a convergent to a divergent one. After the regularization of the model, by assigning a zero probability to multiplicities greater that $`n_E`$, one can show that for $`n_0>n_c`$ a Bose–Einstein condensation develops more and more with increasing values of $`n_0`$. Utmost care is required when evaluating the results published in the literature regarding the nature of coherence and Bose–Einstein condensation in the pion-laser model: some papers identify the “Bose–Einstein condensation” with the “infinitely hot” $`n_0=n_c`$ limiting case. At this point, however, the condensate just appears with non-zero probability (and one has to introduce the cut multiplicity distribution to describe it with a $`p_{n_E}>0`$), but the number of quanta in the condensate is rather small, $`p_{n_E}1/n_E`$ at the $`n=n_c`$ critical point. The nature of the Bose–Einstein condensation was discussed and clarified in Ref. , where it was shown that the condensate will fully develop and dominate the density matrix in the $`R0`$ and $`T0`$ simultaneous limiting cases, confirming the intuitive picture that Bose–Einstein condensation happens in very cold and very small systems. In the highly condensed limiting case, the multiplicity distribution of the produced particles will be sub-Poissonian, a very narrow, cut power-law distribution that increases with $`n`$ as $$p_n\left(\frac{n_0}{n_c}\right)^n\mathrm{\Theta }(n_En)\text{for}n_0n_c,$$ (207) and vanishes after $`n>n_E`$. In the limit when the number of particles in the condensate is very large, the exclusive and the inclusive correlation functions become unity , $`C(𝐤_1,𝐤_2)`$ $`=`$ $`C^{(n)}(𝐤_1,𝐤_2)=1`$ (the highly condensed limiting case, $`n_0n_c`$). By definition, the above equalities imply optical coherence in the highly condensed limiting case . It is worthwhile to emphasize, that optical coherence is not to be confused with the appearance of the coherent states of the annihilation operator . Instead of being an eigenstate of the annihilation operator, the fully developed Bose–Einstein condensate is an eigenstate of the creation operator, with zero eigenvalue. This is due the cutoff induced by the conservation of energy: it is not possible to add one more pion to the condensate if already all the pions allowed by the constraints are in the condensate. Rare gas limiting case. In contrast, the large source sizes or large effective temperatures correspond to a rare Boltzmann gas, the $`x1`$ limiting case. The general analytical solution of the model becomes particularly simple in this limiting case. The leading order multiplicity distribution can be found from Eqs (203,204), corresponding to independently distributed particles with a small admixture of independently distributed particle pairs : $$p_n=\frac{n_0^n}{n!}\mathrm{exp}(n_0)\left[1+\frac{n(n1)n_0^2}{2(2x)^{\frac{3}{2}}}\right].$$ (209) The mean multiplicity, the factorial cumulant moments of the multiplicity distribution, the inclusive and exclusive momentum distributions were obtained to leading order terms in $`1/x`$ in Ref. . Figure 24 indicates that the radius parameter of the exclusive correlation function becomes mean momentum momentum-dependent, even for static sources! This genuine multi-particle symmetrization effect is more pronounced for higher values of the fixed multiplicity $`n`$, in contrast to the momentum dependence of $`\lambda _𝐊`$ that is independent of $`n`$ . One finds that multi-boson symmetrization effects lead to the development of a Bose–Einstein condensate. Before the onset of the Bose–Einstein condensation, the stimulated emission becomes significant in the low momentum modes earlier than in the high momentum modes. This is the reason, why even the exclusive correlation functions develop a mean momentum dependent radius parameter, as well as a direction-dependent radius component and a mean momentum dependent intercept parameter. ## 18 Summary and Outlook In this review, new kind of similarities were highlighted between stellar astronomy and intensity interferometry in high energy physics. The model independent characterization of short-range correlation was given in terms of expansions in complete orthonormal sets of polinomials, the core/halo model and the recently found Coulomb wave-function correction method was reviewed, for Bose–Einstein $`n`$-particle correlations. The invariant Buda–Lund (BL) parameterization of Bose–Einstein correlation functions was derived in a general form, and compared to the Bertsch–Pratt and the Yano–Koonin–Podgoretskii parameterization in the particular Gaussian limiting case. The Buda–Lund hydrodynamical parameterization, BL-H was fitted to hadron–proton and Pb + Pb collisions at CERN SPS energies. Larger mean freeze-out proper-times and larger transverse radii were found in the Pb + Pb reactions. Although the central values of freeze-out temperatures were rather similar in both reactions, the transverse temperature gradient is larger while the transversal flow is smaller in h + p reactions, than in the Pb + Pb system. This resulted in different shapes for the transverse density profiles, that were approximately reconstructed assuming the applicability of a new family of solutions to fireball hydrodynamics . Although Pb + Pb reactions were found to be rather homogenous expanding fireballs, the h + p reactions were found to be similar to a cold and expanding ring of fire when viewed in the transverse plane. The central freeze-out temperature is about $`T_0=140`$ MeV in both reactions, the surface temperature after the emission of particles is over seems to be also similar, about $`T_s=82`$ MeV, the duration of the particle emission is also about $`\mathrm{\Delta }\overline{\tau }1.5`$ fm in both cases. Inspecting the results of a non-relativistic version of BL-H to <sup>40</sup>Ar + <sup>197</sup>Au proton and neutron correlations and spectra, an indirect signal was observed for the formation of a shell of fire, made of protons, while the neutrons seems to come from an ordinary fireball. The hydrodynamics of cooling and expanding shells of low energy heavy ion reactions was shown to be similar to that of spherical planetary nebulae, indicating a new connection between stellar astronomy and particle interferometry in heavy ion physics. Another similarity between stellar astronomy and high energy physics was discussed in terms of the interferometry of binary sources: the binary stars in stellar astronomy create oscillations in the HBT effect similarly to the oscillations that were shown to exist in the Buda–Lund type of hydrodynamical parameterization in heavy ion physics and to the expected oscillations of pion correlations in particle interferometry in W<sup>+</sup>W<sup>-</sup> decays at LEP2. The first positive evidence for the existence of such binary sources in heavy ion physics seems to be the recent measurement of oscillating proton–proton correlations by the NA49 Collaboration , which may be a consequence of the existence of the two maxima in the proton rapidity distribution and the attractive final-state interactions of protons, that enhance the large $`Q`$ part of the pp intensity correlation function and make these oscillations clearly visible. The question of non-Gaussian oscillations of three-dimensional Bose–Einstein correlation functions in heavy ion physics has not yet been experimentally investigated. I think it is time to start the experimental search for non-Gaussian structures in multi-dimensional Bose–Einstein correlation functions in high energy heavy ion and particle physics. I hope that experiments will decide to publish in the future not only the (Gaussian) fit parameters of (multi-dimensional) Bose–Einstein correlation functions, but, most importantly, the measured data points and the corresponding the error bars. It was shown already in Refs , that the reconstruction of the space-time picture of the particle emission: the extraction of density, flow and temperature profiles requires the simultaneous analysis of the double-differential single-particle spectra and the momentum-dependent multi-dimensional Bose–Einstein correlation functions. These data sets should be made public in as much detail as possible, including multi-dimensional data and error-bar tables. In the chiral limit, when the up, down and strange quarks become massless and the $`U_A(1)`$ symmetry is fully restored, the mass of the $`\eta ^{}`$ meson vanishes in the $`U_A(1)`$ symmetric, new phase. The appearance of such a phase implies that the intercept parameter of the two-pion correlation function vanishes in the $`p_t150`$ MeV region. In this sense, the transverse mass-dependent intercept parameter $`\lambda _{}(m_t)`$, was interpreted as an effective order parameter of partial $`U_A(1)`$ symmetry restoration . Bosonic mass shifts in medium were shown to result in unlimitedly large back-to-back correlations of the observable boson–anti-boson pairs. Although a finite time suppression factor may reduce the strength of these correlations substantially, the magnitude of the back-to-back correlations is estimated to be observably strong for typical mass shifts and freeze-out time distributions in ultra-relativistic heavy ion collisions. Multi-boson symmetrization effects were shown to generate momentum-dependent radius and intercept parameters even for static sources. The proposed $`\lambda _{}`$-hole signal of the $`U_A(1)`$ symmetry restoration, the new kind of back-to-back correlations and optically coherent, effectively lasing pion sources could be searched for in future in heavy ion experiments at CERN SPS and at RHIC. The oscillations in multi-dimensional Bose–Einstein and Fermi–Dirac correlations could be searched for in e<sup>+</sup>e<sup>-</sup> annihilation experiments at LEP2, as well as in heavy ion collisions at CERN SPS and at RHIC. Note that this paper is a review of particle interferometry before 2000; a substantially shortened version of the present material has been published in Ref. . ## Acknowledgments I would like to thank to my co-authors: M. Asakawa, J. Beier, M. Gyulassy, R. Hakobyan, S. Hegyi, J. Helgesson, D. Kiang, W. Kittel, D. Kharzeev, B. Lörstad, S. Nickersson, A. Ster, S. Vance and J. Zimányi, and to the NA22 Collaboration, for their various contributions to some of the sections in this review. I also would like to thank the Organizers of the NATO School Nijmegen’99 for creating a pleasent atmosphere and an inspiring working environment. I am grateful to Professors Kittel, Hama and Padula for inspiration and for stimulating working environment. This research was supported by the grants Hungarian OTKA T024094, T026435, T029158 and T034269, the US–Hungarian Joint Fund MAKA grant 652/1998, NWO–OTKA N025186, OMFB–Ukraine S&T grant 45014 and FAPESP 98/2249-4 and 99/09113-3.