diff --git "a/T9E2T4oBgHgl3EQfCga9/content/tmp_files/2301.03615v1.pdf.txt" "b/T9E2T4oBgHgl3EQfCga9/content/tmp_files/2301.03615v1.pdf.txt" new file mode 100644--- /dev/null +++ "b/T9E2T4oBgHgl3EQfCga9/content/tmp_files/2301.03615v1.pdf.txt" @@ -0,0 +1,3726 @@ +Study of the de Almeida-Thouless (AT) line in the one-dimensional diluted power-law +XY spin glass +Bharadwaj Vedula,1 M. A. Moore,2 and Auditya Sharma1 +1Department of Physics, Indian Institute of Science Education and Research, Bhopal, Madhya Pradesh 462066, India +2Department of Physics and Astronomy, University of Manchester, Manchester M13 9PL, United Kingdom +(Dated: January 11, 2023) +We study the AT line in the one-dimensional power-law diluted XY spin glass model, in which +the probability that two spins separated by a distance r interact with each other, decays as 1/r2σ. +Tuning the exponent σ is equivalent to changing the space dimension of a short-range model. We +develop a heat bath algorithm to equilibrate XY spins; using this in conjunction with the standard +parallel tempering and overrelaxation sweeps, we carry out large scale Monte Carlo simulations. +For σ = 0.6, which is in the mean-field regime above six dimensions – it is similar to being in 10 +dimensions – we find clear evidence for an AT line. For σ = 0.75 and σ = 0.85, which are in the +non-mean-field regime and similar to four and three dimensions respectively, our data is like that +found in previous studies of the Ising and Heisenberg spin glasses when reducing the temperature +at fixed field. For σ = 0.75, there is evidence from finite size scaling studies for an AT transition +but for σ = 0.85, the evidence for a transition is non-existent. We have also studied these systems +at fixed temperature varying the field and discovered that at both σ = 0.75 and at σ = 0.85 there +is evidence of an AT transition! Confusingly, the correlation length and spin glass susceptibility as +a function of the field are both entirely consistent with the predictions of the droplet picture and +hence the non-existence of an AT line. In the usual finite size critical point scaling studies used to +provide evidence for an AT transition, there is seemingly good evidence for an AT line at σ = 0.75 +for small values of the system size N, which is strengthening as N is increased, but for N > 2048 the +trend changes and the evidence then weakens as N is further increased. We have also studied with +fewer bond realizations the system at σ = 0.70, which is the analogue of a system with short-range +interactions just below six dimensions, and found that it is similar in its behavior to the system at +σ = 0.75 but with larger finite size corrections. The evidence from our simulations points to the +complete absence of the AT line in dimensions outside the mean-field region and to the correctness +of the droplet picture. Previous simulations which suggested there was an AT line can be attributed +to the consequences of studying systems which are just too small. The collapse of our data to the +droplet scaling form is poor for σ = 0.75 and to some extent also for σ = 0.85, when the correlation +length becomes of the order of the length of the system, due to the existence of excitations which +only cost a free energy of O(1), just as envisaged in the TNT picture of the ordered state of spin +glasses. However, for the case of σ = 0.85 we can provide evidence that for larger system sizes, +droplet scaling will prevail even when the correlation length is comparable to the system size. +I. +INTRODUCTION +While the spin glass problem at mean-field level is now +well-understood [1], questions remain as to the nature of +the ordered state in three dimensional spin glasses. A +key question is whether the ordered phase of real spin +glasses has the broken replica symmetry features found +in mean-field theory. +This question is most easily an- +swered by finding whether on application of a magnetic +field hr there is a line, the so-called de Almeida Thou- +less (AT) line [2], below which in the hr − T plane there +is replica symmetry breaking. This line exists at mean- +field level (see Fig. 1) and its possible existence in three +dimensions can be studied experimentally and with sim- +ulations. Simulational studies of the existence of replica +symmetry breaking within the zero-field spin glass state +itself are plagued by finite size effects: it is expected that +the difference between the predictions of droplet scaling +and those of replica symmetry breaking will only become +visible for very large systems (for a review see [3]). A +recent review of simulations, including studies of the ex- +istence of the AT line, can be found in Ref. [4]. +Right from the early days of spin glass studies there +have been doubts raised as to whether the AT line existed +below six dimensions. For example Bray and Roberts [6] +attempted to do an expansion in 6 − ϵ dimensions for +the critical exponents at the AT line but failed to find +a stable fixed point. They suggested that maybe that +indicated that there might be no AT line below six di- +mensions. A renormalization group calculation also gave +indications that the AT line was going away as d → 6 +from above +[7]. As it is difficult to do simulations in +dimensions around 6 to check these speculations, simula- +tors have had to turn instead to one-dimensional models +with long-range power-law interactions. +These models go back to Kotliar, Anderson and Stein +[8], who in turn were inspired by the long-range ferro- +magnet that was studied by Dyson [9, 10]. +The long- +range power-law model has the advantage that by tuning +the power-law exponent σ, one has access to both the +mean-field and the regimes with non-mean-field critical +behavior. However, the full power-law model is expen- +sive for numerics. Fortunately a clever workaround was +introduced by Leuzzi et al. [11] where instead of the in- +arXiv:2301.03615v1 [cond-mat.dis-nn] 9 Jan 2023 + +2 +0.0 +0.2 +0.4 +0.6 +0.8 +1.0 +T/Tc +0.0 +0.2 +0.4 +0.6 +0.8 +1.0 +1.2 +1.4 +hr/J +exact AT +approximate AT +SK data +σ = 0.6 (T∗ data) +σ = 0.6 (h∗ data) +Figure 1. The AT line. The solid line is the exact AT line +for the SK model, calculated as in Ref. [5]. The dashed line +is the approximation to it of Eq. (8). We have marked on +the diagram the results of our simulations on the SK model, +which were done to check our Monte Carlo procedures. The +points in red and green are the results of our simulations at +σ = 0.6, which lie in the mean-field region. The data on the +horizontal axis for σ = 0.6 are normalized to the transition +temperature Tc in zero field for that value of σ. For the XY +SK model the AT line goes to infinity as T → 0. +teractions falling off as a power law, it is the probability +of there being a bond between two spins that falls off as +a power law. The fewer bonds in the model means that a +significantly smaller computational cost is involved, thus +allowing for the simulation of larger system sizes. +While the vast literature on spin glasses is mostly fo- +cussed on Ising spins [11–16], there has been a revival of +interest in classical m-component vector spin glass mod- +els [5, 17–26] in the last decade or so. The XY model +has m = 2 and the Heisenberg model has m = 3. One +of the triggers for this revival has been the finding that +the infinite-range vector spin glass exhibits an AT line +provided a magnetic field that is random in all the com- +ponent directions is applied [5]. Furthermore analytical +studies of the AT transition in m-vector models shows +that the field theory of these AT transitions is that of the +Ising spin glass [5]. Thus it has become possible to study +the question of whether or not an Ising AT transition ex- +ists in various dimensions by studying one-dimensional +vector spin glasses with long-range interactions [18]! +In this paper, we study the one-dimensional diluted +XY spin glass subjected to a random vector magnetic +field, with the aid of large scale Monte Carlo simula- +tions. While Monte Carlo simulations are a time-tested +tool for the study of phase transitions in spin glasses, the +exorbitant cost of equilibration makes them rather chal- +lenging in practice. It has been argued that vector spins +tend to equilibrate faster compared to Ising spins [27], be- +cause of the soft nature of the spins involved, even though +the presence of more components adds to the cost. The +Heisenberg spin glass [5, 17–19, 27–32] has been the pop- +ular vector spin to have been considered, because of the +availability of the heatbath algorithm [28], which works +very efficiently to equilibrate it. The XY spin glass is +less effectively handled by the heatbath algorithm [33] +because of the technicalities involved in inverting a prob- +ability distribution for which a simple closed form ex- +pression is unavailable in the XY case. In this paper, we +develop a method, which is outlined in Appendix A to +perform this inversion numerically with the hope of ben- +efiting from the vector nature of XY spins, while simul- +taneously reducing the components to as small a number +as possible. +The improved algorithm yields mixed fruits. The gains +from the reduced number of components seems to be +largely counterbalanced by the additional resources con- +sumed by the numerical inversion. However, with the aid +of extensive computational power, we are able to access +system sizes comparable to those in the corresponding +study with Heisenberg spins. Our findings for the XY +diluted spin glass closely mimic those obtained for the +Heisenberg version of the same model [5, 17, 18] and the +Ising spin glass in three dimensions [34] when we investi- +gate crossing the possible AT line by varying T at fixed +values of hr. In the mean-field regime, (we studied here +the case of σ = 0.6, which corresponds to 10 dimen- +sions, which is above the upper critical dimension of 6), +there is clear evidence of an Almeida-Thouless line (see +Sec. V A). There is rather weak evidence for an Almeida- +Thouless line for σ = 0.85 using the commonly employed +finite size critical point scaling methods of analysis (see +Sec. V C). At this value of σ, our system should be simi- +lar to the Edwards-Anderson model in three dimensions +with short-range interactions. For the in-between case +at σ = 0.75 which lies in the non-mean-field regime, but +closer to the mean-field boundary at σ = 2/3, our data +do provide stronger evidence for a phase transition in +the presence of small magnetic fields than at σ = 0.85. +However, by varying the magnetic field hr at fixed tem- +perature T we find in Sec. V that at both σ = 0.75 and +at σ = 0.85 there is quite decent evidence for an AT +line. Confusingly, the field dependence of the correlation +length is very well-described by the Imry-Ma prediction +of the droplet picture, which implies the complete ab- +sence of the AT transition! In the droplet picture the +correlation length in a field remains finite and only di- +verges as hr → 0. However, when this correlation length +becomes comparable to the system size L the Imry-Ma +formula needs to be modified and we give in Sec. +VI +a scaling form for this modification. +It is based upon +the usual finite size scaling approach used in studying +critical phenomena, and just as for critical phenomena +we find that there are finite size corrections to this scal- +ing form. In addition to these scaling corrections there + +3 +are corrections which arise when ξSG ∼ L < L∗ which +are of different origin and are connected to TNT effects +[35, 36]. TNT effects arise from droplets of free energy +cost of O(1), which certainly exist in systems whose sizes +L < L∗ [3]. The length scale L∗ is always large and is +expected to diverge as d → 6 or as σ → 2/3. It is only +for the case of σ = 0.85 that we can reach sizes where +TNT effects seem to be getting small. These matters are +discussed in Sec. VII. +Furthermore, we can use the droplet scaling picture to +explain some of the features of the apparent AT transi- +tion which arise on performing the usual finite size crit- +ical scaling analyses, and show that these are the conse- +quence of not studying large enough systems. Unfortu- +nately these arguments will only become compelling for +system sizes which we cannot reach. Our chief evidence +for the droplet picture is its very successful prediction +of the correlation length as a function of the field in the +region when finite size and TNT effects are unimportant. +Our claim that the evidence favors the absence of the +AT line for values of σ outside the mean-field region is +consistent with the attempt [21] to calculate the AT field +at T = 0 using an expansion in 1/m. This indicated that +as d → 6 from above in the Edwards-Anderson model, the +AT field would go to zero, implying the absence of the AT +line below 6 dimensions (which in the one-dimensional +long-range model corresponds to 2/3 < σ < 1). +For +σ > 1 there is no finite temperature spin glass phase. +The plan of this paper is as follows. In Sec. II we de- +scribe the model in detail. In Sec. III we describe the +quantities which were studied in our Monte Carlo simu- +lations, the details of which are given in Appendix A. Our +data is analysed in Sec. V on the assumption that there +is an AT transition, while in Sec. VI the data is analysed +according to droplet scaling assumptions. In Sec. VII we +discuss the effect of TNT behavior on our results. Finally +in Sec. VIII we summarize our conclusions. +II. +MODEL HAMILTONIAN +The general Hamiltonian for vector spin glasses is: +H = − +� +⟨i,j⟩ +JijSi · Sj − +� +i +hi · Si , +(1) +where Si is the spin on the ith lattice site (i += +1, 2, . . . , N), which is chosen to be a unit vector. m rep- +resents the number of components of the vector Si. In +this work we concentrate on XY spins, and set m = 2. +The Cartesian components hµ +i (µ = 1, 2) of the on-site +external magnetic field are i.i.d random variables drawn +from a Gaussian distribution of zero mean and variance +h2 +r and satisfy the relation: +� +hµ +i hν +j +� +av = h2 +rδijδµν. +(2) +We use the notation ⟨· · · ⟩ for thermal average and [· · · ]av +for an average over quenched disorder throughout this +paper. +The spins are arranged on a circle so the geometric +distance between a pair of spins (i, j) is given by [16] +rij = N +π sin +� π +N |i − j| +� +, +(3) +which is the length of the chord connecting the ith and +jth spins. The interactions Jij are independent random +variables such that the probability of having a non-zero +interaction between a pair of spins (i, j) falls with the +distance rij between the spins as a power law: +P(Jij) ∝ +1 +r2σ +ij +. +(4) +If the spins i and j are linked the magnitude of the inter- +action between them is drawn from a Gaussian distribu- +tion whose mean is zero and whose standard deviation is +unity, i.e: +[Jij]av = 0 +and +� +J2 +ij +� +av = 1. +(5) +The mean number of non-zero bonds from a site is fixed +to be ˜z (co-ordination number). So, the total number +of bonds among all the spins on the lattice is fixed to +be Nb = N ˜z/2. +When ˜z = 6 this model mimics the +3D simple cubic lattice model and we use this value for +˜z for all the σ values studied. +(For σ = 0 and ˜z = +N −1, the model becomes the infinite-range Sherrington- +Kirkpatrick (SK) model [37]). +To generate the set of interaction pairs [11, 17] (i, j) +with the desired probability we pick a site i randomly and +uniformly and then choose a second site j with probabil- +ity given by: +pij = +r−2σ +ij +� +j̸=i +r−2σ +ij +. +(6) +If the spins at i and j are already connected we repeat +this process until we find a pair of sites (i, j) which have +not been connected. Once we find such a pair of spins, +we connect them with a bond whose strength Jij is a +Gaussian random variable with attributes given by Eq. +(5). We repeat this process exactly Nb times to generate +Nb pairs of interacting spins. +The advantage of the diluted model over the fully con- +nected model is that, in a fully connected model, there +are N(N − 1)/2 interactions. The ratio of the number of +interactions of the diluted model to the fully connected +model is ˜z/N which is a very small value as N becomes +large. Hence it is possible to go to much larger system +sizes with a diluted model as compared to a fully con- +nected model. +At zero-field, the mean-field spin glass transition tem- +perature for the m-component vector spin glass is given +by [17, 18, 38] +T MF +c += 1 +m +� +�� +j +� +J2 +ij +� +av +� +� +1/2 += +√ +˜z +m J. +(7) + +4 +The approximate location of the AT line for an m- +component infinite-range spin glass near the zero-field +transition temperature Tc is [5] +�hr +J +�2 += +4 +m(m + 2) +� +1 − T +Tc +�3 +. +(8) +The accuracy of this approximation for the SK model can +be judged from Fig. 1. +A one-dimensional chain with power law diluted in- +teractions for a particular value of σ is equivalent to a +short-range model [14] of effective dimension deff, where +deff = +2 +2σ − 1, +(9) +i.e., there is a one-to-one mapping between a long-range +diluted network with exponent σ and a short-range model +with space dimension deff, at least when 1/2 < σ < 2/3. +Thus when σ = 0.60, deff = 10. +For the interval +2/3 < σ < 1 other relations are required [15, 39]. For +example, for Ising spin glasses, it was suggested in Ref. +[39] that d = 4 corresponded to σ ≈ 0.790, while d = 3 +corresponded to σ ≈ 0.896. Unfortunately, the mapping +for the XY model has been less studied. +III. +CORRELATION LENGTHS AND +SUSCEPTIBILITIES +In this section we discuss the quantities which were +obtained from our Monte Carlo simulations and used to +extract a correlation length ξSG and the spin glass suscep- +tibility χSG. The simulations themselves are described in +detail in Appendix A. +The thermal average of a quantity is calculated using +multiple replicas in the following standard way: +⟨A⟩⟨B⟩⟨C⟩⟨D⟩ = ⟨A(1)B(2)C(3)D(4)⟩ +(10) +where (1),(2),(3), and (4) are four copies of the system at +the same temperature. The wave-vector-dependent spin +glass susceptibility is given by [5] +χSG(k) = 1 +N +� +i,j +1 +m +� +µ,ν +�� +χµν +ij +�2� +av eik(i−j), +(11) +where +χµν +ij = +� +Sµ +i Sν +j +� +− ⟨Sµ +i ⟩ +� +Sν +j +� +. +(12) +The spin glass correlation length is then determined from +ξSG = +1 +2 sin(kmin/2) +� +χSG(0) +χSG (kmin) − 1 +�1/(2σ−1) +(13) +where kmin = (2π/N). +IV. +FINITE-SIZE ANALYSES ASSUMING A +TRANSITION EXISTS +In this section we detail the method of finite-size anal- +ysis when a transition is assumed to exist. When study- +ing the AT line, which is a line of phase transitions in +the hr − T plane, it can be crossed on an infinite number +of trajectories. The most commonly used trajectory is +the one where hr is kept constant and the temperature +T is varied. In this work we also consider the trajectory +in which T is kept constant and hr is varied. We refer +to the zero-field transition temperature as Tc while we +denote a generic transition temperature on the AT line +by TAT(hr). Similarly we denote the field on the AT line +by hAT(T). +The spin glass susceptibility χSG ≡ χSG(0) of a finite +system of N spins has the finite size scaling form (near +the transition temperature TAT(hr)) [5]: +χSG +N 2−η = C +� +N 1/ν (T − TAT(hr)) +� +, +(2/3 ≤ σ < 1), +(14a) +χSG +N 1/3 = C +� +N 1/3 (T − TAT(hr)) +� +, +(1/2 < σ ≤ 2/3), +(14b) +where η is given by 2 − η = 2σ − 1. These forms are +examples of finite size scaling expressions which would +be expected to hold in the critical region when N → ∞, +(T − TAT(hr)) → 0, with (say) N 1/ν(T − TAT(hr)) finite. +The scaling function C will depend on the value of σ. +There are always finite size corrections to these forms. +For example, the corrections to Eq. (14b) will be of the +form +χSG +N 1/3 = C +� +N 1/3 (T − TAT(hr)) +� ++ N −ωG +� +N 1/3(T − TAT(hr)) +� +. +(15) +It has been suggested [17, 40] that the correction to scal- +ing exponent is given at least in the mean-field region +by +ω = 1/3 − (2σ − 1). +(16) +Curves of χSG/N 2−η +(χSG/N 1/3 +in the mean-field +regime) plotted for different system sizes should inter- +sect at the transition temperature TAT(hr). In reality, +finite-size corrections to Eq. (14) are always present and +cause the intersection point between the curves for size N +and 2N to depend on N. The intersection temperatures +vary as [40–43] +T ∗(N, 2N) = TAT(hr) + A +N λ , +(17) +where A is the amplitude of the leading correction, and +the exponent λ is +λ = 1/3 + ω, +(1/2 < σ ≤ 2/3), +(18a) +λ = 1/ν + ω, +(2/3 < σ < 1), +(18b) + +5 +where ω is the leading correction to the scaling exponent. +When σ = 0.6, ω = −2σ + 4/3, so λ = 5/3 − 2σ = 0.467 +[40]. In the regime when σ > 2/3 the values of both ν +and λ are not well-determined, so there we shall treat λ +as a fitting parameter. +The spin glass correlation length has a similar finite +size scaling form in the critical region +ξSG +N += X +� +N 1/ν (T − TAT(hr)) +� +, +(2/3 ≤ σ < 1), +(19a) +ξSG +N deff/6 = X +� +N 1/3 (T − TAT(hr)) +� +, +(1/2 < σ ≤ 2/3). +(19b) +ν, the correlation length critical exponent, has to be de- +termined numerically in the interval 2/3 < σ < 1. +We have also studied crossing the AT line at fixed T +and varying hr. Then Eq. (19) takes the form, +ξSG +N += X +� +N 1/ν (hr − hAT(T)) +� +, +(2/3 ≤ σ < 1), +(20a) +ξSG +N deff/6 = X +� +N 1/3 (hr − hAT(T)) +� +, +(1/2 < σ ≤ 2/3), +(20b) +where hAT(T) denotes the field at the AT line at tem- +perature T. Similarly, the spin glass susceptibility χSG +of the finite system near the AT transition line takes the +form +χSG +N 2−η = C +� +N 1/ν (hr − hAT(T)) +� +, +(2/3 ≤ σ < 1), +(21a) +χSG +N 1/3 = C +� +N 1/3 (hr − hAT(T)) +� +, +(1/2 < σ ≤ 2/3). +(21b) +In the thermodynamic limit, Eq. (19) is similar to +Eq. (20); the effect of finite size corrections to the two +can differ. For example, while the correction to scaling +exponent λ does not depend on the choice of the trajec- +tory, the magnitude of the scaling corrections can differ. +Thus in the intersection formulae when applied to fields +h∗(N, 2N) = hAT(T) + +˜A +N λ , +(22) +the coefficient ˜A will be different from A in Eq. (17). Cor- +rections to scaling of, say, Eq. (21a), are more generally +of the form +χSG +N 2−η = C +� +N 1/ν (hr − hAT(T)) +� ++ N −ωG +� +N 1/ν (hr − hAT(T)) +� +, +(23) +where ω is the correction to scaling exponent, and G is +another scaling function. This type of scaling form holds +in the limit where N 1/ν(hr −hAT(T)) is fixed as N → ∞, +which of course can only be realized approximately in +numerical studies. +A key feature of the finite size critical point scaling +analysis is that right on the AT line itself, that is when +hr = hAT(T), R = χSG/N 2−η (χSG/N 1/3 for σ ≤ 2/3) +should be finite as N → ∞. We find (see Sec. VI) that R +is at least not increasing with N, and perhaps finite (see +Fig. 37), for σ = 0.60 but for σ = 0.70, 0.75 and 0.85 it is +in fact increasing with N, at the crossing field h∗(N, 2N). +We deduce from this observation that at these values of +σ the crossings at h∗(N, 2N) are not associated with a +true critical point at all but are consequences of droplet +scaling. At a true critical point R would tend to a finite +constant as N increases, but we find it increases with N, +provided N > 1024 (or system sizes 2N > 2048) for the +case of σ = 0.75 (see Fig. 38). +In Sec. V we shall present our attempts at analysing +the data for σ = 0.6, σ = 0.75, σ = 0.85 at fixed values of +hr but varying T, and also at a fixed value of T and vary- +ing hr, on the assumption that there is an AT line and +using the finite-size scaling methods of this subsection. +We have also obtained data at fixed T and varying hr +for σ = 0.60, 0.70, 0.75, 0.85 and analysed them using +finite size generalizations of well-known droplet scaling +relations. In this case the droplet picture provides a sim- +ple set of formulae for analysing the data in the assumed +absence of an AT line. +V. +ANALYSES OF THE SIMULATION DATA +ASSUMING THERE IS AN AT LINE +We shall study the phase transitions at hr += 0, +and determine the zero-field transition temperature Tc +(= TAT(hr = 0)), and seek evidence of an AT transition +at non-zero hr using the standard critical point finite size +scaling method of determining the “crossings” or inter- +sections of the curves of, say, χSG/N z (with z = 1/3 +when σ ≤ 2/3, and with z = 2 − η = 2σ − 1 for σ ≥ 2/3) +at values of N and 2N as we reduce T through the AT +transition temperature at fixed hr, or the field hr at fixed +T in the vicinity of the AT field hAT(T) as outlined in +Sec. IV. There seems no reason to doubt the existence +of an AT line for any value of σ in the mean-field region +σ < 2/3, and our results are entirely consistent with the +existence of an AT transition at σ = 0.60. They serve +as a useful comparison for the studies in the non-mean +regime σ > 2/3, where the evidence will be found to favor +the droplet picture. We have studied N values 128, 256, +512, 1024, 2048, 4096, 8192 and 16384 for both σ = 0.60 +and σ = 0.75, but went up to N = 32768 for the case +of σ = 0.75 when the field hr was varied at fixed T. +When σ = 0.85 the largest N values used was 4096. In +this case the zero-field transition temperature Tc is quite +low and as a consequence all the investigations have to +be done also at low temperatures, where equilibration +times are long, preventing the study of larger systems. +We are mainly interested in the question as to whether + +6 +outside the mean-field region, that is for σ > 2/3, an AT +transition actually exists and whether (say) the depen- +dence of ξSG on the field hr can be understood as will +be suggested in Sec. VI on the droplet picture without +invoking an AT transition at all. If it can, this would +provide support to the argument that the droplet scaling +picture rather than replica symmetry breaking describes +spin glasses below 6 dimensions. We have analysed the +data for σ = 0.70, 0.75 and for σ = 0.85 using the usual +“crossing” method, (the finite size scaling approach out- +lined in Sec. IV), which indeed works well for σ = 0.6. +The evidence for the existence of an AT transition at +σ = 0.75 and 0.85 will be contrasted with the evidence +against an AT transition at these values of σ. +Our main focus was the case σ = 0.75. +We looked +briefly at the case σ = 0.70 to find whether or not it +might be practical to study whether σ = 2/3 is the value +of σ above which the AT line might disappear. We found +that it was similar to σ = 0.75, but that the corrections +to scaling were larger. This means that for a given level +of accuracy, larger N values are required. We studied +σ = 0.85 because it should behave similarly to physical +systems in three dimensions but we could not equilibrate +systems at the larger N values in this case because the +temperatures T of interest have to be less than Tc, which +is rather small. +A. +σ = 0.6 +We shall focus on σ = 0.60 in this subsection. It corre- +sponds according to Eq. (9) to an effective dimension of +10 dimensions, which is in the mean-field region; it lies +above the upper critical dimension of spin glasses, which +is 6 (or in the mean-field region σ < 2/3 in the one- +dimensional long-range model). It is natural to expect +that for this value of σ there will be an AT line and this +is amply confirmed by our simulations. For this value of +σ, simulations of the corresponding Ising model [12, 15] +and the Heisenberg model [17, 18] also found an AT line. +Our results for hr = 0 are given in Figs. 2 and 3. Ac- +cording to Eq. (14b), the data for χSG/N 1/3 when plot- +ted for different system sizes should intersect at the tran- +sition temperature Tc. Similarly, according to Eq. (19b), +the data of ξSG/N deff/6 with deff = 2/(2σ − 1) should in- +tersect at the same transition temperature. Fig. 2 shows +the data for different system sizes. We find the tempera- +ture T ∗(N, 2N) at which the curves corresponding to the +system sizes N and 2N intersect. We then fit this data +with Eq. (17) to find the transition temperature. The +exponent λ ≡ 5/3 − 2σ is known to equal 0.467 in this +case [17, 40]. The result is displayed in Fig. 3, where the +T ∗(N, 2N) data obtained from intersections of χSG are +fitted against N −λ with a straight line for the largest 6 +pairs of system sizes to give Tc = 0.8873 ± 0.0017. The +corresponding intersections of the ξSG data (omitting the +two smallest system sizes) gives Tc = 0.8893 ± 0.0046. +The values of Tc obtained from χSG data and ξSG data +0.75 +0.80 +0.85 +0.90 +0.95 +T +0.1 +1 +χSG/N 1/3 +σ = 0.6 +hr = 0 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +N = 16384 +0.7 +0.8 +0.9 +T +10−3 +10−2 +10−1 +100 +101 +ξSG/N deff /6 +Figure 2. +The main figure shows the plot of χSG/N 1/3 as +a function of the temperature T for different system sizes, +for σ = 0.6 with hr = 0. The inset figure shows the corre- +sponding data for ξSG/N deff/6, with deff = 2/(2σ − 1) in the +mean-field regime (Eq. (9)). +The exponents of N are cho- +sen according to Eqs. (14b) and (19b). Both the plots show +that the curves for different system sizes intersect. The data +for the intersection temperatures T ∗(N, 2N) between pairs of +adjacent system sizes are presented in Fig. 3. +are in agreement with each other. The mean-field pre- +diction of Eq. (7) is much higher, T MF +c += +√ +6/2 = 1.2247. +Fluctuation effects not present in the SK limit must be +responsible for this large difference. +For hr = 0.1, the data is as shown in Figs. 4 and 5. +When the T ∗(N, 2N) data obtained from χSG are fitted +against N −λ with a straight line for the largest 4 pairs +of system sizes we get TAT(hr = 0.1) = 0.6735 ± 0.0120. +The corresponding ξSG data (omitting the two smallest +system sizes) gives TAT(hr = 0.1) = 0.6745 ± 0.0148. +Thus we have found that the AT line passes through +the point (T, hr) = (0.674, 0.1). To compare that with +the predictions from the SK model, we use the zero-field +transition temperature Tc = 0.887 obtained above. Then +for hr = 0.1, the predicted value of the AT transition +temperature ratio of the SK model would be TAT(hr = +0.1)/Tc = 0.74, while the Monte Carlo determined value +at σ = 0.6 is 0.7590 ± 0.0113. (For the SK model, the +Monte Carlo value of the ratio is 0.7641 ± 0.0341). Thus +while the zero-field transition temperature at σ = 0.6 is +not close to the mean-field value of Eq. (7), the SK form +of the AT line is a good approximation provided it is ex- +pressed in terms of the renormalized zero-field transition +temperature Tc (see also Fig. 1). +The AT line can be approached not only by reducing +the temperature T but also by reducing the field at fixed +T. In Figs. 6 and 7 we have constructed the crossing plots + +7 +0.000 +0.025 +0.050 +0.075 +0.100 +N −λ +0.70 +0.75 +0.80 +0.85 +T ∗(N, 2N) +σ = 0.6 +hr = 0 +χSG +−0.72N −λ + 0.89 +ξSG +−1.56N −λ + 0.89 +Figure 3. A plot of the intersection temperatures T ∗(N, 2N) +for χSG/N 1/3 and ξSG/N deff/6 obtained from the data in +Fig. 2, as a function of N −λ, for σ = 0.6 with hr = 0. The +value of the exponent λ is fixed to be 0.467 which is known +exactly [17, 40]. The fits give Tc = 0.8873 ± 0.0017 from χSG +and Tc = 0.8893 ± 0.0046 from ξSG. +0.60 +0.65 +0.70 +0.75 +T +0.2 +0.3 +0.4 +0.5 +0.6 +χSG/N 1/3 +σ = 0.6 +hr = 0.1 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +N = 16384 +0.55 +0.60 +0.65 +0.70 +0.75 +T +0.1 +1 +10 +ξSG/N deff /6 +Figure 4. A finite size scaling plot of χSG (main figure) and +ξSG (inset figure), for σ = 0.6 in a magnetic field of hr = +0.1. Both the datasets clearly indicate that a phase transition +occurs. +The transition temperature in the thermodynamic +limit is estimated in Fig. 5. +as a function of hr for ξSG and χSG respectively. Analysis +of the crossing plots of h∗(N, 2N) in Fig. 8 shows that +the behavior is again consistent with the existence of an +AT line at least at σ = 0.60. The same value of λ was +used as when plotting T ∗(N, 2N). The h∗(N, 2N) data +for all the pairs of system sizes are fitted against N −λ +to give hAT(T = 0.6) = 0.1569 ± 0.0061 from χSG and +0.000 +0.025 +0.050 +0.075 +0.100 +N −λ +0.58 +0.60 +0.62 +0.64 +0.66 +0.68 +T ∗(N, 2N) +σ = 0.6 +hr = 0.1 +χSG +−0.23N −λ + 0.67 +ξSG +−0.58N −λ + 0.67 +Figure 5. +The intersection temperatures T ∗(N, 2N), for +σ = 0.6 with hr = 0.1 (look at Fig. 4). Both the datasets +are consistent with a spin glass transition temperature of +TAT(hr = 0.1) = 0.67. +10−2 +10−1 +100 +hr +10−7 +10−5 +10−3 +10−1 +101 +103 +ξSG/N deff/6 +σ = 0.6 +T = 0.6 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +N = 16384 +Figure 6. A finite size scaling plot of ξSG as a function of +magnetic field hr, for σ = 0.6 at a temperature of T = 0.6. +hAT(T = 0.6) = 0.1571 ± 0.0067 from ξSG. We found +two points on the AT line, (T, hr) = (0.674, 0.1) from +T ∗(N, 2N), and (T, hr) = (0.6, 0.157) from h∗(N, 2N) +data. These points are plotted in Fig. 1 for comparison +with the exact AT line for the SK model. +B. +σ = 0.75 +The case σ = 0.75 corresponds to the non-mean-field +regime: the long-range diluted model for this value of σ +is equivalent to a short-range model with d ≈ 4 dimen- + +8 +10−2 +10−1 +100 +hr +10−2 +10−1 +100 +χSG/N 1/3 +σ = 0.6 +T = 0.6 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +N = 16384 +Figure 7. A finite size scaling plot of χSG as a function of +magnetic field hr, for σ = 0.6 at a temperature of T = 0.6. +0.000 +0.025 +0.050 +0.075 +0.100 +N −λ +0.08 +0.10 +0.12 +0.14 +0.16 +h∗(N, 2N) +σ = 0.6 +T = 0.6 +λ = 5 +3 − 2σ = 0.467 +χSG +−0.36N −λ + 0.16 +ξSG +−0.61N −λ + 0.16 +Figure 8. The intersection fields h∗(N, 2N), for σ = 0.6 with +T = 0.6. Both the datasets are consistent with a spin glass +transition at hAT(T = 0.6) ≈ 0.16. +sions. In this regime, simulations of the corresponding +Heisenberg model [17, 18] were thought consistent with +an AT transition. +According to Eq. (14a), the data for χSG/N 2−η,where +2 − η = 2σ − 1, plotted for different system sizes should +intersect at the transition temperature Tc. Similarly, ac- +cording to Eq. (19a), the curves of ξSG/N should also +intersect at the transition temperature. The main plots +of Figs. 9 and 11 show the finite-size-scaled data of χSG, +and the corresponding inset plots show the finite-size- +scaled data of ξSG. The curves for different system sizes +show a clear tendency to intersect close to the same tem- +perature. The data for T ∗(N, 2N) are then fitted with +0.50 +0.55 +0.60 +0.65 +0.70 +T +0.1 +χSG/N 2−η +σ = 0.75 +hr = 0 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +0.5 +0.6 +0.7 +T +0.1 +1 +ξSG/N +Figure 9. The main figure shows data for χSG/N 2−η (with +2 − η = 2σ − 1) for different system sizes, for σ = 0.75 with +hr = 0. The inset shows the data for ξSG/N. According to +Eqs. (14a) and (19a) the data should intersect at Tc, which is +shown in Fig. 10. +0.0 +0.1 +0.2 +0.3 +0.4 +N −λ +0.54 +0.56 +0.58 +0.60 +0.62 +0.64 +T ∗(N, 2N) +σ = 0.75 +hr = 0 +λ = 0.1583 +χSG +−0.15N −λ + 0.64 +ξSG +−0.15N −λ + 0.64 +Figure 10. A plot of the intersection temperatures T ∗(N, 2N) +obtained from the data in Fig. 9, for σ = 0.75 with hr = 0. +Using the value of the scaling exponent λ = 0.1583 (ob- +tained from the h∗(N, 2N) data), the T ∗(N, 2N) data are +fitted against N −λ using a straight line. The resulting values +for the transition temperature are Tc = 0.6397 ± 0.0051 from +χSG and Tc = 0.6440 ± 0.0154 from ξSG. +Eq. (17) where the value of the exponent λ is not known +in the non-mean-field regime and hence should be con- +sidered as a fitting parameter. +If there were an AT transition there would be a unique +value of λ, the same for both the ξSG and χSG intersec- +tions corresponding to both h∗(N, 2N) and T ∗(N, 2N). +The h∗(N, 2N) data obtained from χSG intersections in + +9 +0.30 +0.35 +0.40 +0.45 +0.50 +T +0.3 +0.4 +χSG/N 2−η +σ = 0.75 +hr = 0.05 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +0.35 +0.40 +0.45 +0.50 +0.55 +T +1 +ξSG/N +Figure 11. +A finite size scaling plot of χSG (main figure) +and ξSG (inset figure) for σ = 0.75. +The magnetic field is +hr = 0.05. +0.0 +0.1 +0.2 +0.3 +0.4 +N −λ +0.2 +0.3 +0.4 +0.5 +T ∗(N, 2N) +σ = 0.75 +hr = 0.05 +λ = 0.1583 +χSG +−0.35N −λ + 0.48 +ξSG +0.73N −λ + 0.19 +Figure 12. The intersection temperatures T ∗(N, 2N) for σ = +0.75 with hr = 0.05. The data from Fig. 11 did not fit well +with Eq. (17). So we used the value of λ obtained from the +h∗(N, 2N) data and did a linear fitting which gives TAT(hr = +0.05) = 0.4832 ± 0.0394 from χSG and TAT(hr = 0.05) = +0.1894 ± 0.0383 from ξSG, and the values do not agree with +each other. +Fig. 14 (which is described later) are fitted with Eq. (22) +by considering λ, Tc and ˜A as fitting parameters. This +is a non-linear fitting procedure for which we use effi- +cient methods like the Trusted Region Reflective (TRF) +algorithm and the Levenberg-Marquardt(LM) algorithm +(for which packages are available in python) to determine +the fitting parameters, and we obtain λ = 0.1583. Since +the exponent giving the leading correction to scaling λ is +universal, we use the same value of λ with both intersec- +10−3 +10−2 +10−1 +100 +hr +10−4 +10−3 +10−2 +10−1 +100 +ξSG/N +σ = 0.75 +T = 0.55 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +N = 16384 +N = 32768 +10−3 +10−2 +10−1 +100 +hr +10−3 +10−1 +ξSG/N +Figure 13. A finite size scaling plot of ξSG as a function of +magnetic field hr, for σ = 0.75 at a temperature of T = 0.55. +The inset shows our two largest system sizes. +10−3 +10−2 +10−1 +100 +hr +10−3 +10−2 +10−1 +χSG/N 2−η +σ = 0.75 +T = 0.55 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +N = 16384 +N = 32768 +Figure 14. A finite size scaling plot of χSG as a function of +magnetic field hr, for σ = 0.75 at a temperature of T = 0.55. +tions h∗(N, 2N) and T ∗(N, 2N) obtained from χSG and +ξSG data. We substitute the value of λ obtained above +in Eq. (17) and fit the T ∗(N, 2N) data against N −λ with +a straight line. +As shown in Fig. 10, for hr = 0, the +χSG fit (considering all the pairs of system sizes) gives +Tc = 0.6397 ± 0.0051. The corresponding ξSG fit (omit- +ting the smallest system size) gives Tc = 0.6440±0.0154. +For hr = 0.05, the intersection temperatures data +are shown in Fig. 12. +Omitting the smallest system + +10 +0.0 +0.1 +0.2 +0.3 +0.4 +N −λ +0.005 +0.010 +0.015 +0.020 +h∗(N, 2N) +σ = 0.75 +T = 0.55 +λ = 0.1583 +χSG +0.03N −λ + 0.00 +ξSG +−0.01N −λ + 0.02 +Figure 15. The intersection fields h∗(N, 2N) for σ = 0.75 with +T = 0.55. +size, the T ∗(N, 2N) data are fitted with Eq. +(17) to +give TAT(hr = 0.05) = 0.4832 ± 0.0394 from χSG and +TAT(hr = 0.05) = 0.1894 ± 0.0383 from ξSG. Compared +to Fig. +5 which gives the equivalent plot for the case +with σ = 0.60, the data in Fig. 12 does not look like +data which is converging to the same asymptotic limit +when N is large. If the crossings were actually due to a +genuine AT transition, then the asymptotic limit should +be the same for both. +We have also studied ξSG and χSG at fixed T, but vary- +ing hr and the finite size scaling plots for these are given +in Figs. +13 and 14. +There appears to be good inter- +sections in the curves, supporting therefore the possible +existence of an AT transition at the temperature studied +T = 0.55. A plot of h∗(N, 2N) versus 1/N λ is in Fig. 15, +using the same value of λ = 0.1583. In the intersections +of ξSG there is a clear rising trend of h∗(N, 2N) with in- +creasing N until N = 1024, followed by decreasing values +of h∗(N, 2N) for N > 2048. For the case of σ = 0.60, +where there is almost certainly a genuine AT transition, +(Fig. 8) only the rising trend is seen. It is as if for the +smaller systems N < 2048 the system at σ = 0.75 is +behaving similarly to its mean-field cousin at σ = 0.60. +Note that this change of trend cannot be attributed to the +correction to scaling terms of Eq. (23). These only apply +in the limit N → ∞ with N 1/ν(hr −hAT(T)) fixed. For a +genuine AT transition the intersections h∗(N, 2N) from +both ξSG and χSG should both extrapolate as N → ∞ to +the same field hAT(T). It is hard to argue that Fig. 15 +provides good evidence for this. On the other hand, on +the droplet picture, it would be expected that h∗(N, 2N) +should extrapolate to zero. The evidence that is happen- +ing is also weak. +0.25 +0.30 +0.35 +0.40 +T +0.1 +0.2 +0.3 +χSG/N 2−η +σ = 0.85 +hr = 0 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +0.3 +0.325 +0.35 +0.375 +0.4 +T +1 +ξSG/N +Figure 16. A finite size scaling plot of χSG (main figure) and +ξSG (inset figure), for σ = 0.85 in a magnetic field of hr = 0 +(with 2 − η = 2σ − 1). Both the datasets clearly indicate that +a phase transition occurs. The transition temperature in the +thermodynamic limit is estimated in Fig. 17. +0.000 +0.002 +0.004 +0.006 +N −λ +0.32 +0.34 +0.36 +0.38 +T ∗(N, 2N) +σ = 0.85 +hr = 0 +λ = 1.0315 +χSG +−2.93N −λ + 0.33 +ξSG +9.78N −λ + 0.33 +Figure 17. The intersection temperatures T ∗(N, 2N), for σ = +0.85 with hr = 0. A non-linear fit of the χSG data from Fig. 16 +with the Eq. (17) using the Levenberg-Marquadt algorithm +gives λ = 1.0315. A linear fit of the data using this value of λ +gives Tc = 0.3336±0.0013 from χSG and Tc = 0.3297±0.0036 +from ξSG. +C. +σ = 0.85 +For σ = 0.85 we are further into the non-mean-field +region. +According to Eq. (9), σ = 0.85 corresponds +to a short-range model close to three dimensions. +In +this regime, simulations of the corresponding Heisenberg +model [17, 18] did not find an AT line. + +11 +0.1 +0.2 +0.3 +0.4 +T +0.1 +χSG/N 2−η +σ = 0.85 +hr = 0.05 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +Figure 18. A finite size scaling plot of χSG for σ = 0.85. The +magnetic field is hr = 0.05. +0.1 +0.2 +0.3 +0.4 +T +0.1 +0.2 +χSG/N 2−η +σ = 0.85 +hr = 0.02 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +Figure 19. +A finite size scaling plot of χSG, for σ = 0.85 +with hr = 0.02. +The data do not intersect even at very +low temperatures (much lower than the mean field value of +TAT(hr = 0.02) = 0.2997 obtained using Eq. (8)) indicating +that there is no phase transition in this regime. +For hr = 0, Fig. 16 clearly shows that the curves for +different system sizes are intersecting. The data for in- +tersection temperatures are shown in Fig. 17. Similar to +the case of σ = 0.75, the T ∗(N, 2N) data obtained from +χSG are fitted with Eq. (17) by considering λ, Tc, and A +as fitting parameters. We obtain λ = 1.0315 from both +TRF and LM methods. The fit using the χSG data for +0.1 +0.2 +0.3 +0.4 +T +1 +ξSG/N +σ = 0.85 +hr = 0.05 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +Figure 20. +A finite size scaling plot of ξSG, for σ = 0.85 +with hr = 0.05. +The data show merging behavior at low +temperatures. +0.1 +0.2 +0.3 +0.4 +T +100 +101 +ξSG/N +σ = 0.85 +hr = 0.02 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +Figure 21. +A finite size scaling plot of ξSG, for σ = 0.85 +with hr = 0.02. +The data show merging behavior at low +temperatures. +all the pairs of system sizes gives Tc = 0.3336 ± 0.0013. +The corresponding ξSG fit (omitting the smallest system +size) gives Tc = 0.3297 ± 0.0036. The two values of Tc +are quite close. +For hr = 0.05 the χSG/N 2−η data do not intersect as +shown in Fig. 18. Such a field could conceivably be above +the largest AT field even at T = 0 so we also studied a +smaller field: hr = 0.02 shown in Fig. 19. There is no + +12 +10−3 +10−2 +10−1 +100 +hr +10−2 +10−1 +100 +ξSG/N +σ = 0.85 +T = 0.3 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +10−3 +10−2 +10−1 +100 +hr +10−2 +10−1 +100 +ξSG/N +Figure 22. A finite size scaling plot of ξSG as a function of +magnetic field hr, for σ = 0.85 at a temperature of T = 0.3. +The inset shows our two largest system sizes. +10−3 +10−2 +10−1 +100 +hr +10−3 +10−2 +10−1 +χSG/N 2−η +σ = 0.85 +T = 0.3 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +Figure 23. A finite size scaling plot of χSG as a function of +magnetic field hr, for σ = 0.85 at a temperature of T = 0.3. +sign of any crossing at this field either!. The ξSG data is +less clearcut. Fig. 20 shows there are no intersections at +a field of hr = 0.05 while a merging behavior is seen for +the larger systems at hr = 0.02, as shown in Fig. 21. In +our simulations we went to very low temperatures such as +T = 0.1, which is small in comparison with the mean-field +values of TAT for hr = 0.02 and hr = 0.05 using Eq. (8), +but we still could not find any clear intersections in the +χSG or ξSG data. This suggests that there is no phase +0.000 +0.002 +0.004 +0.006 +N −λ +0.01 +0.02 +0.03 +h∗(N, 2N) +σ = 0.85 +T = 0.3 +λ = 1.0315 +χSG +0.36N −λ + 0.00 +ξSG +3.96N −λ + 0.00 +Figure 24. The intersection fields h∗(N, 2N) for σ = 0.85 with +T = 0.3. We substitue the value of the exponent λ = 1.0315 +obtained from the T ∗(N, 2N) data at hr = 0 in Eq. (22) +and fit h∗(N, 2N) data agianst N −λ to get hAT(T = 0.3) = +0.0046±0.0006 from χSG and hAT(T = 0.3) = 0.0047±0.0016 +from ξSG. +transition in this regime in the presence of a magnetic +field. Our data are consistent with the scenario where +the external magnetic field destroys the phase transition, +just as happens for a ferromagnet when a uniform field +is turned on. +Very similar features were seen for the +Heisenberg version of this model [17, 18] and in the three +dimensional Ising model [34]. +Confusingly, intersections are seen at fixed T = 0.3 +as hr is varied in the plots of ξSG/N in Fig. 22 and of +χSG/N 2−η in Fig. 23. The usual analysis of h∗(N, 2N) +is given in Fig. 24. Thus in crossing the AT line along a +trajectory of fixed T we have seen intersections, suggest- +ing there might be an AT transition. However, the large +N limit of h∗(N, 2N) in Fig. 24 in the case of σ = 0.85, +suggests that hAT(T) might actually be zero, consistent +with the droplet scaling picture. In the next section the +dependence of ξSG and χSG on hr will be explained using +the droplet scaling approach. +VI. +DATA ANALYSES ON THE DROPLET +PICTURE +In this section we give the field dependence of ξSG and +χSG according to the droplet picture [44–46], including +also their finite size modifications, and compare these +with our simulation data. +In the droplet picture one uses an Imry-Ma argument +[47] for the correlation length ξ and identifies it with +the size of the region or domain within which the spins +become re-oriented in the presence of the random field. +The free energy gained from such a reorientation by the +the random field is of order +� +q(T)hrξd/2. The size of +such domains ξ is determined by equating this free energy + +13 +10−3 +10−2 +10−1 +100 +101 +hr +10−1 +101 +103 +105 +107 +ξSG +σ = 0.75 +T = 0.55 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +N = 16384 +N = 32768 +fit for N = 32768 +(∼ h−2.71 +r +) +Figure 25. A plot of ξSG as a function of magnetic field hr, +for σ = 0.75 at a temperature of T = 0.55. +10−3 +10−2 +10−1 +100 +101 +hr +100 +102 +104 +106 +ξSG +σ = 0.85 +T = 0.3 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +fit for N = 4096 +(∼ h−2.20 +r +) +Figure 26. A plot of ξSG as a function of magnetic field hr, +for σ = 0.85 at a temperature of T = 0.3. +to the free energy cost of the interface of this domain of +re-ordered spins with the rest of the system, which is of +the form Υ(T)ξθ [48]. Equating these two free energies +gives +ξ ∼ +� +Υ(T) +� +q(T)hr +�1/(d/2−θ) +. +(24) +While there is a considerable literature on the depen- +dence of the interface exponent θ on σ for the case of +10−2 +10−1 +100 +101 +hr +10−6 +10−3 +100 +103 +106 +109 +1012 +ξSG +σ = 0.6 +T = 0.6 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +N = 16384 +fit for N = 16384 +(∼ h−6.88 +r +) +Figure 27. A plot of ξSG as a function of magnetic field hr, +for σ = 0.6 at a temperature of T = 0.6. +10−1 +101 +N 1/xhr +10−6 +10−4 +10−2 +100 +ξSG/N +σ = 0.75 +T = 0.55 +x = 2.7077 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +N = 16384 +N = 32768 +Figure 28. +A complete finite size scaling plot of ξSG as a +function of magnetic field hr, plotted on a log-log scale, for +σ = 0.75 at a temperature of T = 0.55. +Ising spin glasses [49], we know of no equivalent studies +for the case of the XY spin glass. (Our data suggests +that its θ might be close to that of the Ising spin glass). +Eq. (24) shows that as hr → 0, the length scale be- +comes infinite; ξ diverges as ξ ∼ 1/hx +r, where +x = +1 +d/2 − θ. +(25) +The exponent x is the analogue of ν at the AT transition; + +14 +10−1 +101 +N 1/xhr +10−5 +10−4 +10−3 +10−2 +10−1 +100 +ξSG/N +σ = 0.85 +T = 0.3 +x = 2.2019 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +Figure 29. +A complete finite size scaling plot of ξSG as a +function of magnetic field hr, plotted on a log-log scale, for +σ = 0.85 at a temperature of T = 0.3. +10−1 +100 +101 +102 +N 1/xhr +10−6 +10−4 +10−2 +100 +ξSG/N +σ = 0.75 +T = 0.55 +x = 2.7077 +N = 16384 +N = 32768 +Figure 30. +A complete finite size scaling plot of ξSG as a +function of magnetic field hr, for σ = 0.75 at a temperature +of T = 0.55, showing our two largest system sizes. +it is as if the AT transition hAT(T) = 0. +We would +expect this formula to apply until finite size effects limit +its growth, which will occur when ξ is of O(L) (or O(N) +in our one-dimensional system). Identifying ξSG with ξ, +Figs. 25 and 26 show that the Imry-Ma fit indeed works +well at the larger fields; the data for the larger hr collapse +nicely onto a power law form as predicted by Eq. (24) +for all sizes N. It only departs from this formula when +10−1 +100 +101 +102 +N 1/xhr +10−5 +10−4 +10−3 +10−2 +10−1 +100 +ξSG/N +σ = 0.85 +T = 0.3 +x = 2.2019 +N = 2048 +N = 4096 +Figure 31. +A complete finite size scaling plot of ξSG as a +function of magnetic field hr, for σ = 0.85 at a temperature +of T = 0.3, showing our two largest system sizes. +10−1 +101 +N 1/xhr +10−5 +10−4 +10−3 +10−2 +10−1 +χSG/N z +σ = 0.75 +T = 0.55 +x = 2.7077 +xχ = 1.6114 +z = 0.5951 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +N = 16384 +N = 32768 +Figure 32. +A complete finite size scaling plot of χSG as a +function of magnetic field hr, for σ = 0.75 at a temperature +of T = 0.55. +ξSG becomes of order N, when finite size corrections to +the Imry-Ma formula are needed. Also TNT effects (see +Sec. VII) produce corrections to the Imry-Ma formula +when ξSG is of O(N) unless N = L > L∗. The crossover +scale L∗ is thought to be large, especially as σ approaches +2/3 (or d → 6) [3]. +To allow for finite size effects on the Imry-Ma formula +we use the analogue of Eq. (21a) with hAT = 0 and ν = x + +15 +10−1 +100 +101 +102 +N 1/xhr +10−5 +10−4 +10−3 +10−2 +10−1 +χSG/N z +σ = 0.75 +T = 0.55 +x = 2.7077 +xχ = 1.6114 +z = 0.5951 +N = 16384 +N = 32768 +Figure 33. +A complete finite size scaling plot of χSG as a +function of magnetic field hr, for σ = 0.75 at a temperature +of T = 0.55, for our two largest system sizes. +10−1 +101 +N 1/xhr +10−6 +10−5 +10−4 +10−3 +10−2 +10−1 +χSG/N z +σ = 0.85 +T = 0.3 +x = 2.2019 +xχ = 1.8919 +z = 0.8592 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +Figure 34. +A complete finite size scaling plot of χSG as a +function of magnetic field hr, for σ = 0.85 at a temperature +of T = 0.3. +to write: +ξSG/N = X(N 1/xhr). +(26) +Our results for σ = 0.75 are shown in Fig. 28 and for σ = +0.85 are shown in Fig. +29. There are clearly finite size +corrections to this formula. It is a formula which formally +would be expected to hold in the scaling limit of N → ∞ +with N 1/xhr fixed. The crossover function X(y) ∼ 1/yx +10−1 +100 +101 +102 +N 1/xhr +10−6 +10−5 +10−4 +10−3 +10−2 +10−1 +χSG/N z +σ = 0.85 +T = 0.3 +x = 2.2019 +xχ = 1.8919 +z = 0.8592 +N = 2048 +N = 4096 +Figure 35. +A complete finite size scaling plot of χSG as a +function of magnetic field hr, for σ = 0.85 at a temperature +of T = 0.3, for our two largest system sizes. +0.3 +0.4 +0.5 +0.6 +0.7 +1/N z−(2σ−1) +0.2 +0.4 +0.6 +0.8 +h∗(N, 2N) N 1/x +χSG(σ = 0.7) +ξSG(σ = 0.7) +χSG(σ = 0.75) +ξSG(σ = 0.75) +χSG(σ = 0.85) +ξSG(σ = 0.85) +Figure 36. A plot of h∗(N, 2N)N 1/x versus 1/N ω, for σ = +0.70, 0.75 and 0.85. The values of x and ω, which is obtained +from Eq. (36) were taken from Table I. +when y is large, in order to recover Eq. (24). It goes to +a constant when y → 0. However, a closer look at our +two largest system sizes N = 16384 and N = 32768 at +σ = 0.75 (Fig. 30) and our two largest system sizes at +σ = 0.85, N = 2048 and N = 4096 (Fig. 31) shows that +the finite size corrections are becoming small, and are +smaller the further the system is away from the mean- +field region. If one moves in the other direction, towards + +16 +103 +104 +N +0.24 +0.26 +0.28 +0.30 +0.32 +0.34 +0.36 +R +σ = 0.6, T = 0.6 +Figure 37. A plot of R = χSG(h∗(N, 2N), N)/N 1/3 versus N, +for σ = 0.6 at a temperature of T = 0.6. +103 +104 +N +0.28 +0.29 +0.30 +0.31 +0.32 +0.33 +0.34 +R +σ = 0.75, T = 0.55 +Figure 38. A plot of R = χSG(h∗(N, 2N), N)/N 2σ−1 versus +N, for σ = 0.75 at a temperature of T = 0.55. +the start of the mean-field region σ = 2/3, the finite size +corrections are larger, as seen in Fig. 40 for σ = 0.70. +The finite size scaling form for these corrections to the +scaling of Eq. (26) will be of the form +ξSG +N += X(N 1/xhr) + N −ωH(N 1/xhr), +(27) +where ω is the correction to scaling exponent. However, +TNT effects (see Sec. VII) produce large further correc- +tions to these asymptotic forms when L < L∗. Since in +103 +N +0.2200 +0.2225 +0.2250 +0.2275 +0.2300 +0.2325 +0.2350 +0.2375 +R +σ = 0.85, T = 0.3 +Figure 39. A plot of R = χSG(h∗(N, 2N), N)/N 2σ−1 versus +N, for σ = 0.85 at a temperature of T = 0.3. +our studies L∗ is probably larger than the length N of +our system, at least for σ = 0.75, the scaling form of +Eq. (27) does not work in the region where ξSG is of or- +der N (see Fig. 42). For σ = 0.85 where L∗ is expected +to be smaller, Fig. 43 hints that Eq. (27) might apply as +the plots at adjacent sizes for the larger N values seem +to be getting closer together as N is increased, which is +a feature predicted by Eq. (27). +In Fig. +27 we show a similar plot to those in Figs. +25 and 26 but for the case of σ = 0.60. Notice however +that because of the AT transition at this value of σ, at +which ξSG would diverge to infinity as N → ∞ at some +finite field hr = hAT(T), a shoulder above the dashed +line has started to appear which is the beginning of this +divergence. Such a feature is absent in the figures for +both σ = 0.75 and at σ = 0.85. +The spin glass susceptibility according to the droplet +picture is a similar generalization of the finite size scaling +form of Eq. (14a): +χSG +N z = C(N 1/xhr), +(2/3 ≤ σ < 1). +(28) +The crossover function C(y) ∼ 1/yxχ when y is large, so +that then χSG ∼ ξz and becomes independent of N. Its +form is then +χSG ∼ 1/hxχ +r , +(29) +which implies that xχ = xz. +In the opposite limit as +y → 0, C(y) goes to a finite constant. The exponent z +depends upon whether we are dealing with short-range +interactions, (such as nearest-neighbor interactions) or +with the long-range interactions employed in this paper. +For short-range interactions, the average value of χ2 +ij falls + +17 +Table I. Values of the exponents xχ, x, and z obtained from +our simulations for different values of σ and T. +σ +T +xχ +x +z +0.7 +0.6 +1.5747 ± 0.0009 +3.3220 ± 0.0413 +0.4740 ± 0.0062 +0.75 +0.55 +1.6114 ± 0.0005 +2.7077 ± 0.0531 +0.5951 ± 0.0119 +0.85 +0.3 +1.8919 ± 0.0014 +2.2019 ± 0.0152 +0.8592 ± 0.0066 +off with spin separation rij as +χ2 +ij ∼ q(T)2T +Υ(T)rθ +ij +, +(30) +[45, 46]. This result applies in the zero-field spin glass +state. Then as, +χSG = 1 +N +N +� +i,j=1 +χ2 +ij, +(31) +so in d dimensions for the zero-field spin glass χSG ∼ +Ld−θ. Hence +z = d − θ +d +, +(32) +in order to recover the result χSG → N z as N 1/xhr goes +to zero. We caution that this formula for z will only hold +for short-range interactions. +With long-range interactions a “droplet” is not a single +connected region but a set of isolated islands of flipped +spins [49] and this will make the decay of χ2 +ij with rij +faster than in Eq. (30). This is an effect which has not +been studied before, and so in our problem the exponent +z has to be determined by fitting the data. The results +of our determinations of the droplet exponents x, xχ and +z for the different values of σ which we have studied are +summarised in Table 1. +The resulting excellent data collapse (at least when +ξSG < N), is shown in Figs. 32, 33, +34, and +35. The +value of z was determined from the observation that when +N 1/xhr is large, χSG should be independent of N. It is +remarkable that z determined at large values of N 1/xhr +results in a decent collapse of the data in the opposite +limit where N 1/xhr → 0. Nevertheless corrections to the +Imry-Ma scaling form are visible in the figures (and are +sizeable in the region where N 1/xhr is small when viewed +in a linear plot rather than a log scale plot, (just as in the +ξSG plots Figs. 28 and 29). In the limit when N 1/xhr is +held fixed with N → ∞ the leading correction to scaling +will be +χSG +N z = C(N 1/xhr) + N −ωG(N 1/xhr), +(33) +where G(y) is an unknown scaling function and the cor- +rection to scaling exponent ω is not known with any cer- +tainty (but see Eq. (36)). +Let us suppose that the droplet picture is correct and +that (say) the spin glass susceptibility χSG is described +by Eq. (28). This equation predicts that there will be a +crossing in the plots of χSG/N 2σ−1 used in AT line criti- +cal scaling studies. (Note we are setting 2 − η = 2σ − 1). +The correction to scaling term of Eq. (33) is not needed +for this, but this correction does strongly influence where +the crossings take place for the N values which are +reached in our simulations. The crossing arises as fol- +lows. At small values of hrN 1/x, the function C goes to a +constant. It turns out that z > (2σ − 1), so χSG/N 2σ−1 +diverges as N is increased as N z−(2σ−1) as hr → 0. On +the other hand when hrN 1/x is large, χSG → 1/hz +r, so +χSG/N 2σ−1 → 1/N 2σ−1hz +r → 0 as N goes to infinity. +Because at small fields, χSG/N 2σ−1 is larger for large N, +but at bigger hr fields it is smaller at the larger N val- +ues, so there must be a crossing point. We shall denote +the crossing value between the lines at N and 2N by +h∗(N, 2N) = H. Then H is determined by the solution +of the following +χSG(H, N) +N 2σ−1 += C(N 1/xH)N z−(2σ−1) = +χSG(H, 2N) +(2N)2σ−1 += C((2N)1/xH)(2N)z−(2σ−1). +(34) +Assuming C(y) → a − by, when y → 0, it is easy to +show then that the N dependence of h∗(N, 2N) at very +large N will be as 1/N 1/x. In reality we have no data +in this region of very large N where the corrections to +scaling term in Eq. (33) can be ignored. The corrections +to scaling are numerically small but are very important +in determining the values of h∗(N, 2N). +There is a similar crossing predicted in the plots of +ξSG/N as a function of hr when Eq. (27) holds, using the +analogue of Eq. (34). In this case it is the scaling cor- +rection which causes the curves to cross, (which requires +H(0) to be negative), and for these curves the crossings +h∗(N, 2N) at very large N will decrease as 1/N 1/x+ω, +(compare with Eq. (18b)) on taking χ(y) = c − dy and +H(y) → constant as y → 0. Once again we have no data +in this very large N regime. In Fig. 36 we have plotted +h∗(N, 2N)N 1/x versus 1/N ω, assuming that ω is given by +Eq. (36). Note that the size of the corrections to scaling +∼ 1/N ω is simply not small for the values of N which we +can study, contrary to what was assumed in the above. +h∗(N, 2N)N 1/x should go to a constant as N goes to in- +finity and it is only for the case of σ = 0.85, where the +corrections to scaling are the smallest of the three cases +studied, does that look remotely possible. For the case +of σ = 0.70 the corrections look to be very large. We +conclude that for the values of σ = 0.70 and σ = 0.75, +the crossing data on h∗(N, 2N) is not close to the large +N asymptotic form predicted by the droplet picture. But +the droplet picture does predict that the existence of such +intersections. +If we only had information on the values of the cross- +ing fields h∗(N, 2N) it would be difficult to really be sure +whether the droplet picture or the RSB picture best de- +scribed the data. The results on h∗(N, 2N) alone are in- +conclusive as regards both the AT transition line picture + +18 +100 +101 +102 +N 1/xhr +10−7 +10−5 +10−3 +10−1 +101 +ξSG/N +σ = 0.7 +T = 0.6 +x = 3.3220 +N = 8192 +N = 16384 +Figure 40. +A complete finite size scaling plot of ξSG as a +function of magnetic field hr, for σ = 0.70 at a temperature +of T = 0.6 for our two largest system sizes. +and the droplet picture. While on the droplet picture +h∗(N, 2N) are predicted to go to zero as N → ∞, the +values of h∗(N, 2N) are not convincingly going to zero as +N is increased (see Fig. 36). Fortunately, there is another +way of distinguishing the two approaches, which does +not require us to reach the N values at which h∗(N, 2N) +starts to approach zero. We define +R ≡ χSG(h∗(N, 2N), N)/N 2σ−1. +(35) +(Because we only determine χSG(hr, N) at a finite num- +ber of values of hr, we use linear interpolation to calcu- +late χSG(h∗(N, 2N), N) using the χSG(hr, N) values at +the two determined values of hr which lie on either side +of h∗(N, 2N)). On the phase transition picture, R should +approach a finite constant as N → ∞. On the droplet +picture R should increase as N z−(2σ−1) as N → ∞. For +σ = 0.60 where an AT line is expected R should go to +a constant but at the N values studied it actually still +appears to be decreasing (see Fig. 37) and has yet to be- +come constant, presumably due to finite size effects. This +indicates that trying to determine whether σ = 2/3 is the +exact value at which the crossover to droplet scaling be- +havior will also be challenging from the side below 2/3. +However, for σ = 0.75, Fig. 38 shows that R is clearly +increasing with N for large N values. But if we had had +only data for system sizes < 2048 we might have indeed +concluded that there was good evidence for an AT transi- +tion in that R seemed to be an N independent constant. +While at the sizes we can reach R is clearly increasing +with N it has yet to reach its asymptotic form of in- +crease as N z−(2σ−1). The quantity R also increases with +N for σ = 0.70 and σ = 0.85, (see for example Fig. 39). +In order for χSG to match as σ → 2/3 from either the +100 +101 +102 +N 1/xhr +10−4 +10−3 +10−2 +10−1 +χSG/N z +σ = 0.7 +T = 0.6 +x = 3.3220 +xχ = 1.5747 +z = 0.4740 +N = 8192 +N = 16384 +Figure 41. A complete finite size scaling plot of χSG a function +of magnetic field hr, for σ = 0.70 at a temperature of T = 0.6 +for our two largest system sizes. +mean-field side, (where z = 1/3) with its value in the +non-mean field region, we would expect that z should +approach 1/3 as σ → 2/3 from above. +At σ = 0.85, +z ≈ 0.8409, at σ = 0.75, z ≈ 0.6065, while at σ = 0.70, +we find z ≈ 0.4737. Thus it seems quite plausible that +z could approach 1/3 as σ → 2/3 from above. Then the +combination z − (2σ − 1) would approach zero in this +limit, which means that the divergence of R with N will +become harder and harder to see as σ approaches 2/3. +We conclude that it will be challenging to do numerical +work which shows that the AT line disappears at precisely +σ = 2/3. On the mean-field side of 2/3 the correction to +scaling exponent ω = 1/3 − (2σ − 1). It therefore seems +natural to expect that on the non-mean field regime +ω = z − (2σ − 1). +(36) +If valid, this would imply that corrections to scaling +should be larger at σ = 0.70 than at σ = 0.75, and this +is what we observed in Figs. 40 and 41, in comparison +with Figs. 30 and 33. +In the presence of a genuine AT transition, as hr is +reduced one would pass through three regions: first the +paramagnetic state at larger values of hr, then the criti- +cal region, then the low-temperature phase with RSB at +smaller values of hr. The good data collapse for all val- +ues of hr using Eq. (26), and Eq. (28) shows that at any +finite value of hr there is just one region, the paramag- +netic region. Studying “intersections��� as in Sec. V is an +attempt to find the critical region. But the intersections +at finite values of hr for σ = 0.75 and σ = 0.85 are not +signs of a genuine phase transition, but at least in the +case of χSG these crossings are also just a consequence of +droplet scaling. The behavior of h∗(N, 2N) as a function + +19 +10−2 +10−1 +100 +N 1/xhr +0.5 +1.0 +1.5 +2.0 +2.5 +3.0 +ξSG/N +σ = 0.75 +T = 0.55 +x = 2.7077 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +N = 8192 +N = 16384 +N = 32768 +Figure 42. A finite size scaling plot of ξSG as a function of +magnetic field hr, for σ = 0.75 at a temperature of T = 0.55. +of N is greatly complicated by finite size effects and will +only become clear at much larger N values than those +which we have been able to study. +Because on the droplet picture there is no AT line and +so one is always in the paramagnetic phase at any non- +zero field (just as in a ferromagnet). +However, length +scales like ξSG become very large as hr → 0 for temper- +atures T < Tc(hr = 0). Once they become comparable +to the system dimensions L and one is in the regime +hr < h∗(N, 2N), the system will have many of the fea- +tures which might be associated with being in the bro- +ken replica symmetric phase which is envisaged to exist +below the AT line. For physical systems in three dimen- +sions the relevant length scale is not the linear dimension +of the system L, but the linear dimension of a fully equi- +librated region. This may explain why both simulations +and experiments have failed for many years to resolve the +debate. +Might it be possible to find by simulations whether the +borderline between RSB ordering and droplet ordering is +at σ = 2/3, which is the equivalent of d = 6 with short- +range interactions? To this end we looked at the case +of σ = 0.70. We found from studying the crossings of +ξSG and χSG for the zero field case that the zero field +transition temperature is ≈ 0.724. Figs. 40 and 41 show +our attempt to collapse the data with the droplet scaling +forms. Clearly the effects of corrections to scaling are +larger than was the case at σ = 0.75 in Figs. 30 and +33. This is in accord with Eq. (36) which predicts that +the correction to scaling exponent ω will go to zero as +σ → 2/3 if also z → 1/3 as expected. We conclude that +it will be difficult to provide good numerical evidence +that σ = 2/3 is the lower critical dimension of the AT +transition. +10−2 +10−1 +100 +101 +N 1/xhr +0.5 +1.0 +1.5 +2.0 +ξSG/N +σ = 0.85 +T = 0.3 +x = 2.2019 +N = 128 +N = 256 +N = 512 +N = 1024 +N = 2048 +N = 4096 +Figure 43. A finite size scaling plot of ξSG as a function of +magnetic field hr, for σ = 0.85 at a temperature of T = 0.3. +VII. +TNT VERSUS THE DROPLET SCALING +PICTURE +Newman and Stein [50] (see also the recent review +[51]), have suggested that the ordered phase of spin +glasses in finite dimensions will fall into one of 4 cate- +gories, (and which one might depend on the dimension- +ality d of the system): The RSB state is one of these, +and is somewhat similar to that envisaged by Parisi for +the SK model, but there is also the chaotic pairs state +picture of Newman and Stein. In both of these pictures +there is an AT transition. The other two pictures are +the so-called TNT picture of Krzakala and Martin [35] +and Palassini and Young [36] and the droplet scaling pic- +ture [44–46]. In neither the TNT picture nor the droplet +scaling picture is there an AT transition. In the droplet +picture the Parisi overlap function P(q) is trivial, consist- +ing of two delta functions at ±qEA in zero field, whereas +in the TNT picture the form of P(q) is quite similar to +the non-trivial (NT) form which Parisi found for the SK +model. +The TNT picture accounts for the non-trivial +form of the Parisi overlap function by postulating that +there exist droplets of the linear size L of the system, +which contain O(Ld) spins, and which do not have a free +energy of order Lθ (as they would in the droplet scaling +picture), but which have instead a free energy of O(1). It +is the presence of such droplets which makes P(q) non- +trivial, which is a feature observed in all simulations of +it to date. +In a recent paper [3] one of us argued that once the +linear dimension of the system became larger than a +crossover length L∗ the non-trivial behavior observed in +P(q) will change to the trivial form predicted by droplet +scaling. Estimates of L∗ in d = 3 suggest it might be + +20 +large, of the order of several hundred lattice spacings +and it is probably the case that to date the regime where +L > L∗ has not been reached. Furthermore it was sug- +gested that as d → 6, L∗ would grow towards infinity, +as the droplets of O(1) evolve to the O(1) excitations in +the Parisi RSB solution, where the pure states have free +energies which differ from each other by O(1). In our +one dimensional proxy system we would therefore expect +to find that L∗ is much larger when σ = 0.75 than it is +when σ = 0.85. +In this paper there are TNT-like effects visible in the +behavior of ξSG/N in the region where ξSG is of O(N) +(see Figs. +28 and 29). +When ξSG is of order N the +droplets which are important are those of size N and if +L < L∗ some of these will have free energy of O(1) rather +than Lθ. As a consequence the good scaling collapse of +the data visible when ξSG/N ≪ 1 will be lost. In Figs. +42 and 43 we have plotted ξSG/N on a linear scale versus +N 1/xhr focussing only on the region where ξSG/N is of +O(1). If the droplet scaling collapse had been good and +of the form of Eq. (27) then as N is increased the collapse +should get better and better. In fact due to TNT effects +the data in Fig. 42 for σ = 0.75 show the opposite trend, +and the lines get further apart with increasing N in the +region where ξSG is of O(N). +However, for σ = 0.85 +Fig. 43 shows the lines seem to be getting closer with +increasing N. +It suggests that for this value of σ we +are getting into the region where L > L∗ when droplet +scaling applies even when ξSG is of O(N). Data at larger +values of N than 4096 would be nice to confirm this trend +but because these simulations have to be done at quite +low temperatures compared to those for σ = 0.75 it will +be challenging to do this. Despite this limitation on the +size of N which can be reached for σ = 0.85, there is +evidence that for it, TNT and finite size scaling effects +are less troublesome than for σ = 0.75, despite the fact +that much larger values of N can be studied at this σ +value. +VIII. +SUMMARY AND CONCLUSIONS +In this paper, we have studied the phase transitions +in the one-dimensional power-law diluted XY spin glass, +both in the zero-field limit, and in the presence of a mag- +netic field random in the component directions. Whether +or not an AT line exists for various values of the param- +eter σ is a question of fundamental interest. To address +this, we have performed large scale Monte-Carlo sim- +ulations using a new heatbath algorithm, described in +Appendix A. This algorithm hopefully speeds up equi- +libration, so cutting computational costs. We certainly +do gain some advantage in terms of computational time +due to the smaller number of components of XY spins +compared to those of the Heisenberg model. Alas, the +heatbath algorithm for XY spins suffers from an intrin- +sic disadvantage. Because our algorithm has to generate +two random numbers during each Monte Carlo step, the +benefits of the smaller number of components are largely +counterbalanced by the additional labor involved in the +heatbath step. We were unable to go to larger system +sizes than in the corresponding work with Heisenberg +spins [18]. +The largest system sizes that we are able +to simulate are: N = 16384 for σ = 0.6, N = 32768 for +σ = 0.75, while the largest N for σ = 0.85 was 4096. The +total CPU time spent in generating all the data that we +presented at fixed hr and varying T was 1183636.2 hrs, +which is 135.12 years. The total CPU time consumed in +generating the data at fixed T and varying hr was 96101.6 +days which is 263.29 years. Thus despite the algorithm +not producing significant dividends, we are able to study +fairly large system sizes owing to the expenditure of a +large amount of computer time. +The results from our work are broadly in accord with +those for the corresponding Heisenberg spin glass model. +For σ = 0.6, which is in the mean-field regime, we find +a phase transition in the absence of an external mag- +netic field, and in the presence of a magnetic field, which +indicates the existence of an AT line. The location of +the AT line is close to the mean-field predictions. For +σ = 0.75, which is in the non-mean-field regime, the con- +ventional data collapse suggests the existence of an AT +line, but the behavior of the intersections as a function +of N indicate that the data is not close to its large N +asymptotic form. The estimated location of the AT field +based upon intersections that we get from our data at +σ = 0.75 is strikingly smaller than estimates based on +the mean-field theory formulas. For σ = 0.85, which is +deep in the non-mean-field regime and corresponds to a +space dimension of about 3, our data are consistent with +the absence of an AT line. In this case there is no crossing +of the curves of χSG/N 2−η versus T at various N values. +But confusingly intersections h∗(N, 2N), as a function of +hr, seem to exist, whereas intersections T ∗(N, 2N) are +absent at least for σ = 0.85. +However, for σ = 0.75 and for σ = 0.85 we found that +the droplet picture provided a much better description +of our data from that obtained assuming the existence +of an AT transition line. The Imry-Ma formula for the +field dependence of ξSG works well until ξSG becomes +comparable to the system size. A similar behavior was +reported for the Ising spin glass at σ = 0.75 in Ref. [48]. +A finite-size scaling formulation was developed to treat +the data at small fields when ξSG is comparable to the +system size N, and with it an excellent collapse of all +our data on ξSG and χSG was obtained. We showed that +droplet scaling predicts the existence of the intersections +h∗(N, 2N). Our data unfortunately does not extend to +values of N large enough to be in the asymptotic region +where the N-dependence of h∗(N, 2N) is simple. Fortu- +nately there exists a way of testing whether the intersec- +tions are due to an AT transition or are just those pre- +dicted by droplet scaling, which is to study the N depen- +dence of R = χSG/N 2σ−1, calculated at h∗(N, 2N), and +this test supports the droplet picture provided N > 1024 +at σ = 0.75. Thus it is only for large systems that one + +21 +can obtain good evidence for the droplet picture. +We now summarize our main results. The strongest +evidence for droplet scaling is the success of the Imry-Ma +formula for the field dependence of ξSG for σ = 0.75 and +σ = 0.85 (see Figs. 30 and 31). If droplet scaling works, +then no AT line is to be expected. When ξSG ∼ N there +are visible sizeable corrections to the Imry-Ma formula +which are related to TNT effects. However for σ = 0.85 +there is tentative evidence in Fig. 43 that if even larger +systems could be studied then the TNT effects might be +absent, and so there could exist a length scale L∗ above +which TNT effects become unimportant (see Ref. [3]). +If instead of droplet scaling one assumes that there is +an AT phase transition then the usual finite size scaling +plots used to determine hAT as in Fig. 15 for σ = 0.75 +are unsatisfactory: for example the values of hAT which +would be derived from the crossings of ξSG and χSG as +N becomes large look to be significantly different. In the +equivalent data plot for σ = 0.60 (see Fig. +8) they are +in good agreement. Furthermore the quantity R of Eq. +(35) should approach a constant as N → ∞ if there is a +genuine AT transition, but instead for the cases σ = 0.75 +(Fig. 38) and 0.85 (Fig. 39), it is increasing with N once +N becomes large enough. +The simulations of this paper provide numerical evi- +dence that the AT line and hence RSB is absent in spin +glasses below six dimensions. What is now needed is an +explanation of why this might be the case. Better still +would be a rigorous proof that the lower critical dimen- +sion for replica symmetry breaking is six. Our work indi- +cates that showing that σ = 2/3 is the precise value of the +critical value of σ will be challenging using simulations +as finite size effects are large in its vicinity. +ACKNOWLEDGMENTS +We are grateful to the High Performance Comput- +ing (HPC) facility at IISER Bhopal, +where large- +scale +calculations +in +this +project +were +run. +We +thank Peter Young and Dan Stein for helpful discus- +sions. +B.V is grateful to the Council of Scientific +and Industrial Research (CSIR), India, for his PhD +fellowship. +A.S acknowledges financial support from +SERB via the grant (File Number: CRG/2019/003447), +and from DST via the DST-INSPIRE Faculty Award +[DST/INSPIRE/04/2014/002461]. +Appendix A: The simulation method +We now give some technical aspects of how the simula- +tions are run. In the simulations we start with a random +initial configuration and allow it to evolve according to +the prescription given in this section. To incorporate par- +allel tempering, we simultaneously simulate NT copies of +the system over NT different temperatures ranging from +Tmin ≡ T1 to Tmax ≡ TNT . +In order to facilitate the +computation of the observables outlined in this section, +it is convenient to simulate 4 sets of NT copies (2 for +hr = 0), which we label (1),(2),(3), and (4). We perform +overrelaxation, heatbath and parallel tempering sweeps +over all these copies keeping track of the labels appropri- +ately. +For every 10 overrelaxation sweeps we perform +1 heatbath and 1 parallel tempering sweep, since the +overrelaxation sweep involves a significantly lower com- +putational cost, and is known to speed up equilibration. +The parameters of the simulations are shown in Tables II +and III. Once the system reaches equilibrium, we perform +the same number of sweeps in the measurement phase, +so Nsweep is the total number of sweeps over which the +simulation is run, inclusive of both the equilibration and +measurement phases. The last column in the table shows +the amount of computer time expended to generate the +data corresponding to the parameters in that row. +In +the measurement phase, we perform one measurement +on the system for every 4 sweeps. The following sections +contain the details of our Monte Carlo simulation pro- +cedures. +In order to equilibrate the system as quickly +as possible, we perform three kinds of sweeps: overre- +laxation or microcanonical sweeps, heatbath sweeps, and +parallel tempering sweeps. +1. +Overrelaxation sweep +We sweep sequentially through all the lattice sites and +compute the local field Hi = � +j JijSj+hi at a particular +lattice site. The new spin direction S′ +i at the ith lattice +site is taken to be the mirror image of the vector Si about +Hi, i.e., +S′ +i = −Si + 2Si · Hi +H2 +i +Hi. +(A1) +Since S′ +i · Hi = Si · Hi, the energy of the system does +not change due to these sweeps. Hence these sweeps are +also called microcanonical sweeps. These sweeps help us +in sampling out the microstates with the same energy. +The process of equilibration speeds up when we include +overrelaxation sweeps along with the other sweeps [33, +52]. +2. +Heatbath sweep +The overrelaxation sweeps generate states with the +same energy and hence they cannot directly equilibrate +the system. Therefore, we also perform a heatbath sweep +for every 10 microcanonical sweeps. Similar to the micro- +canonical case, we sweep sequentially through the lattice. +To equilibrate the system, the angle θ between Hi and +S +′ +i should be sampled out from the Boltzmann distribu- +tion given by +fΘ(θ) = e−βEi +Z += eβHiSi cos θ +Z += ew cos θ +Z +, +(A2) + +22 +where w = βHiSi and +Z = +π +� +−π +eβHiSi cos θ dθ +(A3) +is the normalizing constant. The simplest way to do this +is to equate the cumulative density function (CDF) of θ, +FΘ(θ), to that of a uniform distribution: +FΘ(θ) = +θ +� +−π +fΘ(θ′) dθ′ = Π(r1) = r1, +(A4) +where r1 is a random variable sampled from a uniform +distribution in the interval (0, 1). The value of θ can be +obtained by simply inverting this function to get +θ = F −1 +Θ (r1). +(A5) +This method works well with the Heisenberg spins +as fΘ(θ) is integrable, which gives an invertible CDF +FΘ(θ) [18, 28]. Since the probabililty density function +(PDF) fΘ(θ) for the XY spin glasses given by Eq. (A2) +is not exactly integrable, this method cannot be used. +To overcome this problem and to sample out θ from the +Boltzmann distribution (Eq. (A2)) in as few a number of +sweeps as possible, we develop a heatbath sweep based +on the rejection method [53]. We generate two random +numbers r1 ∈ uniform(−π, π) and r2 ∈ uniform(0, fmax). +If r2 < fθ(r1), we accept the move, i.e., take θ = r1. Else, +we reject the move and generate another pair of random +numbers (r1, r2). This process is repeated until we find +an acceptable value of r1. A graphical representation for +this method is shown in Fig. 44. The new spin direction +S′ +i in Cartesian co-ordinates is given by: +S′ +x = cos(θ + θH), +(A6a) +S′ +y = sin(θ + θH), +(A6b) +where θH is the angle made by the Hi vector with the X- +axis. Since the generation of random numbers is involved, +this sweep is computationally costlier than others. Hence +we perform more microcanonical sweeps than heatbath +sweeps. +3. +Parallel tempering sweep +Spin glasses have a complex free energy landscape due +to which, at low temperatures, they tend to get stuck in- +side metastable valleys, and true equilibration consumes +a lot of time. At high temperatures, the system can eas- +ily escape the valley due to thermal fluctuations, and +so equilibration is quick. To equilibrate the system in +as small a number of moves as possible, we perform +one parallel tempering sweep for every 10 overrelaxation +sweeps [28, 33]. To benefit from the parallel tempering +algorithm [54, 55], we simultaneously run the simulation +−2 +0 +2 +θ +0.0 +0.1 +0.2 +0.3 +0.4 +fΘ(θ) +fmax +(r1, r2) +(r1, fΘ(r1)) +(r1, r2) +(r1, fΘ(r1)) +w = βHS = 1.5 +Rejected +Accepted +Figure 44. Graphical representation of the rejection method. +We randomly pick a point (r1, r2) within the rectangle from a +uniform distribution. If the point lies under the fΘ(θ) curve +given by Eq. (A2), then the point is accepted, and θ is taken +to be r1. Otherwise, the point is rejected. +for NT copies of the system at NT different temperatures +T1 < T2 < T3 < · · · < TNT . The minimum temperature +T1 is the low temperature at which we are interested in +studying the behavior of the system, and the maximum +temperature TNT is high enough that the system equi- +librates very fast. We perform overrelaxation and heat- +bath sweeps separately on each of the NT copies of the +system. In the parallel tempering sweep, we compare the +energies of two spin configurations at adjacent tempera- +tures, Ti and Ti+1, starting from the smallest tempera- +ture T1. We swap these two spin configurations such that +the detailed balance condition is satisfied. The Metropo- +lis probability for such a swap is +P(T swap) = min{1, exp(∆β∆E)} +(A7) += +� +exp(∆β∆E) +(if ∆β∆E < 0), +1 +(otherwise), +(A8) +where ∆β += +1/Ti − 1/Ti+1 and ∆E += +Ei(Ti) − +Ei+1(Ti+1). In this way, a given set of spins performs +a random walk in temperature space. +4. +Checks for equilibration +In order to check whether the system has reached equi- +librium, we have used a convenient test [56] which is pos- +sible because of the Gaussian nature of the interactions +and the onsite external magnetic field. The relation +U = zJ2 +2T (ql − qs) + h2 +r +T +� +q − |S|2� +, +(A9) + +23 +Table II. Parameters of the simulations. Nsamp is the number of disorder samples, Nsweep is the number of over-relaxation +Monte Carlo sweeps for a single disorder sample. The system is equilibrated over the first half of the sweeps, and measurements +are done over the last half of the sweeps with a measurement performed every four over-relaxation sweeps. Tmin and Tmax are +the lowest and highest temperatures simulated, and NT is the number of temperatures used for parallel tempering. +σ +hr +N +Nsamp +Nsweep +Tmin +Tmax +NT +ttot(hrs) +0.6 +0 +128 +10000 +512 +0.6 +1 +18 +0.49 +0.6 +0 +256 +8000 +1024 +0.6 +1 +22 +2.23 +0.6 +0 +512 +6400 +2048 +0.6 +1 +22 +6.46 +0.6 +0 +1024 +8000 +4096 +0.6 +1 +26 +40.74 +0.6 +0 +2048 +3840 +8192 +0.6 +1 +24 +105.41 +0.6 +0 +4096 +3200 +16384 +0.6 +1 +27 +571.49 +0.6 +0 +8192 +3200 +32768 +0.6 +1 +30 +3776.85 +0.6 +0 +16384 +2600 +65536 +0.64 +0.98 +32 +18225.5 +0.6 +0.1 +128 +9600 +2048 +0.5 +0.8 +21 +7.75 +0.6 +0.1 +256 +9600 +2048 +0.5 +0.8 +21 +15.89 +0.6 +0.1 +512 +9600 +8192 +0.5 +0.8 +22 +85.67 +0.6 +0.1 +1024 +8000 +16384 +0.5 +0.8 +22 +414.47 +0.6 +0.1 +2048 +7200 +32768 +0.5 +0.8 +26 +2029.23 +0.6 +0.1 +4096 +7200 +65536 +0.5 +0.8 +24 +10014.8 +0.6 +0.1 +8192 +4380 +131072 +0.55 +0.8 +25 +34810.6 +0.6 +0.1 +16384 +7128 +262144 +0.55 +0.8 +28 +224425 +0.75 +0 +128 +12800 +1024 +0.35 +0.85 +21 +1.6 +0.75 +0 +256 +12800 +2048 +0.35 +0.85 +24 +7.21 +0.75 +0 +512 +8000 +8192 +0.35 +0.85 +24 +35.22 +0.75 +0 +1024 +8000 +16384 +0.35 +0.85 +24 +196.9 +0.75 +0 +2048 +6400 +32768 +0.35 +0.85 +25 +774.55 +0.75 +0 +4096 +4880 +65536 +0.35 +0.85 +27 +3405.2 +0.75 +0 +8192 +3000 +131072 +0.38 +0.82 +30 +14290.9 +0.75 +0.05 +128 +19200 +8192 +0.28 +0.6 +21 +45.27 +0.75 +0.05 +256 +16000 +16384 +0.28 +0.6 +20 +133.35 +0.75 +0.05 +512 +13600 +32768 +0.28 +0.6 +20 +464.77 +0.75 +0.05 +1024 +11000 +65536 +0.28 +0.6 +21 +2075.57 +0.75 +0.05 +2048 +10920 +262144 +0.28 +0.6 +24 +21314.3 +0.75 +0.05 +4096 +10800 +524288 +0.3 +0.58 +26 +123093 +0.75 +0.05 +8192 +5320 +1048576 +0.32 +0.54 +32 +364358 +0.85 +0 +128 +12800 +8192 +0.2 +0.5 +30 +17.97 +0.85 +0 +256 +12800 +16384 +0.2 +0.5 +32 +72.15 +0.85 +0 +512 +12800 +65536 +0.2 +0.5 +30 +752.36 +0.85 +0 +1024 +12800 +131072 +0.2 +0.5 +30 +3219.93 +0.85 +0 +2048 +8000 +262144 +0.2 +0.5 +30 +9504.05 +0.85 +0 +4096 +6480 +524288 +0.24 +0.48 +30 +40322.4 +0.85 +0.02 +128 +8000 +65536 +0.1 +0.4 +30 +194.33 +0.85 +0.02 +256 +4000 +131072 +0.1 +0.4 +32 +470.39 +0.85 +0.02 +512 +4400 +524288 +0.1 +0.4 +34 +4780.6 +0.85 +0.02 +1024 +3000 +2097152 +0.1 +0.4 +35 +30356.9 +0.85 +0.02 +2048 +1800 +4194304 +0.16 +0.4 +36 +84056 +0.85 +0.05 +128 +2000 +65536 +0.1 +0.4 +30 +67.07 +0.85 +0.05 +256 +4000 +131072 +0.1 +0.4 +32 +604.8 +0.85 +0.05 +512 +3500 +524288 +0.1 +0.4 +36 +4958.17 +0.85 +0.05 +1024 +3120 +2097152 +0.1 +0.4 +36 +28028.1 +0.85 +0.05 +2048 +3240 +4194304 +0.16 +0.4 +36 +151503 +is valid in equilibrium. Here +U = 1 +N [⟨H⟩]av += − 1 +N +� +�� +⟨i,j⟩ +ϵijJij ⟨Si · Sj⟩ + +� +i,µ +hµ +i ⟨Sµ +i ⟩ +� +� +av +(A10) +is the average energy per spin, q = +1 +N +� +i +[⟨Si⟩ · ⟨Si⟩]av +is +the +Edwards-Anderson +order +parameter, +ql += +1 +Nb +� +⟨i,j⟩ +� +ϵij ⟨Si · Sj⟩2� +av is the “link overlap”, and qs = +1 +Nb +� +⟨i,j⟩ +� +ϵij +� +(Si · Sj)2�� +av is the “spin overlap”, where + +24 +Table III. Parameters of the simulations done at fixed temperature T and varying field hr. N(hr) is the number of values of +field taken in the range hr(min,max). The equilibration times are different for different values of the field hr, which lie in the +range Nsweep(min,max). The number of disorder samples for different fields lie in the range Nsamp(min,max). ttot is the total +CPU time consumed in hours to generate data for a particular system size. +σ +T +N +hr(min,max) +N(hr) +Nsweep(min,max) +Nsamp(min,max) +ttot(hrs) +0.6 +0.6 +128 +(0.010, 9.000) +32 +(2048, 2048) +(2000, 64000) +11.44 +0.6 +0.6 +256 +(0.010, 9.000) +32 +(4096, 4096) +(2000, 48000) +55.5 +0.6 +0.6 +512 +(0.010, 9.000) +32 +(8192, 16384) +(2000, 80000) +163.91 +0.6 +0.6 +1024 +(0.010, 9.000) +32 +(4096, 65536) +(2000, 80000) +2375 +0.6 +0.6 +2048 +(0.010, 9.000) +32 +(16384, 131072) +(1200, 60000) +10259.8 +0.6 +0.6 +4096 +(0.010, 9.000) +32 +(65536, 2097152) +(960, 28600) +92052.1 +0.6 +0.6 +8192 +(0.010, 9.000) +32 +(65536, 4194304) +(400, 19428) +310948 +0.6 +0.6 +16384 +(0.010, 9.000) +32 +(131072, 4194304) +(488, 10689) +1021481 +0.7 +0.6 +128 +(0.010, 9.000) +31 +(1024, 1024) +(4000, 4000) +1.86 +0.7 +0.6 +256 +(0.010, 9.000) +31 +(2048, 2048) +(1000, 8000) +8.53 +0.7 +0.6 +512 +(0.010, 9.000) +31 +(4096, 4096) +(1000, 8000) +39.46 +0.7 +0.6 +1024 +(0.010, 9.000) +31 +(8192, 8192) +(1500, 8000) +288.41 +0.7 +0.6 +2048 +(0.010, 9.000) +31 +(16384, 16384) +(500, 8000) +559.99 +0.7 +0.6 +4096 +(0.010, 9.000) +31 +(32768, 32768) +(400, 8000) +3020.62 +0.7 +0.6 +8192 +(0.010, 9.000) +31 +(16384, 131072) +(2000, 6800) +17500.2 +0.7 +0.6 +16384 +(0.010, 9.000) +31 +(32768, 262144) +(640, 6400) +77734.3 +0.75 +0.55 +128 +(0.001, 9.000) +42 +(512, 1024) +(2000, 40000) +6.59 +0.75 +0.55 +256 +(0.001, 9.000) +42 +(1024, 2048) +(2000, 40000) +56.67 +0.75 +0.55 +512 +(0.001, 9.000) +42 +(4096, 4096) +(1000, 24000) +165.51 +0.75 +0.55 +1024 +(0.001, 9.000) +43 +(8192, 8192) +(1000, 12000) +255.75 +0.75 +0.55 +2048 +(0.001, 9.000) +43 +(16384, 16384) +(1000, 12000) +1000.52 +0.75 +0.55 +4096 +(0.001, 9.000) +43 +(32768, 32768) +(800, 12000) +5269.11 +0.75 +0.55 +8192 +(0.001, 9.000) +42 +(65536, 65536) +(800, 8000) +21410.3 +0.75 +0.55 +16384 +(0.001, 9.000) +42 +(131072, 131072) +(760, 6280) +62062 +0.75 +0.55 +32768 +(0.001, 9.000) +42 +(131072, 262144) +(512, 4995) +146627 +0.85 +0.3 +128 +(0.001, 9.000) +36 +(32768, 131072) +(2000, 30000) +264.97 +0.85 +0.3 +256 +(0.001, 9.000) +36 +(65536, 262144) +(1000, 25000) +1052.21 +0.85 +0.3 +512 +(0.001, 9.000) +36 +(131072, 524288) +(1000, 34800) +6161.52 +0.85 +0.3 +1024 +(0.001, 9.000) +36 +(262144, 1048576) +(1000, 20000) +17668 +0.85 +0.3 +2048 +(0.001, 9.000) +36 +(524288, 8388608) +(320, 3372) +68933.6 +0.85 +0.3 +4096 +(0.001, 9.000) +36 +(262144, 16777216) +(312, 5760) +439004 +Nb = Nz/2, and ϵij = 1 if the ith and jth spins are in- +teracting and is zero otherwise. The [· · · ]av in Eq. (A10) +is analytically evaluated by performing integration over +Jij and hµ +i [57] since they have Gaussian distributions. +On evaluating this integral using integration by parts, +we get Eq. (A9). As the system reaches equilibrium, the +two sides of Eq. (A9) approach their common equilibrium +value from opposite directions. +In simulations, we evaluate both sides of Eq. (A9) for +different number of Monte-Carlo sweeps (MCSs), which +increase in an exponential manner, each value being twice +the previous one. The averaging is done over the last half +of the sweeps. We initially start with a random spin con- +figuration, so the LHS of Eq. (A9) is small and the RHS +is very large. As the system gets closer to equilibrium, +these two values come closer to each other from opposite +directions. +When we notice that the averaged quanti- +ties satisfy Eq. (A9) within error bars, consistently for +at least the last two points, we declare that our system +has reached equilibrium. Once the system reaches equi- +librium, we perform the same number of sweeps in the +measurement phase, where we evaluate different quanti- +ties (given below) used to study the possible phase tran- +sitions of the system. +[1] Marc M´ezard, Giorgio Parisi, +and Miguel Angel Vira- +soro, Spin glass theory and beyond: An Introduction to +the Replica Method and Its Applications, Vol. 9 (World +Scientific Publishing Company, 1986). +[2] J R L de Almeida and D J Thouless, “Stability of the +sherrington-kirkpatrick solution of a spin glass model,” +Journal of Physics A: Mathematical and General 11, +983–990 (1978). + +25 +[3] M. A. Moore, “Droplet-scaling versus replica symmetry +breaking debate in spin glasses revisited,” Phys. Rev. E +103, 062111 (2021). +[4] V. Martin-Mayor, J. J. Ruiz-Lorenzo, B. Seoane, +and +A. P. Young, “Numerical simulations and replica sym- +metry breaking,” (2022). +[5] Auditya Sharma and A. P. Young, “de Almeida–Thouless +line in vector spin glasses,” Phys. Rev. E 81, 061115 +(2010). +[6] A J Bray and S A Roberts, “Renormalisation-group ap- +proach to the spin glass transition in finite magnetic +fields,” Journal of Physics C: Solid State Physics 13, +5405–5412 (1980). +[7] M. A. Moore and A. J. Bray, “Disappearance of the de +Almeida-Thouless line in six dimensions,” Phys. Rev. B +83, 224408 (2011). +[8] G. Kotliar, P. W. Anderson, +and D. L. Stein, “One- +dimensional spin-glass model with long-range random in- +teractions,” Phys. Rev. B 27, 602–605 (1983). +[9] Freeman J. Dyson, “Non-existence of spontaneous mag- +netization in a one-dimensional Ising ferromagnet,” +Communications in Mathematical Physics 12, 212–215 +(1969). +[10] Freeman J. Dyson, “An Ising ferromagnet with discontin- +uous long-range order,” Communications in Mathemati- +cal Physics 21, 269–283 (1971). +[11] L. Leuzzi, G. Parisi, F. Ricci-Tersenghi, and J. J. Ruiz- +Lorenzo, “Dilute One-Dimensional Spin Glasses with +Power Law Decaying Interactions,” Phys. Rev. Lett. 101, +107203 (2008). +[12] L. Leuzzi, G. Parisi, F. Ricci-Tersenghi, and J. J. Ruiz- +Lorenzo, “Ising Spin-Glass Transition in a Magnetic Field +Outside the Limit of Validity of Mean-Field Theory,” +Phys. Rev. Lett. 103, 267201 (2009). +[13] A. P. Young and Helmut G. Katzgraber, “Absence of +an Almeida-Thouless Line in Three-Dimensional Spin +Glasses,” Phys. Rev. Lett. 93, 207203 (2004). +[14] Helmut G. Katzgraber and A. P. Young, “Probing the +Almeida-Thouless line away from the mean-field model,” +Phys. Rev. B 72, 184416 (2005). +[15] Helmut G. Katzgraber, Derek Larson, and A. P. Young, +“Study of the de Almeida–Thouless Line Using Power- +Law Diluted One-Dimensional Ising Spin Glasses,” Phys. +Rev. Lett. 102, 177205 (2009). +[16] Helmut G. Katzgraber and A. P. Young, “Monte Carlo +studies of the one-dimensional Ising spin glass with +power-law interactions,” Phys. Rev. B 67, 134410 (2003). +[17] Auditya Sharma and A. P. Young, “Phase transitions in +the one-dimensional long-range diluted Heisenberg spin +glass,” Phys. Rev. B 83, 214405 (2011). +[18] Auditya Sharma and A. P. Young, “de Almeida–Thouless +line studied using one-dimensional power-law diluted +Heisenberg spin glasses,” Phys. Rev. B 84, 014428 (2011). +[19] Auditya Sharma, Joonhyun Yeo, +and M. A. Moore, +“Metastable minima of the Heisenberg spin glass in a +random magnetic field,” Phys. Rev. E 94, 052143 (2016). +[20] Frank Beyer, Martin Weigel, +and M. A. Moore, “One- +dimensional infinite-component vector spin glass with +long-range interactions,” Phys. Rev. B 86, 014431 (2012). +[21] M. A. Moore, “1/m expansion in spin glasses and the de +Almeida-Thouless line,” Phys. Rev. E 86, 031114 (2012). +[22] Auditya Sharma, Alexei Andreanov, and Markus M¨uller, +“Avalanches and hysteresis in frustrated superconductors +and XY spin glasses,” Phys. Rev. E 90, 042103 (2014). +[23] Silvio Franz, Flavio Nicoletti, Giorgio Parisi, +and Fed- +erico Ricci-Tersenghi, “Delocalization transition in low +energy excitation modes of vector spin glasses,” SciPost +Phys. 12, 16 (2022). +[24] Cosimo Lupo and Federico Ricci-Tersenghi, “Compari- +son of Gabay–Toulouse and de Almeida–Thouless insta- +bilities for the spin-glass XY model in a field on sparse +random graphs,” Phys. Rev. B 97, 014414 (2018). +[25] Cosimo Lupo, +Giorgio Parisi, +and Federico Ricci- +Tersenghi, “The random field XY model on sparse +random graphs shows replica symmetry breaking and +marginally stable ferromagnetism,” Journal of Physics A: +Mathematical and Theoretical 52, 284001 (2019). +[26] M. Baity-Jesi, V. Mart´ın-Mayor, G. Parisi, and S. Perez- +Gaviro, “Soft Modes, Localization, and Two-Level Sys- +tems in Spin Glasses,” Phys. Rev. Lett. 115, 267205 +(2015). +[27] L. A. Fernandez, V. Martin-Mayor, S. Perez-Gaviro, +A. Tarancon, +and A. P. Young, “Phase transition in +the three dimensional Heisenberg spin glass: Finite-size +scaling analysis,” Phys. Rev. B 80, 024422 (2009). +[28] L. W. Lee and A. P. Young, “Large-scale Monte Carlo +simulations of the isotropic three-dimensional Heisenberg +spin glass,” Phys. Rev. B 76, 024405 (2007). +[29] I. Campos, M. Cotallo-Aban, V. Martin-Mayor, S. Perez- +Gaviro, and A. Tarancon, “Spin-Glass Transition of the +Three-Dimensional Heisenberg Spin Glass,” Phys. Rev. +Lett. 97, 217204 (2006). +[30] J. A. Olive, A. P. Young, and D. Sherrington, “Computer +simulation of the three-dimensional short-range Heisen- +berg spin glass,” Phys. Rev. B 34, 6341–6346 (1986). +[31] K. Hukushima and H. Kawamura, “Chiral-glass tran- +sition +and +replica +symmetry +breaking +of +a +three- +dimensional Heisenberg spin glass,” Phys. Rev. E 61, +R1008–R1011 (2000). +[32] Dao Xuan Viet and Hikaru Kawamura, “Numerical +Evidence of Spin-Chirality Decoupling in the Three- +Dimensional Heisenberg Spin Glass Model,” Phys. Rev. +Lett. 102, 027202 (2009). +[33] J. H. Pixley and A. P. Young, “Large-scale Monte Carlo +simulations of the three-dimensional XY +spin glass,” +Phys. Rev. B 78, 014419 (2008). +[34] M Baity-Jesi, +R A Ba˜nos, +A Cruz, +L A Fernan- +dez, J M Gil-Narvion, A Gordillo-Guerrero, D I˜niguez, +A Maiorano, F Mantovani, E Marinari, V Martin- +Mayor, J Monforte-Garcia, A Mu˜noz Sudupe, D Navarro, +G Parisi, S Perez-Gaviro, M Pivanti, F Ricci-Tersenghi, +J J Ruiz-Lorenzo, S F Schifano, B Seoane, A Tarancon, +R Tripiccione, +and D Yllanes, “The three-dimensional +Ising spin glass in an external magnetic field: the role +of the silent majority,” Journal of Statistical Mechanics: +Theory and Experiment 2014, P05014 (2014). +[35] F. Krzakala and O. C. Martin, “Spin and Link Overlaps +in Three-Dimensional Spin Glasses,” Phys. Rev. Lett. 85, +3013–3016 (2000). +[36] Matteo Palassini and A. P. Young, “Nature of the Spin +Glass State,” Phys. Rev. Lett. 85, 3017–3020 (2000). +[37] David Sherrington and Scott Kirkpatrick, “Solvable +Model of a Spin-Glass,” Phys. Rev. Lett. 35, 1792–1796 +(1975). +[38] J R L de Almeida, R C Jones, J M Kosterlitz, +and +D J Thouless, “The infinite-ranged spin glass with m- +component spins,” Journal of Physics C: Solid State +Physics 11, L871–L875 (1978). + +26 +[39] R. A. Ba˜nos, L. A. Fernandez, V. Martin-Mayor, +and +A. P. Young, “Correspondence between long-range and +short-range spin glasses,” Phys. Rev. B 86, 134416 +(2012). +[40] Derek Larson, Helmut G. Katzgraber, M. A. Moore, and +A. P. Young, “Numerical studies of a one-dimensional +three-spin spin-glass model with long-range interac- +tions,” Phys. Rev. B 81, 064415 (2010). +[41] K. Binder, “Finite size scaling analysis of Ising model +block distribution functions,” Zeitschrift f¨ur Physik B +Condensed Matter 43, 119–140 (1981). +[42] H.G. Ballesteros, L.A. Fern´andez, V. Mart´ın-Mayor, and +A. Mu˜noz Sudupe, “Finite size effects on measures of +critical exponents in d = 3 O(N) models,” Physics Letters +B 387, 125–131 (1996). +[43] Martin Hasenbusch, Andrea Pelissetto, +and Ettore Vi- +cari, “Critical behavior of three-dimensional Ising spin +glass models,” Phys. Rev. B 78, 214205 (2008). +[44] W L McMillan, “Scaling theory of Ising spin glasses,” +Journal of Physics C: Solid State Physics 17, 3179–3187 +(1984). +[45] A. J. Bray and M. A. Moore, “Scaling Theory of the +ordered phase of spin glasses,” in Heidelberg Collo- +quium on Glassy Dynamics and Optimization, edited +by L. Van Hemmen and I. Morgenstern (Springer, New +York, 1986) p. 121. +[46] Daniel S. Fisher and David A. Huse, “Equilibrium be- +havior of the spin-glass ordered phase,” Phys. Rev. B +38, 386–411 (1988). +[47] Yoseph Imry and Shang-keng Ma, “Random-Field Insta- +bility of the Ordered State of Continuous Symmetry,” +Phys. Rev. Lett. 35, 1399–1401 (1975). +[48] T. Aspelmeier, Helmut G. Katzgraber, Derek Larson, +M. A. Moore, Matthew Wittmann, and Joonhyun Yeo, +“Finite-size critical scaling in Ising spin glasses in the +mean-field regime,” Phys. Rev. E 93, 032123 (2016). +[49] T. Aspelmeier, Wenlong Wang, M. A. Moore, and Hel- +mut G. Katzgraber, “Interface free-energy exponent in +the one-dimensional Ising spin glass with long-range in- +teractions in both the droplet and broken replica sym- +metry regions,” Phys. Rev. E 94, 022116 (2016). +[50] C. M. Newman and D. L. Stein, “Finite-Dimensional Spin +Glasses: States, Excitations, and Interfaces,” Ann. Henri +Poincar´e 4, 497 (2003). +[51] C. M. Newman, N. Read, and D. L. Stein, “Metastates +and Replica Symmetry Breaking,” (2022). +[52] J. L. Alonso, A. Taranc´on, H. G. Ballesteros, L. A. +Fern´andez, V. Mart´ın-Mayor, +and A. Mu˜noz Sudupe, +“Monte Carlo study of O(3) antiferromagnetic models in +three dimensions,” Phys. Rev. B 53, 2537–2545 (1996). +[53] Richard C Larson and Amedeo R Odoni, Urban opera- +tions research, Monograph (1981). +[54] Koji Hukushima and Koji Nemoto, “Exchange monte +carlo method and application to spin glass simulations,” +Journal of the Physical Society of Japan 65, 1604–1608 +(1996). +[55] J. Machta, “Strengths and weaknesses of parallel tem- +pering,” Phys. Rev. E 80, 056706 (2009). +[56] Helmut G. Katzgraber, Matteo Palassini, +and A. P. +Young, “Monte Carlo simulations of spin glasses at low +temperatures,” Phys. Rev. B 63, 184422 (2001). +[57] A J Bray and M A Moore, “Some observations on the +mean-field theory of spin glasses,” Journal of Physics C: +Solid State Physics 13, 419–434 (1980). +