diff --git "a/QNFPT4oBgHgl3EQfojUh/content/tmp_files/2301.13134v1.pdf.txt" "b/QNFPT4oBgHgl3EQfojUh/content/tmp_files/2301.13134v1.pdf.txt" new file mode 100644--- /dev/null +++ "b/QNFPT4oBgHgl3EQfojUh/content/tmp_files/2301.13134v1.pdf.txt" @@ -0,0 +1,4138 @@ +arXiv:2301.13134v1 [math.RA] 30 Jan 2023 +The fundamental theorem of calculus in +differential rings +Clemens G. Raaba,∗ and Georg Regensburgera,b +aInstitute for Algebra, Johannes Kepler University Linz, Austria +bInstitute of Mathematics, University of Kassel, Germany +clemensr@algebra.uni-linz.ac.at +regensburger@mathematik.uni-kassel.de +Abstract +In this paper, we study the fundamental theorem of calculus and its +consequences from an algebraic point of view. For functions with singu- +larities, this leads to a generalized notion of evaluation. We investigate +properties of such integro-differential rings and discuss many examples. +We also construct corresponding integro-differential operators and pro- +vide normal forms via rewrite rules. They are then used to derive several +identities and properties in a purely algebraic way, generalizing well-known +results from analysis. In identities like shuffle relations for nested integrals +and the Taylor formula, additional terms are obtained that take singular- +ities into account. Another focus lies on treating basics of linear ODEs +in this framework of integro-differential operators. These operators can +have matrix coefficients, which allow to treat systems of arbitrary size in a +unified way. In the appendix, using tensor reduction systems, we give the +technical details of normal forms and prove them for operators including +other functionals besides evaluation. +Keywords +Integro-differential rings, integro-differential operators, normal +forms, generalized shuffle relations, generalized Taylor formula +1 +Introduction +Differential rings are a well-established algebraic structure for modelling dif- +ferentiation by derivations, i.e. linear operations satisfying the Leibniz rule. +More recently, integro-differential algebras have been introduced to additionally +model integration and point evaluation of continuous univariate functions by +∗Corresponding author +1 + +linear operations satisfying corresponding algebraic identities. In contrast, the +integro-differential rings introduced in this paper only use the two identities +d +dx +� x +a +f(t) dt = f(x) +and +� x +a +f ′(t) dt = f(x) − f(a) +of the fundamental theorem of calculus and the Leibniz rule as axioms. +In +particular, this results in a generalized notion of evaluation that is only required +to map to constants. This allows to deal with evaluations of functions even if +singularities or discontinuities are present. For example, it is natural to consider +integro-differential rings containing the rational functions leading to so-called +hyperlogarithms. +It turns out that many analytic identities and general statements about +ODEs, like variation of constants, have a purely algebraic proof in integro- +differential rings that is independent of the analytic properties of concrete +functions. Likewise, computations with and transformations of linear integro- +differential equations and initial conditions can be done in this algebraic setting +as well. +Moreover, exploring algebraic consequences of the Leibniz rule and +the fundamental theorem of calculus, we also find new results that introduce +additional evaluation terms into identities like shuffle relations for nested inte- +grals and the Taylor formula. In short, we investigate the analytic operations +of differentiation, integration, and evaluation from a purely algebraic point of +view. In this context, we implement in some sense the somewhat provocative +statement of Rota [33, p. 57] that “the algebraic structure sooner or later comes +to dominate [. . . ]. Algebra dictates the analysis.” +In order to study linear integro-differential equations and identities, we use +the operator point of view, which treats identities of functions as identities of +corresponding linear operators acting on them. To this end, we algebraically +construct the ring of integro-differential operators. For simplifying expressions +for operators, we use identities as rewrite rules. As a key result, we work out a +particular rewrite system that can be applied straightforwardly to obtain nor- +mal forms and to prove any algebraic identity of integro-differential operators. +Our construction of the ring of operators can be used for operators with scalar +coefficients and for operators with matrix coefficients. In particular, it allows to +uniformly deal with scalar equations as well as with systems, even of undeter- +mined size. Preliminary versions of some results presented in this paper have +already been presented by the authors at the conference “Differential Algebra +and Related Topics” (DART VII) in 2016. +A related approach was taken in the work of Danuta Przeworska-Rolewicz, +see for example [23]. In short, she considers a right invertible linear operator on +a vector subspace as a generalization of derivation. Right inverses and projec- +tions onto the kernel take the role of integration and evaluation, respectively. In +this linear setting, she develops algebraic generalizations of results for calculus +and linear differential equations. She refers to this as algebraic analysis, see +also [24] for historic context and references. Several results, e.g. a Taylor for- +mula, can be formulated already in this purely linear setting. For other results, +multiplication in a commutative algebra is considered. In some statements, the +2 + +Leibniz rule or weakened versions of it are needed. However, she does not con- +sider shuffle relations for nested integrals or additional evaluation terms in the +Taylor formula. Her treatment of operators is limited to properties and iden- +tities of given operators and she does not consider rings generated by them or +normal forms. +Integro-differential algebras and operators over a field of constants were al- +ready introduced in [29, 30], see [31] for a detailed overview and further refer- +ences. More general differential algebras with integration over rings were intro- +duced in [11], see [12] for a unified presentation and comparison. In contrast, the +integro-differential rings defined in [14, 13] require the integration to be linear +over all constants, but the construction of corresponding operators introduced +there allows noncommutative coefficients and constants. Further references to +the literature can be found in the respective sections of the present paper. +So far, all algebraic treatments of integro-differential operators in the litera- +ture restrict to multiplicative evaluations, i.e. evaluation of a product is the prod- +uct of the individual evaluations. Assuming only the Leibniz rule and the iden- +tities of the fundamental theorem of calculus, we deal with non-multiplicative +evaluations quite naturally in this paper. In Section 2, we introduce (general- +ized) integro-differential rings following this principle. Analyzing the relations +imposed, we show for example that any linear projection onto constants may be +used as the evaluation of such an integro-differential ring. We also present many +examples, both with multiplicative and with non-multiplicative evaluation. Re- +mark 2.12 discusses the differences among our definition and the definitions in +the literature. +In Section 3, we investigate identities satisfied by integrals and their prod- +ucts. This reveals a generalization of the Rota-Baxter identity for integration +that contains an additional evaluation term. Focusing on nested integrals, we +discover generalized shuffle relations with elaborate additional terms involving +nested integrals of lower depth. We also characterize properties of repeated in- +tegrals of 1, which form the smallest integro-differential ring with given ring of +constants. +Linear integro-differential operators with coefficients in arbitrary (gener- +alized) integro-differential rings are constructed algebraically in Section 4 by +generators and relations. Working at the operator level enables statements of +broader applicability, since operators not only act on integro-differential rings +but also on more general modules. We compute a complete set of rewrite rules +to simplify such operators to normal form. A precise analysis of the uniqueness +of these normal forms is presented in the appendix only, since it requires a re- +fined construction relying on tensor rings (like in [14, 13] for the multiplicative +case). After a largely self-contained introduction to tensor reduction systems in +the appendix, this is carried out in Section A.4 allowing also other functionals +besides evaluation. In Section 4.1, we collect properties of integro-differential +operators with coefficients from an integral domain (e.g. analytic functions). In +particular, we also characterize the action of such operators when evaluation is +multiplicative. +In the remaining sections, we illustrate how computations in the ring of +3 + +integro-differential operators can be used to prove and generalize well-known +results from analysis. For example, variation of constants remains valid for ar- +bitrary integro-differential rings, which is the focus of Section 5. In particular, +we also detail how integro-differential operators with noncommutative coeffi- +cients can be used for proving statements about systems of arbitrary or even +undetermined size. In Section 6, we discuss how results from analysis need to +be modified for allowing the induced evaluation to be non-multiplicative. First, +we look at the formula for variation of constants, which for multiplicative eval- +uations automatically satisfies homogeneous initial conditions, and include an +extra term to retain this property in general. Finally, we present a version of the +Taylor formula with integral remainder term that is valid also for generalized +evaluations. +Conventions +Throughout the paper, rings are implicitly assumed to have +a unit element and to be different from the zero ring. +Unless stated other- +wise, rings are not assumed to be commutative and can be of arbitrary char- +acteristic. Nevertheless, for easier reading, we use the notions of modules and +linear maps from the commutative setting to refer to bimodules and bimodule- +homomorphisms over noncommutative rings. In addition, we use operator no- +tation for linear maps, e.g. the Leibniz rule for the derivative of products then +reads ∂fg = (∂f)g + f∂g. +2 +Integro-differential rings +To uniformly deal with differentiation of various kinds of functions, we use a +few basic abstract notions. Recall from differential algebra that a derivation on +a ring R is an additive map ∂ : R → R that satisfies the Leibniz rule +∂fg = (∂f)g + f∂g +(1) +for all f, g ∈ R. Then, (R, ∂) is called a differential ring and f ∈ R is called +a constant in this differential ring if and only if ∂f = 0. It is easy to see that +the set of constants forms a subring of R and ∂ is linear w.r.t. the ring of its +constants. +For further theory of differential rings see e.g. [15]. +In addition, +we introduce the following notions of integration and evaluation in differential +rings. +Definition 2.1. Let (R, ∂) be a differential ring and let C be its ring of con- +stants. We call a C-linear map +� +: R → R an integration on R, if +∂ +� +f = f +(2) +holds for all f ∈ R. A C-linear functional e: R → C which acts on C as the +identity is called an evaluation on R. +In other words, integrations on differential rings are right inverses of the +derivation that are linear over the constants and evaluations on differential rings +are C-linear projectors onto the ring of constants C. +4 + +Definition 2.2. Let (R, ∂) be a differential ring and let +� +: R → R be an +integration on R. We call (R, ∂, +� +) a (generalized) integro-differential ring and +we define the (induced) evaluation E on R by +Ef := f − +� +∂f. +(3) +If in addition R is a field or skew field, then we also call (R, ∂, +� +) a (generalized) +integro-differential (skew) field, respectively. +This extends the definition of integro-differential rings in [14] by dropping +the additional requirement that the induced evaluation should be multiplicative. +In the present paper, the notion of integro-differential rings always refers to +Definition 2.2. The following lemma shows that in any integro-differential ring, +the (induced) evaluation E is indeed an evaluation as defined in Definition 2.1. +Moreover, the ring R can be decomposed as direct sum of constant and non- +constant “functions”. +Lemma 2.3. Let (R, ∂, +� +) be an integro-differential ring with constants C. +Then, for all f ∈ R and c ∈ C, we have Ef ∈ C, E +� +f = 0, and Ec = c. +Moreover, +R = C ⊕ +� +R +as direct sum of C-modules. +Proof. First, we compute ∂Ef = ∂(f − � ∂f) = ∂f − ∂f = 0 and E� f = +� +f − +� +∂ +� +f = 0 for f ∈ R as well as Ec = c− +� +∂c = c for c ∈ C. For any f ∈ R, +we have f = Ef + f − Ef = Ef + +� +∂f and hence R = C + +� +R. Let f ∈ C ∩ +� +R +and g ∈ R such that f = +� +g. Then, 0 = ∂f = ∂ +� +g = g, which implies f = 0. +Hence, the sum R = C + +� +R is direct. +By the previous lemma, any integration induces an evaluation by id − +� +∂. +Conversely, any evaluation e can be used to define an integration that has e as +its induced evaluation, as the following theorem shows. It is easy to see that +two different integrations cannot have the same induced evaluation. Altogether, +on differential rings with ∂R = R, there is a one-to-one correspondence of +integrations and evaluations. +Theorem 2.4. Let (R, ∂) be a differential ring such that ∂R = R and let e be +an evaluation on R. Define +� +e : R → R by +� +ef := g − eg +for all f ∈ R, where g ∈ R is such that ∂g = f. Then (R, ∂, +� +e) is an integro- +differential ring and the induced evaluation is E = e. +Moreover, any integration +� +on R can be obtained from its induced evaluation +E via this construction: +� += +� +E. +Proof. Let C be the ring of constants of (R, ∂). First, we show that +� +e is well- +defined. +If g, ˜g ∈ R are such that ∂g = ∂˜g, then with c := ˜g − g ∈ C we +5 + +have ˜g − e˜g = g + c − eg − ec = g − eg, since ec = c by definition of e. For +showing C-linearity of +� +e, we let c ∈ C and f1, f2, g1, g2 ∈ R with ∂gi = fi. +Then, ∂(cg1 + g2) = cf1 + f2 together with C-linearity of id − e implies C- +linearity of +� +e. +Consequently, (R, ∂, +� +e) is an integro-differential ring, since +we also have ∂ +� +ef = f by construction. The induced evaluation is given by +Ef = f − +� +e∂f = f − (f − ef) = ef for f ∈ R. +Let +� +: R → R be any C-linear right inverse of ∂, E its induced evaluation, +and f ∈ R. Then, we have � +Ef = � f − E� f = � f by definition of � +E and by +Lemma 2.3. +In particular, if (R, ∂, +� +) is an integro-differential ring and e is any evaluation +on R, then the integration that induces e can be given in terms of +� +by +� +e := +� +− e +� +. +(4) +This implies that the difference of two integrations +� +1, +� +2 on the same differential +ring can be given as +� +1− +�� +2 = E2 +� +1 = −E1 +� +2 in terms of the induced evaluations +E1, E2. More generally, if (R, ∂, +� +) is an integro-differential ring with constants +C and e : R → C is only C-linear, then +� +e defined by (4) can be easily seen to be +an integration on R and its induced evaluation is given by e + (id − e)E, which +agrees with e if and only if the latter is an evaluation (i.e. e1 = 1). +By Lemma 2.3, +� +R is a direct complement of C in an integro-differential ring +R. Conversely, any direct complement of C gives rise to an evaluation on R, +which in turn induces an integration by Theorem 2.4. More precisely, we have +the following characterization of integro-differential rings. +Corollary 2.5. Let (R, ∂) be a differential ring with ring of constants C. Then, +(R, ∂) can be enriched into an integro-differential ring if and only if ∂R = R and +C is a complemented C-module in R. Moreover, if ∂R = R, there exists a one- +to-one correspondence between direct complements of C in R and integrations +on (R, ∂). +This characterization shows that on an integro-differential ring (R, ∂, +� +), +in general, there are many other integrations that make R into an integro- +differential ring with the same derivation. In contrast, the following character- +ization shows that, in general, ∂ is the only derivation that turns R into an +integro-differential ring with the same integration. +Lemma 2.6. Let R be a ring, let C be a subring of R, and let +� +: R → R be a +C-linear map. Then, there exists a derivation ∂ on R such that (R, ∂, � ) is an +integro-differential ring with constants C if and only if the following conditions +hold. +1. +� +is injective. +2. R = C ⊕ +� +R +3. ( +� +f) +� +g − +� +( +� +f)g − +� +f +� +g ∈ C for all f, g ∈ R. +6 + +Moreover, this derivation is unique if it exists. +Proof. First, it is easy to see, from injectivity of +� +and by R = C ⊕ +� +R, that +there exists a C-linear map ∂ : R → R such that ker ∂ = C and ∂ +� += id and +that this map is unique. To show that ∂ is indeed a derivation, we verify the +Leibniz rule on two arbitrary elements of R. By R = C ⊕ +� +R, we write these +two elements as c+ +� +f and d+ +� +g with c, d ∈ C and f, g ∈ R. Now, we compute +∂(c + � f)(d + � g) − (∂(c + � f))(d + � g) − (c + � f)∂(d + � g) += ∂( +� +f) +� +g − f +� +g − ( +� +f)g += ∂ +� +( +� +f) +� +g − +� +( +� +f)g − +� +f +� +g +� +, +which is zero by ker ∂ = C and the last assumption on +� +. +Conversely, if (R, ∂, +� +) is an integro-differential ring with constants C, then +injectivity of +� +follows from the definition (2) and the other two conditions on +� +follow from Lemma 2.3 and Theorem 3.1. +It is straightforward to equip the matrix ring over an integro-differential +ring with an integro-differential ring structure. Such noncommutative integro- +differential rings are relevant when working with linear systems, see Section 5.2. +Lemma 2.7. Let (R, ∂, +� +) be an integro-differential ring with constants C and +let n ≥ 1. +Then, (Rn×n, ∂, � ) is an integro-differential ring with constants +Cn×n, where operations ∂, +� +, E act on matrices by applying the corresponding +operation entrywise in R. +Proof. Clearly, ∂, +� +, E are Cn×n-linear on Rn×n satisfying ∂ +� +A = A and +� +∂A = +A − EA for all A ∈ Rn×n. Moreover, it is straightforward to verify that ∂ is a +derivation on R with constants Cn×n. +Example 2.8. Basic examples for commutative integro-differential rings are +univariate polynomials C[x] and formal power series C[[x]] over a commutative +ring C with Q ⊆ C. The derivation is given by ∂ = +d +dx with ring of constants C +and integration is defined C-linearly by +� +xn = xn+1 +n + 1 +for all n ∈ N. +The induced evaluation extracts the constant coefficient and +corresponds to evaluation at 0. +If C ⊆ C, the integration +� +corresponds to +integration +� x +0 from 0. Also the rings of complex-valued smooth or analytic +functions on a (possibly unbounded) interval I ⊆ R together with derivation +∂ = +d +dx and integration +� += +� x +a , for fixed a ∈ I, are integro-differential rings. +Then, the induced evaluation is the evaluation of functions at the point a. +In particular, the ring of exponential polynomials on the real line generated +by polynomials and exponential functions is closed under differentiation and +integration and hence is an integro-differential ring as well. Algebraically, for any +field C of characteristic zero, we can consider the ring of exponential polynomials +7 + +C[x, eCx], where ecxedx = e(c+d)x for all c, d ∈ C and e0x = 1, together with +derivation ∂ = +d +dx and with the integration that is induced by evaluation at 0 +based on Theorem 2.4. +Example 2.9. A basic example of integro-differential rings of arbitrary charac- +teristic are Hurwitz series, which are closely related to formal power series and +have been defined in [16, 17], with derivation ∂(a0, a1, . . . ) = (a1, a2, . . . ) and +integration given by +� +(a0, a1, . . . ) = (0, a0, a1, . . . ), see also [18]. +The examples mentioned so far have the special property that the evalua- +tion of the integro-differential ring is multiplicative, as in the usual definition of +integro-differential algebras. However, for certain differential rings (in particu- +lar for differential fields, cf. Corollary 5 in [31]), it is not possible to define a +multiplicative evaluation for the following reason. +Remark 2.10. If the induced evaluation is multiplicative, one can see easily +that no element of +� +R can have a multiplicative inverse. Since otherwise we +would have Ef 1 +f = E1 = 1 and (Ef)E 1 +f = 0E 1 +f = 0 for such f ∈ � R. +Analytically, if evaluation should correspond to evaluation of functions at a +fixed point a for functions that are continuous at a, then requiring multiplica- +tivity of evaluation means that functions with poles at a cannot be considered. +In particular, if a function f has a pole of order m at a, then evaluation of the +product (x − a)mf gives a nonzero value, but the factor (x − a)m evaluates to +zero at a. +In the following theorem, based on results from the literature, we briefly +characterize when the induced evaluation is multiplicative. From (3) it imme- +diately follows that the identity (7) is equivalent to multiplicativity of E. Since +in integro-differential rings +� +is C-linear by definition, we have the following +characterization of integro-differential rings with multiplicative evaluation. +Theorem 2.11. Let (R, ∂, +� +) be an integro-differential ring. Then the following +properties are equivalent. +1. E is multiplicative, i.e. for all f, g ∈ R we have +Efg = (Ef)Eg. +(5) +2. +� +satisfies the Rota-Baxter identity, i.e. for all f, g ∈ R we have +(� f)� g = � (� f)g + � f� g. +(6) +3. The hybrid Rota-Baxter identity holds, i.e. for all f, g ∈ R we have +( +� +∂f) +� +∂g = ( +� +∂f)g + f +� +∂g − +� +∂fg. +(7) +Proof. Since (R, ∂) is a differential Z-algebra of weight 0 and +� +is Z-linear, this +immediately follows from items (b), (g), and (a) of Theorem 2.5 in [12]. +8 + +Remark 2.12. In the literature, integro-differential K-algebras (of weight 0) +over a commutative ring K with unit element are defined as differential K- +algebras where the additional map +� +is only required to be K-linear, but has +to satisfy the hybrid Rota-Baxter axiom (7) in addition to (2), see [12]. Analo- +gously, the more general notion of differential Rota-Baxter K-algebras (of weight +0) imposes the Rota-Baxter identity (6) instead of (7) in addition to (2). On a +differential Rota-Baxter K-algebra with constants C, a K-linear map E is defined +by (3) as well and, by Theorem 2.5 in [12], properties (5), (7), and C-linearity +of +� +are equivalent, see also Proposition 10 in [31]. +By the previous theorem, any (generalized) integro-differential ring with +constants C is an integro-differential K-algebra for K = Z and for K = C ∩Z(R) +(i.e. K = C, if R is commutative) if any of the equivalent conditions holds. +Conversely, any differential Rota-Baxter K-algebra (of weight 0) is an integro- +differential ring if and only if +� +is linear over the constants C. In particular, +this is automatically the case for integro-differential K-algebras (of weight 0) by +Proposition 10 in [31]. For concrete differential Rota-Baxter algebras (of weight +0) where +� +is not C-linear, see Example 3 in [30] and the algebraic analog of +piecewise functions constructed in [32]. +As a basic example for integro-differential rings with non-multiplicative eval- +uation, we extend the polynomial ring C[x] over a commutative ring C with +Q ⊆ C by adjoining the multiplicative inverse x−1. In order to have a surjective +derivation ∂ = +d +dx, we also need to adjoin the logarithm ln(x) as in the following +example. +Example 2.13. On C[x, x−1, ln(x)], with Q ⊆ C and ∂ = +d +dx, we can define the +C-linear integration recursively as follows. +� xk ln(x)n := + + + + + + + +xk+1 +k+1 +k ̸= −1 ∧ n = 0 +xk+1 +k+1 ln(x)n − +n +k+1 +� +xk ln(x)n−1 +k ̸= −1 ∧ n > 0 +ln(x)n+1 +n+1 +k = −1 +The same recursive definition also works on the larger ring C((x))[ln(x)] of formal +Laurent series with logarithms, where every element can be written in the form +�∞ +k=−m +�m +n=0 ck,nxk ln(x)n for some m ∈ N and ck,n ∈ C. In both cases, the +induced evaluation acts by +E +∞ +� +k=−m +m +� +n=0 +ck,nxk ln(x)n = c0,0 +and is not multiplicative as expected by Remark 2.10. Moreover, for the integro- +differential subrings of polynomials or formal power series, this evaluation cor- +responds to the usual multiplicative evaluation at 0. +Example 2.14. Rational functions together with nested integrals of rational +functions also form an integro-differential ring with non-multiplicative evalua- +tion. Algebraically, if C is a field of characteristic zero, this can be understood +9 + +as an integro-differential subring of C((x))[ln(x)] with integration +� +as in the +previous example. In fact, this ring is the smallest integro-differential ring con- +taining C(x) and is generated as a C(x)-vector space by 1 and all nested inte- +grals +� +f1 +� +f2 +� +. . . +� +fn of arbitrary depth n ≥ 1, where fi ∈ C(x) are proper +and have irreducible denominators. In particular, if C = C, the integrands can +be chosen as fi = +1 +x−ai with ai ∈ C. These kind of nested integrals are called +hyperlogarithms [21] and have been investigated already in [20]. In [12], the +free integro-differential algebra (having multiplicative evaluation) generated by +C(x) has been constructed at the somewhat unnatural expense that +� +1 ̸∈ C(x) +and the constants of the resulting differential ring contain much more than just +C. +Example 2.15. Another example of an integro-differential ring that contains +the rational functions are the D-finite functions [35]. They are characterized +as solutions of linear differential equations with rational function coefficients +and indeed form a differential ring with surjective derivation +d +dx. +Since the +constants of this differential ring are given by C, there exists an integration by +Corollary 2.5. +Example 2.16. All the examples considered above are just rings, not fields. +In contrast, transseries R[[[x]]] are an explicit construction of a differential +field (with field of constants R) that is closed under taking antiderivatives, +see [36, 8, 1] and references therein. +Thus, R[[[x]]] can be turned into an +integro-differential field by Corollary 2.5 whose evaluation necessarily is non- +multiplicative by Remark 2.10. +As shown by Corollary 2.5, in general, there are many different choices for +an integration +� +in order to turn a differential ring with ∂R = R into an +integro-differential ring. On the same differential ring, for some integrations +the induced evaluation is multiplicative and for others it is not. It may even +be the case that a canonical choice of +� +yields a non-multiplicative evaluation +while there are other choices that would give a multiplicative evaluation. In +particular, this is the case for exponential polynomials, for example. They were +mentioned above with a multiplicative evaluation, while the following canonical +definition of +� +gives rise to a non-multiplicative one. +Example 2.17. The ring of exponential polynomials C[x, eCx] over a field C of +characteristic zero is C-linearly generated by terms of the form xkecx with k ∈ N +and c ∈ C. Apart from the evaluation-based integration on C[x, eCx] mentioned +above, it is quite natural to define a C-linear integration +� +recursively as follows. +� +xkecx := + + + + + + + +xk+1 +k+1 +c = 0 +1 +cecx +k = 0 ∧ c ̸= 0 +1 +cxkecx − k +c +� +xk−1ecx +k > 0 ∧ c ̸= 0 +In terms of the Pochhammer symbol (a)k := a·(a + 1)· . . . ·(a + k − 1), +� +can be +10 + +given explicitly as +� +xkecx = +k +� +i=0 +(−k)k−ici−k−1xiecx. +Then, the induced evaluation Ef := f − +� +∂f acts by +Exkecx = +� +1 +k = c = 0 +0 +otherwise +and is not multiplicative since we have Eecxe−cx = E1 = 1 but Ee±cx = 0 for any +c ̸= 0, for example. On the other hand, C[x, ex] is an integro-differential subring +with multiplicative evaluation, for example. With this integration, for instance, +the subset C[x, ex]ex is closed under addition, multiplication, derivation, and +integration and, hence, could be viewed as an integro-differential subring with- +out unit element and having multiplicative evaluation and the zero ring as its +constants. +All the examples with explicit integration discussed so far contain an integro- +differential subring on which the induced evaluation is multiplicative. In general, +however, this need not be the case as the following example shows. +Example 2.18. On C[x] with the usual derivation and Q ⊆ C, for example, +we can define a C-linear integration by +� +xn = xn+1 +n+1 + c for all n ∈ N for any +fixed c ∈ Z(C). Such an integration induces the evaluation Ef = f(0) − cf ′(1) +on C[x], which is not multiplicative if c ̸= 0 (e.g. f = x and g = x2 − 2x yield +Eg = 0 and Efg = c). +3 +Products of nested integrals +In integro-differential rings with multiplicative evaluation the standard Rota- +Baxter identity (6) allows to write the product of integrals as a sum of two +nested integrals. For nested integrals, this leads to shuffle identities [27] where +a product of two nested integrals is expressed as a sum of nested integrals. +More generally, for Rota-Baxter operators with weight and corresponding shuffle +products involving additional terms, see [10] and references therein. In general +integro-differential rings, i.e. if E is not multiplicative, additional terms involving +the evaluation arise in the identities (6) and (7) and also in the shuffle identi- +ties. Note that all evaluation terms are evaluations of products of integrals. +Therefore, they vanish if E is multiplicative, since E� f = 0 for all f ∈ R. +Theorem 3.1. Let (R, ∂, +� +) be an integro-differential ring. Then the Rota- +Baxter identity with evaluation +( +� +f) +� +g = +� +f +� +g + +� +( +� +f)g + E( +� +f) +� +g +(8) +holds for all f, g ∈ R as well as +( +� +∂f) +� +∂g = ( +� +∂f)g + f +� +∂g − +� +∂fg − E( +� +∂f) +� +∂g. +(9) +11 + +Proof. Using (3), we can effect the decomposition of R shown in Lemma 2.3. +For ( +� +f) +� +g, we thereby obtain the decomposition +� +∂( +� +f) +� +g + E( +� +f) +� +g, which +implies (8) by the Leibniz rule for the derivation ∂. By +� +∂f = f − Ef, +� +∂g = +g − Eg, and +� +∂fg = fg − Efg, one can write (9) in a form that can be easily +verified using the fact that E is an evaluation. +As a first application, we show that in every integro-differential ring the re- +peated integrals of 1 give rise to an integro-differential subring. It is the smallest +integro-differential ring with the same constants. In characteristic zero, this ring +consists of the univariate polynomials with coefficients in the constants and, for +nonzero characteristic, it consists of a finite version of Hurwitz series [17]. +Theorem 3.2. Let (R, ∂, +� +) be an integro-differential ring with constants C. +For all n ≥ 1, let xn := +� n1 ∈ R and let x0 := 1. Then, 1, x1, x2, . . . commute +with all elements of C and are C-linearly independent. The C-module +P := spanC{1, x1, x2, . . .} +is an integro-differential subring of R. If Q ⊆ R, then P = C[x1]. +Moreover, E is multiplicative on P if and only if Exmxn = 0 for m, n ≥ 1. +Assuming E is multiplicative on P, then we have xmxn = +�m+n +m +� +xm+n and, if +in addition Q ⊆ R, xn = 1 +n!xn +1 for m, n ∈ N. +Proof. Since +� +is C-linear, every element of C commutes with xi for every i ∈ N, +even if R or C is noncommutative. +So, every element of P is of the form +�n +i=0 cixi, for some ci ∈ C. To show C-linear independence of x0, x1, . . . , let n ∈ +N be minimal such that there are c0, . . . , cn ∈ C with cn ̸= 0 and �n +i=0 cixi = 0. +Then, �n−1 +i=0 ci+1xi = ∂ �n +i=0 cixi = 0 would imply n = 0 by minimality of +n. Because this would yield c0 = 0, we conclude that x0, x1, . . . are C-linearly +independent. +Obviously, P is closed under ∂ and +� +since both operations are C-linear with +∂xn ∈ P and +� +xn = xn+1 for all n ∈ N. For showing that P is closed under +multiplication, it suffices to show that xmxn ∈ P for all m, n ≥ 1. We proceed +by induction on the sum n + m. For m = n = 1 we have x2 +1 = 2x2 + Ex2 +1 by (8). +For m + n > 2 we have +xmxn = +� +xm−1xn + +� +xmxn−1 + Exmxn +by (8). By the induction hypothesis, xm−1xn and xmxn−1 are in P. Hence, the +same is also true after applying +� +, which completes the induction. Altogether, +P is an integro-differential subring of R. +For showing P = C[x1], it is sufficient to prove that every xn is contained +in C[x1]. By (8), we obtain xn+1 = x1xn − +� +xn−1x1 − Ex1xn for all n ≥ 1. +Therefore, xn ∈ C[x1] follows by induction, if C[x1] is closed under +� +. Assuming +Q ⊆ R, we verify that +� +xn +1 − +1 +n+1xn+1 +1 +∈ C for all n ≥ 1 by applying ∂ to it, +which shows +� +xn +1 ∈ C[x1] for all n ≥ 1. +Moreover, any xn with n ≥ 1 satisfies Exn = 0. So, if E is multiplicative on +P, then trivially Exmxn = (Exm)Exn = 0 for m, n ≥ 1. Conversely, since P +12 + +is generated by 1, x1, x2, . . . as a C-module, we know that +� +P is generated by +x1, x2, . . . as a C-module. Hence, applying E to the product of two elements of +� +P gives 0, if Exmxn = 0 for all m, n ≥ 1. Altogether, using the decomposition +P = C ⊕ +� +P given by Lemma 2.3, we conclude that E is multiplicative on P, if +Exmxn = 0 for all m, n ≥ 1. +Now, we assume E is multiplicative on P and let m, n ∈ N. If m + n ≤ 1, +then xmxn = +�m+n +m +� +xm+n and xn = +1 +n!xn +1 hold trivially. For m + n ≥ 2, it fol- +lows inductively by (8) that xmxn = � �m+n−1 +m−1 +� +xm+n−1 + � �m+n−1 +m +� +xm+n−1 = +�m+n +m +� +xm+n. In particular, for n ≥ 2, x1xn−1 = nxn implies xn = +1 +n!xn +1 induc- +tively, if Q ⊆ R. +Even if E is not multiplicative on P, we can analyze some properties of the +sequence of constants cm,n := Exmxn with n, m ≥ 1. By C-linearity of � and +E, it trivially follows that all cm,n are in Z(C). It can be shown that all xn +commute with each other w.r.t. multiplication if and only if the sequence cm,n +is symmetric. +If Q ⊆ R, it can be shown by lengthy computation that the +constants cm,n are determined by all c1,n via the recursion +cm,n = 1 +m +��m+n−1 +m−1 +� +c1,m+n−1 + +m−2 +� +j=0 +n−1 +� +k=1 +�j+k +j +� +c1,j+kcm−j−1,n−k +� +. +To express xn in terms of powers of x1 in general, for Q ⊆ R, we obtain the +recursion xn = 1 +n!xn +1 − �n +i=2 +1 +i!xn−iExi +1 from Theorem 6.2.2 in [23]. +3.1 +Generalized shuffle relations +In this section, we let (R, ∂, +� +) be a commutative integro-differential ring. Iter- +ating the standard Rota-Baxter identity (6) leads to shuffle relations for nested +integrals expressing a product of two nested integrals of depth m and n as a +sum of nested integrals of depth exactly m + n. Also by recursively applying +the Rota-Baxter identity with evaluation (8), products of nested integrals can +be rewritten in R as sums of nested integrals where also terms of lower depth +may occur. For convenient notation of the formulae involved, it is standard to +work in tensor products of R and to use the shuffle product, which we recall in +the following (see [10] for example). +We consider the C-module C⟨R⟩ := �∞ +n=0 R⊗n, where the tensor prod- +uct is taken over C and the empty tensor is denoted by ε. Let pure tensors +a1⊗ . . . ⊗an ∈ C⟨R⟩ represent nested integrals +� +a1 +� +a2 . . . +� +an ∈ R. More for- +mally, by C-linearity of +� +, we consider the unique C-module homomorphism +ϕ : C⟨R⟩ → R such that +ϕ(a1⊗ . . . ⊗an) = +� +a1 +� +a2 . . . +� +an ∈ R +and ϕ(ε) = 1 ∈ R. For pure tensors a ∈ C⟨R⟩, we denote shortened versions +of them by aj +i := ai⊗ai+1⊗ . . . ⊗aj, where aj +i := ε ∈ R⊗0 if i = j + 1. The +13 + +shuffle product on C⟨R⟩ can be recursively defined as follows. For pure tensors +a, b ∈ C⟨R⟩ of length m and n, respectively, we set +a +� b := +� +a ⊗ b +if m = 0 ∨ n = 0 +a1 ⊗ (am +2 +� b) + b1 ⊗ (a +� bn +2) +otherwise +in R⊗(m+n). Extending this definition to C⟨R⟩ by C-linearity, the shuffle product +turns C⟨R⟩ into a commutative C-algebra. +Using the shuffle product for pure tensors, the product of nested integrals in +R can now be represented as sum of nested integrals as follows. The constant +coefficients of nested integrals of lower depth are evaluations of products of +integrals. Consequently, if E is multiplicative, then we recover the standard +shuffle relations [27] with all these constant coefficients equal zero. +Theorem 3.3. Let (R, ∂, +� +) be a commutative integro-differential ring with +constants C. Let f, g ∈ C⟨R⟩ be pure tensors of length m and n, respectively. +Then, the product of the nested integrals ϕ(f) = +� +f1 +� +f2 . . . +� +fm and ϕ(g) = +� +g1 +� +g2 . . . +� +gn is given by +ϕ(f)ϕ(g) = ϕ(f +� g) + +m−1 +� +i=0 +n−1 +� +j=0 +e(f m +i+1, gn +j+1)ϕ(f i +1 +� gj +1) ∈ R +(10) +with constants e(f m +i+1, gn +j+1) := Eϕ(f m +i+1)ϕ(gn +j+1) ∈ C. +Proof. Without loss of generality, assume m ≤ n. We proceed by induction +on m. If m = 0, then f = cε for some c ∈ C and the equation (10) reads +cϕ(g) = ϕ(cε ⊗ g), which is trivially true since cε ⊗ g = cg and ϕ is C-linear. +For m ≥ 1, we proceed by induction on n. By virtue of (8), for n ≥ m, we have +ϕ(f)ϕ(g) = � f1ϕ(f m +2 )ϕ(g) + � ϕ(f)g1ϕ(gn +2 ) + e(f, g). The product ϕ(f m +2 )ϕ(g) +is covered by the induction hypothesis on m so that we obtain +� +f1ϕ(f m +2 )ϕ(g) = +� +f1ϕ(f m +2 +� g) + +m−1 +� +i=1 +n−1 +� +j=0 +e(f m +i+1, gn +j+1) +� +f1ϕ(f i +2 +� gj +1) +by (10). The product ϕ(f)ϕ(gn +2 ) is covered by the induction hypothesis on n +(or on m, if n = m) so that (10) yields +� ϕ(f)g1ϕ(gn +2 ) = � g1ϕ(f +� gn +2 ) + +m−1 +� +i=0 +n−1 +� +j=1 +e(f m +i+1, gn +j+1)� g1ϕ(f i +1 +� gj +2). +By definition of ϕ and +�, we have +� +f1ϕ(f m +2 +� g) + +� +g1ϕ(f +� gn +2 ) = ϕ(f +� g) +and similarly +� +f1ϕ(f i +2 +� gj +1) + +� +g1ϕ(f i +1 +� gj +2) = ϕ(f i +1 +� gj +1) for i, j ≥ 1 as well +as +� +f1ϕ(f i +2 +� g0 +1) = ϕ(f m +1 +� g0 +1) and +� +g1ϕ(f 0 +1 +� gj +2) = ϕ(f 0 +1 +� gj +1). Altogether, +14 + +this yields +ϕ(f)ϕ(g) = ϕ(f +� g) + +m−1 +� +i=1 +n−1 +� +j=1 +e(f m +i+1, gn +j+1)ϕ(f i +1 +� gj +1) ++ +m−1 +� +i=1 +e(f m +i+1, gn +1 )ϕ(f m +1 +� g0 +1) + +n−1 +� +j=1 +e(f m +i+1, gn +j+1)ϕ(f 0 +1 +� gj +1) + e(f, g), +which proves (10). +4 +Integro-differential operators +In the following, starting from a given integro-differential ring (R, ∂, +� +), we +define the corresponding ring of operators by generators ∂, +� +, E and relations. +As additive maps on R, any f ∈ R acts as multiplication operator g �→ fg and +satisfies certain identities together with the maps ∂, +� +, E. Those identities of +additive maps that correspond to the defining properties of the operations on +R will be used as defining relations for the abstract ring of operators below. +In particular, the Leibniz rule ∂fg = f∂g + (∂f)g of the derivation ∂ on R +implies the identity ∂ ◦ f = f ◦ ∂ + ∂f of additive maps for every multiplication +operator f ∈ R. This motivates the identity (11) in the definition below. Sim- +ilarly, the identities (2) and (3) in R give rise to the identities (12) and (13). +Moreover, from C-linearity of the operations ∂, +� +, E we obtain +� +cg = c +� +g for all +c ∈ C and g ∈ R, for example. In addition, we also obtain +� +fEg = ( +� +f)Eg for +all f, g ∈ R, since Eg ∈ C. Hence, we also impose commutativity of ∂, +� +, E with +elements of C and the identities (14)–(16) in the following definition. +Definition 4.1. Let (R, ∂, +� +) be an integro-differential ring and let C be its ring +of constants. We let +R⟨∂, +� +, E⟩ +be the (noncommutative unital) ring extension of R generated by indeterminates +∂, +� +, E, where ∂, +� +, E commute with all elements of C and the following identities +hold for all f ∈ R. +∂ · f = f · ∂ + ∂f +(11) +∂ · +� += 1 +(12) +� +· ∂ = 1 − E +(13) +∂ · f · E = ∂f · E +(14) +� +· f · E = +� +f · E +(15) +E · f · E = Ef · E +(16) +We call R⟨∂, +� +, E⟩ the ring of (generalized) integro-differential operators (IDO). +Note that, for multiplication in R⟨∂, +� +, E⟩, we always explicitly write · when +one of ∂, +� +, E is involved. This is necessary in order to distinguish the product +15 + +∂ · f of operators from the multiplication operator ∂f, for example. By con- +struction, the ring R⟨∂, +� +, E⟩ has a natural action on R, where the elements of +R act as multiplication operators and ∂, +� +, E act as the corresponding opera- +tions. With this action, R becomes a left R⟨∂, +� +, E⟩-module and multiplication +in R⟨∂, +� +, E⟩ corresponds to composition of additive maps on R. +Since elements of C always commute with ∂, +� +, E in R⟨∂, +� +, E⟩, it follows +that R⟨∂, +� +, E⟩ is a C-algebra whenever R is commutative. For computing in +the ring R⟨∂, � , E⟩ of IDO, we use the identities (11)–(16) as rewrite rules in +the following way. If the left hand side of one of these identities appears in an +expression of an operator, we replace it by the right hand side to obtain a new +expression for the same operator. +If rewrite rules can be applied to a given expression in different ways, then +it may happen that useful consequences of the defining relations (11)–(16) are +discovered. A simple instance starts with the expression +� +· ∂ · +� +, to which we +can apply either (12) or (13) to obtain the expressions +� +and +� +− E · +� +for the +same operator. Hence, by taking their difference, we see that the identity +E · +� += 0 +(17) +holds in R⟨∂, +� +, E⟩. Moreover, for every f ∈ R, the expression +� +· ∂ · f can be +rewritten by (11) and by (13). Thereby we obtain the expressions +� +·f ·∂+ +� +·∂f +and f − E · f for the same operator, which implies the identity +� +· f · ∂ = f − E · f − +� +· ∂f. +(18) +By letting both sides of this identity in R⟨∂, +� +, E⟩ act on any g ∈ R, we show +that integration by parts +� +f∂g = fg − Efg − +� +(∂f)g +holds in R. Furthermore, by considering also the newly obtained identity (18) +as rewrite rule, for every f ∈ R, we can rewrite the expression +� +·f ·∂ · +� +by (12) +and by (18). Substituting f by � f in the difference f ·� −E·f ·� −� ·∂f ·� −� ·f +of the results, we obtain the identity +� +· f · +� += +� +f · +� +− +� +· +� +f − E · +� +f · +� +. +(19) +By acting with both sides of this identity on any g ∈ R, we obtain an alternative +proof for the Rota-Baxter identity with evaluation (8). Note that, for f = 1, we +also obtain the following identities from (14)–(16) and (19). +∂ · E = 0, +� +· E = +� +1 · E, +E · E = E, +(20) +� +· +� += +� +1 · +� +− +� +· +� +1 − E · +� +1 · +� +(21) +In Table 1, we collect the identities (11)–(21) as a rewrite system for expres- +sions of operators in the ring of IDO. In fact, we drop (14) since it is redundant +in the presence of (11) and (20). +16 + +Theorem 4.2. Let (R, ∂, +� +) be an integro-differential ring. Then, by repeatedly +applying the rewrite rules of Table 1 in any order, every element of the ring +R⟨∂, +� +, E⟩ can be written as a sum of expressions of the form +f · ∂j, +f · +� +· g, +f · E · g · ∂j, +or +f · E · h · +� +· g +where j ∈ N0, f, g ∈ R, and h ∈ +� +R. +Note that the expressions specified in the above theorem are irreducible in +the sense that they cannot be rewritten any further by any rewrite rules from +Table 1. The above derivation of identities (17)–(21) is similar to Knuth-Bendix +completion [19] and Buchberger’s algorithm for computing Gröbner bases [4, 22]. +So, one can show that Table 1 represents all consequences of Definition 4.1 +in the sense that every identity in R⟨∂, +� +, E⟩ can be proven by applying the +rewrite rules in the table and by exploiting identities in R. +Moreover, the +irreducible forms of operators specified in the above theorem are unique up to +multiadditivity and commutativity. +In the appendix, we will give a precise statement (Theorem A.3) of this +by giving an explicit construction of the ring R��∂, +� +, E⟩ as a quotient of an +appropriate tensor ring. Then, the translation of Table 1 into a tensor reduction +system facilitates the proof. In fact, this proof is carried out in Theorem A.5 for +a more general class of operators including linear functionals, which are useful +for dealing with boundary problems, for instance. +Remark 4.3. In the literature, integro-differential operators were considered +only with multiplicative evaluation so far. Integro-differential operators were +first introduced in [29, 30] over a field of constants using a parametrized Gröb- +ner basis in infinitely many variables and a basis of the commutative coefficient +algebra. Integro-differential operators with polynomial coefficients over a field +of characteristic zero were also studied using generalized Weyl algebras [2], skew +polynomials [28], and noncommutative Gröbner bases [26]. A general construc- +tion of rings of linear operators over commutative operated algebras is presented +in [9]. In particular, integro-differential operators are discussed in that setting +and also differential Rota-Baxter operators are investigated. Tensor reduction +systems have already been used in [14, 13] for the construction of IDO includ- +ing functionals in the special case that the evaluation and all functionals are +multiplicative. +There, also additional operators arising from linear substitu- +tions were included. +In particular, these cover integro-differential-time-delay +operators, which were already constructed algebraically in [25], see also [5]. +∂ · f = f · ∂ + ∂f +� +· f · ∂ = f − E · f − +� +· ∂f +∂ · E = 0 +� · f · E = � f · E +∂ · +� += 1 +� +· f · +� += +� +f · +� +− +� +· +� +f − E · +� +f · +� +E · f · E = Ef · E +� +· ∂ = 1 − E +E · E = E +� +· E = +� +1 · E +E · +� += 0 +� +· +� += +� +1 · +� +− +� +· +� +1 − E · +� +1 · +� +Table 1: Rewrite rules for operator expressions +17 + +To refer to integro-differential operators of special form, we use the following +notions. +Definition 4.4. Let L ∈ R⟨∂, +� +, E⟩, then we call L +1. a differential operator, if there are f0, . . . , fn ∈ R such that +L = +n +� +i=0 +fi · ∂i, +where we call L monic, if fn = 1, +2. an integral operator, if there are f1, . . . , fn, g1, . . . , gn ∈ R such that +L = +n +� +i=1 +fi · +� +· gi, +3. an initial operator, if there are f1, . . . , fn ∈ R and differential and integral +operators L1, . . . , Ln such that +L = +n +� +i=1 +fi · E · Li. +In particular, we call an initial operator L monic, if there is a differential +operator L1 and an integral operator L2 such that +L = E · (L1 + L2). +In R⟨∂, � , E⟩, using Table 1, one can check that the differential operators +are the elements of the subring generated by R and ∂, the integral operators +are the elements of the R-bimodule generated by +� +, the initial operators are the +elements of the two-sided ideal generated by E, and the monic initial operators +are the elements of the right ideal generated by E. Theorem 4.2 says that every +integro-differential operator can be written as the sum of a differential operator, +an integral operator, and an initial operator. In fact, by the stronger results in +the appendix, this decomposition of integro-differential operators even is unique. +Remark 4.5. We outline how the construction of integro-differential opera- +tors changes when the evaluation is multiplicative, see also [31, 14, 13]. For +multiplicative evaluation, we have Efg = (Ef)Eg for all f, g ∈ R. So, for the +algebraic construction of corresponding operators as in Definition 4.1, we need +to impose in addition that +E · f = Ef · E +(22) +for all f ∈ R. This does not give rise to new consequences other than (17)–(21), +but together with (20) it makes (16) redundant. Hence, in Table 1, we can +replace (16) by (22). Moreover, (22) allows to reduce the evaluation term in +(18) and, since E +� +f = 0 for all f ∈ R, to omit the evaluation terms in (19) and +18 + +(21). Consequently, the irreducible forms of operators given in Theorem 4.2 can +be simplified to +f · ∂j, +f · +� +· g, +f · E · ∂j, +where j ∈ N0 and f, g ∈ R. Evidently, this ring of operators is isomorphic to +R⟨∂, +� +, E⟩ factored by the two-sided ideal (E · f − Ef · E | f ∈ R). +4.1 +IDO over integral domains +In many concrete situations, when computing with differential operators or dif- +ferential equations with scalar coefficients, the order of a product of differential +operators is the sum of the orders of the factors. +This is equivalent to the +ring R of coefficients being an integral domain, i.e. a commutative ring without +nontrivial zero divisors. For the rest of this section, we only consider integro- +differential rings that are integral domains. Under this assumption, we can in- +vestigate further properties of computations in the ring of IDO. For instance, an +integro-differential equation can be reduced to a differential equation by differ- +entiation, see Lemma 4.8 below. Additionally, investigating the action of IDOs +on integro-differential rings algebraically leads to analyzing two-sided ideals. +As explained earlier, elements of R⟨∂, +� +, E⟩ naturally act as additive maps +on R. Likewise, they act naturally on any integro-differential ring extension of +R. In Theorem 4.11, we also provide conditions when this action is faithful, i.e. +0 ∈ R⟨∂, +� +, E⟩ is the only element that induces the zero map. +Let (S, +� +, ∂) be the integro-differential ring defined in Example 2.13 and as- +sume C is an integral domain. Then, S⟨∂, � , E⟩ does not act faithfully on S, since +E · ln(x) acts like zero. However, with the integro-differential subring R := C[x] +of S, R⟨∂, +� +, E⟩ acts faithfully on S by Theorem 4.11 below. To see this, let +L := �n +i=0 fi ·∂i ∈ R⟨∂, +� +, E⟩ and let k ∈ N such that coeff(fn, xk) ̸= 0. Apply- +ing E·L to xn−k ln(x)n ∈ S, we obtain (E·L)xn−k ln(x)n = n! coeff(fn, xk) ̸= 0. +As a preparatory step, we need some basic statements involving multiplica- +tion with constants that are valid for integro-differential rings that are integral +domains. Linear independence over constants is tied to the Wronskian. Follow- +ing the analytic definition, the Wronskian of elements f1, . . . , fn of a commuta- +tive differential ring is defined by +W(f1, . . . , fn) := det +�� +∂i−1fj +� +i,j=1,...,n +� +. +Lemma 4.6. Let (R, ∂) be a differential ring that is an integral domain with +ring of constants C and let f1, . . . , fn ∈ R, n ≥ 1. +1. f1, . . . , fn are linearly independent over C if and only if their Wronskian +W(f1, . . . , fn) is zero. +2. If f1, . . . , fn are linearly independent over C, then g1, . . . , gn−1 are linearly +independent over C, where gi := W(fi, fn). +Proof. For showing the first statement, we note that neither linear independence +nor zeroness of the Wronskian changes, if we replace R and hence C by their +19 + +quotient fields. For differential fields, a proof can be found in [15], which implies +the statement given here. +For showing the second statement, we assume that f1, . . . , fn are linearly +independent over C and we let c1, . . . , cn−1 ∈ C such that �n−1 +i=1 cigi = 0. +By definition of gi and multilinearity of the Wronskian over C, we conclude +W(�n−1 +i=1 cifi, fn) = 0. By assumption on f1, . . . , fn, this implies �n−1 +i=1 cifi = 0 +and hence c1 = . . . = cn−1 = 0. +Lemma 4.7. Let (R, ∂, +� +) be an integro-differential ring that is an integral +domain and let C be its ring of constants. Then, elements of C commute with +all elements of R⟨∂, +� +, E⟩. Moreover, for any L ∈ R⟨∂, +� +, E⟩ and any nonzero +c ∈ C, we have that L = 0 if and only if c · L = 0. +Proof. By construction, constants commute with ∂, +� +, E ∈ R⟨∂, +� +, E⟩. Since +R is commutative, it follows that elements of C commute with all elements +of R⟨∂, +� +, E⟩. Therefore, R⟨∂, +� +, E⟩ is a unital C-algebra and hence Q(C) ⊗C +R⟨∂, +� +, E⟩ is a unital C-algebra as well, with multiplication (c1⊗L1)·(c2⊗L2) = +(c1c2) ⊗ (L1 · L2). Now, 1 ⊗ L = c−1 ⊗ (c · L) implies L = 0 if c · L = 0. +Now, we are ready to have a closer look at certain computations with IDO. +The following two lemmas construct left or right multiples of IDO by differential +operators such that the product does not involve the integration operator any- +more. The tricky part will be to ensure that the product is a nonzero operator +again. +Lemma 4.8. Let (R, ∂, +� +) be an integro-differential ring that is an integral +domain. +Let C be the ring of constants of R and let L ∈ R⟨∂, � , E⟩ be not +an initial operator. +Then, there exist nonzero h1, . . . , hn ∈ R such that the +following product is a nonzero differential operator. +(h1 · ∂ − ∂h1) · . . . · (hn · ∂ − ∂hn) · L +Proof. There are n ∈ N, a nonzero c ∈ C, differential operators L0, . . . , Ln ∈ +R⟨∂, +� +, E⟩, and f1, . . . , fn, g1, . . . , gn ∈ R such that f1, . . . , fn are linearly inde- +pendent over C and c · L = L0 + �n +i=1 fi · +�� +· gi + E · Li +� +. Since L is not an +initial operator, L0 and g1, . . . , gn are not all zero by Lemma 4.7. If L0 = 0, +then we assume without loss of generality that g1 ̸= 0. +To remove the sum �n +i=1 fi· +�� +· gi + E · Li +� +, we will multiply c·L iteratively +by n first-order differential operators from the left as follows. If n > 0, then by +(11) and (12) we have (fn ·∂ − ∂fn)·fi · +� +·gi = fnfigi + (fn∂fi − (∂fn)fi)· +� +·gi +for all i. Therefore, using (14), we obtain +(fn · ∂ − ∂fn) · c · L = (fn · ∂ − ∂fn) · L0 + +n +� +i=1 +fnfigi ++ +n−1 +� +i=1 +(fn∂fi − (∂fn)fi) · +�� +· gi + E · Li +� +, +20 + +which has the form ˜L0+�n−1 +i=1 ˜fi · +�� +· gi + E · Li +� +similar to c·L. Note that also +the following two properties of the above representation of c · L are preserved. +First, if L0 is nonzero, the differential operator ˜L0 := (fn · ∂ − ∂fn) · L0 + +�n +i=1 fnfigi is nonzero too, since fn ̸= 0. Second, ˜fi := fn∂fi − (∂fn)fi, for +i ∈ {1, . . . , n − 1}, are again linearly independent over C by Lemma 4.6. Now, +we let hn := fnc such that (hn · ∂ − ∂hn) · L = (fn · ∂ − ∂fn) · c · L. +If n = 1, we only need to show that (hn · ∂ − ∂hn) · L = ˜L0 is nonzero. As +remarked above, if L0 ̸= 0 then ˜L0 ̸= 0. On the other hand, if L0 = 0, then +˜L0 = f 2 +1 g1, which is nonzero as well due to the assumptions. +If n > 1, we continue by repeating what we did with c · L above. That is, +noting ˜f1, . . . , ˜fn−1 are linearly independent over C, we multiply (hn · ∂ − ∂hn) · +L = ˜L0 + �n−1 +i=1 ˜fi · +�� +· gi + E · Li +� +from the left by hn−1 · ∂ − ∂hn−1, where +hn−1 := ˜fn−1. We iterate this a total of n − 1 times, the result is a differential +operator. To see that it is also nonzero, we focus on the last step, i.e. we refer +to the case n = 1 above. +An immediate consequence of the previous lemma is that a left ideal in +R⟨∂, +� +, E⟩ is nontrivial if and only if it contains a nonzero differential operator +or a nonzero initial operator. Left ideals in R⟨∂, +� +, E⟩ arise, for instance, as sets +of operators annihilating a given subset of a left R⟨∂, � , E⟩-module. As another +consequence, for two-sided ideals, we obtain Lemma 4.10 below. +Lemma 4.9. Let (R, ∂, +� +) be an integro-differential ring that is an integral +domain. Let C be the ring of constants of R and let L ∈ R⟨∂, +� +, E⟩ be a nonzero +monic initial operator. Then, there exist nonzero h1, . . . , hn ∈ R and a nonzero +differential operator ˜L ∈ R⟨∂, +� +, E⟩ such that +L · (hn · ∂ + 2∂hn) · . . . · (h1 · ∂ + 2∂h1) = E · ˜L. +Proof. There are n ∈ N, a nonzero c ∈ C, a differential operator L0 ∈ R⟨∂, +� +, E⟩, +nonzero f1, . . . , fn ∈ � R, and nonzero g1, . . . , gn ∈ R such that g1, . . . , gn are +linearly independent over C and L·c = E·L0+�n +i=1 E·fi· +� +·gi. By Lemma 4.7, +L · c = c · L is nonzero. +To remove the sum �n +i=1 E · fi · +� +· gi, we will multiply L · c iteratively by +n first-order differential operators from the right as follows. If n > 0, then we +have +E · fi · +� +· gign · ∂ = E · figign − Efi · E · gign − E · fi · +� +· ((∂gi)gn + gi∂gn) +for all i by (18) and by (16). Therefore, by Efi = 0, we obtain +L · c · (gn · ∂ + 2∂gn) = E · L0 · (gn · ∂ + 2∂gn) + +n +� +i=1 +E · fi · +� +· (gign · ∂ + 2gi∂gn) += E · L1 + +n−1 +� +i=1 +E · fi · +� +· (gi∂gn − (∂gi)gn), +21 + +with L1 := L0 · (gn · ∂ + 2∂gn) + �n +i=1 figign. +Note that gi∂gn − (∂gi)gn, +for i ∈ {1, . . . , n − 1}, are again linearly independent over C by Lemma 4.6. +Consequently, with hn := cgn being nonzero, the operator L · (hn · ∂ + 2∂hn) is +expressed like L · c above, just with L0 replaced by L1, with different gi, and +without the last summand. +By iterating this process of multiplying by a differential operator that is +chosen as above, we obtain a right multiple of L that is of the form E · Ln with +a differential operator Ln. It remains to show that Ln is nonzero. After the +last step, i.e. passing from E · Ln−1 + E · f1 · +� +· ˜g1, with differential operator +Ln−1 and nonzero ˜g1 ∈ R, to E · Ln, we have Ln = Ln−1 · (˜g1 · ∂ + 2∂˜g1) + +f1˜g2 +1. If the differential operator Ln−1 is nonzero, then also Ln is nonzero, since +˜g1 ̸= 0. Otherwise, we have Ln = f1˜g2 +1, which is nonzero as well due to the +assumptions. +Lemma 4.10. Let (R, ∂, +� +) be an integro-differential ring that is an integral +domain and let I ⊆ R⟨∂, +� +, E⟩ be a two-sided ideal. If I contains an operator +that is not an initial operator, then I contains an operator of the form f ·E with +nonzero f ∈ R. +Proof. By Lemma 4.8, I contains a nonzero differential operator L = �n +i=0 fi·∂i. +Let k ∈ {0, . . ., n} be minimal such that fk ̸= 0. Then, fk · E ∈ I since +L · +� k · E = +n +� +i=k +fi · ∂i · +� k · E = +n +� +i=k +fi · ∂i−k · E = fk · E. +Based on the above lemmas, we can find conditions when the action of +R⟨∂, +� +, E⟩ on an integro-differential ring is faithful. In this case, the annihilator +AnnR⟨∂,� +,E⟩(S) = {L ∈ R⟨∂, +� +, E⟩ | ∀f ∈ S : Lf = 0} +is a two-sided ideal in R⟨∂, +� +, E⟩. In particular, we show that the annihilator +only contains initial operators and, if C is a field, can be generated by monic +initial operators. +Theorem 4.11. Let (S, ∂, +� +) be an integro-differential ring with its ring of +constants C being an integral domain. Let R be an integro-differential subring of +S that is an integral domain and contains C. Then, R⟨∂, +� +, E⟩ acts faithfully on +S if and only if there is no nonzero differential operator L ∈ R⟨∂, +� +, E⟩ such that +E · L vanishes on all of S. In any case, the two-sided ideal I = AnnR⟨∂,� +,E⟩(S) +only contains initial operators. Moreover, if C is a field, I has a generating set +consisting only of monic initial operators. +Proof. Since operators of the form f · E, with nonzero f ∈ R, do not vanish on +1 ∈ S, the ideal I only contains initial operators by Lemma 4.10. If there is a +nonzero differential operator L ∈ R⟨∂, +� +, E⟩ such that E · L vanishes on all of +S, then R⟨∂, +� +, E⟩ obviously does not act faithfully on S, since E · L is nonzero +as well. +22 + +Now, we assume that there is no nonzero differential operator L such that +E · L ∈ I. +Let L ∈ I, then there are nonzero c ∈ C, f1, . . . , fn ∈ R, and +nonzero monic initial operators L1, . . . , Ln ∈ R⟨∂, +� +, E⟩ such that f1, . . . , fn are +C-linearly independent and c · L = �n +i=1 fi · Li. If we would have n ̸= 0, then +we would conclude L1, . . . , Ln ∈ I, since �n +i=1 fiLig = cLg = 0 and Lig ∈ C +for all g ∈ S. Therefore, by Lemma 4.9, there would be a nonzero differential +operator ˜L ∈ R⟨∂, +� +, E⟩ such that E · ˜L ∈ I in contradiction to the assumption. +Hence, n = 0, which implies c · L = 0. By Lemma 4.7, it follows that L = 0, +which shows that R⟨∂, +� +, E⟩ acts faithfully on S. +Finally, if C is a field, we show that I has a generating set consisting only +of monic initial operators. Let L ∈ I, then L is an initial operator. Now, there +are nonzero f1, . . . , fn ∈ R, and nonzero monic initial operators L1, . . . , Ln ∈ +R⟨∂, +� +, E⟩ such that f1, . . . , fn are C-linearly independent and L = �n +i=1 fi · Li. +For all g ∈ S, we have Lig ∈ C and �n +i=1 fiLig = Lg = 0, which implies Lig = 0 +for all i. Therefore, L1, . . . , Ln ∈ I, which shows that I is generated by nonzero +monic initial operators. +Assuming additional properties of R resp. of the action of R⟨∂, � , E⟩, gen- +erating sets of annihilators can be narrowed down even more. In particular, +the following two corollaries consider the situation when R is a field or E is +multiplicative. +Corollary 4.12. Let (S, ∂, � ) be an integro-differential ring with its ring of +constants C being a field. Let R be an integro-differential subring of S that is a +field and contains C. Then, the two-sided ideal AnnR⟨∂, +� +,E⟩(S) has a generating +set consisting only of operators of the form E · �n +i=0 fi · ∂i, with f0, . . . , fn ∈ R +where f0 ∈ +� +R is nonzero and fn = 1. +Proof. By Theorem 4.11, I := AnnR⟨∂, +� +,E⟩(S) has a generating set consisting +only of monic initial operators. Now, let L ∈ I be a monic initial operator and +let A be the set of all elements in I that have the form specified in the statement +of the corollary. By Lemma 4.9, we obtain nonzero h1, . . . , hm ∈ R such that +L · M = E · ˜L with M := (hm · ∂ + 2∂hm) · . . . · (h1 · ∂ + 2∂h1) ∈ R⟨∂, +� +, E⟩ +and some nonzero differential operator ˜L = �n +i=0 ˜fi · ∂i ∈ R⟨∂, +� +, E⟩ with +˜fn ̸= 0. Then, by repeated use of (11), there are f0, . . . , fn ∈ R with fn = 1 +such that ˜L · ˜f −1 +n += �n +i=0 fi · ∂i. Let k be minimal such that fk ̸= 0, then +˜L · ˜f −1 +n +· � k = �n−k +i=0 fi+k · ∂i. Since E · ˜L · ˜f −1 +n +· � k ∈ I vanishes on 1 ∈ S, +we conclude fk ∈ +� +R and hence E · �n−k +i=0 fi+k · ∂i ∈ A. Since R is a field, we +can verify by (11) and (12) that h−1 +i +· +� +· hi ∈ R⟨∂, +� +, E⟩ is a right inverse of +hi·∂+2∂hi for every i. Hence, there exists ˜ +M ∈ R⟨∂, +� +, E⟩ such that M · ˜ +M = 1. +Then, E · +��n−k +i=0 fi+k · ∂i� +· ∂k · ˜fn · ˜ +M = E · ˜L · ˜ +M = L · M · ˜ +M = L, i.e. L is in +the ideal generated by A. Altogether, this shows that I is generated by A. +Whenever E is multiplicative on R, the action of R⟨∂, +� +, E⟩ cannot be faithful +on R since the operator E · f acts as zero map for any f ∈ R with Ef = 0. +23 + +By Lemma 2.3, Ef = 0 is equivalent to f ∈ +� +R and one can always choose +f = +� +1, for example. In Remark 4.5, we already discussed factoring the ring of +operators by the relations (22) immediately arising from multiplicativity of E. +The following corollary shows that the resulting quotient always acts faithfully. +Corollary 4.13. Let (S, ∂, +� +) be an integro-differential ring with its ring of +constants C being an integral domain. Let R be an integro-differential subring +of S that is an integral domain and contains C. If Efg = (Ef)Eg for all f ∈ R +and g ∈ S, then the two-sided ideal AnnR⟨∂, +� +,E⟩(S) is generated by the set +{E · f − Ef · E | f ∈ R}. +Proof. Let I := AnnR⟨∂,� +,E⟩(S) and let J ⊆ R⟨∂, +� +, E⟩ be the two-sided ideal +generated by the set {E · f − Ef · E | f ∈ R}. By assumption on E, we have +that J ⊆ I, so it only remains to show I ⊆ J. +By Theorem 4.11, every L ∈ I is an initial operator. Following the form +of initial operators in Remark 4.5, there are f0, . . . , fn ∈ R such that L = +�n +i=0 fi ·E·∂i modulo J. By J ⊆ I, we have that �n +i=0 fi ·E·∂i ∈ I. Therefore, +�n +i=0 fi·E·∂i vanishes on +� k1 ∈ S for all k ∈ {0, . . . , n}. This implies inductively +that f0, . . . , fn are all zero. Consequently, we have L ∈ J. +5 +Equational prover in calculus +In this section, we illustrate how results from analysis can be proven via compu- +tations in the ring of integro-differential operators. The general approach con- +sists in formulating an analytic statement as an identity of integro-differential +operators and then prove this identity algebraically in R⟨∂, +� +, E⟩, with minimal +assumptions on the integro-differential ring (R, ∂, +� +) used in the coefficients. +Note that we prove results directly by a computation with integro-differential +operators, instead of doing the whole computation with elements of R or any +other left R⟨∂, +� +, E⟩-module. +An identity in R⟨∂, +� +, E⟩ can be proven by comparing irreducible forms +of the left hand side and right hand side. Recall that such irreducible forms +are given by Theorem 4.2 and can be computed systematically by the rewrite +rules in Table 1. In this sense, the rewrite rules provide an equational prover +for integro-differential operators provided one can decide equality of irreducible +forms in R⟨∂, +� +, E⟩, which includes deciding identities in R. In practice, this is +often possible for concrete irreducible forms. +Once an identity L1 = L2 is proven in R⟨∂, +� +, E⟩, we immediately infer the +corresponding identity in R, i.e. L1f = L2f for all f ∈ R, by the canonical +action of R⟨∂, +� +, E⟩ on R. Moreover, by acting on any other left R⟨∂, +� +, E⟩- +module, we obtain the analogous identity also in those modules. Furthermore, +any concrete computation with operators acting on functions uses only finitely +many derivatives. Consequently, by inspecting every step of the computation, +an identity proven for infinitely differentiable functions can also be proven for +functions that are only sufficiently often differentiable. +24 + +5.1 +Variation of constants for scalar equations +In this section, we deal with the method of variation of constants for comput- +ing solutions of inhomogeneous ODEs in terms of integro-differential operators. +First, we recall the analytic statement, see e.g. Theorem 6.4 in Chapter 3 of [6]. +Consider the inhomogeneous linear ODE +y(n)(x) + an−1(x)y(n−1)(x) + · · · + a0(x)y(x) = f(x). +Assume that z1(x), . . . , zn(x) is a fundamental system of the homogeneous equa- +tion, i.e. z(n) +i +(x) + an−1(x)z(n−1) +i +(x) + · · · + a0(x)zi(x) = 0 and the Wronskian +w(x) := W(z1(x), . . . , zn(x)) is nonzero. Then, +z∗(x) := +n +� +i=1 +(−1)n−izi(x) +� x +x0 +W(z1(t), . . . , zi−1(t), zi+1(t), . . . , zn(t)) +w(t) +f(t) dt +(23) +is a particular solution of the inhomogeneous equation above. +Algebraically, we model scalar functions by fixing a commutative integro- +differential ring (R, ∂, +� +) and, as indicated above, we model computations with +scalar equations by computations in the corresponding ring of integro-differential +operators R⟨∂, +� +, E⟩. From the operator viewpoint, mapping the inhomogeneous +part f to a solution of the equation Ly = f amounts to constructing a right +inverse of the differential operator L. +In general, for an operator L and a +right inverse H, a particular solution of the inhomogeneous equation is given by +z∗ = Hf, since +Lz∗ = L(Hf) = (L · H)f = 1f = f. +Recall that +� +is a right inverse of ∂ by definition. +Now, we consider an +arbitrary monic first-order differential operator +L = ∂ + a, +with a ∈ R, and assume that z ∈ R is a solution of Ly = 0, i.e. ∂z + az = 0. +Then, using (11), we compute +(∂ + a) · z = ∂ · z + a · z = z · ∂ + ∂z + az = z · ∂. +If, moreover, we assume that z has a multiplicative inverse z−1 ∈ R, we obtain +(∂ + a) · (z · +� +· z−1) = z · ∂ · +� +· z−1 = z · z−1 = 1. +Hence +H = z · +� +· z−1 +is a right inverse of L in R⟨∂, +� +, E⟩. +We also outline the computation for a second order differential operator +L = ∂2 + a1 · ∂ + a0, +25 + +with a1, a0 ∈ R. If z ∈ R is a solution of Ly = 0, then +L · z = z · ∂2 + (2∂z + a1z) · ∂. +Assume that there exist two solutions z1, z2 ∈ R of Ly = 0 such that their +Wronskian +w = z1∂z2 − z2∂z1 +has a multiplicative inverse 1 +w ∈ R. Let +H = −z1 · +� +· z2 +w + z2 · +� +· z1 +w . +In the ring of operators, we can compute +∂ · zi +w = zi +w · ∂ + ∂zi +w − zi∂w +w2 . +Then, using the normal forms of L · zi and ∂ · zi +w , we obtain +L · H = −z1 · ∂ · z2 +w − 2(∂z1) z2 +w − a1z1 z2 +w + z2 · ∂ · z1 +w + 2(∂z2) z1 +w + a1z2 z1 +w += −z1 · ∂ · z2 +w + z2 · ∂ · z1 +w + 2 w +w = −z1∂ z2 +w + z2∂ z1 +w + 2 = 1. +Hence H is a right inverse of L in R⟨∂, +� +, E⟩. +More generally, we have the following formulation of the method of variation +of constants for integro-differential operators. +Recall from Section 4.1, that +the Wronskian W(f1, . . . , fn) of elements f1, . . . , fn ∈ R is defined completely +analogous to the analytic situation. +Theorem 5.1. Let (R, ∂, +� +) be a commutative integro-differential ring and let +L = ∂n + �n−1 +i=0 ai · ∂i ∈ R⟨∂, +� +, E⟩, n ≥ 1, with a0, . . . , an−1 ∈ R. Assume that +z1, . . . , zn ∈ R are such that Lzi = 0 and w := W(z1, . . . , zn) has a multiplicative +inverse +1 +w ∈ R. Then, with +H := +n +� +i=1 +(−1)n−izi · +� +· W(z1, . . . , zi−1, zi+1, . . . , zn) +w +we have that L · H = 1 in R⟨∂, +� +, E⟩. +Proof. The cases n = 1 and n = 2 have been shown above. In principle, an +analogous computation could be done for any concrete n ≥ 3 as well. To obtain +a finite proof for all n ≥ 3 at once, we will utilize a more general framework in +Section 5.3. +5.2 +Linear systems and operators with matrix coefficients +Algebraically, computing with linear systems of differential equations can be +modelled by integro-differential operators over some noncommutative integro- +differential ring, whose elements correspond to matrices. Such noncommutative +integro-differential rings can be obtained from any scalar integro-differential +26 + +ring, since one can equip the ring of n × n matrices with a derivation and an +integration defined entrywise, see Lemma 2.7. +Recall that Definition 4.1 and Theorem 4.2 hold also for operators with +coefficients in noncommutative integro-differential rings, in particular for oper- +ators with matrix coefficients. Consequently, any computation using a general +noncommutative integro-differential ring is automatically valid for concrete ma- +trices of any size, without the need for entrywise computations. This allows for +compact proofs for matrices of general size. For switching to entrywise com- +putations, one can view operators with matrix coefficients equivalently also as +matrices of operators with scalar coefficients by identifying operators ∂, +� +, E +with corresponding diagonal matrices of operators. The following lemma shows +that also computations are equivalent in both viewpoints. Here, we use the +notation Ei,j(L) := (δi,kδj,lL)k,l=1,...,n for matrices with only one nonzero entry +L ∈ R⟨∂, +� +, E⟩. +Lemma 5.2. Let (R, ∂, +� +) an integro-differential ring and let n ≥ 1. Then, +there is exactly one ring homomorphism +ϕ : Rn×n⟨∂, +� +, E⟩ → R⟨∂, +� +, E⟩n×n +with ϕ(A) = A for A ∈ Rn×n and ϕ(L) = diag(L, . . . , L) for L ∈ {∂, +� +, E}. +This ϕ is an isomorphism and its inverse homomorphism +ψ : R⟨∂, � , E⟩n×n → Rn×n⟨∂, � , E⟩ +can be given by ψ(Ei,j(f)) = Ei,j(f) and ψ(Ei,j(L)) = Ei,j(1) · L for all f ∈ R +resp. L ∈ {∂, +� +, E}. +Proof. First, we check that the definition of ϕ indeed provides a unique and +well-defined homomorphism of rings. Uniqueness follows from the fact that ϕ is +defined on a generating set of Rn×n⟨∂, +� +, E⟩. For proving well-definedness, we +need to verify that the definition of ϕ respects all identities of generators given +in Definition 4.1 as follows. Evidently, we have ϕ(1) = In. It is immediate to see +that ϕ(∂) · ϕ( +� +) = ϕ(1) − ϕ(E) and ϕ( +� +) · ϕ(∂) = ϕ(1) hold. For L ∈ {∂, +� +, E}, +we verify in R⟨∂, +� +, E⟩n×n that ϕ(L) commutes with all elements of Cn×n and +that, for all A ∈ Rn×n, we have ϕ(L) · ϕ(A) · ϕ(E) = ϕ(LA) · ϕ(E). +More +explicitly, using (14)–(16) in R⟨∂, +� +, E⟩, we compute +ϕ(L) · ϕ(A) · ϕ(E) = diag(L, . . . , L) · A · diag(E, . . . , E) += (L · ai,j · E)i,j=1,...,n = (Lai,j · E)i,j=1,...,n = ϕ(LA) · ϕ(E). +Analogously, using (11) in R⟨∂, +� +, E⟩, we can also verify that ϕ(∂) · ϕ(A) = +ϕ(A) · ϕ(∂) + ϕ(∂A) for all A ∈ Rn×n. +Similarly, we need to verify that the definition of ψ on a generating set of +R⟨∂, +� +, E⟩n×n gives rise to a well-defined homomorphism. For example, using +27 + +(14)–(16) in Rn×n⟨∂, +� +, E⟩, we compute +ψ(Ei,j(L)) · ψ(Ep,q(f)) · ψ(Ek,l(E)) = Ei,j(1) · L · Ep,q(f) · Ek,l(1) · E += Ei,j(1) · L · δq,kEp,l(f) · E = Ei,j(1) · δq,kLEp,l(f) · E = δj,pδq,kEi,l(Lf) · E += Ei,q(δj,pLf) · Ek,l(1) · E = ψ(Ei,q(δj,pLf)) · ψ(Ek,l(E)) +for all f ∈ R, L ∈ {∂, +� +, E}, and all i, j, k, l, p, q ∈ {1, . . ., n}. +Finally, ψ is the inverse of ϕ, since we have ψ(ϕ(A)) = A, ϕ(ψ(Ei,j(f))) = +Ei,j(f), ψ(ϕ(L)) = L, and ϕ(ψ(Ei,j(L))) = Ei,j(L) for the generators. +Integro-differential operators with coefficients from Rn×n constructed this +way have natural actions on Rn×n and on Rn. +By the above lemma, for +any left R⟨∂, +� +, E���-module M, there is a natural way of viewing M n as a +left Rn×n⟨∂, � , E⟩-module. +It is straightforward to show that the action of +Rn×n⟨∂, +� +, E⟩ on M n is faithful if and only if R⟨∂, +� +, E⟩ acts faithfully on M. +5.3 +Variation of constants for first-order systems +Instead of higher order scalar ODEs, we now consider variation of constants for +first-order systems, see Theorem 3.1 in Chapter 3 of [6] for example. Analyti- +cally, it can be stated as follows. For an n × n matrix A(x) and a vector f(x) +of size n, we consider the first-order system given by +y′(x) + A(x)y(x) = f(x). +If Φ(x) is a fundamental matrix of the homogeneous system, i.e. it satisfies +Φ′(x) + A(x)Φ(x) = 0 and det(Φ(x)) ̸= 0, then a particular solution of the +inhomogeneous system is given by +z∗(x) = Φ(x) +� x +x0 +Φ(t)−1f(t) dt. +Theorem 5.3. Let (R, ∂, +� +) be an integro-differential ring. Assume that a ∈ R +is such that there exists z ∈ R that satisfies ∂z +az = 0 and has a multiplicative +right inverse z−1 ∈ R. +Then, in R⟨∂, +� +, E⟩, the operators L := ∂ + a and +H := z · +� +· z−1 satisfy +L · H = 1. +Proof. The same computation as in the commutative case above can be done +also for noncommutative R without any changes. Note that zz−1 = 1 was the +only property of z−1 used there, so it suffices that z−1 is a right inverse of z. +Observe that from this statement over an abstract integro-differential ring +(R, ∂, +� +), which was proven without referring to matrices at all, the analytic +statement follows for arbitrary size n of the matrix A(x), provided we assume +sufficient regularity of the functions involved. +Based on Theorem 5.3, we now can complete the proof of Theorem 5.1 for +arbitrary n ≥ 3. Before doing so, we first detail the required translation from +28 + +a first-order system to the scalar equation entirely at the operator level. For +shorter notation, in the following, we again use the symbol ∂ also for the matrix +diag(∂, . . . , ∂) ∈ R⟨∂, +� +, E⟩n×n. +Lemma 5.4. Let (R, ∂, +� +) be an integro-differential ring and let +L = ∂n + +n−1 +� +i=0 +ai · ∂i ∈ R⟨∂, +� +, E⟩ +with a0, . . . , an−1 ∈ R. Moreover, let H ∈ R⟨∂, +� +, E⟩n×n satisfy (∂+A)·H = In +in R⟨∂, +� +, E⟩n×n, where +A := + + + + + + + + +0 +−1 +0 +· · · +0 +... +... +... +... +... +... +... +... +0 +0 +· · · +· · · +0 +−1 +a0 +a1 +· · · +· · · +an−1 + + + + + + + + +. +Then, the upper right entry H1,n of H satisfies L · H1,n = 1 in R⟨∂, +� +, E⟩ and +Hi,n = ∂i−1 · H1,n for i ∈ {1, . . . , n}. +Proof. For n = 1, the statement is trivial. So, we let n ≥ 2 in the following. In +short, starting from the identity (∂ + A)·H = In of n× n matrices of operators, +we multiply both sides from the left with a suitable S ∈ R⟨∂, +� +, E⟩n×n and +inspect the last column. In particular, for elimination in the last n − 1 columns +of ∂ + A, we use +S := + + + + + + + + + + + +1 +0 +· · · +· · · +· · · +0 +∂ +... +... +... +∂2 +... +... +... +... +... +... +... +... +... +... +∂n−2 +· · · +∂2 +∂ +1 +0 +s1 +· · · +· · · +· · · +sn−1 +1 + + + + + + + + + + + +with sj := ∂n−j+�n−j−1 +i=0 +ai+j ·∂i ∈ R⟨∂, +� +, E⟩ for j ∈ {1, . . ., n−1}. Observing +that, in R⟨∂, � , E⟩, we have −sj + sj+1 · ∂ = −aj for j ∈ {1, . . ., n − 2}, we +29 + +compute S · (∂ + A) explicitly: +S · + + + + + + + + +∂ +−1 +0 +· · · +0 +0 +... +... +... +... +... +... +... +... +0 +0 +· · · +0 +∂ +−1 +a0 +a1 +· · · +an−2 +∂ + an−1 + + + + + + + + += + + + + + + + + +∂ +−1 +0 +· · · +0 +∂2 +0 +−1 +... +... +... +... +... +... +0 +∂n−1 +0 +· · · +0 +−1 +s1 · ∂ + a0 +0 +· · · +0 +−sn−1 + ∂ + an−1 + + + + + + + + += + + + + + + + + + +∂ +−1 +0 +· · · +0 +∂2 +0 +−1 +... +... +... +... +... +... +0 +∂n−1 +... +... +−1 +L +0 +· · · +· · · +0 + + + + + + + + + +. +Altogether, we obtain that + + + + + + + + + +∂ +−1 +0 +· · · +0 +∂2 +0 +−1 +... +... +... +... +... +... +0 +∂n−1 +... +... +−1 +L +0 +· · · +· · · +0 + + + + + + + + + +· H = S · (∂ + A) · H = S, +where comparison of the entries in the last column of the left hand side and +right hand side yields ∂ · H1,n − H2,n = 0, . . . , ∂n−1 · H1,n − Hn,n = 0 and +L · H1,n = 0. +Finally, we proceed with the proof of Theorem 5.1. Recalling the assump- +tions, we fix a commutative integro-differential ring (R, ∂, +� +) and we let L = +∂n+�n−1 +i=0 ai·∂i ∈ R⟨∂, +� +, E⟩ with a0, . . . , an−1 ∈ R. We also let z1, . . . , zn ∈ R +such that Lzi = 0 and such that w := W(z1, . . . , zn) has a multiplicative inverse +1 +w ∈ R. +Proof of Theorem 5.1. Let n ≥ 3. In the (noncommutative) integro-differential +30 + +ring (Rn×n, ∂, +� +), we consider the matrices +A := + + + + + + + + +0 +−1 +0 +· · · +0 +... +... +... +... +... +... +... +... +0 +0 +· · · +· · · +0 +−1 +a0 +a1 +· · · +· · · +an−1 + + + + + + + + +and +Z := + + + + + +z1 +· · · +zn +∂z1 +· · · +∂zn +... +... +∂n−1z1 +· · · +∂n−1zn + + + + + +and note that (∂ + A)Z = 0 and that Z−1 ∈ Rn×n exists, since det(Z) = w +was assumed to be invertible in R. +Then, we apply Theorem 5.3 to obtain +(∂ + A) · (Z · +� +· Z−1) = 1 in Rn×n⟨∂, +� +, E⟩. +Instead of integro-differential +operators with matrix coefficients in Rn×n, we now consider the objects as +n × n matrices whose entries are integro-differential operators with coefficients +in R, cf. Lemma 5.2. Then, by Lemma 5.4, we obtain that L·(Z ·� ·Z−1)1,n = 1 +holds in R⟨∂, +� +, E⟩. In order to compute the entry (Z · +� +·Z−1)1,n, we can easily +determine a general form of the entries of the matrix product Z · +� +· Z−1 ∈ +R⟨∂, +� +, E⟩n×n: +Z · + + + + + + +� +0 +· · · +0 +0 +... +... +... +... +... +... +0 +0 +· · · +0 +� + + + + + + +· Z−1 = Z · + + + +� +· (Z−1)1,1 +· · · +� +· (Z−1)1,n +... +... +� +· (Z−1)n,1 +· · · +� +· (Z−1)n,n + + + += +� n +� +k=1 +(∂i−1zk) · � · (Z−1)k,j +� +i,j=1,...,n +. +Then, we compute all entries +(Z−1)i,n = (−1)n+i W(z1, . . . , zi−1, zi+1, . . . , zn) +w +in the last column of Z−1 ∈ Rn×n via Cramer’s rule (or via the cofactor matrix) +so that we recognize H ∈ R⟨∂, +� +, E⟩ defined in the statement of Theorem 5.1 +as the top right entry (Z · +� +· Z−1)1,n of the above matrix. This concludes the +proof that L · H = 1. +6 +Generalizing identities from calculus +The generalizations of well-known identities presented in this section introduce +additional terms involving the induced evaluation, which vanish if the evaluation +is multiplicative. As explained in the previous section, the proofs mostly rely +on computing irreducible forms for IDO. +31 + +6.1 +Initial value problems +Recall that the formula given in (23) provides a particular solution of the inho- +mogeneous linear ODE +y(n)(x) + an−1(x)y(n−1)(x) + · · · + a0(x)y(x) = f(x). +Moreover, this particular solution also satisfies the homogeneous initial condi- +tions y(x0) = 0, y′(x0) = 0, . . . , y(n−1)(x0) = 0, see e.g. Theorem 6.4 in Chap- +ter 3 of [6]. +The induced evaluation of an integro-differential ring allows us +to model such properties algebraically. Similarly to the general proof of Theo- +rem 5.1, we first investigate the general solution formula for homogeneous initial +value problems of first-order systems. To this end, we fix a (not necessarily com- +mutative) integro-differential ring R and we work in the corresponding ring of +IDO R⟨∂, +� +, E⟩. +As a general principle, not only in the ring R⟨∂, +� +, E⟩ but in any ring with +unit element, if some H is a right inverse of some L and B, P are such that +L · P = 0 and B · P = B, then G := (1 − P) · H satisfies +L · G = 1 +and +B · G = 0. +So, by choosing an appropriate operator P ∈ R⟨∂, +� +, E⟩ that satisfies (∂+a)·P = +0 and E · P = E, we can obtain the following version of Theorem 5.3 that also +solves the homogeneous initial condition. +Theorem 6.1. Let (R, ∂, +� +) be an integro-differential ring and let L = ∂ + a ∈ +R⟨∂, +� +, E⟩ with a ∈ R. Assume that z ∈ R is such that ∂z + az = 0 and in +addition to a multiplicative right inverse z−1 ∈ R also a right inverse (Ez)−1 ∈ C +exists. Then, in R⟨∂, � , E⟩, the operator +G := (1 − z(Ez)−1 · E) · z · +� +· z−1 +satisfies L · G = 1 and E · G = 0. +Proof. By Theorem 5.3, we have that H := z · +� +· z−1 satisfies L · H = 1. Now, +letting P := z(Ez)−1 · E, we compute the normal forms of L · P and E · P: +(∂ + a) · z(Ez)−1 · E = +� +z(Ez)−1 · ∂ + ∂z(Ez)−1� +· E + az(Ez)−1 · E += z(Ez)−1 · ∂ · E = 0, +E · z(Ez)−1 · E = Ez(Ez)−1 · E = E. +Then, from L · P = 0 and E · P = E, it follows straightforwardly that G = +(1 − P) · H has the properties claimed. +Remark 6.2. If the evaluation E is multiplicative, then the existence of (Ez)−1 ∈ +C follows form the existence of z−1 ∈ R and, with (22), we also obtain E·z · +� += +(Ez)E · +� += 0, which implies P · H = 0 and hence G = H. Therefore, the right +inverse H obtained in Theorem 5.3 satisfies the initial value condition E ·H = 0 +already. +32 + +For conversion to the scalar case, we need to compute the element in the top +right corner of G = (1 − Z(EZ)−1 · E) · Z · +��� +· Z−1 and show that it has the +desired properties. +Theorem 6.3. Let (R, ∂, +� +) be a commutative integro-differential ring and let +L = ∂n + �n−1 +i=0 ai · ∂i ∈ R⟨∂, +� +, E⟩, n ≥ 1, with a0, . . . , an−1 ∈ R. Assume +that z1, . . . , zn ∈ R are such that Lzi = 0 and w := W(z1, . . . , zn) has a mul- +tiplicative inverse +1 +w ∈ R. Moreover, assume that there are ci,j ∈ C such that +E∂k �n +i=1 ci,jzi = δj,k for all j, k ∈ {0, . . . , n − 1}. Then, with these ci,j and +G := +n +� +k=1 +(−1)n−k +� +zk − +n +� +i,j=1 +zici,j−1 · E · (∂j−1zk) +� +· +� +· W(z1,...,zk−1,zk+1,...,zn) +w +we have that L · G = 1 and E · G = 0 in R⟨∂, +� +, E⟩. +Proof. As in the proof of Theorem 5.1, we consider the matrices +A := + + + + + + + + +0 +−1 +0 +· · · +0 +... +... +... +... +... +... +... +... +0 +0 +· · · +· · · +0 +−1 +a0 +a1 +· · · +· · · +an−1 + + + + + + + + +and +Z := + + + + + +z1 +· · · +zn +∂z1 +· · · +∂zn +... +... +∂n−1z1 +· · · +∂n−1zn + + + + + +and note that (∂ + A)Z = 0 and that Z−1 ∈ Rn×n exists, since det(Z) = w was +assumed to be invertible in R. Furthermore, we note that + + + +c1,0 +· · · +c1,n−1 +... +... +cn,0 +· · · +cn,n−1 + + + +is the multiplicative (right) inverse of EZ in Cn×n. By Theorem 6.1, we conclude +that ˜G := (1−Z(EZ)−1 ·E)·Z · +� +·Z−1 ∈ Rn×n⟨∂, +� +, E⟩ satisfies (∂ +A)· ˜G = 1 +and E · ˜G = 0. Passing to R⟨∂, +� +, E⟩n×n via Lemma 5.2, the latter identity +implies E · ˜Gi,j = 0 and the former implies L · ˜G1,n = 1 and ˜Gi,n = ∂i−1 · ˜G1,n +by Lemma 5.4. Hence, we also have E · ∂i−1 · ˜G1,n = 0 for i ∈ {1, . . . , n}. +Finally, we verify that ˜G1,n = G. With H := Z · +� +· Z−1 ∈ R⟨∂, +� +, E⟩n×n, +we have ˜G1,n = H1,n − �n +i,j=1 Z1,i((EZ)−1)i,j · E · Hj,n. From the proof of +Theorem 5.1, we obtain that Hj,n = (−1)n+k(∂j−1zk)· +� +· W(z1,...,zi−1,zi+1,...,zn) +w +. +Altogether, this yields ˜G1,n = G. +6.2 +Taylor formula +Usually, Taylor’s theorem is only considered for sufficiently smooth functions. +In integro-differential rings, an analog of the Taylor formula with integral re- +mainder term +f(x) = +n +� +k=0 +f (k)(x0) +k! +xk + +� x +x0 +(x − t)n +n! +f (n+1)(t) dt +33 + +can be formulated. While the formula arising from the identity of operators in +Theorem 6.6 is more complicated, it is also valid if singularities are present. +We start by giving a first version of the Taylor formula where the remainder +term is given as repeated integral. It simply follows by iterating (3) resp. (13). +See also Corollary 2.2.1 in [23]. +Lemma 6.4. Let (R, ∂, +� +) be an integro-differential ring. In R⟨∂, +� +, E⟩ we have +for any n ∈ N that +1 = +n +� +i=0 +� i · E · ∂i + +� n+1 · ∂n+1. +Proof. For any n ∈ N, we can use (13) to rewrite the right hand side: +n +� +i=0 +� i · E · ∂i + +� n+1 · ∂n+1 = +n +� +i=0 +� i · E · ∂i + +� n · (1 − E) · ∂n += +n−1 +� +i=0 +� i · E · ∂i + +� n · ∂n. +Setting n = 0 here gives E+ +� +·∂ = 1. Hence, the claim follows by induction. +Using (15) and the related identity in (20), we can always write operators +� i · E as xi · E, where xi := � i1 as in Theorem 3.2. By (19) and (21), we can +always write +� n+1 without higher powers of +� +. To make the resulting expressions +simpler, we restrict to the case that E is multiplicative on polynomials, i.e. +Exmxn = 0 for all m, n ≥ 1. +Lemma 6.5. Let (R, ∂, +� +) be an integro-differential ring such that the induced +evaluation E satisfies Exmxn = 0 for all m, n ≥ 1. Then, in R⟨∂, +� +, E⟩, we +have for any n ∈ N that +� n+1 = +n +� +k=0 +(−1)n−kxk · +� +· xn−k − +n−1 +� +k=0 +n−k +� +j=1 +(−1)n−k−jxk · E · xj · +� +· xn−k−j. +Proof. We prove this identity by induction on n ∈ N. For n = 0, the right +hand side directly yields +� +in agreement with the left hand side. Assuming the +identity holds for some n ∈ N, we multiply both sides by +� +from the left and we +rewrite the right hand side using (19) and (15) to obtain +� n+2 = +n +� +k=0 +(−1)n−k�� xk · � − � · � xk − E · � xk · � � +· xn−k +− +n−1 +� +k=0 +n−k +� +j=1 +(−1)n−k−j� +xk · E · xj · +� +· xn−k−j. +34 + +We have xk+1xn−k = +�n+1 +k+1 +� +xn+1 by Theorem 3.2, so we can expand the right +hand side into +n +� +k=0 +(−1)n−kxk+1 · +� +· xn−k + +n +� +k=0 +(−1)n−k+1 +�n + 1 +k + 1 +�� +· xn+1 +− +n +� +k=0 +(−1)n−kE · xk+1 · � · xn−k − +n−1 +� +k=0 +n−k +� +j=1 +(−1)n−k−jxk+1 · E · xj · � · xn−k−j. +Exploiting �n +k=0(−1)n−k+1�n+1 +k+1 +� += (−1)n+1 in the second sum, we can regroup +terms to obtain the right hand side of the claimed identity for n + 1. +This +completes the induction. +Altogether, we obtain the following identity in R⟨∂, +� +, E⟩ generalizing the +usual Taylor formula with integral remainder term. This identity holds over any +integro-differential ring in which the induced evaluation is multiplicative on the +integro-differential subring generated by 1. +Theorem 6.6 (Taylor formula). Let (R, ∂, +� +) be an integro-differential ring +such that E is multiplicative on the integro-differential subring generated by 1, +i.e. Exmxn = 0 for all m, n ≥ 1. Then, for all f ∈ R and all n ∈ N we have +1 = +n +� +k=0 +xk · E · ∂k + +n +� +k=0 +(−1)n−kxk · +� +· xn−k · ∂n+1 +− +n−1 +� +k=0 +n−k +� +j=1 +(−1)n−k−jxk · E · xj · � · xn−k−j · ∂n+1 +Proof. Follows from Lemma 6.4 using +� i·E = xi·E and Lemma 6.5, as explained +above. +In particular, if in addition to the assumptions of the theorem we have +Q ⊆ R, then, with operators acting on some f, we have that +f = +n +� +k=0 +xk +1 +k! E∂kf + +n +� +k=0 +(−1)n−k +k!(n − k)!xk +1 +� +xn−k +1 +∂n+1f +− +n−1 +� +k=0 +n−k +� +j=1 +(−1)n−k−j +k!j!(n − k − j)!xk +1Exj +1 +� +xn−k−j +1 +∂n+1f +While the first and the second sum correspond to the Taylor polynomial and the +integral remainder term, the third sum corresponds to an additional polynomial +that arises from our general setting allowing non-multiplicative evaluations. It +can be viewed as an integro-differential algebraic version of the analytic formula +− +n−1 +� +k=0 +(x − x0)k +k! +�� x +x0 +(x − t)n−k − (x0 − t)n−k +(n − k)! +f (n+1)(t)dt +� +x=x0 +, +35 + +which vanishes for smooth functions f(x). Although in practice these additional +terms yield zero even for many elements f that model singular functions in con- +crete integro-differential rings, they cannot be dropped in general, as illustrated +in C[x, x−1, ln(x)] by considering n = 1 and f = ln(x), for example. With in- +tegration defined as in Example 2.13, we have x1 = x and with f = ln(x) the +Taylor polynomial Ef + xE∂f vanishes. By ∂2f = − 1 +x2 , the second sum yields +− +� +x∂2f + x +� +∂2f = ln(x) + 1 and the third sum −Ex +� +∂2f = −1 compensates +the constant term. More generally, we can characterize the integro-differential +rings where the additional polynomial does not play a role in the Taylor formula. +Corollary 6.7. Let (R, ∂, +� +) be an integro-differential ring such that E is mul- +tiplicative on the integro-differential subring generated by 1, i.e. Exmxn = 0 for +all m, n ≥ 1. Then, we have Exng = 0 for all n ≥ 1 and g ∈ R if and only if +we have +f = +n +� +k=0 +xkE∂kf + +n +� +k=0 +(−1)n−kxk +� +xn−k∂n+1f +for all n ∈ N and f ∈ R. +Proof. If Exng = 0 for all n ≥ 1 and g ∈ R, then we have in particular +Exj +� +xn−k−j∂n+1f = 0 for all j, k, n ∈ N and f ∈ R with 1 ≤ j ≤ n − k. +Hence, Theorem 6.6 implies the claimed identity for all n ∈ N and f ∈ R. +For the converse, let n ≥ 1 be minimal such that Exng ̸= 0 for some g ∈ R. +With such g, we let f := +� ng and, by minimality of n, we obtain that +n−1 +� +k=0 +n−k +� +j=1 +(−1)n−k−jxkExj +� +xn−k−j∂n+1f = Exn +� +∂g = Exng − ExnEg = Exng +is nonzero. Consequently, Theorem 6.6 implies that the claimed identity does +not hold for this n and f. +Acknowledgements +This work was supported by the Austrian Science Fund (FWF): P 27229, +P 31952, and P 32301. Part of this work was done while both authors were +at the Radon Institute of Computational and Applied Mathematics (RICAM) +of the Austrian Academy of Sciences. The authors would like to thank Alban +Quadrat for bringing the book [23] to the attention of the second author during +his stay at INRIA Saclay. +A +Normal forms for IDO in tensor rings +The goal of this appendix is to state and prove a refinement of Theorem 4.2 +providing uniqueness of normal forms. Uniqueness is achieved by representing +operators by elements of a tensor ring, which is formed on a module of basic +operators generated by R, ∂, +� +, E. The ring of operators can be constructed +36 + +as quotient of the tensor ring, where relations of basic operators are encoded by +tensor reduction rules. The main technical tool for proving uniqueness of nor- +mal forms is a generalization of Bergman’s Diamond Lemma in tensor rings [3]. +For the convenience of the reader, we give a formal and largely self-contained +summary of tensor reduction systems and we explain the translation of identi- +ties of operators into this framework. In this appendix, K denotes a ring (not +necessarily commutative) with unit element. +In Section A.1, we start by recalling basic properties of bimodules and tensor +rings on them. For further details on tensor rings and proofs see, for example, +[34, 7]. Then, we recall decompositions with specialization, which are used for +defining reduction rules. In Section A.2, we state the Diamond Lemma for tensor +reduction systems with specialization from [14] and provide a summary of the +relevant notions. In Section A.3, we construct an appropriate tensor ring along +with a tensor reduction system for dealing with IDO. We use these to state The- +orem A.3, which provides a precise formulation of uniqueness of normal forms +of IDO in terms of tensors. In Section A.4, we give a complete proof of the +theorem, where the necessary computations for verifying uniqueness of normal +forms in the tensor ring are done in an automated way by our package TenReS in +the computer algebra system Mathematica. These computations are contained +in the Mathematica file accompanying this paper, which includes a log of the re- +duction steps and is available at http://gregensburger.com/softw/tenres/. +The theorem proved in that section even covers more general rings of operators, +which allow to deal with additional functionals besides the induced evaluation. +A.1 +Tensor rings on bimodules and decompositions +A K-bimodule is a left K-module M which is also a right K-module satisfying +the associativity condition (km)l = k(ml) for all m ∈ M and k, l ∈ K. A ring R +that is a K-bimodule such that (xy)z = x(yz) for any x, y, z in R or K is called +a K-ring. In particular, if K is a subring of some ring R, then R is a K-ring. +We first recall basic properties of the tensor product on K-bimodules. For +K-bimodules M, N, their K-tensor product M ⊗ N is a K-bimodule generated +by the pure tensors {m ⊗ n | m ∈ M, n ∈ N} with relations +(m + ˜m) ⊗ n = m ⊗ n + ˜m ⊗ n, +m ⊗ (n + ˜n) = m ⊗ n + m ⊗ ˜n, +and +mk ⊗ n = m ⊗ kn +having scalar multiplications +k(m ⊗ n) = (km) ⊗ n +and +(m ⊗ n)k = m ⊗ (nk) +for all m, ˜m ∈ M, n, ˜n ∈ N, and k ∈ K. +We denote the tensor product of M with itself over K by M ⊗n = M ⊗· · ·⊗M +(n factors). In particular, M ⊗1 = M and we interpret M ⊗0 as the K-module +Kε, where ε denotes the empty tensor and right scalar multiplication satisfies +(k1ε)k2 = (k1k2)ε for k1, k2 ∈ K. As a K-bimodule, the tensor ring K⟨M⟩ is +37 + +defined as the direct sum +K⟨M⟩ = +∞ +� +n=0 +M ⊗n. +It can be turned into a K-ring with unit element ε where multiplication M ⊗r × +M ⊗s → M ⊗(r+s) is defined via +(m1 ⊗ · · · ⊗ mr, ˜m1 ⊗ · · · ⊗ ˜ms) �→ m1 ⊗ · · · ⊗ mr ⊗ ˜m1 ⊗ · · · ⊗ ˜ms. +Via the tensor product, any decomposition of the module M carries over to +a decomposition of the tensor ring K⟨M⟩. In particular, we use decompositions +with specialization, which were introduced in [14]. These are given by a family +(Mz)z∈Z of K-subbimodules of M and a subset X ⊆ Z with M = � +z∈Z Mz = +� +x∈X Mx such that every module Mz, z ∈ Z, satisfies +Mz = +� +x∈S(z) +Mx +where S(z) := {x ∈ X | Mx ⊆ Mz} is the set of specializations of z. Note that +this definition implies S(x) = {x} for x ∈ X. For words W = w1 . . . wn in the +word monoid ⟨Z⟩, we define the corresponding K-subbimodule of K⟨M⟩ by +MW := Mw1 ⊗ · · · ⊗ Mwn. +The notion of specialization extends from the alphabet Z to the whole word +monoid ⟨Z⟩ by +S(W) := {v1 . . . vn ∈ ⟨X⟩ | ∀i : vi ∈ S(wi)} +such that S(W) = {V ∈ ⟨X⟩ | MV ⊆ MW }. We have the following generaliza- +tion +MW = +� +V ∈S(W) +MV +of the direct sum above and the decomposition +K⟨M⟩ = +� +W∈⟨Z⟩ +MW = +� +W∈⟨X⟩ +MW +of the tensor ring. +A.2 +Tensor reduction systems with specialization +Fixing a decomposition with specialization of M, a reduction rule for K⟨M⟩ is +given by a pair r = (W, h) of a word W ∈ ⟨Z⟩ and a K-bimodule homomorphism +h: MW → K⟨M⟩. It acts on tensors of the form a⊗w⊗b with a ∈ MA, w ∈ MW , +and b ∈ MB for some A, B ∈ ⟨Z⟩ by a ⊗ w ⊗ b →r a ⊗ h(w) ⊗ b. Later, we will +specify homomorphisms h in concrete reduction rules (W, h) via their values on +a generating set of MW . Formally, well-definedness of such homomorphisms can +38 + +be ensured by the universal property of the tensor product. A set Σ of such +reduction rules is called a reduction system over Z on K⟨M⟩ and induces the +two-sided reduction ideal +IΣ := (t − h(t) | (W, h) ∈ Σ and t ∈ MW ) ⊆ K⟨M⟩. +For computing in the factor ring K⟨M⟩/IΣ, we apply the reduction relation →Σ +induced by Σ on K⟨M⟩. It reduces a tensor t ∈ K⟨M⟩ to a tensor s ∈ K⟨M⟩ +if there is an r ∈ Σ such that t →r s. We say that t can be reduced to s by Σ +if t = s or there exists a finite sequence of reduction rules r1, . . . , rn in Σ such +that +t →r1 t1 →r2 · · · →rn−1 tn−1 →rn s. +If one tensor can be reduced to another, then their difference is contained in +IΣ and they represent the same element of K⟨M⟩/IΣ. The K-subbimodule of +irreducible tensors +K⟨M⟩irr = +� +W∈⟨X⟩irr +MW +can be characterized by the set of irreducible words ⟨X⟩irr ⊆ ⟨X⟩, which consists +of those words that avoid subwords arising as specializations S(W) of words +occurring in reduction rules (W, h) ∈ Σ. The irreducible tensors to which a +given tensor t can be reduced, are called its normal forms. If t has a unique +normal form, it is denoted by t↓Σ. +An ambiguity is a minimal situation where two (not necessarily distinct) re- +duction rules can be applied to tensors in different ways. For each pure tensor +of this kind, the corresponding S-polynomial is the difference of the results of +the two reduction steps. For example, an overlap ambiguity arises from two +reduction rules (AB1, h), (B2C, ˜h) ∈ Σ, where A, B1, B2, C ∈ ⟨Z⟩ are nonempty +such that B1, B2 are equal or have a common specialization, and corresponding +S-polynomials are referred to by SP(AB1, B2C). An ambiguity is called resolv- +able, if all its S-polynomials can be reduced to zero by Σ. If all ambiguities of Σ +are resolvable, then the reduction relation induced by Σ on K⟨M⟩ is confluent +and, by abuse of language, we also call Σ confluent. This means that there are +no hidden consequences implied in K⟨M⟩/IΣ by the identities explicitly specified +by Σ. +The following theorem relies on the existence of a partial order of words in +⟨Z⟩ that has certain properties, which are briefly explained now. A partial order +≤ on ⟨Z⟩ is called a semigroup partial order if it is compatible with concatenation +of words. If in addition the empty word ǫ is the least element of ⟨Z⟩, then ≤ +is called a monoid partial order. It is called Noetherian if there are no infinite +descending chains. We call a partial order ≤ on ⟨Z⟩ consistent with specialization +if every strict inequality V < W implies ˜V < ˜W for all specializations ˜V ∈ S(V ) +and ˜W ∈ S(W). A partial order ≤ on ⟨Z⟩ is compatible with a reduction system +Σ over Z on K⟨M⟩ if for all (W, h) ∈ Σ the image of h is contained in the sum +of modules MV where V ∈ ⟨Z⟩ satisfies V < W. +39 + +Theorem A.1. [14, Thm. 20] Let M be a K-bimodule and let (Mz)z∈Z be +a decomposition with specialization. +Let Σ be a reduction system over Z on +K⟨M⟩ and let ≤ be a Noetherian semigroup partial order on ⟨Z⟩ consistent with +specialization and compatible with Σ. Then, the following are equivalent: +1. All ambiguities of Σ are resolvable. +2. Every t ∈ K⟨M⟩ has a unique normal form t↓Σ. +3. K⟨M⟩/IΣ and K⟨M⟩irr are isomorphic as K-bimodules. +Moreover, if these conditions are satisfied, then we can define a multiplication +on K⟨M⟩irr by s · t := (s ⊗ t)↓Σ so that K⟨M⟩/IΣ and K⟨M⟩irr are isomorphic +as K-rings. +A.3 +Tensor reduction systems for IDO +Before using the tensor setting to construct the ring of integro-differential op- +erators R⟨∂, +� +, E⟩, we illustrate this construction on the well-known ring of +differential operators R⟨∂⟩ to highlight some of the special properties of the +construction. Usually, the ring of differential operators with coefficients from R +is constructed via skew polynomials � +i fi∂i over R in one indeterminate ∂, with +commutation rule ∂ · f = f · ∂ + ∂f. Let (R, ∂, +� +) be an integro-differential ring +and let C denote its ring of constants. In the following, we consider K-tensor +rings with K := C. +Example A.2. The module of basic operators that generates all differential +operators is given by +M := MR ⊕ MD, +with K-bimodules +MR := R +(24) +and MD defined as the free left K-module +MD := K∂ +(25) +generated by the symbol ∂, which we view as a K-bimodule with the right +multiplication c∂ · d = cd∂ for all c, d ∈ K. This definition is based on left +K-linearity of the derivation ∂ on R. The commutation rule ∂ · f = f · ∂ + ∂f +coming from the Leibniz rule in R translates to the tensor reduction rule +(DR, ∂⊗f �→ f⊗∂ + ∂f). +This rule is formalized by the K-bimodule homomorphism MDR = MD ⊗ MR → +K⟨M⟩ defined by ∂⊗f �→ f⊗∂ + ∂f on tensors ∂⊗f generating MDR. Note that +this homomorphism represents a parameterized family of identities. Reduction +of tensors by the rule above allows to syntactically move all multiplication op- +erators to the left of any differentiation operator. +40 + +Note that, in order to correctly model differential operators as equivalence +classes of tensors in K⟨M⟩, other relations among operators need to be phrased +as tensor reduction rules as well. This is because the tensor ring K⟨M⟩ itself +is constructed without respecting relations coming from multiplication in R. +For instance, the composition of two multiplication operators f and g is a mul- +tiplication operator again, which leads to the reduction rule (RR, f⊗g �→ fg) +defined on the module MRR = MR ⊗ MR. Moreover, the multiplication operator +that multiplies by 1 acts like the identity operator, which is represented by the +empty tensor ε. To define reduction rules that act only on C instead of all of R, +we need a direct decomposition of the K-bimodule +MR = MK ⊕ M˜R, +(26) +which in our case can be given by +MK := K +and +M˜R := +� +R +(27) +based on Lemma 2.3. Then, we can define a reduction rule on MK = C by 1 �→ ε. +Altogether, the tensor reduction system ΣDiff = {rK, rRR, rDR} for differential +operators is given by the three reduction rules +{(K, 1 �→ ε), +(RR, f⊗g �→ fg), +(DR, ∂⊗f �→ f⊗∂ + ∂f)} +defined above. It induces the two-sided ideal +IDiff := (t − h(t) | (W, h) ∈ ΣDiff and t ∈ MW ) += (1 − ε, f⊗g − fg, ∂⊗f − f⊗∂ − ∂f | f, g ∈ R) +in the ring K⟨M⟩ and computations with differential operators are modelled in +the quotient ring +R⟨∂⟩ := K⟨M⟩/IDiff. +Tensors that are not reducible w.r.t. ΣDiff are precisely K-linear combinations +of pure tensors of the form ∂⊗i and f ⊗ ∂⊗i, where i ∈ N0 and f ∈ +� +R. With +alphabets X = {K, ˜R, D} and Z = X ∪{R}, one can check that all ambiguities of +ΣDiff are resolvable. Using an appropriate ordering of words, one can show that +the conditions of Theorem A.1 are satisfied by this construction of R⟨∂⟩. +Proceeding to an analogous construction of the ring R⟨∂, +� +, E⟩ from Defini- +tion 4.1, we also require tensor reduction rules corresponding to the identities +(12)–(16), in addition to the three rules from the example above. To this end, +analogous to MD above, we introduce the K-bimodules +MI := K +� +and +ME := KE, +(28) +which are freely generated as left K-modules by the symbols +� +and E, respec- +tively. Then, we consider the K-tensor ring on the K-bimodule +M := MR ⊕ MD ⊕ MI ⊕ ME. +41 + +K +1 �→ ε +ID +� +⊗∂ �→ ε − E +RR +f⊗g �→ fg +DRE +∂⊗f⊗E �→ ∂f⊗E +DR +∂⊗f �→ f⊗∂ + ∂f +IRE +� +⊗f⊗E �→ +� +f⊗E +DI +∂⊗ +� +�→ ε +ERE +E⊗f⊗E �→ (Ef)E +Table 2: Defining reduction system for integro-differential operators +Altogether, we obtain the tensor reduction system given in Table 2. The two- +sided ideal IIDO induced by it allows to construct the ring R⟨∂, +� +, E⟩ as the +quotient K⟨M⟩/IIDO. However, this reduction system is not confluent. In order +to obtain normal forms that are unique as tensors in K⟨M⟩, we need a confluent +tensor reduction system on K⟨M⟩ that induces the same ideal IIDO. The con- +fluent tensor reduction system given in Table 3 can be obtained by turning the +identities of Table 1 into tensor reduction rules and including the rules rK and +rRR from above. This allows us to state the following more precise version of +K +1 �→ ε +EI +E⊗ +� +�→ 0 +RR +f⊗g �→ fg +IRD +� ⊗f⊗∂ �→ f − E⊗f − � ⊗∂f +DR +∂⊗f �→ f⊗∂ + ∂f +IRE +� +⊗f⊗E �→ +� +f⊗E +DE +∂⊗E �→ 0 +IRI +� +⊗f⊗ +� +�→ +� +f⊗ +� +− E⊗ +� +f⊗ +� +− +� +⊗ +� +f +DI +∂⊗ +� +�→ ε +ID +� +⊗∂ �→ ε − E +ERE +E⊗f⊗E �→ (Ef)E +IE +� +⊗E �→ +� +1⊗E +EE +E⊗E �→ E +II +� +⊗ +� +�→ +� +1⊗ +� +− E⊗ +� +1⊗ +� +− +� +⊗ +� +1 +Table 3: Confluent reduction system ΣIDO for integro-differential operators +Theorem 4.2. Note that multiples like f · ∂ are treated differently now, due to +the fact that we are working in the tensor ring K⟨M⟩ and we have the reduction +rule (K, 1 �→ ε), which splits f ∈ R according to Lemma 2.3. +Theorem A.3. Let (R, ∂, +� +) be an integro-differential ring with constants K = +C. Let M be given as above in terms of the modules defined in Eqs. (26), (27), +(25), and (28) and let the tensor reduction system ΣIDO be defined by Table 3. +Then every t ∈ K⟨M⟩ has a unique normal form t↓ΣIDO∈ K⟨M⟩, which can +be written as a K-linear combination of pure tensors of the form +f ⊗ ∂⊗j, +f ⊗ +� +⊗ g, +f ⊗ E ⊗ g ⊗ ∂⊗j, +or +f ⊗ E ⊗ h ⊗ +� +⊗ g +where j ∈ N0, f, g, h ∈ +� +R, and each f and g may be absent. Moreover, +R⟨∂, +� +, E⟩ ∼= K⟨M⟩irr +as K-rings, where multiplication on K⟨M⟩irr is defined by s · t := (s ⊗ t)↓ΣIDO. +Instead of proving this theorem, we will prove a more general one below, +which allows to include additional functionals into the construction of the ring +of operators. Theorem A.3 follows from Theorem A.5 by specializing Φ = {E}. +42 + +A.4 +Proof of normal forms for IDO with functionals +To treat more general problems than the initial value problems in Section 6.1, it +is useful to include additional functionals into the ring of operators. For instance, +dealing with boundary problems requires evaluations at more than one point. +In general, we consider a set Φ of K-linear functionals R → K including E. We +consider the K-bimodule MΦ defined as free left K-module +MΦ := KΦ +(29) +generated by the elements of Φ, where we define right multiplication in terms +of Φ by +�� +ϕ∈Φ cϕϕ +� +· d = � +ϕ∈Φ cϕdϕ for cϕ, d ∈ K. +Note that K-linear +combinations of K-linear maps are not necessarily K-linear again, if K is not +commutative. +Since ∂, +� +, and all elements of Φ are K-linear, we have, for +instance, that ϕfψg = (ϕf)ψg for all f, g ∈ R and ϕ, ψ ∈ Φ. +This allows +to extend the last three reduction rules of Table 2 to cover all elements of Φ +instead of the evaluation E only. Altogether, we consider the K-tensor ring on +the K-bimodule +M := MR ⊕ MD ⊕ MI ⊕ MΦ +(30) +and we have the defining reduction system given in Table 4. +K +1 �→ ε +ID +� ⊗∂ �→ ε − E +RR +f⊗g �→ fg +DRΦ +∂⊗f⊗ϕ �→ ∂f⊗ϕ +DR +∂⊗f �→ f⊗∂ + ∂f +IRΦ +� +⊗f⊗ϕ �→ +� +f⊗ϕ +DI +∂⊗ +� +�→ ε +ΦRΦ +ϕ⊗f⊗ψ �→ (ϕf)ψ +Table 4: Defining reduction system for integro-differential operators with func- +tionals +Definition A.4. Let (R, ∂, +� +) be an integro-differential ring with constants K = +C and let Φ be a set of K-linear functionals R → K including E. Let the K- +bimodule M be defined as above in Eqs. (24), (25), (28), (29), and (30). We +call +R⟨∂, +� +, Φ⟩ := K⟨M⟩/IIDOΦ +the ring of integro-differential operators with functionals Φ, where IIDOΦ is the +two-sided ideal induced by the reduction system obtained from Table 4 using also +submodules of M defined in Eq. (27). +We use Theorem A.1 above to determine unique normal forms of tensors. +Again, the reduction system given by Table 4 is not confluent and we need +to construct a confluent reduction system, like ΣIDOΦ given in Table 5, by a +completion process similar to how Table 3 was obtained. Observe that, whenever +Φ = {E}, the ring R⟨∂, � , Φ⟩ and the relation →ΣIDOΦ specialize to R⟨∂, � , E⟩ +and →ΣIDO, respectively. +43 + +K +1 �→ ε +EI +E⊗ +� +�→ 0 +RR +f⊗g �→ fg +IRD +� +⊗f⊗∂ �→ f − E⊗f − +� +⊗∂f +DR +∂⊗f �→ f⊗∂ + ∂f +IRΦ +� +⊗f⊗ϕ �→ +� +f⊗ϕ +DΦ +∂⊗ϕ �→ 0 +IRI +� +⊗f⊗ +� +�→ +� +f⊗ +� +− E⊗ +� +f⊗ +� +− +� +⊗ +� +f +DI +∂⊗ +� +�→ ε +ID +� +⊗∂ �→ ε − E +ΦRΦ +ϕ⊗f⊗ψ �→ (ϕf)ψ +IΦ +� +⊗ϕ �→ +� +1⊗ϕ +ΦΦ +ϕ⊗ψ �→ (ϕ1)ψ +II +� +⊗ +� +�→ +� +1⊗ +� +− E⊗ +� +1⊗ +� +− +� +⊗ +� +1 +Table 5: Confluent reduction system ΣIDOΦ for integro-differential operators +with functionals +Theorem A.5. Let (R, ∂, +� +) be an integro-differential ring with constants K = +C and let Φ be a set of K-linear functionals R → K including E. Let M and +R⟨∂, +� +, Φ⟩ be defined as in Definition A.4 above and let the tensor reduction +system ΣIDOΦ be defined by Table 5. +Then every t ∈ K⟨M⟩ has a unique normal form t↓ΣIDOΦ∈ K⟨M⟩, which can +be written as a K-linear combination of pure tensors of the form +f ⊗ ∂⊗j, +f ⊗ +� +⊗ g, +f ⊗ ϕ ⊗ h ⊗ ∂⊗j, +or +f ⊗ ϕ ⊗ h ⊗ +� +⊗ g +where j ∈ N0, f, g, h ∈ +� +R, ϕ ∈ Φ, and each f, g, h may be absent such that +ϕ ⊗ h ⊗ +� +does not specialize to E ⊗ +� +. Moreover, +R⟨∂, +� +, Φ⟩ ∼= K⟨M⟩irr +as K-rings, where multiplication on K⟨M⟩irr is defined by s · t := (s ⊗ t)↓ΣIDOΦ. +Proof. We use the alphabets X := {K, ˜R, D, I, E, ˜Φ} and Z := X ∪ {R, Φ}, which +turns (Mz)z∈Z into a decomposition with specialization for the module M, +where S(R) = {K, ˜R} and S(Φ) = {E, ˜Φ}. For defining a Noetherian monoid +partial order ≤ on ⟨Z⟩ consistent with specialization that is compatible with +ΣIDOΦ, it is sufficient to require the order to satisfy +DR > RD +and +I > E˜R. +For instance, we could first define a monoid total order on ⟨{R, D, I, Φ}⟩ ⊆ +⟨Z⟩ by counting occurrences of the letter I and breaking ties with any degree- +lexicographic order satisfying D > R and then generate from it a partial order +on ⟨Z⟩ that is consistent with specialization. +By our Mathematica package TenReS, we generate all ambiguities of ΣIDOΦ +and verify that they are resolvable. +There are 54 ambiguities and indeed +all S-polynomials reduce to zero, see the accompanying Mathematica file at +http://gregensburger.com/softw/tenres/. Here, we just give two short ex- +amples. The first one illustrates in its last step of computation that also iden- +44 + +tities in MR, like the Leibniz rule, need to be used. +SP(IRD, DR) = (f − E ⊗ f − +� +⊗ ∂f) ⊗ g − +� +⊗ f ⊗ (g ⊗ ∂ + ∂g) +→rRR fg − E ⊗ fg − +� +⊗ (∂f)g − +� +⊗ fg ⊗ ∂ − +� +⊗ f∂g +→rIRD − +� +⊗ (∂f)g + +� +⊗ ∂fg − +� +⊗ f∂g += +� +⊗ (−(∂f)g + ∂fg − f∂g) = 0 +The second one illustrates that ambiguities involving specialization also need to +be considered. +SP(ΦΦ, EI) = (ϕ1)E ⊗ +� +− ϕ ⊗ 0 →rEI 0 +Since all ambiguities are resolvable, by Theorem A.1 every element in K⟨M⟩ has +a unique normal form and R⟨∂, +� +, Φ⟩ ∼= K⟨M⟩irr as K-rings. +It remains to determine the explicit form of elements in K⟨M⟩irr. In order to +do so, we determine the set of irreducible words ⟨X⟩irr in ⟨X⟩. Irreducible words +containing only the letters K, ˜R, E, ˜Φ have to avoid the subwords arising from +the reduction rules K, S(RR) = {KK, K˜R, ˜RK, ˜R˜R}, S(ΦΦ) = {EE, E˜Φ, ˜ΦE, ˜Φ˜Φ}, +and S(ΦRΦ). Hence they are given by +ǫ, ˜R, E, ˜Φ, ˜RE, ˜R˜Φ, E˜R, ˜Φ˜R, ˜RE˜R, ˜R˜Φ˜R +Allowing also the letter D, we have to avoid the subwords coming from S(DR) = +{DK, D˜R} and S(DΦ) = {DE, D˜Φ}. Therefore, we can only append words Dj +with j ∈ N0 to the irreducible words determined so far, in order to obtain all +elements of ⟨X⟩irr not containing the letter I. Finally, we also consider the letter +I. Since we have to avoid the subwords S(IΦ) = {IE, I˜Φ}, ID, and II, any letter +immediately following I has to be ˜R. In addition, we have to avoid the subwords +S(IRΦ) = {IKE, IK˜Φ, I˜RE, I˜R˜Φ}, S(IRD) = {IKD, I˜RD}, and S(IRI) = {IKI, I˜RI}, +so the letter I cannot be followed by a subword of length greater than one. +Therefore, the letter I can appear at most once in an element of ⟨X⟩irr and, since +subwords EI and DI have to be avoided, it can only be immediately preceded by +the letters ˜R or ˜Φ. Altogether, the elements of ⟨X⟩irr are precisely of the form +˜RUDj +or +˜RV I˜R, +where j ∈ N0 and each of ˜R and U ∈ {E, ˜Φ, E˜R, ˜Φ˜R} and V ∈ {˜Φ, E˜R, ˜Φ˜R} may +be absent. +Remark A.6. As discussed in Remark 2.12, the induced evaluation of an +integro-differential ring often is multiplicative in concrete examples. Also when +considering several point evaluations of regular functions, the resulting func- +tionals are multiplicative, i.e. ϕfg = (ϕf)ϕg for all f, g ∈ R. We discuss how +reflecting this additional property of some functionals in the ring of operators +influences the normal forms of operators. To this end, we consider a subset +Φm ⊆ {ϕ ∈ Φ | ϕ is multiplicative and ϕ1 = 1} +45 + +of Φ, which may or may not include the evaluation E. For the corresponding +elements ϕ ∈ Φm in R⟨∂, +� +, Φ⟩, we impose +ϕ · f = (ϕf)ϕ +for all f ∈ R. +To model this identity by reduction rules, we consider the +submodule +MΦm := KΦm +of MΦ. Evidently, we have the decomposition +MΦ = ME ⊕ M˜Φm ⊕ M˜Φ +where the submodules M˜Φm and M˜Φ are generated by Φm \ {E} and Φ \ ({E} ∪ +Φm), respectively. We include the reduction rule +(ΦmR, ϕ⊗f �→ (ϕf)ϕ) +into Tables 4 and 5, where ϕ in the formula defining this K-bimodule homomor- +phism on MΦmR is not a general element of MΦm but of Φm and the definition +needs to be extended by left K-linearity to all of MΦmR. +Theorem A.5 generalizes to this situation with the additional restriction on +normal forms in K⟨M⟩ that h needs to be absent whenever ϕ ∈ Φm. +The +proof is analogous by adapting the alphabets and the order accordingly. The +additional reduction rule gives rise to additional ambiguities, whose resolvability +is also checked in the Mathematica file. Determination of irreducible words also +needs to be adapted accordingly. The resulting theorem includes Theorem A.5 +for Φm = ∅ and it includes Theorem 27 from [14] for Φm = Φ, i.e. when all +functionals are multiplicative. +References +[1] Matthias Aschenbrenner, Lou van den Dries, and Joris van der Hoeven, +Asymptotic differential algebra and model theory of transseries, Annals of +Mathematics Studies, vol. 195, Princeton University Press, Princeton, NJ, +2017. +[2] Vladimir V. Bavula, The algebra of integro-differential operators on a poly- +nomial algebra, J. Lond. Math. Soc. (2) 83 (2011), 517–543. +[3] George M. Bergman, The diamond lemma for ring theory, Adv. in Math. +29 (1978), 178–218. +[4] Bruno Buchberger, An algorithm for finding the bases elements of the +residue class ring modulo a zero dimensional polynomial ideal (German), +Ph.D. thesis, University of Innsbruck, 1965. +46 + +[5] Thomas Cluzeau, Jamal Hossein Poor, Alban Quadrat, Clemens G. Raab, +and Georg Regensburger, Symbolic computation for integro-differential- +time-delay operators with matrix coefficients, IFAC-PapersOnLine 51 +(2018), no. 14, 153–158. +[6] Earl A. Coddington and Norman Levinson, Theory of ordinary differential +equations, McGraw-Hill Book Co., Inc., New York-Toronto-London, 1955. +[7] Paul M. Cohn, Further algebra and applications, Springer-Verlag London, +Ltd., London, 2003. +[8] Gerald A. Edgar, Transseries for beginners, Real Anal. Exchange 35 (2010), +no. 2, 253–309. +[9] Xing Gao, Li Guo, and Markus Rosenkranz, On rings of differential Rota- +Baxter operators, Internat. J. Algebra Comput. 28 (2018), no. 1, 1–36. +[10] Li Guo, An introduction to Rota-Baxter algebra, Surveys of Modern Math- +ematics, vol. 4, International Press, Somerville, MA; Higher Education +Press, Beijing, 2012. +[11] Li Guo and William Keigher, On differential Rota-Baxter algebras, J. Pure +Appl. Algebra 212 (2008), 522–540. +[12] Li Guo, Georg Regensburger, and Markus Rosenkranz, On integro- +differential algebras, J. Pure Appl. Algebra 218 (2014), 456–473. +[13] Jamal Hossein Poor, Tensor reduction systems for rings of linear operators, +Ph.D. thesis, Johannes Kepler University Linz, Austria, 2018. +[14] Jamal Hossein Poor, Clemens G. Raab, and Georg Regensburger, Algo- +rithmic operator algebras via normal forms in tensor rings, J. Symbolic +Comput. 85 (2018), 247–274. +[15] Irving Kaplansky, An introduction to differential algebra, Publ. Inst. Math. +Univ. Nancago, No. 5, Hermann, Paris, 1957. +[16] William F. Keigher, Adjunctions and comonads in differential algebra, Pa- +cific J. Math. 59 (1975), no. 1, 99–112. +[17] +, On the ring of Hurwitz series, Comm. Algebra 25 (1997), no. 6, +1845–1859. +[18] William F. Keigher and F. Leon Pritchard, Hurwitz series as formal func- +tions, J. Pure Appl. Algebra 146 (2000), no. 3, 291–304. +[19] Donald E. Knuth and Peter B. Bendix, Simple word problems in universal +algebras, Computational Problems in Abstract Algebra, Pergamon, Oxford, +1970, pp. 263–297. +47 + +[20] Ernst E. Kummer, Ueber die Transcendenten, welche aus wiederholten Inte- +grationen rationaler Formeln entstehen, J. Reine Angew. Math. 21 (1840), +74–90, 193–225, and 328–371. +[21] Ivan A. Lappo-Danilevski, Résolution algorithmique des problèmes réguliers +de Poincaré et de riemann (Mémoire premier: Le problème de Poincaré, +concernant de la construction d’un groupe de monodromie d’un système +donné d’équations différentielles linéaires aux intégrales régulières), J. Soc. +Phys.-Math. Léningrade 2 (1928), no. 1, 94–120. +[22] Teo Mora, An introduction to commutative and noncommutative Gröbner +bases, Theoret. Comput. Sci. 134 (1994), no. 1, 131–173. +[23] Danuta Przeworska-Rolewicz, Algebraic analysis, PWN—Polish Scientific +Publishers, Warsaw; D. Reidel Publishing Co., Dordrecht, 1988. +[24] +, Two centuries of the term “algebraic analysis”, Algebraic analysis +and related topics (Warsaw, 1999), Banach Center Publ., vol. 53, Polish +Acad. Sci. Inst. Math., Warsaw, 2000, pp. 47–70. +[25] Alban Quadrat, A constructive algebraic analysis approach to Artstein’s +reduction of linear time-delay systems, IFAC-PapersOnLine 48 (2015), +no. 12, 209–214. +[26] Alban Quadrat and Georg Regensburger, Computing polynomial solutions +and annihilators of integro-differential operators with polynomial coeffi- +cients, Algebraic and symbolic computation methods in dynamical systems, +Adv. Delays Dyn., vol. 9, Springer, Cham, 2020, pp. 87–114. +[27] Rimhak Ree, Lie elements and an algebra associated with shuffles, Ann. of +Math. (2) 68 (1958), 210–220. +[28] Georg Regensburger, Markus Rosenkranz, and Johannes Middeke, A skew +polynomial approach to integro-differential operators, Proceedings of ISSAC +’09 (New York, NY, USA), ACM, 2009, pp. 287–294. +[29] Markus Rosenkranz, A new symbolic method for solving linear two-point +boundary value problems on the level of operators, J. Symbolic Comput. 39 +(2005), 171–199. +[30] Markus Rosenkranz and Georg Regensburger, Solving and factoring bound- +ary problems for linear ordinary differential equations in differential alge- +bras, J. Symbolic Comput. 43 (2008), 515–544. +[31] Markus Rosenkranz, Georg Regensburger, Loredana Tec, and Bruno +Buchberger, Symbolic analysis for boundary problems: From rewriting to +parametrized Gröbner bases, Numerical and Symbolic Scientific Computing: +Progress and Prospects (Ulrich Langer and Peter Paule, eds.), Springer Vi- +enna, Vienna, 2012, pp. 273–331. +48 + +[32] Markus Rosenkranz and Nitin Serwa, An integro-differential structure for +Dirac distributions, J. Symbolic Comput. 92 (2019), 156–189. +[33] Gian-Carlo Rota, Twelve problems in probability no one likes to bring up, +Algebraic combinatorics and computer science, Springer Italia, Milan, 2001, +pp. 57–93. +[34] Louis H. Rowen, Ring theory, student ed., Academic Press, Inc., Boston, +MA, 1991. +[35] Richard P. Stanley, Differentiably finite power series, European J. Combin. +1 (1980), no. 2, 175–188. +[36] Joris van der Hoeven, Transseries and real differential algebra, Lecture +Notes in Mathematics, vol. 1888, Springer-Verlag, Berlin, 2006. +49 +