Greenrgb0,.8,0 \definecolorGreenrgb0.20,0.43,0.09
Differential Galois Theory and Integration
Abstract
In this chapter, we present methods to simplify reducible linear differential systems before solving. Classical integrals appear naturally as solutions of such systems. We will illustrate the methods developed in DW (19) on several examples to reduce the differential system. This will give information on potential algebraic relations between integrals.
Keywords:
Ordinary Differential Equations, Differential Galois Theory, Computer Algebra, Integrals, Lie Algebras, -finite functions.Introduction
In this chapter, we will review properties of block triangular linear differential systems and their use to compute properties of integrals.
Let and . We will study the corresponding linear differential system . More generally, we might consider linear differential systems over a differential field of characteristic zero, that is a field equipped with an additive morphism satisfying the Leibniz rule .
The Galois theory of linear differential equations aims at understanding what are the algebraic relations between the solutions of . We attach to a group that measures the relations. The computation of this differential Galois group is a hard task in full generality. The goal of this chapter is to illustrate on examples the method described in DW (19) that focuses on the reduction of block-triangular linear differential systems. This approach is powerful enough to understand the desired relations on the solutions.
Given an invertible matrix , the linear change of variables produces a new differential system denoted where
Two linear differential systems and are called (gauge) equivalent over if there exists a gauge transformation, an invertible matrix , such that . A linear differential system is called reducible (over ) when it is gauge equivalent to a linear differential system in block triangular form:
(1) |
It turns out that computing properties of integrals of -finite functions111A function is -finite when its is a solution of a linear differential equations with coefficients in or of some types of iterated integrals may be reduced to computing solutions of such reducible linear differential systems. Computing the differential Galois groups of block triangular systems gives, in turn, information on properties of their solutions. This idea was promoted by Bertrand Ber (01) or Berman and Singer BS (99) who showed how to compute Galois groups of some reducible systems and how this would reveal algebraic properties of integrals.
Our aim, in this chapter, is to show how to compute such algebraic properties. The underlying theory is developed in DW (19) and CW (18); general references for differential Galois theory are for example PS (03); CH (11); Sin (09); general references for constructive theory of reduced forms of differential systems are AMCW (13); MS (02); AMDW (16); BCDVW (20); BCWDV (16). We will rely on many examples rather than on cumbersome theory and provide references for interested readers. For example, we will consider the dilogarithm function defined by and our -dimensional example (in the next sections) will provide a simple algorithmic proof that it is not only transcendent but algebraically independent of and ; the proof will only use rational solutions of linear first order differential equations.
The chapter is organized as follows. We begin by some examples to illustrate what the method will provide. Then we give a review of differential Galois theory and reduced forms of differential systems. We finish by explaining the strategy of DW (19) on two examples that are chosen so that almost all calculations can be reproduced easily.
Acknowledgment This project has received funding from the ANR project De Rerum Natura ANR-19-CE40-0018. We warmly thank the organizers of the conference Antidifferentiation and the Calculation of Feynman Amplitudes for stimulating exchanges and lectures. We specially thank the referee for many observations and precisions that have enhanced the quality of this text.
1 Examples
1.1 A first toy example
We consider two confluent Heun222See https://dlmf.nist.gov/31.12 or Mot (18). The notation is the syntax in Maple. functions
and
They form a basis of solutions of the second order linear differential equation
The Wronskian relation gives us the algebraic relation . The equation has order two so the Kovacic algorithm Kov (86); UW (96); vHW (05) can be used to compute the differential Galois group and we find that no other algebraic relations exist between , , and . Now let be a primitive. We want to determine whether and are algebraically independent of the and or not. The techniques explained below will show that this question reduces to asking whether there is a rational solution to the linear differential system
or, equivalently, whether the following linear differential equation (the left hand side turns out to be the adjoint operator of ) has a rational solution:
It can be seen directly or with a computer algebra system that this equation has no rational solution. The underlying theoretical tools are from the constructive differential Galois theory. However, in operational terms, it is rather easy to compute and check : no hard theory is required for the calculation. Let us unveil a corner of the underlying tools.
We have in fact studied the differential Galois group of the differential system with
which admits the fundamental solution matrix
Our calculation of rational solution above shows (with the tools displayed below) that the differential Galois group of the system will have dimension . This in turn shows that the integrals are algebraically independent of , and their derivatives. In fact, it even shows that both integrals are algebraically independent.
1.2 A second toy example
Recall that the hypergeometric function is given by the formula
We now take two hypergeometric functions (with nice modular properties)
and
The vectors are solutions of below. To study properties of their integrals, we set and we have where
We will see in the sequel how we may find a suitable change of variables
such that
This shows that
and the differential Galois group of has dimension 3. We note that none of these two examples is new and that efficient methods to handle these questions have been developed by Abramov and van Hoeij in AvH (99, 97).
1.3 Integrals via reducible systems
These two examples have shown how, by augmenting the dimension of a linear differential system, we can study integrals of its solutions. Conversely, block triangular systems give rise to integrals via variation of constants; we review this for clarification. For a factorized reducible system of the form
with and , we have a fundamental solution matrix of the form
Once and are known, is given by integrals : .
So it may seem that no further theory is required. However, and are fundamental solution matrices for systems so the relation may involve integrals of complicated -finite functions.
Our approach will be to first “reduce” as much as possible, using manipulations with rational functions, to prepare the system for easier solving. In return, we will obtain algebraic information on all (possibly iterated) integrals that may occur in the relation .
We note, for the record, that factorized differential operators also correspond to block triangular differential systems.
1.4 Example of situations involving reducible linear differential systems
Operators from statistical physics and combinatorics
In many models attached to statistical mechanics, quantities are expressed as multiple integrals depending on a parameter. They are generally holonomic in this parameter, meaning that they are -finite, i.e they are solutions to linear differential operators (for example, the so-called Ising operators). A description of this setting may be found in the books of Baxter Bax (82) and McCoy McC (10) (notably Chapters 10 and 12) or in the surveys MM (12); MAB+ (11) (and references therein). Similarly, in combinatorics, sequences that satisfy recurrence relations may be studied through their generating series, which are often -finite.
Experimentally (see, for example, BHM+ (07); BBH+ (09, 10, 11); BHMW (14, 15) or HKMZ (16); Kou (13); ABKM (20, 18)), it turns out that many differential operators coming from these processes are factored into a product of smaller order factors - and the corresponding companion systems are reducible.
Other cases of reducible systems are the ones admitting reducible monodromies such as ST (16) or in the works of Kalmykov on Feynman integrals via Mellin-Barnes integrals Kal (06); BKK (10); KK (12, 17, 11); about differential equations for Feynman integrals, one may consult the works of Smirnov Smi (12); LSS (18).
Finally, we mention the paper ABRS (14) with other examples of integrals where the techniques presented below may offer alternative approaches for some of the computations.
Variational equations of nonlinear differential systems
Another natural source of reducible systems is the old method of variational equations of nonlinear differential systems along a given particular solution . One can form the linear differential equation describing perturbations along this solution . The general principle is that obstructions to integrability of the nonlinear system can be read on this linear differential system. Ziglin Zig (82) linked non-integrability to non-commutations in the monodromy group with concrete versions given e.g. in CR (91) and Sal (14). It was generalized in the theory of Morales-Ruiz and Ramis and extended with Simó MRR (01); MRRS (07) for Hamiltonian systems; they prove that a Hamiltonian system is completely integrable only if all its variational equations have a virtually abelian differential Galois group. Extensions to nonhamiltonian differential equations can be found in AZ (10); CW (18). Applications to specific problems have been occasions to establish effective criteria. For example: the three body problem Tsy (01); BW (03), -body problems MRS (09); Com (12), Hill systems (movements of the moon) MRSS (05); AMW (12) or a swinging Atwood machine PPR+ (10).
These variational equations can be written in the form of reducible linear differential systems of big dimension. The simplification techniques outlined below are hence particularly relevant to make computation on such systems practical (see AMDW (16)).
We note that the Morales-Ramis theory has had a spectacular recent development, initiated in MR (20) where this variational approach is applied to path integrals thus establishing a beautiful and unexpected bridge with the previous subsection.
Reducible linear differential systems appear naturally in another type of perturbative approach: in -expansions of solutions for perturbed systems like the ones that appear in works of J. Blümlein, C. Raab, C. Schneider, J. Henn and others for example.
2 Reduced forms of linear differential systems
2.1 Ingredient #1 : Differential Galois-Lie Algebra
In what follows, is a differential field of characteristic . We outline a brief exposition of the Galois theory of linear differential equations, see PS (03); CH (11); Sin (09) for expositions with proofs. We consider a linear differential system
(2) |
Let be the field of constants of the differential field , that is . We will assume that is algebraically closed, i.e, every non constant polynomial equation has a solution in .
A Picard-Vessiot extension is a field , where is a fundamental solution matrix333An invertible matrix such that . of , such that the field of constants of is still . This can be constructed algebraically; alternatively, when is a field of meromorphic functions, one may consider a local matrix of power series solutions at a regular point and this gives a Picard-Vessiot extension. A Picard-Vessiot extension is unique modulo differential field isomorphisms.
The differential Galois group is the set of automorphisms of which leave the base field fixed and commute with the derivation. Let . By construction, is also a fundamental solution matrix in and we find that there exists a matrix such that . The map provides a faithful representation of as a subgroup of , actually a linear algebraic group. If we change the fundamental solution, we obtain a conjugate representation.
We recall that, given an invertible matrix , the linear change of variables produces a new differential system denoted where
We note that such a gauge transformation , with , does not change the Galois group.
Given a Picard-Vessiot extension , the polynomial relations among all entries of (over ) form an ideal . The Galois group can then be viewed as the set of matrices which stabilize this ideal of relations. Thus the computation of is strongly related to the understanding of the algebraic relations among the solutions.
The Galois-Lie algebra of is the Lie algebra of the differential Galois group . It is defined as the tangent space of at the identity . The dimension of measures the transcendence degree of over , that is
One way of computing the Lie algebra (PS (03)) is the following : is the set of matrices such that with . In other terms, satisfies the defining equations of the group modulo . For example, (set of such that ) gives the Lie algebra of matrices such that . The symplectic group is the set of such that ; its Lie algebra is found to the set of such that , with . The additive group admits the Lie algebra ; the multiplicative group admits the Lie algebra .
The reduction technique exposed in this chapter aims at computing directly the Galois-Lie algebra before computing itself. Although the theory is not obvious, the resulting calculations are reasonably simple.
2.2 Ingredient #2 : Lie algebra associated to
The Lie algebra associated to the matrix is defined as follows. Let be a basis of the -vector space generated by the coefficients of . We can then decompose as where the are constant matrices. Now, we consider the smallest Lie algebra containing all the : this is the vector space generated by the and all their iterated Lie brackets (); then we take its algebraic envelope.
Definition 1
The decomposition is not unique but the vector space generated by the is unique. Thus, the associated Lie algebra does not depend on the chosen decomposition.
This Lie algebra appears in works of Magnus Mag (54) or Feynman who use the Baker-Campbell-Hausdorf formula to write solutions of as (infinite) products of exponentials constructed with Lie brackets. Wei and Norman give in WN (63, 64) a finite formula to solve the system when is solvable. This formula is well-known in physics and control theory but not as well among mathematicians.
The terminology of Lie algebra associated to appears555For this reason, some authors, including ourselves, call the decomposition a Wei-Norman decomposition of . in WN (63, 64) (in there, it is defined as the Lie algebra generated by all values of for spanning all constants minus singularities and the algebraic envelope is missing). In the sequel, we will study a -dimensional example and an -dimensional example where our technique will have some relations to the Wei-Norman approach; namely, we will change to obtain an associated Lie algebra of minimal dimension so that solving formulas become optimal in some sense.
Example 1 (A 4-dimensional example)
Let
Note that this system is upper triangular, contrary to (1). This will illustrate that our method can be equivalently applied to upper and lower triangular systems. We obtain a Wei-Norman decomposition , where
We have
All the other brackets in are zero and we find that . It is solvable of depth : the first derived algebra (the set of all matrices in Lie(A) which can be written as a Lie bracket) is and the second derived algebra is . We will continue below with this example.
2.3 Linear differential systems in reduced form
We now turn to the link between and differential Galois theory, based on two important results of Kolchin and Kovacic. Proofs can be found in PS (03), Proposition 1.31 and Corollary 1.32; see also BSMR (10), Theorem 5.8, and AMCW (13), § 5.3 after Remark 31. Let ; let be the differential Galois group of and its Lie algebra, the Galois-Lie algebra of the system . The first result is
So, the Lie algebra associated to , an object which is very easy to compute, provides an “upper bound” on . When we perform a gauge transformation to obtain a new system , and are invariant while may vary. This shows that is a lower bound on all , for all gauge transformations . The second result of Kolchin and Kovacic is that this lower bound is reached. By definition, is the Lie algebra of an algebraic connected group . Then there exists a gauge transformation666The notation denotes matrices whose entries are in and satisfy all the equations defining the algebraic group . such that . Furthermore, if is connected and under the very mild additional condition that is a -field777 A field is a -field when every non-constant homogeneous polynomial over has a non-trivial zero provided that the number of its variables is more than its degree. For example, is a -field and any algebraic extension of a -field is a -field (Tsen’s theorem). then we may choose (no algebraic extension).
Definition 2
A system is in reduced form when the Lie algebra associated to is equal to the Lie algebra of the differential Galois group of .
The results of Kolchin and Kovacic show that a reduced form always exists. We provide, in the sequel, constructive methods to obtain them when the systems are in block-triangular form. They will be illustrated on our -dimensional example in the next section.
Example 2 (4-dimensional example, continued)
We will show, in the next section, that is in reduced form. This system is easily integrated step by step and we find a fundamental solution matrix
where dilog is defined by . We may take . When computing , the terms requiring one integration ( and ) correspond to the terms in and in the Wei-Norman decomposition of . The term dilog comes from the existence of , the Lie bracket of and in .
The corresponding Galois group is a semi-direct product of a -dimensional torus (giving rise to the in the solution) and of a vector group generated by , (giving rise to the terms in ) and (giving rise to the dilog). Its Lie algebra is .
The ideas behind this notion of reduced form have been used for inverse problems in differential Galois theory: given an algebraic group , construct a differential system having as its differential Galois group. It is also a technique known in differential geometry.
Its use for direct problems in differential Galois theory is more recent. A remark in PS (03) suggests that this would be a good idea. In the context of Lie-Vessiot systems, Blazquez and Morales exploit this idea in BSMR (10). It is developed in AMDW (16); AMW (12, 11) in order to study variational equations in the context of integrability of Hamiltonian systems and the Morales-Ramis-Simó theory (and later in CW (18) to study algebraic properties of Painlevé equations). For irreducible systems (or systems in block diagonal form), a criterion for reduced forms is established in AMCW (13) with a decision procedure. Another, much more efficient approach is given in BCWDV (16); BCDVW (20) together with generalizations of the criterion of AMCW (13).
The approach described below allows, given the above results, to compute a reduced form of a block triangular linear differential system (the last case remaining after all the above contributions). It is based upon these works, notably CW (18), and is constructed in DW (19).
3 How to Compute a Reduced Form of a Reducible System
Assume now that , where is algebraically closed of characteristic zero where the derivation acts trivially. It is in particular a -field. We consider a block triangular system over the differential field in the same form as (1), that is
Let . In what follows, we will assume that the block diagonal part is in reduced form and we will show how to find a gauge transformation such that is in reduced form. By DW (19), Lemma 2.7, the differential Galois group is connected and the reduction matrix we are looking for has coefficients in . Instead of reproving all the theory (which, in this case, can be mostly found in DW (19)), we will work out in details a simple example where most of the required algorithmic elements appear. This may help convince the reader of how the method works (the details in DW (19) may be technical, at least in a first reading).
3.1 Shape of the gauge transformation
Let be the the set of off-diagonal constant matrices of the form (same sizes as in relation (1)). We will extend the scalars to , the off-diagonal matrices with coefficients in . Our first step is that we may find a reduction matrix in a very particular shape.
Lemma 1 (AMDW (16), Lemma 3.4)
There exists a gauge transformation such that is in reduced form.
The following is is based on an observation from AMW (12, 11). Let , . Suppose that, for all , we have ; then, is in reduced form. In other terms, no rational gauge transformation can turn it into a system with a smaller associated Lie algebra. In this case, will be the Lie algebra of the differential Galois group and this will give us transcendence relations and algebraic relations on the solutions; this will be seen on the main example of this section.
More generally, as we can see in DW (19), Section 5, if our method can reduce a system with two diagonal blocks then we can iterate this method to obtain a reduced form of a block-triangular system with an arbitrary number of blocks on the diagonal. Let us illustrate this iteration on a system with three diagonal blocks of the form
where the block diagonal part is in reduced form (see BCDVW (20); BCWDV (16) for this). We will first reduce the south-east part (which is of the same form as (1)) into a form
Let be the reduction matrix. By DW (19), Lemma 5.1, the following system is automatically in reduced form
Now we perform the gauge transformation to obtain a system of the form
(the may have changed after the first reduction step). We now see that this system in the same form as (1) with as the block diagonal matrix. So a second reduction of a two-blocks triangular system allows to reduce the initial three-blocks triangular system.
This iteration is well seen in our -dimensional example below.
Example 3 (4-dimensional example, continued)
Let
A simple application of a factorization algorithm shows that it is reducible. Indeed, letting
we have
This example is, of course, particularly simple. We use it to show how to apply the iteration procedure and Lemma 1 to simplify the system or prove that it cannot be simplified further.
Since we consider an upper triangular system, we start with the “north-west” corner. We let
The diagonal part is in reduced form (solutions are and cannot be simplified using rational functions). The associated Lie algebra has dimension 2. Reduction would imply to have dimension . By Lemma 1, a reduction matrix would have the form
The north-east coefficient of is . The coefficient could never be constant (the equation has no rational solution, the simple pole cannot be canceled by the derivative of a rational function). For any choice of , will have dimension . It follows that is in reduced form.
So we iterate.
We now pick a bigger matrix :
By DW (19), Lemma 5.1, and by the above calculation, we find that is in reduced form. Lemma 1 thus shows that a reduction matrix would have the simple form
Now has dimension and has dimension 3. A reduction matrix should therefore map to a matrix with an associated Lie algebra of dimension .
We have
so there should exist constants and a rational function such that
. A necessary condition is and .
Similarly, there should exist constants such that .
We now plug our condition on into this relation and find that there should be a rational function such that
Now, because of the pole of order at , this can never have a rational solution (whatever the values of the unknown constants). It follows that, for any choice of , will have dimension .
So is in reduced form.
Note that our main ingredient here has been to look for a rational solution of an inhomogeneous linear differential equation whose right-hand side contains parameters. There exist algorithms to compute conditions on the parameters (from the right-hand side) so that such an equation has rational solutions, see subsection 3.2 below or Sin (91), and this will be the key to what follows.
We continue iterating the reduction process. Now we will have
Using again DW (19), Lemma 5.1, we see that is in reduced form. Furthermore, has dimension and has dimension . We compute .
The relation gives conditions and . The same study on gives us . Without finishing with the last coefficient, we see that contains matrices of the following forms (respectively because of terms in and ):
whose Lie bracket is
So we see that, whatever our future choices may be, will contain and hence have dimension . This shows that our system cannot be reduced so it is in reduced form. Furthermore, this suggests that our reduction conditions might have been stronger : requiring was not enough. We could have imposed and . Both these relations are easily seen to be impossible to fulfill with rational functions so our system is again seen to be in reduced form.
To summarize what this example suggests: we need to “cancel” terms in the purely triangular part; this reduces to finding rational solutions of linear differential equations with parametrized right-hand sides. And the order of the computations matters: here, one needs to study the relations on and before studying relations on . We will show, in the sequel, how to systematize these ideas, using an isotypical decomposition and an adapted flag structure, and how to make them algorithmic so that a computer algebra system may perform the calculations.
In this example, we had seen that a fundamental solution matrix could be written using and dilog(x). As is in reduced form, is the Lie algebra of the Galois group and it has dimension . This shows that these four functions are transcendent and algebraically independent. So our calculation above (long but not hard) gave us a simple proof that dilog(x) is algebraically independent of .
3.2 The adjoint action of the diagonal
We recall our notations so far. We have matrices with coefficients in and
If we take two off-diagonal matrices and in , we have . This allows two simple calculations. First, let , with , . Then
(3) |
Furthermore, . These two calculations show that reduction will be governed by the adjoint action of the block diagonal part on . This adjoint action is a linear map. Its matrix, on the canonical basis of , is
When has an abelian Lie algebra we may easily compute a Jordan normal form of . Furthermore the eigenvalues of belong to . This is the idea behind AMDW (16). In our case, we will need a more subtle structure, an isotypical decomposition into -invariant subspaces of .
Example 4 (An 8-dimensional example)
We consider a matrix given by
The block diagonal part is given by two copies of our -dimensional example (here, ) and we have shown that it was in reduced form. The off-diagonal part is given by
As , the matrix of the adjoint action of the diagonal on is a (sparse) matrix given by .
Isotypical decomposition
Recall that is the -vector space of off-diagonal matrices. We now show how the adjoint action of the diagonal will govern the reduction strategy on .
Definition 3
A vector space will be called a -space if .
The importance of these -spaces is stated in the following lemma.
Lemma 2 (DW (19), Lemma 2.11)
Let and assume that is in reduced form. Then, is a -space.
So our reduction strategy will be to try to project onto the smallest possible -space using rational gauge transformations. In DW (19), we provide references to algorithms to decompose and factor into -spaces. This is obtained using an isotypical decomposition (eigenring methods) and a flag structure.
Lemma 3 (Krull-Schmidt)
The -vector space admits a unique isotypical decomposition
where
-
•
each is a -space;
-
•
, a direct sum of -spaces that are all isomorphic to an indecomposable -space which admits a flag decomposition
and is a sum of isomorphic irreducible -spaces for ;
-
•
For , the -spaces and are not isomorphic.
Once this decomposition and flag structure are computed, we perform, at each stage, a projection on a minimal -subspace in . For some vectors and a matrix with coefficients in (obtained by linear algebra), this reduces to computing all tuples , with and constants, such that
The resulting system will be “minimal”: it will be in reduced form. The proof of this result is technical and can be found in DW (19). We will illustrate the process on our main example.
Example 5 (8-dimensional example, continued)
In this example,
decomposes as a direct sum of three indecomposable
-spaces.
We first study the adjoint action of on .
We find (see DW (20)) an adapted basis given by off-diagonal matrices with south-west blocks
The matrix of the adjoint action on this basis of is
The flag structure on suggests the following reduction path. Try to remove elements in if possible; then in ; then in . How to do this will be made clear in the next section; the flag structure guides the order in which computations should be handled.
We turn to . We find (see DW (20)) a basis adapted to the flag structure given by off-diagonal matrices whose south-west blocks are:
The matrix of the adjoint action on this adapted basis is:
Intermezzo : Reduction and Rational Solutions
Before we continue, let us make a quick excursion into our main algorithmic toolbox. we start with a simple case. We look for a condition on to have
A simple calculation shows that should be a rational solution of the matrix linear differential system . If we let vec denote the operator transforming a matrix into a vector by stacking its rows, we find (see DW (19)) that , where is again the adjoint action of the diagonal defined above. So reduction will be governed by computing rational solutions of linear differential systems. When , a computer algebra algorithm for this task has been given by Barkatou in Bar (99), see BCEBW (12) for a generalization to linear partial differential systems and a Maple implementation.
Now, our general tool (also found in the above references) will be an apparently more complicated problem. Given a matrix and vectors , we will look for
tuples , with and constant, such that
. Such tuples form a computable vector space and the algorithms in Bar (99); BCEBW (12) provide this when . Results and algorithms for general fields can be found in Sin (91).
We now pick concrete coefficients to show how to perform the reduction on our -dimensional example. A Maple worksheet888The reader may also find a pdf version at
http://www.unilim.fr/pages_perso/jacques-arthur.weil/DreyfusWeilReductionExamples.pdf with this example and the chosen coefficients may be found at DW (20).
Reduction on (8-dimensional example)
To remove all of , it would be enough to have a rational solution to the system
and is given in Example 5 (page 5). This gives us reduction equations
The first two equations correspond to the highest level of the flag. To remove an element from , there should be a rational solution to the equation . The -vector space of pairs such that there exists with is found to be -dimensional; for , we have ; for , we have , where the are arbitrary constants (their importance will soon be visible). Our gauge transformation is and does not contain any terms from .
Now is -dimensional. The equation for the reduction on is now
We have necessary and sufficient conditions on the parameters to have a rational solution, namely , and then a general rational solution . Our new gauge transformation is and does not contain any term from any more.
Finally, we look for all such that is rational: we look for non-zero pairs such that there exists a rational solution of
This integral is rational if and only if both residues are zero. As the solution is not admissible, we see that a necessary and sufficient condition is . The set of desired pairs is of dimension . For , we have , for , we have , where the are constants and can be chosen arbitrarily. So our last gauge transformation matrix will be and the reduction matrix on is
The resulting matrix contains no terms from .
Reduction on (8-dimensional example)
The matrix is given in Example 5 page 5. The reduction equations are now
We will let the reader solve this iteratively following the method from the previous section. This will give the following successive reductions
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/615bd67e-fb6d-46f4-b399-a859d020eec3/flag3.jpg)
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/615bd67e-fb6d-46f4-b399-a859d020eec3/flag4.jpg)
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/615bd67e-fb6d-46f4-b399-a859d020eec3/flag5.jpg)
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/615bd67e-fb6d-46f4-b399-a859d020eec3/flag6.jpg)
where the green denotes parts that have been successfully removed.
However, we reach an obstruction when trying to remove
(once the equation for has a rational solution, the equation for cannot have a rational solution).
The reduction matrix is
and we obtain the reduced form :
The associated Lie algebra is spanned by
This gives us the Lie algebra of the differential Galois group. Note that, during the reduction process, we found the two incompatible equations and , where is a constant. There were two mutually exclusive paths: either remove or remove . We removed here by setting ; the choice of removing (by setting ) gives a different reduced form whose associated Lie algebra is conjugated to the one we just found. We refer to DW (19) for the computations in that other path. We also remark that two of the matrices that could not be removed from are “absorbed” as lower triangular parts of matrices coming from . It is -dimensional, whereas the Lie algebra associated to the original matrix had dimension . This shows that the Picard-Vessiot extension is obtained from the Picard-Vessiot extension for by adding only one integral and the system has indeed been transformed into a form where solving is much simpler than before - and we also have proofs of transcendence properties for the remaining objects.
References
- ABKM (18) Y. Abdelaziz, S. Boukraa, C. Koutschan, and J.-M. Maillard, Diagonals of rational functions, pullbacked hypergeometric functions and modular forms, J. Phys. A 51 (2018), no. 45, 455201, 30.
- ABKM (20) , Heun functions and diagonals of rational functions, J. Phys. A 53 (2020), no. 7, 075206, 24.
- ABRS (14) J. Ablinger, J. Blümlein, C. G. Raab, and C. Schneider, Iterated binomial sums and their associated iterated integrals, J. Math. Phys. 55 (2014), no. 11, 112301, 57.
- AMCW (13) A. Aparicio-Monforte, É. Compoint, and J.-A. Weil, A characterization of reduced forms of linear differential systems, Journal of Pure and Applied Algebra 217 (2013), no. 8, 1504–1516.
- AMDW (16) A. Aparicio-Monforte, T. Dreyfus, and J.-A. Weil, Liouville integrability: an effective Morales–Ramis–Simó theorem, Journal of Symbolic Computation 74 (2016), 537 – 560.
- AMW (11) A. Aparicio-Monforte and J.-A. Weil, A reduction method for higher order variational equations of Hamiltonian systems, Symmetries and Related Topics in Differential and Difference Equations, Contemporary Mathematics, vol. 549, Amer. Math. Soc., Providence, RI, September 2011, pp. 1–15.
- AMW (12) A. Aparicio-Monforte and J.-A. Weil, A reduced form for linear differential systems and its application to integrability of Hamiltonian systems, Journal of Symbolic Computation 47 (2012), no. 2, 192 – 213.
- AvH (97) S. A. Abramov and M. van Hoeij, A method for the integration of solutions of Ore equations, Proceedings of the 1997 International Symposium on Symbolic and Algebraic Computation (Kihei, HI), ACM, New York, 1997, pp. 172–175.
- AvH (99) , Integration of solutions of linear functional equations, Integral Transform. Spec. Funct. 8 (1999), no. 1-2, 3–12.
- AZ (10) M. Ayoul and Nguyen Tien Zung, Galoisian obstructions to non-Hamiltonian integrability, C. R. Math. Acad. Sci. Paris 348 (2010), no. 23-24, 1323–1326.
- Bar (99) M. A. Barkatou, On rational solutions of systems of linear differential equations, J. Symbolic Comput. 28 (1999), no. 4-5, 547–567.
- Bax (82) R. J. Baxter, Exactly solved models in statistical mechanics, Academic Press Inc. [Harcourt Brace Jovanovich Publishers], London, 1982.
- BBH+ (09) A. Bostan, S. Boukraa, S. Hassani, J.-M. Maillard, J.-A. Weil, and N. Zenine, Globally nilpotent differential operators and the square Ising model, J. Phys. A 42 (2009), no. 12, 125206, 50.
- BBH+ (10) A. Bostan, S. Boukraa, S. Hassani, J.-M. Maillard, J.-A. Weil, N. Zenine, and N. Abarenkova, Renormalization, isogenies, and rational symmetries of differential equations, Adv. Math. Phys. (2010), 44p.
- BBH+ (11) A. Bostan, S. Boukraa, S. Hassani, M. van Hoeij, J.-M. Maillard, J.-A. Weil, and N. Zenine, The Ising model: from elliptic curves to modular forms and Calabi-Yau equations, J. Phys. A: Math. Theor. 44 (2011), no. 4, 045204, 44.
- BCDVW (20) M. Barkatou, T. Cluzeau, L. Di Vizio, and J.-A. Weil, Reduced forms of linear differential systems and the intrinsic Galois-Lie algebra of Katz, SIGMA Symmetry Integrability Geom. Methods Appl. 16 (2020), Paper No. 054, 13.
- BCEBW (12) M. A. Barkatou, T. Cluzeau, C. El Bacha, and J.-A. Weil, Computing closed form solutions of integrable connections, Proceedings of the 36th international symposium on Symbolic and algebraic computation (New York, NY, USA), ISSAC ’12, ACM, 2012.
- BCWDV (16) M. Barkatou, T. Cluzeau, J.-A. Weil, and L. Di Vizio, Computing the lie algebra of the differential galois group of a linear differential system, Proceedings of the ACM on International Symposium on Symbolic and Algebraic Computation, 2016, pp. 63–70.
- Ber (01) D. Bertrand, Unipotent radicals of differential Galois group and integrals of solutions of inhomogeneous equations, Math. Ann. 321 (2001), no. 3, 645–666.
- BHM+ (07) S. Boukraa, S. Hassani, J.-M. Maillard, B. M. McCoy, J.-A. Weil, and N. Zenine, Fuchs versus Painlevé, J. Phys. A 40 (2007), no. 42, 12589–12605.
- BHMW (14) S. Boukraa, S. Hassani, J.-M. Maillard, and J.-A. Weil, Differential algebra on lattice green and calabi-yau operators, J. Phys. A: Math. Theor. 47 (2014), no. 9, 095203.
- BHMW (15) S. Boukraa, S. Hassani, J.-M. Maillard, and J.-A. Weil, Canonical decomposition of irreducible linear differential operators with symplectic or orthogonal differential galois groups, Journal of Physics A: Mathematical and Theoretical 48 (2015), no. 10, 105202.
- BKK (10) V. V. Bytev, M. Yu. Kalmykov, and B. A. Kniehl, Differential reduction of generalized hypergeometric functions from Feynman diagrams: one-variable case, Nuclear Phys. B 836 (2010), no. 3, 129–170.
- BS (99) P. H. Berman and M. F. Singer, Calculating the Galois group of , completely reducible operators, J. Pure Appl. Algebra 139 (1999), no. 1-3, 3–23, Effective methods in algebraic geometry (Saint-Malo, 1998).
- BSMR (10) D. Blázquez-Sanz and J.-J. Morales-Ruiz, Differential Galois theory of algebraic Lie-Vessiot systems, Differential algebra, complex analysis and orthogonal polynomials, Contemp. Math., vol. 509, Amer. Math. Soc., Providence, RI, 2010, pp. 1–58.
- BW (03) D. Boucher and J.-A. Weil, Application of J.-J. Morales and J.-P. Ramis’ theorem to test the non-complete integrability of the planar three-body problem, From combinatorics to dynamical systems, IRMA Lect. Math. Theor. Phys., vol. 3, de Gruyter, Berlin, 2003, pp. 163–177.
- CH (11) T. Crespo and Z. Hajto, Algebraic groups and differential Galois theory, Graduate Studies in Mathematics, vol. 122, American Mathematical Society, Providence, RI, 2011.
- Com (12) T. Combot, Non-integrability of the equal mass -body problem with non-zero angular momentum, Celestial Mech. Dynam. Astronom. 114 (2012), no. 4, 319–340.
- CR (91) R. C. Churchill and D. L. Rod, On the determination of Ziglin monodromy groups, SIAM J. Math. Anal. 22 (1991), no. 6, 1790–1802.
- CW (18) G. Casale and J.-A. Weil, Galoisian methods for testing irreducibility of order two nonlinear differential equations, Pacific J. Math. 297 (2018), no. 2, 299–337.
- DW (19) T. Dreyfus and J.-A. Weil, Computing the Lie algebra of the differential Galois group: the reducible case, ArXiv 1904.07925 (2019).
- DW (20) , Maple worksheet with the examples for this paper: http://www.unilim.fr/pages_perso/jacques-arthur.weil/DreyfusWeilReductionExamples.mw, 2020.
- FdG (07) C. Fieker and W. A. de Graaf, Finding integral linear dependencies of algebraic numbers and algebraic Lie algebras, LMS J. Comput. Math. 10 (2007), 271–287.
- HKMZ (16) S. Hassani, Ch. Koutschan, J.-M. Maillard, and N. Zenine, Lattice Green functions: the -dimensional face-centered cubic lattice, , J. Phys. A 49 (2016), no. 16, 164003, 30.
- Kal (06) M. Yu. Kalmykov, Gauss hypergeometric function: reduction, -expansion for integer/half-integer parameters and Feynman diagrams, J. High Energy Phys. (2006), no. 4, 056, 21.
- KK (11) M. Yu. Kalmykov and B. A. Kniehl, Counting master integrals: integration by parts vs. differential reduction, Phys. Lett. B 702 (2011), no. 4, 268–271.
- KK (12) , Mellin-Barnes representations of Feynman diagrams, linear systems of differential equations, and polynomial solutions, Phys. Lett. B 714 (2012), no. 1, 103–109.
- KK (17) , Counting the number of master integrals for sunrise diagrams via the Mellin-Barnes representation, J. High Energy Phys. (2017), no. 7, 031, front matter+27.
- Kou (13) C. Koutschan, Lattice Green functions of the higher-dimensional face-centered cubic lattices, J. Phys. A 46 (2013), no. 12, 125005, 14.
- Kov (86) J. J. Kovacic, An algorithm for solving second order linear homogeneous differential equations, J. Symbolic Comput. 2 (1986), no. 1, 3–43.
- LSS (18) R. N. Lee, A. V. Smirnov, and V. A. Smirnov, Solving differential equations for Feynman integrals by expansions near singular points, J. High Energy Phys. (2018), no. 3, 008, front matter+14.
- MAB+ (11) B. M. McCoy, M. Assis, S. Boukraa, S. Hassani, J.-M. Maillard, W. P. Orrick, and N. Zenine, The saga of the Ising susceptibility, New trends in quantum integrable systems, World Sci. Publ., Hackensack, NJ, 2011, pp. 287–306.
- Mag (54) W. Magnus, On the exponential solution of differential equations for a linear operator, Comm. Pure Appl. Math. 7 (1954), 649–673.
- McC (10) B. M. McCoy, Advanced statistical mechanics, International Series of Monographs on Physics, vol. 146, Oxford University Press, Oxford, 2010.
- MM (12) B. M. McCoy and J-M. Maillard, The importance of the ising model, Prog. Theor. Phys. 127 (2012), 791-817, 2012.
- Mot (18) O. V. Motygin, On evaluation of the confluent heun functions, 2018.
- MR (20) J.-J. Morales-Ruiz, A differential Galois approach to path integrals, J. Math. Phys. 61 (2020), no. 5, 052103, 12.
- MRR (01) J.-J. Morales-Ruiz and J.-P. Ramis, Galoisian obstructions to integrability of Hamiltonian systems. I, II, Methods Appl. Anal. 8 (2001), no. 1, 33–95, 97–111.
- MRRS (07) J.-J. Morales-Ruiz, J.-P. Ramis, and C. Simo, Integrability of Hamiltonian systems and differential Galois groups of higher variational equations, Ann. Sci. École Norm. Sup. (4) 40 (2007), no. 6, 845–884.
- MRS (09) J.-J. Morales-Ruiz and S. Simon, On the meromorphic non-integrability of some -body problems, Discrete Contin. Dyn. Syst. 24 (2009), no. 4, 1225–1273.
- MRSS (05) J.-J. Morales-Ruiz, C. Simó, and S. Simon, Algebraic proof of the non-integrability of Hill’s problem, Ergodic Theory Dynam. Systems 25 (2005), no. 4, 1237–1256.
- MS (02) C. Mitschi and M. F. Singer, Solvable-by-finite groups as differential Galois groups, Ann. Fac. Sci. Toulouse Math. (6) 11 (2002), no. 3, 403–423.
- PPR+ (10) O. Pujol, J.-P. Pérez, J.-P. Ramis, C. Simó, S. Simon, and J.-A. Weil, Swinging Atwood Machine: experimental and numerical results, and a theoretical study, Physica D: Nonlinear Phenomena 239 (2010), no. 12, 1067–1081.
- PS (03) M. van der Put and M. F. Singer, Galois theory of linear differential equations, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 328, Springer-Verlag, Berlin, 2003.
- Sal (14) V. Salnikov, Effective algorithm of analysis of integrability via the Ziglin’s method, Journal of Dynamical and Control Systems 20 (2014), no. 4, 465–474 (English).
- Sin (91) M. F. Singer, Liouvillian solutions of linear differential equations with Liouvillian coefficients, J. Symbolic Comput. 11 (1991), no. 3, 251–273.
- Sin (09) , Introduction to the Galois theory of linear differential equations, Algebraic theory of differential equations, London Math. Soc. Lecture Note Ser., vol. 357, Cambridge Univ. Press, Cambridge, 2009, pp. 1–82.
- Smi (12) V. A. Smirnov, Analytic tools for Feynman integrals, Springer Tracts in Modern Physics, vol. 250, Springer, Heidelberg, 2012.
- ST (16) T. M. Sadykov and S. Tanabé, Maximally reducible monodromy of bivariate hypergeometric systems, Izv. Ross. Akad. Nauk Ser. Mat. 80 (2016), no. 1, 235–280.
- Tsy (01) A. Tsygvintsev, The meromorphic non-integrability of the three-body problem, J. Reine Angew. Math. 537 (2001), 127–149.
- UW (96) F. Ulmer and J.-A. Weil, Note on Kovacic’s algorithm, J. Symbolic Comput. 22 (1996), no. 2, 179–200.
- vHW (05) M. van Hoeij and J.-A. Weil, Solving second order differential equations with Klein’s theorem, ISSAC 2005 (Beijing), ACM, New York, 2005.
- WN (63) J. Wei and E. Norman, Lie algebraic solution of linear differential equations, J. Mathematical Phys. 4 (1963), 575–581.
- WN (64) J. Wei and E. Norman, On global representations of the solutions of linear differential equations as a product of exponentials, Proc. Amer. Math. Soc. 15 (1964), 327–334.
- Zig (82) S. L. Ziglin, Branching of solutions and nonexistence of first integrals in hamiltonian mechanics. I, Functional Analysis and Its Applications 16 (1982), no. 3, 181–189.