Department of Computer Science, University of Oxford, United Kingdom
11email: bertrand.teguia@cs.ox.ac.uk
Arithmetic of D-Algebraic Functions
Abstract
We are concerned with the arithmetic of solutions to ordinary or partial nonlinear differential equations which are algebraic in the indeterminates and their derivatives. We call these solutions D-algebraic functions, and their equations are algebraic (ordinary or partial) differential equations (ADEs). The general purpose is to find ADEs whose solutions contain specified rational expressions of solutions to given ADEs. For univariate D-algebraic functions, we show how to derive an ADE of smallest possible order. In the multivariate case, we introduce a general algorithm for these computations and derive conclusions on the order bound of the resulting algebraic PDE. Using our accompanying Maple software, we discuss applications in physics, statistics, and symbolic integration.
Keywords:
Gröbner bases triangular set differential algebra symbolic computation ranking1 Introduction
Let be a field of characteristic zero, and a differential field. We denote by , the differential polynomial ring of the differential indeterminate . Elements of that ring are called differential polynomials. Differential ideals are ideals that are closed under taking derivatives.
Definition 1
A function is called differentially algebraic, or simply D-algebraic, in if there exists a differential polynomial such that , i.e., .
Therefore D-algebraic functions can be seen as zeros of differential polynomials. We will often use this terminology. What we call functions are always elements of some -algebra which is closed under the derivation used; they could be analytic functions, formal power series, or distributions in the appropriate domains. They are mainly differentiable objects that are suitable for multiplications by the indeterminates.
Example 1 (Univariate D-algebraic)
The Weierstrass function is D-algebraic since it is a zero of , with .
The equation resulting from equating a differential polynomial to zero is called an algebraic differential equation (ADE). For a given differential polynomial, such an equation will be called associated ADE. Similarly, given an ADE, the differential polynomial obtained by removing the equality to zero is called its associated differential polynomial. The order of a differential polynomial is the one of its associated ADE, and its total degree, which we simply call degree, is the total degree of the differential polynomial seen as a multivariate polynomial over .
Definition 1 generalizes to the multivariate case with the partial differential ring , , and , . In that case, a D-algebraic function is seen as a zero of a partial differential polynomial in , where is a tuple with components of .
Example 2 (Multivariate D-algebraic)
The so-called harmonic functions are D-algebraic since they are zeros of the partial differential polynomial
(1) |
associated to the Laplace equation .
Note that the notion of D-algebraicity used here was already defined in [25, Chapter IV]; it relaxes the definitions given in [11, 35] as it does not require D-algebraicity in each independent variable. We will manipulate D-algebraic functions as generic zeros [25, Chapter IV.2] of the general component of some differential ideal defined from the ADEs under consideration. An essential ingredient for the correctness of our algorithms is the notion of triangular sets, which is discussed in detail in [18].
The set of D-algebraic functions is perhaps the most general structural class of functions that encompasses most functions encountered in the sciences. One prominent area where they naturally occur is enumerative combinatorics where D-algebraic functions are regarded as generating functions of counting problems. Recently in [35], Bernardi, Bousquet-Mélou, and Raschel classified some D-algebraic quadrant models that are not differentially finite (D-finite), i.e., the generating functions of the corresponding counting problems are D-algebraic but their derivatives do not span a finite dimensional vector space (see [40, 22]). This motivates to explore beyond D-finiteness, a structure that interacts favorably with linear algebra techniques, unlike D-algebraic functions. It is thus important to study the effectiveness of the closure properties of D-algebraic functions. These functions form a field, and are closed under many other operations like composition and differentiation [1] [35, Section 6] [32]. The theory behind D-algebraicity began earlier in the last century, see for instance [32, 30, 36], [25, Chapter II], [10], and [28]. However, these studies primarily focus on theoretical aspects regarding operations that may or may not preserve D-algebraicity, rather than their practical effectiveness. One had to wait until the end of the last century, when algorithms like the Rosendfeld-Gröbner algorithm for representing radical differential ideals appeared [6]. This algorithm is implemented in the DifferentialAlgebra package [5] in Maple. Of a similar flavor is the Thomas decomposition algorithm which views the generators of radical differential ideals as the “simple systems” of a system of algebraic PDEs [2] [37, Chapter II]. It is available in the DifferentialThomas package of Maple.
Both algorithms may serve as alternatives to our computations in certain contexts. We also mention the connection between ADEs and biochemical reaction networks which arise from particular types of dynamical systems [17, 12]. As we will see later, the theory developed in [17] is essential to our algorithmic approach. Other interesting topics related to D-algebraic functions are that of differentially definable functions [19], and solutions to ADEs of fixed degrees [44]. We are confident that our results will enable certain progress around the study of these D-algebraic sub-classes. There is a method for constructing power series that are not D-algebraic, as discussed in [32] and [28]. The idea is based on gaps between non-zeros coefficients of the series. One well-known example of such power series is [30]. While these specific series are not the focus of our paper, it is worth noting that a similar strategy, as outlined in [42], adapted to higher degree ADEs in conjunction with our findings, may assist in deciding the D-algebraicity of series like [28, 29].
To our knowledge, this paper and the work in [1] are the only ones that are entirely dedicated to effective computations of the closure properties of D-algebraic functions. With regards to power series solutions, effective results appeared in [11]. Like in [1], compared to that work of van der Hoeven, which for the zero-equivalence problem requires investigations on the initial conditions, in our case we focus on what he called “global computations”, which essentially rely on the framework of differential algebra and computational commutative algebra. The reason is that instead of dealing with a particular solution of an ADE, we deal with all solutions of that ADE for which the arithmetic operation performed is meaningful. Thus zero-testing and initial conditions are neglected, though our results may apply to them.
This paper begins with the arithmetic of univariate D-algebraic functions. Based on the construction of systems of ordinary differential equations (ODEs), we show how to compute an ADE of order at most the sum of the orders of any given ADEs, satisfied by some rational expression of the solutions of those given ADEs. In Section 3, we develop an algorithm for the arithmetic of multivariate D-algebraic functions and prove that the order of the resulting algebraic PDEs is not always bounded by the sum of the orders of the given algebraic PDEs. Section 4 presents applications using our Maple implementation of Algorithm 1 and Algorithm 2. We mention that all our results also work for differential rational expressions of D-algebraic functions due to the natural action of differentiation for D-algebraic functions.
Some background and terminology
Let us now assume the univariate setting. To define our target ODE system, we use the multivariate differential polynomial ring , where for some . Our main interest is in systems of the form
where . In the context of differential algebra, we consider the following polynomials
(2) |
where is the least common multiple of the denominators in , and . With an appropriate monomial ordering over the differential polynomial ring , multiplying by in (2) is equivalent to clearing denominators and considering as the least common multiple of the initials of the resulting differential polynomials. We make the above choice for convenience as it allows us to defer the selection of a monomial ordering to a later stage, which also appears more suitable for the algorithmic part (see also [12]). Next, we consider the differential ideal
(3) |
where denotes the saturation with . We define as the algebraic ideal containing the polynomials in (2) and their first derivatives of order at most saturated with . The system may be seen as a single-output state-space model without input. General setting and computations of such models are discussed in [17, 34]. There it is also explained how so-called input-output equations of , i.e. ADEs that relate the output variable and the input variable (usually denoted ), are deduced from the elimination ideal
Hence in our setting, the input-output equations correspond to ADEs satisfied by the rational function .
Remark 1
Recall that the separant of a differential polynomial is its derivative with respect to its highest order term. Thus for all positive integers , the separant of is , for . Therefore saturation by suffices to target the generic solutions of .
A system of the type of is called rational dynamical system of dimension (the dimension of ). In their work [33], Pavlov and Pogudin developed an algorithmic approach to derive such systems from first-order ADEs of two dependent variables, which represent the output and the input. The relationship between these systems and the closure properties of D-algebraic functions is discussed in [1]. Specifically, when the dynamical system models ADEs fulfilled by some D-algebraic functions , such that in represents a rational expression in the dependent and independent variables, the corresponding input-output equations are ADEs satisfied by . In particular, thanks to this construction and the non-trivial nature of the elimination ideal that generates the input-output equations of , we can reestablish that the set of D-algebraic functions is a field.
Definition 2
A differential polynomial of order is called linear in its highest order term (l.h.o.) if, in the expression of , appears only linearly.
Example 3
The differential polynomials and are l.h.o., whereas is not l.h.o. But its derivative is l.h.o. In general, the derivative of a differential polynomial is always l.h.o.
We will say that a computed algebraic (partial or ordinary) differential equation is of the least order if its order is the minimal order differential polynomial in the corresponding differential ideal. The arithmetic of two D-algebraic functions in the l.h.o. and non-l.h.o. cases were treated in [1]. In this paper, we improve the computational bound given in that paper for non-l.h.o. algebraic ODEs. We demonstrate that the bound established for l.h.o. ADEs also applies to non-l.h.o. ADEs. In essence, the sum of the orders of some given ADEs serves as an upper bound for the minimal order of an ADE satisfied by rational expressions formed from generic solutions of those ADEs.
2 Modeling the arithmetic of univariate D-algebraic functions
Throughout this section, we consider D-algebraic functions , and , where . We assume that is a zero of the differential polynomial , of order . The main goal of this section is to complement the work in [1]. As we will see in Example 4, using the implementation of Algorithm 1, one can compute least-order ADEs in all situations.
2.1 The l.h.o. case
We here recall the result from [1] for the arithmetic of D-algebraic functions that are zeros of l.h.o. differential polynomials. Since each is l.h.o, we can write its highest derivative of the dependent variable in terms of the lower ones from the associated ADE. We write
(4) |
Let us now construct the ODE system (here a dynamical system) that models the operation .
Define the indeterminates
(5) |
This implies that
(6) |
Furthermore, let
(7) |
Set . The following dimensional dynamical system models :
We denote the corresponding differential ideal by . The above construction sustains the following proposition, which extends [1, Proposition 4.9.] for arbitrarily many differential polynomials.
Proposition 1
We use the above notation. There is an algorithm that computes an ADE of order at most , fulfilled by the D-algebraic function .
Proof
Considering a lexicographic monomial ordering that rank the differential indeterminate higher than the other indeterminates, the proof relies on the fact that the elimination ideal
(8) |
is not trivial, and contains the least-order differential polynomial that has as a zero (see [17, Corollary 3.21] and [34, Prop. 1.27]). Computationally, among the generators (Gröbner basis) of the elimination ideal, we select a differential polynomial of the lowest degree among those of the smallest order.
2.2 The non-l.h.o. case
Suppose there are differential polynomials that are non-l.h.o. among the ’s, . In [1] this situation is overcome by replacing those differential polynomials with their first derivatives. This brings the problem back to the l.h.o. case, but increase the bound for the order of the differential equation sought by . The main issue is that without taking derivatives, we do not easily relate the arithmetic in this case to a rational dynamical system.
Our main observation is that, provided some algorithmic changes and a little theoretical adaptation, the arguments over which the computations from the dynamical systems rely are still valid. Indeed, suppose that , is a non-l.h.o. differential polynomial, and let be the degree of in . We can rewrite as
(9) |
where , such that . The coefficient is often called the initial of . We make the same change of differential indeterminates as in Section 2.1 and build an ODE system of the form
where . The rational functions result from solving for , i.e., with the same variable names in the expressions of and , we can write
Observe that the denominators are free of the indeterminate representing . This allows us to define in as in by using all the denominators of the system. To ease the theoretical understanding, let us simplify the notation by rewriting in the following abstract form
(10) |
where
-
•
, is either or one of the in ;
-
•
(with numerator ) is either or the part of , that is free of in , ;
-
•
(with numerator ) is either or the part of that contains (in its numerator) in , ;
-
•
with numerator .
We call a system of the form of (10) a radical-rational dynamical system. Note that the shape of imposes certain constraints on the radical-rational dynamical systems we use. In the context of differential algebra, we consider the differential ideal
(11) |
where is the least common multiple of and the separants of . These separants are explicitely given by
where derivations are taken component wise.
Let us now present our algorithmic steps for computing ADEs satisfied by rational expressions of univariate D-algebraic functions.
-
1.
Construct from the input ADEs as in (10).
-
2.
Denote by the set of polynomials
in the polynomial ring .
-
3.
Compute the first derivatives (w.r.t. ) of all polynomials in and add them to .
-
4.
Compute the th derivative of and add it to . We are now in the ring .
-
5.
Let be the ideal generated by the elements of .
-
6.
If some of the input are not l.h.o., let be the least common multiple of and the separants of . Otherwise .
-
7.
Update by its saturation with , i.e, .
-
8.
Compute the elimination ideal . From the Gröbner basis, choose a polynomial of the lowest degree among those of the lowest order.
-
9.
Return (or its writing with ).
To prove the correctness of Algorithm 1, we must show that the elimination ideal in item 8 is non-trivial. This fact is established by the following theorem.
Theorem 2.1
On the commutative ring , seen as a polynomial ring in infinitely many variables, consider the lexicographic monomial ordering corresponding to any ordering on the variables such that
-
(i.)
for all ,
-
(ii.)
and for all .
Then the set is a triangular set with respect to this ordering. Moreover,
(12) |
Proof
The leading monomials of and in the ring (with derivation seen in ) have highest variables and , respectively. Since these variables are all distinct, by definition (see [18, Definition 4.1]), we deduce that is a consistent triangular set. We shall see the coefficients in the field . As a triangular set, defines the ideal . Therefore by [18, Theorem 4.4]) all associated primes of share the same transcendence basis given by the non-leading variables in . Thus the transcendence degree of over is . However, the transcendence degree of is . Hence we must have .
From the proof of Theorem 2.1, one should observe that is the minimal integer for which the used arguments hold. This relates to the minimality of the order of the ADE obtained after the computations. Moreover, the l.h.o and the non-l.h.o cases are unified in (10). We make explicit to illustrate the difference with the construction of in Section 2.1.
The main point behind Algorithm 1 is the systematic construction of ADEs for rational expressions of D-algebraic functions using ODE systems (or triangular sets). Note that the saturation with respect to the least common multiple of the denominators and the separants implies that the D-algebraic functions we deal with are the generic solutions of ADEs.
2.3 Implementation
We have implemented Algorithm 1 in the NLDE package (see [43, 41]). It complements the implementation for arithmetic operations from [1]. For full details on the syntax, we refer the reader to the documentation at MathRepo NLDE Doc (see also [41]). We here present examples to illustrate what can be seen as an improvement of the implementation in [1]. The commands for arithmetic operations are NLDE:-arithmeticDalg and NLDE:-unaryDalg. The former is for several input ADEs, and the latter is for a single input ADE. By default, NLDE:-arithmeticDalg follows the second method in [1], which implies that in the non-l.h.o. case, the order of the returned ADE may be higher than expected. When some of the input ADEs are not l.h.o., one specifies lho=false to use Algorithm 1 and obtain an ADE of the smallest order possible. Regarding Gröbner bases elimination, by default,
-
•
when lho=true, we use the pure lexicographic ordering plex from the Maple package Groebner; this is the same ordering used in Theorem 2.1, which guarantees the desired result.
-
•
When lho=false, we use the lexdeg elimination order of Bayer and Stillman [4] implemented by the EliminationIdeal command of the Maple package PolynomialIdeals. This is a block order where blocks are ordered with plex. In our case, there are two blocks: the variables of the elimination ideal constitute one block, and the remaining variables constitute the other one. Both blocks are internally ordered with a suitable order like degrevlex for instance. Although this ordering generally returns the desired result, it can also fail to find it sometimes. Nevertheless, lexdeg tends to provide a better efficiency. To use plex when lho=false, one further specifies lhoplex=true.
To use the lexdeg ordering when lho=true, one specifies ordering=lexdeg. For unary operations, NLDE:-unaryDalg was updated to only implement the method of this paper.
Example 4
Consider the ADE
(13) |
Its solutions are and for arbitrary such that . As a second ADE, we take the ODE of the exponential function
(14) |
We want to compute ADEs for the sum of solutions of these two ADEs. Using Method II from [1], one gets the output:
(15) |
Since (13) is not l.h.o., its first derivative is used; which explains why the output ADE is of order . Let us now use the method of this paper. The corresponding ODE system is
(16) |
Thus . We use our package as follows:
(17) |
We obtain a second-order ADE satisfied by sums of all generic solutions to (13) and (14). Notice that is not a solution of (15) and (17). As explained in [1], in general, the saturation often neglects polynomial solutions of degrees less than the order of non-l.h.o. ADEs; this applies to (15). Similarly, for (17), saturation at the separants eliminates those same polynomials.
On the other hand, our implementation provides an option for the user to avoid saturation by the separants. The optional argument is separantsZeros=true, which is false by default. The option is exploratory and does not rely on Theorem 2.1 when separantsZeros=true. In this case, we have no guarantee that the corresponding elimination ideal is non-trivial. Nevertheless, we have not yet found an example where this does not happen.
(18) |
The particular solutions and , are zeros of the factors and , respectively.
Example 5
Let us take the third ADE:
(19) |
Its solutions are , , where is an arbitrary constant, and the ’s are roots of the polynomial . We want to find an ADE for , where and are solutions of (13), (14), and (19), respectively.
The corresponding ODE system is
(20) |
Thus . The implementation of Algorithm 1 yields:
(21) |
For this example, [1, Method II] gives the same output. The latter is one of the three factors of the ADE obtained with the option separantsZeros=true. The other two factors are the following:
(22) |
These two factors vanish at , , .
Note that either of the implementations of the Rosendfeld-Gröbner algorithm and the Thomas decomposition could not find (21) within a reasonable time. The former runs out of memory, and the latter keeps running even after an hour of computations. They tend to behave this way for ADEs of higher degrees. We also mention that even in the situation where one of these algorithms would return a correct result, the exponent in (22) could not be found since those algorithms output generators of a radical differential ideal, unlike our algorithm which computes in the corresponding differential ideal. It would be interesting to know what these exponents mean.
3 Arithmetic of multivariate D-algebraic functions
We now focus on solutions of algebraic PDEs. We aim to present an algorithm in the spirit of Method I in [1]. For simplicity, we start with bivariate D-algebraic functions. Thus the functions are now bivariate in , and each is a zero of the partial differential polynomial of order w.r.t. , and of order w.r.t. .
3.1 Ordering on the semigroup of derivative operators
We are going to introduce a total ordering on the set of derivative operators with the aim of using them in increasing order up to a given limit. Algorithmically, this will entail to defining a map that sends every non-negative integer to a derivative operator. In [25] and many references in differential algebra (see also [15]), the notion of order for partial differential rings is different from the one we use. The prevalent definition does not easily adapt to our computational goal. Indeed, given the set of derivations , one considers the free multiplicative semigroup generated by the elements of . To any derivative operator
one defines the order . The derivative of , is called the derivative of of order (supposedly w.r.t. ). The order of a differential polynomial (also called -polynomial ring) is then the maximum such that occurs in . What we wish to add in this definition is a total ordering (or ranking) of the derivative operators in . Observe that for , there is a natural ranking of the derivative operators:
allowing to uniquely associate a non-negative integer with the order of a particular derivative. With the above definition of order for bivariate partial differential polynomial rings, there are two possible choices for the derivative of of order , namely, and . We combine total order and classical orderings (lexicographic or reverse-lexicographic).
The total-order is used to ensure the possibility to move from to by some applications of the chosen derivation rule. In fact, the bijective map between and should be “obvious” by construction. Thus what we call order for is either the non-negative integer whose image (or preimage) is through that bijection, or simply , assuming is well-ordered (might be implicit for non-computational purposes). We will use the latter, and define a derivation rule based on a bijective map between and .
Our idea is to use a “very” classical map in set theory. Consider the following start of a ranking for the application of derivative operators
This corresponds to the following ranking of derivative orders
One can associate each ordered pair of integers with a unique positive integer. Indeed, the underlying map is the so-called Cantor pairing function, which is a bijection from to . The underlying ordering of the ordered pairs is the graded (or degree) lexicographic ordering; however, to avoid confusion with the lexicographic ordering that we will use for monomials, we use the terminology of Cantor pairing function or Cantor -tuple functions for arbitrarily many independent variables. As proven by Rudolf Fueter and George Pólya [16], the Cantor pairing function is the only quadratic polynomial that maps to . Therefore for two variables, it appears as the most “natural and efficient” choice for processing speed on a regular computer.
We see that such a ranking is inherent to the definition of a total order on to rank the variable in . Note that this ranking is in line with the terminology introduced by Ritt and should not be confused with monomial orderings. Let be the Cantor pairing function, its functional inverse such that . For , the th derivation, denoted , is:
(23) |
We view as
This implicitly defines an operator which we denote by .
Remark 2
-
•
For all , , and therefore satisfies the linearity and the Leibniz rules.
-
•
On the operator side, may wrongly be seen as the th power of the operator . Although commutativity is induced by the semigroup , exponentiation does not follow. In general, for ,
(24) The equality is easy to establish since for , and , we have
Just to give an example, let . Then we have
(25) We will consider the th derivation w.r.t. as the application of , and not successive applications of .
We introduced primarily to avoid continuous referencing to the chosen ordering. This will significantly simplify the arguments in the upcoming sections. Needless to say, one can easily avoid explicit links to this operator.
3.2 Arithmetic of bivariate D-algebraic functions
Observe that, unlike other definitions of multivariate D-algebraic functions that restrict their study to D-algebraic functions in each independent variable (see [11, Section 5.2],[35, Section 6]), we here have a more general notion. Indeed, a multivariate function can be D-algebraic without being D-algebraic in some of its variables. One can derive proofs for the closure properties by providing bounds for the transcendence degree, as shown in [11, Section 5.2]. Note, however, that the bounds may not be as sharp as in the ordinary case.
Example 6
The incomplete function defined as
(26) |
where denotes the confluent hypergeometric function, is D-algebraic, and satisfies the linear ADE:
(27) |
Getting back to our algorithmic goal, we can now write every differential polynomial in in terms of derivations.
Example 7
Since every is the result of , for , we can write , replacing by and omitting the superscript. One can even remove and the subscript if there is no ambiguity for the variables involved (and get ‘polynomials’ in ).
(28) | |||
(29) |
Remark 3
-
•
.
-
•
.
Proposition 2 (Composition property)
For all non-negative integers , and a differential indeterminate , we have
(30) |
where addition for ordered pairs is taken component wise.
Proof
Let be the result of . Then according to (23),
To illustrate our use of , let us take an explanatory example.
Example 8
Consider
With , we have
(31) | |||
(32) |
Thus w.r.t. , and are of order and , respectively. Suppose we want to find an algebraic PDE for the sum , where is a zero of . Our knowledge of the ordinary case suggests that we look for an ADE of order at most in and in . This is the same as looking for an ADE of -order at most . We define (representing ) as the dependent variable for the targeted ADE, and consider the differential polynomial
We take the first -derivatives of and remove those that involve derivatives of orders greater than w.r.t. (first component) and greater than w.r.t. (second component). We get
(33) |
We now need to take some -derivatives of and . The idea is to make the highest -derivatives of and from (33) occur. Here we wish to have among the derivatives. Our bound for the number of derivations to be taken is . For this example, it turns out that we only need to take the first -derivatives. For , we obtain the differential polynomials
(34) |
And for we obtain
(35) |
We write these differential polynomials with their dependent variables and not in terms of -derivations. This is a preliminary step for the Gröbner basis elimination that will follow. Note that these computations can be performed without any explicit use of and , or any derivative operator in . Indeed, we have the natural isomorphism
(36) |
by the correspondence between -derivations and elements of .
In a similar way as defined in Theorem 2.1, we consider a lexicographic monomial ordering such that
-
(I.)
for all ,
-
(II.)
and for all .
We want to eliminate the lower variables . We compute the elimination ideal
(37) |
where is the ideal containing , , , and their derivatives from (34), (35), and (33). This may be viewed as a ramification of a truncation of the corresponding differential ideal in with the ranking imposed by ; however, the truncation would be of -order . We obtain the principal ideal
(38) |
Therefore the sum of zeros of and are solutions of the associated linear algebraic PDE given by
(39) |
In Example 8, we obtained an ADE whose order is the sum of the orders of and w.r.t. each independent variable. A natural question that we might want to answer is to know whether, like in the ordinary case, the sum of orders of given differential polynomials constitutes a bound for the order of ADEs fulfilled by functions defined as arithmetic expressions of zeros of those polynomials. We will show that this does not hold in general. We give another example to that aim, a tiny modification of Example 8.
Example 9 (The orders do not add up)
Let
(40) |
As previously, we can expect to find a partial differential polynomial of order at most in and in for the sum of zeros of and . However, for this example, the elimination ideal is trivial, even when we do the elimination step with the truncation of -order of the corresponding differential ideal.
We now look for an algebraic PDE of order in the first component and order in the second. This corresponds to the -order . For the polynomials and , we take their first -derivatives and proceed as before. Then we compute the elimination ideal
(41) |
and obtain a principal ideal whose defining polynomial is given by
(42) |
The notation denotes all -derivatives of of (-)order at most .
Hence for this second example, we obtained an ADE whose order (according to any of the definitions mentioned in Section 3.1) is not the sum of the orders of and from (40).
To conclude that this is a general fact, we need to prove that the result of this example does not change if we consider higher -order truncations of the corresponding differential ideal. Prior to that, we state a useful lemma.
Lemma 1 (Increasing property)
Let be a differential indeterminate. Consider the -ranking in . For all non-negative integers , and , we have
(43) |
In other words, the application of derivations preserves the ranking.
Proof
Suppose . By construction of the -ranking, this is equivalent to
We have and . Therefore
Hence
We can now state our theorem.
Theorem 3.1
We use the same notations of this section. For all , , the order of the least-order algebraic PDE fulfilled by the bivariate D-algebraic function given by is not bounded by the sum of the orders of their defining differential polynomials .
Proof
We prove this by providing a counter-example, which is nothing else but the addition of zeros of and from (40). By Lemma 1 it holds that the leading terms of the polynomials
(44) |
are all distinct. Since those leading terms are linear over , we deduce that each of their higher -derivative linearly occurs starting from a specific -truncation of the differential ideal
Therefore any (algebraic) elimination ideal computed from higher -truncations of would either contain (42), a -derivative of it, or another differential polynomial of higher -order; thus, supporting the fact that no higher -order truncation of contains a differential polynomial in of lower order than (42).
Note that the arguments used in the proof of Theorem 3.1 are valid for any l.h.o. partial differential polynomials with initials in . Thus it implies that our algorithm always yields the least-order algebraic PDEs for the arithmetic of zeros of l.h.o. differential polynomials with initials in , and can therefore be used to define lower bounds for the order in those cases. It would be interesting to study the behavior of the algorithm for arbitrary algebraic PDEs with an analysis of the transcendence basis when the -derivatives are taken. We think there is a connection between Kolchin polynomials and the growth of the number of partial derivatives involved in the truncated ideals (see [25, Section 12], [27]).
3.3 Generalization and algorithm
We now want to consider arbitrarily many independent variables. Suppose there are independent variables. As we have seen in the previous two subsections, the definition of a total ordering encoded by the operator using a bijective mapping from to is inherent to our algorithmic computations. One must always define how partial derivatives of differential indeterminates are selected because there is no natural way to do that like in the ordinary setting. A naive bijection can be defined in a recursive fashion with the Cantor pairing function in the base case. Indeed, one can show by induction that the maps defined by the recursion
(45) |
, , are bijective. However, these maps do not satisfy our desired properties. What is important for us in the choice of this bijection is the fulfillment of a property like Lemma 1 that gives theoretical meanings to all outputs of our algorithm for solutions of l.h.o. algebraic PDEs whose initials are polynomials in . For the functions above, one verifies, for instance, that but , which is not desired.
Having said that, there could still be several possible choices. For this paper, the choice is the Cantor -tuple function which generalizes our chosen in dimensions. This corresponds to the inverse lexicographic ordering, often called co-lexicographic ordering. For monomial ordering, this is understood as the degree lexicographic ordering with the variable ranked in the inverse order. Precisely, we consider as the -tuple function that ranks -tuples of integers as follows:
(46) |
We then take as the derivation rule that uses and its reciprocal similarly as uses and its reciprocal . Note that an explicit algebraic formula may be derived for this Cantor -tuple function. We will not dive into such details; the combinatorial investigation can be found in [38]. We only require fulfillment of properties like Proposition 2 and Lemma 1. The increasing property generalizes to arbitrarily many independent variables as we show in the next proposition.
Proposition 3 (Increasing property continued)
Let be a differential indeterminate. Consider the -ranking in . For all non-negative integer-valued tuples and an integer , we have
(47) |
Proof
Let such that w.r.t. the -ranking. By definition . Thus the application of to and adds to each and , . This preserves the inequality by the definition of the ranking in (46).
One easily verifies that a similar property as Proposition 2 also holds. Hence using we can compute ADEs for arithmetic of multivariate D-algebraic functions. Our algorithm that summarizes all the necessary steps is given by Algorithm 2.
-
1.
For each differential polynomial , let be the orders of w.r.t. .
-
2.
Set .
-
3.
Set . //Comment: can be optionally given in the input.
-
4.
Let be the numerator of the rational expression .
-
5.
Let be the set
-
6.
Let be the minimum order (w.r.t. the -ranking) that appears among the orders .
-
7.
Let .
-
8.
Update the set by adding the polynomials , and their first -derivatives to it,i.e.,
-
9.
Let , be the maximum order w.r.t. the -ranking of the differential indeterminate in , .
-
10.
Let
be the ideal defined by the polynomials in .
-
11.
Compute the elimination ideal
using the lexicographic monomial ordering as defined in (I.) and (II.) from page 36.
-
12.
If then:
-
•
Let be the polynomial of lowest degree among those of lowest -order in .
-
•
Stop and return the writing of in terms of partial derivatives.
-
•
-
13.
() While do:
-
•
(update by )
-
•
Repeat steps 8 to 12
-
•
-
14.
Return (meaning that “No ADE of order component-wisely less than found”).
Theorem 3.2
Algorithm 2 is correct, and when successful (non-zero output) with input ADEs that are l.h.o. algebraic PDEs with initials in , it returns an algebraic PDE of least-order w.r.t. the total ordering induced by .
Proof
As we have seen in the previous section, Algorithm 2 is an exhaustion of the possibilities to find a least-order ADE with a given bound. Therefore correctness easily follows. However, when the algorithm halts, there is no evidence that the output is necessarily of minimal order. In the case of l.h.o. algebraic PDEs whose initials are polynomials in the independent variables, the justification is the same as for the counter-example given in the proof of Theorem 3.1: the first non-trivial elimination always yields the least-order algebraic PDE possible.
Remark 4
We may speed up Algorithm 2 by checking the algebraic dependency with a Jacobian condition (see [14, Theorem 2.2]) instead of always computing Gröbner bases.
Let us say a few words on the reason why the triangular set arguments (see Theorem 2.1) do not apply the same way to guarantee the minimality of the order for arbitrary algebraic PDEs like in the ordinary case. When applying the -derivations, the transcendence degree increases because some variables occur independently of the algebraic dependencies encoded by a single PDE. This may explain why partial differential equations are often studied as systems instead of single equations like in the ordinary setting. To illustrate this phenomenon, consider our example of bivariate differential polynomials , , for which we were looking for an ADE fulfilled by sums of their zeros. Using the -ranking, the following system can be used to portray how the algebraic dependencies occur.
(48) |
Since these are linear equations, substitution easily does the elimination. The main question is to know whether we can always substitute higher -derivatives of and in higher -derivatives of by some rational expression in . The answer is no since applying to the last equation yields
and there is no link to . Thus could be seen as a new transcendent element to consider in the elimination process. Such a situation never happens with ordinary differential equations because . Note that an algebraic dependency between -derivatives will certainly arise at some point; however, it does not seem obvious to predict when that will occur, like in the ODE case.
3.4 Implementation
Our package NLDE contains a subpackage called MultiDalg that is being designed to perform operations with multivariate D-algebraic functions. The current implementation of Algorithm 2 is arithmeticMDalg in MultiDalg for both unary and -ary arithmetic of multivariate D-algebraic functions. The syntax is similar to that of NLDE:-arithmeticDalg. The monomial ordering used is either the lexicographic (default) or the lexdeg ordering with the corresponding -ranking, where is the number of independent variables.
Example 10
Consider the three PDEs:
(49) |
We view them as three independent PDEs and would like to compute a fourth PDE for the sum of their solutions. Using our package, this can be done with the code below.
(50) |
One can use Maple’s pdsolve command to check that sum of solutions to the PDEs in (49) are solutions of (50) and that the solutions of (50) are sums of solutions to the PDEs in (49). A nice algebraic perspective for solving such linear PDEs is given in [31, Section 3.3].
Example 11
We show how to recover the results of our explanatory examples with our implementation. To get (39), one uses the code below. We do not display the output as it is identical to (39). The last line gives the result.
For the second example, to check that there is no ADE of order less than , component wise, one uses the following code.
(51) |
The in the output should be interpreted as “no ADE of order less than the sum of the orders was found”. To get the ADE associated to (42), one supplies the optional argument maxord=[4,1] and gets:
(52) |
4 Discussions: applications and future developments
4.1 Applications
Let us discuss the potential application of our results, and in particular, the use of our software. First of all, it is worthwhile to remember that these elimination techniques are familiar to many scientists from many disciplines. However, it is often the case that the software they use, when the computations involved are tedious to do by hand, is tailored to a very specific problem and cannot be easily adapted when the underlying (D-algebraic) functions are changed. Usually one can notice that their accompanying codes for related problems are adaptations of an old code that was designed for some popular D-algebraic functions. See for instance, the recurrent use of elimination to find the fourth Painlevé transcendent of Hamiltonian representations in [13]. Since several functions in the sciences such as Physics, Biology, or Statistics are defined by algebraic ordinary or partial differential equations, usually with some initial values, one can use our software to compute differential equations satisfied by rational expressions of solutions to given ADEs. Let us demonstrate this with some concrete examples.
Example 12
In [45, Problem 6.24], the following algebraic ODE is derived from the Korteweg-de Vries equation:
(53) |
Physicists are interested in solutions which satisfy . This limits the admissible parameter on the shape of . One looks for a linear transformation of that eliminates the term from (53). Using NLDE, we compute an ADE for any such linear transformation and deduce a desired equation.
(54) |
From this output, one deduces that for the transformation eliminates the first derivative in the resulting equation. Thus we have found infinitely many transformations to eliminate the first derivative. A simple one would be , which yields
(55) |
Note that we recover the same result of the book by using the linear transformation (given in the book as . This gives the ADE .
Example 13
Observe that (53) can be integrated once to get
(56) |
where is the constant of integration. Note that the shape of also depends on . In [45, Problem 6.24], it is mentioned that satisfies (56), where is the Weierstrass elliptic function. Let us verify this fact. We use our software to get an ADE fulfilled by the rational expression from the ADE of the Weierstrass elliptic function.
The above two examples present how one can apply our result to “basic” (symbolic) calculations in Physics. Note that the Korteweg-de Vries equation is an algebraic PDE that models shallow water waves. One of its outstanding features is the existence of so-called soliton, i.e., waves that travel in time without changing their shape (see [45, Problem 6.24]).
For algebraic PDEs, there seems to be more interest in numerical methods than symbolic ones. The usual non-expressiveness of general solutions of PDEs may explain this. Nevertheless, as for algebraic ODEs, our software can compute ADEs satisfied by rational expressions of solutions to algebraic PDEs (see Example 16). However, note that so far, we only considered additions of solutions of linear algebraic PDEs. The main reason behind that is because finding least-order ADEs for addition of solutions to linear ADEs can be performed with Gauss elimination, and so carrying out the same computations using Gröbner bases is essentially the same. When it comes to other operations like the product, Gauss elimination will not generally give the desired least-order ADE, and then the computations with Gröbner bases pay the price of dealing with higher degree equations. For multivariate D-algebraic functions, this price is notoriously visible by the high CPU time encountered when running our implementation for product and division using the pure lexicographic monomial ordering. This behavior could be predicted by the number of variables involved in the elimination steps of Algorithm 2. We recommend the specification ordering=lexdeg which often (not always) enables computations in shorter timings.
There are many situations in which the algorithm proves useful. Let us give an idea of applications of Algorithm 2 related to solving PDEs. We construct PDEs out of solutions to known ODEs. This may be seen as a reverse engineering approach of the method of separating variables. The latter consists of finding solutions that can be expressed rationally in terms of solutions of ODEs in each independent variable. For instance, given a PDE in the independent variables and , we commonly look for solutions of the product form , where and are solutions of some ODEs in and , respectively. The product form is not very interesting for linear ODEs with constant coefficients, because viewing as a constant for a linear ODE in , implies already that is a solution of that ODE. Therefore the algorithm may just return a derivative of that ODE. We would like to consider more tricky examples.
Example 14
Suppose we want to construct an algebraic PDE whose solutions are of the form
(59) |
with arbitrary constants . Note that this is a well-known form in Statistics for the logistic distribution (see for instance [3]), where one of the variables would be interpreted as the mean.
We use the linear ADEs for and for . Observe that these ADEs could also be taken as PDEs, i.e, and ; however, using these PDEs would lead to an algebraic PDE whose solutions have a more general form. For instance, a solution of is of the form , which is not desired for our target form (59).
We here reveal a feature of our implementation that enables algebraic ODE inputs. The computations still use derivations as described in Algorithm 2. This decreases the number of variables used in the elimination step and makes the code get the result more efficiently. Indeed, when we apply derivations, and its derivatives w.r.t. are treated as constants for derivation w.r.t. , while and its derivatives w.r.t. are treated as constants for derivation w.r.t. . This results in the cancellation of all variables and , , with . Let us now compute the desired algebraic PDE.
(60) |
The obtained equation is the logistic differential equation, which could also be seen as an ODE in . This corresponds to solutions of the form
(61) |
where is an arbitrary function in . This is a more restricted form compared to the result obtained with PDEs as inputs.
Example 15
Taking the same algebraic ODEs as inputs, we now look for a PDE whose solutions have the form
(62) |
Our package yields:
Remark that the calling syntax of arithmeticMDalg in the above two examples contains a bracket list with all the independent variables, here , given as the last argument. This is how the code is called when there are algebraic ODEs in the input.
Example 16 (Bonus example)
We give a PDE fulfilled by the square of the probability density function
of the univariate normal distribution , seen as a bivariate function in the mean and the indeterminate variable. We use FPS:-HolonomicPDE (see [24]) to derive a partial PDE w.r.t. each independent variable, and use arithmeticMDalg to find an algebraic PDE for the product of their solutions.
(65) |
(66) |
(67) |
Solutions to this PDE have the form
(68) |
with arbitrary functions and .
4.2 Conclusion and future work
We completed the exposition of [1] on univariate D-algebraic functions by providing an algorithm (see Algorithm 1) to compute least-order ADEs for rational expressions of solutions to ADEs that are not linear in their highest occurring derivatives. This was accompanied by Theorem 2.1, which established its correctness. We have also proposed an algorithm for doing these computations with multivariate D-algebraic functions using an ordering on the semigroup of derivations that allows us to keep track of the application of partial derivatives (see Theorem 3.2 and Algorithm 2). From that, we established Theorem 3.1, which states that the sum of orders of given algebraic PDEs is not always a bound for the order of an algebraic PDE fulfilled by rational expressions of solutions to those ADEs. We implemented the given algorithm using the computer algebra system Maple. The accompanying software can be downloaded from MathRepo NLDE or [41] (with the source code) as a subpackage of the NLDE package. In certain calculations, our software may find similar results with the known packages [39], [8], [20]; however, we are here in a more general setting as we mentioned earlier. In the univariate setting, Algorithm 1 compares better with [2, 6]. Example 5 illustrates one such example where our algorithm outperforms those existing methods. We developed our approach from the works in [17, 12]. Our software can be used in many scientific disciplines where eliminations of variables in partial or ordinary differential equations are of interest. We have given some in Section 4.1. A Maple worksheet with the examples of the article can be downloaded from Arithmetic of D-Algebraic functions or [41].
We remark from references on partial differential equations that the compositions of solutions of algebraic PDEs with solutions of algebraic ODEs is often used to simplify algebraic PDEs. For instance, to get the equation in (53), one considers the composition with the Korteweg-de Vries equation
(69) |
We think that an adaptation of Algorithm 2 for compositions can be used to make such transformations of algebraic PDEs into algebraic ODEs more systematic.
We also highlight the potential use of -derivations for symbolic integration, which is also concerned with elimination. Indeed, given a symbolic expression in several variables , it is possible to develop a Fasenmyer-like algorithm (see [23, Theorem 11.4]) that looks for linear algebraic PDEs with eliminated variables among the coefficients, such that a differential equation for the integral can be deduced as a generalization of Feynman’s method. A preliminary implementation of this technique is provided in the FPS package by the HolonomicPDE command (see [24]). For a general knowledge of symbolic integration, we refer the reader to the non-exhaustive list of references [46, 9, 21, 26, 7].
Acknowledgment. The author thanks Bernd Sturmfels for encouraging this work. He appreciates Gleb Pogudin for reading and commenting on earlier versions of the article, and the anonymous referees for their valuable comments. He also thanks Rida Ait El Manssour and Anna-Laura Sattelberger for useful discussions.
References
- [1] Ait El Manssour, R., Sattelberger, A.L., Teguia Tabuguia, B.: D-algebraic functions. arXiv preprint arXiv:2301.02512 (2023)
- [2] Bächler, T., Gerdt, V., Lange-Hegermann, M., Robertz, D.: Algorithmic Thomas decomposition of algebraic and differential systems. Journal of Symbolic Computation 47(10), 1233–1266 (2012)
- [3] Balakrishnan, N.: Handbook of the Logistic Distribution. Marcel Dekker, New York (1992)
- [4] Bayer, D., Stillman, M.: A theorem on refining division orders by the reverse lexicographic order. Duke J. Math. 55(2), 321–328 (1987)
-
[5]
Boulier, F.: DifferentialAlgebra project: A C library for differential
elimination. Available at
https://codeberg.org/francois.boulier/DifferentialAlgebra - [6] Boulier, F., Lazard, D., Ollivier, F., Petitot, M.: Representation for the radical of a finitely generated differential ideal. In: Proceedings of the 1995 International Symposium on Symbolic and Algebraic Computation. pp. 158–166 (1995)
- [7] Chen, S., Kauers, M.: Some open problems related to creative telescoping. Journal of Systems Science and Complexity 30(1), 154–172 (2017)
-
[8]
Chyzak, F.: Mgfun project: a Maple package for systems of holonomic
functions. Available at
https://specfun.inria.fr/chyzak/mgfun.html - [9] Chyzak, F.: An extension of Zeilberger’s fast algorithm to general holonomic functions. Discrete Mathematics 217(1-3), 115–134 (2000)
- [10] Denef, J., Lipshitz, L.: Power series solutions of algebraic differential equations. Mathematische Annalen 267, 213–238 (1984)
- [11] van Der Hoeven, J.: Computing with D-algebraic power series. Appl. Algebra Engrg. Comm. Comput. 30(1), 17–49 (2019)
- [12] Dong, R., Goodbrake, C., Harrington, H.A., Pogudin, G.: Differential elimination for dynamical models via projections with applications to structural identifiability. SIAM Journal on Applied Algebra and Geometry 7(1), 194–235 (2023)
- [13] Dzhamay, A., Filipuk, G., Ligȩza, A., Stokes, A.: Different Hamiltonians for the Painlevé equation and their identification using a geometric approach. arXiv preprint arXiv:2109.06428 (2021)
- [14] Ehrenborg, R., Rota, G.C.: Apolarity and canonical forms for homogeneous polynomials. European Journal of Combinatorics 14(3), 157–181 (1993)
- [15] Freitag, J., León Sánchez, O., Li, W.: Effective definability of Kolchin polynomials. Proceedings of the American Mathematical Society 148(4), 1455–1466 (2020)
- [16] Fueter, R., Pólya, G.: Rationale abzählung der gitterpunkte. Vierteljschr. Naturforsch. Ges. Zürich 58, 380–386 (1923)
- [17] Hong, H., Ovchinnikov, A., Pogudin, G., Yap, C.: Global identifiability of differential models. Communications on Pure and Applied Mathematics 73(9), 1831–1879 (2020)
- [18] Hubert, E.: Notes on triangular sets and triangulation-decomposition algorithms I: Polynomial systems. In: Symbolic and Numerical Scientific Computation: Second International Conference, SNSC 2001, Hagenberg, Austria, September 12–14, 2001. Revised Papers. pp. 1–39. Springer (2003)
- [19] Jiménez-Pastor, A., Pillwein, V., Singer, M.F.: Some structural results on -finite functions. Advances in Applied Mathematics 117, 102027 (2020)
- [20] Kauers, M., Mezzarobba, M.: Multivariate ore polynomials in SageMath. ACM Communications in Computer Algebra 53(2), 57–60 (2019)
- [21] Kauers, M., Paule, P.: The Concrete Tetrahedron. Symbolic Sums, Recurrence Equations, Generating Functions, Asymptotic Estimates. Springer-Verlag, Wien (2011). https://doi.org/10.1007/978-3-7091-0445-3
- [22] Kauers, M., Yatchak, R.: Walks in the quarter plane with multiple steps. In: Proceedings of FPSAC’15, DMTCS. pp. 25–36 (2015)
- [23] Koepf, W.: Computer Algebra: An Algorithm-Oriented Introduction. Springer Nature (2021)
-
[24]
Koepf, W., Teguia Tabuguia, B.: FPS: A Maple package for symbolic computations
with power series and linear PDEs. Available at
http://www.mathematik.uni-kassel.de/~bteguia/FPS_webpage/FPS.htm - [25] Kolchin, E.R.: Differential Algebra & Algebraic Groups. Academic press (1973)
- [26] Koutschan, C.: Advanced Applications of the Holonomic Systems Approach. Ph.D. thesis, RISC, Johannes Kepler University, Linz (September 2009)
- [27] Levin, A.: Bivariate Kolchin-type dimension polynomials of non-reflexive prime difference-differential ideals. the case of one translation. Journal of Symbolic Computation 102, 173–188 (2021)
- [28] Lipshitz, L., Rubel, L.A.: A gap theorem for power series solutions of algebraic differential equations. American Journal of Mathematics 108(5), 1193–1213 (1986)
- [29] Lipshitz, L., Rubel, L.A.: A gap theorem for differentially algebraic power series. In: Chudnovsky, D.V., Chudnovsky, G.V., Cohn, H., Nathanson, M.B. (eds.) Number Theory. pp. 211–214. Springer New York, New York, NY (1991)
- [30] Maillet, E.: Sur les séries divergentes et les équations différentielles. Annales Scientifiques de l’École Normale Supérieure 20(3), 487–518 (1903)
- [31] Michałek, M., Sturmfels, B.: Invitation to Nonlinear Algebra, vol. 211. American Mathematical Soc. (2021)
- [32] Moore, E.H.: Concerning transcendentally transcendental functions. Mathematische Annalen 48(1-2), 49–74 (1896)
- [33] Pavlov, D., Pogudin, G.: On realizing differential-algebraic equations by rational dynamical systems. In: Proceedings of the 2022 International Symposium on Symbolic and Algebraic Computation. pp. 119–128 (2022)
-
[34]
Pogudin, G.: Differential algebra. Lecture notes, available at
http://www.lix.polytechnique.fr/Labo/Gleb.POGUDIN/files/da_notes.pdf - [35] Raschel, K., Bousquet-Mélou, M., Bernardi, O.: Counting quadrant walks via Tutte’s invariant method. Discrete Math. Theor. Comput. Sci. (2020)
- [36] Ritt, J.F.: Differential Algebra, vol. 33. American Mathematical Soc. (1950)
- [37] Robertz, D.: Formal Algorithmic Elimination for PDEs, vol. 2121. Springer (2014)
-
[38]
Ruskey, F.: Combinatorial Generation. Book available at
https://page.math.tu-berlin.de/~felsner/SemWS17-18/Ruskey-Comb-Gen.pdf (2003), preliminary working draft. University of Victoria, Victoria, BC, Canada - [39] Salvy, B., Zimmermann, P.: GFUN: a Maple package for the manipulation of generating and holonomic functions in one variable. ACM Transactions on Mathematical Software 20(2), 163–177 (1994)
- [40] Stanley, R.P.: Differentiably finite power series. European Journal of Combinatorics 1(2), 175–188 (1980)
- [41] Teguia Tabuguia, B.: NLDE: NonLinear algebra and Differential Equations: source and lattest update. Available at https://github.com/T3gu1a/D-algebraic-functions
- [42] Teguia Tabuguia, B.: Guessing with quadratic differential equations. Software Demo at ISSAC’22. To appear in ACM Communications in Computer Algebra (2022)
- [43] Teguia Tabuguia, B.: Operations for D-algebraic functions. ACM Communications in Computer Algebra 57(2), 51–56 (2023)
- [44] Teguia Tabuguia, B., Koepf, W.: On the representation of non-holonomic univariate power series. Maple Trans. 2(1) (2022), article 14315
- [45] Teschl, G.: Ordinary Differential Equations and Dynamical Systems, vol. 140. American Mathematical Soc. (2012)
- [46] Zeilberger, D.: The method of creative telescoping. J. Symb. Comput. 11(3), 195–204 (1991)