Computation of The Unipotent Radical of The Differential Galois Group For A Parameterized Second-Order Linear Differential Equation
Computation of The Unipotent Radical of The Differential Galois Group For A Parameterized Second-Order Linear Differential Equation
a r t i c l e i n f o a b s t r a c t
http://dx.doi.org/10.1016/j.aam.2014.03.001
0196-8858/© 2014 Elsevier Inc. All rights reserved.
C.E. Arreche / Advances in Applied Mathematics 57 (2014) 44–59 45
1. Introduction
n−1
δxn Y + ri δxi Y = 0, (1.1)
i=0
δx2 Y − qY = 0, (1.2)
where q ∈ F (x) =: K, and F is a Π-field. In [11], Dreyfus applies results from [10]
to develop algorithms to compute H (see also [1] for a detailed discussion of Dreyfus’
results, in the setting of one parametric derivation). This algorithm is extended to arbi-
trary second-order linear differential equations over K in [3], by reinterpreting a classical
change-of-variables procedure in Galois-theoretic terms. In [11, §2.1], the computation
of Ru (H) is reduced to the computation of a finite set of linear differential operators in
F [Π], and a method is given for their computation that is known to halt under the as-
sumption that the maximal reductive quotient H/Ru (H) is Π-constant (cf. [24, Alg. 1]).
46 C.E. Arreche / Advances in Applied Mathematics 57 (2014) 44–59
∂2γ ∂γ
−p = 0, (1.3)
∂x2 ∂x
where p := 1−t−x
x . This is an example of a parameterized linear differential equation over
C(x, t), where δx := ∂x ∂
, and Π := { ∂t
∂
} is the parametric derivation. In [15] it was shown
that γ(t, x) does not satisfy any polynomial differential equations over C(x, t, ex , log(x))
∂
with respect to ∂t . This result was necessary in [6, Ex. 7.2] for the computation of Gγ ;
specifically, to conclude that Ru (Gγ ) = Ga , the additive group. In [2, Thm. 3.2], whose
proof follows the ideas of [14, Cor. 3.4.1], the differential transcendence of the solutions to
(1.3) was characterized in terms of two conditions on the coefficient p. Letting G (resp.,
G ) denote the PPV-group (resp., the non-parameterized PV-group) corresponding to
(1.3) for an arbitrary p ∈ C(x, t), the first condition is equivalent to the statement that
G/Ru (G) is not ∂t ∂
-constant, and the second condition says that Ru (G ) = Ga . When
the first condition is satisfied, one proves as in [14, Lem. 3.6(2)] that Ru (G) is either 0
or Ga . Since G is Zariski-dense in G by [6, Prop. 3.6(2)], it follows that Ru (G) = Ga
precisely when Ru (G ) = Ga . In other words, whenever G/Ru (G) is not ∂t ∂
-constant,
then Ru (G) is defined by the same (non-differential) equations as Ru (G ). Since G is
technical results used in the proof of [2, Thm. 3.2]. This is done in Lemmas 3.5 and 3.7
(see also Remarks 3.6 and 3.6).
An algorithm to compute the unipotent radical Ru (G) of the PPV-group G cor-
responding to an nth-order parameterized linear differential equation over K is given
in [24], under the familiar assumption that G/Ru (G) is Π-constant. We expect that
an elaboration of the methods presented in this paper will be successful in extending
the procedure of [24] to compute Ru (G) in cases where G/Ru (G) is not necessarily
Π-constant. Algorithms to compute telescopers for rational functions, algebraic func-
tions, and hyper-exponential functions, are given in [8,9], and [4], respectively. The
notion of parallel telescoping investigated in [7] leads to algorithms [7, §5] to compute
PPV-groups in the setting of several principal derivations and one parametric derivation
(see Section 2.1). The creative telescoping problems solved by these algorithms lie at
the core of the algorithms presented in [11] (see also [1]) to compute the PPV-group
of (1.2). The hyper-exponential assumption, which is necessary in order to apply the
algorithms of [7,8] to compute Ru (H), coincides in this case with the requirement that
H/Ru (H) be Π-constant (see [7, §4] and Remark 3.1). We refer to [7, §1] for more details
and references concerning the connection between creative telescoping problems and the
computation of PPV-groups.
2. Preliminaries
See [19,26] for more details concerning the following definitions. A Δ-ring is a ring A
equipped with a finite set Δ := {δ1 , . . . , δm } of commuting derivations (that is, δi (ab) =
aδi (b) + δi (a)b and δi δj = δj δi for each a, b ∈ A and 1 i, j m). We often omit the
parentheses, and write δa for δ(a). For Π ⊆ Δ, we denote the subring of Π-constants
of A by AΠ := {a ∈ K | δa = 0, δ ∈ Π}. When Π = {δ} is a singleton, we write Aδ
instead of AΠ . If A = K happens to be a field, we say that (K, Δ) is a Δ-field. Every
field is assumed to be of characteristic zero.
The ring of differential polynomials over K (in m differential indeterminates) is de-
noted by K{Y1 , . . . , Ym }Δ . Algebraically, it is the free K-algebra in the countably infinite
set of variables {θYi | 1 i m, θ ∈ Θ}, where
Θ := δ1r1 · · · δnrn ri ∈ Z0 for 1 i n
is the free commutative monoid on the set Δ. The ring K{Y1 , . . . , Ym }Δ carries a natural
structure of Δ-ring, given by δi (θYj ) := (δi · θ)Yj .
We say p ∈ K{Y1 , . . . , Ym }Δ is a linear differential polynomial if it belongs to the
K-vector space spanned by the θYj , for θ ∈ Θ and 1 j m. The K-vector space of
linear differential polynomials will be denoted by K{Y1 , . . . , Ym }1Δ .
The ring of linear differential operators K[Δ] is the K-span of Θ. Its (non-
commutative) ring structure is defined by composition of additive endomorphisms
48 C.E. Arreche / Advances in Applied Mathematics 57 (2014) 44–59
of K. The canonical identification of (left) K-vector spaces K[Δ] K{Y }1Δ given by
θ aθ θ ↔ θ aθ θY will be assumed implicitly in what follows.
If M is a Δ-field and K ⊆ M is a subfield such that δ(K) ⊂ K for each δ ∈ Δ, we say
K is a Δ-subfield of M and M is a Δ-field extension of K. If y1 , . . . , yn ∈ M , we denote
the Δ-subfield of M generated over K by all the derivatives of the yi by
K y1 , . . . , yn
Δ ⊆ M.
We recall some standard facts from the parameterized Picard–Vessiot theory [6] (see
also [12,14,22]) and the theory of linear differential algebraic groups [5] (see also [20,25]).
Let F be a Π-field, where Π := {∂1 , . . . , ∂m }, and let K := F (x) be the field of rational
functions in x with coefficients in F . Let Δ := ({δx }∪Π), and consider K as a Δ-field by
setting δx x = 1, K δx = F , and ∂i x = 0 for each i. We will sometimes refer to δx as the
main derivation, and to Π as the set of parametric derivations. Consider the following
linear differential equation with respect to the main derivation, where ri ∈ K for each
0 i n − 1:
n−1
δxn Y + ri δxi Y = 0. (2.1)
i=0
If F is Π-closed, it is shown in [6] that a PPV-extension and PPV-group for (2.1) over
K exist and are unique up to K-Δ-isomorphism. Although this assumption allows for a
simpler exposition of the theory, several authors [12,30] have shown that, in many cases
of practical interest, the parameterized Picard–Vessiot theory can be developed without
C.E. Arreche / Advances in Applied Mathematics 57 (2014) 44–59 49
assuming that F is Π-closed. In any case, we may always embed F in a Π-closed field
[16,29].
The action of GalΔ (M/K) is determined by its restriction to S, whence a choice of
F -basis for S determines an embedding GalΔ (M/K) → GLn (F ). It is shown in [6] that
this embedding identifies the PPV-group with a linear differential algebraic group.
Definition 2.2. (See [5,20].) Let F be a Π-closed field. We say that a subgroup G ⊆
GLn (F ) is a linear differential algebraic group if G is defined as a subset of GLn (F )
by the vanishing of a system of polynomial differential equations in the matrix entries,
with coefficients in F . We say that G is Π-constant if it is conjugate to a subgroup of
GLn (F Π ).
The study of linear differential algebraic groups was pioneered in [5], where the differ-
ential algebraic subgroups of the additive group Ga (F ) and the multiplicative group
Gm (F ) were classified in terms of finite sets of linear differential operators (see [5,
Prop. 11, Prop. 31 and its Corollary]). The differential algebraic subgroups of SL2 (F )
were classified in [28]. When the PPV-group GalΔ (M/K) is Π-constant, it is proved in [6,
Prop. 3.9(1)] that (2.1) is completely integrable [6, Defn. 3.8], and that M is a Picard–
Vessiot-extension (or PV-extension) of K for (2.1), in the non-parameterized sense of
[17]. The algorithms developed in [13] drastically reduce the number of conditions that
one has to check in order to decide whether (2.1) is completely integrable.
There is a parameterized Galois correspondence between the differential algebraic
subgroups Γ of GalΔ (M/K) and the intermediate Δ-fields K ⊆ L ⊆ M , given by
Γ
→ M Γ and L
→ GalΔ (M/L). Under this correspondence, an intermediate Δ-field
L is a PPV-extension of K (for some linear differential equation with respect to δx ) if
and only if GalΔ (M/L) is normal in GalΔ (M/K); in which case the homomorphism
GalΔ (M/K) GalΔ (L/K), defined by σ
→ σ|L , is surjective with kernel GalΔ (M/L).
See [6, Thm. 3.5] and [12, §8.1] for more details.
3. Main result
δx2 Y − qY = 0, (3.1)
In [11] (see also [1]), Dreyfus develops a procedure to compute the PPV-group H corre-
sponding to (3.1). We begin with a brief summary of this procedure in the non-reductive
case (see [23,24]).
When H is not reductive [23, Defn. 2.2.6], it is proved in [21] that there exists an
F -basis {η, ξ} for the solution space S of (3.1) such that δx η = uη for some u ∈ K, and
δx ( ηξ ) = η −2 . The embedding H → SL2 (F ) is defined by
δx Y − uY = 0, (3.3)
then the unipotent radical (see [23, Defn. 2.2.5]) B Ru (H) coincides with the
PPV-group GalΔ (M/L), and the maximal reductive quotient A H/Ru (H) is canoni-
cally isomorphic to GalΔ (L/K).
We refer to [11] for the computation of A (see also [4,8,21,23,27]). By the classification
result of [5, Prop. 11], B is described by a finite set of linear differential operators
p1 , . . . , ps ∈ F [Π], and it is shown in [11, §2] that they satisfy pi (η −2 ) ∈ δx (L) for
1 i s. When A ⊆ Gm (F Π ), it is proved in [24] that M is of finite algebraic
transcendence degree over K, whence there exist bounds on the orders of the pi . But
if A Gm (F Π ) and B = 0, then M is of infinite algebraic transcendence degree over
L (see [2,24], and cf. Corollary 3.3). We are not aware of a priori bounds on the orders
of the pi in this case (cf. [24, p. 13]), which raises the problem of deciding whether all
the pi have already been found, or whether it is still necessary to do more prolongations
(see [23–25]).
In the setting of one parametric derivation Π = {∂}, the problem of computing
Ru (H) when H/Ru (H) fails to be ∂-constant was solved completely in [1,2]. In this case,
it follows from [14, proof of Lem. 3.6(2)] that either Ru (H) = Ga (F ) or Ru (H) = 0, and
there are no other possibilities. In light of [24], this has the counterintuitive consequence
that the computation of Ru (H) is actually easier when H/Ru (H) is not ∂-constant (see
[2, Thm. 3.2]), since in this case the parametric derivation ∂ is barred from appearing
in the defining equations for Ru (H). Theorem 3.2 describes the generalization of this
phenomenon to the setting of several parametric derivations.
C.E. Arreche / Advances in Applied Mathematics 57 (2014) 44–59 51
L := {∂ ∈ D | ∂a = 0, ∀a ∈ A}. (3.4)
Remark 3.1. Let us briefly describe how to compute Π . Observe that, for every ∂ ∈ D
and σ ∈ H,
∂η ∂η ∂aσ ∂η
σ = + and δx = ∂u. (3.5)
η η aσ η
Theorem 3.2 (Main result). The reductive quotient H /Ru (H ) is Π -constant, and the
linear differential operators {pi }si=1 ⊂ F [Π ] defining Ru (H ) are also defining operators
for Ru (H), under the natural inclusion F [Π ] ⊆ F [Π].
Lemma 3.4, we have that B ⊆ B . By Lemma 3.5, there is a finite set {pi }si=1 ⊂ F [Π ]
such that B coincides with the set of those b ∈ Ga (F ) such that pi (b) = 0 for each
1 i s. By Lemma 3.7, pi (b ) = 0 for each b ∈ B and 1 i s, whence B ⊆ B. 2
52 C.E. Arreche / Advances in Applied Mathematics 57 (2014) 44–59
Corollary 3.3. (See [14, Lem. 3.6(2)] and [2, Prop. 4.4].) Suppose that Ru (H) = {0}.
Then,
Ru (H) Ga (F ) ⇐⇒ L = {0}.
The following three lemmas were used in the proof of Theorem 3.2.
The fact that B is the unipotent radical of (3.2), and not just any differential algebraic
subgroup of Ga (F ), allows us to produce a set of defining operators for B from F [Π ] ⊆
F [Π], which sharpens the classification result of [5, Prop. 11] in this very particular
case. The following structural result, which was inspired by the results of [28] cited in its
proof, holds true for any linear differential algebraic group G of the form (3.2), whether
or not it happens to be a PPV-group over K.
Lemma 3.5. (See [14, Lem. 3.6(2)], [28, Thm. II.1.3 and Thm. II.1.4].) There exist
finitely many linear differential operators p1 , . . . , ps ∈ F [Π ] ⊆ F [Π] such that
B = b ∈ F pi (b) = 0, 1 i s .
Proof. By [20, Prop. 0.7] the F -basis Π for L can be extended to a commuting F -basis
Π̃ := {∂1 , . . . , ∂m
} for all of D. We denote by Θ̃ (resp., Θ ) the free commutative monoid
generated by Π̃ (resp., Π ). Consider the orderly ranking on F {Y }Π̃ determined by the
lexicographic order on Θ̃ defined by setting δi δj if j i. In other words, to compare
two elements θ, θ in Θ̃, we first compare their total orders, and then the exponents of
∂1 , . . . , ∂m
, in that order.
C.E. Arreche / Advances in Applied Mathematics 57 (2014) 44–59 53
By [28, Thm. II.1.3(b) and Thm. II.1.4], there is a characteristic set {p1 , . . . , ps } for
the defining ideal of B (with respect to this ranking) such that pi (aY ) = api (Y ) for
each a ∈ A and 1 i s. Therefore, to show that {pi }si=1 ⊂ F [Π ], it suffices to prove
that if p ∈ F [Π̃] does not belong to the image of F [Π ] under the natural embedding
F [Π ] ⊆ F [Π], then there exists an element a ∈ A such that p(aY ) − ap(Y ) = 0.
So suppose that p ∈ F [Π̃] and p ∈ / F [Π ], and let cθ θY be the monomial in p of
highest rank such that cθ = 0 and θ contains a derivation
∂ ∈ Π̃\Π = ∂k+1
, . . . , ∂m .
Assume that ∂ is the derivation of lowest rank appearing effectively in θ, and let θ̃ denote
the element of Θ̃ obtained from θ by decreasing the order of ∂ by 1. Since θ (aY ) = aθ Y
for every a ∈ A and θ ∈ Θ , the leader of p(aY ) − ap(Y ) is cθ ∂ (a)θ̃Y whenever a ∈ A.
Since ∂ ∈
/ L, there exists a ∈ A such that ∂ (a) = 0, whence p(aY ) − ap(Y ) = 0. 2
Remark 3.6. When A is Π-constant, we may take Π = Π, and Lemma 3.5 coincides with
[5, Prop. 11]. In case that Π = {∂} is a singleton and Π = ∅, Lemma 3.5 is equivalent
to [14, Lem. 3.6(2)].
The previous result shows that B can be defined as a subset of Ga (F ) using derivations
from Π only. The following lemma rules out the possibility that B could somehow be
defined by more Π -differential equations than B is.
Lemma 3.7. (See [6, Prop. 3.6(2)].) If p ∈ F [Π ] is such that p(b) = 0 for every b ∈ B,
then p(b ) = 0 for every b ∈ B . In other words, B ⊆ B is Π -dense.
Proof. Suppose that p ∈ F [Π ] is such that p(b) = 0 for each b ∈ B. Then by [11, §2.1,
p. 7], we have p(η −2 ) ∈ δx (L). Moreover, since p ∈ F [Π ],
p η −2 ∈ K η
Δ =: L ,
the fixed field of Ru (H ). We will show that in fact p(η −2 ) ∈ δx (L ). Again by [11, §2.1,
p. 7], this will imply that p(b ) = 0 for each b ∈ B . Assume that η is algebraically
transcendental over K, since otherwise A μk , the group of kth roots of unity (see [5,
Prop. 31]), in which case H = H and in particular B = B .
By [6, Prop. 3.9] (cf. Remark 3.1), the fact that A is Π -constant implies that
∂j η
vj := ∈K (3.7)
η
for each ∂j ∈ Π , and therefore L = K(η) is algebraically isomorphic to the field of
rational functions in η with coefficients in K. It also follows from (3.7) that
−2vj = η 2 ∂j η −2 ∈ K. (3.8)
54 C.E. Arreche / Advances in Applied Mathematics 57 (2014) 44–59
proves the induction step, and our claim. Hence, η 2 p(η −2 ) ∈ K for every p ∈ F [Π ].
Since
∂1 η ∂m η
L := K η
Δ = K(η) ∂1 η, . . . , ∂m η
Δ = K(η) ,..., ,
η η Δ
where Θ is the free commutative monoid on Π. By [23, Cor. 5.1.2] and [24, Prop. 3.2] (see
also [24, §3.2.1]), if we consider L and K as δx -fields, then L is a (non-parameterized)
PV-extension of K, and the algebraic transcendence degree of L over K is finite. Hence,
we may choose a finite set β1 , . . . , βs of F -linearly independent generators for L over
L from the set (3.9). It follows from (3.5) that δx βi ∈ K for each 1 i s. By the
Kolchin–Ostrowski theorem [18], the βi are then algebraically independent over L . We
define
N := K(β1 , . . . , βs ),
and observe that L = N (η). Since A is abelian, the subgroup GalΔ (L/N ) A is normal
and consequently N is a PPV-extension of K, by the parameterized Galois correspon-
dence [6, Thm. 3.5]. Again by [23, Cor. 5.1.2] and [24, Prop. 3.2], the δx -field N is a
(non-parameterized) PV-extension of the δx -field K. Since
δx η
=u∈K and δx βi ∈ K
η
P := K[β1 , . . . , βs ] ⊂ N. (3.10)
Let f ∈ L such that δx (f ) = p(η −2 ). We claim that there exist elements g ∈ N and
c ∈ F such that f = gη −2 + c. To see this, let h ∈ K be such that
δx f = p η −2 = hη −2 ,
C.E. Arreche / Advances in Applied Mathematics 57 (2014) 44–59 55
Let e0 = 0, and apply δx on both sides of (3.11) to obtain (cf. [1, Lem. 2.1]):
f = g2,0 η −2 + c0 ,
g2,0 = η 2 (f − c0 ) ∈ L = N (η),
(δx g − 2ug)η −2 = δx f = hη −2
δx g − 2ug = h. (3.13)
the coefficient of β I in the right-hand side of (3.13) is 0, we see that δx rI = 2urI , which
implies that rI = aη 2 for some 0 = a ∈ F , a contradiction. Hence, no such monomial
rI β I appears in g, which means that g ∈ K and gη −2 + c = f ∈ L , concluding the proof
of the lemma. 2
Remark 3.8. In case Π = ∅, or in other words when the Lie subspace L defined in (3.4)
is {0}, then H is the (non-parameterized) PV-group for (3.1), and Lemma 3.7 reduces
to a special case of [6, Prop. 3.6(2)].
4. An example
We let K = F (x) denote the Δ-field defined in the previous section, where Π :=
{∂1 , ∂2 }, ∂j := ∂t∂j for j = 1, 2, and F denotes a Π-closed field containing Q(t1 , t2 )
(see [16,29]). In this section, we will apply Theorem 3.2 to compute the PPV-group H
corresponding to the parameterized linear differential equation
t1 t2 − 1 − x
u := .
x
Therefore, by [21] there is a basis {η, ξ} for the solution space of (4.1) such that δx η = uη
and δx ( ηξ ) = η −2 , and by [11, §2.1] there exist differential algebraic subgroups A
Gm (F ) and B Ga (F ) such that H is given by (3.2). Since ∂1 u = tx2 and ∂2 u = tx1 , we
have that
is isomonodromic [13] (or completely integrable, in the terminology of [6, Defn. 3.8]).
Therefore, by [6, Prop. 3.9] L = K(η) is a (non-parameterized) PV-extension of K for
(4.3), and (cf. Theorem 3.2)
GalΔ L /K H /Ru H A = Gm F ∂1 .
By [2, Lem. 4.3] and [6, Prop. 2.6(2)], to see that B = 0, it suffices to show that there
is no f ∈ K such that δx f + 2uf = 1. We prove this by contradiction, along the lines of
[2, proof of Cor. 3.3]. Assume that f ∈ K satisfies
δx f + 2uf = 1. (4.4)
First, note that f cannot be δx -constant, whence it must a have a pole somewhere in
P1 (F ). But f cannot have a pole outside of {0, ∞}, for otherwise the left-hand side of
(4.4) would have a pole. If f had a pole at 0, the residue of 2u at 0 would have to be an
integer, which is clearly false. Therefore, f can only have a pole at ∞, i.e., f must be a
polynomial in x. Moreover, f must be divisible by x, because otherwise the left-hand side
of (4.4) would have a pole at 0. But then the degree of the polynomial on the left-hand
side of (4.4) is equal to the degree of f , which is at least 1, since f is not constant. This
contradiction concludes the proof that there is no solution in K for (4.4), and therefore
that B = 0. Since
ξ ξ
δx ∂1 = ∂1 η −2 = 0 =⇒ ∂1 ∈ F = δx (F · x) ⊂ δx (K),
η η
it follows from [11, §2.1, p. 7] that B = Ga (F ∂1 ). Therefore, by Theorem 3.2,
a b a, b ∈ F ; a = 0; t1 ∂a1 a = t2 ∂a2 a ;
H .
0 a−1 ∂1 ( ∂a1 a ) = 0 = ∂2 ( ∂a2 a ); t1 ∂1 b = t2 ∂2 b
Acknowledgments
This work owes very much to Alexey Ovchinnikov’s essential suggestions and cor-
rections, as well as his patience, guidance and advice over the course of this project.
Richard Churchill’s generosity with his time and ideas has helped me better understand
numerous points in Picard–Vessiot theory and Kovacic’s algorithm. Phyllis Cassidy and
58 C.E. Arreche / Advances in Applied Mathematics 57 (2014) 44–59
William Sit have been invaluable guides in almost every aspect of differential algebra
with which I have struggled. Also indispensable were many helpful discussions about
this work in particular, and so much else, with: the members of the Kolchin Seminar
in New York, Raymond Hoobler, Michael Singer, Sergey Gorchinskiy, William Keigher,
Jacques-Arthur Weil, Thomas Dreyfus, Michael Wibmer, James Freitag, and Omar de
León Sánchez. Manuel Kauers and the referee did excellent editorial work on behalf of
the journal. I sincerely thank all of them for all their help and for their kind interest in
my work.
This material is based upon work supported by the National Science Foundation
(NSF) Graduate Research Fellowship Program under grant 40017-04-06 and by NSF
grant CCF-0952591.
References
[1] C.E. Arreche, Computing the differential Galois group of a one-parameter family of second order
linear differential equations, preprint, arXiv:1208.2226 (2012).
[2] C.E. Arreche, A Galois-theoretic proof of the differential transcendence of the incomplete Gamma
function, J. Algebra 389 (2013) 119–127, http://dx.doi.org/10.1016/j.jalgebra.2013.04.037.
[3] C.E. Arreche, Computing the differential Galois group of a parameterized second-order linear dif-
ferential equation, To appear in Proceedings of ISSAC, 2014, arXiv:1401.5127.
[4] A. Bostan, S. Chen, F. Chyzak, Z. Li, G. Xin, Hermite Reduction and Creative Telescoping for
Hyperexponential Functions, Proceedings of ISSAC, 2013, pp. 77–84.
[5] P.J. Cassidy, Differential algebraic groups, Amer. J. Math. 94 (1972) 891–954.
[6] P.J. Cassidy, M.F. Singer, Galois theory of parameterized differential equations and linear differential
algebraic groups, in: D. Bertrand, B. Enriquez, C. Mitschi, C. Sabbah, R. Schaefke (Eds.), IRMA
Lect. Math. Theor. Phys., vol. 9, Eur. Math. Soc. Publishing House, Zürich, 2007, pp. 113–155,
http://dx.doi.org/10.4171/020-1/7.
[7] S. Chen, R. Feng, Z. Li, M.F. Singer, Parallel telescoping and parameterized Picard–Vessiot theory,
To appear in Proceedings of ISSAC, 2014, arXiv:1401.4666.
[8] S. Chen, M. Kauers, M.F. Singer, Telescopers for Rational and Algebraic Functions via Residues,
Proceedings of ISSAC, 2012, pp. 130–137.
[9] S. Chen, M.F. Singer, Residues and telescopers for bivariate rational functions, Adv. in Appl. Math.
49 (2012) 111–133, http://dx.doi.org/10.1016/j.aam.2012.04.003.
[10] T. Dreyfus, A density theorem for parameterized differential Galois theory, Pacific J. Math. (2014),
in press arXiv:1203.2904.
[11] T. Dreyfus, Computing the Galois group of some parameterized linear differential equa-
tion of order two, Proc. Amer. Math. Soc. 142 (2014) 1194–1207, http://dx.doi.org/10.1090/
S0002-9939-2014-11826-0.
[12] H. Gillet, S. Gorchinskiy, A. Ovchinnikov, Parameterized Picard–Vessiot extensions and Atiyah
extensions, Adv. Math. 238 (2013) 322–411.
[13] S. Gorchinskiy, A. Ovchinnikov, Isomonodromic differential equations and differential Tannakian
categories, J. Math. Pures Appl. (2014), http://dx.doi.org/10.1016/j.matpur.2013.11.001, in press.
[14] C. Hardouin, M.F. Singer, Differential Galois theory of linear difference equations, Math. Ann. 342
(2008) 333–377, http://dx.doi.org/10.1007/s00208-008-0238-z.
[15] J. Johnson, G. Reinhart, L. Rubel, Some counterexamples to separation of variables, J. Differential
Equations 121 (1995) 42–66.
[16] E. Kolchin, Constrained extensions of differential fields, Adv. Math. 12 (1974) 141–170,
http://dx.doi.org/10.1016/S0001-8708(74)80001-0.
[17] E.R. Kolchin, Algebraic matric groups and the Picard–Vessiot theory of homogeneous linear ordinary
differential equations, Ann. of Math. 49 (1948) 1–42.
[18] E.R. Kolchin, Algebraic groups and algebraic dependence, Amer. J. Math. 90 (1968) 1151–1164.
[19] E.R. Kolchin, Differential Algebra and Algebraic Groups, Academic Press, New York, 1976.
[20] E.R. Kolchin, Differential Algebraic Groups, Pure Appl. Math., vol. 114, Academic Press, Orlando,
FL, 1985.
C.E. Arreche / Advances in Applied Mathematics 57 (2014) 44–59 59
[21] J.J. Kovacic, An algorithm for solving second order linear homogeneous differential equations,
J. Symbolic Comput. 2 (1986) 3–43, http://dx.doi.org/10.1016/S0747-7171(86)80010-4.
[22] P. Landesman, Generalized differential Galois theory, Trans. Amer. Math. Soc. 360 (2008)
4441–4495, http://dx.doi.org/10.1090/S0002-9947-08-04586-8.
[23] A. Minchenko, A. Ovchinnikov, M.F. Singer, Reductive linear differential algebraic groups and
the Galois groups of parameterized linear differential equations, Int. Math. Res. Not. (2014),
http://dx.doi.org/10.1093/imrn/rnt344, in press.
[24] A. Minchenko, A. Ovchinnikov, M.F. Singer, Unipotent differential algebraic groups as param-
eterized differential Galois groups, J. Inst. Math. Jussieu (2014), http://dx.doi.org/10.1017/
S1474748013000200, in press.
[25] A. Ovchinnikov, Differential Tannakian categories, J. Algebra 321 (10) (2009) 3043–3062,
http://dx.doi.org/10.1016/j.jalgebra.2009.008.
[26] M. van der Put, M.F. Singer, Galois Theory of Linear Differential Equations, Grundlehren Math.
Wiss., Springer-Verlag, New York, 2003.
[27] M.F. Singer, Linear algebraic groups as parameterized Picard–Vessiot Galois groups, J. Algebra 373
(2013) 151–161, http://dx.doi.org/10.1016/j.jalgebra.2012.09.037.
[28] W.Y. Sit, Differential algebraic subgroups of SL(2) and strong normality in simple extensions, Amer.
J. Math. 97 (1975) 627–698.
[29] D. Trushin, Splitting fields and general differential Galois theory, Sb. Math. 201 (2010) 1323–1353,
http://dx.doi.org/10.1070/SM2010v201n09ABEH004114.
[30] M. Wibmer, Existence of ∂-parameterized Picard–Vessiot extensions over fields with algebraically
closed constants, J. Algebra 361 (2012) 163–171, http://dx.doi.org/10.1016/j.jalgebra.2012.03.035.