This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

From algebra to analysis: new proofs of
theorems by Ritt and Seidenberg

D. Pavlov Faculty of Mechanics and Mathematics, Moscow State University, Russia, e-mail: [email protected]    G. Pogudin LIX, CNRS, École Polytechnique, Institute Polytechnique de Paris, Palaiseau, France, email: [email protected]    Yu.P. Razmyslov Faculty of Mechanics and Mathematics, Moscow State University, Russia, e-mail: [email protected]
Abstract

Ritt’s theorem of zeroes and Seidenberg’s embedding theorem are classical results in differential algebra allowing to connect algebraic and model-theoretic results on nonlinear PDEs to the realm of analysis. However, the existing proofs of these results use sophisticated tools from constructive algebra (characteristic set theory) and analysis (Riquier’s existence theorem).

In this paper, we give new short proofs for both theorems relying only on basic facts from differential algebra and the classical Cauchy-Kovalevskaya theorem for PDEs.

1 Introduction

The algebraic theory of differential equations, also known as differential algebra [19], aims at studying nonlinear differential equations using methods of algebra and algebraic geometry. For doing this, one typically abstracts from functions (analytic, meromorphic, etc) to elements of differential fields (fields equipped with a derivation or several commuting derivations). This approach turned out to be fruitful yielding interesting results from theoretical and applied perspectives (see, e.g., [5, 2, 21, 16, 6]). Furthermore, one can additionally use powerful tools from model theory to study differential fields (see, e.g., [14, 15]).

In this context, a fundamental question is how to transfer results about differential fields back to the realm of analysis. There are two classical theorems in differential algebra typically used for this purpose:

  • Ritt’s theorem of zeroes [19, p. 176] which can be viewed as an analogue of Hilbert’s Nullstellensatz. The theorem implies that any system of nonlinear PDEs having a solution in some differential field has a solution in a field of meromorphic functions on some domain.

  • Seidenberg’s embedding theorem [20] which is often used as a differential analogue of the Lefschetz principle (e.g. [6, 7, 3, 1, 10]). The theorem says that any countably generated differential field with several commuting derivations can be embedded into a field of meromorphic functions on some domain.

In [20], Seidenberg gave a complete proof of his theorem for the case of a single derivation (see also [14, Appendix A]). For the PDE case, he gave a sketch which reuses substantial parts of Ritt’s proof of Ritt’s zero theorem from [19]. The latter proof concludes the whole monograph and heavily relies on the techniques developed there. In particular, Ritt’s proof uses the machinery of characteristic sets [19, Chapter V] which is a fundamental tool in differential algebra but not so well-known in the broader algebra community and quite technical existence theorem for PDEs due to Riquier [19, Chapter VIII] (see also [18]) which, to the best of our knowledge, is not discussed in the standard PDE textbooks.

Due to the importance of the theorems of Ritt and Siedenberg as bridges between the algebraic and analytic theories of nonlinear PDEs, we think that it is highly desirable to have short proofs of these theorems accessible to people with some general knowledge in algebra and PDEs. In the present paper, we give such proofs. Our proofs rely only on some basic facts from differential algebra and the classical Cauchy-Kovalevskaya theorem for PDEs.

Our proof strategy is inspired by the argument from [8, Theorem 3.1] for the case of one derivation. However, the techniques from [8] had to be substantially developed in order to tackle the PDE case (which is quite subtle [13]) and to prove both Ritt’s and Seidenberg’s theorem (not only the Ritt’s as in [8]). The key ingredients of the argument are an auxiliary change of derivations (Lemma 1) which helps us to bring a system of PDEs into the form as in the Cauchy-Kovalevaskaya theorem, Taylor homomorpishms (Definition 7) allowing to build formal power series solutions, and a characterization of differentially simple algebras (Lemma 5).

The paper is organized as follows. Section 2 contains the basic definitions used to state the main results in Section 3. Section 4 contains relevant notions and facts from algebra and analysis used in the proofs. The proofs are located in Section 5. Section 6 contains a remark on the special case of algebras over \mathbb{C}.

2 Preliminaries

2.1 Algebra

Throughout the paper, all algebras are assumed to be unital (that is, with a multiplicative identity element).

Notation 1 (Multi-indices).

For every α=(α1,,αm)0m\alpha=(\alpha_{1},\ldots,\alpha_{m})\in\mathbb{Z}_{\geqslant 0}^{m} and for every tuple t=(t1,,tm)t=(t_{1},\ldots,t_{m}) of elements of a ring, we denote

tα:=t1α1tαm and α!:=α1!αm!.t^{\alpha}:=t_{1}^{\alpha_{1}}\cdot\ldots\cdot t^{\alpha_{m}}\quad\text{ and }\quad\alpha!:=\alpha_{1}!\cdot\ldots\cdot\alpha_{m}!.
Definition 1 (Differential rings and algebras).

Let Δ={δ1,,δm}\Delta=\{\delta_{1},\ldots,\delta_{m}\} be a set of symbols.

  • Let RR be a commutative ring. An additive map δ:RR\delta\colon R\to R is called derivation if δ(ab)=δ(a)b+aδ(b)\delta(ab)=\delta(a)b+a\delta(b) for any a,bRa,b\in R.

  • A commutative ring RR is called Δ\Delta-ring if δ1,,δm\delta_{1},\ldots,\delta_{m} act on RR as pairwise commuting derivations. If RR is a field, it is called Δ\Delta-field.

  • Let AA be a commutative algebra over ring RR. If AA and RR are Δ\Delta-rings and the action of Δ\Delta on RR coincides with the restriction of the action of Δ\Delta on R1AAR\cdot 1_{A}\subseteq A, then AA is called Δ\Delta-algebra over RR.

Definition 2 (Differential generators).

Let AA be a Δ\Delta-algebra over a Δ\Delta-ring RR. A set SAS\subseteq A is called a set of Δ\Delta-generators of AA over RR if the set

{δαssS,α0m}\{\delta^{\alpha}s\mid s\in S,\;\alpha\in\mathbb{Z}_{\geqslant 0}^{m}\}

of all the derivatives of all the elements of SS generates AA as RR-algebra. A Δ\Delta-algebra is said to be Δ\Delta-finitely generated if it has a finite set of Δ\Delta-generators.

Δ\Delta-generators for Δ\Delta-fields are defined analogously.

Definition 3 (Differential homomorphisms).

Let AA and BB be Δ\Delta-algebras over Δ\Delta-ring RR. A map f:ABf\colon A\rightarrow B is called Δ\Delta-homomorphism if ff is a homomorphism of commutative RR-algebras and f(δa)=δf(a)f(\delta a)=\delta f(a) for all δΔ\delta\in\Delta and aAa\in A. An injective Δ\Delta-homomorphism is called Δ\Delta-embedding.

Definition 4 (Differential algebraicity).

Let AA be a Δ\Delta-algebra over a Δ\Delta-ring RR. An element aAa\in A is said to be Δ\Delta-algebraic over RR if the set {δαaα0m}\{\delta^{\alpha}a\mid\alpha\in\mathbb{Z}_{\geqslant 0}^{m}\} of all the derivatives of aa is algebraically dependent over RR.

In other words, aa satisfies a nonlinear PDE with coefficients in RR.

2.2 Analysis

Definition 5 (Multivariate holomorphic functions).

Let UmU\subseteq\mathbb{C}^{m} be a domain. A function f:Uf:U\rightarrow\mathbb{C} is called a holomorphic function in mm variables on UU if it is holomorphic on UU with respect to each individual variable. The set of all holomorphic functions on UU will be denoted by 𝒪m(U)\mathcal{O}_{m}(U)

Notation 2.

Let ff be a holomorphic function on UmU\subseteq\mathbb{C}^{m}. By V(f)V(f) we denote the set of zeroes of ff.

Definition 6 (Multivariate meromorphic functions, [4, Chapter IV, Definition 2.1]).

Let UmU\subseteq\mathbb{C}^{m} be a domain. A meromorphic function on UU is a pair (f,M)(f,M), where MM is a thin set in UU and f𝒪m(UM)f\in\mathcal{O}_{m}(U\setminus M) with the following property: for every z0Uz_{0}\in U, there is a neighbourhood U0U_{0} of z0z_{0} and there are functions g,h𝒪m(U0)g,h\in\mathcal{O}_{m}(U_{0}), such that V(h)MV(h)\subseteq M and

f(z)=g(z)h(z) for every zU0M.f(z)=\dfrac{g(z)}{h(z)}\leavevmode\nobreak\ \text{ for every }\leavevmode\nobreak\ z\in U_{0}\setminus M.

The set of meromorphic functions on a domain UU is denoted erm(U)\mathcal{M}er_{m}(U). By convention we define er0(U)=𝒪0(U)=\mathcal{M}er_{0}(U)=\mathcal{O}_{0}(U)=\mathbb{C}.

For every domain UmU\subseteq\mathbb{C}^{m}, the field erm(U)\mathcal{M}er_{m}(U) has a natural structure of Δ\Delta-field with δiΔ\delta_{i}\in\Delta acting as zi\frac{\partial}{\partial z_{i}}, where z1,,zmz_{1},\ldots,z_{m} are the coordinates in m\mathbb{C}^{m}. Furthermore, if UVU\subseteq V, then there is a natural Δ\Delta-embedding erm(V)erm(U)\mathcal{M}er_{m}(V)\subseteq\mathcal{M}er_{m}(U).

3 Main Results

Theorem 1 (Seidenberg’s embedding theorem).

Let WmW\subseteq\mathbb{C}^{m} be a domain and let Kerm(W)K\subseteq\mathcal{M}er_{m}(W) be at most countably Δ\Delta-generated Δ\Delta-field (over \mathbb{Q}). Let LKL\supset K be a Δ\Delta-field finitely Δ\Delta-generated over KK.

Then there exists a domain UWU\subseteq W and a Δ\Delta-embedding f:Lerm(U)f\colon L\rightarrow\mathcal{M}er_{m}(U) over KK.

Theorem 2 (Ritt’s theorem of zeroes).

Let WmW\subseteq\mathbb{C}^{m} be a domain and let Kerm(W)K\subseteq\mathcal{M}er_{m}(W) be a Δ\Delta-field. Let AA be a finitely generated Δ\Delta-algebra over KK.

Then there exists a non-trivial Δ\Delta-homomorphism f:Aerm(U)f:A\rightarrow\mathcal{M}er_{m}(U) for some domain UWm{U\subseteq W\subseteq\mathbb{C}^{m}} such that f(a)f(a) is Δ\Delta-algebraic over KK for any aAa\in A.

Corollary 1.

Let AA be a finitely Δ\Delta-generated Δ\Delta-algebra over \mathbb{C}. Then there exists a non-trivial Δ\Delta-homomorphism f:A𝒪m(U)f\colon A\rightarrow\mathcal{O}_{m}(U) for some domain Um{U\subseteq\mathbb{C}^{m}}.

Proof.

Ritt’s theorem yields the existence of a Δ\Delta-homomorphism f:Aerm(W)f\colon A\to\mathcal{M}er_{m}(W). Let a1,,ana_{1},\ldots,a_{n} be a set of Δ\Delta-generators of AA. There is a domain UWU\subseteq W such that f(a1),,f(an)f(a_{1}),\ldots,f(a_{n}) are holomorphic in UU. Therefore, the restriction of ff to UU yields a Δ\Delta-homomorphism A𝒪m(U)A\to\mathcal{O}_{m}(U). ∎

4 Notions and results used in the proofs

4.1 Algebra

Notation 3.

Let RR be a Δ\Delta-ring. By R[[z1,,zm]]R[[z_{1},\ldots,z_{m}]] we denote the ring of formal power series over RR in variables z1,,zmz_{1},\ldots,z_{m}. It has a natural structure of Δ\Delta-algebra over RR with δiΔ\delta_{i}\in\Delta acting as zi\frac{\partial}{\partial z_{i}}.

Definition 7 (Taylor homomorphisms).

Let AA be a Δ\Delta-algebra over Δ\Delta-field KK, LKL\supseteq K be a Δ\Delta-field and the action of Δ\Delta on LL be trivial. Let ψ:AL\psi\colon A\to L be a (not necessarily differential) homomorphism of KK-algebras. Let wLmw\in L^{m}. Then we define a map called Taylor homomorphism Tψ,w:AL[[t1,,tm]]T_{\psi,w}\colon A\to L[[t_{1},\ldots,t_{m}]] by the formula

Tψ,w(a):=α0mψ(δαa)(tw)αα! for every aA.T_{\psi,w}(a):=\sum\limits_{\alpha\in\mathbb{Z}^{m}_{\geqslant 0}}\psi(\delta^{\alpha}a)\dfrac{(t-w)^{\alpha}}{\alpha!}\quad\text{ for every }a\in A.

Direct computation shows [17, §44.3] that that Tψ,wT_{\psi,w} is a Δ\Delta-homomorphism.

Notation 4.

Let RR be a Δ\Delta-ring. For every subset SRS\subseteq R, by ΔS\Delta^{\infty}S we denote the set {δαs|α0m,sS}\{\delta^{\alpha}s|\alpha\in\mathbb{Z}^{m}_{\geqslant 0},s\in S\} of all derivatives of the elements of SS.

Definition 8 (Differential polynomials).

Let RR be a Δ\Delta-ring. Consider an algebra of polynomials

R[Δx1,,Δxn]:=R[δαxi|α0m,i=1,,n]R[\Delta^{\infty}x_{1},\ldots,\Delta^{\infty}x_{n}]:=R[\delta^{\alpha}x_{i}|\alpha\in\mathbb{Z}^{m}_{\geqslant 0},i=1,\ldots,n]

in infinitely many variables δαxi\delta^{\alpha}x_{i}. We define the structure of Δ\Delta-algebra over RR by

δi(δαxj):=(δiδα)xj for every 1im, 1jn,α0m.\delta_{i}(\delta^{\alpha}x_{j}):=(\delta_{i}\delta^{\alpha})x_{j}\text{ for every }1\leqslant i\leqslant m,\;1\leqslant j\leqslant n,\;\alpha\in\mathbb{Z}_{\geqslant 0}^{m}.

The resulting algebra is called the the algebra of Δ\Delta-polynomials in x1,,xnx_{1},\ldots,x_{n} over RR.

Definition 9 (Separants).

Let RR be a Δ\Delta-ring. Let P(x)R[Δx]P(x)\in R[\Delta^{\infty}x]. We introduce an ordering on the derivatives of xx as follows:

δαx<δβxα<grlexβ,\delta^{\alpha}x<\delta^{\beta}x\iff\alpha<_{\operatorname{grlex}}\beta, (1)

where grlex\operatorname{grlex} is the graded lexicographic ordering of 0m\mathbb{Z}^{m}_{\geqslant 0}. Let δμx\delta^{\mu}x be the highest (w.r.t. the introduced ordering) derivative appearing in PP. Consider PP as a univariate polynomial in δμx\delta^{\mu}x over R[δαx|α<grlexμ]R[\delta^{\alpha}x|\alpha<_{\operatorname{grlex}}\mu]. We define the separant of PP by

sepxΔ(P):=(δμx)P.\operatorname{sep}_{x}^{\Delta}(P):=\frac{\partial}{\partial(\delta^{\mu}x)}P.
Remark 1.

Throughout the rest of the paper, we assume that the ordering of a set of derivatives of an element of a Δ\Delta-algebra is the one defined in (1).

Definition 10 (Differential algebraicity and transcendence).

Let RR be a Δ\Delta-ring and let AA be a Δ\Delta-algebra over RR.

  • A subset SAS\subseteq A is said to be Δ\Delta-dependent over RR if ΔS\Delta^{\infty}S is algebraically dependent over RR. Otherwise, SS is called Δ\Delta-independent over RR.

  • An element aAa\in A is said to be Δ\Delta-algebraic over RR if the set {a}\{a\} is Δ\Delta-dependent over RR. Otherwise, aa is called Δ\Delta-transcendental over RR.

Definition 11 (Differential transcendence degree).

Let AA be a Δ\Delta-algebra over field KK. Any maximal Δ\Delta-independent over KK subset of AA is called a Δ\Delta-transcendence basis of AA over KK. The cardinality of a Δ\Delta-transcendence basis does not depend on the choice of the basis [11, II.9, Theorem 4] and is called the Δ\Delta-transcendence degree of AA over KK (denoted by difftrdegKΔA\operatorname{difftrdeg}_{K}^{\Delta}A).

Definition 12 (Differential ideals).

Let RR be a Δ\Delta-ring. A subset IRI\subseteq R is called a differential ideal if it is an ideal of AA considered as a commutative algebra and δaI\delta a\in I for any δΔ\delta\in\Delta and aIa\in I.

Notation 5.

Throughout the rest of the paper, we use the notation Δ0:=Δ{δ1}\Delta_{0}:=\Delta\setminus\{\delta_{1}\}.

4.2 Analysis

The following is a special case of the Cauchy-Kovalevskaya theorem [9, Chapter V, §94] which is sufficient for our purposes.

Theorem 3 (Cauchy-Kovalevskaya).

Consider holomorphic functions in variables z1,,zmz_{1},\ldots,z_{m}. The operator of differentiation with respect to ziz_{i} will be denoted by δi\delta_{i} for i=1,,mi=1,\ldots,m. For a positive integer rr, we introduce a set of multi-indices Mr:={α0m|α|r,α1<r}M_{r}:=\{\alpha\in\mathbb{Z}_{\geqslant 0}^{m}\mid|\alpha|\leqslant r,\alpha_{1}<r\}. Consider a PDE in an unknown function uu

δ1ru=F(z1,,zm;δαuαMr),\delta_{1}^{r}u=F(z_{1},\ldots,z_{m};\delta^{\alpha}u\mid\alpha\in M_{r}), (2)

where FF is a rational function over \mathbb{C} in z1,,zmz_{1},\ldots,z_{m} and derivatives {δαuαMr}\{\delta^{\alpha}u\mid\alpha\in M_{r}\}.

Consider complex numbers a1,,ama_{1},\ldots,a_{m} and functions φ0,,φr1\varphi_{0},\ldots,\varphi_{r-1} in variables z2,,zmz_{2},\ldots,z_{m} holomorphic in a neighbourhood of (a2,,am)(a_{2},\ldots,a_{m}) such that FF is well-defined under the substitution:

  1. 1.

    aia_{i} for ziz_{i} for every 1im1\leqslant i\leqslant m

  2. 2.

    and (δ(α2,,αm)φα1)(a2,,am)(\delta^{(\alpha_{2},\ldots,\alpha_{m})}\varphi_{\alpha_{1}})(a_{2},\ldots,a_{m}) for δαu\delta^{\alpha}u for every αMr\alpha\in M_{r}.

Then there is a unique function uu holomorphic in a neighborhood of (a1,,am)(a_{1},\ldots,a_{m}) satisfying (2) and

(δ1iu)|z1=a1=φi for every 0i<r.(\delta_{1}^{i}u)|_{z_{1}=a_{1}}=\varphi_{i}\quad\text{ for every }\quad 0\leqslant i<r.

5 Proofs

This section is structured as follows. In Section 5.1, we introduce the notion of Δ\Delta-integral elements which is an algebraic way saying that an element satisfies a PDE as in the Cauchy-Kovalevskaya theorem. We prove that there always exists a linear change of derivations making a fixed element Δ\Delta-integral (Lemma 1) and prove Lemma 2 which is a key tool for reducing the problem to the same problem in fewer derivations.

Section 5.2 contains the proof of Seidenberg’s embedding theorem which proceeds by induction on the number of derivations using Lemma 2. We deduce Ritt’s theorem of zeroes in Section 5.3 from Seidenberg’s theorem and Lemma 5 characterizing Δ\Delta-simple algebras.

5.1 Differentially integral generators

Definition 13 (Δ\Delta-integral elements).

Let RR be a Δ\Delta-ring and let AA be a Δ\Delta-algebra over RR. An element aAa\in A is said to be Δ\Delta-integral over RR if there exists P(x)R[Δx]P(x)\in R[\Delta^{\infty}x] such that

  • P(a)=0P(a)=0, sepxΔ(P)(a)0\operatorname{sep}_{x}^{\Delta}(P)(a)\neq 0;

  • the highest (w.r.t. the ordering (1)) derivative in PP is of the form δ1rx\delta_{1}^{r}x.

Remark 2.

If aAa\in A is Δ\Delta-integral over RR, then the equality δ1(P(a))=0\delta_{1}(P(a))=0 can be rewritten as

sepxΔP(a)δ1r+1a=q(a),where qR[δαxα<grlex(r+1,0,,0)].\operatorname{sep}_{x}^{\Delta}P(a)\cdot\delta_{1}^{r+1}a=q(a),\quad\text{where }q\in R[\delta^{\alpha}x\mid\alpha<_{\operatorname{grlex}}(r+1,0,\ldots,0)].

Therefore, if sepxΔP(a)\operatorname{sep}_{x}^{\Delta}P(a) is invertible in AA, we have δ1r+1a=q(a)sepxΔ(P)(a)\delta_{1}^{r+1}a=\dfrac{q(a)}{\operatorname{sep}_{x}^{\Delta}(P)(a)}.

Lemma 1.

Let RR be a Δ\Delta-ring and let AA be a Δ\Delta-algebra over RR. Let AA be Δ\Delta-generated over RR by Δ\Delta-algebraic over RR elements a1,,ana_{1},\ldots,a_{n}. Then there exists an invertible \mathbb{Z}-linear change of derivations transforming Δ\Delta to Δ\Delta^{\ast} such that a1,,ana_{1},\ldots,a_{n} are Δ\Delta^{\ast}-integral over RR.

Proof.

Fix 1in1\leqslant i\leqslant n. Since aia_{i} is Δ\Delta-algebraic over RR, there exists nonzero fiR[Δx]f_{i}\in R[\Delta^{\infty}x] such that fi(ai)=0f_{i}(a_{i})=0. We will choose this fif_{i} so that its highest (w.r.t. (1)) derivative is minimal and, among such polynomials, the degree is minimal. We will call such fif_{i} a minimal polynomial for aia_{i}.

We introduce variables λ2,,λm\lambda_{2},\ldots,\lambda_{m} algebraically independent over AA and extend the derivations from AA to A[λ2,,λm]A[\lambda_{2},\ldots,\lambda_{m}] by δiλj=0\delta_{i}\lambda_{j}=0 for all i=1,,mi=1,\ldots,m and j=2,,mj=2,\ldots,m. Consider a set of derivations D:={d1,d2,,dm}D:=\{\operatorname{d}_{1},\operatorname{d}_{2},\ldots,\operatorname{d}_{m}\} defined by

d1:=δ1,dj:=δj+λiδ1 for j=2,,m.\operatorname{d}_{1}:=\delta_{1},\quad\operatorname{d}_{j}:=\delta_{j}+\lambda_{i}\delta_{1}\text{ for }j=2,\ldots,m.

Consider any 1in1\leqslant i\leqslant n. We rewrite fif_{i} in terms of DD replacing δ1\delta_{1} with d1\operatorname{d}_{1} and δj\delta_{j} with djλid1\operatorname{d}_{j}-\lambda_{i}\operatorname{d}_{1} for j=2,,mj=2,\ldots,m. We denote the order of the highest derivative appearing in fif_{i} by rir_{i} and the partial derivative (d1rix)fi\frac{\partial}{\partial(\operatorname{d}_{1}^{r_{i}}x)}f_{i} by sis_{i}. We will show that si(ai)0s_{i}(a_{i})\neq 0. We write

si(x)=fi(d1rix)=q1++qm=riλ2q2λmqmfi(δ1q1δmqmx).s_{i}(x)=\dfrac{\partial f_{i}}{\partial(\operatorname{d}_{1}^{r_{i}}x)}=\sum\limits_{q_{1}+\ldots+q_{m}=r_{i}}\lambda_{2}^{q_{2}}\ldots\lambda_{m}^{q_{m}}\dfrac{\partial f_{i}}{\partial(\delta_{1}^{q_{1}}\ldots\delta_{m}^{q_{m}}x)}.

Due to the minimality of fif_{i} as a vanishing polynomial of aa and the algebraic independence of λj\lambda_{j}, the latter expression does not vanish at x=aix=a_{i}. So, si(ai)0s_{i}(a_{i})\neq 0.

Since, for every 1in1\leqslant i\leqslant n, si(ai)s_{i}(a_{i}) is a nonzero polynomial in λ2,,λm\lambda_{2},\ldots,\lambda_{m} over AA, it is possible to choose the values λ2,,λmR\lambda^{\ast}_{2},\ldots,\lambda_{m}^{\ast}\in\mathbb{Z}\subset R so that neither of si(ai)s_{i}(a_{i}) vanishes at (λ2,,λm)(\lambda_{2}^{\ast},\ldots,\lambda_{m}^{\ast}). Let Δ={δ1,,δm}\Delta^{\ast}=\{\delta_{1}^{\ast},\ldots,\delta_{m}^{\ast}\} be the result of plugging these values to DD. Then we have sepxΔf(ai)=fi((δ1)rix)(ai)0\operatorname{sep}_{x}^{\Delta^{\ast}}f(a_{i})=\dfrac{\partial f_{i}}{\partial((\delta_{1}^{\ast})^{r_{i}}x)}(a_{i})\neq 0 for every i=1,,ni=1,\ldots,n, so a1,,ana_{1},\ldots,a_{n} are Δ\Delta^{\ast}-integral over RR. ∎

Lemma 2.

Let RR be a Δ\Delta-ring and let AA be a Δ\Delta-algebra over RR. Assume that AA is a domain and is Δ\Delta-generated by a1,,ana_{1},\ldots,a_{n} over RR which are Δ\Delta-integral over RR.

Then there exists aAa\in A such that A[1/a]A[1/a] is finitely Δ0\Delta_{0}-generated over RR.

Proof.

We will prove the lemma by induction on the number nn of Δ\Delta-generators of AA. If n=0n=0, then A=RA=R and AA is clearly finitely Δ0\Delta_{0}-generated.

Assume that the lemma is proved for all extensions Δ\Delta-generated by less than nn elements. Applying the induction hypothesis to Δ\Delta-algebra A0:=R[Δa1,,Δan1]A_{0}:=R[\Delta^{\infty}a_{1},\ldots,\Delta^{\infty}a_{n-1}], we obtain b0A0b_{0}\in A_{0} such that A0[1/b0]A_{0}[1/b_{0}] is a finitely Δ0\Delta_{0}-generated RR-algebra.

Since ana_{n} is Δ\Delta-integral over RR, there exists P(x)R[Δx]P(x)\in R[\Delta^{\infty}x] such that P(an)=0P(a_{n})=0, the highest derivative in PP is δ1rx\delta_{1}^{r}x, and b2:=sepxΔ(P)(an)0b_{2}:=\operatorname{sep}^{\Delta}_{x}(P)(a_{n})\neq 0. We claim that

A[1b1b2]=A0[1b1,Δ0(1b2),Δ0(δ1ran)],A\left[\frac{1}{b_{1}b_{2}}\right]=A_{0}\left[\frac{1}{b_{1}},\Delta_{0}^{\infty}\left(\frac{1}{b_{2}}\right),\Delta_{0}^{\infty}(\delta_{1}^{\leqslant r}a_{n})\right], (3)

where δ1ran:={an,δ1an,,δnran}\delta_{1}^{\leqslant r}a_{n}:=\{a_{n},\delta_{1}a_{n},\ldots,\delta_{n}^{r}a_{n}\}. Since A0[1/b1]A_{0}[1/b_{1}] is finitely Δ0\Delta_{0}-generated over RR, this would imply that A[1/(b1b2)]A[1/(b_{1}b_{2})] is finitely Δ0\Delta_{0}-generated over RR as well.

In order to prove (3), it is sufficient to show that the images of {δ1ran,1/b2}\{\delta_{1}^{\leqslant r}a_{n},1/b_{2}\} under δ1\delta_{1} belong to B:=A0[1/b1,Δ0(1/b2),Δ0(δ1ran)]B:=A_{0}[1/b_{1},\Delta_{0}^{\infty}(1/b_{2}),\Delta_{0}^{\infty}(\delta_{1}^{\leqslant r}a_{n})]. This is clear for δ1<ran\delta_{1}^{<r}a_{n}, so it remains to show that δ1r+1an,δ1(1/b2)B\delta_{1}^{r+1}a_{n},\delta_{1}(1/b_{2})\in B:

  • For δ1r+1an\delta_{1}^{r+1}a_{n} we use Remark 2 to write

    δ1r+1an=q(an)sepxΔ(P)(an)=q(an)b2B, where qR[Δ0(δ1rx)].\delta_{1}^{r+1}a_{n}=\dfrac{-q(a_{n})}{\operatorname{sep}^{\Delta}_{x}(P)(a_{n})}=\dfrac{-q(a_{n})}{b_{2}}\in B,\text{ where }q\in R[\Delta_{0}^{\infty}(\delta_{1}^{\leqslant r}x)].
  • For δ1(1/b2)\delta_{1}(1/b_{2}), we observe that

    δ1(1b2)1b22A0[Δ0(δ1ran)]B.\delta_{1}\left(\dfrac{1}{b_{2}}\right)\in\dfrac{1}{b_{2}^{2}}A_{0}[\Delta_{0}^{\infty}(\delta_{1}^{\leqslant r}a_{n})]\subseteq B.\qed

5.2 Proof of Seidenberg’s Theorem

Lemma 3.

Let WmW\subseteq\mathbb{C}^{m} be a domain and KK be a countably Δ\Delta-generated subfield of erm(W)\mathcal{M}er_{m}(W). Then there exist cc\in\mathbb{C} and a domain VW{z1=c}V\subseteq W\cap\{z_{1}=c\} such that, for every fKf\in K, f|{z1=c}f|_{\{z_{1}=c\}} is a well-defined element of erm1(V)\mathcal{M}er_{m-1}(V) and, therefore, the restriction to {z1=c}\{z_{1}=c\} defines a Δ0\Delta_{0}-embedding Kerm1(V)K\to\mathcal{M}er_{m-1}(V).

Proof.

Let KK be Δ\Delta-generated by {bi}i=1\{b_{i}\}_{i=1}^{\infty}. For every i0i\geqslant 0, denote by SiS_{i} the set of singularities of bib_{i}. By definition, SiS_{i} is a nowhere dense subset of m\mathbb{C}^{m}. Therefore, the union S=i=1SiS=\bigcup\limits_{i=1}^{\infty}S_{i} is a meagre set. As WW is a domain in m\mathbb{C}^{m}, the difference WSW\setminus S is non-empty. Choose any point (w1,,wm)WS(w_{1},\ldots,w_{m})\in W\setminus S. Then all the restrictions of bib_{i} to {z1=w1}\{z_{1}=w_{1}\} are holomorphic at (w2,,wm)(w_{2},\ldots,w_{m}) and meromorphic in some vicinity VW{t1=w1}V\subseteq W\cap\{t_{1}=w_{1}\} of (w2,,wm)(w_{2},\ldots,w_{m}). Since every element fKf\in K is a rational function in bib_{i}’s and their partial derivatives, its restriction to {z1=w1}\{z_{1}=w_{1}\} is also a well-defined meromorphic function on VV. ∎

Lemma 4.

Let UmU\subseteq\mathbb{C}^{m} be a domain. For every countably Δ\Delta-generated Δ\Delta-field Kerm(U)K\subseteq\mathcal{M}er_{m}(U), difftrdegKΔerm(U)\operatorname{difftrdeg}_{K}^{\Delta}\mathcal{M}er_{m}(U) is infinite.

Proof.

Suppose difftrdegKΔerm(U)=l<\operatorname{difftrdeg}_{K}^{\Delta}\mathcal{M}er_{m}(U)=l<\infty. Let τ1,,τl\tau_{1},\ldots,\tau_{l} be a Δ\Delta-transcendence basis of erm(U)\mathcal{M}er_{m}(U) over KK. Let LL be a field Δ\Delta-generated by KK and τ1,,τl\tau_{1},\ldots,\tau_{l}. Note that LL is still at most countably Δ\Delta-generated and erm(U)\mathcal{M}er_{m}(U) is Δ\Delta-algebraic over LL. Choose an arbitrary point cUc\in U and denote by FF a subfield of \mathbb{C} generated by the values at cc of those elements of LL that are holomorphic at cc. Clearly FF is at most countably generated and the transcendence degree of \mathbb{C} over FF is infinite. Now any function in erm(U)\mathcal{M}er_{m}(U) that is holomorphic at cc and such that the set of values of its derivatives at cc is transcendental over FF is Δ\Delta-transcendental over LL, which contradicts to the assumption that erm(U)\mathcal{M}er_{m}(U) is Δ\Delta-algebraic over LL. ∎

Notation 6.

Let AA be a Δ\Delta-algebra without zero divisors. By Frac(A)\operatorname{Frac}(A) we denote the field of fractions of AA.

We are now ready to prove Seidenberg’s theorem.

Theorem 1 (Seidenberg’s embedding theorem).

Let WmW\subseteq\mathbb{C}^{m} be a domain and let Kerm(W)K\subseteq\mathcal{M}er_{m}(W) be at most countably Δ\Delta-generated Δ\Delta-field (over \mathbb{Q}). Let LKL\supset K be a Δ\Delta-field finitely Δ\Delta-generated over KK.

Then there exists a domain UWU\subseteq W and a Δ\Delta-embedding f:Lerm(U)f\colon L\rightarrow\mathcal{M}er_{m}(U) over KK.

Proof.

We will first reduce the theorem to the case when LL is Δ\Delta-algebraic over KK. Assume that it is not, and let u1,,uu_{1},\ldots,u_{\ell} be a Δ\Delta-transcendence basis of LL over KK. Lemma 4 implies that there exist functions f1,,ferm(W)f_{1},\ldots,f_{\ell}\in\mathcal{M}er_{m}(W) Δ\Delta-transcendental over KK. Let K~\tilde{K} be a Δ\Delta-field generated by KK and f1,,ff_{1},\ldots,f_{\ell}. The embedding KLK\to L can be extended to an embedding K~L\tilde{K}\to L by sending fif_{i} to uiu_{i} for every 1i1\leqslant i\leqslant\ell. Therefore, by replacing KK with K~\tilde{K} we will further assume that LL is Δ\Delta-algebraic over KK.

We will prove the theorem by induction on the number mm of derivations. If m=0m=0, then LL can be embedded into \mathbb{C} by [12, Chapter V, Theorem 2.8].

Suppose m>0m>0. Let a1,,ana_{1},\ldots,a_{n} be a set of Δ\Delta-generators of LL over KK. Let A:=K[Δa1,,Δan]A:=K[\Delta^{\infty}a_{1},\ldots,\Delta^{\infty}a_{n}]. Since AA is Δ\Delta-algebraic over KK, by Lemma 1, there exist and invertible m×mm\times m matrix MM over \mathbb{Q} such that, for a new set of derivations

Δ={δ1,,δm}:=MΔ,\Delta^{\ast}=\{\delta_{1}^{\ast},\ldots,\delta_{m}^{\ast}\}:=M\Delta,

a1,,ana_{1},\ldots,a_{n} are Δ\Delta^{\ast}-integral over KK. Due to the invertibility of MM, every Δ\Delta^{\ast}-embedding Lerm(U)L\to\mathcal{M}er_{m}(U) over KK yields a Δ\Delta-embedding. Therefore, by changing the coordinates in the space m\mathbb{C}^{m} from (z1,,zm)(z_{1},\ldots,z_{m}) to M1(z1,,zm)M^{-1}(z_{1},\ldots,z_{m}), we can further assume that Δ=Δ\Delta=\Delta^{\ast}, so a1,,ana_{1},\ldots,a_{n} are Δ\Delta-integral over KK.

Lemma 2 implies that there exists aAa\in A such that B:=A[a1]B:=A[a^{-1}] is finitely Δ0\Delta_{0}-generated. By Δ\Delta-integrality of a1,,ana_{1},\ldots,a_{n} and Remark 2, for every 1in1\leqslant i\leqslant n, there exists a positive integer rir_{i} and a rational function giK(δαyα<grlex(r,0,,0))g_{i}\in K(\delta^{\alpha}y\mid\alpha<_{\operatorname{grlex}}(r,0,\ldots,0)) such that aia_{i} satisfies

δriai=gi(ai).\delta^{r_{i}}a_{i}=g_{i}(a_{i}). (4)

Since KK is at most countably Δ\Delta-generated, Lemma 3 implies that there exist w1w_{1}\in\mathbb{C} and VW{z1=w1}m1V\subseteq W\cap\{z_{1}=w_{1}\}\subseteq\mathbb{C}^{m-1} such that the restriction to {z1=w1}\{z_{1}=w_{1}\} induces a Δ0\Delta_{0}-embedding ρ:Kerm1(V)\rho\colon K\to\mathcal{M}er_{m-1}(V). We apply the induction hypothesis to Δ0\Delta_{0}-fields ρ(K)\rho(K) and Frac(B)=L\operatorname{Frac}(B)=L. This yields Δ0\Delta_{0}-embedding h:Lerm1(V~)h:L\rightarrow\mathcal{M}er_{m-1}(\widetilde{V}) for some V~V\widetilde{V}\subseteq V.

Choose a point v=(w2,,wm)V~v=(w_{2},\ldots,w_{m})\in\widetilde{V} such that all the h(ai)h(a_{i}) are holomorphic at vv and all the gi(ai)g_{i}(a_{i}) are holomorphic at w=(w1,w2,,wm)Ww=(w_{1},w_{2},\ldots,w_{m})\in W. Consider the Taylor homomorphism Th,w:Aerm1(V~)[[z1]]T_{h,w}\colon A\rightarrow\mathcal{M}er_{m-1}(\widetilde{V})[[z_{1}]] defined as follows (see Definition 7):

Th,w(a):=k=0h(δ1ka)(z1w1)kk! for every aA.T_{h,w}(a):=\sum\limits_{k=0}^{\infty}h(\delta_{1}^{k}a)\dfrac{(z_{1}-w_{1})^{k}}{k!}\quad\text{ for every }a\in A.

Note that Th,wT_{h,w} is a Δ\Delta-homomorphism.

Fix 1in1\leqslant i\leqslant n. Since a1a_{1} is a solution of (4) and Th,wT_{h,w} is a Δ\Delta-homomorphism, Th,wT_{h,w} is a formal power series solution of δ1riy=gi(y)\delta_{1}^{r_{i}}y=g_{i}(y) corresponding to holomorphic initial conditions

y|z1=w1=h(ai),(δ1y)|z1=w1=h(δ1ai),,(δ1ri1y)|z1=w1=h(δ1ri1ai).y|_{z_{1}=w_{1}}=h(a_{i}),\;(\delta_{1}y)|_{z_{1}=w_{1}}=h(\delta_{1}a_{i}),\;\ldots,\;(\delta_{1}^{r_{i}-1}y)|_{z_{1}=w_{1}}=h(\delta_{1}^{r_{i}-1}a_{i}).

By the Cauchy-Kovalevskaya theorem, this solution is holomorphic in some vicinity UiU_{i} of ww. We set U:=i=1nUiU:=\bigcap_{i=1}^{n}U_{i}. Thus, Th,wT_{h,w} induces a non-trivial Δ\Delta-homomorphism from AA to erm(U)\mathcal{M}er_{m}(U). Since hh is injective, Th,wT_{h,w} is also injective, so it can be extended to a Δ\Delta-embedding Lerm(U)L\to\mathcal{M}er_{m}(U) over KK.

5.3 Proof of Ritt’s theorem

Definition 14 (Differentially simple rings).

Δ\Delta-ring RR is called Δ\Delta-simple if it contains no proper Δ\Delta-ideals.

Lemma 5.

Let AA be a Δ\Delta-simple Δ\Delta-algebra Δ\Delta-generated by a1,,ana_{1},\ldots,a_{n} over a Δ\Delta-field K. Then AA does not contain zero divisors.

Furthermore, assume that there exists an integer \ell such that

  • a1,,aa_{1},\ldots,a_{\ell} form a Δ\Delta-transcendence basis of AA over KK;

  • a+1,,ana_{\ell+1},\ldots,a_{n} are Δ\Delta-integral over K[Δa1,,Δa]K[\Delta^{\infty}a_{1},\ldots,\Delta^{\infty}a_{\ell}].

Then AA has finite Δ0\Delta_{0}-transcendence degree over KK. In particular, =0\ell=0.

Proof.

Consider any non-zero (not necessarily differential) homomorphism ψ:AF\psi:A\rightarrow F (FKF\supseteq K is a field) and the corresponding Taylor homomorphism Tψ,0:AF[[z1,,zm]]T_{\psi,0}\colon A\rightarrow F[[z_{1},\ldots,z_{m}]], which is a Δ\Delta-homomorphism. Since AA is Δ\Delta-simple, the kernel of Tψ,0T_{\psi,0} is zero. Therefore, Tψ,0T_{\psi,0} is a Δ\Delta-embedding of AA into F[[z1,,zm]]F[[z_{1},\ldots,z_{m}]]. Since the latter does not contain zero divisors, the same is true for AA.

Assume that AA has infinite Δ0\Delta_{0}-transcendence degree, that is, >0\ell>0. Since a+1,,ana_{\ell+1},\ldots,a_{n} are Δ\Delta-integral over R:=K[Δa1,,Δa]R:=K[\Delta^{\infty}a_{1},\ldots,\Delta^{\infty}a_{\ell}], Lemma 2 implies that there exists an element bAb\in A such that A0:=A[1/b]A_{0}:=A[1/b] is a finitely Δ0\Delta_{0}-generated algebra over RR. Note that A0A_{0} is also Δ\Delta-simple. Let A0=R[Δ0b1,,Δ0bs]A_{0}=R[\Delta_{0}^{\infty}b_{1},\ldots,\Delta_{0}^{\infty}b_{s}]. For every j0j\geqslant 0, consider Δ0\Delta_{0}-algebra

Bj:=K[Δ0(δ1(<j)a1),,Δ0(δ1(<j)al),Δ0b1,,Δ0bs].B_{j}:=K[\Delta_{0}^{\infty}(\delta_{1}^{(<j)}a_{1}),\ldots,\Delta_{0}^{\infty}(\delta_{1}^{(<j)}a_{l}),\Delta_{0}^{\infty}b_{1},\ldots,\Delta_{0}^{\infty}b_{s}].

For every j0j\geqslant 0, we have

jldifftrdegKΔ0Bjjl+s.jl\leqslant\operatorname{difftrdeg}^{\Delta_{0}}_{K}B_{j}\leqslant jl+s.

This inequality implies that there exists NN such that, for every j>Nj>Nδ1ja1,,δ1jal\delta_{1}^{j}a_{1},\ldots,\delta_{1}^{j}a_{l} are Δ0\Delta_{0}-independent over BjB_{j}. Consider any non-zero Δ0\Delta_{0}-homomorphism

φ~:BNL,\tilde{\varphi}\colon B_{N}\rightarrow L,

where LKL\supseteq K is a Δ0\Delta_{0}-field. Due to the Δ0\Delta_{0}-independence of the rest of the elements δ1jai\delta^{j}_{1}a_{i} for 1il1\leqslant i\leqslant l and j>Nj>N, φ~\tilde{\varphi} can be extended to a homomorphism φ:A0L\varphi\colon A_{0}\rightarrow L so that φ(δ1jai)=0\varphi(\delta_{1}^{j}a_{i})=0 for every 1il1\leqslant i\leqslant l and j>Nj>N.

Consider a Taylor homomorphism Tφ,0:A0L[[z]]T_{\varphi,0}\colon A_{0}\to L[[z]] with respect to δ1\delta_{1}. Since φ\varphi was a Δ0\Delta_{0}-homomorphism, Tφ,0T_{\varphi,0} is a Δ\Delta-homomorphism. It remains to observe that the kernel of Tφ,0T_{\varphi,0} contains δ1N+1a1,,δ1N+1al\delta_{1}^{N+1}a_{1},\ldots,\delta_{1}^{N+1}a_{l} contradicting to the fact that A0A_{0} is Δ\Delta-simple. ∎

Theorem 2 (Ritt’s theorem of zeroes).

Let WmW\subseteq\mathbb{C}^{m} be a domain and let Kerm(W)K\subseteq\mathcal{M}er_{m}(W) be a Δ\Delta-field. Let AA be a finitely generated Δ\Delta-algebra over KK.

Then there exists a non-trivial Δ\Delta-homomorphism f:Aerm(U)f:A\rightarrow\mathcal{M}er_{m}(U) for some domain UWm{U\subseteq W\subseteq\mathbb{C}^{m}} such that f(a)f(a) is Δ\Delta-algebraic over KK for any aAa\in A.

Proof.

We can represent AA as A=R/JA=R/J, where R:=K[Δx1,,Δxn]R:=K[\Delta^{\infty}x_{1},\ldots,\Delta^{\infty}x_{n}] and JRJ\subseteq R is a differential ideal. Since RR is a countable-dimensional KK-space, JJ can be generated by at most countable set of generators. Pick any such set and denote the Δ\Delta-field generated by the coefficients of the generators by K0K_{0}. Then K0K_{0} is countably Δ\Delta-generated. Let R0:=K0[Δx1,,Δxn]R_{0}:=K_{0}[\Delta^{\infty}x_{1},\ldots,\Delta^{\infty}x_{n}]. Since JJ is defined over K0K_{0}, for A0:=R0/(JR0)A_{0}:=R_{0}/(J\cap R_{0}), we have A=KK0A0A=K\otimes_{K_{0}}A_{0}.

Let II be a maximal differential ideal of A0A_{0}, and consider the canonical projection π:A0A0/I\pi\colon A_{0}\rightarrow A_{0}/I. Let a1,,ana_{1},\ldots,a_{n} be a set of Δ\Delta-generators of A0/IA_{0}/I. Since A0/IA_{0}/I is differentially simple, by Lemma 5, A0/IA_{0}/I does not have zero divisors and is Δ\Delta-algebraic over K0K_{0}. We apply Theorem 1 to the Δ\Delta-fields K0Frac(A0/I)K_{0}\subseteq\operatorname{Frac}(A_{0}/I) and obtain a Δ\Delta-embedding h:A0/Ierm(U)h\colon A_{0}/I\rightarrow\mathcal{M}er_{m}(U) over KK. Since A0/IA_{0}/I is Δ\Delta-algebraic over K0K_{0}h(a)h(a) is also Δ\Delta-algebraic over K0K_{0} for any aA0/Ia\in A_{0}/I. Let f0:=hπf_{0}:=h\circ\pi, then f0(a)f_{0}(a) is Δ\Delta-algebraic over K0K_{0} for any aA0a\in A_{0}. Since Kerm(U)K\subseteq\mathcal{M}er_{m}(U), we can construct a nontrivial Δ\Delta-homomorphism f:KK0A0erm(U)f\colon K\otimes_{K_{0}}A_{0}\to\mathcal{M}er_{m}(U) as the tensor product of the embedding Kerm(U)K\to\mathcal{M}er_{m}(U) and ff. The Δ\Delta-algebraicity of the image of f0f_{0} over K0K_{0} implies the Δ\Delta-algebraicity of the image of the image of ff over KK. ∎

6 Remarks on the analytic spectrum

Definition 15 (Analytic spectrum).

Consider \mathbb{C} as a Δ\Delta-field with the zero derivations. Let AA be a finitely Δ\Delta-generated Δ\Delta-algebra over \mathbb{C}. A homomorphism (not necessarily differential) ψ:A\psi\colon A\to\mathbb{C} of \mathbb{C}-algebras is called analytic if, for every aAa\in A, the formal power series Tψ,0(a)T_{\psi,0}(a) has a positive radius of convergence. The set of the kernels of analytic \mathbb{C}-homomorphisms is called the analytic spectrum of AA.

Corollary 1 implies the following.

Corollary 2.

Let AA be a finitely generated  Δ\Delta-algebra with identity over \mathbb{C}. Then the analytic spectrum of AA is a Zarisky-dense subset of its maximal spectrum.

Proof.

Assume that analytic spectrum of AA is not Zarisky-dense in specA\operatorname{spec}A. Then it is contained in some maximal proper closed subset F={MspecA|aM}F=\{M\in\operatorname{spec}A|a\in M\} for some non-nilpotent aAa\in A.

Consider the localization A[a1]A[a^{-1}] and a non-trivial Δ\Delta-homomorphism f:A[a1]𝒪m(U)f\colon A[a^{-1}]\rightarrow\mathcal{O}_{m}(U) given by Corollary 1. Then f(a)0f(a)\neq 0. We fix uUu\in U such that f(a)(u)0f(a)(u)\neq 0 and consider a homomorphism g:Ag\colon A\rightarrow\mathbb{C} defined by g(b):=f(b)(u)g(b):=f(b)(u) for every bAb\in A. Note that I:=Ker(g)I:=\operatorname{Ker}(g) is an analytic ideal such that aIa\not\in I, so we have arrived at the contradiction. ∎

Acknowledgements

GP was partially supported by NSF grants DMS-1853482, DMS-1760448, and DMS-1853650 and by the Paris Ile-de-France region.

References

  • Binyamini [2017] G. Binyamini. Bezout-type theorems for differential fields. Compositio Mathematica, 153(4):867–888, 2017. URL https://doi.org/10.1112/s0010437x17007035.
  • [2] F. Boulier and F. Lemaire. Differential algebra and system modeling in cellular biology. In Algebraic Biology, pages 22–39. Springer Berlin Heidelberg. URL https://doi.org/10.1007/978-3-540-85101-1_3.
  • Buium [1995] A. Buium. Geometry of differential polynomial functions, III: Moduli spaces. American Journal of Mathematics, 117(1):1, 1995. URL https://doi.org/10.2307/2375035.
  • Fischer and Lieb [2012] W. Fischer and I. Lieb. A Course in Complex Analysis. Vieweg+Teubner Verlag, 2012. URL https://doi.org/10.1007/978-3-8348-8661-3.
  • Fliess [1987] M. Fliess. Nonlinear control theory and differential algebra: Some illustrative examples. IFAC Proceedings Volumes, 20(5):103–107, 1987. URL https://doi.org/10.1016/s1474-6670(17)55072-7.
  • Freitag and Scanlon [2017] J. Freitag and T. Scanlon. Strong minimality and the jj-function. Journal of the European Mathematical Society, 23(1):119–136, 2017. URL https://doi.org/10.4171/jems/761.
  • Gauchman and Rubel [1989] H. Gauchman and L. A. Rubel. Sums of products of functions of x times functions of y. Linear Algebra and its Applications, 125:19–63, 1989. URL https://doi.org/10.1016/0024-3795(89)90031-1.
  • Gerasimova et al. [2017] O. V. Gerasimova, Y. P. Razmyslov, and G. A. Pogudin. Rolling simplexes and their commensurability. III (Capelli identities and their application to differential algebras). Journal of Mathematical Sciences, 221:315–325, 2017. URL https://doi.org/10.1007/s10958-017-3229-3.
  • Goursat [1945] E. Goursat. A Course in Mathematical Analysis, volume II, part two. Dover Publications, Inc., 1945.
  • Hardouin and Singer [2008] C. Hardouin and M. F. Singer. Differential Galois theory of linear difference equations. Mathematische Annalen, 342(2):333–377, 2008. URL https://doi.org/10.1007/s00208-008-0238-z.
  • Kolchin [1973] E. Kolchin. Differential Algebra and Algebraic Groups. Academic Press, New York, 1973.
  • Lang [2002] S. Lang. Algebra. Springer-Verlag New York, Inc., 2002.
  • Lemaire [2003] F. Lemaire. An orderly linear PDE system with analytic initial conditions with a non-analytic solution. Journal of Symbolic Computation, 35(5):487–498, 2003. URL https://doi.org/10.1016/s0747-7171(03)00017-8.
  • [14] D. Marker, D. Marker, M. Messmer, and A. Pillay. Model theory of differential fields. In Model Theory of Fields, pages 38–113. Cambridge University Press. URL https://doi.org/10.1017/9781316716991.003.
  • Nagloo [2021] J. Nagloo. Model theory and differential equations. Notices of the American Mathematical Society, 68(02):1, 2021. URL https://doi.org/10.1090/noti2221.
  • Pila and Tsimerman [2016] J. Pila and J. Tsimerman. Ax-Schanuel for the j-function. Duke Mathematical Journal, 165(13), 2016. URL https://doi.org/10.1215/00127094-3620005.
  • Razmyslov [1994] Y. P. Razmyslov. Identities of Algebras and their Representations. Translations of Mathematical Monographs. American Mathematical Society, 1994.
  • Riquier [1910] C. Riquier. Les systèmes d’équations aux dérivées partielles. Gauthier-Villars, 1910.
  • Ritt [1950] J. F. Ritt. Differential Algerba. 1950.
  • Seidenberg [1958] A. Seidenberg. Abstract differential algebra and the analytic case. Proceedings of the American Mathematical Society, 9(1):159–164, 1958. URL https://doi.org/10.2307/2033416.
  • van der Put and Singer [2003] M. van der Put and M. F. Singer. Galois Theory of Linear Differential Equations. Springer Berlin Heidelberg, 2003. URL https://doi.org/10.1007/978-3-642-55750-7.