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

Finite Splittings of Differential Matrix Algebras

Amit Kulshrestha [email protected]  and  Kanika Singla [email protected] IISER Mohali, Knowledge City, Sector 81, Mohali 140 306 INDIA
Abstract.

It is well known that central simple algebras are split by suitable finite Galois extensions of their centers. In [JM08] a counterpart of this result was studied in the set up of differential matrix algebras, wherein Picard-Vessiot extensions that split matrix differential algebras were constructed. In this article, we exhibit instances of differential matrix algebras which are split by finite extensions. In some cases, we relate the existence of finite splitting extensions of a differential matrix algebra to the triviality of its tensor powers, and show in these cases, that orders of differential matrix algebras divide their degrees.

Key words and phrases:
differential algebras, splitting of central simple algebras
2010 Mathematics Subject Classification:
12H05, 16H05
This work is supported by DST, India through the research grant EMR/2016/001516.

1. Introduction

Let (K,δ)(K,\delta) be a differential field of characteristic zero. Let AA be a central simple algebra over KK and 𝒟:AA\mathcal{D}:A\to A be a derivation such that 𝒟|K=δ\mathcal{D}|_{K}=\delta. Such a pair (A,𝒟)(A,\mathcal{D}) is called a differential central simple algebra over (K,δ)(K,\delta). If AA is a matrix algebra over KK and 𝒟\mathcal{D} is the derivation δc\delta^{c} obtained by applying δ\delta to each matrix entry, then (A,𝒟)(A,\mathcal{D}) is called a split differential central simple algebra. A differential extension (E,δE)(E,\delta_{E}) of (K,δ)(K,\delta) is said to be a splitting field of (A,𝒟)(A,\mathcal{D}) if the differential central simple algebra (A,𝒟)(E,δE)(A,\mathcal{D})\otimes(E,\delta_{E}) obtained after a base change to (E,δE)(E,\delta_{E}) is split. In [JM08], Juan and Magid prove that if the field of constants of δ\delta is algebraically closed, then every differential central simple algebra (A,𝒟)(A,\mathcal{D}) admits a Picard-Vessiot extension (E,δE)(E,\delta_{E}) that is also a splitting field of (A,𝒟)(A,\mathcal{D}). Picard-Vessiot extensions have a large transcendence degree. However, in this article, we are interested in finite splitting fields of differential matrix algebras. A splitting extension (E,δE)(E,\delta_{E}) of (K,δ)(K,\delta) is called finite if the degree of EE over KK is finite.

If 𝒟\mathcal{D} is a derivation on a matrix algebra Mn(K)M_{n}(K) such that 𝒟|K=δ\mathcal{D}|_{K}=\delta, then by an application of Skolem-Noether theorem, 𝒟(X)=δc(X)+PXXP\mathcal{D}(X)=\delta^{c}(X)+PX-XP, where PP is a traceless matrix. We prove three types of results in this article.

  1. (1)

    Conditions on PP that ascertain the existence of finite splitting extensions. (Theorem 4.2)

  2. (2)

    Calculation of the minimal transcendence degree of splitting extensions, if finite splitting extensions do not exist. (see §4.2)

  3. (3)

    Relating the existence of finite splitting extensions to the existence of an mm\in\mathbb{N} such that the mm-fold tensor power (Mn(K),𝒟)m(M_{n}(K),\mathcal{D})^{\otimes m} is split. (Theorem 5.1, Corollary 5.3)

Since a central simple algebra is split by a finite extension, the case of differential matrix algebras is a significant step while dealing with splitting fields of arbitrary differential central simple algebras.

The organization of the article is as follows. In §2, we set up notations, briefly recall the results of [JM08] and formulate lemmas on which our proofs are based. We develop tools for the proofs of the main results of the article in §3. Given a finite subset SK{0}S\subseteq K\setminus\{0\}, we find differential extensions, called splitting fields of SS, that contain solutions of δ(y)=ay\delta(y)=ay for every aSa\in S. The crucial part of this section is the decomposition 𝒜S+𝒜S=S\mathcal{A}_{S}+\mathcal{A}_{S}^{\prime}=\langle S\rangle, where 𝒜\mathcal{A} is the \mathbb{Q}-span of the set of logarithmic derivatives of δ\delta, S\langle S\rangle is the additive subgroup of KK generated by SS, 𝒜S=𝒜S\mathcal{A}_{S}=\mathcal{A}\cap\langle S\rangle and 𝒜S\mathcal{A}_{S}^{\prime} is a \mathbb{Z}-complement of 𝒜S\mathcal{A}_{S} in S\langle S\rangle. The group 𝒜S\mathcal{A}_{S} governs the algebraic part, and the group 𝒜S\mathcal{A}_{S}^{\prime} governs the transcendental part of a splitting field of SS. The following theorem makes it more precise.

Theorem A.

Let SK\{0}S\subseteq K\backslash\{0\} be a finite subset. If 𝒜S=0\mathcal{A}_{S}^{\prime}=0, then there exists a splitting field (E,δE)(E,\delta_{E}) of SS such that the field extension E/KE/K is finite. If 𝒜S0\mathcal{A}_{S}^{\prime}\neq 0 and (E,δE)(E,\delta_{E}) is a splitting field of SS, then EE is transcendental over KK and trdegK(E)rank(𝒜S){\rm trdeg}_{K}(E)\geq{\rm rank}(\mathcal{A}_{S}^{\prime}). Further, this lower bound on trdegK(E){\rm trdeg}_{K}(E) is sharp. (Theorem 3.3, Corollary 3.5, Proposition 3.6)

In the last two sections of this article, we fix the differential field to be (K,δ)(K,\delta), where K=F(t)K=F(t) and δ\delta is the derivation on KK determined by δ(t)=t\delta(t)=t and δ(a)=0\delta(a)=0 for aFa\in F. Here, FF is a field of characteristic 0. The main section of the paper is §4, where we prove the following theorem concerning the existence of a finite splitting field of matrix differential algebras over (K,δ)(K,\delta).

Theorem B.

Let PMn(K)P\in M_{n}(K) be a traceless matrix and 𝒟P(X)=δc(X)+PXXP\mathcal{D}_{P}(X)=\delta^{c}(X)+PX-XP.

  1. (i).

    If PMn(F)P\in M_{n}(F) then (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) admits a finite splitting field if and only if PP is diagonalizable and every eigenvalue of PP is rational. (Theorem 4.2)

  2. (ii).

    If PMn(F)P\notin M_{n}(F) then (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) admits a finite splitting field if PP has one of the following forms. (Theorem 4.5)

    1. (a)

      P=diag(d1,d2,,dn)P={\rm diag}(d_{1},d_{2},\dots,d_{n}) and each di𝒜d_{i}\in\mathcal{A}.

    2. (b)

      P=(pij)P=(p_{ij}), pijF[t]p_{ij}\in F[t] is an upper triangular matrix satisfying piip_{ii}\in\mathbb{Q} and piipjj/p_{ii}\mathbb{Z}\neq p_{jj}\mathbb{Z}\in\mathbb{Q}/\mathbb{Z}, unless i=ji=j.

Subsequently, in §4.2, over a more general derivation on K=F(t)K=F(t) defined by δc,m(a)=0\delta_{c,m}(a)=0 for aFa\in F and δc,m(t)=ctm\delta_{c,m}(t)=ct^{m}, cF{0}c\in F\setminus\{0\}, mm\in\mathbb{Z}, we compute the minimum transcendence degree of a splitting field of (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}), where PMn(F)P\in M_{n}(F). We notice that the only case when finite splitting extensions exist for such derivations is m=1m=1.

Finally, in §5, we relate the existence of finite splitting fields to the splitting of a tensor power of differential matrix algebras over the differential field (K,δ)(K,\delta). Smallest integer rr such that (Mn(K),𝒟P)r(M_{n}(K),\mathcal{D}_{P})^{\otimes r} is split is called the order of (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}).

Theorem C.

Let PMn(F)P\in M_{n}(F) be a traceless matrix. If the order of (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is finite, then (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) admits a finite splitting field over (K,δ)(K,\delta). Further, the order is a divisor of nn. (Theorem 5.1, Corollary 5.3)

Acknowledgement.  We thank Varadharaj R. Srinivasan for helpful discussions and his interest in this work. We also thank B. Sury for valuable comments on a draft of this paper.

2. Preliminaries

In this section, we recall relevant definitions and record some lemmas to be used subsequently in the paper.

2.1. Basic definitions

Throughout the paper, KK will denote a field of characteristic zero. A specific instance of KK will be explicitly mentioned, wherever needed. Let AA be an associative algebra over KK. An additive map 𝒟:AA\mathcal{D}:A\to A is called a derivation if 𝒟(ab)=𝒟(a)b+a𝒟(b)\mathcal{D}(ab)=\mathcal{D}(a)b+a\mathcal{D}(b). The field KK together with a derivation δ:KK\delta:K\to K is called a differential field. For a differential field (K,δ)(K,\delta), the collection CK={aK:δ(a)=0}C_{K}=\{a\in K:\delta(a)=0\} forms a subfield of KK and is called the field of constants of (K,δ)(K,\delta). A derivation of the form u\partial_{u} for some uAu\in A, where u(x)=uxxu\partial_{u}(x)=ux-xu, for all xAx\in A, is called an inner derivation of AA. If (A,𝒟)(A,\mathcal{D}) and (K,δ)(K,\delta) are such that δ=𝒟|K\delta=\mathcal{D}|_{K} then (A,𝒟)(A,\mathcal{D}) is called a differential algebra over (K,δ)(K,\delta) and 𝒟\mathcal{D} is called an extension of δ\delta.

If EKE\supseteq K is a field extension with a derivation δE\delta_{E} such that δE|K=δ\delta_{E}\big{|}_{K}=\delta then (E,δE)(E,\delta_{E}) is called an extension of the differential field (K,δ)(K,\delta) or a differential extension of (K,δ)(K,\delta). A differential extension (E,δE)(E,\delta_{E}) of (K,δ)(K,\delta) is called algebraic (respectively, transcendental or finite) if the field extension E/KE/K is so. If (K,δ)(K,\delta) is a differential field and E/KE/K is an algebraic extension of fields, then a derivation δ\delta of KK extends uniquely to a derivation of EE. A differential extension (E,δE)(E,\delta_{E}) over (K,δ)(K,\delta) is called a Picard-Vessiot extension if it does not admit new constants, i.e., CE=CKC_{E}=C_{K}, and EE can be differentially generated by adjoining the solutions of a homogeneous linear ordinary differential equation to KK.

If AA is a central simple KK-algebra, then every derivation δ\delta of KK extends to a derivation 𝒟\mathcal{D} of AA. (see [Ami82, Theorem A]). The pair (A,𝒟)(A,\mathcal{D}), where AA is a central simple algebra over KK and 𝒟\mathcal{D} is a derivation on AA that extends a derivation δ\delta of KK, is called a differential central simple algebra over (K,δ)(K,\delta). If (A1,𝒟1)(A_{1},\mathcal{D}_{1}) and (A2,𝒟2)(A_{2},\mathcal{D}_{2}) are two differential central simple algebras over a differential field (K,δ)(K,\delta) then we define their tensor product by (A1,𝒟1)(A2,𝒟2)=(A1KA2,𝒟1𝒟2)(A_{1},\mathcal{D}_{1})\otimes(A_{2},\mathcal{D}_{2})=(A_{1}\otimes_{K}A_{2},\mathcal{D}_{1}\otimes\mathcal{D}_{2}), where 𝒟1𝒟2(a1a2)=𝒟1(a1)a2+a1𝒟2(a2).\mathcal{D}_{1}\otimes\mathcal{D}_{2}(a_{1}\otimes a_{2})=\mathcal{D}_{1}(a_{1})\otimes a_{2}+a_{1}\otimes\mathcal{D}_{2}(a_{2}).

If 𝒟1\mathcal{D}_{1} and 𝒟2\mathcal{D}_{2} are two derivations on a central simple algebra AA which extend δ\delta, then by an application of Skolem-Noether Theorem 𝒟1𝒟2\mathcal{D}_{1}-\mathcal{D}_{2} is inner (see [Jac89, Theorem 4.9]), and hence there exists uAu\in A such that 𝒟1𝒟2=u\mathcal{D}_{1}-\mathcal{D}_{2}=\partial_{u}. Consequently, for a matrix algebra Mn(K)M_{n}(K), every derivation 𝒟\mathcal{D} that extends δ\delta is of the form 𝒟P:=δc+P\mathcal{D}_{P}:=\delta^{c}+\partial_{P}, where PP is a traceless matrix.

2.2. Splitting of differential central simple algebras

Let (A,𝒟)(A,\mathcal{D}) be a differential central simple algebra of degree nn over (K,δ)(K,\delta). If there exists a differential extension (E,δE)(E,\delta_{E}) over (K,δ)(K,\delta) such that (A,𝒟)(E,δE)(A,\mathcal{D})\otimes(E,\delta_{E}) is differentially isomorphic to (Mn(E),δEc)(M_{n}(E),\delta_{E}^{c}), where δEc(B)=(δE(bij))\delta_{E}^{c}(B)=(\delta_{E}(b_{ij})) for every B=(bij)Mn(E)B=(b_{ij})\in M_{n}(E), then (A,𝒟)(A,\mathcal{D}) is said to be split by (E,δE)(E,\delta_{E}) or trivialized by (E,δE)(E,\delta_{E}). If (A,𝒟)(A,\mathcal{D}) is isomorphic to (Mn(K),δc)(M_{n}(K),\delta^{c}), then (A,𝒟)(A,\mathcal{D}) is called a split differential central simple algebra over (K,δ)(K,\delta).

We now briefly recall the results of Juan and Magid [JM08].

Proposition 2.1.

[JM08, Proposition 2] Let (K,δ)(K,\delta) be a differential field and P,QMn(K)P,Q\in M_{n}(K) be traceless matrices. Differential matrix algebras (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) and (Mn(K),𝒟Q)(M_{n}(K),\mathcal{D}_{Q}) are isomorphic if and only if there exists HGLn(K)H\in{\rm GL}_{n}(K) such that H1δc(H)+H1QH=PH^{-1}\delta^{c}(H)+H^{-1}QH=P. In particular, a differential matrix algebra (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is split if and only if P=H1δc(H)P=H^{-1}\delta^{c}(H) for some HGLn(K)H\in{\rm GL}_{n}(K).

Corollary 2.2.

If δ\delta is the zero derivation on KK, then (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is isomorphic to (Mn(K),𝒟Q)(M_{n}(K),\mathcal{D}_{Q}) if and only if PP and QQ are similar matrices.

Corollary 2.3.

Let (E,δE)(E,\delta_{E}) be a differential extension of (K,δ)(K,\delta). Then (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is split by (E,δE)(E,\delta_{E}) if and only if GLn(E){\rm GL}_{n}(E) contains a solution of the equation δc(Y)=PtY\delta^{c}(Y)=P^{t}Y.

Later in this paper, we will frequently use the above corollary in conjunction with the following lemma.

Lemma 2.4.

Let (K,δ)(K,\delta) be a differential field, and CKC_{K} be its field of constants. Let A,BMn(K)A,B\in M_{n}(K), and QMn(CK)Q\in M_{n}(C_{K}) be such that A=Q1BQA=Q^{-1}BQ. Let (E,δE)(K,δ)(E,\delta_{E})\supseteq(K,\delta) be an extension of differential fields. Then, δc(Y)=AtY\delta^{c}(Y)=A^{t}Y has a solution for YY in GLn(E){\rm GL}_{n}(E) if and only if δc(Y)=BtY\delta^{c}(Y)=B^{t}Y has a solution for YY in GLn(E){\rm GL}_{n}(E).

Proof.

Suppose that δc(Y)=AtY\delta^{c}(Y)=A^{t}Y has a solution ZGLn(E)Z\in{\rm GL}_{n}(E). Then

δEc(Z)=AtZ=QtBt(Q1)tZ=QtBt(Q1)tZQt(Q1)t.\delta_{E}^{c}(Z)=A^{t}Z=Q^{t}B^{t}(Q^{-1})^{t}Z=Q^{t}B^{t}(Q^{-1})^{t}ZQ^{t}(Q^{-1})^{t}.

Let Z1=(Q1)tZQtZ_{1}=(Q^{-1})^{t}ZQ^{t}. Since δc(Q)=0\delta^{c}(Q)=0, we have

δEc(Z1)=δEc((Q1)tZQt)=(Q1)tδEc(Z)Qt=Bt(Q1)tZQt=BtZ1.\delta_{E}^{c}(Z_{1})=\delta_{E}^{c}((Q^{-1})^{t}ZQ^{t})=(Q^{-1})^{t}\delta_{E}^{c}(Z)Q^{t}=B^{t}(Q^{-1})^{t}ZQ^{t}=B^{t}Z_{1}.

Thus a solution of δc(Y)=BtY\delta^{c}(Y)=B^{t}Y is Z1=(Q1)tZQtGLn(E)Z_{1}=(Q^{-1})^{t}ZQ^{t}\in{\rm GL}_{n}(E). Converse holds because the hypothesis of the lemma is symmetric in AA and BB. ∎

2.3. Order of (A,𝒟)(A,\mathcal{D})

A differential central simple algebra (A,𝒟)(A,\mathcal{D}) over (K,δ)(K,\delta) is said to be of finite order if there is an integer m1m\geq 1 such that (A,𝒟)m:=(A,𝒟)(A,𝒟)(A,𝒟)mtimes(A,\mathcal{D})^{\otimes m}:=\underbrace{(A,\mathcal{D})\otimes(A,\mathcal{D})\otimes\dots\otimes(A,\mathcal{D})}_{m\,\,times} is split over (K,δ)(K,\delta) itself. If (A,𝒟)(A,\mathcal{D}) is of finite order, then the smallest mm for which (A,𝒟)m(A,\mathcal{D})^{\otimes m} is split is called the order of (A,𝒟)(A,\mathcal{D}).

The lemmas recorded in this subsection would be referred to while proving results concerning orders of differential matrix algebras.

Lemma 2.5.

Let (A1,𝒟P1)(A_{1},\mathcal{D}_{P_{1}}) and (A2,𝒟P2)(A_{2},\mathcal{D}_{P_{2}}) be two differential matrix algebras over (K,δ)(K,\delta). Then 𝒟P1𝒟P2=𝒟P11+1P2\mathcal{D}_{P_{1}}\otimes\mathcal{D}_{P_{2}}=\mathcal{D}_{P_{1}\otimes 1+1\otimes P_{2}}, where P1A1,P2A2P_{1}\in A_{1},P_{2}\in A_{2} respectively.

Proof.

For X1A1X_{1}\in A_{1} and X2A2X_{2}\in A_{2},

(𝒟P1𝒟P2)(X1X2)\displaystyle(\mathcal{D}_{P_{1}}\otimes\mathcal{D}_{P_{2}})(X_{1}\otimes X_{2}) =(𝒟P11+1𝒟P2)(X1X2)\displaystyle=(\mathcal{D}_{P_{1}}\otimes 1+1\otimes\mathcal{D}_{P_{2}})(X_{1}\otimes X_{2})
=(𝒟P11)(X1X2)+(1𝒟P2)(X1X2)\displaystyle=(\mathcal{D}_{P_{1}}\otimes 1)(X_{1}\otimes X_{2})+(1\otimes\mathcal{D}_{P_{2}})(X_{1}\otimes X_{2})
=(δc(X1)+P1X1X1P1)X2+X1(δc(X2)+P2X2X2P2)\displaystyle=(\delta^{c}(X_{1})+P_{1}X_{1}-X_{1}P_{1})\otimes X_{2}+X_{1}\otimes(\delta^{c}(X_{2})+P_{2}X_{2}-X_{2}P_{2})
=δc(X1)X2+X1δc(X2)+P1X1X2+X1P2X2\displaystyle=\delta^{c}(X_{1})\otimes X_{2}+X_{1}\otimes\delta^{c}(X_{2})+P_{1}X_{1}\otimes X_{2}+X_{1}\otimes P_{2}X_{2}
X1P1X2X1X2P2\displaystyle\quad-X_{1}P_{1}\otimes X_{2}-X_{1}\otimes X_{2}P_{2}
=δc(X1X2)+(P11+1P2)(X1X2)\displaystyle=\delta^{c}(X_{1}\otimes X_{2})+(P_{1}\otimes 1+1\otimes P_{2})(X_{1}\otimes X_{2})
(X1X2)(P11+1P2)\displaystyle\quad-(X_{1}\otimes X_{2})(P_{1}\otimes 1+1\otimes P_{2})
=𝒟P11+1P2(X1X2).\displaystyle=\mathcal{D}_{P_{1}\otimes 1+1\otimes P_{2}}(X_{1}\otimes X_{2}).

If 𝒟Q=𝒟P1𝒟P2\mathcal{D}_{Q}=\mathcal{D}_{P_{1}}\otimes\mathcal{D}_{P_{2}} then Q=P11+1P2Q=P_{1}\otimes 1+1\otimes P_{2}, where \otimes is the Kronecker product of matrices. By [HJ94, Theorem 4.4.5] eigenvalues of QQ are the sums of the eigenvalues of P1P_{1} and P2P_{2}, i.e., if Λ1\Lambda_{1} and Λ2\Lambda_{2} denote the sets of eigenvalues of P1P_{1} and P2P_{2}, respectively, then the set of eigenvalues of QQ is Λ={λ1+λ2:λ1Λ1,λ2Λ2}.\Lambda=\{\lambda_{1}+\lambda_{2}:\lambda_{1}\in\Lambda_{1},\lambda_{2}\in\Lambda_{2}\}.

Lemma 2.6.

Let (Ai,𝒟Pi)(A_{i},\mathcal{D}_{P_{i}}), i{1,2,,n}i\in\{1,2,\dots,n\} be differential matrix algebras over (K,δ)(K,\delta). Let

(A,𝒟Q)=(A1,𝒟P1)(A2,𝒟P2)(An,𝒟Pn).(A,\mathcal{D}_{Q})=(A_{1},\mathcal{D}_{P_{1}})\otimes(A_{2},\mathcal{D}_{P_{2}})\otimes\dots\otimes(A_{n},\mathcal{D}_{P_{n}}).

For each i{1,2,,n}i\in\{1,2,\dots,n\}, let Λi\Lambda_{i} be the set of eigenvalues of PiP_{i} and Λ\Lambda be the set of eigenvalues of QQ. Then

  1. (i)

    Λ={λ1+λ2++λn:λiΛi}\Lambda=\{\lambda_{1}+\lambda_{2}+\dots+\lambda_{n}:\lambda_{i}\in\Lambda_{i}\}. In particular, the matrix QQ has all its eigenvalues in \mathbb{{Q}} if and only if each PiP_{i} has all its eigenvalues in \mathbb{{Q}}.

  2. (ii)

    QQ is diagonalizable if and only if each PiP_{i} is diagonalizable.

Proof.

(i)(i).  Since the Kronecker product of matrices is associative (see [HJ94, Property 4.2.6]), Λ={λ1+λ2++λn:λiΛi}\Lambda=\{\lambda_{1}+\lambda_{2}+\dots+\lambda_{n}:\lambda_{i}\in\Lambda_{i}\} holds inductively, with n=2n=2 as the base case that is discussed immediately before this lemma. Therefore, if each Λi\Lambda_{i}\subset\mathbb{Q} then Λ\Lambda\subset\mathbb{Q}. Conversely, suppose that Λ\Lambda\subset\mathbb{Q}. Then for a given λΛi\lambda\in\Lambda_{i} and arbitrary λjΛj\lambda_{j}\in\Lambda_{j} we have

(2.1) λ+jiλj.\displaystyle\lambda+\sum_{j\neq i}\lambda_{j}\in\mathbb{Q}.

Since each PjP_{j} is a traceless matrix, λjΛjλj=0\sum_{\lambda_{j}\in\Lambda_{j}}{\lambda_{j}}=0. Using this after adding up all instances of (2.1) as λj\lambda_{j} varies over Λj\Lambda_{j}, we have (ji|Λj|)λ\left(\prod_{j\neq i}|\Lambda_{j}|\right)\lambda\in\mathbb{Q}, where |Λj||\Lambda_{j}| is the cardinality of the set Λj\Lambda_{j}. Clearly, λ\lambda\in\mathbb{Q}.

(ii)(ii).  As the tensor product (A1,𝒟P1)(A2,𝒟P2)(An,𝒟Pn)(A_{1},\mathcal{D}_{P_{1}})\otimes(A_{2},\mathcal{D}_{P_{2}})\otimes\dots\otimes(A_{n},\mathcal{D}_{P_{n}}) can be constructed repeatedly by taking one tensor factor each time, it suffices to demonstrate the proof for n=2n=2 case.

Suppose n=2n=2. By Lemma 2.5, Q=P11+1P2Q=P_{1}\otimes 1+1\otimes P_{2}. Let JP1J_{P_{1}} and JP2J_{P_{2}} be Jordan forms of P1P_{1} and P2P_{2}, respectively. Let R1R_{1} and R2R_{2} be such that R11P1R1=JP1R_{1}^{-1}P_{1}R_{1}=J_{P_{1}} and R21P2R2=JP2R_{2}^{-1}P_{2}R_{2}=J_{P_{2}}. Note that R1R2R_{1}\otimes R_{2} is invertible and Q=(R1R2)(JP11+1JP2)(R1R2)1Q=(R_{1}\otimes R_{2})(J_{P_{1}}\otimes 1+1\otimes J_{P_{2}})(R_{1}\otimes R_{2})^{-1}. Thus if P1P_{1} and P2P_{2} are diagonalizable, the Jordan forms JP1J_{P_{1}} and JP2J_{P_{2}} are diagonal matrices, and so is JP11+1JP2J_{P_{1}}\otimes 1+1\otimes J_{P_{2}}. In other words, QQ is diagonalizable.

Conversely, if at least one of the matrices, say P2P_{2}, is not diagonalizable. Then there exists an eigenvalue β\beta of P2P_{2} whose algebraic multiplicity is strictly greater than its geometric multiplicity. We may assume, by permuting Jordan blocks of JP2J_{P_{2}} if needed, that the first Jordan block of JP2J_{P_{2}} corresponds to β\beta. Let α\alpha be the eigenvalue corresponding to the first Jordan block of JP1J_{P_{1}}. We claim that the algebraic multiplicity of the eigenvalue α+β\alpha+\beta of Q:=P11+1P2Q:=P_{1}\otimes 1+1\otimes P_{2} is strictly greater than its geometric multiplicity. To see this, we look at the conjugate Q:=JP11+1JP2=(aij)r×rQ^{\prime}:=J_{P_{1}}\otimes 1+1\otimes J_{P_{2}}=(a_{ij})_{r\times r} of QQ and observe that QQ^{\prime} is an upper triangular matrix with a11=a22=α+βa_{11}=a_{22}=\alpha+\beta, a12=1a_{12}=1. Let ss denote the algebraic multiplicity of α+β\alpha+\beta. To estimate its geometric multiplicity, we find a lower bound on the rank of Q(α+β)IrQ^{\prime}-(\alpha+\beta)I_{r}, where IrI_{r} is the r×rr\times r identity matrix. In Q(α+β)IrQ^{\prime}-(\alpha+\beta)I_{r} the first row and rsr-s rows corresponding to eigenvalues different from α+β\alpha+\beta constitute a linearly independent set of size rs+1r-s+1. Thus, rank(Q(α+β)Ir)rs+1{\rm rank}(Q^{\prime}-(\alpha+\beta)I_{r})\geq r-s+1. By rank-nullity theorem, the geometric multiplicity of α+β\alpha+\beta is at most r(rs+1)=s1<sr-(r-s+1)=s-1<s. Thus QQ is not diagonalizable. ∎

As a consequence, we derive the following lemma that will be used later while dealing with tensor powers of differential matrix algebras.

Lemma 2.7.

Let (K,δ)(K,\delta) be a differential field and PMn(K)P\in M_{n}(K) be a traceless matrix. Let QQ be a traceless matrix such that (Mnm(K),𝒟Q)=(Mn(K),𝒟P)m(M_{n^{m}}(K),\mathcal{D}_{Q})=(M_{n}(K),\mathcal{D}_{P})^{\otimes m}. Then the following hold for QQ.

  1. (i)

    QQ is diagonalizable if and only if PP is diagonalizable.

  2. (ii)

    QQ has all its eigenvalues in \mathbb{{Q}} if and only if PP has all its eigenvalues in \mathbb{{Q}}.

3. Splitting fields of SK\{0}S\subseteq K\backslash\{0\}

Let (K,δ)(K,\delta) be a differential field. Consider the set

𝒜={aK:δ(y)=nay has a nonzero solution in K for some n}.\mathcal{A}=\{a\in K:\delta(y)=nay\text{ has a nonzero solution in }K\text{ for some }n\in\mathbb{N}\}.

Then 𝒜\mathcal{A} is an additive subgroup of KK that is closed under division by natural numbers. Let a𝒜a\in\mathcal{A}. Let m,nm,n\in\mathbb{N} and u,vKu,v\in K be such that δ(u)=mau\delta(u)=mau and δ(v)=nav\delta(v)=nav. Then δ(uv1)=(mn)auv1\delta(uv^{-1})=(m-n)auv^{-1}. Thus, for each a𝒜a\in\mathcal{A}, there exists a unique positive integer nan_{a} with the following property: δ(y)=nay\delta(y)=nay has a solution in K\{0}K\backslash\{0\} if and only if nn is an integral multiple of nan_{a}.

An element aKa\in K is called a logarithmic derivative if a=δ(b)b1a=\delta(b)b^{-1} for some bK\{0}b\in K\backslash\{0\}. It is evident that if aa is a logarithmic derivative, then a𝒜a\in\mathcal{A}. We denote the set of logarithmic derivatives by δ(K)\mathcal{L}_{\delta}(K). The following lemma provides a description of 𝒜\mathcal{A} in terms of δ(K)\mathcal{L}_{\delta}(K).

Lemma 3.1.

The subgroup 𝒜\mathcal{A} of KK is equal to the \mathbb{Q}-span of δ(K)\mathcal{L}_{\delta}(K).

Proof.

Since δ(K)\mathcal{L}_{\delta}(K) is a subset of 𝒜\mathcal{A} and 𝒜\mathcal{A} is an additive subgroup of KK which is closed under division by natural numbers, span(δ(K))𝒜{\rm span}_{\mathbb{Q}}(\mathcal{L}_{\delta}(K))\subseteq\mathcal{A}.

Conversely, let a𝒜a\in\mathcal{A}. Then there exists vaK{0}v_{a}\in K\setminus\{0\} and nan_{a}\in\mathbb{N} such that δ(va)=naava\delta(v_{a})=n_{a}av_{a}. Hence a=(1/na)δ(va)va1span(δ(K))a=(1/n_{a})\delta(v_{a})v_{a}^{-1}\in{\rm span}_{\mathbb{Q}}(\mathcal{L}_{\delta}(K)). ∎

Let SK\{0}S\subseteq K\backslash\{0\}. We are interested in extensions (E,δE)(E,\delta_{E}) of (K,δ)(K,\delta) which contain a solution of δ(y)=ay\delta(y)=ay for every aSa\in S. Such extensions are to be called splitting fields of SS. Let a,bSa,b\in S and (E,δE)(E,\delta_{E}) be an extension of (K,δ)(K,\delta) such that δE(ua)=aua\delta_{E}(u_{a})=au_{a} and δE(ub)=bub\delta_{E}(u_{b})=bu_{b} for some ua,ubEu_{a},u_{b}\in E. Then uaub±1Eu_{a}u_{b}^{\pm 1}\in E is a solution of δ(y)=(a±b)y\delta(y)=(a\pm b)y. From this, we derive the following lemma. In the lemma, and thereafter, S\langle S\rangle denotes the additive subgroup of KK generated by SS.

Lemma 3.2.

Let (E,δE)(E,\delta_{E}) be a differential extension of (K,δ)(K,\delta). Then (E,δE)(E,\delta_{E}) is a splitting field of SS if and only if it is a splitting field of S\langle S\rangle.

Let SK\{0}S\subseteq K\backslash\{0\} be a finite subset. Since char(K)=0{\rm char}(K)=0, the group S\langle S\rangle and its subgroups are finitely generated free abelian groups. Let us denote 𝒜S\mathcal{A}\cap\langle S\rangle by 𝒜S\mathcal{A}_{S}. Let 𝒜S\mathcal{A}_{S}^{\prime} denote a \mathbb{Z}-complement of 𝒜S\mathcal{A}_{S} in S\langle S\rangle so that the sum 𝒜S+𝒜S=S\mathcal{A}_{S}+\mathcal{A}_{S}^{\prime}=\langle S\rangle is direct.

Theorem 3.3.

Let SK\{0}S\subseteq K\backslash\{0\} be a finite subset.

  1. (i).

    If 𝒜S=0\mathcal{A}_{S}^{\prime}=0, then there exists a splitting field (E,δE)(E,\delta_{E}) of SS such that the field extension E/KE/K is finite.

  2. (ii).

    If 𝒜S0\mathcal{A}_{S}^{\prime}\neq 0 and (E,δE)(E,\delta_{E}) is splitting field of SS, then EE is transcendental over KK.

Proof.

(i).(i). Since 𝒜S=0\mathcal{A}_{S}^{\prime}=0, S𝒜S\subseteq\mathcal{A}. Let aS𝒜a\in S\subseteq\mathcal{A}. By definition of 𝒜\mathcal{A}, there exist vaK{0}v_{a}\in K\setminus\{0\} and nn\in\mathbb{N} such that δ(va)=nava\delta(v_{a})=nav_{a}. Let EaE_{a} be a field extension of KK that contains a root of the polynomial xnvaK[x]x^{n}-v_{a}\in K[x]. Then a solution of xnva=0x^{n}-v_{a}=0 in EaE_{a} is also a solution of δ(y)=ay\delta(y)=ay in EaE_{a}. Varying aa over SS and repeating this procedure, one may construct a finite extension EE of KK as required.

(ii).(ii). Assume that 𝒜S0\mathcal{A}_{S}^{\prime}\neq 0. It is enough to exhibit an element aSa\in\langle S\rangle such that no algebraic extension of (K,δ)(K,\delta) contains a solution of δ(y)=ay\delta(y)=ay. We claim that any nonzero element of 𝒜S\mathcal{A}_{S}^{\prime} serves the purpose. Let 0a𝒜S0\neq a\in\mathcal{A}_{S}^{\prime}. Let (E,δE)(E,\delta_{E}) be an extension of (K,δ)(K,\delta) and uaEu_{a}\in E be such that δE(ua)=aua\delta_{E}(u_{a})=au_{a}. If uau_{a} is algebraic over KK then we take the minimal monic polynomial f(x)=xn+an1xn1++a1x+a0f(x)=x^{n}+a_{n-1}x^{n-1}+\dots+a_{1}x+a_{0}; a00a_{0}\neq 0 of uau_{a} over KK. Differentiating the relation

uan+an1uan1++a1ua+a0=0,u_{a}^{n}+a_{n-1}u_{a}^{n-1}+\dots+a_{1}u_{a}+a_{0}=0,

we get

(nuan1+(n1)an1uan2++a0)aua+δ(an1)uan1++δ(a0)=0.(nu_{a}^{n-1}+(n-1)a_{n-1}u_{a}^{n-2}+\dots+a_{0})au_{a}+\delta(a_{n-1})u_{a}^{n-1}+\dots+\delta(a_{0})=0.

Thus the polynomial g(x):=naxn+((n1)an1a+δ(an1))xn1++δ(a0)g(x):=nax^{n}+((n-1)a_{n-1}a+\delta(a_{n-1}))x^{n-1}+\dots+\delta(a_{0}) of degree nn is satisfied by uau_{a}. Therefore, g(x)=naf(x)g(x)=naf(x). Now comparing constant terms of these polynomials, a=(1/n)δ(a0)a01span(δ(K))=𝒜a=(1/n)\delta(a_{0})a_{0}^{-1}\in{\rm span}_{\mathbb{Q}}(\mathcal{L}_{\delta}(K))=\mathcal{A}. Thus a(𝒜S)𝒜S=0a\in(\mathcal{A}\cap\langle S\rangle)\cap\mathcal{A}_{S}^{\prime}=0. This is a contradiction. ∎

Now we shall compute the minimal transcendence degree over KK amongst the splitting fields of SS. If 𝒜S=0\mathcal{A}_{S}^{\prime}=0; the minimal transcendence degree is 0, as is evident from Theorem 3.3(i)(i). Thus we would be interested in 𝒜S0\mathcal{A}_{S}^{\prime}\neq 0 case. For this we introduce some notation.

Let SK\{0}S\subseteq K\backslash\{0\} be a finite subset with 𝒜S0\mathcal{A}_{S}^{\prime}\neq 0 and (E,δE)(E,\delta_{E}) be an extension of (K,δ)(K,\delta) that contains elements saEs_{a}\in E such that δE(sa)=asa\delta_{E}(s_{a})=as_{a} for a𝒜Sa\in\mathcal{A}_{S}^{\prime}. Scaling each saEs_{a}\in E by a suitable element of CEC_{E}, if required, we may assume that sasb=sa+bs_{a}s_{b}=s_{a+b} for each a,b𝒜Sa,b\in\mathcal{A}_{S}^{\prime}. We may also assume s0=1s_{0}=1.

Proposition 3.4.

With the notation as above, we have the following.

  1. (i).

    The subset {sa:a𝒜S}\{s_{a}:a\in\mathcal{A}_{S}^{\prime}\} of EE is linearly independent over KK.

  2. (ii).

    Let a1,a2,,ar𝒜Sa_{1},a_{2},\dots,a_{r}\in\mathcal{A}_{S}^{\prime}. Then {a1,a2,,ar}\{a_{1},a_{2},\dots,a_{r}\} is a \mathbb{Q}-linearly independent subset of KK if and only if sa1,sa2,,sarEs_{a_{1}},s_{a_{2}},\dots,s_{a_{r}}\in E are algebraically independent over KK.

Proof.

(i).(i). Let rr\in\mathbb{N} be the smallest integer such that

(3.1) i=1risai=0\sum_{i=1}^{r}\ell_{i}s_{a_{i}}=0

for 0iK0\neq\ell_{i}\in K. Applying δE\delta_{E} on both sides of the equation 3.1, we get

(3.2) (δ(1)+1a1)sa1+i=2r(δ(i)+iai)sai=0.(\delta(\ell_{1})+\ell_{1}a_{1})s_{a_{1}}+\sum_{i=2}^{r}\left(\delta(\ell_{i})+\ell_{i}a_{i}\right)s_{a_{i}}=0.

Multiplying equation 3.1 by δ(1)+1a1\delta(\ell_{1})+\ell_{1}a_{1} and equation 3.2 by 1\ell_{1}, and subtracting

(3.3) i=2r((δ(1)+1a1)i(δ(i)+iai)1)sai=0.\sum_{i=2}^{r}((\delta(\ell_{1})+\ell_{1}a_{1})\ell_{i}-(\delta(\ell_{i})+\ell_{i}a_{i})\ell_{1})s_{a_{i}}=0.

By minimality of rr, each coefficient of sais_{a_{i}} in equation 3.3 must be 0. Thus

(δ(1)+1a1)i(δ(i)+iai)1=0(\delta(\ell_{1})+\ell_{1}a_{1})\ell_{i}-(\delta(\ell_{i})+\ell_{i}a_{i})\ell_{1}=0

for each i{2,,r}i\in\{2,\dots,r\}. Rearranging these terms, a1ai=δ(i)i1δ(1)11a_{1}-a_{i}=\delta(\ell_{i})\ell_{i}^{-1}-\delta(\ell_{1})\ell_{1}^{-1}. Evidently a1ai𝒜Sa_{1}-a_{i}\in\mathcal{A}_{S}^{\prime} and δ(i)i1𝒜\delta(\ell_{i})\ell_{i}^{-1}\in\mathcal{A}. Thus a1ai𝒜S𝒜={0}a_{1}-a_{i}\in\mathcal{A}_{S}^{\prime}\cap\mathcal{A}=\{0\}, and hence a1=aia_{1}=a_{i} for each i{2,,r}i\in\{2,\dots,r\}. This is a contradiction to the minimality of rr, and we conclude that {sa:a𝒜S}\{s_{a}:a\in\mathcal{A}_{S}^{\prime}\} is a linearly independent set over KK.

(ii).(ii). We first assume that {a1,a2,,ar}\{a_{1},a_{2},\dots,a_{r}\} is a \mathbb{Q}-linearly dependent set. Then there exist nonzero nin_{i}\in\mathbb{Z}, such that iniai=0\sum_{i}n_{i}a_{i}=0. Hence isaini=siniai=s0=1\prod_{i}s_{a_{i}}^{n_{i}}=s_{\sum_{i}n_{i}a_{i}}=s_{0}=1 is an algebraic relation over \mathbb{Q} that is satisfied by sa1,sa2,,sars_{a_{1}},s_{a_{2}},\dots,s_{a_{r}}.

Conversely, assume that {a1,a2,,ar}\{a_{1},a_{2},\dots,a_{r}\} is a \mathbb{Q}-linearly independent set. If possible, let

f(x1,x2,,xr)=(i1,i2,,ir)(i1,i2,,ir)x1i1x2i2xrirK[x1,x2,,xr]f(x_{1},x_{2},\dots,x_{r})=\sum_{(i_{1},i_{2},\dots,i_{r})\in\mathcal{I}}\ell_{(i_{1},i_{2},\dots,i_{r})}x_{1}^{i_{1}}x_{2}^{i_{2}}\dots x_{r}^{i_{r}}\in K[x_{1},x_{2},\dots,x_{r}]

be a polynomial such that f(sa1,sa2,,sar)=0Ef(s_{a_{1}},s_{a_{2}},\dots,s_{a_{r}})=0\in E. Then

(i1,i2,,ir)(i1,i2,,ir)sa1i1sa2i2sarir=0=(i1,i2,,ir)(i1,i2,,ir)si1a1+i2a2++irar=0E\sum_{(i_{1},i_{2},\dots,i_{r})\in\mathcal{I}}\ell_{(i_{1},i_{2},\dots,i_{r})}s_{a_{1}}^{i_{1}}s_{a_{2}}^{i_{2}}\dots s_{a_{r}}^{i_{r}}=0=\sum_{(i_{1},i_{2},\dots,i_{r})\in\mathcal{I}}\ell_{(i_{1},i_{2},\dots,i_{r})}s_{i_{1}a_{1}+i_{2}a_{2}+\dots+i_{r}a_{r}}=0\in E

Since {a1,a2,,ar}\{a_{1},a_{2},\dots,a_{r}\} is a \mathbb{Q}-linearly independent set, i1a1+i2a2++irari_{1}a_{1}+i_{2}a_{2}+\dots+i_{r}a_{r} are all distinct as (i1,i2,,ir)(i_{1},i_{2},\dots,i_{r}) varies over \mathcal{I}. Now from part (i)(i) of this proposition, the set

{si1a1+i2a2++irar:(i1,i2,,ir)}\{s_{i_{1}a_{1}+i_{2}a_{2}+\dots+i_{r}a_{r}}:(i_{1},i_{2},\dots,i_{r})\in\mathcal{I}\}

is linear independent over KK. Thus (i1,i2,,ir)=0\ell_{(i_{1},i_{2},\dots,i_{r})}=0 for each (i1,i2,,ir)(i_{1},i_{2},\dots,i_{r})\in\mathcal{I} and

f(x1,x2,,xr)=0K[x1,x2,,xr].f(x_{1},x_{2},\dots,x_{r})=0\in K[x_{1},x_{2},\dots,x_{r}].

Hence sa1,sa2,,sars_{a_{1}},s_{a_{2}},\dots,s_{a_{r}} are algebraically independent over KK. ∎

As a corollary to Proposition 3.4 we derive the following.

Corollary 3.5.

Let (K,δ)(K,\delta) be a differential field and SK\{0}S\subseteq K\backslash\{0\} be a finite set. Let (E,δE)(E,\delta_{E}) be a splitting field of SS. Then trdegK(E)rank(𝒜S){\rm trdeg}_{K}(E)\geq{\rm rank}(\mathcal{A}_{S}^{\prime}).

Proof.

Since (E,δE)(E,\delta_{E}) is a splitting field of SS, by Lemma 3.2 it is also a splitting field of S\langle S\rangle. In particular, the equation δ(y)=ay\delta(y)=ay has a solution saEs_{a}\in E for each a𝒜Sa\in\mathcal{A}_{S}^{\prime}. Let S\mathcal{B}_{S}^{\prime} be a \mathbb{Z}-basis of 𝒜S\mathcal{A}_{S}^{\prime}. By Proposition 3.4, the elements sas_{a}, where aSa\in\mathcal{B}_{S}^{\prime}, are algebraically independent over KK. Thus trdegK(E)|S|=rank(𝒜S){\rm trdeg}_{K}(E)\geq|\mathcal{B}_{S}^{\prime}|={\rm rank}(\mathcal{A}_{S}^{\prime}). ∎

We assert that there indeed exists a splitting field (E,δE)(E,\delta_{E}) of SS such that trdegK(E)=rank(𝒜S){\rm trdeg}_{K}(E)={\rm rank}(\mathcal{A}_{S}^{\prime}). To establish this assertion, we construct a differential extension (K(tS),δK(tS))(K(t_{S}),\delta_{K(t_{S})}) of (K,δ)(K,\delta) as follows: Fix a \mathbb{Z}-basis S\mathcal{B}_{S}^{\prime} of 𝒜S\mathcal{A}_{S}^{\prime}. Let tS:={ta:aS}t_{S}:=\{t_{a}:a\in\mathcal{B}_{S}^{\prime}\} be a set, in bijection with S\mathcal{B}_{S}^{\prime}, consisting of algebraically independent mutually commuting indeterminates over KK. We denote by K(tS)K(t_{S}) the field generated by KK and tSt_{S}. We extend the derivation δ\delta of KK to K(tS)K(t_{S}) by defining δK(tS)(ta)=ata\delta_{K(t_{S})}(t_{a})=at_{a} for each aSa\in\mathcal{B}_{S}^{\prime}. It is easy to check that

δK(tS)(tamtbn)=(ma+nb)tamtbn,\delta_{K(t_{S})}(t_{a}^{m}t_{b}^{n})=(ma+nb)t_{a}^{m}t_{b}^{n},

for m,nm,n\in\mathbb{Z} and a,bSa,b\in\mathcal{B}_{S}^{\prime}. Therefore (K(tS),δK(tS))(K(t_{S}),\delta_{K(t_{S})}) has a solution of the equation δ(y)=ay\delta(y)=ay for each a𝒜Sa\in\mathcal{A}_{S}^{\prime} and the transcendence degree of K(tS)K(t_{S}) over KK is equal to the rank of 𝒜S\mathcal{A}_{S}^{\prime}.

Proposition 3.6.

Let (K,δ)(K,\delta) be a differential field and SK\{0}S\subseteq K\backslash\{0\}. Then there exists a splitting field (E,δE)(E,\delta_{E}) of SS such that trdegK(E)=rank(𝒜S){\rm trdeg}_{K}(E)={\rm rank}(\mathcal{A}_{S}^{\prime}).

Proof.

We claim that there exists an algebraic extension of the field K(tS)K(t_{S}) containing solutions of δ(y)=ay\delta(y)=ay for all aSa\in S. To see this, let aSa\in S. Then a=α+βa=\alpha+\beta for suitable α𝒜S\alpha\in\mathcal{A}_{S} and β𝒜S\beta\in\mathcal{A}_{S}^{\prime}. Fix \mathbb{Z}-bases S\mathcal{B}_{S} of 𝒜S\mathcal{A}_{S} and S\mathcal{B}_{S}^{\prime} of 𝒜S\mathcal{A}_{S}^{\prime}. Let α=i=1rkiαi\alpha=\sum_{i=1}^{r}k_{i}\alpha_{i} and β=j=1sjβj\beta=\sum_{j=1}^{s}\ell_{j}\beta_{j}, where αiS\alpha_{i}\in\mathcal{B}_{S}, βjS\beta_{j}\in\mathcal{B}_{S}^{\prime} and ki,jk_{i},\ell_{j}\in\mathbb{Z}. Since αiS𝒜\alpha_{i}\in\mathcal{B}_{S}\subset\mathcal{A}, there exist nin_{i}\in\mathbb{N} and viKv_{i}\in K such that δ(vi)=niαivi\delta(v_{i})=n_{i}\alpha_{i}v_{i}. Consider the field E=K(tS)(v11/n1,v21/n2,,vr1/nr)E=K(t_{S})(v_{1}^{1/n_{1}},v_{2}^{1/n_{2}},\dots,v_{r}^{1/n_{r}}). The derivation δK(tS)\delta_{K(t_{S})} extends uniquely to a derivation δE\delta_{E} of EE, and

δE(i=1rviki/ni)=i=1rδE(viki/ni)jivjkj/nj=i=1r(ki/ni)viki/nivi1δE(vi)jivjkj/nj=i=1r(ki/ni)viki/niniαijivjkj/nj=i=1rkiαij=1rvjkj/nj=αi=1rviki/ni\begin{split}\delta_{E}\left(\prod_{i=1}^{r}v_{i}^{k_{i}/n_{i}}\right)&=\sum_{i=1}^{r}\delta_{E}(v_{i}^{k_{i}/n_{i}})\prod_{j\neq i}v_{j}^{k_{j}/n_{j}}\\ &=\sum_{i=1}^{r}(k_{i}/n_{i})v_{i}^{k_{i}/n_{i}}v_{i}^{-1}\delta_{E}(v_{i})\prod_{j\neq i}v_{j}^{k_{j}/n_{j}}\\ &=\sum_{i=1}^{r}(k_{i}/n_{i})v_{i}^{k_{i}/n_{i}}n_{i}\alpha_{i}\prod_{j\neq i}v_{j}^{k_{j}/n_{j}}\\ &=\sum_{i=1}^{r}k_{i}\alpha_{i}\prod_{j=1}^{r}v_{j}^{k_{j}/n_{j}}=\alpha\prod_{i=1}^{r}v_{i}^{k_{i}/n_{i}}\end{split}

and a similar calculation yields

δE(j=1stβjj)=βj=1stβjj.\begin{split}\delta_{E}\left(\prod_{j=1}^{s}t_{\beta_{j}}^{\ell_{j}}\right)=\beta\prod_{j=1}^{s}t_{\beta_{j}}^{\ell_{j}}.\end{split}

Now,

δE(i=1rviki/nij=1stβjj)=δE(i=1rviki/ni)j=1stβjj+i=1rviki/niδE(j=1stβjj)=αi=1rviki/nij=1stβjj+i=1rviki/niβj=1stβjj=(α+β)i=1rviki/nij=1stβjj=ai=1rviki/nij=1stβjj.\begin{split}\delta_{E}\left(\prod_{i=1}^{r}v_{i}^{k_{i}/n_{i}}\prod_{j=1}^{s}t_{\beta_{j}}^{\ell_{j}}\right)&=\delta_{E}\left(\prod_{i=1}^{r}v_{i}^{k_{i}/n_{i}}\right)\prod_{j=1}^{s}t_{\beta_{j}}^{\ell_{j}}+\prod_{i=1}^{r}v_{i}^{k_{i}/n_{i}}\delta_{E}\left(\prod_{j=1}^{s}t_{\beta_{j}}^{\ell_{j}}\right)\\ &=\alpha\prod_{i=1}^{r}v_{i}^{k_{i}/n_{i}}\prod_{j=1}^{s}t_{\beta_{j}}^{\ell_{j}}+\prod_{i=1}^{r}v_{i}^{k_{i}/n_{i}}\beta\prod_{j=1}^{s}t_{\beta_{j}}^{\ell_{j}}\\ &=(\alpha+\beta)\prod_{i=1}^{r}v_{i}^{k_{i}/n_{i}}\prod_{j=1}^{s}t_{\beta_{j}}^{\ell_{j}}=a\prod_{i=1}^{r}v_{i}^{k_{i}/n_{i}}\prod_{j=1}^{s}t_{\beta_{j}}^{\ell_{j}}.\end{split}

Thus for each aSa\in S, the equation δ(y)=ay\delta(y)=ay has a solution i=1rviki/nij=1stβjjE\prod_{i=1}^{r}v_{i}^{k_{i}/n_{i}}\prod_{j=1}^{s}t_{\beta_{j}}^{\ell_{j}}\in E. Clearly EE is an algebraic extension of K(tS)K(t_{S}) and trdegK(E)=trdegK(K(tS))=rank(𝒜S){\rm trdeg}_{K}(E)={\rm trdeg}_{K}(K(t_{S}))={\rm rank}(\mathcal{A}_{S}^{\prime}). ∎

We note that the degree of the finite extension EE over K(tS)K(t_{S}) is at most n1n2nrn_{1}n_{2}\dots n_{r}. We may choose ni=nain_{i}=n_{a_{i}}, where nain_{a_{i}} is the smallest positive integer such that δ(y)=naiaiy\delta(y)=n_{a_{i}}a_{i}y has a solution in KK, to obtain a better bound on the degree of EE over K(tS)K(t_{S}).

Thus far, we have constructed a differential field extension of (K,δ)(K,\delta) which contains solutions of differential equations of the form δ(y)=ay\delta(y)=ay where aSa\in S, and SK\{0}S\subseteq K\backslash\{0\} is a finite set. We shall now focus on solving the system of equations δ(y1)=ay1,δ(yi)=yi1+ayi,for i2\delta(y_{1})=ay_{1},\delta(y_{i})=y_{i-1}+ay_{i},\text{for }i\geq 2.

Lemma 3.7.

Let aKa\in K and (L,δL)(L,\delta_{L}) be a differential extension of (K,δ)(K,\delta) that contains a solution uau_{a} of δ(y)=ay\delta(y)=ay. Let n2n\geq 2 be an integer. Consider the following system of equations over LL.

(3.4) y1=ua,δL(yi)=yi1+ayi, for 2iny_{1}=u_{a},\delta_{L}(y_{i})=y_{i-1}+ay_{i},\text{ for }2\leq i\leq n

Then the following are equivalent.

  1. (i).

    There exists wLw\in L such that δL(w)=1\delta_{L}(w)=1.

  2. (ii).

    The differential field (L,δL)(L,\delta_{L}) contains a solution of 3.4.

  3. (iii).

    An algebraic extension of (L,δL)(L,\delta_{L}) contains a solution of 3.4.

Further, if these equivalent conditions do not hold, then there exists a differential extension (E,δE)(E,\delta_{E}) over (L,δL)(L,\delta_{L}) such that trdegL(E)=1{\rm trdeg}_{L}(E)=1 and EE contains a solution of 3.4.

Proof.

Suppose wLw\in L is such that δL(w)=1\delta_{L}(w)=1. The equation corresponding to i=2i=2 in 3.4 is δL(y2)=ua+ay2\delta_{L}(y_{2})=u_{a}+ay_{2}. This can be rewritten as δL(y2ua1)=1\delta_{L}(y_{2}u_{a}^{-1})=1. Since wLw\in L is such that δL(w)=1\delta_{L}(w)=1, there exists c2CLc_{2}\in C_{L} such that y2ua1=w+c2y_{2}u_{a}^{-1}=w+c_{2}. Thus y2=(w+c2)uaLy_{2}=(w+c_{2})u_{a}\in L is a solution of δL(y2)=ua+ay2\delta_{L}(y_{2})=u_{a}+ay_{2}. Now for i=3i=3, the equation is δL(y3)=y2+ay3\delta_{L}(y_{3})=y_{2}+ay_{3}. Substituting y2=(w+c2)uaLy_{2}=(w+c_{2})u_{a}\in L, the equation is δL(y3)=(w+c2)ua+ay3\delta_{L}(y_{3})=(w+c_{2})u_{a}+ay_{3}. Again, this can be rewritten as δL(y3ua1)=w+c2\delta_{L}(y_{3}u_{a}^{-1})=w+c_{2}. Since w+c2=δL(w2/2+c2w)w+c_{2}=\delta_{L}(w^{2}/2+c_{2}w), there exists c3CLc_{3}\in C_{L} such that y3ua1=w2/2+c2w+c3y_{3}u_{a}^{-1}=w^{2}/2+c_{2}w+c_{3}. Thus y3=(w2/2+c2w+c3)uaLy_{3}=(w^{2}/2+c_{2}w+c_{3})u_{a}\in L is a solution of δL(y3)=y2+ay3\delta_{L}(y_{3})=y_{2}+ay_{3}. Proceeding this way,

yi=(j=0i1cij(j!)1wj)uaL,y_{i}=\left(\sum_{j=0}^{i-1}c_{i-j}(j!)^{-1}w^{j}\right)u_{a}\in L,

where c1=1c_{1}=1 and cijc_{i-j} are suitable constants in LL, is a solution set of the system of equations 3.4.

Now, suppose (M,δM)(M,\delta_{M}) is an algebraic extension of (L,δL)(L,\delta_{L}) that contains a solution of 3.4. In particular, there exists z2Mz_{2}\in M such that δM(z2)=ua+az2\delta_{M}(z_{2})=u_{a}+az_{2}. Let u=z2ua1Mu=z_{2}u_{a}^{-1}\in M. Then δM(u)=1\delta_{M}(u)=1. If uLu\in L then w=uw=u satisfies δL(w)=1\delta_{L}(w)=1. Thus we may assume that uLu\notin L. Since MM is an algebraic extension of LL, we consider the minimal monic polynomial f(x)=xk+ak1xk1++a1x+a0f(x)=x^{k}+a_{k-1}x^{k-1}+\dots+a_{1}x+a_{0} of ww over LL. Thus,

f(u)=uk+ak1uk1++a1u+a0=0.f(u)=u^{k}+a_{k-1}u^{k-1}+\dots+a_{1}u+a_{0}=0.

Applying δM\delta_{M} on both sides of this equation,

δM(f(u))=(k+δL(ak1))uk1++(2a2+δL(a1))u+a1+δL(a0)=0.\delta_{M}(f(u))=(k+\delta_{L}(a_{k-1}))u^{k-1}+\dots+(2a_{2}+\delta_{L}(a_{1}))u+a_{1}+\delta_{L}(a_{0})=0.

If k+δL(ak1)0k+\delta_{L}(a_{k-1})\neq 0, then we arrive at a contradiction to the minimality of f(x)f(x), unless k=1k=1. However, k=1k=1 too, leads to a contradiction since uLu\notin L. Thus k+δL(ak1)=0k+\delta_{L}(a_{k-1})=0 and w=k1ak1Lw=-k^{-1}a_{k-1}\in L satisfies δL(w)=1\delta_{L}(w)=1.

This establishes the equivalence of three conditions.

If these equivalent conditions do not hold then we extend the derivation δL\delta_{L} to the field E=L(t)E=L(t), by defining δE(t)=1\delta_{E}(t)=1. Since w=tEw=t\in E satisfies δE(w)=1\delta_{E}(w)=1, by the above equivalent conditions, the system of equations 3.4 has a solution in (E,δE)(E,\delta_{E}). ∎

Corollary 3.8.

Let (K,δ)(K,\delta) be a differential field and (E,δE)(E,\delta_{E}) be an algebraic extension of (K,δ)(K,\delta). Then there exists wEw\in E such that δE(w)=1\delta_{E}(w)=1 if and only if there exists vKv\in K such that δ(v)=1\delta(v)=1.

Lemma 3.9.

Let (K,δ)(K,\delta) be a differential field such that δ(z)1\delta(z)\neq 1 for every zKz\in K. Let L=K(w)L=K(w) be a simple transcendental extension of KK. Let δL\delta_{L} be an extension of δ\delta to LL such that δL(w)=1\delta_{L}(w)=1. Let

𝒜K={aK:δ(y)=nay has a nonzero solution in K for some n},and\mathcal{A}_{K}=\{a\in K:\delta(y)=nay\text{ has a nonzero solution in }K\text{ for some }n\in\mathbb{N}\},and
𝒜L={aL:δ(y)=nay has a nonzero solution in L for some n}.\mathcal{A}_{L}=\{a\in L:\delta(y)=nay\text{ has a nonzero solution in }L\text{ for some }n\in\mathbb{N}\}.

Then 𝒜K=𝒜LK\mathcal{A}_{K}=\mathcal{A}_{L}\cap K.

Proof.

Clearly 𝒜K𝒜LK\mathcal{A}_{K}\subseteq\mathcal{A}_{L}\cap K. For the other inclusion, let a𝒜LKa\in\mathcal{A}_{L}\cap K. Let vLv\in L and nn\in\mathbb{N} be such that δL(v)=nav\delta_{L}(v)=nav. We write v=fg1v=fg^{-1}, where f=i=0fiwiK[w]f=\sum_{i=0}^{\ell}f_{i}w^{i}\in K[w] and g=j=0mgjwjK[w]g=\sum_{j=0}^{m}g_{j}w^{j}\in K[w] are polynomials with f0f_{\ell}\neq 0 and gm0g_{m}\neq 0. From δL(v)=nav\delta_{L}(v)=nav, we have δL(f)gfδL(g)=nafg\delta_{L}(f)g-f\delta_{L}(g)=nafg. Substituting f=i=0fiwi,g=j=0mgjwjf=\sum_{i=0}^{\ell}f_{i}w^{i},~{}g=\sum_{j=0}^{m}g_{j}w^{j} and comparing coefficients of wl+mw^{l+m}, we get δ(f)gmfδ(gm)=nafgm\delta(f_{\ell})g_{m}-f_{\ell}\delta(g_{m})=naf_{\ell}g_{m}. Consequently, δ(fgm1)=na(fgm1)\delta(f_{\ell}g_{m}^{-1})=na(f_{\ell}g_{m}^{-1}). Thus fgm1Kf_{\ell}g_{m}^{-1}\in K is a solution of equation δ(y)=nay\delta(y)=nay, hence a𝒜Ka\in\mathcal{A}_{K}. ∎

The next corollary is a counterpart of Corollary 3.5.

Corollary 3.10.

Let (K,δ)(K,\delta) be a differential field such that δ(w)1\delta(w)\neq 1 for every wKw\in K. Let SK\{0}S\subseteq K\backslash\{0\} be a finite set. Let (E,δE)(E,\delta_{E}) be a differential extension of (K,δ)(K,\delta) such that for each aS\{0}a\in S\backslash\{0\}, the field EE contains solutions of the set of equations

(3.5) δ(y1)=ay1,δ(y2)=y1+ay2.\delta(y_{1})=ay_{1},\delta(y_{2})=y_{1}+ay_{2}.

Then trdegK(E)rank(𝒜S)+1{\rm trdeg}_{K}(E)\geq{\rm rank}(\mathcal{A}_{S}^{\prime})+1.

Proof.

Since (E,δ)(E,\delta) is a splitting field of SS, for each aSa\in S the field EE contains solutions of the equations δ(y2)=y1+ay2\delta(y_{2})=y_{1}+ay_{2}. Thus, by Lemma 3.7, there exists a transcendental element wEw\in E such that δE(w)=1\delta_{E}(w)=1. Let L=K(w)L=K(w) and δL\delta_{L} be the restriction of δE\delta_{E} to LL. Since KK does not contain an element whose derivative is 11, by Lemma 3.9, 𝒜=𝒜K=𝒜LK\mathcal{A}=\mathcal{A}_{K}=\mathcal{A}_{L}\cap K. Taking intersection with SS and noting that SKS\subseteq K, we obtain 𝒜S=(𝒜L)S\mathcal{A}_{S}=(\mathcal{A}_{L})_{S}. Therefore rank(𝒜S)=rank((𝒜L)S){\rm rank}(\mathcal{A}_{S}^{\prime})={\rm rank}((\mathcal{A}_{L})^{\prime}_{S}). Now, by Corollary 3.5, trdegL(E)rank((𝒜L)S)=rank(𝒜S){\rm trdeg}_{L}(E)\geq{\rm rank}((\mathcal{A}_{L})_{S}^{\prime})={\rm rank}(\mathcal{A}_{S}^{\prime}). Thus, trdegK(E)=trdegL(E)+trdegK(L)rank(𝒜S)+trdegK(L)=rank(𝒜S)+1{\rm trdeg}_{K}(E)={\rm trdeg}_{L}(E)+{\rm trdeg}_{K}(L)\geq{\rm rank}(\mathcal{A}_{S}^{\prime})+{\rm trdeg}_{K}(L)={\rm rank}(\mathcal{A}_{S}^{\prime})+1. ∎

4. Finite Splitting Fields of Differential Central Simple Algebras

Throughout this section, KK denotes the field F(t)F(t), where FF is a field of characteristic 0 and δ\delta denotes the derivation on KK determined by δ(t)=t\delta(t)=t and δ(a)=0\delta(a)=0 for each aFa\in F. In this section, we discuss the splitting of differential matrix algebras (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) over (K,δ)(K,\delta), where PP is a traceless matrix over KK and 𝒟P=δc+P\mathcal{D}_{P}=\delta^{c}+\partial_{P} as in §2. First, we shall look at the case when PMn(F)P\in M_{n}(F), and later consider other cases. Following lemma will be useful in proofs of main theorems.

Lemma 4.1.

Let 𝒜={aK:δ(y)=nay has a nonzero solution in K for some n}\mathcal{A}=\{a\in K:\delta(y)=nay\text{ has a nonzero solution in }K\text{ for some }n\in\mathbb{N}\}. Then F𝒜=F\cap\mathcal{A}=\mathbb{Q}.

Proof.

Let aa\in\mathbb{Q}. Since δ(t)=t\delta(t)=t and 𝒜\mathcal{A} is a \mathbb{Q}-vector space, 𝒜\mathbb{Q}\subseteq\mathcal{A}. Since char(F)=0{\rm char}(F)=0, F𝒜\mathbb{Q}\subseteq F\cap\mathcal{A}.

Now, suppose aF𝒜a\in F\cap\mathcal{A}. Then there exist f,gF[t]f,g\in F[t]; g0g\neq 0, such that δ(f/g)=naf/g\delta(f/g)=naf/g for some nn\in\mathbb{N}. We rewrite it as

δ(f)gfδ(g)g2=nafg.\frac{\delta(f)g-f\delta(g)}{g^{2}}=\frac{naf}{g}.

Let f=at++a1t+a0f=a_{\ell}t^{\ell}+\dots+a_{1}t+a_{0} and g=bmtm++b1t+b0g=b_{m}t^{m}+\dots+b_{1}t+b_{0}, where ai,bjFa_{i},b_{j}\in F. Then δ(f)=at++a1t\delta(f)={\ell}a_{\ell}t^{\ell}+\dots+a_{1}t and δ(g)=mbmtm++b1t\delta(g)=mb_{m}t^{m}+\dots+b_{1}t. Substituting it in the above equation,

(at++a1t)(bmtm++b1t+b0)(at++a1t+a0)(mbmtm++b1t)=na(at++a1t+a0)(bmtm++b1t+b0).\begin{split}({\ell}a_{\ell}t^{\ell}+\dots+a_{1}t)(b_{m}t^{m}+\dots+b_{1}t+b_{0})-(a_{\ell}t^{\ell}+\dots+a_{1}t+a_{0})(mb_{m}t^{m}+\dots+b_{1}t)\\ =na(a_{\ell}t^{\ell}+\dots+a_{1}t+a_{0})(b_{m}t^{m}+\dots+b_{1}t+b_{0}).\end{split}

Let ii be the smallest index such that ai0a_{i}\neq 0 and jj be the smallest index such that bj0b_{j}\neq 0. Comparing coefficients of ti+jt^{i+j},

iaibjjaibj=naaibj.ia_{i}b_{j}-ja_{i}b_{j}=naa_{i}b_{j}.

Thus a=(ij)/na=(i-j)/n\in\mathbb{Q}. ∎

Theorem 4.2.

Let PMn(F)P\in M_{n}(F) be a traceless matrix. The differential matrix algebra (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) admits a finite splitting field if and only if PP is diagonalizable and every eigenvalue of PP is rational.

Proof.

We first assume that PP is diagonalizable, and every eigenvalue of PP is in \mathbb{Q}. Let D=diag(λ1,λ2,,λn)D={\rm diag}(\lambda_{1},\lambda_{2},\cdots,\lambda_{n}) be a diagonal matrix with eigenvalues of PP, with multiplicity, as diagonal entries. Let S={λ1,,λn}S=\{\lambda_{1},\dots,\lambda_{n}\} and

𝒜={aK:δ(y)=nay has a nonzero solution in K for some n}.\mathcal{A}=\{a\in K:\delta(y)=nay\text{ has a nonzero solution in }K\text{ for some }n\in\mathbb{N}\}.

Since δ(t)=t\delta(t)=t, by Lemma 4.1 S𝒜S\subseteq\mathcal{A}. Hence 𝒜S=0\mathcal{A}_{S}^{\prime}=0. We recall that 𝒜S\mathcal{A}_{S}^{\prime} is determined by the direct sum 𝒜S+𝒜S=S\mathcal{A}\cap\langle S\rangle+\mathcal{A}_{S}^{\prime}=\langle S\rangle. Thus by Theorem 3.3, there exists a splitting field EE of SS such that E/KE/K is finite. Thus δ(y)=λiy\delta(y)=\lambda_{i}y has a solution uiE{0}u_{i}\in E\setminus\{0\}, for each ii. Consequently, the equation δc(Y)=DY\delta^{c}(Y)=DY has a solution diag(u1,u2,,un)GLn(E){\rm diag}(u_{1},u_{2},\dots,u_{n})\in{\rm GL}_{n}(E), where Y=(yij)Y=(y_{ij}) is an n×nn\times n matrix of indeterminates. Now, from Lemma 2.4, the equation δc(Y)=PtY\delta^{c}(Y)=P^{t}Y has a solution in GLn(E){\rm GL}_{n}(E). By Corollary 2.3, the extension E/KE/K splits (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}).

Conversely, suppose that PP is not diagonalizable or there is an eigenvalue of PP that is not in \mathbb{Q}. We consider two cases.

Case 1 : Each eigenvalue λi\lambda_{i} of PP is rational.

Let (E,δE)(E,\delta_{E}) be a finite differential extension of (K,δ)(K,\delta). Let J=diag(J1,J2,,Jk)J={\rm diag}(J_{1},J_{2},\cdots,J_{k}) be a Jordan canonical form of PP, where JiJ_{i} is an upper triangular Jordan block. As PP is not diagonalizable, at least one JiJ_{i} has its length strictly greater than 11. We assume that the length of J1J_{1} is >1\ell>1 and that the eigenvalue that corresponds to J1J_{1} is λ1\lambda_{1}.

Since E/KE/K is finite and there does not exist an element vKv\in K such that δ(v)=1\delta(v)=1, by Corollary 3.8 there does not exist wEw\in E such that δE(w)=1\delta_{E}(w)=1. By Lemma 3.7, the equations δ(y1)=λ1y1\delta(y_{1})=\lambda_{1}y_{1} and δ(y2)=y1+λ1y2\delta(y_{2})=y_{1}+\lambda_{1}y_{2} do not have solutions in EE.

Thus the equation δc(Y)=JtY\delta^{c}(Y)=J^{t}Y has no solution over EE, where Y=(yij)Y=(y_{ij}) is an n×nn\times n matrix of indeterminates. Now, from Lemma 2.4, the equation δc(Y)=PtY\delta^{c}(Y)=P^{t}Y has no solution over EE. By Corollary 2.3, the extension EE does not split (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}).

Case 2 : At least one eigenvalue of PP is not rational.

Let F¯\overline{F} be an algebraic closure of FF. Let SF¯S\subseteq\overline{F} be the set of eigenvalues of PP. Suppose (E,δE)(E,\delta_{E}) is a finite differential extension of (F¯(t),δF¯(t))(\overline{F}(t),\delta_{\overline{F}(t)}). Let

𝒜={aF¯(t):δ(y)=nay has a nonzero solution in F¯(t) for some n}.\mathcal{A}=\{a\in\overline{F}(t):\delta(y)=nay\text{ has a nonzero solution in }\overline{F}(t)\text{ for some }n\in\mathbb{N}\}.

By Lemma 4.1, F¯𝒜=\overline{F}\cap\mathcal{A}=\mathbb{Q}. Thus, if λ\lambda is a nonrational eigenvalue of PP, then λ𝒜\lambda\notin\mathcal{A}. Therefore, 𝒜S0\mathcal{A}_{S}^{\prime}\neq 0. Since EE is a finite extension over F¯(t)\overline{F}(t), by Theorem 3.3, the equation δ(y)=λy\delta(y)=\lambda y does not have a solution in EE. Consequently, the equation δc(Y)=JtY\delta^{c}(Y)=J^{t}Y has no solution over EE, where JJ is a Jordan canonical form of PP and Y=(yij)Y=(y_{ij}) is an n×nn\times n matrix of indeterminates. Now, from Lemma 2.4, the equation δc(Y)=PtY\delta^{c}(Y)=P^{t}Y has no solution over EE. Therefore, by Corollary 2.3, (E,δE)(E,\delta_{E}) does not split (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}).

Now, if (L,δL)(L,\delta_{L}) is a finite extension of (K,δ)(K,\delta) that splits (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}), then the compositum of (L,δL)(L,\delta_{L}) and (F¯(t),δF¯(t))(\overline{F}(t),\delta_{\overline{F}(t)}) is a finite extension of (F¯(t),δF¯(t))(\overline{F}(t),\delta_{\overline{F}(t)}) splitting (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}), which is a contradiction. Thus a finite extension of (K,δ)(K,\delta) does not split (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}). ∎

Corollary 4.3.

Let PMn(F)P\in M_{n}(F) be a traceless matrix. If the differential matrix algebra (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) admits a finite splitting field (E,δE)(E,\delta_{E}) over (K,δ)(K,\delta), then degK(E)[S{1}:]{\rm deg}_{K}(E)\geq[\langle S\cup\{1\}\rangle:\mathbb{Z}], where SS\subseteq\mathbb{Q} is the set of eigenvalues of PP.

Proof.

Let SF¯S\subseteq\overline{F} be the set of eigenvalues of PP. Since (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) admits a finite splitting field, by Theorem 4.2, SS\subseteq\mathbb{Q}. Thus, S\langle S\rangle\subseteq\mathbb{Q} is a cyclic group. Let qq\in\mathbb{N} be such that 1/q1/q generates S\langle S\rangle. If (E,δE)(E,\delta_{E}) is a finite extension of (K,δ)(K,\delta) that splits (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}), the equation δ(y)=(1/q)y\delta(y)=(1/q)y has a solution zEz\in E. Thus δE(zq)=zq\delta_{E}(z^{q})=z^{q} and consequently, zq=ctz^{q}=ct for some cCEc\in C_{E}. Since E/KE/K is algebraic, by [Mag94, Lemma 3.19], cc is algebraic over CK=FC_{K}=F.

Let II be the ideal of F(c)[t]F(c)[t] generated by tt. Then ctIct\in I but ctI2ct\notin I^{2} and by generalized Eisenstein criterion, the polynomial XqctK(c)[X]X^{q}-ct\in K(c)[X] is irreducible. Therefore, degK(E)degK(c)(E)q=[S{1}:]{\rm deg}_{K}(E)\geq{\rm deg}_{K(c)}(E)\geq q=[\langle S\cup\{1\}\rangle:\mathbb{Z}]. ∎

Let us examine n=2n=2 case. If PM2(F)P\in M_{2}(F) is a traceless matrix, then (M2(K),𝒟P)(M_{2}(K),\mathcal{D}_{P}) is split by a finite extension of (K,δ)(K,\delta) if and only if det(P)-\det(P) is a square in \mathbb{Q}^{*}. Further, in that case, if det(P)=p2/q2-\det(P)=p^{2}/q^{2} is in reduced fractional form, then there exists a finite extension of degree qq that splits (M2(K),𝒟P)(M_{2}(K),\mathcal{D}_{P}). We observe that q=[S{1}:]q=[\langle S\cup\{1\}\rangle:\mathbb{Z}], where S={p/q,p/q}S=\{-p/q,p/q\}. More generally, in the above corollary the lower bound [S{1}:][\langle S\cup\{1\}\rangle:\mathbb{Z}] on degK(E){\rm deg}_{K}(E) is sharp, which is attained when cKc\in K and E=K(z)E=K(z). The following corollary can be derived from q=1q=1 case.

Corollary 4.4.

Let PMn(F)P\in M_{n}(F) be a traceless matrix and (A,𝒟P)(A,\mathcal{D}_{P}) be the differential matrix algebra of degree nn over KK. The following are equivalent.

  1. (i).

    (A,𝒟P)(A,\mathcal{D}_{P}) is split.

  2. (ii).

    (A,𝒟P)(A,\mathcal{D}_{P}) is split over (F¯(t),δF¯(t))(\overline{F}(t),\delta_{\overline{F}(t)}).

  3. (iii).

    PP is diagonalizable and all the eigenvalues of PP are in \mathbb{{Z}}.

4.1. Finite splitting of (A,𝒟P)(A,\mathcal{D}_{P}) when δc(P)0\delta^{c}(P)\neq 0

In this section also, we assume the field (K,δ)(K,\delta) to be (F(t),δ)(F(t),\delta) such that δ(a)=0\delta(a)=0 for every aFa\in F and δ(t)=t\delta(t)=t. Let (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) be the differential matrix algebra, where δc(P)0\delta^{c}(P)\neq 0, i.e., at least one entry of PP is a nonconstant element in F(t)F(t).

In the following theorem, we find some conditions on PP which ensure that (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is split by a finite extension.

Theorem 4.5.

Let PMn(K)P\in M_{n}(K) be a traceless matrix. Then (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is split by a finite extension if PP is of one of the following forms.

  1. (i).

    P=diag(d1,d2,,dn)P={\rm diag}(d_{1},d_{2},\dots,d_{n}) and each di𝒜d_{i}\in\mathcal{A}.

  2. (ii).

    P=(pij)P=(p_{ij}), pijF[t]p_{ij}\in F[t] is an upper triangular matrix satisfying piip_{ii}\in\mathbb{Q} and piipjj/p_{ii}\mathbb{Z}\neq p_{jj}\mathbb{Z}\in\mathbb{Q}/\mathbb{Z}, unless i=ji=j.

Proof.

(i).(i). Since P=diag(d1,d2,,dn)P=\mathrm{diag}(d_{1},d_{2},\dots,d_{n}), the set of eigenvalues of PP is S={d1,d2,,dn}S=\{d_{1},d_{2},\dots,d_{n}\}, which, by hypothesis, is contained in 𝒜\mathcal{A}. Thus 𝒜S=0\mathcal{A}_{S}^{\prime}=0 and from Theorem 3.3, there exists a finite extension EE over KK such that the equation δ(y)=diy\delta(y)=d_{i}y has a solution uiE{0}u_{i}\in E\setminus\{0\}. Thus the equation δc(Y)=PtY\delta^{c}(Y)=P^{t}Y has a solution diag(u1,u2,,un)GLn(E){\rm diag}(u_{1},u_{2},\dots,u_{n})\in{\rm GL}_{n}(E), where Y=(yij)Y=(y_{ij}) is an n×nn\times n matrix of indeterminates. By Corollary 2.3, the extension E/KE/K splits (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}).

(ii).(ii). Since PP is upper triangular, S:={p11,p22,,pnn}S:=\{p_{11},p_{22},\cdots,p_{nn}\} is the set of eigenvalues of PP. By hypothesis, SS\subseteq\mathbb{Q} and by Lemma 4.1, =F𝒜\mathbb{Q}=F\cap\mathcal{A}. Thus, S𝒜S\subseteq\mathcal{A} and 𝒜S=0\mathcal{A}_{S}^{\prime}=0. Now by Theorem 3.3, there exists a finite extension E/KE/K and uiE{0}u_{i}\in E\setminus\{0\} such that δE(ui)=piiui\delta_{E}(u_{i})=p_{ii}u_{i}. We show that EE splits (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}). By Corollary 2.3, it suffices to show that the equation δc(Y)=PtY\delta^{c}(Y)=P^{t}Y has a solution in GLn(E){\rm GL}_{n}(E). It is enough to find a lower triangular matrix with nonzero diagonal entries satisfying δc(Y)=PtY\delta^{c}(Y)=P^{t}Y. Let Y=(yij)Y=(y_{ij}), where yijy_{ij} are indeterminates, and write the jthj^{\rm th} column of the above system of equations

(4.1) δ(yij)=k=jipkiykj\delta(y_{ij})=\sum_{k=j}^{i}p_{ki}y_{kj}

for jinj\leq i\leq n. We construct a lower triangular matrix Z=(zij)Z=(z_{ij}), with zjj=ujE{0}z_{jj}=u_{j}\in E\setminus\{0\} as a solution of this system.

For each jnj\leq n, we use induction on ii, with jinj\leq i\leq n, to show that there exist rijCE[t]r_{ij}\in C_{E}[t] such that the elements zjj,,zij,,znjEz_{jj},\cdots,z_{ij},\cdots,z_{nj}\in E defined by zij:=rijujEz_{ij}:=r_{ij}u_{j}\in E, are solutions of the system of equations in 4.1. For i=ji=j, this is evident with rjj=1r_{jj}=1 and zjj=ujz_{jj}=u_{j}, since the corresponding equation in the system 4.1 is δ(yjj)=pjjyjj\delta(y_{jj})=p_{jj}y_{jj}. By induction hypothesis, assume that for jin1j\leq i\leq n-1, there exist rijCE[t]r_{ij}\in C_{E}[t] such that zij:=rijujEz_{ij}:=r_{ij}u_{j}\in E satisfies first njn-j equations of 4.1. The last equation is

δ(ynj)=k=jnpknykj.\delta(y_{nj})=\sum_{k=j}^{n}p_{kn}y_{kj}.

Substituting ykj=zkj:=rkjujy_{kj}=z_{kj}:=r_{kj}u_{j}, whenever jkn1j\leq k\leq n-1, we get

δ(ynj)pnnynj=k=jn1pknrkjuj.\delta(y_{nj})-p_{nn}y_{nj}=\sum_{k=j}^{n-1}p_{kn}r_{kj}u_{j}.

We multiply this equation by un1u_{n}^{-1} and use δE(un1)=pnnun1\delta_{E}(u_{n}^{-1})=-p_{nn}u_{n}^{-1}, to obtain

(4.2) δ(ynjun1)=δ(ynj)un1pnnynjun1=(k=jn1pknrkj)ujun1.\delta(y_{nj}u_{n}^{-1})=\delta(y_{nj})u_{n}^{-1}-p_{nn}y_{nj}u_{n}^{-1}=\left(\sum_{k=j}^{n-1}p_{kn}r_{kj}\right)u_{j}u_{n}^{-1}.

We expand the polynomial k=jn1pknrkjCE[t]\sum_{k=j}^{n-1}p_{kn}r_{kj}\in C_{E}[t] in powers of tt and write

k=jn1pknrkj=l=1manjltl,\sum_{k=j}^{n-1}p_{kn}r_{kj}=\sum_{l=1}^{m}a_{njl}t^{l},

where anjlCEa_{njl}\in C_{E}. Thus

(4.3) δ(ynjun1)=l=1manjltlujun1.\delta(y_{nj}u_{n}^{-1})=\sum_{l=1}^{m}a_{njl}t^{l}u_{j}u_{n}^{-1}.

Since for each jn1j\leq n-1, pnnpjjp_{nn}\neq p_{jj} modulo \mathbb{Z}, the rational number l+pjjpnnl+p_{jj}-p_{nn} is nonzero for each ll\in\mathbb{Z}. We compute

δE((l+pjjpnn)1anjltlujun1)=anjltlujun1.\delta_{E}((l+p_{jj}-p_{nn})^{-1}a_{njl}t^{l}u_{j}u_{n}^{-1})=a_{njl}t^{l}u_{j}u_{n}^{-1}.

Making this substitution in 4.3,

δ(ynjun1)=δE(l=1m(l+pjjpnn)1anjltlujun1).\delta(y_{nj}u_{n}^{-1})=\delta_{E}\left(\sum_{l=1}^{m}(l+p_{jj}-p_{nn})^{-1}a_{njl}t^{l}u_{j}u_{n}^{-1}\right).

We define rnj:=l=1m(l+pjjpnn)1anjltlCE[t]r_{nj}:=\sum_{l=1}^{m}(l+p_{jj}-p_{nn})^{-1}a_{njl}t^{l}\in C_{E}[t] and rewrite the above equation as

δ(ynjun1)=δE(rnjujun1).\delta(y_{nj}u_{n}^{-1})=\delta_{E}\left(r_{nj}u_{j}u_{n}^{-1}\right).

Thus znj:=rnjuj0z_{nj}:=r_{nj}u_{j}\neq 0 is a solution of 4.3.

Finally, we vary jj to get zij:=rijujEz_{ij}:=r_{ij}u_{j}\in E, for each 1jin1\leq j\leq i\leq n, with rjj=1r_{jj}=1. such that the lower triangular matrix Z:=(zij)Mn(E)Z:=(z_{ij})\in M_{n}(E) satisfies δc(Y)=ptY\delta^{c}(Y)=p^{t}Y. Since each zjj=ujz_{jj}=u_{j} is nonzero, ZGLn(E)Z\in{\rm GL}_{n}(E). ∎

4.2. Minimal transcendence degree of a splitting field.

In this subsection, we are concerned about finding transcendence degrees of splitting fields of differential matrix algebras which do not admit a finite splitting field.

In the following theorem, we assume our field to be K=F(t)K=F(t) with derivation δ\delta such that δ(a)=0\delta(a)=0 for each aFa\in F and δ(t)=t\delta(t)=t.

Theorem 4.6.

Let 0PMn(F)0\neq P\in M_{n}(F) be a traceless matrix. Let SF¯S\subseteq\overline{F} be the set of eigenvalues of PP and k=dim(span(S{1}))k={\rm dim}_{\mathbb{Q}}({\rm span}_{\mathbb{Q}}(S\cup\{1\})). Let (E,δE)(E,\delta_{E}) be a differential field extension of (K,δ)(K,\delta) that splits the differential matrix algebra (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}).

  1. (i).

    If PP is diagonalizable over F¯\overline{F}, then trdegK(E)k1{\rm trdeg}_{K}(E)\geq k-1.

  2. (ii).

    If PP is not diagonalizable over F¯\overline{F}, then trdegK(E)k{\rm trdeg}_{K}(E)\geq k.

Proof.

Since trdegK(E)=trdegF¯(t)(EF¯){\rm trdeg}_{K}(E)={\rm trdeg}_{\overline{F}(t)}(E\overline{F}), where EF¯E\overline{F} is the subfield of E¯\overline{E} generated by EE and F¯\overline{F}, we may assume that FF is algebraically closed.

(i).(i). Since PP is diagonalizable over F¯=F\overline{F}=F and (E,δE)(E,\delta_{E}) splits (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}), as in the proof of Theorem 4.2, for each λS\lambda\in S, the field EE contains solutions of δ(y)=λy\delta(y)=\lambda y. Further, 𝒜S=S𝒜F𝒜=\mathcal{A}_{S}=\langle S\rangle\cap\mathcal{A}\subseteq F\cap\mathcal{A}=\mathbb{Q}, where the last equality holds by Lemma 4.1. Thus by Corollary 3.5, trdegK(E)rank(𝒜S)=dim(span(S{1}))1=k1{\rm trdeg}_{K}(E)\geq{\rm rank}(\mathcal{A}_{S}^{\prime})={\rm dim}_{\mathbb{Q}}({\rm span}_{\mathbb{Q}}(S\cup\{1\}))-1=k-1.

(ii).(ii). Suppose that PP is not diagonalizable over F¯=F\overline{F}=F. Then, as in the proof of Theorem 4.2, the field EE contains a solution of the set of equations δ(y)=λy\delta(y)=\lambda y for all λS\lambda\in S and for some aSa\in S, EE contains solution of the system δ(y)=ay\delta(y)=ay, δ(y1)=y+ay1\delta(y_{1})=y+ay_{1}. Thus by Corollary 3.10, trdegK(E)1+rank(𝒜S)=dim(span(S{1}))=k{\rm trdeg}_{K}(E)\geq 1+{\rm rank}(\mathcal{A}_{S}^{\prime})={\rm dim}_{\mathbb{Q}}({\rm span}_{\mathbb{Q}}(S\cup\{1\}))=k. ∎

From the above theorem, the only case where (Mn(K),𝒟P)(M_{n}(K),{\mathcal{D}}_{P}) admits a finite splitting field is when PP is diagonalizable over F¯\overline{F} and k=1k=1. However, if we extend the zero derivation of FF to KK, by defining δc,m(t)=ctm\delta_{c,m}(t)=ct^{m}, where m1m\neq 1 is an integer and 0cF0\neq c\in F, then such a possibility is not realized. In the following theorem we find sharp bounds on transcendence degrees of the extensions that split differential matrix algebra (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) over (F(t),δc,m)(F(t),\delta_{c,m}), where PMn(F)P\in M_{n}(F) is a traceless matrix.

Theorem 4.7.

Let K=F(t)K=F(t), 0cF0\neq c\in F and m1m\neq 1 be an integer. Let δc,m\delta_{c,m} be the derivation on KK defined by δc,m(a)=0\delta_{c,m}(a)=0, if aFa\in F, and δc,m(t)=ctm\delta_{c,m}(t)=ct^{m}. Let 0PMn(F)0\neq P\in M_{n}(F) be a traceless matrix and (E,δE)(E,\delta_{E}) be a differential field extension of (K,δc,m)(K,\delta_{c,m}) that splits (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}). Then EE is not algebraic over KK. In fact, trdegK(E)rank(S){\rm trdeg}_{K}(E)\geq{\rm rank}(\langle S\rangle), where SF¯S\subseteq\overline{F} is the set of eigenvalues of PP.

Proof.

Let.

𝒜={aF¯(t):δc,m(y)=nay has a nonzero solution in F¯(t) for some n}.\mathcal{A}=\{a\in\overline{F}(t):\delta_{c,m}(y)=nay\text{ has a nonzero solution in }\overline{F}(t)\text{ for some }n\in\mathbb{N}\}.

We claim that 𝒜S:=𝒜S=0\mathcal{A}_{S}:=\mathcal{A}\cap S=0. Since 𝒜=span(δc,m(F¯(t)))\mathcal{A}={\rm span}_{\mathbb{Q}}({\mathcal{L}}_{\delta_{c,m}}(\overline{F}(t))) and SF¯S\subseteq\overline{F}, it is enough to show that δc,m(a)a1F¯\delta_{c,m}(a)a^{-1}\notin\overline{F} for every aF¯(t)F¯a\in\overline{F}(t)\setminus\overline{F}. Let a=fg1F¯(t)F¯a=fg^{-1}\in\overline{F}(t)\setminus\overline{F}, where f,gF¯[t]f,g\in\overline{F}[t] are polynomials of degrees rr and ss, respectively, and gcd(f,g)=1{\rm gcd}(f,g)=1. Then

δc,m(a)=δc,m(f)gfδc,m(g)g2=(fgfg)ctmg2,\delta_{c,m}(a)=\frac{\delta_{c,m}(f)g-f\delta_{c,m}(g)}{g^{2}}=\frac{(f^{\prime}g-fg^{\prime})ct^{m}}{g^{2}},

where ff^{\prime} and gg^{\prime} denote the images of the derivation defined by t=1t^{\prime}=1. If δc,m(a)a1=kF¯\delta_{c,m}(a)a^{-1}=k\in\overline{F}, then the above equation may be rewritten as

(4.4) (fgfg)ctm=kfg.(f^{\prime}g-fg^{\prime})ct^{m}=kfg.

We compare the degrees on both sides of the equation 4.4 and obtain

r+s+m1r+s.r+s+m-1\geq r+s.

If m0m\leq 0, the above inequality cannot hold. Thus δc,m(a)a1F¯\delta_{c,m}(a)a^{-1}\notin\overline{F}.

Consider the case where m>1m>1. Let \ell be the highest power of tt that divides ff. Since gcd(f,g)=1\gcd(f,g)=1 and t1t^{\ell-1} divides fgfgf^{\prime}g-fg^{\prime}, comparing highest powers of tt dividing the two sides of the equation 4.4, 1+m\ell-1+m\leq\ell. This is a contradiction, since m>1m>1. Thus δc,m(a)a1F¯\delta_{c,m}(a)a^{-1}\notin\overline{F}.

We conclude that 𝒜S=0\mathcal{A}_{S}=0 in either case. Consequently, 𝒜S=S\mathcal{A}_{S}^{\prime}=\langle S\rangle. We now proceed to the proof of the theorem. Let (E,δE)(E,\delta_{E}) be a differential extension of (K,δc,m)(K,\delta_{c,m}) that splits (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}).

First, assume that PP is diagonalizable. By Corollary 2.3, EE contains solutions of differential equations δc,m(y)=ay\delta_{c,m}(y)=ay for each aSa\in S. Now by Corollary 3.5, trdegKF¯(EF¯)rank(𝒜S)=rank(S){\rm trdeg}_{K\overline{F}}(E\overline{F})\geq{\rm rank}(\mathcal{A}_{S}^{\prime})={\rm rank}(\langle S\rangle). As F¯(t)/K\overline{F}(t)/K is algebraic, trdegK(E)=trdegF¯(t)(EF¯)rank(S){\rm trdeg}_{K}(E)={\rm trdeg}_{\overline{F}(t)}(E\overline{F})\geq{\rm rank}(\langle S\rangle).

Now assume that PP is not diagonalizable. By Corollary 2.3, EE contains solutions of equations δc,m(y)=ay\delta_{c,m}(y)=ay and δc,m(y1)=y+ay1\delta_{c,m}(y_{1})=y+ay_{1} for some aSa\in S. Observe that w:=(c(m1))1t1mKw:=(c(m-1))^{-1}t^{1-m}\in K is such that δ(w)=1\delta(w)=1. We now use Lemma 3.7, Corollary 3.5 and argue as in the case when PP is diagonalizable to show that trdegK(E)rank(𝒜S)=rank(S){\rm trdeg}_{K}(E)\geq{\rm rank}(\mathcal{A}_{S}^{\prime})={\rm rank}(\langle S\rangle). ∎

We assert that the bounds on trdegK(E){\rm trdeg}_{K}(E) in Theorem 4.6 and Theorem 4.7 are sharp. Using Proposition 3.6, we may construct a field EE whose transcendence degree is equal to these bounds. Using the solutions of δ(y)=λy\delta(y)=\lambda y; λS\lambda\in S it is easy to construct a block diagonal matrix ZGLn(E)Z\in{\rm GL}_{n}(E) consisting of lower triangular invertible blocks such that δc(Z)=JPtZ\delta^{c}(Z)=J_{P}^{t}Z, where JPJ_{P} is a Jordan form of PP. This, in conjunction with Corollary 2.3 and Lemma 2.4, proves the assertion.

5. Splitting of tensor powers

Let K=F(t)K=F(t) and δ\delta the derivation on KK defined by δ(a)=0\delta(a)=0 for aFa\in F and δ(t)=t\delta(t)=t. In this section, we are concerned about the splitting of tensor powers of a matrix differential algebra (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}). We also relate it to the existence of finite splitting fields of (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}).

Theorem 5.1.

Let PMn(F)P\in M_{n}(F) be a traceless matrix. Let (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) be a differential matrix algebra over KK. If (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) does not admit a finite splitting field over (K,δ)(K,\delta), then the order of (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is infinite.

Proof.

Let mm be a positive integer such that (A,𝒟Q):=(Mn(K),𝒟P)m(A,\mathcal{D}_{Q}):=(M_{n}(K),\mathcal{D}_{P})^{\otimes m} is split over (K,δ)(K,\delta). By Theorem 4.2 and Corollary 4.4, QQ is diagonalizable, and all its eigenvalues are integers. Hence by Lemma 2.7, PP is diagonalizable, and all the eigenvalues of PP are rational. By Theorem 4.2, (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is split by a finite extension over (K,δ)(K,\delta), which is a contradiction to the hypothesis. ∎

To construct an example of a differential matrix algebra of finite order, let us consider the product (M2(K),𝒟P1)(M2(K),𝒟P2)(M_{2}(K),\mathcal{D}_{P_{1}})\otimes(M_{2}(K),\mathcal{D}_{P_{2}}), where P1,P2M2(F)P_{1},P_{2}\in M_{2}(F) are traceless diagonalizable matrices with rational eigenvalues ±λ1\pm\lambda_{1} and ±λ2\pm\lambda_{2}, respectively. Consider the differential matrix algebra (M2(K),𝒟P1)(M2(K),𝒟P2)(M_{2}(K),\mathcal{D}_{P_{1}})\otimes(M_{2}(K),\mathcal{D}_{P_{2}}). Derivation on this algebra is 𝒟P11+1𝒟P2=𝒟P11+1P2\mathcal{D}_{P_{1}}\otimes 1+1\otimes\mathcal{D}_{P_{2}}=\mathcal{D}_{P_{1}\otimes 1+1\otimes P_{2}}. From Lemma 2.6, eigenvalues of P11+1P2P_{1}\otimes 1+1\otimes P_{2} are ±λ1±λ2\pm\lambda_{1}\pm\lambda_{2}. Thus by Theorem 4.2 and Corollary 4.3, there exists a finite extension E/KE/K of degree dd that splits (M2(K),𝒟P1)(M2(K),𝒟P2)(M_{2}(K),\mathcal{D}_{P_{1}})\otimes(M_{2}(K),\mathcal{D}_{P_{2}}), where dd is given by

d={lcm(q1,q2)/2 if lcm(q1,q2) is even and p1,p2 are both oddlcm(q1,q2) otherwise,d=\begin{cases}{\rm lcm}(q_{1},q_{2})/2&\text{ if ${\rm lcm}(q_{1},q_{2})$ is even and $p_{1},p_{2}$ are both odd}\\ {\rm lcm}(q_{1},q_{2})&\text{ otherwise},\end{cases}

where λi=pi/qi\lambda_{i}=p_{i}/q_{i} and pi,qip_{i},q_{i} are relatively prime integers. Thus, if each λi=12\lambda_{i}=\frac{1}{2} then (M2(K),𝒟P1)(M2(K),𝒟P2)(M_{2}(K),\mathcal{D}_{P_{1}})\otimes(M_{2}(K),\mathcal{D}_{P_{2}}) is split over (K,δ)(K,\delta) itself.

In the next theorem, we use these ideas to deal with tensor powers of a differential matrix algebra (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}). We find conditions on the eigenvalues of PP so that the mm-fold tensor product of (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is trivial.

Theorem 5.2.

Let PMn(F)P\in M_{n}(F) be a traceless matrix such that the differential matrix algebra (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is split by a finite extension. Then the following are equivalent:

  1. (i).

    (A,𝒟Q):=(Mn(K),𝒟P)m(A,\mathcal{D}_{Q}):=(M_{n}(K),\mathcal{D}_{P})^{\otimes m} is split over (K,δ)(K,\delta).

  2. (ii).

    If λ\lambda and μ\mu are any two eigenvalues of PP, then m{λ}=m{μ}m\{\lambda\}=m\{\mu\}, where {λ}\{\lambda\} and {μ}\{\mu\} denote fractional parts of λ\lambda and μ\mu, respectively.

We note that, since (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is split by a finite extension, eigenvalues of PP are rational and it makes sense to talk about fractional parts.

Proof.

Let S:={λ1,λ2,,λn}S:=\{\lambda_{1},\lambda_{2},\dots,\lambda_{n}\} be the set of eigenvalues of PP. Since (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is split by a finite extension, by Theorem 4.2, the matrix PP is diagonalizable and SS\subseteq\mathbb{Q}. Since 𝒟Q=𝒟Pm\mathcal{D}_{Q}=\mathcal{D}_{P}^{\otimes m}, by Lemma 2.6, the set of eigenvalues of QQ is

Λ:={λi1+λi2++λim:(i1,i2,,im){1,2,,n}m}.\Lambda:=\{\lambda_{i_{1}}+\lambda_{i_{2}}+\dots+\lambda_{i_{m}}\,:\,(i_{1},i_{2},\cdots,i_{m})\in\{1,2,\dots,n\}^{m}\}.

(i)\Rightarrow(ii): Since (A,𝒟Q)(A,\mathcal{D}_{Q}) is split over (K,δ)(K,\delta), by Corollary 4.4 each μΛ\mu\in\Lambda is an integer. Thus, from the particular case i1=i2==imi_{1}=i_{2}=\dots=i_{m}, each mλim\lambda_{i}\in\mathbb{Z}. Let αi=mλi\alpha_{i}=m\lambda_{i}. Now, from the case i1==im1imi_{1}=\dots=i_{m-1}\neq i_{m}, we get that aij:=(m1)λi+λja_{ij}:=(m-1)\lambda_{i}+\lambda_{j}\in\mathbb{Z}, whenever iji\neq j. Thus λjλi=aijαi\lambda_{j}-\lambda_{i}=a_{ij}-\alpha_{i}\in\mathbb{{Z}}, and therefore, {λi}\{\lambda_{i}\} is the same for each ii.

(ii)\Rightarrow(i): Let m{λi}=am\{\lambda_{i}\}=a\in\mathbb{{Z}} for each ii. Then each eigenvalue λi1+λi2++λim\lambda_{i_{1}}+\lambda_{i_{2}}+\dots+\lambda_{i_{m}} of QQ is an integer and by Corollary 4.4, (A,𝒟Q)(A,\mathcal{D}_{Q}) is split over (K,δ)(K,\delta). ∎

From Theorems 5.1 and Theorem 5.2, we conclude that if (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) has finite order, then this order is precisely the smallest integer mm such that m{λi}m\{\lambda_{i}\}\in\mathbb{N}, where {λi}\{\lambda_{i}\} is the constant fractional part of eigenvalues of PP. In the following corollary we show that the order of (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) is necessarily a divisor of nn.

Corollary 5.3.

Let PMn(F)P\in M_{n}(F) be a traceless matrix. If the differential matrix algebra (Mn(K),𝒟P)(M_{n}(K),\mathcal{D}_{P}) has finite order mm, then mm divides nn.

Proof.

Let S:={λ1,λ2,,λn}S:=\{\lambda_{1},\lambda_{2},\dots,\lambda_{n}\}, be the set of eigenvalues of PP. Let aij,αia_{ij},\alpha_{i}\in\mathbb{Z} be as in the proof of Theorem 5.2. Then (m1)αi+αj=maij(m-1)\alpha_{i}+\alpha_{j}=ma_{ij}. Since PP is a traceless matrix, 0=m(j=1nλj)=j=1nαj0=m\left(\sum_{j=1}^{n}\lambda_{j}\right)=\sum_{j=1}^{n}\alpha_{j}. Substituting αj=m(a1jα1)+α1\alpha_{j}=m(a_{1j}-\alpha_{1})+\alpha_{1} whenever j1j\neq 1,

j=1nαj=α1+j=2n(α1+m(a1jα1))=nα1+mc=0,\sum_{j=1}^{n}\alpha_{j}=\alpha_{1}+\sum_{j=2}^{n}(\alpha_{1}+m(a_{1j}-\alpha_{1}))=n\alpha_{1}+mc=0,

where c=j=2n(a1jα1)c=\sum_{j=2}^{n}(a_{1j}-\alpha_{1})\in\mathbb{Z}. Now α1=mλ1=m([λ1]+{λ1})=m[λ1]+a\alpha_{1}=m\lambda_{1}=m([\lambda_{1}]+\{\lambda_{1}\})=m[\lambda_{1}]+a, where aa is as in the proof of Theorem 5.2 and [λ1][\lambda_{1}] is the integer part of λ1\lambda_{1}. It is now clear that gcd(α1,m)=gcd(a,m)=1\gcd(\alpha_{1},m)=\gcd(a,m)=1 and nα1+mc=0n\alpha_{1}+mc=0. Thus mm divides nn. ∎

We conclude this article by asking the following intriguing question. ”Let (A,𝒟)(A,\mathcal{D}) be a differential central simple algebra over a differential field (K,δ)(K,\delta). If (A,𝒟)m(A,\mathcal{D})^{\otimes m} is split for some mm\in\mathbb{N}, then does mm necessarily divide deg(A){\rm deg}(A)?”

References

  • [Ami82] S. A. Amitsur, Extension of derivations to central simple algebras, Comm. Algebra 10 (1982), no. 8, 797–803. MR 651973
  • [HJ94] Roger A. Horn and Charles R. Johnson, Topics in matrix analysis, Cambridge University Press, Cambridge, 1994, Corrected reprint of the 1991 original. MR 1288752
  • [Jac89] Nathan Jacobson, Basic algebra. II, second ed., W. H. Freeman and Company, New York, 1989. MR 1009787
  • [JM08] Lourdes Juan and Andy R. Magid, Differential central simple algebras and Picard-Vessiot representations, Proc. Amer. Math. Soc. 136 (2008), no. 6, 1911–1918. MR 2383496
  • [Mag94] Andy R. Magid, Lectures on differential Galois theory, University Lecture Series, vol. 7, American Mathematical Society, Providence, RI, 1994. MR 1301076