429
Views
0
CrossRef citations to date
0
Altmetric
Pure Mathematics

A characterization of real matrix semigroups

& | (Reviewing editor:)
Article: 2289203 | Received 11 Jul 2023, Accepted 24 Nov 2023, Published online: 12 Dec 2023

ABSTRACT

We characterize all real matrix semigroups, indexed by the non-negative reals, which satisfy a mild boundedness assumption, without assuming continuity. Besides the continuous solutions of the semigroup functional equation, we give a description of solutions arising from non-measurable solutions of Cauchy’s functional equation. To do so, we discuss the primary decomposition and the Jordan—Chevalley decomposition of a matrix semigroup. Our motivation stems from a characterization of all multi-dimensional self-similar Gaussian Markov processes, which is given in a companion paper.

MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

It is a classical fact that all continuous matrix-valued functions g:[0,)Rn×n satisfying the semigroup property

(1.1) g(x+y)=g(x)g(y),x,y0,(1.1)
(1.2) g(0)=id,(1.2)

where id is the identity matrix, are given by the maps g(x)=exp(Mx), where MRn×n is arbitrary. See, e.g. Problem 2.1 in  (Engel & Nagel, Citation2000) and the subsequent discussion. To the best of our knowledge, few authors have considered non-continuous solutions of (Equation1.1) in the multidimensional case d > 1. Kuczma and Zajtz  (Kuczma & Zajtz, Citation1966) determine all matrix semigroups (g(x))xR where xg(x) is measurable. Zajtz  (Zajtz, Citation1971) characterizes the matrix semigroups (g(x))xQ indexed by rational numbers. The one-dimensional case, which is equivalent to Cauchy’s functional equation,

(1.3) f(x)+f(y)=f(x+y),f:RR,(1.3)

is well studied, on the other hand (see  (Aczél, Citation1966; Bingham et al., Citation1987)). Our motivation to investigate non-continuous matrix semigroups stems from probability theory: In a companion paper, we develop a characterization of all self-similar Gaussian Markov processes. By self-similarity, the bivariate covariance function of such a multi-dimensional stochastic process can be transformed to a matrix function of a single argument, which must satisfy (Equation1.1). See Section 2 for some more details. Our main result (Theorem 2.6) determines all solutions of (Equation1.1) which satisfy a mild boundedness assumption. In Section 3, we prepare the proof by providing a decomposition of the vector space into invariant subspaces, and establish some useful properties of the decomposition. The main proofs (in Section 5) are preceded by a discussion of the Jordan—Chevalley decomposition of a semigroup, in Section 4. In Appendix A, we give two auxiliary results on Cauchy’s functional equation.

2. Preliminaries and main result

To motivate our investigation, we first recall some facts about Gaussian stochastic processes (see, e.g. Lifshits, Citation2012; Nourdin, Citation2012, for more information). Let X=(Xt)t0 be a d-dimensional real centered Gaussian process. The covariance of X is a matrix-valued function (s,t)R(s,t)Rd×d satisfying

vTR(s,t)u=E[XsTvXtTu],s,t0, u,vRd,

and uniquely characterizes the law of the process. Suppose that X is self-similar, and that R(1,1)=id is the identity matrix. In terms of the covariance function, self-similarity means that R(as,at)=a2HR(s,t) for a,s,t>0 and some self-similarity parameter H > 0. By a classical criterion  (Doob, Citation1953) [Theorem V.8.1], X is a Markov process if and only if its covariance function satisfies

(2.1) R(s,t)R(t,t)1R(t,u)=R(s,u),0stu.(2.1)

Upon introducing g(x):=R(ex,1), self-similarity allows to reduce (Equation2.1) to

g(x)g(y)=g(x+y),x,y,0.

This observation has been used in dimension d = 1 to prove that certain Gaussian processes do not have the Markov property (see, e.g. (Nourdin, Citation2012), [Theorem 2.3]). Unifying and generalizing these isolated results in a companion paper, we obtain a classification of all d-dimensional self-similar Gaussian Markov processes. A full classification requires finding all solutions of (Equation1.1), without assuming continuity. The boundedness assumption 2.5 stated below causes no problems, though. To state our main result, define the rotation matrix

Q(θ)=(cos(θ)sin(θ)sin(θ)cos(θ)),θR,

and, for even k and a function ν:RR, the block-diagonal matrix

(2.2) Qkν(x):=(Q(ν(x))00Q(ν(x)))Rk×k(2.2)

consisting of k/2 rotation matrices. In our statements, ν will denote some non-measurable (equivalently, non-continuous) solution of (Equation1.3). The following assumption is in force throughout the paper.

Assumption 2.1.

In the following, V denotes a real d-dimensional vector space equipped with an inner product ,.

Definition 2.2.

We write L(V) for the set of linear maps from V to itself. The operator norm induced by the inner product , is denoted by op. If the basis is clear from the context, we will identify elements of L(V) and d × d matrices.

Definition 2.3.

We say that g:R0L(V) is a semigroup if g(0)=id and

g(x+y)=g(x)g(y)

for all x,y0. Semigroups acting on VC, the complexification of V, are defined by the same property.

Definition 2.4.

We say that a semigroup (g(x))x0 in L(V) is elementary if there exists an orthonormal basis such that

g(x)=exp(Mx), x0,org(x)=Qdν(x)exp(Mx), x0,

for some non-continuous ν satisfying Cauchy’s EquationEquation (1.3) and some matrix MRd×d.

Assumption 2.5.

The semigroup g(x)x0 satisfies g(x)opf(x) for x0, where the function f:[0,)R is locally bounded, right-continuous at 0 and satisfies f(0)=1.

We can now state our main theorem, which will be proven at the end of Section 5.

Theorem 2.6.

Let (g(x))x0 be a semigroup satisfying Assumption 2.5. Then, there exists an orthogonal decomposition V=i=1nVi such that each Vi is invariant under g(x), and either g(x) is elementary on Vi or g(x)|Vi=0 for x > 0.

We call g((x))x0 degenerate if g(x)=0 for x > 0.

Example 2.7.

Let cR, and let f:RR be a non-continuous solution of Cauchy’s functional EquationEquation (1.3). Then, the semigroup given by

g(x):=(cos(f(x))sin(f(x))sin(f(x))cos(f(x)))ecx,x0,

is an example of a semigroup covered by Theorem 2.6 (by putting d = 2 and M=cid in Definition 2.4), but not the previous results mentioned at the beginning of the introduction. It illustrates that, in contrast to dimension one, a two-dimensional locally bounded semigroup need not be continuous.

3. Primary decomposition for semigroups

In this section, we discuss a decomposition of V into subspaces which are invariant for the given semigroup.

Definition 3.1.

Let S be a linear operator on the vector space V. A subspace UV is called S-invariant if S maps U into U.

The primary decomposition theorem from linear algebra [O’Meara et al., Citation2011, Theorem 1.5.1] decomposes a vector space into invariant subspaces for a given operator. On each subspace, the operator has a single real eigenvalue or a pair of conjugate complex eigenvalues. Instead of a single operator, we need the following version for semigroups.

Theorem 3.2.

Primary decomposition For any semigroup (g(x))x0 of linear maps acting on V, there exists a decomposition V=i=1nVi with dim(Vi)1 such that each Vi is g(x)-invariant for all x0, and for all i one of the following holds:

(1)

For all x0, g(x) has one eigenvalue λ(x)0 on Vi,

(2)

For all x0, g(x) has eigenvalues λ(x),λ(x)C on Vi. They may coincide (and thus be real) for some values of x, but not for all x0.

Proof.

It is known that the primary decomposition extends to commuting sets of matrices. Indeed, the decomposition V=i=1nVi into invariant subspaces follows from Theorem 5 on p. 40 in  (Jacobson, Citation1962), applied to the span of the semigroup (g(x))x0. Thus, we only need to argue why (1) or (2) follows from the semigroup property. Assume that each g(x) has only one eigenvalue λ(x)R on some Vi. Since the g(x) commute, they share a common eigenvector vi, and thus we have λ(x)λ(y)=λ(x+y) for x,y0. Suppose λ(x0)<0 for some x0>0. Then, we have λ(x0)=λ(x0/2)2, and hence λ(x0/2)R, contradicting λ(x)R for all x0. Hence λ(x)R0 for all x0.

Theorem 3.3.

Consider a semigroup g=(g(x))0 acting on VC, the complexification of V. Then, there exists a decomposition VC=i=1nVi with dim(Vi)1 such that, for each i and all x0, the space Vi is g(x)-invariant, and g(x) has only one eigenvalue λ(x)C on Vi.

Proof.

This is an immediate consequence of Theorem 5 on p. 40 in  (Jacobson, Citation1962) (cf. the preceding proof).

Definition 3.4.

We call the decomposition from Theorem 3.2 simultaneous real primary decomposition (SRPD) of V, omitting “w.r.t. g” if the semigroup is clear from the context. The component Vi is of first type in case (1), and of second type in case (2). Similarly, the simultaneous primary decomposition (SPD) is the decomposition from Theorem 3.3.

Lemma 3.5.

Let g be a semigroup acting on V, and let ViV be a subspace of first type from the SRPD of V. Then, there exists a common eigenvector vV such that g(x)v=λ(x)v for all x > 0.

Proof.

We present an algorithm which yields the subspace of common eigenvectors. If Vi=:Vi(0) itself is this subspace, then we are done. Otherwise there exists x0>0 such that the eigenspace Vi(1) of g(x0) is a strict subspace of Vi. For any y > 0 and any vVi(1) we have

0=g(y)0=g(y)[λ(x0)vg(x0)v]=λ(x0)[g(y)v]g(x0)[g(y)v].

It follows that Vi(1) is g(x)-invariant for all x > 0. Now either Vi(1) consists of common eigenvectors, or we can again find x1>0 such that the eigenspace Vi(2) of g(x1) in Vi(1) is a strict subspace of Vi(1). Repeating this argument yields a sequence of nontrivial subspaces Vi(n) whose dimensions are strictly decreasing, hence it has to terminate. Clearly, the final vector space in this sequence is the space of common eigenvectors of the semigroup.

Lemma 3.6.

Let g be a semigroup acting on V, and let Vi be a subspace from the SRPD of V such that there exists x0>0 with λ(x0)=0. Then, we have λ(x)=0 for all x > 0. Furthermore Viker(g(x)) for all x > 0.

Proof.

By the previous lemma, there exists a common eigenvector vVi. (Since commuting matrices are simultaneously triangularizable it follows that they share a common eigenvector vVC.) We have

λ(x+y)v=g(x+y)v=g(x)g(y)v=λ(x)λ(y)v.

Hence λ satisfies λ(x)λ(y)=λ(x+y). Since λ(x0)=0 we have λ(x0/n)=0 for all nN since λ(x0/n)n=0. Let y > 0, then there exists nN such that x0/n<y. We obtain

λ(y)=λ(x0/n)λ(yx0/n)=0.

Since λ(y)=0 for all y > 0 it follows that the characteristic polynomial of g(y) satisfies χg(y)(Z)=Zdim(Vi) and hence, by the Cayley—Hamilton theorem, g(y)dim(Vi)0. For vVi we obtain

g(y)v=g(y/dim(Vi))dim(Vi)v=0.

Corollary 3.7.

Let g be a semigroup acting on V, with SRPD V=i=1nVi. For any x > 0 we have ker(g(x))=iIVi, where

I={i|Viof type 1 with λ0}.

Proof.

If Viker(g(x)), then 0=λ(x)R for all x0 and hence Vi is of type 1. By the previous lemma, we have

ker(g(x))iIVi.

If the inclusion was strict, then there would exist Vj with jI such that ker(g(x))Vj. But since g(x) is invertible on Vj this gives a contradiction, hence we have equality.

Corollary 3.8.

Let g be a semigroup acting on V. Then, there exists a decomposition V=V1V2, such that g(x)|V1 is invertible for all x0 and g(x)|V20 for x > 0.

4. Multiplicative Jordan–Chevalley decomposition of semigroups

Due to Corollary 3.8, from now on we assume in most of our statements that g(x) is invertible for all x0. A standard result from linear algebra, the Jordan—Chevalley decomposition, asserts that any matrix A can be uniquely decomposed as A=D+N, where D is diagonalizable, N is nilpotent and D and N commute. If A is invertible, then we can express it as A=D(id+D1N):=DT with T unipotent and commuting with D.

Definition 4.1.

For an invertible linear map A on Kd with K{R,C}, the multiplicative decomposition A = DT into commuting factors with D diagonalizable and T unipotent, is called the multiplicative Jordan—Chevalley decomposition.

For background on the (multiplicative) Jordan—Chevalley decomposition, we refer to Section 15.1 in  (Humphreys, Citation1975). We now analyze the structure of the multiplicative Jordan—Chevalley decomposition of a semigroup.

Theorem 4.2.

Let (g(x))x0 be a semigroup of invertible linear maps acting on Kd with K{R,C} and let g(x)=D(x)T(x) be the multiplicative Jordan—Chevalley decomposition of each g(x). Then ((D(x))x0) and (T(x))x0 each form a semigroup, and the two families commute with each other, i.e. T(x)D(y)=D(y)T(x) for all x,y0.

Proof.

We can w.l.o.g. assume that K=C, since by uniqueness of the Jordan—Chevalley decomposition D(x)Rd×d and T(x)Rd×d if g(x)Rd×d for x0. Take the SPD from Theorem 3.3, V=i=1kVi, so that each g(x) has only one eigenvalue on Vi. Denote by g(x)|Vi=Di(x)Ti(x) the multiplicative Jordan—Chevalley decomposition g(x) restricted to Vi. Denote by λi(x) the eigenvalue of g(x) on Vi. Clearly, Di(x)=λi(x)id. Since the (gi(x))x0 are a commuting family of matrices, they share a common eigenvector viVi. We have

λi(x)λi(y)vi=gi(x)gi(y)vi=gi(x+y)vi=λi(x+y)vi,

and hence (Di(x))x0 is a semigroup. Since each Di(x) is a multiple of the identity, it commutes with every linear map and hence

Ti(x)Ti(y)=1λi(x)λi(y)gi(x)gi(y)=1λi(x+y)gi(x+y)=Ti(x+y),

which shows that (Ti(x))x0 is also a semigroup. The result then follows, since by uniqueness (Di(x))(Ti(x)) is the multiplicative Jordan—Chevalley decomposition of g(x).

Theorem 4.3.

Let (g(x))x0 be a semigroup of invertible linear maps acting on V and let g(x)=D(x)T(x) be its multiplicative Jordan—Chevalley decomposition. Then, there exist commuting real diagonalizable linear maps J(x) and commuting real nilpotent linear maps N(x) satisfying

(1)

J(x)+J(y)=J(x+y)

(2)

N(x)+N(y)=N(x+y)

(3)

J(x)N(y)=N(y)J(x) for all x,y0

such that

D(x)=exp(J(x)),T(x)=exp(N(x))

and

g(x)=exp(J(x)+N(x)).

Proof.

Let V=i=1nVi be the SRPD and let gi(x):=g(x)|Vi. Assume first that Vi is of first type and λi(x) is the single positive eigenvalue of gi(x). Then, the multiplicative Jordan—Chevalley decomposition on Vi is gi(x)=λi(x)idTi(x). Define Ji(x):=log(λi(x))id. Since Ti(x)id is nilpotent, we have

Ni(x):=log(Ti(x))=k=1d11k(idTi(x))k.

Notice that since the logarithm converges for all unipotent matrices, the exponential map between the Lie algebra of nilpotent matrices and the Lie group of unipotent matrices is bijective (see p. 35 in  (Goodman & Wallach, Citation1998)). Since Ti(x) is a semigroup we have

exp(Ni(x))exp(Ni(y))=exp(Ni(y))exp(Ni(x))=exp(Ni(x+y)).

Rewrite this as

exp(Ni(y))1exp(Ni(x))exp(Ni(y))=exp(exp(Ni(y))1Ni(x)exp(Ni(y)))=exp(Ni(x)).

Since exp(Ni(y))1Ni(x)exp(Ni(y)) is nilpotent and exp is bijective, we have

exp(Ni(y))1Ni(x)exp(Ni(y))=Ni(x).

For any tR set Nit(x):=tid+Ni(x), which is invertible for t0. It is clear that

Nit(x)exp(Ni(y))=exp(Ni(y))Nit(x).

Applying the same idea again yields

exp(Nit(x)1Ni(y)Nit(x))=exp(Ni(y)).

Again by uniqueness we obtain

Nit(x)Ni(y)=Ni(y)Nit(x)

or equivalently

Ni(x)Ni(y)=Ni(y)Ni(x).

By commutativity we obtain

exp(Ni(x)+Ni(y))=exp(Ni(x))exp(Ni(y))=exp(Ni(x+y))

and hence, again by uniqueness, we have Ni(x)+Ni(y)=Ni(x+y). It follows that Ji(x) and Ni(x) have the desired properties.

In the second case Vi is of second type. By Theorem 3.3, and since gi(x) is real, Vi decomposes over C as Vi=U1U2, where gi(x) has only one eigenvalue on each Uj and U1 is isomorphic to U2 with isomorphism given by uU1uU2. By taking the principal branch of the logarithm, in the same manner as in the real case, we obtain commuting J,N on U1 and J,N on U2 satisfying Cauchy’s equation such that

gi(x)=exp(J(x)+N(x))exp(J(x)+N(x)).

Notice that J(x)=(μ(x)+iν(x))id, where i=1 and ν(x) satisfies Cauchy’s functional equation on R/[π,π). By Lemma A.1, we can lift any solution on R/[π,π) to a solution on R such that linearity is preserved. From now on denote this lift by ν(x). Hence, if we choose any basis of U1 and its complex conjugate on U2 we obtain that gi(x) is similar to

gi(x)exp[(J(x)+N(x)00J(x)+N(x))].

Taking the similarity transform with the matrix

A:=12(idiidiidid),

where id=iddim(U1), we obtain

A(J(x)+N(x)00J(x)+N(x))A1=(Re(J(x)+N(x))Im(J(x)+N(x))Im(J(x)+N(x))Re(J(x)+N(x)))L(Vi).

Since matrix similarity over C is equivalent to matrix similarity over R for two real matrices, there exists a real matrix Bi on Vi such that

(4.1) gi(x)=exp[Bi(Re(J(x))Im(J(x))Im(J(x))Re(J(x)))Bi1+Bi(Re(N(x))Im(N(x))Im(N(x))Re(N(x)))Bi1].(4.1)

Setting

Ji(x):=Bi(Re(J(x))Im(J(x))Im(J(x))Re(J(x)))Bi1=Bi(idμ(x)idν(x)idν(x)idμ(x))Bi1

and

Ni(x):=Bi(Re(N(x))Im(N(x))Im(N(x))Re(N(x)))Bi1,

we see that Ji and Ni satisfy the desired conditions, with Di(x)=exp(Ji(x)) and Ti(x)=exp(Ni(x)) by uniqueness of the Jordan—Chevalley decomposition. The direct sums i=1nJi(x) and i=1nNi(x) give the required matrices. Furthermore, in the case where, there exists x > 0 for which gi(x) has a complex eigenvalue λ(x)=μ(x)+iν(x) we have

gi(x)=Bi(cos(ν(x))idsin(ν(x))idsin(ν(x))idcos(ν(x))id)Ui(x)Bi1exp(μ(x)id+BiN(x)Bi1),

where Ui(x)SO(dim(Vi)). Recall the matrix defined in (Equation2.2). By changing the order of the basis, we have that Ui is similar to the block diagonal matrix (recall the notation (Equation2.2))

(4.2) Ui(x)(Q(ν(x))Q(ν(x)))=Qdim(Vi)ν(x).(4.2)

Hence

(4.3) gi(x)=B~iQdim(Vi)ν(x)B~i1exp(μ(x)id+BiN(x)Bi1),(4.3)

where B~i is the composition of Bi with some permutation matrix P.

5. Proof of the main result

After providing some final preparatory results, this section ends with the proof of Theorem 2.6. Consider again the SRPD from Theorem 3.2. Now on each Vi, g(x) has either one positive eigenvalue λi(x)=eμi(x) or two complex conjugate eigenvalues λi(x)=eμi(x)+iνi(x) and λi(x)=eμi(x)iνi(x), where it is possible that νi(x)=π for some values of x. If Vi is of first type, set νi(x)=0. Each νi is a solution to Cauchy’s functional EquationEquation (1.3). Consider then the set {f|f=νior f=νifor some 1in} and partition it into equivalent solutions, according to Definition A.2. This then gives a partition of the index set {1,,n}=l=1kIl in the following manner: If νiνj or νiνj, then i,j are in the same subset of the partition. This is well-defined, since if fg then fg. Set Wl:=iIlVi.

Definition 5.1.

We call the decomposition V=i=1kWi the partitioned SRPD of V.

Furthermore associate with each Wj one solution ηj of Cauchy’s equation such that ηjνi with iIj. If ηj is linear we always take ηj0. Notice that for i ≠ j we have ηiηj, hence there can be at most one Wi with ηi=0, and furthermore this is the only Wi which can have odd dimension since it contains all Vj of type 1 (recall Definition 3.4).

Theorem 5.2.

Let (g(x))x0 be a semigroup of invertible linear maps acting on V, and denote by V=i=1kWi its partitioned SRPD. Denote by ηi the solution associated with Wi. If ηi is non-continuous, then there exists a change of basis Ai on Wi such that

g(x)|Wi=AiQdiηi(x)Ai1exp(Gi(x)),

where di=dim(Wi), and

AiQdiηi(x)Ai1Gi(y)=Gi(y)AiQdiνi(x)Ai1

for all x,y0. If ηi=0, then

g(x)|Wi=exp(Gi(x)).

In both cases Gi(x):WiWi is a commuting family of matrices on Wi satisfying Cauchy’s functional equation. Furthermore, if v is a common eigenvector of (Gi(x))x0 such that Gi(x)v=[μ(x)+iν(x)]v, then ν is linear in x.

Proof.

Assume first that ηi, the associated solution of Cauchy’s functional equation, is non-continuous. For each Wi we have the decomposition Wi=jIiVj where Vj are subspaces from the SRPD. Since ηi is non-continuous, each Vj in the direct sum has to be of second type, hence by (Equation4.3) on each Vj we have

gj(x)=B~jQdim(Vj)νj(x)B~j1exp(μ(x)id+BjN(x)Bj1).

By definition of ηi we have ηiνj and hence there exists cjR such that νj(x)=ηi(x)+cjx. Let

Cj:=cj(LL)

with

L:=(0110),

where Cj is of dimension dim(Vj)×dim(Vj). Then, we have

gj(x)=B~jQdim(Vj)νj(x)B~j1exp(μj(x)id+BjNj(x)Bj1)=B~jQdim(Vj)ηi(x)exp(Cjx)B~j1exp(μj(x)id+BjNj(x)Bj1)=B~jQdim(Vj)ηi(x)B~j1exp(B~jCjxB~j1+μj(x)id+BjNj(x)Bj1).

Set G~j(x):=B~jCjxB~j1+μj(x)id+BjNj(x)Bj1. Then

(5.1) Ai:=jIiB~jandGi(x):=jIiG~j(x)(5.1)

satisfy

g(x)|Wi=jIigj(x)=AiQdiηi(x)Ai1exp(Gi(x)).

By construction the common eigenvalues of exp(Gi(x)) satisfy λ(x)=eμj(x)±icjx. If ηi0, then by Theorem 4.3 there exists Gi(x) such that g(x)|Wi=exp(Gi(x)). By the definition of Wi, it follows that the imaginary parts of the eigenvalues of Gi(x) have to be equivalent to ηi0, hence they have to be linear.

Theorem 5.3.

Let g be a non-degenerate semigroup acting on V which satisfies Assumption 2.5. Then, there exists a matrix M and a semigroup S(x)SO(d), where SO(d) is the set of special orthogonal matrices, such that

g(x)=S(x)exp(Mx),x0,

with M commuting with S(x).

Proof.

Consider the SRPD V=i=1nVi from Theorem 3.2, and define gi(x):=g(x)|Vi. Clearly, gi(x)opg(x)opf(x). For each eigenvalue λi(x)=eμi(x)+iνi(x) on Vi, we have

|λi(x)|gi(x)opf(x).

Since λi(x)λi(y)=λi(x+y), we have |λi(x)||λi(y)|=|λi(x+y)|. Moreover, μi(x)=log(|λi(x)|), and so μi(x)log(f(x)) for x0 and μi(x)+μi(y)=μi(x+y). It follows that µi is locally bounded and hence that μi(x)=aix for some ai0. As in the proof of Theorem 4.3, on Vi we have gi(x)=exp(Ji(x))Ti(x) where Ji(x) is diagonalizable with eigenvalues aix±iνi(x), and thus

Ti(x)opf(x)eaix.

As in Theorem 4.3, set

Ni(x):=log(Ti(x))=k=1d11k(idTi(x))k.

Then, since idTi(x)op1+f(x)eaix, we obtain

(5.2) Ni(x)opk=1d11kidTi(x)opkk=1d11k(1+f(x)eaix)k=:F(x),(5.2)

where F(x) is again locally bounded. Since the operator norm in finite dimensions is equivalent to the L-norm, it follows that each entry of Ni(x) is locally bounded and satisfies Cauchy’s functional equation. Thus, there exists a nilpotent linear map Pi such that Ni(x)=Pix. Let V=i=1kWi be the partitioned SRPD of V. Assume first that the solution associated with Wi is ηi0. Since Wi=jIiVj we have g(x)|Wi=jIiexp(Jj(x)+Pjx). The real part of the eigenvalues of Jj(x) is linear in x and by Theorem 5.2 also the complex parts have to be linear since νi0. Since Jj(x) is diagonalizable and all its eigenvalues are continuous in x, Jj(x) is continuous in x and hence Jj(x)=M~jx. Setting Mi:=jIi(M~j+Pj) and Si(x)=id yields g(x)|Wi=Si(x)exp(Mix). By Theorem 5.2 and (Equation5.1), we have for Wi with ηi non-continuous

g(x)|Wi=AiQdiηi(x)Ai1exp(Gi(x))

with

Gi(x)=jIi(B~jCjxB~j1+μj(x)id+BjNj(x)Bj1)=jIi(B~jCjB~j1+ajid+BjPjBj1)x.

Hence, setting Mi:=jIi(B~jCjB~j1+ajid+BjPjBj1), we have Gi(x)=Mix. Thus

g(x)|Wi=AiQdiηi(x)Ai1exp(Mix).

Next we show that Si(x):=AiQdiηi(x)Ai1 is an isometry on Wi. Since g(x)opf(x), we have

(5.3) v,v1f(x)2g(x)v,g(x)v0(5.3)

for any vV. Hence, for vWi we have

v,v1f(x)2exp(Mix)Si(x)v,exp(Mix)Si(x)v0.

Fix an arbitrary x00, and set θ0:=ηi(x0). The graph of ηi is dense in R×R (see Appendix A), and so there is a sequence of positive reals xn with limnxn=0 such that limηi(xn)=θ0. Then, for vWi we obtain

0lim(v,v1f(xn)2exp(Mixn)Si(xn)v,exp(Mixn)Si(xn)v)=v,vSi(x0)v,Si(x0)v,

where lim1f(xn)=1 since f is right-continuous at 0. Choose (x~n) such that limx~n=0 and limηi(x~n)=θ0. Then, since S(x0)vWi, we obtain

0lim(S(x0)v,S(x0)v1f(x~n)2exp(Mix~n)Si(x~n)Si(x0)v,exp(Mix~n)Si(x~n)Si(x0)v)=Si(x0)v,Si(x0)vv,v,

since Qdi(θ0)Qdi(θ0)=id. Hence, we obtain

v,v=S(x0)v,S(x0)v,

which implies that S(x0) is an isometry on Wi. Next we show that all Wi are pairwise orthogonal. Let vWi and uWj for i ≠ j. By Lemma A.3, w.l.o.g. there is a sequence x^n such that limx^n=0, limηi(x^n)=π and limηj(x^n)=θ(π,π). Hence limS(x^n)|Wi=id and limS(x^n)|Wj=AjQdj(θ)Aj1 with θ(π,π). We have the identity Qdj(ϕ)+Qdj(ϕ)=2cos(ϕ)id. Applying this to ϕ:=θ/2 we obtain

id+Qdj(θ)=2cos(θ/2)Qdj(θ/2).

Hence we have (id+Qdj(θ))1=(2cos(θ/2))1Qdj(θ/2). Set u~:=Aj1(id+Qdj(θ))1Aju. Then u~Wj, and applying (Equation5.3) to v+u~ and taking the limit along x^n yields

0v+u~,v+u~limS(x^n)(v+u~),limS(x^n)(v+u~)=v2+u~2limS(x^n)v2limS(x^n)u~2+2v,u~2limS(x^n)v,limS(x^n)u~=2v,u~2limS(x^n)v,limS(x^n)u~,

where the last equality follows from the fact that each S(x^n) is an isometry on Wi and Wj. Since the inequality above also holds for vu~, we obtain the equality

0=v,u~limS(x^n)v,limS(x^n)u~=v,u~+v,AjQdj(θ)Aj1u~=v,Aj(id+Qdj(θ))Aj1u~=v,u.

Since v,u were arbitrary, this shows orthogonality. Hence, the decomposition V=i=1kWi is orthogonal, and since Si(x) is an isometry on Wi, it follows that S(x):=i=1kSi(x) is in SO(d). Setting M=i=1kMi, we obtain

g(x)=S(x)exp(Mx).

Corollary 5.4.

Assume that S(x)=AQdν(x)A1 is a semigroup with A being an invertible matrix such that S(x) is an isometry for each x0. Then, there exists an orthogonal matrix U such that

S(x)=UQdν(x)UT.

Proof.

One can easily verify the identities S(x)+S(x)T=2cos(ν(x))id and

sin(ν(y))S(x)sin(ν(x))S(y)=sin(ν(x)ν(y))id.

Choose x > 0 such that sin(ν(x))0. Such an x clearly exists since ν is linear on Qx. Choose any vV of unit length and set

u:=S(x)vcos(ν(x))vsin(ν(x)).

The definition of u is invariant under the choice of x as long as sin(ν(x))0. This can be seen by noticing that

sin(ν(y))S(x)sin(ν(x))S(y)=sin(ν(x)ν(y))idS(x)vcos(ν(x))vsin(ν(x))=S(y)vcos(ν(y))vsin(ν(y)).

Hence we obtain S(x)v=cos(ν(x))v+sin(ν(x))u. We have

u,v=1sin(ν(x))S(x)vcos(ν(x))v,v=1sin(ν(x))(S(x)v,vcos(ν(x)))=0,

where the last equality follows from

2S(x)v,v=S(x)v,v+S(x)Tv,v=S(x)v+S(x)Tv,v=2cos(ν(x)).

Similarly, we can show that u is also of unit length. Set H:=span(u,v). Clearly, H is invariant under (S(x))x and hence so is H since each S(x)SO(d). In this manner, we can construct an orthonormal basis, and we denote by U the matrix associated with this change of basis. Then, we have

Ug(x)UT=UAS(x)A1UTexp(UMUTx)=Qdν(x)exp(UMUTx).

Lemma 5.5.

Suppose that the semigroup g, acting on V, satisfies Assumption 2.5. Let V=V1V2 be the decomposition of Corollary 3.8 such that g|V1 is invertible. Then, for any vV1 we have

limx0g(x)vv=1.

Proof.

By Theorem 5.3, we have g(x)|V1=S(x)exp(Mx). Since S(x)SO(V1), we obtain

g(x)v,g(x)v=exp(Mx)v,exp(Mx),vV1,

which is continuous in x. Hence

limx0g(x)v=limx0exp(Mx)v=v.

Corollary 5.6.

Suppose that the semigroup g, acting on V, satisfies Assumption 2.5. Then, the decomposition from Corollary 3.8 is orthogonal.

Proof.

Let V=V1V2 with g(x)|V1 being non-degenerate and g(x)|V20. Assume V1 is not orthogonal to V2. Then, there exists vV1 such that pV2(v)0, where pV2 denotes the orthogonal projection onto V2. By Lemma 5.5, we have

limx0g(x)vv=1.

Calculating the same limit for vpV2(v) we obtain

limx0g(x)(vpV2(v))vpV2(v)=limx0g(x)vvvvpV2(v)=vvpV2(v).

Since v2=vpV2(v)2+pV2(v)2 and pV2(v)0, we have vvpV2(v)>1. Hence

limx0g(x)(vpV2(v))vpV2(v)>1,

but this contradicts g(x)opf(x). Hence V1V2.

Proof.

Proof of Theorem 2.6

By Corollary 5.6, we have the orthogonal decomposition V=V~1V~2 with g(x)|V~20 and g(x)|V~1 non-degenerate. Applying Theorem 5.3 to g(x)|V~1 yields the result.

Supplemental material

Disclosure statement

No potential conflict of interest was reported by the author(s).

Supplemental data

Supplemental data for this article can be accessed online at https://doi.org/10.1080/27684830.2023.2289203.

Additional information

Funding

Financial support from the Austrian Science Fund (FWF) under grants P 30750 and Y 1235 is gratefully acknowledged.

References

  • Aczél, J. (1966). Lectures on functional equations and their applications. In Mathematics in science and engineering, (Vol. 19). Academic Press.
  • Bingham, N. H., Goldie, C. M., & Teugels, J. L. (1987). Regular variation, vol. 27 of encyclopedia of Mathematics and its applications. Cambridge University Press.
  • Doob, J. L. (1953). Stochastic processes. John Wiley and Sons, Inc.
  • Engel, K.-J., & Nagel, R. (2000). One-parameter semigroups for linear evolution equations, vol. 194 of graduate texts in Mathematics. Springer-Verlag. ( With contributions by S. Brendle, M. Campiti, T. Hahn, G. Metafune, G. Nickel, D. Pallara, C. Perazzoli, A. Rhandi, S. Romanelli and R. Schnaubelt).
  • Goodman, R., & Wallach, N. R. (1998). Representations and invariants of the classical groups, vol. 68 of encyclopedia of Mathematics and its applications. Cambridge University Press.
  • Humphreys, J. E. (1975). Linear algebraic groups, Graduate Texts in Mathematics, no. 21. Springer-Verlag.
  • Jacobson, N. (1962). Lie algebras, interscience tracts in Pure and applied Mathematics, no. 10. Interscience Publishers (a division of John Wiley and Sons, Inc.).
  • Kuczma, M., & Zajtz, A. (1966). On the form of real solutions of the matrix functional equation Φ(x)Φ(y) = Φ(xy) for non-singular matrices Φ. Publicationes Mathematicae Debrecen, 13(1–4), 257–262. https://doi.org/10.5486/PMD.1966.13.1-4.31
  • Lifshits, M. (2012). Lectures on Gaussian processes, Springer briefs in Mathematics. Springer.
  • Nourdin, I. (2012). Selected aspects of fractional Brownian motion, vol. 4 of Bocconi and Springer series. Springer, Milan; Bocconi University Press.
  • O’Meara, K. C., Clark, J., & Vinsonhaler, C. I. (2011). Advanced topics in linear algebra. Oxford University Press.
  • Zajtz, A. (1971) . On semigroups of linear operators. Zeszyty Nauk. Uniw. Jagielloń. Prace Mat, 181–184.

Appendix A.

Cauchy’s functional equation

It is classical that all continuous solutions of the EquationEquation (1.3), f(x)+f(y)=f(x+y), are linear, and that the non-linear solutions are not continuous, even not Lebesgue measurable, and have dense graphs. For this, and further references, we refer to [2, Section 1.1]. In this section, we provide two auxiliary results on Cauchy’s equation. They concern lifting solutions from an interval to the real line, resp. the joint behavior of two solutions that differ by a non-linear function.

Lemma A.1.

Let f:RR/[a,a) be a solution to Cauchy’s functional EquationEquation (1.3) on R/[a,a) with a > 0. Then, there exists a solution f~:RR of (Equation1.3) such that f(x)f~(x)mod[a,a). The solution f~ is linear if and only if f is linear.

Proof.

Take a Hamel basis (ri)iI of R such that ri[a,a) for every iI. This is clearly possible by rescaling every basis element if necessary. For any xR there exists a finite subset IxI and ciQ such that x=iIxciri. The function f~(x)=iIxcif(ri) satisfies f(x)f~(x)mod[a,a). If f is linear, then clearly f~ is linear as well. If f is not linear then there exist two basis elements r1 and r2 such that f(r1)r1r2f(r2)0, and hence f~ is also not linear.

Definition A.2.

We say that two solutions ν and η of Cauchy’s functional equation are equivalent if νη is linear.

Lemma A.3.

Let f,g:RR be two non-equivalent solutions of Cauchy’s functional equation. Then there exists a sequence (xn)n0 in R0, converging to 0, such that either limnf(xn)=π and limng(xn)=θ with |θ|<π or vice versa.

Proof.

Choose x,yR>0 such that the two vectors v1=(x,f(x),g(x)) and v2=(y,f(y),g(y)) are linearly independent and f(x)xf(y)yg(x)xg(y)y. This is possible, since

f(x)xf(y)y=g(x)xg(y)y

for all x,y>0 would imply that f − g is linear. Assume w.l.o.g. that |f(x)xf(y)y|>|g(x)xg(y)y|. Since f and g are both linear on Qx and Qy and v1 and v2 are linearly independent, there exist sequences qnQ and rnQ such that qnxrny0 for every n, limqnxrny=0 and limnf(qnxrny)=limnqnf(x)rnf(y)=π. We show that xn=qnxrny has the desired property. Clearly limrnqn=xy and

π=limnqnf(x)rnf(y)=limnqn(f(x)rnqnf(y))=(f(x)xyf(y))limqn.

Hence

|limqng(x)rng(y)|=|(g(x)xyg(y))limqn|=|π(g(x)xg(y)y)/(f(x)xf(y)y)|<π.