Open access peer-reviewed chapter

Linear K-Power Preservers and Trace of Power-Product Preservers

Written By

Huajun Huang and Ming-Cheng Tsai

Reviewed: 15 February 2022 Published: 17 April 2022

DOI: 10.5772/intechopen.103713

From the Edited Volume

Matrix Theory - Classics and Advances

Edited by Mykhaylo Andriychuk

Chapter metrics overview

126 Chapter Downloads

View Full Metrics

Abstract

Let V be the set of n×n complex or real general matrices, Hermitian matrices, symmetric matrices, positive definite (resp. semi-definite) matrices, diagonal matrices, or upper triangular matrices. Fix k∈Z\\01. We characterize linear maps ψ:V→V that satisfy ψAk=ψAk on an open neighborhood S of In in V. The k-power preservers are necessarily k-potent preservers, and k=2 corresponds to Jordan homomorphisms. Applying the results, we characterize maps ϕ,ψ:V→V that satisfy “trϕAψBk=trABk for all A∈V, B∈S, and ψ is linear” or “trϕAψBk=trABk for all A,B∈S and both ϕ and ψ are linear.” The characterizations systematically extend existing results in literature, and they have many applications in areas like quantum information theory. Some structural theorems and power series over matrices are widely used in our characterizations.

Keywords

  • k-power
  • power preserver
  • trace preserver
  • power series of matrices

1. Introduction

Preserver problem is one of the most active research areas in matrix theory (e.g. [1, 2, 3, 4]). Researchers would like to characterize the maps on a given space of matrices preserving certain subsets, functions or relations. One of the preserver problems concerns maps ψ on some sets V of matrices which preserves k-power for a fixed integer k2, that is, ψAk=ψAk for any AV (e.g. [3, 5, 6]). The k-power preservers form a special class of polynomial preservers. One important reason of this problem lies on the fact that the case k=2 corresponds to Jordan homomorphisms. Moreover, every k-power preserver is also a k-potent preserver, that is, Ak=A imply that ψAk=ψA for any AV. Some researches on k-potent preservers can be found in [6, 7, 8].

Given a field F, let MnF, SnF, DnF, NnF, and TnF denote the set of n×n general, symmetric, diagonal, strictly upper triangular, and upper triangular matrices over F, respectively. When F is the complex field C, we may write Mn instead of MnC, and so on. Let Hn, Pn, and Pn¯ denote the set of complex Hermitian, positive definite, and positive semidefinite matrices, and HnR=SnR, PnR, and PnR¯ the corresponding set of real matrices, respectively. A matrix space is a subspace of Mm,nF for certain m,nZ+. Let At (resp. A) denote the transpose (resp. conjugate transpose) of a matrix A.

In 1951, Kadison [9] showed that a Jordan -isomorphism on Mn, namely, a bijective linear map with ψA2=ψA2 and ψA=ψA for all AMn, is the direct sum of a -isomorphism and a -anti-isomorphism. Hence ψA=UAU for all AMn or ψA=UATU for all AMn by [[3], Theorem A.8]. Let k2 be a fixed integer. In 1992, Chan and Lim ([5]) determined a nonzero linear operator ψ:MnFMnF (resp. ψ:SnFSnF) such that ψAk=ψAk for all AMnF (resp. SnF) (See Theorems 3.1 and 5.1). In 1998, Brešar, Martindale, and Miers considered additive maps of general prime rings to solve an analogous problem by using the deep algebraic techniques ([10]). Monlár [[3], P6] described a particular case of their result which extends Theorem 3.1 to surjective linear operators on BH. In 2004, Cao and Zhang determined additive k-power preserver on MnF and SnF ([11]). They also characterized injective additive k-power preserver on TnF ([12] or [[6], Theorem 6.5.2]), which leads to injective linear k-power preserver on TnF (see Theorem 8.1). In 2006, Cao and Zhang also characterized linear k-power preservers from MnF to MmF and from SnF to MmF (resp. SmF) [8].

In this article, given an integer kZ\01, we show that a unital linear map ψ:VW between matrix spaces preserving k-powers on a neighborhood of identity must preserve all integer powers (Theorem 2.1). Then we characterize, for F=C and R, linear operators on sets V=MnF, Hn, SnF, Pn, PnR, DnF, and TnF that satisfy ψAk=ψAk on an open neighborhood of In in V. In the following descriptions, PMnF is invertible, UMnF is unitary, OMnF is orthogonal, and λF satisfies that λk1=1.

  1. V=MnF (Theorem 3.4): ψA=λPAP1 or ψA=λPAtP1.

  2. V=Hn (Theorem 4.1): When k is even, ψA=UAU or ψA=UAtU. When k is odd, ψA=±UAU or ψA=±UAtU.

  3. V=SnF (Theorem 5.2): ψA=λOAOt.

  4. V=Pn or PnR (Theorem 6.1): ψA=UAU or ψA=UAtU.

  5. V=DnF (Theorem 7.1): ψA=ψIndiagfp1AfpnA, in which ψInk=ψIn, p:1n01n is a function, and fi:DnFF i=01n satisfy that, for A=diaga1an, f0A=0 and fiA=ai for i=1,,n.

  6. V=TnF (Theorem 8.4 for n3): ψA=λPAP1 or ψA=λPAP1, in which PTnF and A=an+1j,n+1i if A=aij.

Our results on MnF and SnF extend Chan and Lim’s results in Theorems 3.1 and 5.1, and result on TnF extend Cao and Zhang’s linear version result in [12].

Another topic is the study of a linear map ϕ from a matrix set S to another matrix set T preserving trace equation. In 1931, Wigner’s unitary-antiunitary theorem [[3], p. 12] says that if ϕ is a bijective map defined on the set of all rank one projections on a Hilbert space H satisfying

trϕAϕB=trAB,E1

then there is an either unitary or antiunitary operator U on H such that ϕP=UPU or ϕP=UPtU for all rank one projections P. In 1963, Uhlhorn generalized Wigner’s theorem to show that the same conclusion holds if the equality trϕPϕQ=trPQ is replaced by trϕPϕQ=0trPQ=0 (see [13]).

In 2002, Molnár (in the proof of [[14], Theorem 1]) showed that maps ϕ on the space of all bounded linear operators on a Banach space BX satisfying (1) for ABX, rank one operator BBX are linear. In 2012, Li, Plevnik, and Šemrl [15] characterized bijective maps ϕ:SS satisfying trϕAϕB=ctrAB=c for a given real number c, where S is Hn, SnR, or the set of rank one projections.

In [[16], Lemma 3.6], Huang et al. showed that the following statements are equivalent for a unital map ϕ on Pn:

  1. trϕAϕB=trAB for A,BPn;

  2. trϕAϕB1=trAB1 for A,BPn;

  3. ϕA=UAU or UAtU for a unitary matrix U.

The authors also determined the cases if ϕ is not assuming unital, the set Pn is replaced by another set like Mn, Sn, Tn, or Dn. In [[17], Theorem 3.8], Leung, Ng, and Wong considered the relation (1) on infinite dimensional space.

Let S denote the subspace spanned by a subset S of a vector space. Recently, Huang and Tsai studied two maps preserving trace of product [18]. Suppose two maps ϕ:V1W1 and ψ:V2W2 between subsets of matrix spaces over a field F under some conditions satisfy

trϕAψB=trABE2

for all AV1, BV2. The authors showed that these two maps can be extended to bijective linear maps ϕ˜:V1W1 and ψ˜:V2W2 that satisfy trϕ˜Aψ˜B=trAB for all AV1, BV2 (see Theorem 2.2). Hence when a matrix space V is closed under conjugate transpose, every linear bijection ϕ:VV corresponds to a unique linear bijection ψ:VV that makes (2) hold (see Corollary 2.3). Therefore, each of ϕ and ψ has no specific form.

One natural question is to ask when the following equality holds for a fixed kZ\01:

trϕAψBk=trABk.E3

The second major work of this paper is to use our descriptions of linear k-power preservers on an open neighborhood S of In in V to characterize maps ϕ,ψ:VV under one of the assumptions:

  1. equality (3) holds for all AV, BS, and ψ is linear, or

  2. equality (3) holds for all A,BS and both ϕ and ψ are linear,

for the sets V=Mn, Hn, Pn, Sn, Dn, Tn, and their real counterparts. These results, together with Theorem 2.2 and the characterizations of maps ϕ1,,ϕm:VV (m3) that satisfy trϕ1A1ϕmAm=trA1Am in [18], make a comprehensive picture of the preservers of trace of matrix products in the related matrix spaces and sets.

In the following characterizations, F=C or R, P,QMnF are invertible, UMnF is unitary, OMnF is orthogonal, and cF\0.

  1. V=MnF (Theorem 3.5):

    1. When k=1, ϕA=PAQ and ψB=PBQ, or ϕA=PAtQ and ψB=PBtQ.

    2. When kZ\1,0,1, ϕA=ckPAP1 and ψB=cPBP1, or ϕA=ckPAtP1 and ψB=cPBtP1.

  2. V=Hn (Theorem 4.2):

    1. When k=1, ϕA=cPAP and ψB=cPBP, or ϕA=cPAtP and ψB=cPBtP, for c11.

    2. When kZ\1,0,1, ϕA=ckUAU and ψB=cUBU, or ϕA=ckUAtU and ψB=cUBtU, for cR\0.

  3. V=SnF (Theorem 5.3):

    1. When k=1, ϕA=cPAPt and ψB=cPBPt.

    2. When kZ\1,0,1, ϕA=ckOAOt and ψB=cOBOt.

  4. V=Pn and PnR (Theorem 6.4): ϕA=ckUAU and ψB=cUBU, or ϕA=ckUAtU and ψB=cUBtU, in which cR+. Characterizations under some other assumptions are also given as special cases of Theorem 6.2 (Huang, Tsai [18]).

  5. V=DnF (Theorem 7.2): ϕA=PCkAP1,ψB=PCBP1 where P is a permutation matrix and C=DnF is diagonal and invertible.

  6. V=TnF (Theorem 8.5): ϕ and ψ send NnF to NnF, DϕDnF and DψDnF are characterized by Theorem 7.2, and Dϕ=DϕD. Here D denotes the map that sends ATnF to the diagonal matrix with the same diagonal as A.

The sets Mn, Hn, Pn, Sn, Dn, and their real counterparts are closed under conjugate transpose. In these sets, trAB=AB for the standard inner product. Our trace of product preservers can also be interpreted as inner product preservers, which have wide applications in research areas like quantum information theory.

Advertisement

2. Preliminary

2.1 Linear operators preserving powers

We show below that: given kZ\01, a unital linear map ψ:VW between matrix spaces preserving k-powers on a neighborhood of identity in V must preserve all integer powers. Let Z+ (resp. Z) denote the set of all positive (resp. negative) integers.

Theorem 2.1. Let F=C or R. Let VMpF and WMqF be matrix spaces. Fix kZ\01.

  1. Suppose the identity matrix IpV and AkV for all matrices A in an open neighborhood SV of Ip in V consisting of invertible matrices. Then

AB+BA:ABVV,E4
A1:AVis invertibleV.E5

In particular,

Ar:AVV,rZ+,andE6
Ar:AVis invertibleV,rZ.E7

  1. 2. Suppose IpV, IqW, and AkV for all matrices A in an open neighborhood SV of Ip in V consisting of invertible matrices. Suppose ψ:VW is a linear map that satisfies the following conditions:

ψIp=Iq,E8
ψAk=ψAk,ASV.E9

Then

ψAB+BA=ψAψB+ψBψA,A,BV,E10
ψA1=ψA1,invertibleAV.E11

In particular,

ψAr=ψAr,AV,rZ+,andE12
ψAr=ψAr,invertibleAV,rZ.E13

Proof. We prove the complex case. The real case is done similarly.

  1. 1. For each AV\0, there is ϵ>0 such that Ip+xASV for all xC with x<minϵ1A. Thus

Ip+xAk=Ip+xkA+x2kk12A2+V.E14

The second derivative

d2dx2Ip+xAkx=0=kk1A2V.E15

Since k01, we have A2V for all AV. Therefore, for A,BV,

AB+BA=A+B2A2B2V.E16

In particular, AV implies that ArV for all rZ+.

Cayley-Hamilton theorem implies that every invertible matrix A satisfies that A1=fA for a certain polynomial fxFx. Therefore, A1V, so that ArV for all rZ.

  1. 2. Now suppose (8) and (9) hold. The proof is proceeded similarly to the proof of part (1). For every AV, there is ϵ>0 such that for all xC with x<minϵ1A1ψA,

ψIp+xAk=Iq+xkψA+x2kk12ψA2+W,E17
ψIp+xAk=Iq+xkψA+x2kk12ψA2+W.E18

since (17) and (18) equal, we have

ψA2=ψA2,AV.E19

Therefore, for A,BV,

ψA+B2=ψA+B2E20

We get (10): ψAB+BA=ψAψB+ψBψA. In particular ψAr=ψAr for all AV and rZ+.

Every invertible AV can be expressed as A1=fA for a certain polynomial fxFx. Then ψA1=ψfA=fψA is commuting with ψA. Hence

2ψA1ψA=ψA1ψA+ψAψA1=ψA1A+AA1=2Iq.E21

We get ψA1=ψA1. Therefore, ψAr=ψAr for all rZ.

Theorem 2.1 is powerful in exploring k-power preservers in matrix spaces. Note that every k-power preserver is a k-potent preserver. Theorem 2.1 can also be used to investigate k-potent preservers in matrix spaces.

2.2 Two maps preserving trace of product

We recall two results about two maps preserving trace of product in [18]. They are handy in proving linear bijectivity of maps preserving trace of products. Recall that if S is a subset of a vector space, then S denotes the subspace spanned by S.

Theorem 2.2 (Huang, Tsai [18]). Let ϕ:V1W1 and ψ:V2W2 be two maps between subsets of matrix spaces over a field F such that:

dimV1=dimV2maxdimW1dimW2.

  1. AB are well-defined square matrices for ABV1×V2W1×W2.

  2. If AV1 satisfies that trAB=0 for all BV2, then A=0.

  3. ϕ and ψ satisfy that

trϕAψB=trAB,AV1,BV2.E22

Then dimV1=dimV2=dimW1=dimW2 and ϕ and ψ can be extended to bijective linear map ϕ˜:V1W1 and ψ˜:V2W2, respectively, such that

trϕ˜Aψ˜B=trAB,AV1,BV2.E23

A subset V of Mn is closed under conjugate transpose if A:AVV. A real or complex matrix space V is closed under conjugate transpose if and only if V equals the direct sum of its subspace of Hermitian matrices and its subspace of skew-Hermitian matrices.

Corollary 2.3 (Huang, Tsai [18]). Let V be a subset of Mn closed under conjugate transpose. Suppose two maps ϕ,ψ:VV satisfy that

trϕAψB=trAB,A,BV.E24

Then ϕ and ψ can be extended to linear bijections on V. Moreover, when V is a vector space, every linear bijection ϕ:VV corresponds to a unique linear bijection ψ:VV such that (24) holds. Explicitly, given an orthonormal basis A1A of V with respect to the inner product AB=trAB, ψ is defined by ψAi=Bi in which B1B is a basis of V with trϕAiBj=δi,j for all i,j1.

Corollary 2.3 shows that when a matrix space V is closed under conjugate transpose, every linear bijection ϕ:VV corresponds to a unique linear bijection ψ:VV that makes (24) hold. The next natural thing is to determine ϕ and ψ that satisfy trϕAψBk=trABk for a fixed kZ\01.

From now on, we focus on the fields F=C or R.

Advertisement

3. k-power linear preservers and trace of power-product preservers on Mn and MnR

3.1 k-power preservers on Mn and MnR

Chan and Lim described the linear k-power preservers on Mn and MnR for k2 in [7, Theorem 1] as follows.

Theorem 3.1. (Chan, Lim [5]) Let an integer k2. Let F be a field with charF=0 or charF>k. Suppose that ψ:MnFMnF is a nonzero linear operator such that ψAk=ψAk for all AMnF. Then there exist λF with λk1=1 and an invertible matrix PMnF such that

ψA=λPAP1,AMnF,orE25
ψA=λPAtP1,AMnF.E26

(25) and (26) need not hold if ψ is zero or is a map on a subspace of MnF. The following are two examples. Another example can be found in maps on DnF (Theorem 7.1).

Example 3.2. The zero map ψA0 clearly satisfies ψAk=ψAk for all AMn but they are not of the form (25) or (26).

Example 3.3. Let n=k+m, k,m2, and consider the operator ψ on the subspace W=MkMm of Mn defined by ψAB=ABt for AMk and BMm. Then ψAk=ψAk for all AW and kZ+, but ψ is not of the form (25) or (26).

We now generalize Theorem 3.1 to include negative integers k and to assume the k-power preserving condition ψAk=ψAk only on matrices nearby the identity.

Theorem 3.4. Let F=C or R. Let an integer kZ\01. Suppose that ψ:MnFMnF is a nonzero linear map such that ψAk=ψAk for all A in an open neighborhood of In consisting of invertible matrices. Then there exist λF with λk1=1 and an invertible matrix PMnF such that

ψA=λPAP1,AMnF,orE27
ψA=λPAtP1,AMnF.E28

Proof. We prove for the case F=C. The case F=R can be done similarly. Obviously, ψIn=ψInk=ψInk.

  1. First suppose k2. For each AMn, there exists ϵ>0 such that for all xC with x<ϵ, the following two power series converge and equal:

ψIn+xAk=ψIn+xi=0k1ψIniψAψInk1i+x2i=0k2j=0k2iψIniψAψInjψAψInk2ij+E29
ψIn+xAk=ψIn+xkψA+x2kk12ψA2+E30

Equating degree one terms above, we get

A=i=0k1ψIniψAψInk1i.E31

Applying (31), we have

InψAAψIn=ψInkψAψAψInk=ψInψAψAψIn.E32

Hence ψInψA=ψAψIn for AMn, that is, ψIn commutes with the range of ψ.

Now equating degree two terms of (29) and (30) and taking into account that k01, we have

ψInk2ψA2=ψA2.E33

Define ψ1A=ψInk2ψA for AMn. Then ψ1A2=ψ1A2 for all AMn. (31) and the assumption that ψ is nonzero imply that ψIn0. So ψ1InψIn=ψInk=ψIn0. Thus ψ1In0 and ψ1 is nonzero. By Theorem 3.1, there exists an invertible PMn such that ψ1A=PAP1 for AMn or ψ1A=PAtP1 for AMn. Moreover, ψIn commutes with all ψ1A, so that ψIn=λIn for a λC. By In=ψ1In=ψInk1, we get λk1=1. Therefore, ψA=λψ1A. We get (27) and (28).

  1. Next Suppose k<0. For every AMn, the power series expansions of ψIn+xAk and ψIn+xAk1 are equal when x is sufficiently small:

ψIn+xAk=ψIn1+xi=0k1ψIniψAψInk1i+E34
ψIn+xAk1=ψIn1xkψIn1ψAψIn1+E35

Equating degree one terms of (34) and (35), we get

In1ψAψIn1=i=0k1ψIniψAψInk1i.E36

Therefore,

kψAψIn1ψIn1ψA=ψIni=0k1ψIniψAψInk1ii=0k1ψIniψAψInk1iψIn=ψInkψAψAψInk=ψIn1ψAψAψIn1.E37

We get ψIn1ψA=ψAψIn1 for AMn. So ψIn1 and ψIn commute with the range of ψ. The following power series are equal for every AMn when x is sufficiently small:

ψIn+xAk=ψIn+xkψInk1ψA+x2kk12ψInk2ψA2+E38
ψIn+xAk=ψIn+xkψA+x2kk12ψA2+E39

Equating degree two terms of (38) and (39), we get ψInk2ψA2=ψA2. Let ψ1AψInk2ψA=ψIn1ψA. Then ψ1A2=ψ1A2 and ψ1 is nonzero. Using Theorem 3.1, we can get (27) and (39).

3.2 Trace of power-product preserers on Mn and MnR

Corollary 2.3 shows that every linear bijection ϕ:MnFMnF corresponds to another linear bijection ψ:MnFMnF such that trϕAψB=trAB for all A,BMnF. When m3, maps ϕ1,,ϕm on MnF that satisfy trϕ1A1ϕmAm=trA1Am for A1,,AmMnF are determined in [18].

If two maps on MnF satisfy the following trace condition about k-powers, then they have specific forms.

Theorem 3.5. Let F=C or R. Let kZ\01. Let S be an open neighborhood of In consisting of invertible matrices. Then two maps ϕ,ψ:MnFMnF satisfy that

trϕAψBk=trABk,E40

  1. for all AMnF,BS, and ψ is linear, or

  2. for all A,BS and both ϕ and ψ are linear,

if and only if ϕ and ψ take the following forms:

  1. a. When k=1, there exist invertible matrices P,QMnF such that

ϕA=PAQψB=PBQorϕA=PAtQψB=PBtQA,BMnF.E41

  1. b. When kZ\1,0,1, there exist cF\0 and an invertible matrix PMnF such that

ϕA=ckPAP1ψB=cPBP1orϕA=ckPAtP1ψB=cPBtP1A,BMnF.E42

Proof. We prove the case F=C; the case F=R can be done similarly.

Suppose assumption (2) holds. Then for every AMnF, there exists cF\0 such that IncAS, so that for all BS:

trBk=trϕIncA+AψBk=trIncABk+ctrϕAψBk.E43

Thus trϕAψBk=trABk for AMnF and BS, which leads to assumption (1).

Now we prove the theorem under assumption (1), that is, (40) holds for all AMnF and BS, and ψ is linear. Only the necessary part is needed to prove.

Let S=BPn¯:B1/kS, which is an open neighborhood of In in Pn¯. Define ψ˜:SMn such that ψ˜B=ψB1/kk. Then (40) implies that

trϕAψ˜B=trϕAψB1/kk=trAB,AMn,BS.E44

The complex span of S is Mn. By Theorem 2.2, ϕ is bijective linear, and ψ˜ can be extended to a linear bijection on Mn.

The linearity of ψ and (40) imply that for every BMn, there exists ϵ>0 such that In+xBS and the power series of In+xBk converges whenever x<ϵ. Then

trϕAψIn+Bk=trAIn+xBk,AMn,x<ϵ.E45

  1. 1. First suppose k2. Equating degree one terms and degree k1 terms on both sides of (45) respectively, we get the following identities for A,BMn:

trϕAi=0k1ψInk1iψBψIni=trkAB,E46
trϕAi=0k1ψBiψInψBk1i=trkABk1.E47

Let Ci:i=1n2 be a basis of projection matrices (i.e. Ci2=Ci) in Mn. For example, we may choose the following basis of rank 1 projections:

Eii:1in12Eii+Ejj+δEij+δ¯Eji:1i<jnδ{1i}.E48

By (40) and (47), for AMn and i=1,,n2,

trAψCik=trkACi=trϕAj=0k1ψCijψInψCik1j.E49

By the bijectivity of ϕ,

Cik=j=0k1ψCijψIψCik1j.E50

Therefore, for i=1,,n2,

0=ψCij=0k1ψCijψInψCik1jj=0k1ψCijψInψCik1jψCi=ψCikψInψInψCik.E51

Since

trCik=trϕ1ACik=trϕ1ACi,AMn,i=1,,n2,E52

the only matrix AMn such that trCik=0 for all i1n2 is the zero matrix. So ψCik:i=1n2 is a basis of Mn. (51) implies that ψIn=cIn for certain cC\0.

(46) shows that

ck1trϕAψB=trAB,A,BMn.E53

Therefore,

ck1trϕAψBk=trABk=trϕAψBk,AMn,BS.E54

The bijectivity of ϕ shows that ck1ψBk=ψBk for BS, that is,

c1ψBk=c1ψBk,BS.E55

Notice that c1ψIn=In. By Theorem 3.4, there is an invertible PMn such that ψ is of the form ψB=cPBP1 or ψB=cPBtP1 for BMn. Consequently, we get (42).

  1. 2. Now suppose k<0. Then ψIn is invertible. For every BMn and sufficiently small x, we have the power series expansion:

ψIn+Bk=In+In1ψB1ψIn1k=InIn1ψB+x2ψIn1ψBψIn1ψB+ψIn1k=ψInkxi=1kψIniψBψInk1+i+x2i=1kj=1k+1iψIniψBψInjψBψInk2+i+j+E56

Equating degree one terms and degree two terms of (45) respectively and using (56), we get the following identities for A,BMn:

trϕAi=1kψIniψBψInk1+i=trkAB,E57
trϕAi=1kj=1k+1iψIniψBψInjψBψInk2+i+j=trkk12AB2.E58

(57) and (40) imply that

i=1kψIniψBkψInk1+i=kψBk,BS.E59

Let FrB denote the degree r coefficient in the power series of ψIn+Bk. Then (57) and (58) show that:

k12F1B2=F2B,BMn.E60

Denote ψ1BψBψIn1. We discuss the cases k=1 and k1.

  1. When k=1, (60) leads to

ψB2=ψBψIn1ψB,BMn.E61

So ψ1B2=ψ1B2 for BMn. Note that ψ1In=In. By Theorem 3.4, there exists an invertible PMn such that ψ1B=PBP1 or ψ1B=PBtP1 for BMn. Let QP1ψIn. Then Q is invertible, and ψB=PBQ or ψB=PBtQ for BMn. Using (40), we get (41).

  1. b. Suppose the integer k<1. Then (60) implies that

k12ψIn1F1B2F2B2ψIn1=ψIn1F2BF2BψIn1,E62

which gives

1k2ψInkψB2ψB2ψInk=ψInkψBψIn1ψBψBψIn1ψBψInk.E63

In other words, for BMn:

ψInk1k2ψ1B2ψ1B2=1k2ψ1B2ψ1B2ψInk.E64

Let B=In+xE for an arbitrary matrix EMn. Then (64) becomes

ψInkx1kψ1E+x21k2ψ1E2ψ1E2=x1kψ1E+x21k2ψ1E2ψ1E2ψInk.E65

The equality on degree one terms shows that ψInk commutes with all ψ1E. Hence ψInk commutes with the range of ψ. (??) can be rewritten as

tri=1kψInk1+iϕAψIniψB=trkAB,A,BMn.E66

By Theorem 2.2, ψ is a linear bijection and its range is Mn. So ψInk=μIn for certain μC.

Now by (59), for BS:

kψInψBkkψBkψIn=ψIni=1kψIniψBkψInk1+ii=1kψIniψBkψInk1+iψIn=ψBkψInkψInkψBk=0E67

So ψIn commutes with ψBk for BS. In particular, ψIn commutes with ψ˜B=ψB1/kk for BS. The complex span of S is Mn, and ψ˜ can be extended to a linear bijection on Mn. Hence ψIn=cIn for certain cC\0. By (59), we get ψ1Bk=ψ1Bk for BS. Note that ψ1In=In. By Theorem 3.4, there is an invertible PMn such that ψ1B=PBP1 or ψ1B=PBtP1. Then ψB=cPBP1 or ψB=cPBtP1. Using (40), we get (42).

Remark 3.6 The following modifications could be applied to the proof of Theorem 3.5 for F=R:

  1. Let S be the collection of matrices A=QDQ1, in which D is nonnegative diagonal and QMnR is invertible, such that A1/k=QD1/kQ1S.

  2. We may choose the following basis of rank 1 projections of MnR to substitute (48):

Eii:1in12Eii+Ejj+Eij+Eji:1i<jnω1Eii+ω2Ejj+EijEji:1i<jn,E68

in which ω1,ω2 are distinct roots of x2x1=0..

The arguments in the above proof will be applied analogously to maps on the other sets discussed in this paper.

Advertisement

4. k-power linear preservers and trace of power-product preservers on Hn

4.1 k-power linear preservers on Hn

We give a result that determine linear operators on Hn that satisfy ψAk=ψAk on a neighborhood of In in Hn for certain kZ\01

Theorem 4.1. Fix kZ\01. A nonzero linear map ψ:HnHn satisfies that

ψAk=ψAkE69

on an open neighborhood of In consisting of invertible matrices if and only if ψ is of the following forms for certain unitary matrix UMn:

  1. When k is even,

ψA=UAU,AHn;orψA=UAtU,AHn.E70

  1. 2. When k is odd,

ψA=±UAU,AHn;orψA=±UAtU,AHn.E71

Proof. It suffices to prove the necessary part. Suppose (69) holds on an open neighborhood S of In in Hn.

  1. 1. First assume k2. Replacing Mn by Hn in part (1) of the proof of Theorem 3.4 up to (33), we can prove that ψIn commutes with the range of ψ, and ψ1AψInk2ψA is a nonzero linear map that satisfies ψ1A2=ψ1A2 for AHn.

Every matrix in Mn can be uniquely expressed as A+iB for A,BHn. Extend ψ1 to a map ψ˜:MnMn such that

ψ˜A+iB=ψ1A+iψ1B,A,BHn.E72

It is straightforward to check that ψ˜ is a complex linear bijection. Moreover, for A,BHn,

ψ1AB+BA=ψ1A+B2ψ1A2ψ1B2=ψ1A+B2ψ1A2ψ1B2=ψ1Aψ1B+ψ1Bψ1A.

It implies that

ψ˜A+iB2=ψ˜A+iB2,A,BHn.

By Theorem 3.1, there is an invertible matrix UMn such that

  1. ψ˜A=UAU1 for all AMn, or

  2. ψ˜A=UAtU1 for all AMn.

First suppose ψ˜A=UAU1. The restriction of ψ˜ on Hn is ψ1:HnHn. Hence for AHn, we have UAU1=UAU1=UAU; then UUA=AUU for all AHn, which shows that UU=cIn for certain cR+. By adjusting a scalar if necessary, we may assume that U is unitary. So ψInk2ψA=UAU. Then ψInk1=In, so that ψIn=In when k is even and ψInInIn when k is odd. Thus ψA=UAU when k is even and ψA=±UAU when k is odd. Similarly for the case ψ˜A=UAtU1. Therefore, (70) or (71) holds.

  1. 2. Now assume that k<0. Replacing Mn by Hn in part (2) of the proof of Theorem 3.4, we can show that ψIn commutes with the range of ψ, and furthermore the nonzero linear map ψ1AψIn1ψA satisfies that ψ1A2=ψ1A2. By arguments in the preceding paragraphs, we get (70) or (71).

4.2 Trace of power-product preservers on Hn

By Corollary 2.3, every linear bijection ϕ:HnHn corresponds to another linear bijection ψ:HnHn such that trϕAψB=trAB for all A,BHn. When m3, linear maps ϕ1,,ϕm:HnHn that satisfy trϕ1A1ϕmAm=trA1Am are characterized in [18].

Theorem 4.2. Let kZ\01. Let S be an open neighborhood of In in Hn consisting of invertible Hermitian matrices. Then two maps ϕ,ψ:HnHn satisfy that

trϕAψBk=trABk,E73

  1. for all AHn,BS, and ψ is linear, or

  2. for all A,BS and both ϕ and ψ are linear,

if and only if ϕ and ψ take the following forms:

  1. When k=1, there exist an invertible matrix PMn and c11 such that

ϕA=cPAPψB=cPBPA,BHn;orϕA=cPAtPψB=cPBtPA,BHn.E74

  1. b. When kZ\1,0,1, there exist a unitary matrix UMn and cR\0 such that

ϕA=ckUAUψB=cUBUA,BHn;orϕA=ckUAtUψB=cUBtUA,BHn.E75

Proof. Assumption (2) leads to assumption (1) (cf. the proof of Theorem 3.5). We prove the theorem under assumption (1). It suffices to prove the necessary part.

  1. When k2, in the part (1) of proof of Theorem 3.5, through replacing Mn by Hn, complex numbers by real numbers, and Theorem 3.1 or Theorem 3.4 by Theorem 4.1, we can prove that ϕψ has the forms in (75).

  2. When k<0, in the part (2) of proof of Theorem 3.5, through replacing Mn by Hn and complex numbers by real numbers, we can get the corresponding equalities of (56) (60) on Hn. The case k<1 can be proved completely analogously with the help of Theorem 4.1.

For the case k=1, the equality corresponding to (60) can be simplified as

ψB2=ψBψIn1ψB,BHn.E76

Let ψ1BψIn1ψB. Then ψ1:HnMn is a nonzero real linear map that satisfies ψ1B2=ψ1B2 for BHn. Extend ψ1 to a complex linear map ψ˜:MnMn such that

ψ˜A+iBψ1A+iψ1B,A,BHn.E77

Similarly to the arguments in part (1) of the proof of Theorem 4.1, we have ψ˜A+iB2=ψ˜A+iB2 for all A,BHn. Using Theorem 3.4 and the fact that ψ˜In=ψ1In=In, we can prove that there is an invertible PMn such that for all BHn, either ψ1B=P1BP or ψ1B=P1BtP. So

ψB=ψInP1BP,BHn;orE78
ψB=ψInP1BtP,BHn.E79

If ψB=ψInP1BP for BHn, then ψInP1BP=ψInP1BP=PBPψIn, which gives

PψInP1B=BPψInP1,BHn.E80

Hence PψInP1=cIn for certain cR\0. We have ψIn=cPP so that ψB=cPBP for BHn. Similarly for the case ψB=ψInP1BtP. Adjusting c and P by scalar factors simultaneously, we may assume that c11. It implies (74).

Remark 4.3. Theorem 4.2 does not hold if ψ is not assumed to be linear. Let k be a positive even integer. Let ψ˜:HnHn be any bijective linear map such that ψ˜Pn¯Pn¯. For example, ψ˜ may be a completely positive map of the form ψ˜B=i=1rNiBNi for r2, N1,,NrMn linearly independent, and at least one of N1,,Nr is invertible. By Corollary 2.3, there is a linear bijection ϕ:HnHn such that trϕAψ˜B=trAB for all A,BHn. Let ψ:HnHn be defined by ψB=ψ˜Bk1/k. Then

trϕAψBk=trϕAψ˜Bk=trABk,A,BHn.

Obviously, ψ may be non-linear, and the choices of pairs ϕψ are much more than those in (74) and (75).

Advertisement

5. k-power linear preservers and trace of power-product preservers on Sn and SnR

5.1 k-power linear preservers on Sn and SnR

Chan and Lim described the linear k-power preservers on SnF for k2 in [7, Theorem 2] as follows.

Theorem 5.1. (Chan, Lim [5]) Let an integer k2. Let F be an algebraic closed field with charF=0 or charF>k. Suppose that ψ:SnFSnF is a nonzero linear operator such that ψAk=ψAk for all ASnF. Then there exist λF with λk1=1 and an orthogonal matrix OMnF such that

ψA=λOAOt,ASn.E81

We generalize Theorem 5.1 to include the case SnR, to include negative integers k, and to assume the k-power preserving condition only on matrices nearby the identity.

Theorem 5.2. Let kZ\01. Let F=C or R. Suppose that ψ:SnFSnF is a nonzero linear map such that ψAk=ψAk for all A in an open neighborhood of In in SnF consisting of invertible matrices. Then there exist λF with λk1=1 and an orthogonal matrix OMnF such that

ψA=λOAOt,ASnF.E82

Proof. It suffices to prove the necessary part. In both k2 and k<0 cases, using analogous arguments as parts (1) and (2) of the proof of Theorem 3.4, we get that ψIn commutes with the range of ψ, and the nonzero map ψ1AψInk2ψA satisfies that ψ1A2=ψ1A2 for ASnF. Then

ψ1Aψ1B+ψ1Bψ1A=ψ1A+B2ψ1A2ψ1B2=ψ1A+B2ψ1A2ψ1B2=ψ1AB+BA.E83

In particular, ψ1Aψ1Ar+ψ1Arψ1A=2ψ1Ar+1 for rZ+. Using induction, we get ψ1A=ψ1A for all ASnF and Z+. By [26, Corollary 6.5.4], there is an orthogonal matrix OMnF such that ψ1A=OAOt. Since ψIn commutes with the range of ψ1, we have ψIn=λIn for certain λF in which λk1=1. So ψA=λOAOt as in (82).

Obviously, in F=R case, (82) has λ=1 when k is even and λ11 when k is odd.

5.2 Trace of power-product preservers on Sn and SnR

Corollary 2.3 shows that every linear bijection ϕ:SnFSnF corresponds to another linear bijection ψ:SnFSnF such that trϕAψB=trAB for all A,BSnF. When m3, maps ϕ1,,ϕm:SnFSnF that satisfy trϕ1A1ϕmAm=trA1Am are determined in [18].

We characterize the trace of power-product preserver for SnF here.

Theorem 5.3. Let F=C or R. Let kZ\01. Let S be an open neighborhood of In in SnF consisting of invertible matrices. Then two maps ϕ,ψ:SnFSnF satisfy that

trϕAψBk=trABk,E84

  1. for all ASnF,BS, and ψ is linear, or

  2. for all A,BS and both ϕ and ψ are linear,

if and only if ϕ and ψ take the following forms:

  1. When k=1, there exist an invertible matrix PMnF and cF\0 such that

ϕA=cPAPt,ψB=cPBPt,A,BSnF.E85

We may choose c=1 for F=C and c11 for F=R.

  1. b. When kZ\1,0,1, there exist cF\0 and an orthogonal matrix OMnF such that

ϕA=ckOAOt,ψB=cOBOt,A,BSnF.E86

Proof. Assumption (2) leads to assumption (1) (cf. the proof of Theorem 3.5). We prove the theorem under assumption (1). It suffices to prove the necessary part.

Obviously, SnHn=SnR and SnPn=PnR. Let SBPnR:B1/kS, which is an open neighborhood of In in PnR and whose real (resp. complex) span is SnR (resp. Sn). Using an analogous argument of the proof of Theorem 3.5, and replacing Mn by SnF, replacing the basis (48) of Mn by the following basis of rank 1 projections in SnF:

Eii:1in12Eii+Ejj+Eij+Eji:1i<jn,E87

and replacing the usage of Theorem 3.4 by that of Theorem 5.2, we can prove the case k2, and for k<0, we can get the corresponding equalities up to (60).

Define a linear map ψ1:SnFMnF by ψ1BψBψIn1.

When k=1, we get the corresponding equality of (61), so that ψ1B2=ψ1B2 for BSnF. Similar to the proof of Theorem 5.2, we get ψ1Br=ψ1Br for all rZ+. By [26, Theorem 6.5.3], there is an invertible matrix PMnF such that ψ1B=PBP1, so that ψB=PBP1ψIn for BSnF. Since ψB=ψBt, we get

P1ψInPtB=BP1ψInPt,BSnF.E88

Therefore, P1ψInPt=cIn for certain cF\0, so that ψB=cPBPt for all BSnF. Consequently, we get (85). The remaining claims are obvious.

When k<1, using analogous argument as in the proof of k<1 case of Theorem 3.5 and applying Theorem 5.2, we can get (86).

Advertisement

6. k-power linear preservers and trace of power-product preservers on Pn and PnR

In this section, we will determine k-power linear preservers and trace of power-product preservers on maps PnPn¯ (resp. PnRPnR¯). Properties of such maps can be applied to maps PnPn and Pn¯Pn¯ (resp. PnRPnR and PnR¯PnR¯).

6.1 k-power linear preservers on Pn and PnR

Theorem 6.1. Fix kZ\01. A nonzero linear map ψ:PnPn¯ (resp. ψ:PnRPnR¯) satisfies that

ψAk=ψAkE89

on an open neighborhood of In in Pn (resp. PnR) if and only if there is a unitary (resp. real orthogonal) matrix UMn such that

ψA=UAU,APn;orψA=UAtU,APn.E90

Proof. We prove the case ψ:PnPn¯. The sufficient part is obvious. About the necessary part, the nonzero linear map ψ:PnPn¯ can be easily extended to a linear map ψ˜:HnHn that satisfies ψ˜Ak=ψ˜Ak on an open neighborhood of In. By Theorem 4.1, we immediately get (90).

The case ψ:PnRPnR¯ can be similarly proved using Theorem 5.2.

6.2 Trace of powered product preservers on Pn and PnR

Now consider the maps PnPn¯ (resp. PnRPnR¯) that preserve trace of powered products. Unlike Mn and Hn, the set Pn (resp. PnR) is not a vector space. The trace of powered product preservers of two maps have the following forms.

Theorem 6.2 (Huang, Tsai [18]). Let a,b,c,dR\0. Two maps ϕ,ψ:PnPn¯ satisfy

trϕAaψBb=trAcBd,A,BPn,E91

if and only if there exists an invertible PMn such that

ϕA=PAcP1/aψB=P1BdP1/borϕA=PAtcP1/aψB=P1BtdP1/bA,BPn.E92

Theorem 6.3 (Huang, Tsai [18]). Given an integer m3 and real numbers α1,,αm,β1,,βmR\0, maps ϕi:PnPn¯ (i=1,,m) satisfy that

trϕ1A1α1ϕmAmαm=trA1β1Amβ1,A1,,AmPn,E93

if and only if they have the following forms for certain c1,,cmR+ with c1cm=1:

  1. When m is odd, there exists a unitary matrix UMn such that for i=1,,m:

    ϕiA=ci1/αiUAβi/αiU,APn.E94

  2. When m is even, there exists an invertible MMn such that for i=1,,m:

    ϕiA=ci1/αiMAβiM1/αi,iisodd,ci1/αiM1AβiM1/αi,iis even,APn.E95

    Both Theorems 6.2 and 6.3 can be analogously extended to maps PnRPnR¯ without difficulties.

    Theorem 6.2 determines maps ϕ,ψ:PnPn¯ that satisfy (91) throughout their domain. If we only assume the equality (91) for AB in certain subset of Pn×Pn and assume certain linearity of ϕ and ψ, then ϕ and ψ may have slightly different forms. We determine the case a=c=1 and b=d=kZ\0 here.

    Theorem 6.4. Let kZ\0. Let S be an open neighborhood of In in Pn. Two maps ϕ,ψ:PnPn¯ satisfy

    trϕAψBk=trABk,E96

  3. for all A,BPn, or

  4. for all AS,BPn, and ϕ is linear,

if and only if there exists an invertible PMn such that

ϕA=PAPψB=P1BkP1/korϕA=PAtPψB=P1BtkP1/kA,BPn.E97

The maps ϕ and ψ satisfy (96)

  1. for all APn,BS, and ψ is linear, or

  2. for all A,BS and both ϕ and ψ are linear,

if and only if when k11, ϕ and ψ take the form (97), and when kZ\1,0,1, there exist a unitary matrix UMn and cR+ such that

ϕA=ckUAUψB=cUBUorϕA=ckUAtUψB=cUBtUA,BPn.E98

Proof. It suffices to prove the necessary part.

The case of assumption (1) has been proved by Theorem 6.2.

Similar to the proof of Theorem 3.5, assumption (2) implies assumption (1); assumption (4) implies assumption (3). It remains to prove the case with assumption (3).

When k=1, assumption (3) is analogous to assumption (2), and we get (97).

Suppose kZ\10. Let ψ1:PnPn¯ be defined by ψ1BψB1/kk. Let S1BPn:B1/kS. Then (97) with assumption (2) becomes

trϕAψ1B=trAB,APn,BS1.E99

Let ψ˜:HnHn be the linear extension of ψ. By Theorem 2.2, ϕ can be extended to a linear bijection ϕ˜:HnHn such that

trϕ˜Aψ˜Bk=trϕ˜Aψ1Bk=trABk,AHn,BS.E100

By Theorem 4.2 and taking into account the ranges of ϕ and ψ, we see that when k=1, ϕ and ψ take the form of (97), and when kZ\1,0,1, ϕ and ψ take the form of (98).

Theorem 6.4 has counterpart results for ϕ,ψ:PnRPnR¯ and the proof is analogous using Theorem 5.3 instead of Theorem 4.2.

Advertisement

7. k-power linear preservers and trace of power-product preservers on Dn and DnR

Let F=C or R. Define the function diag:FnDnF to be the linear bijection that sends each c1cnt to the diagonal matrix with c1,,cn (in order) as the diagonal entries. Define diag1:DnFFn the inverse map of diag.

With the settings, every linear map ψ:DnFDnF uniquely corresponds to a matrix LψMnF such that

ψA=diagLψdiag1A,ADnF.E101

7.1 k-power linear preservers on Dn and DnR

We define the linear functionals fi:DnFF i=01n, such that for each A=diaga1anDnF,

f0A=0;fiA=ai,i=1,,n.E102

Theorem 7.1. Let F=C or R. Let kZ\01. Let S be an open neighborhood of In in DnF. A linear map ψ:DnFDnF satisfies that

ψAk=ψAk,AS,E103

if and only if

ψA=ψIndiagfp1AfpnA,ADnF,E104

in which ψInk=ψIn and p:1n01n is a function such that pi0 when k<0 for i=1,,n. In particular, a linear bijection ψ:DnFDnF satisfies (103) if and only if there is a diagonal matrix CMnF with Ck1=In and a permutation matrix PMnF such that

ψA=PCAP1,ADnF.E105

Proof. For every A=diaga1anDnF, when xF is sufficiently close to 0, we have In+xAS and the power series of In+xAk converges, so that ψIn+xAk=ψIn+xAk.

ψIn+xAk=ψIn+xkψA+x2kk12ψA2+E106
ψIn+xAk=ψIn+xkψInk1ψA+x2kk12ψInk2ψA2+E107

So for all ADnF:

ψA=ψInk1ψA,E108
ψA2=ψInk2ψA2.E109

The linear map ψ1AψInk2ψA satisfies that

ψ1A2=ψ1A2,ADnF.E110

By (101), let Lψ1=ijMnF such that diag1ψ1A=Lψ1diag1A for ADnF. Then (110) implies that for all A=diaga1anDnF:

j=1nijaj2=j=1nijaj2,i=1,2,,n.E111

Therefore, each row of Lψ1 has at most one nonzero entry and each nonzero entry must be 1. We get

ψ1A=diagfp1AfpnAE112

in which p:1n01n is a function. Suppose ψIn=diagλ1λn. Then (108) implies that ψA=ψInψ1A has the form (104). Obviously, ψInk=ψIn and when k<0, each pi0 for i=1,,n. Moreover, when ψ is a linear bijection, (112) shows that ψ1A=PAP1 for a permutation matrix P. (105) can be easily derived.

7.2 Trace of power-product preservers on Dn and DnR

In [18], we show that two maps ϕ,ψ:DnFDnF satisfy trϕAψB=trAB for A,BDnF if and only if there exists an invertible NMnF such that

ϕA=diagNdiag1A,ψB=diagNtdiag1B,A,BDnF.E113

When m3, the maps ϕ1,,ϕm:DnFDnF satisfying trϕ1A1ϕmAm=trA1Am for A1,,AmDnF are also determined in [18].

Next we consider the trace of power-product preserver on DnF.

Theorem 7.2. Let F=C or R. Let kZ\01. Let S be an open neighborhood of In in DnF. Two maps ϕ,ψ:DnFDnF satisfy that

trϕAψBk=trABk,E114

  1. for all ADnF,BS, and ψ is linear, or

  2. for all A,BS and both ϕ and ψ are linear,

if and only if there exist an invertible diagonal matrix CDnF and a permutation matrix PMnF such that

ϕA=PCkAP1,ψB=PCBP1,A,BDnF.E115

Proof. Assumption (2) leads to assumption (1) (cf. the proof of Theorem 3.5). We prove the theorem under assumption (1).

For every BDnF, In+xBS and the power series of In+xBk converges when xF is sufficiently close to 0, so that

trϕAψIn+xBk=trAIn+xBkE116

Comparing degree one terms and degree two terms in the power series of the above equality, respectively, we get the following equalities for A,BDnF:

trϕAψBψInk1=trAB,E117
trϕAψB2ψInk2=trAB2.E118

Applying Theorem 2.2 to (117), ψIn is invertible and both ϕ and ψ are linear bijections. (117) and (119) imply that ψB2ψInk1=ψB2ψInk2. Let ψ1BψBψIn1. Then ψ1B2=ψ1B2 for BDnF. By Theorem 7.1 and ψ1In=In, there exists a permutation matrix PMnF such that ψ1B=PBP1 for BDnF. So ψB=ψInPBP1=PCBP1 for CP1ψInPDnF. Then (114) implies (115).

Advertisement

8. k-power injective linear preservers and trace of power-product preservers on Tn and TnR

8.1 k-power preservers on TnF

The characterization of injective linear k-power preserver on TnF can be derived from Cao and Zhang’s characterization of injective additive k-power preserver on TnF ([12] or [[6], Theorem 6.5.2]).

Theorem 8.1 (Cao and Zhang [12]). Let k2 and n3. Let F be a field with charF=0 or charF>k. Then ψ:TnFTnF is an injective linear map such that ψAk=ψAk for all ATnF if and only if there exists a k1th root of unity λ and an invertible matrix PTnF such that

ψA=λPAP1,ATnF,orE119
ψA=λPAP1,ATnF,E120

where A=an+1j,n+1i if A=aij.

Example 8.2. When n=2, the injective linear maps that satisfy ψAk=ψAk for AT2F send A=a11a120a22 to the following ψA:

λa11ca120a22,λa22ca120a11,E121

in which λk1=1 and cF\0..

Example 8.3. Theorem 8.1 does not hold if ψ is not assumed to be injective. Let n=3 and suppose ψ:T3FT3F is a linear map that sends A=aij3×3T3F to one of the following ψA (c,dF):

a11ca1200a22da2300a33,a33000a11000a22,a220ca1200000a11.E122

Then each ψ satisfies that ψAk=ψAk for every positive integer k but it is not of the forms in Theorem 8.1.

We extend Theorem 8.1 to the following result that includes negative k-powers and that only assumes k-power preserving in a neighborhood of In.

Theorem 8.4. Let F=C or R. Let integers k0,1 and n3. Suppose that ψ:TnFTnF is an injective linear map such that ψAk=ψAk for all A in an open neighborhood of In in TnF consisting of invertible matrices. Then there exist λF with λk1=1 and an invertible matrix PTnF such that

ψA=λPAP1,ATnF,orE123
ψA=λPAP1,ATnF.E124

where A=an+1j,n+1i=JnAtJn if A=aij, Jn is the anti-diagonal identity.

Proof. Obviously ψ is a linear bijection. Follow the same process in the proof of Theorem 3.4. In both k2 and k<0 cases we have ψIn commutes with the range of ψ, so that ψIn=λIn for λF and λk1=1. Moreover, let ψ1AψIn1ψA, then ψ1 is injective linear and ψ1A2=ψ1A2 for ATnF. Theorem 8.1 shows that ψ1A=PAP1 or ψ1A=PAP1 for certain invertible PTnF. It leads to (123) and (124).

8.2 Trace of power-product preservers on Tn and TnR

Theorem 2.2 or Corollary 2.3 does not work for maps on TnF. However, the following trace preserving result can be easily derived from Theorem 7.2. We have TnF=DnFNnF. Let DA denote the diagonal matrix that takes the diagonal of ATnF.

Theorem 8.5. Let F=C or R. Let kZ\01. Let S be an open neighborhood of In in TnF consisting of invertible matrices. Then two maps ϕ,ψ:TnFTnF satisfy that

trϕAψBk=trABk,E125

  1. for all ATnF,BS, and ψ is linear, or

  2. for all A,BS and both ϕ and ψ are linear,

if and only if ϕ and ψ send NnF to NnF, DϕDnF and DψDnF are linear bijections characterized by (115) in Theorem 7.2, and Dϕ=DϕD.

Proof. The sufficient part is easy to verify. We prove the necessary part here. Let ϕDϕDnF and ψDψDnF. Then ϕ,ψ:DnFDnF satisfy trϕAψBk=trABk for A,BDnF. So they are characterized by (115). The bijectivity of ϕ and ψ implies that ϕ and ψ must send NnF to NnF in order to satisfy (125). Moreover, ϕ should send matrices with same diagonal to matrices with same diagonal, which implies that Dϕ=DϕD.

Advertisement

9. Conclusion

We characterize linear maps ψ:VV that satisfy ψAk=ψAk on an open neighborhood S of In in V, where kZ\01 and V is the set of n×n general matrices, Hermitian matrices, symmetric matrices, positive definite (resp. semi-definite) matrices, diagonal matrices, or upper triangular matrices, over the complex or real field. The characterizations extend the existing results of linear k-power preservers on the spaces of general matrices, symmetric matrices, and upper triangular matrices.

Applying the above results, we determine the maps ϕ,ψ:VV on the preceding sets V that satisfy trϕAψBk=trABk

  1. for all AV, BS, and ψ is linear, or

  2. for all A,BS and both ϕ and ψ are linear.

These results, together with Theorem 2.2 about maps satisfying trϕAψB=trAB and the characterizations of maps ϕ1,,ϕm:VV (m3) satisfying trϕ1A1ϕmAm=trA1Am in [18], make a comprehensive picture of the preservers of trace of matrix products in the related matrix spaces and sets. Our results can be interpreted as inner product preservers when V is close under conjugate transpose, in which wide applications are found.

There are a few prospective directions to further the researches.

First, for a polynomial or an analytic function fx and a matrix set V, we can consider “local” linear f-preservers, that is, linear operators ψ:VV that satisfy ψfA=fψA on an open subset S of V. A linear f-preserver ψ on S also preserves matrices annihilated by f on S, that is, fA=0 (AS) implies fψA=0. When S=V is Mn, BH, or some operator algebras, extensive studies have been done on operators preserving elements annihilated by a polynomial f; for examples, the results on Mn by R. Howard in [19], by P. Botta, S. Pierce, and W. Watkins in [20], and by C.-K. Li and S. Pierce in [21], on BH by P. Šemrl [22], on linear maps ψ:BHBK by Z. Bai and J. Hou in [23], and on some operator algebras by J. Hou and S. Hou in [24]. We may further explore linear f-preservers for a multivariable function fx1xr, that is, operator ψ satisfying ψfA1Ar=fψA1ψAr. The corresponding annihilator preserver problem has been studied in some special cases, for example, on Mn for homogeneous multilinear polynomials by A. E. Guterman and B. Kuzma in [25].

Second, it is interesting to further investigate maps ϕ,ψ:VV that satisfy trfϕAgψB=trfAgB for some polynomials or analytic functions fx and gx. This is equivalent to the inner product preserver problem fϕAgψB=fAgB when V is close under conjugate transpose. More generally, given a multivariable function hx1xm, we can ask what combinations of linear operators ϕ1,,ϕm:VV satisfy that tr(hϕ1A1ϕmAm=trhA1Am. The research on this area seems pretty new. No much has been discovered by the authors.

References

  1. 1. Li CK, Pierce S. Linear preserver problems. American Mathematical Monthly. 2001;108:591-605
  2. 2. Li CK, Tsing NK. Linear preserver problems: A brief introduction and some special techniques. Linear Algebra and its Applications. 1992;162–164:217-235
  3. 3. Molnár L. Selected Preserver Problems on Algebraic Structures of Linear Operators and on Function Spaces, Lecture Notes in Mathematics. Vol. 1895. Berlin: Springer-Verlag; 2007
  4. 4. Pierce S et al. A survey of linear preserver problems. Linear and Multilinear Algebra. 1992;33:1-129
  5. 5. Chan GH, Lim MH. Linear preservers on powers of matrices. Linear Algebra and its Applications. 1992;162–164:615-626
  6. 6. Zhang X, Tang X, Cao C. Preserver Problems on Spaces of Matrices. Beijing: Science Press; 2006
  7. 7. Brešar M, Šemrl P. Linear transformations preserving potent matrices. Proceedings of the American Mathematical Society. 1993;119:81-86
  8. 8. Zhang X, Cao CG. Linear k-power/k-potent preservers between matrix spaces. Linear Algebra and its Applications. 2006;412:373-379
  9. 9. Kadison RV. Isometries of operator algebras. Annals of Mathematics. 1951;54:325-338
  10. 10. Brešar M, Martindale WS, Miers CR. Maps preserving nth powers. Communications in Algebra. 1998;26:117-138
  11. 11. Cao C, Zhang X. Power preserving additive maps (in Chinese). Advances in Mathematics. 2004;33:103-109
  12. 12. Cao C, Zhang X. Power preserving additive operators on triangular matrix algebra (in Chinese). Journal of Mathematics. 2005;25:111-114
  13. 13. Uhlhorn U. Representation of symmetry transformations in quantum mechanics. Arkiv för Matematik. 1963;23:307-340
  14. 14. Molnár L. Some characterizations of the automorphisms of and CX. Proceedings of the American Mathematical Society. 2002;130(1):111-120
  15. 15. Li CK, Plevnik L, Šemrl P. Preservers of matrix pairs with a fixed inner product value. Operators and Matrices. 2012;6:433-464
  16. 16. Huang H, Liu CN, Tsai MC, Szokol P, Zhang J. Trace and determinant preserving maps of matrices. Linear Algebra and its Applications. 2016;507:373-388
  17. 17. Leung CW, Ng CK, Wong NC. Transition probabilities of normal states determine the Jordan structure of a quantum system. Journal of Mathematical Physics. 2016;57:015212
  18. 18. Huang H, Tsai MC. Maps Preserving Trace of Products of Matrices. Available from: https://arxiv.org/abs/2103.12552 [Preprint]
  19. 19. Howard R. Linear maps that preserve matrices annihilated by a polynomial. Linear Algebra and its Applications. 1980;30:167-176
  20. 20. Botta P, Pierce S, Watkins W. Linear transformations that preserve the nilpotent matrices. Pacific Journal of Mathematics. 1983;104(1):39-46
  21. 21. Li C-K, Pierce S. Linear operators preserving similarity classes and related results. Canadian Mathematical Bulletin. 1994;37(3):374-383
  22. 22. Šemrl P. Linear mappings that preserve operators annihilated by a polynomial. Journal of Operator Theory. 1996;36(1):45-58
  23. 23. Bai Z, Hou J. Linear maps and additive maps that preserve operators annihilated by a polynomial. Journal of Mathematical Analysis and Applications. 2002;271(1):139-154
  24. 24. Hou J, Hou S. Linear maps on operator algebras that preserve elements annihilated by a polynomial. Proceedings of the American Mathematical Society. 2002;130(8):2383-2395
  25. 25. Guterman AE, Kuzma B. Preserving zeros of a polynomial. Communications in Algebra. 2009;37(11):4038-4064

Written By

Huajun Huang and Ming-Cheng Tsai

Reviewed: 15 February 2022 Published: 17 April 2022