Loading [MathJax]/jax/output/SVG/jax.js
Research article

Generalizations of Wigner's theorem from rank-1 projections to rank-n projections

  • Received: 12 March 2025 Revised: 15 May 2025 Accepted: 20 May 2025 Published: 23 May 2025
  • We provide generalizations of the classical Wigner's theorem as well as Uhlhorn's version of Wigner's theorem by considering maps that send rank-1 projections to rank-n projections. Namely, we describe the general form of maps ϕ:P1(H)Pn(K) multiplying n times the transition probability and maps ϕ:P1(H)Pn(K) sending each complete orthogonal system of rank-1 projections to some complete orthogonal system of rank-n projections.

    Citation: Yulong Tian, Jinli Xu. Generalizations of Wigner's theorem from rank-1 projections to rank-n projections[J]. Electronic Research Archive, 2025, 33(5): 3201-3209. doi: 10.3934/era.2025140

    Related Papers:

    [1] Yaguo Guo, Shilin Yang . Projective class rings of the category of Yetter-Drinfeld modules over the $ 2 $-rank Taft algebra. Electronic Research Archive, 2023, 31(8): 5006-5024. doi: 10.3934/era.2023256
    [2] Xin Lu, Jinbang Yang, Kang Zuo . Strict Arakelov inequality for a family of varieties of general type. Electronic Research Archive, 2022, 30(7): 2643-2662. doi: 10.3934/era.2022135
    [3] Liu Yang, Yuehuan Zhu . Second main theorem for holomorphic curves on annuli with arbitrary families of hypersurfaces. Electronic Research Archive, 2024, 32(2): 1365-1379. doi: 10.3934/era.2024063
    [4] Vladimir Lazić, Fanjun Meng . On Nonvanishing for uniruled log canonical pairs. Electronic Research Archive, 2021, 29(5): 3297-3308. doi: 10.3934/era.2021039
    [5] Lie Fu, Victoria Hoskins, Simon Pepin Lehalleur . Motives of moduli spaces of rank $ 3 $ vector bundles and Higgs bundles on a curve. Electronic Research Archive, 2022, 30(1): 66-89. doi: 10.3934/era.2022004
    [6] Yimou Liao, Tianxiu Lu, Feng Yin . A two-step randomized Gauss-Seidel method for solving large-scale linear least squares problems. Electronic Research Archive, 2022, 30(2): 755-779. doi: 10.3934/era.2022040
    [7] Rongmin Zhu, Tiwei Zhao . The construction of tilting cotorsion pairs for hereditary abelian categories. Electronic Research Archive, 2025, 33(5): 2719-2735. doi: 10.3934/era.2025120
    [8] Jun Guo, Yanchao Shi, Weihua Luo, Yanzhao Cheng, Shengye Wang . Exponential projective synchronization analysis for quaternion-valued memristor-based neural networks with time delays. Electronic Research Archive, 2023, 31(9): 5609-5631. doi: 10.3934/era.2023285
    [9] Frédéric Campana . Algebraicity of foliations on complex projective manifolds, applications. Electronic Research Archive, 2022, 30(4): 1187-1208. doi: 10.3934/era.2022063
    [10] Yu Wang . Bi-shifting semantic auto-encoder for zero-shot learning. Electronic Research Archive, 2022, 30(1): 140-167. doi: 10.3934/era.2022008
  • We provide generalizations of the classical Wigner's theorem as well as Uhlhorn's version of Wigner's theorem by considering maps that send rank-1 projections to rank-n projections. Namely, we describe the general form of maps ϕ:P1(H)Pn(K) multiplying n times the transition probability and maps ϕ:P1(H)Pn(K) sending each complete orthogonal system of rank-1 projections to some complete orthogonal system of rank-n projections.



    Let H,K be complex or real separable Hilbert spaces and n a positive integer. As usual, the symbols Pn(H) and IH stand for the set of all rank-n self-adjoint projections on H, and the identity operator on H, respectively. For S,TPn(H), we say S is orthogonal to T iff ST=0 and the quantity Tr(ST) is the transition probability between S,T. Plainly, ST is equivalent to Tr(ST)=0. If uH is a unit vector, then the rank-1 projection onto span{u} will be denoted by uu. The transition probability associated with a pair of rank-1 projections (pure states) is the commonly used concept in quantum theory. We call a family {Si}Pn(H) a complete orthogonal system of rank-n projections (briefly, COSPn) iff

    SiSj whenever ij.

    ● There is no rank-1 projection T orthogonal to each Si.

    The celebrated Wigner's theorem [1, pp.251–254] states that if ϕ:P1(H)P1(H) is a bijection satisfying

    Tr(ϕ(S)ϕ(T))=Tr(ST),S,TP1(H), (1.1)

    equivalently, if ϕ preserves the transition probability between S and T, then there exists a unitary or an anti-unitary U:HH such that ϕ(A)=UAU. Recently, there has been considerable interest in improving and reproving this vital result in many ways (referred to in [2,3,4,5,6,7]).

    Wigner's theorem also serves as a frequently used tool for investigating the symmetries in some mathematical structures of quantum mechanics. Suppose that ϕ is a bijection on the set of all observables/the state space/the effect algebra, and such a map preserves a certain property/relation/operation relevant in quantum mechanics. The given problem is to characterize the form of such maps (symmetries), and a classical approach to this problem is to first show that ϕ preserves the rank-1 projections and the corresponding transition probability. This is the crucial step of the proof. Applying Wigner's theorem, one may immediately see that the restriction of ϕ to P1(H) has a nice behavior. Then the final step to prove that ϕ takes the desired form on the entire quantum structure is usually considered as an easier part of the proof. The interested readers are referred to [8, Chapter 2] and references therein for more examples of this approach and some background for the so-called preservers problems.

    When using the above method, sometimes we may not ensure that ϕ maps P1(H) into itself, and quite often we merely know that it preserves the zero-transition probability. This motivates us to search for a stronger version of the classical Wigner's theorem. The main aim of this paper is to provide the generalizations of Wigner's theorem in which instead of assuming that ϕ maps P1(H) into itself, we assume that ϕ maps P1(H) into Pn(K).

    Theorem 1.1. If ϕ:P1(H)Pn(K) is a map satisfying

    Tr(ϕ(S)ϕ(T))=nTr(ST),S,TP1(H), (1.2)

    then there exists a collection {V1,,Vn} of linear or conjugate linear isometries from H into K with mutually orthogonal ranges, such that

    ϕ(A)=ni=1ViAVi,AP1(H).

    Notice that the property (1.2) is equivalent to the following condition:

    ϕ(S)ϕ(T)HS=nSTHS,S,TP1(H),

    where HS represents the Hilbert–Schmidt norm. Namely, our result describes the general form of maps from P1(H) into Pn(K) multiplying n times the distance induced by this special norm. We point out that several papers [9,10] studied the isometries of Pn(H) with respect to the operator norm.

    For the case of dimH3, Uhlhorn [11] significantly generalized Wigner's theorem by replacing the assumption (1.1) with a weaker one: Tr(ST)=0Tr(ϕ(S)ϕ(T))=0. Uhlhorn's result has been further improved in [12,13]: It is proved that the bijectivity assumption can be relaxed when dimH<. Unfortunately, when dimH=, it is shown in [14] that there exist injective maps preserving orthogonality in both directions, which behave quite wildly. Thus, an additional hypothesis will be needed in the infinite-dimensional case.

    Theorem 1.2. Let dimH3. If ϕ:P1(H)Pn(K) is a map that sends each complete orthogonal system of rank-1 projections to some complete orthogonal system of rank-n projections, then there exists a collection {V1,,Vn} of linear or conjugate linear isometries from H into K, which have mutually orthogonal ranges and satisfy ni=1ViVi=I, such that

    ϕ(A)=ni=1ViAVi,AP1(H). (1.3)

    If 3dimH< and dimK=ndimH, then a map ϕ:P1(H)Pn(K) that preserves orthogonality only in one direction automatically sends each COSP1 to some COSPn. Therefore, a generalization (without bijectivity either) of Uhlhorn's theorem in matrix algebra is a direct consequence of Theorem 1.2.

    Corollary 1.3. Let 3dimH< and dimK=ndimH. If ϕ:P1(H)Pn(K) is a map that preserves orthogonality in one direction, then ϕ has the form (1.3).

    In what follows, we denote by C(H), F(H), and Fs(H) the set of compact operators, finite-rank operators, and finite-rank self-adjoint operators on H. The following lemma will be used to prove Theorem 1.1.

    Lemma 2.1. If ϕ:Fs(H)Fs(K) is a linear map that sends rank-1 projections to rank-n projections and satisfies

    Tr(ϕ(S)ϕ(T))=nTr(ST),S,TFs(H), (2.1)

    then there exists a collection {V1,,Vn} of linear or conjugate linear isometries from H into K with mutually orthogonal ranges, such that

    ϕ(A)=ni=1ViAVi,AFs(H).

    To prove Lemma 2.1, we need the following lemmas. For S,TFs(H), we write ST if TS is positive.

    Lemma 2.2. Let ϕ:Fs(H)Fs(K) be a linear map that preserves projections. If S,TFs(H) are projections with ST, then ϕ(S)ϕ(T).

    Proof. Since S,T are projections with ST, there exists some projection R orthogonal to T, such that S=T+R. Thus, ϕ(S)=ϕ(T)+ϕ(R)ϕ(T).

    Lemma 2.3. (see [15, Theorem 1.9.1]) Let M be a dense subspace of a normed space V, and W a Banach space. If ϕ:MW is a continuous linear map, then ϕ has a unique continuous linear extension ϕ:VW.

    Proof of Lemma 2.1. By Eq (2.1), we see that ϕ sends orthogonal rank-1 projections to orthogonal rank-n projections. Clearly, any finite-rank projection is the sum of mutually orthogonal rank-1 projections. Consequently, ϕ preserves the projections.

    Assume that the underlying space H is complex. Extend ϕ to a complex linear map from F(H) into F(K) by setting

    ˜ϕ(A+iB):=ϕ(A)+iϕ(B),A,BFs(H).

    Let A=iαiPi, αiR, PiP1(H), denote the spectral decomposition of any operator AFs(H). Then ˜ϕ(Pi)˜ϕ(Pj)=0 for each ij, and hence ˜ϕ(A2)=˜ϕ(A)2. Replacing A by A+B, with A,BFs(H), we obtain that ˜ϕ(AB+BA)=˜ϕ(A)˜ϕ(B)+˜ϕ(B)˜ϕ(A). Then it follows that

    ˜ϕ((A+iB)2)=˜ϕ(A2)˜ϕ(B2)+i˜ϕ(AB+BA)=˜ϕ(A)2˜ϕ(B)2+i(˜ϕ(A)˜ϕ(B)+˜ϕ(B)˜ϕ(A))=(˜ϕ(A)+i˜ϕ(B))2=˜ϕ(A+iB)2.

    This implies that ˜ϕ is a Jordan homomorphism. Since ˜ϕ preserves the self-adjoint operators, we infer that ˜ϕ is a (continuous) Jordan - homomorphism. It is known that F(H) is dense in the C - algebra C(H). By Lemma 2.3, ˜ϕ can be uniquely extended to a Jordan - homomorphism from C(H) into C(K). According to [8, Theorem A.6], each Jordan - homomorphism of the C - algebra is a direct sum of a -antihomomorphism and a - homomorphism. Every - homomorphism of C(H) is in fact a direct sum of inner homomorphisms (see [16, Theorem 10.4.7]). Then ˜ϕ has the asserted form.

    The case when H is real demands an other approach (this idea is borrowed from [17, Theorem 2.2] below). Assume that {ui}iΩ is an orthonormal basis for H and denote rngϕ(uiui)=Ki, iΩ. For any i,jΩ with ij, since [(ui+uj)(ui+uj)]/2 is a projection with range lying within that of uiui+ujuj, it follows by Lemma 2.2 that

    12ϕ((uiuj+ujui)+(uiui+ujuj))IKiKj0.

    Therefore, we may write

    Pij=ϕ(uiuj+ujui)=[PiiPijPjiPjj]0

    for some linear operator Pij:KjKi. For any nonzero αR, consider

    Q1=(α2+1)1(α2u1u1+α(u1u2+u2u1)+u2u2)P1(H),Q2=(α2+1)1(u1u1α(u1u2+u2u1)+α2u2u2)P1(H).

    By directly computing, Q1Q2=0. It follows that

    0=(α2+1α)2ϕ(Q1)ϕ(Q2)=(ϕ(αu1u1+1αu2u2)+P12)( ϕ(1αu1u1+αu2u2)P12)=([αIK1001αIK2]0+P12)([1αIK100αIK2]0P12)=[IK100IK2]0P212[αP11αP121αP211αP22]0+[1αP11αP121αP21αP22]0=[IK100IK2]0P212[(α1α)P1100(1αα)P22]0.

    Because this equation holds true for any nonzero αR, we see that P11 and P22 are zeroes. Hence, we obtain

    P12P21=IK1 and P21P12=IK2.

    It follows that P12=P211=P21 and we have similar conclusions for all Pij.

    Since all Ki are isomorphic to Rn, there exists an isomorphism from HRn to ΩKi, given by iuiηi(ηi). Write K=(HRn)Ks and replace ϕ by the mapping

    A(IK1(iΩ,i1P1i)IKs)ϕ(A)(IK1(iΩ,i1P1i1)IKs)

    such that

    ϕ(u1ui+uiu1)=[(u1ui+uiu1)IRn]0Ks,iΩ.

    We are going to prove that

    ϕ(uiuj+ujui)=[(uiuj+uiuj)IRn]0Ks wheneveri,jΩwithij.

    To see this, let Z=[(u1+ui+uj)(u1+ui+uj)]/3. Then Z is a rank-1 projection such that, up to unitary similarity, ϕ(Z) is equal to a direct sum of 0 and

    Y=31[IRnIRnIRnIRnIRnPijIRnPij1IRn].

    As Y2=Y, it follows that IRn+2Pij=3Pij. Thus Pij=Pij1=IRn.

    Since spanR{uiuj+ujui:i,jΩ} is dense in Fs(H) when H is a real Hilbert space, we can prove that

    ϕ(A)=U[(AIRn)0Ks]U,AFs(H),

    where U:KK is a unitary. We arrive at the conclusion.

    Proof of Theorem 1.1. Since the whole Fs(H) is real linearly generated by P1(H), we may extend ϕ to a real-linear map ˜ϕ:Fs(H)Fs(K) by setting

    ˜ϕ(iλiSi):=iλiϕ(Si)

    where {λi}R and {Si}P1(H) are finite subsets. We claim that ˜ϕ is well-defined. Assume that iλiSi=jμjTj, {μj}R, {Tj}P1(H). Then for each AP1(H), it follows by Eq (1.2) that

    Tr(iλiϕ(Si)ϕ(A))=iλiTr(ϕ(Si)ϕ(A))=iλinTr(SiA)=nTr(iλiSiA)=nTr(jμjTjA)=jμjnTr(TjA)=jμjTr(ϕ(Tj)ϕ(A))=Tr(jμjϕ(Tj)ϕ(A)).

    This implies that

    Tr((iλiϕ(Si)jμjϕ(Tj))ϕ(A))=0.

    Based on the linearity of the function Tr, we can replace ϕ(A) by its linear combination. Then we obtain

    Tr((iλiϕ(Si)jμjϕ(Tj))(iλiϕ(Si)jμjϕ(Tj)))=0.

    Since the square of Hermitian operator (iλiϕ(Si)jμjϕ(Tj))2 is positive with zero trace, we deduce that

    (iλiϕ(Si)jμjϕ(Tj))2=0=iλiϕ(Si)jμjϕ(Tj).

    It means that ˜ϕ is well-defined. Then the form of this linear map ˜ϕ is given by Lemma 2.1.

    Proof of Theorem 1.2. First, let us recall Gleason's theorem [18]. A positive and trace-class operator σ:HH with Tr(σ)=1 is called a density operator. We restate Gleason's theorem as follows: Suppose that dimH3 and f:P1(H)[0,1] is a function such that for each complete orthogonal system of rank-1 projections {Si}P1(H), one has

    if(Si)=1.

    Then there is a density operator σ:HH for which

    f(S)=Tr(σS),SP1(H).

    To prove Theorem 1.2, we need to choose an arbitrary density operator ϱ:KK and define the function fϱ:P1(H)[0,1] by

    fϱ(S):=Tr(ϱϕ(S)),SP1(H).

    It follows from our assumption that Gleason's theorem can be used. Therefore, for each density operator ϱ:KK, there exists a density operator σ:HH such that

    fϱ(S)=Tr(σS),SP1(H).

    In particular, pick ϱ=ϕ(T)/n for some fixed TP1(H). Then we obtain

    fϱ(S)=1nTr(ϕ(T)ϕ(S))=Tr(σTS),SP1(H),

    where σT is the density operator corresponding to T. Taking S=T, we infer that

    Tr(σTT)=1.

    It is easy to verify that if u is a unit vector such that T=uu, then

    Tr(σTT)=σTu,u.

    As 0σTI, it follows by the operator theory that σTu=u. Therefore, 1 is an eigenvalue of σT and u belongs to the corresponding eigenspace. Under the decomposition H=span{u}{u}, the operator σT has the following matrix representation:

    σT=[100X],

    where X is the positive operator acting on {u} with zero trace. Thus, X=0, which means σT=T.

    Hence, for each SP1(H), we have Tr(ϕ(S)ϕ(T))=nTr(ST). Since T was chosen arbitrarily, we deduce that ϕ multiplies n times the transition probability. Then Theorem 1.1 tells us the form of the map ϕ.

    The conclusion in Theorem 1.2 does not hold when dimH=2, as demonstrated in the following example: In fact, we can identify H with C2 and hence F(H)=M2(C), the set of 2×2 complex matrices. All the rank-1 projections in M2(C) are in 1-to-1 correspondence with the unit vectors in the Bloch sphere in R3, i.e.: $

    P1(C2)={21[1+x1x2+ix3x2ix31x1]:x1,x2,x3R with x21+x22+x23=1}.

    It is straightforward to compute the orthogonal complement of

    A=21[1+x1x2+ix3x2ix31x1] is IA=21[1x1x2ix3x2+ix31+x1].

    Consider the bijective transformation ϕ:P1(C2)P1(C2), which fix all rank-1 projections, but change the role of [1000] and [0001]. Obviously, the only COSP1 in M2(C) that contains [1000] is {[1000],[0001]} and hence ϕ preserves orthogonality. However, this discontinuous transformation ϕ can not be extended to any linear transformation (in fact, any Jordan - homomorphism also) on the whole matrix space M2(C).

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    This study was funded by Fundamental Research Funds for the Central Universities of China (Grant No. 2572022DJ07).

    The authors declare there is no conflicts of interest.



    [1] E. P. Wigner, Gruppentheorie und ihre Anwendung auf die Quantenmechanik der Atomspektren, Frederik Vieweg und Sohn, Braunschweig, 1931.
    [2] L. Molnár, Transformations on the set of all n-dimensional subspaces of a Hilbert space preserving principal angles, Commun. Math. Phys., 217 (2001), 409–421. https://doi.org/10.1007/PL00005551 doi: 10.1007/PL00005551
    [3] G. P. Gehér, Wigner's theorem on Grassmann spaces, J. Funct. Anal., 273 (2017), 2994–3001. https://doi.org/10.1016/j.jfa.2017.06.011 doi: 10.1016/j.jfa.2017.06.011
    [4] P. Šemrl, Maps on Grassmann spaces preserving the minimal principal angle, Acta Sci. Math., 90 (2024), 109–122. https://doi.org/10.1007/s44146-023-00093-8 doi: 10.1007/s44146-023-00093-8
    [5] G. P. Gehér, An elementary proof for the non-bijective version of Wigner's theorem, Phys. Lett. A, 378 (2014), 2054–2057. https://doi.org/10.1016/j.physleta.2014.05.039 doi: 10.1016/j.physleta.2014.05.039
    [6] M. Pankov, L. Plevnik, A non-injective version of Wigner's theorem, Oper. Matrices, 17 (2023), 517–524. https://doi.org/10.7153/oam-2023-17-33 doi: 10.7153/oam-2023-17-33
    [7] P. Šemrl, Automorphisms of Hilbert space effect algebras, Phys. Lett. A, 48 (2015), 195301. https://doi.org/10.1088/1751-8113/48/19/195301 doi: 10.1088/1751-8113/48/19/195301
    [8] L. Molnár, Selected Preserver Problems on Algebraic Structures of Linear Operators and on Function Spaces, Springer, Berlin, 2007.
    [9] U. Uhlhorn, Representation of symmetry transformations in quantum mechanics, Ark. Fysik., 23 (1963), 307–340.
    [10] P. Šemrl, G. P. Gehér, Isometries of Grassmann spaces, J. Funct. Anal., 270 (2016), 1585–1601. https://doi.org/10.1016/j.jfa.2015.11.018 doi: 10.1016/j.jfa.2015.11.018
    [11] L. Molnár, J. Jamison, F. Botelho, Surjective isometries on Grassmann spaces, J. Funct. Anal., 265 (2013), 2226–2238. https://doi.org/10.1016/j.jfa.2013.07.017 doi: 10.1016/j.jfa.2013.07.017
    [12] M. Pankov, T. Vetterlein, A geometric approach to Wigner-type theorems, Bull. London. Math. Soc., 53 (2021), 1653–1662. https://doi.org/10.1112/blms.12517 doi: 10.1112/blms.12517
    [13] P. Šemrl, Wigner symmetries and Gleason's theorem, J. Phys. A, 54 (2021), 315301. https://doi.org/10.1088/1751-8121/ac0d35 doi: 10.1088/1751-8121/ac0d35
    [14] P. Šemrl, Orthogonality preserving transformations on the set of n-dimensional subspaces of a Hilbert space, Illinois J. Math., 48 (2004), 567–573. https://doi.org/10.1215/ijm/1258138399 doi: 10.1215/ijm/1258138399
    [15] E. R. Megginson, An Introduction to Banach Space Theory, Springer, New York, 2012.
    [16] R. V. Kadison, J. R. Ringrose, Fundamentals of the Theory of Operator Algebras. Volume Ⅱ: Advanced Theory, Academic press, New York, 1986.
    [17] C. K. Li, M. C. Tsai, Y. S. Wang, N. C. Wang, Nonsurjective zero product preservers between matrix spaces over an arbitrary field, Linear Multilinear Algebra, 72 (2024), 2406–2425. https://doi.org/10.1080/03081087.2023.2263139 doi: 10.1080/03081087.2023.2263139
    [18] M. A. Gleason, Measures on the closed subspaces of a Hilbert space, Indiana Univ. Math. J., 6 (1957), 885–893.
  • Reader Comments
  • © 2025 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(188) PDF downloads(24) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog