Learning Symmetric Hamiltonian
Abstract
Hamiltonian Learning is a process of recovering system Hamiltonian from measurements, which is a fundamental problem in quantum information processing. In this study, we investigate the problem of learning the symmetric Hamiltonian from its eigenstate. Inspired by the application of group theory in block diagonal secular determination, we have derived a method to determine the number of linearly independent equations about the Hamiltonian unknowns obtained from an eigenstate. This number corresponds to the degeneracy of the associated irreducible representation of the Hamiltonian symmetry group. To illustrate our approach, we examine the XXX Hamiltonian and the XXZ Hamiltonian. We first determine the Hamiltonian symmetry group, then work out the decomposition of irreducible representation, which serves as foundation for analyzing the uniqueness of recovered Hamiltonian. Our numerical findings consistently align with our theoretical analysis.
I introduction
Hamiltonian Learning (HL) refers to recovering the Hamiltonian of a quantum system, typically from observations of its steady state or dynamics. HL plays a pivotal role in quantum physics. In recent years, there has been a rapid development in quantum simulators and related quantum computing devices, including controllable trapped ion [1, 2, 3, 4] and superconducting quantum circuits [5, 6]. The exponential growth of noise with the size of quantum hardware has made HL an essential tool for verification and benchmarking [7, 8, 9, 10, 11, 12]. In the filed of condensed matter physics, HL aids in understanding the behavior of quantum material and may be utilized to explore new material systems [13, 14, 15, 16, 17, 18, 19].
One of the challenge in HL lies in determining when we can uniquely recover a Hamiltonian from its steady state. This question was initially raised and partially addressed by Qi and Ranard , and they present that the most of the generic local Hamiltonian can be recovered from measurements of a single eigenstate [20]. From the perspective of dimension counting, the space of -local Hamiltonian scales linearly with the system size , whereas the space of states scales exponentially with . As a result, it becomes feasible to recover local Hamiltonian from its single eigenstate. Qi and Ranard introduced the concept of “correlation matrix”, which successfully recovered local Hamiltonian from a single eigenstate.
Following Qi’s question, various methods for recovering the Hamiltonian from a steady state have been proposed [21, 22, 23]. Most of these methods aim to simplify the process of measuring the steady state, while ignoring the case when recovering the Hamiltonian from its steady state is not feasible. In fact, there exists critical chain length in spin system with length for generic local Hamiltonian. When is smaller than , we cannot recover Hamiltonian form its steady state. The value of depends on the rank of the steady state and Hamiltonian model. In previous work, we utilize the energy eigen-value equations (EEE) to analytically determine the value of , successfully applying this approach to models of generic local Hamiltonians [24]. Subsequently, we extended this investigation to degenerate steady states using the orthogonal space equations (OSE) method [25].
So far, we focused on the recoverability of a generic local Hamiltonian from its steady state, while many Hamiltonians arouse interests exhibit symmetries [26, 27]. Such as the Heisenberg spin chain models [28]. Symmetric Hamiltonians inherently entail fewer unknowns compared to generic ones. However, it doesn’t necessarily make them easier to recover. According to Noether’s theorem, there is a direct relationship between the Hamiltonian symmetry and the conserved quantities in the associated physical system. It directly reduces the degrees of freedom of associated eigenstates [29]. In other words, the space dimension of eigenstate is reduced, which decreases the number of linearly independent equations (LIE) that can be derived from an eigenstate.
To determine the number of the LIE obtained from an eigenstate of symmetric Hamiltonian, we introduce the block diagonalization of secular determination, which is a commonly used technique in group theory. We expand the EEE in a symmetrized basis, and analyze the repeating structure of linearly equations. We find that the number of LIE equals the degeneracy of the associated irreducible representation (IR) of the Hamiltonian symmetry group. In other words, if we know the symmetry group of a Hamiltonian, we can infer how much information is contained in its eigenstate. To verify our theoretical derivation, we perform simulations on the XXX and XXZ Hamiltonians. Given only the Hamiltonian’s symmetry group and its IRs, we are able to successfully predict the number of LIEs obtained from the eigenstate.
II framework
II.1 Problem setting
The specific form of HL can be reduced to determine the unknowns in a local Hamiltonian
(1) |
where is the Pauli operator that acts on continuously site.
In general, the algorithms of HL from steady state often can be written in the forms of homogeneous linear equations
(2) |
where is the unknown vector for , and is the eigenvalue of corresponding eigenstate, is constraint matrix for unknowns . Eq. (2) extracts information from measurements and can be varied in different algorithms. For the EEE and the OSE method, is deduced from energy eigen-value equations. By calculating the number of LIE in Eq. (2), we can evaluate the degree to which a steady state characterizes a Hamiltonian and thus determine the recoverability of the Hamiltonian from its steady state.
We note that the rank of matrix equals the number of LIE for , which we denote as . The number of LIE for , denoted as , is equal to . In short, we have . Eq. (2) represents a system of homogeneous linear equations, such that the parameter of Hamiltonian can only be determined up to a linear factor , where . As a consequence, when the value of reaches , the Hamiltonian can be uniquely determined.
In this paper, we propose the first algorithm to calculate the number of LIE obtained from an eigenstate of symmetric Hamiltonian.
II.2 The symmetry group of Hamiltonian
The degeneracy of Hamiltonian’s energy level is intimately linked to its symmetry group. Supposing that Hamiltonian is invariant under the transformation operator , such that
(3) |
If is invariant under transformation and , then it follows that also remains invariant under their composition . Therefore, all transformations satisfying (3) constitute the symmetry group of . Eq. (3) can also be written as
(4) |
Now we consider an eigenstate of Hamiltonian , with eigenvalue . Applying the transformation operator to , the resulting state remains an eigenstate of ,
(5) |
Similarly, repeatedly applying on eigenstate , the resulting state also remains an eigenstate of .Consequently, all states , where are eigenstates correspond to the same eigenvalue .
Representation theory plays a vital role in group theory. Here, we introduce two fundamental theorems concerning representations, which reveal the relevance between properties of Hamiltonian and its symmetry group representation[30, 31].
Theorem 1.
The eigenfunctions of the Hamiltonian with the same eigenvalues constitute a set of basis functions for the symmetry group of the Hamiltonian.
Theorem 2.
In the absence of accidental degeneracy, a group of eigenstates that are transformed according to an IR of group , belong to the same energy level.
Through the theorems above, it can be inferred that the degeneracy of the Hamiltonian levels equals the dimension of the IR of the symmetry group . This insight inspires a useful concept: If we determine the symmetry group of an unknown Hamiltonian, we can uncover information about the degeneracy of its energy levels and the transformations between degenerate eigenstates.
Now we ask: Is it possible to uncover Hamiltonian’s symmetry group without knowing its specific form? The answer is yes. In fact, we can determine under which transformations the Hamiltonian ansatz specified in Eq. (1) is invariant. As long as all terms are invariant under certain transformations, the Hamiltonian will also be invariant under those transformations. In other words, the Hamiltonians under the same ansatz in Eq. (1) share the same symmetry group, resulting in consistent energy level degeneracy. This degeneracy, caused by symmetry, is referred to as intrinsic degeneracy. Mathematically, any symmetric transformation in this group is defined by
(6) |
We call the group as the symmetry group on Hamiltonian learning.
Next, we will explore the impact of this symmetry on Hamiltonian Learning.
II.3 The diagonalization of the Schodinger equation
Group theory can be utilized in simplifying the solution of Schodinger equations. In this subsection, we will introduce the application of group theory in the diagonalization of secular determination.
Supposing that Hamiltonian having group as its symmetry group. Starting with the Schodinger equation
(7) |
Expanding in the basis , ,
(8) |
Choosing a set of symmetrized basis for eigenstate , whose transformation is in accordance with IR of group . is the -th base of the -th IR of group , and is the index of the repeating IR. Expanding the eigenstate in the basis , and Eq. (2) can be written as
(9) |
Taking the inner product of with the equation above
(10) | |||
To solve the tedious equations above, we introduce the theorem of matrix elements for the invariant operator[32]:
Theorem 3 ( Theorem of matrix elements for the invariant operator).
If is the symmetry group of Hamiltonian , and is the basis of IR for , then the matrix elements of satisfy that
(11) |
where is independent of .
According to Theorem 3, Eq. (10) becomes
(12) |
which holds for any . Noting that coefficient is independent of . Eq. (12) can be written as
(13) |
which we see as a linear equation for the unknowns . Since is independent of , is also independent of . Consequently, Eq. (12) is rewritten as
(14) |
which reveals that once are fixed, Eqs. (12) are equivalent for all . Therefore Eqs. (12) can be divided into series of subsystem of linear equations, which are composed of repeating unit of equations. Noting that Eqs. (14) are complex equations for generic Hamiltonian. Here we consider a symmetric Hamiltonian whose basis can be written as a real matrix, making Eqs. (14) real equations.
The number of LIE in each unit depends on the degeneracy of IR, and repeating times of units depends on the dimension of IR. An example of such structure of linear equations with base functions which contains one 4-dimensional IR and two degenerate 2-dimensional IR is shown in Fig. 1.

So far, we have shown the repeating structure of system of linear equations and given the possible number of LIE obtained from a eigenstate of Hamiltonian. However, while considering a specific eigenstate, some of the equations in Eqs. (12) could be trivial [33].
Theorem 4.
Eigen-states with the same eigen-value furnish an IR of Hamiltonian symmetry group.
Eigenstate lies in a subspace spanned by the basis of a IR of , that is to say, every eigenstate has a corresponding . Eqs. (12) are nontrivial only when the value of is conform to . Consequently, the number of LIE obtained from an eigenstate is
(15) |
where is the degeneracy of IR in the Hilbert space.
III examples
III.1 The XXX Spin Chain
Considering an 1-dimensional spin chain of sites, with a spin particle at each site. The Hamiltonian of this model can be written as
(16) |
where
(17) |
is the ’th Pauli matrix, and is the identity matrix.
XXX Hamiltonian has time reversal and SU(2) symmetry. Its symmetry group is composed of
(18) |
where is the time reversal operator, and and is a set of anti-unitary operators. The representation of non-unitary group is denoted as co-representation.
To determine whether time-reversal symmetry will cause a doubling of energy degeneracy, it’s instructive to study the co-representation of the symmetry group [34]. can be derived from an IR of subgroup , whose specific forms are illustrated in appendix. A. The dimension of is doubled compared to the basis of . Without the time reversal symmetry, the energy level degeneracy of Hamiltonian is determined by the dimension of the IR of the group . If the co-representation is reducible, the degeneracy will not be doubled. However, if is irreducible, then the degeneracy will be doubled.
The co-representation of group is reducible and the degeneracy will not be doubled (see appendix. B). Therefore the dimension of IR of and SU(2) group is consistent. TABLE. 1 shows the decomposition of the SU(2) symmetry of the -site XXX Hamiltonian into a series of IRs.
Decomposition of representation | |
2 | |
3 | |
4 | |
5 | |
6 | |
7 |
Up to now, we have calculated the dimensions of IR of the symmetry group . Next, we will leverage the properties of these IRs to determine how much information can be extracted from an eigenstate of the XXX Hamiltonian. In other words, we will explore the number of linearly independent equations that can be deduced from a single eigenstate.
The number of LIE obtained from an eigenstate corresponds to the degeneracy of its associated irreducible subspace. For example, when the chain length , the Hilbert space is segmented into one subspace of dimension 4, and two subspaces of dimension 2, resulting in energy level degeneracies of 4, 2 and 2, respectively. For eigenstates lie in the 4-dimensional subspace, the degeneracy of the corresponding IR is 1. Thus we can only derive a single LIE from it. Conversely, eigenstates situated in the 2-dimensional subspaces have a degeneracy of 2, allowing us to obtain two LIEs. This elucidates that determining which subspace an eigenstate lies in enables us to ascertain how much information about the Hamiltonian it encodes.
To determine the subspace occupancy of an eigenstate within the formalism of angular momentum, one often utilizes ladder operators:
(19) |
Upon applying to an eigenstate of with eigenvalue , the resulting state remains an eigenstate of with eigenvalue .
In the context of a finite spin chain, the irreducible subspace is finite. Starting from an eigenstate of the XXX Hamiltonian, one can apply the ladder operators a finite number of times until it reaches 0. Within the irreducible subspace labeled by , the maximum eigenvalue of is , the minimum is . This can be expressed as
(20) |
The space that the ladder operators acting on, is exactly an irreducible subspace with dimension . All vectors within this subspace can be connected through the action of the ladder operators. Therefore, when given an eigenstate, by applying ladder operators, we can traverse its subspace and thereby determine where it lies in.
We perform simulations on recovering the Hamiltonian from a single eigenstate. The results for chain length are shown in TABLE. 2 and 3, respectively. We present the recovery result for all eigenstate, where the symbol ‘O’ indicates the success of HL and ‘X’ signifies a failure. We also record the energy level degeneracy (), the dimension of IR (), the degeneracy of IR (), and the number of LIEs for unknowns of the Hamiltonian ().
eigenstate | ||||||||
2 | 2 | 2 | 2 | 4 | 4 | 4 | 4 | |
2 | 2 | 2 | 2 | 4 | 4 | 4 | 4 | |
2 | 2 | 2 | 2 | 1 | 1 | 1 | 1 | |
1 | 1 | 1 | 1 | 0 | 0 | 0 | 0 | |
success? | O | O | O | O | X | X | X | X |
eigenstate | success? | ||||
3 | 3 | 3 | 2 | O | |
3 | 3 | 3 | 2 | O | |
3 | 3 | 3 | 2 | O | |
1 | 1 | 2 | 1 | X | |
3 | 3 | 3 | 2 | O | |
3 | 3 | 3 | 2 | O | |
3 | 3 | 3 | 2 | O | |
5 | 5 | 1 | 0 | X | |
5 | 5 | 1 | 0 | X | |
5 | 5 | 1 | 0 | X | |
5 | 5 | 1 | 0 | X | |
5 | 5 | 1 | 0 | X | |
1 | 1 | 2 | 1 | X | |
3 | 3 | 3 | 2 | O | |
3 | 3 | 3 | 2 | O | |
3 | 3 | 3 | 2 | O |
When the chain length , the Hamiltonian has two unknown parameters. Its Hilbert space is partitioned into one subspace with dimension 4 and two subspaces with dimension 2. When (1) , , , the Hamiltonian can be uniquely recovered, (2) , , , and the Hamiltonian cannot be uniquely recovered. These results are consistent with Equation (15), which relates the number of LIEs to the IR degeneracy .
When and , the dimension of IR is 4, labeled by quantum number . We can express the XXX Hamiltonian using the ladder operators acting on the -th site:
(21) |
From this expression, we can see that the states and are eigenstates of , with all spins either up or down in the -direction. Additionally, since the operator commutes with , i.e., , applying to the eigenstate results in another eigenstate of .
(22) |
repeatedly acting , until it reaches 0
(23a) |
(23b) |
(23c) |
From this example, we note that for any 3-site Hamiltonian, the following states are degenerate eigenstates, regardless of the specific model parameters: , , and . These four of degenerate eigenstates can be connected through the action of the ladder operators, but they do not provide any information about the Hamiltonian parameters. Therefore, the number of LIE obtained from these eigenstate is 0.
When , for all eigenstates, the relationship between the degeneracy of the IR and number of LIE is given by , as shown in TABLE. 3. When , the Hamiltonian can be uniquely recovered. Here, we point out that, under the case when the value of exceeds the value of Hamiltonian unknowns . The number of LIE is given by
(24) |
III.2 The XXZ Spin Chain
Now we consider 1-dimensional spin- chain with XXZ Hamiltonian give by
(25) |
The system described by XXZ Hamiltonian has U(1) and time reversal symmetry, whose symmetry group can be written as
(26) |
where and represent U(1) group and time reversal operator, respectively.
The representation of the non-unitary group is denoted as co-representation. The degeneracy of Hamiltonian energy level is determined by the reducibility of . In appendix. C, we show that when , is reducible, and the corresponding energy level will not be doubled; when , is irreducible, and the corresponding energy level will be doubled.
Decomposition of representation | |
2 | |
3 | |
4 | |
5 | |
6 | |
7 |
TABLE. 4 shows the decomposition of the symmetry group of the XXZ Hamiltonian into a series of IRs, whose degeneracy is dependent on the value of .This means that for a given eigenstate of the XXZ Hamiltonian, the number of linearly independent eigenstates (LIE) is determined by the value of . Simulations of HL are performed on , the number of LIE as a function of for chain lengths ranging from 2 to 7 are presented in Table 5.
0 | N | ||||||||
2 | 0 | / | 0 | / | / | / | / | / | 2 |
3 | / | 2 | / | 0 | / | / | / | / | 4 |
4 | 2 | / | 3 | / | 0 | / | / | / | 6 |
5 | / | 7 | / | 4 | / | 0 | / | / | 8 |
6 | 9 | / | 9 | / | 5 | / | 0 | / | 10 |
7 | / | 11 | / | 11 | / | 6 | / | 0 | 12 |
When the degeneracy of the IR is smaller than the unknowns , the number of LIE equals . In this case, the Hamiltonian cannot be fully recovered. When is equal to or larger than , equals to , it succeed to recover Hamiltonian. To sum up, the number of LIE is determined by the degeneracy of IR of the corresponding eigenstate or the number of unknowns for the first and the second situation:
(27) |
IV conclusions
In this paper, we investigate the problem of when a local Hamiltonian can be uniquely recovered from its eigenstate, which extends the previous work to the realm of symmetric Hamiltonian. Our work shows that the number of LIEs derived from an eigenstate corresponds to the degeneracy of IR of Hamiltonian symmetry group. An open question remains regarding the potential extension of our results to a broader range of Hamiltonians, such as those featuring all-to-all two-body interactions, or those of indistinguishable particles. Additionally, we ask whether our findings are applicable to systems possessing spatial symmetries, such as crystallographic point groups that are intricately linked to the symmetries of crystals. Addressing these inquiries is pivotal for the practical implementation of HL in the exploration of novel quantum materials.
V acknowledgement
This work is supported by National Key Research and Development Program of China (Grant No. 2021YFA0718302 and No. 2021YFA1402104), National Natural Science Foundation of China (Grants No. 12075310).
References
- Blatt and Roos [2011] R. Blatt and C. Roos, Nature Physics 8, 277 (2011).
- Webb et al. [2018] A. E. Webb, S. C. Webster, S. Collingbourne, D. Bretaud, A. M. Lawrence, S. Weidt, F. Mintert, and W. K. Hensinger, Phys. Rev. Lett. 121, 180501 (2018).
- Ballance et al. [2016] C. J. Ballance, T. P. Harty, N. M. Linke, M. A. Sepiol, and D. M. Lucas, Phys. Rev. Lett. 117, 060504 (2016).
- Cirac and Zoller [1995] J. I. Cirac and P. Zoller, Phys. Rev. Lett. 74, 4091 (1995).
- Barends et al. [2013] R. Barends, J. Kelly, A. Megrant, D. Sank, E. Jeffrey, Y. Chen, Y. Yin, B. Chiaro, J. Mutus, C. Neill, P. O’Malley, P. Roushan, J. Wenner, T. C. White, A. N. Cleland, and J. M. Martinis, Phys. Rev. Lett. 111, 080502 (2013).
- Barends et al. [2014] R. Barends, J. Kelly, A. Megrant, A. Veitia, D. Sank, E. Jeffrey, T. C. White, J. Mutus, A. G. Fowler, B. Campbell, Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, C. Neill, P. O’Malley, P. Roushan, A. Vainsencher, J. Wenner, A. N. Korotkov, A. N. Cleland, and J. M. Martinis, Nature 508, 500 (2014).
- Pastori et al. [2022] L. Pastori, T. Olsacher, C. Kokail, and P. Zoller, PRX Quantum 3, 030324 (2022).
- Ben Av et al. [2020] E. Ben Av, Y. Shapira, N. Akerman, and R. Ozeri, Phys. Rev. A 101, 062305 (2020).
- Wang et al. [2017] J. Wang, S. Paesani, R. Santagati, S. Knauer, A. A. Gentile, N. Wiebe, M. Petruzzella, J. L. O’brien, J. G. Rarity, A. Laing, et al., Nature Physics 13, 551 (2017).
- Motta et al. [2020] M. Motta, C. Sun, A. T. Tan, M. J. O’Rourke, E. Ye, A. J. Minnich, F. G. Brandão, and G. K.-L. Chan, Nature Physics 16, 205 (2020).
- Wiebe et al. [2014] N. Wiebe, C. Granade, C. Ferrie, and D. G. Cory, Physical review letters 112, 190501 (2014).
- Holzäpfel et al. [2015] M. Holzäpfel, T. Baumgratz, M. Cramer, and M. B. Plenio, Phys. Rev. A 91, 042129 (2015).
- Zhu et al. [2019] W. Zhu, Z. Huang, and Y.-C. He, Phys. Rev. B 99, 235109 (2019).
- Huang et al. [2023] H.-Y. Huang, Y. Tong, D. Fang, and Y. Su, Phys. Rev. Lett. 130, 200403 (2023).
- Anshu et al. [2021] A. Anshu, S. Arunachalam, T. Kuwahara, and M. Soleimanifar, Nature Physics 17, 931 (2021).
- Shabani et al. [2011] A. Shabani, M. Mohseni, S. Lloyd, R. L. Kosut, and H. Rabitz, Physical Review A 84, 012107 (2011).
- Burgarth et al. [2009] D. Burgarth, K. Maruyama, and F. Nori, Phys. Rev. A 79, 020305 (2009).
- Rath et al. [2021] A. Rath, R. van Bijnen, A. Elben, P. Zoller, and B. Vermersch, Phys. Rev. Lett. 127, 200503 (2021).
- Anshu et al. [2020] A. Anshu, S. Arunachalam, T. Kuwahara, and M. Soleimanifar, in 2020 IEEE 61st Annual Symposium on Foundations of Computer Science (FOCS) (2020) pp. 685–691.
- Qi and Ranard [2019] X.-L. Qi and D. Ranard, Quantum 3, 159 (2019).
- Chertkov and Clark [2018] E. Chertkov and B. K. Clark, Phys. Rev. X 8, 031029 (2018).
- Bairey et al. [2019] E. Bairey, I. Arad, and N. H. Lindner, Phys. Rev. Lett. 122, 020504 (2019).
- Hou et al. [2020] S.-Y. Hou, N. Cao, S. Lu, Y. Shen, Y.-T. Poon, and B. Zeng, New Journal of Physics 22, 083088 (2020).
- Zhou and Zhou [2022] J. Zhou and D. L. Zhou, Phys. Rev. A 105, 012615 (2022).
- Zhou and Zhou [2024] J. Zhou and D. Zhou, Physics Letters A 493, 129279 (2024).
- Gross [1996] D. Gross, Proceedings of the National Academy of Sciences 93, 14256 (1996), https://www.pnas.org/doi/pdf/10.1073/pnas.93.25.14256 .
- LaBorde and Wilde [2022] M. L. LaBorde and M. M. Wilde, Phys. Rev. Lett. 129, 160503 (2022).
- Joel et al. [2013] K. Joel, D. Kollmar, and L. F. Santos, AMERICAN JOURNAL OF PHYSICS 81, 450 (2013).
- Kosmann-Schwarzbach et al. [2011] Y. Kosmann-Schwarzbach, B. E. Schwarzbach, and Y. Kosmann-Schwarzbach, The Noether Theorems (Springer, 2011).
- Xu and Ka [2016] W. Xu and X. Ka, Group theory and its applications in solid state physics [in Chinese] (Higher Education Press, 2016).
- Burns [1977] G. Burns, in Introduction to Group Theory with Applications, Materials Science and Technology, edited by G. Burns (Academic Press, 1977) pp. 125–155.
- Cornwell [1997] J. F. Cornwell, Group theory in physics: An introduction (Academic press, 1997) pp. 65–91.
- Zee [2016] A. Zee, Group theory in a nutshell for physicists (Princeton University Press, 2016).
- Jacobs [2005] P. W. Jacobs, Group theory with applications in chemical physics (Cambridge University Press, 2005) pp. 252–264.
- El-Batanouny and Wooten [2008] M. El-Batanouny and F. Wooten, Symmetry and condensed matter physics: a computational approach (Cambridge University Press, 2008) pp. 452–453.
Appendix A Co-representation of group containing anti-unitary elements
Considering a non-unitary group that composes half of the anti-unitary elements
(28) |
where is a unitary subgroup and is an anti-unitary operator. The representation of such group is called co-representation. The co-representation can be educed from an IR of unitary subgroup
(29a) |
(29b) |
This co-representation is denoted as , whose representation matrix is unitary. Since is educed from IR , its reducibility is directly related to and can be divided into three conditions:
-
(1)
If , and , then is reducible and the degeneracy will not be doubled,
-
(2)
If , and , then is irreducible and the degeneracy will be doubled,
-
(3)
If two IR and of subgroup are inequivalent, then is irreducible and the degeneracy will be doubled.
Character of representation can be utilized to distinguish these three cases. Supposing that is the character of representation , then
(30) |
where is the rank of subgroup . Through co-representation of Hamiltonian symmetry group, we can educe the irreducible subspace of the system’s Hilbert space and thus the energy level degeneracy.
Appendix B The XXX spin chain
We will first show that the XXX Hamiltonian is invariant under the time reversal transformation. The time reversal is represented by an antiunitary operator
(31) |
where is the complex conjugate operator and is an unitary operator. The spin angular momentum is transformed in the way
(32) |
We take the definition of for a single spin as
(33) |
For a system contains spins, we choose
(34) |
It can be verified that is invariant under the rime reversal transformation
(35) |
We now show that the XXX Hamiltonian has SU(2) symmetry by confirming that it commutes with the total spin operator
(36) |
It can be checked that
(37) |
The spin operators form a SU(2) algebra and consequently the XXX spin chain has SU(2) as a symmetry group.
Each Lie group has a corresponding Lie algebra, the elements of the group can be constructed using the exponential mapping from the algebra elements. Therefore the elements of the SU(2) group can be written as
(38) |
which describes the state transformation corresponding to the rotation with angle around the axis .
So far, we have shown that the XXX Hamiltonian has time reversal and SU(2) symmetry. Thus its symmetry group is composed of
(39) |
where is the time reversal operator, and . Note that is a set of anti-unitary operators. Time reversal symmetry may cause the doubling of energy degeneracy, which is distinguished by Eq. (30). We first calculate the left-hand-side of Eq. (30),
(40) | ||||
In a finite group, the group function averages over the group elements. When this group function is extended to a Lie group, it becomes an integral of the group function over the group elements. Eq. (40) becomes
(41) |
As is known that the IRs of SU(2), are labeled by , where , whose dimension is . The characters of the representation of an arbitrary group element is
(42) |
We have , such that
(43) |
The integration over SU(2) is thus [35]
(44) | ||||
Now it’s time to figure the rank of subgroup SU(2). The rank of the Lie algebra is known as the number of generators that can be simultaneously diagonalized [33]. The rank of SU(2) is thus 1: we can only diagonalize . We can now conclude that
(45) |
which means that the co-representation is reducible and the degeneracy will not be doubled. Therefore the dimension of IR of and group is the same.
Appendix C the XXZ spin chain
Noting that commutes with total -spin operator
(46) |
Additionally, the XXZ Hamiltonian is invariant under time reversal ,
(47) |
Therefore, system of XXZ Hamiltonian possesses U(1) symmetry associated with the conservation of total -spin, as well as time reversal symmetry.
The group consists of all the one dimensional unitary operator
(48) |
The U(1) group is a compact abelian group, and its elements commute with each other. According to Schur’s theorem, all the IRs of a compact abelian group are one-dimensional. Therefore, the IR of U(1) symmetry is one-dimensional. Geometrically, the elements of the U(1) group can be visualized as points on a unit circle in the complex plane. In the context of the spin chain system, the U(1) symmetry can be expressed as
(49) |
where is the continuous parameter, and is the total -spin operator.
Thus the symmetry group of the XXZ Hamiltonian can be written as
(50) |
where represents U(1) group, represents time reversal operator. Similar to the XXX Hamiltonian, co-representation can be utilized to characterize the representation of group . Next, we will investigate the reducibility of co-representation of group .
Recalling that there are three cases of criterion for reducibility of co-representation . To determine which case it belongs, we first calculate :
(51) |
The co-representation is deduced from IR of subgroup , such that the reducibility of is decided by the properties of . We first calculate and in the context of -site spin chain:
(52a) | |||
(52b) |
Noting that the equivalence between and depends on the value of . Therefore, we partition Hilbert space into subspaces by the value of . This immediately leading to two cases:
-
1.
When (only exist for is even), , which is a 1-dim identity representation. The transformation operator is , with . As a result, is reducible, and the degeneracy of the corresponding energy level will not be doubled.
-
2.
When (exist for is even or odd). is not equivalent to . Hence is irreducible, and the degeneracy of the corresponding energy level will be doubled.
Through the discussion above, we have discovered that the time reversal symmetry connects certain subspaces with opposite values of . In the absence of time reversal symmetry, when only the U(1) symmetry is present, all IRs are one-dimensional. However, with the introduction of time reversal symmetry, some of the subspaces become connected, forming two-dimensional IRs. In other words, when the time reversal operator acts on a vector in a subspace with (where ), it results in a vector in the subspace with . On the other hand, when acts on a vector in the subspace with , it leaves the vector unchanged within the same subspace.

