<<

Permutation Representation Quantum Monte Carlo

Lalit Gupta,1 Tameem Albash,2 and Itay Hen1, 3, ∗ 1Department of Physics and Astronomy and Center for Quantum Information Science & Technology, University of Southern California, Los Angeles, California 90089, USA 2Department of Electrical and Computer Engineering, Department of Physics and Astronomy, and Center for Quantum Information and Control, CQuIC, University of New Mexico, Albuquerque, New Mexico 87131, USA 3Information Sciences Institute, University of Southern California, Marina del Rey, California 90292, USA (Dated: August 5, 2020) We present a quantum Monte Carlo algorithm for the simulation of general quantum and classical many-body models within a single unifying framework. The algorithm builds on a power series expansion of the quantum partition function in its off-diagonal terms and is both parameter-free and Trotter error-free. In our approach, the quantum dimension consists of products of elements of a permutation group. As such, it allows for the study of a very wide variety of models on an equal footing. To demonstrate the utility of our technique, we use it to clarify the emergence of the sign problem in the simulations of non-stoquastic physical models. We showcase the flexibility of our algorithm and the advantages it offers over existing state-of-the-art by simulating transverse- field Ising model Hamiltonians and comparing the performance of our technique against that of the stochastic series expansion algorithm. We also study a transverse-field Ising model augmented with randomly chosen two-body transverse-field interactions.

I. INTRODUCTION In this paper, we provide a QMC scheme that has the flexibility to simulate a broad range of quantum Quantum Monte Carlo (QMC) algorithms [1,2] many-body models. The technique we propose here are extremely useful for studying equilibrium prop- builds on a power-series expansion of the canoni- erties of large quantum many-body systems, with cal quantum partition function about the classical applications ranging from superconductivity and partition function first introduced in Refs. [13, 14]). novel quantum materials [3–5] through the physics While strongly inspired by earlier methods that ex- of neutron stars [6] and quantum chromodynam- pand the partition function in powers of the inverse- ics [7,8]. The algorithmic development of QMC re- temperature [9, 15–17], our expansion is more ac- mains an active area of research, with the dual goal curately described as an expansion in off-diagonal of extending the scope of QMC applicability and operators of the Hamiltonian. improving convergence rates of existing algorithms Our formalism enables a very general treatment in order to facilitate the discovery of new phenom- of Hamiltonians, allowing us to develop a QMC ena [9–11]. scheme that is applicable to a wide variety of mod- While QMC algorithms have been adapted to els, ranging from highly interacting models with the simulation of a wide variety of physical sys- multi-body terms to non-interacting ones and from tems, different models typically require the devel- strongly quantum models to purely classical ones, opment of distinct model-specific update rules and using the same updating formalism. In order to measurement schemes. A notable recent example is demonstrate the advantage of the new method, we the transverse-field Ising model (TFIM), which tra- study the TFIM with random two-body longitudi- ditionally includes only single-body X terms (the nal interactions, comparing the performance of our transverse field), supplemented with two-body X technique against that of stochastic series expansion terms. While the updates associated with single- QMC [18]. In addition, we study in detail a vari- body X terms can be implemented by local (in ant of the model with added random XX interac- space) updates, the inherently non-local nature of tions. This model poses a challenge for traditional arXiv:1908.03740v2 [cond-mat.stat-mech] 4 Aug 2020 the two-body X terms requires novel cluster up- QMC approaches [12, 19] due to the random connec- dates [12]. Thus, a proper treatment of the Hamil- tivity of the Ising couplings and XX interactions, tonian with single-body and two-body X terms re- which vary between instances. While we focus most quires updates that are different than if the Hamil- of our attention in what follows on finite-dimensional tonian included only single-body or only two-body Hamiltonians, the technique we present here should X terms. apply with equal rigor to infinite-dimensional sys- tems. The paper is organized as follows. In Sec.II, we describe the ‘ representation’ ∗ [email protected] 2

(PMR) of Hamiltonians on which the partition func- B). This in turn implies that any Hamiltonian H can tion expansion detailed in Sec.III is founded. We be written as discuss the emergence of the sign problem within X iΦj −iΦj −1 the formulation and its sometimes-intricate relation H = Rj e Pj + e Pj , (2) with the concept of non-stoquasticity in Sec.IV and j in Sec.V we present the QMC algorithm we have devised based on the expansion. Results of TFIM where Rj, Φj are real-valued diagonal matrices. In simulations are presented in Sec.VI. We conclude the case where a permutation matrix Pj is its own in Sec.VII with additional discussions and some inverse, the corresponding Φj will necessarily be the caveats. . To further elucidate PMR, we now provide several examples. II. THE PERMUTATION MATRIX REPRESENTATION A. Example I: A single spin-1/2 particle We consider many-body systems whose Hamilto- nians we cast as the sum The Hamiltonian of a single spin-1/2 particle can most generally be written as M M M X ˜ X X H = Pj = DjPj = Hc + DjPj , (1) H = α01 + α1X + α2Y + α3Z, (3) j=0 j=0 j=1 where X,Y and Z are the matrix representations of where {P˜j} is a set of M + 1 distinct generalized the usual Pauli operators in the basis that diagonal- permutation matrices [20], i.e., matrices with pre- izes the Pauli-Z operator. In PMR, the Hamiltonian cisely one nonzero element in each row and each col- becomes umn (this condition can be relaxed to allow for rows and columns with only zero elements). Each opera- H = D0P0 + D1P1 (4) tor P˜j can be written, without loss of generality, as with P = , P = X, D = H = α + α Z and P˜ = D P where D is a diagonal matrix1 and P is 0 1 1 0 c 01 3 j j j j j D = α − iα Z. a permutation matrix with no fixed points (equiva- 1 11 2 lently, no nonzero diagonal elements) except for the P = 1. We will refer to the ba- 0 B. Example II: Two-local spin-1/2 models sis in which the operators {Dj} are diagonal as the computational basis and denote its states by {|zi}. A general two-local n-particle spin-1/2 Hamilto- We will call the D0 the ‘classical nian has similarly the following form Hamiltonian’ and will sometimes denote it by Hc. The permutation matrices appearing in H will be X X treated as a subset of a permutation group, wherein H = αij,Ki,Kj KiKj . (5) i

1 The diagonal matrix Dj will be invertible, i.e., will not contain zero elements along the diagonal, if P˜j is a bonafide generalized permutation matrix. 3

C. Example III: Spin-one particles (qutrits) We begin by replacing the operation Tr[·] P with the explicit sum zhz|·|zi and then expanding The permutation matrix representation general- the exponent in the partition function in a Taylor izes straightforwardly to higher dimensional sys- series in β: tems. The Hamiltonian for a single qutrit can be ∞ n X X β n written as H = D0P0 + D1P1 + D2P2 where Z = hz|(−H) |zi (7) n!       z n=0 1 0 0 0 0 1 0 1 0 n ∞   P = 1 = 0 1 0 ,P = 1 0 0 ,P = 0 0 1 , X X βn X 0   1   2   = hz| −H − D P |zi 0 0 1 0 1 0 1 0 0 n!  c j j z n=0 j=1 ∗ ∞ n and D2 = D1, a condition imposed by the hermitic- X X X β = hz|S |zi . ity of the Hamiltonian. n! in z n=0 {Sin }

n D. Example IV: The Bose-Hubbard model In the last step we have expressed (−H) in terms of all sequences of length n composed of products of basic operators H and D P , which we have de- Another model that can just as easily be rep- c j j noted by the set {Sin }. Here in = (i1, i2, . . . , in) is a resented in permutation matrix form is the Bose- set of indices, each of which runs from 0 to M, that Hubbard model. This discretely infinite dimensional denotes which of the M + 1 operators in H appear model captures the physics of interacting spinless in Sin . bosons on a lattice [21] and is commonly used to de- We proceed by stripping away all the diago- scribe superfluid-insulator transitions [22], bosonic nal Hamiltonian terms from the sequence hz|Sin |zi. atoms in an optical lattice [23] and certain magnetic We do so by evaluating the action of these terms insulators [24]. on the relevant basis states, leaving only the off- The Bose-Hubbard Hamiltonian is given by diagonal operators unevaluated inside the sequence (see Refs. [13, 14] for a more detailed derivation). X ˆ†ˆ U X X H = −t b bj + nˆi (ˆni − 1) − µ nˆi . (6) The partition function may then be written as i 2 hi,ji i i ∞  q  ∞ βn(−1)n X X X Y (ij ) X Here, hi, ji denotes summation over all neighbor- Z =  dz  hz|Siq |zi j n! ˆ† ˆ z q=0 {S } j=1 n=q ing lattice sites i and j, while bi and bi are regu- q lar bosonic creation and annihilation operators such  † X thatn ˆ = ˆb ˆb gives the number of particles at the k0 kq i i i × (Ez0 ) · ... · (Ezq )  , (8) P i-th site. The model is parametrized by the hopping ki=n−q amplitude t and the on-site interaction U.

In the bosonic number basis where states are where Ezi = hzi|Hc|zii and {Siq } denotes the set of described by the number of bosons in each all products of length q of ‘bare’ off-diagonal opera- site |n1i ... |nLi (here L is the number of lat- tors Pj. Also tice sites) we identify the diagonal part to U P P d(ij ) = hz |D |z i , (9) be D0 = 2 i nˆi (ˆni − 1) − µ i nˆi and the off- zj j ij j diagonal (infinite dimensional) permutation opera- ˆ†ˆ which can be considered as the ‘hopping strength’ tors as Phi,ji = bi bj. The diagonal operators asso- of Pi with respect to |zji. Note that while the par- ciated with Phi,ji are Dhi,ji whose entries are −t for j (ij ) states whose nj is positive (the j-th boson can be tition function is positive and real-valued, the dzj annihilated) and zero otherwise. elements do not necessarily have to be so. The term in parentheses in Eq. (8) sums over the

diagonal contribution of all hz|Sin |zi terms that cor-

III. OFF-DIAGONAL PARTITION respond to the same hz|Siq |zi term. The various FUNCTION EXPANSION {|zii} states are the states obtained from the action

of the ordered Pj operators in the product Siq on We are now in a position to discuss the off- |z0i, then on |z1i, and so forth. For example, for

diagonal series expansion of the partition function Siq = Piq ...Pi2 Pi1 , we obtain |z0i = |zi,Pi1 |z0i =  −βH  Z = Tr e as it applies to Hamiltonians cast in |z1i,Pi2 |z1i = |z2i, etc. The proper indexing of

the form given in Eq. (1). the states |zji along the path is |z(i1,i2,...,ij )i to in- dicate that the state at the j-th step depends on 4

all Pi1 ...Pij . We will use the shorthand |zji. The where we have denoted sequence of basis states {|zii} may be viewed as a q ‘walk’ on the hypercube of basis states [13, 14, 25] Y (ij ) D(z,S ) = d . (16) (see Fig.1). iq zj j=1 After a change of variables, n → n + q, we arrive at: We stress that the partition function expanded as such is not an expansion in β. Similar to the tech- ∞   q  X X X q Y (ij ) niques introduced by Handcomb in the 1960s [15, 16] Z = hz|Si |zi (−β) d q   zj  and further developed in the stoquastic series expan- z q=0 {Sq } j=1 sion (SSE) scheme pioneered by Sandvik [9, 17], the  ∞ (−β)n off-diagonal series expansion begins with a Taylor X X k0 kq × (Ez0 ) ··· (Ezq )  (10). series expansion of the exponential function in the (n + q)! P n=0 ki=n inverse temperature β, but the regrouping of terms into the exponent of divided-differences means that Noting that the various {Ezi } are the classical ener- it is no longer a high-temperature expansion. SSE gies of the states |zii states created by the operator writes the trace of products of the Hamiltonian as a product Siq , the partition function is now given by: sum of products of matrix elements written in a suit- ably chosen basis which are then sampled (thereby ∞  q  X X Y (ij ) X overcoming the limitation of Handscomb’s scheme Z = d hz|Si |zi (11)  zj  q which required the evaluation of traces of products z q=0 j=1 {Sq } of the Hamiltonian). This is made possible by break-  (∞,...,∞) q  (−β)q ing up the Hamiltonian into a sum of local bonds. In X Y kj ×  P (−βEzj )  . PMR, this is not the case — all terms may remain (q + ki)! {ki}=(0,...,0) j=0 non-local but are grouped according to their action on basis states. A feature of the above infinite sum is that the term Specially, the diagonal portion of the Hamiltonian in parentheses can be further simplified to give the remains ‘intact’ which then allows us to regroup a exponent of divided differences of the E ’s (a short zi large portion of all terms. A single divided-difference description of divided differences and an accompa- nying proof of the above assertion can be found in Ref. [13]); it can therefore be succinctly rewritten as: q |�⟩ � X (−β)q Y kj −β[Ez0 ,...,Ezq ] P (−βEzj ) = e (q + ki)! {ki} j=0 (12),

� where [Ez0 ,...,Ezq ] is a multiset of energies and where a function F [·] of a multiset of input values is defined by q |� ⟩ X F (Ezj ) F [E ,...,E ] ≡ (13) z0 zq Q (E − E ) j=0 k6=j zj zk and is called the divided differences [26, 27] of the |�⟩ � function F [·] with respect to the list of real-valued () () () [ , , , ] input variables [E ,...,E ]. In our case, F [·] is � = � � � � z0 zq the function

−β[Ez ,...,Ezq ] Figure 1. Diagrammatic representation of a generalized F [Ez ,...,Ez ] = e 0 . (14) 0 q Boltzmann weight, or a GBW, calculated from the classi-

Therefore, the infinite sum over energies in Eq. (11) cal energies Ezj of the classical states |zj i, which form a may be simplified to closed walk on the hypercube of basis states. The walk is determined by the action of the permutation operators ∞ of the configuration, represented by Siq = P3P2P1, on X X X −β[E ,...,E ] Z = hz|S |ziD e z0 zq , the initial basis state |z0i. The walk closes if and only if iq (z,Siq ) the sequence of permutation operators evaluates to the z q=0 {Sq } (15) identity operation. 5

¯ term thus corresponds to a sum of an infinite number gate weight WC¯ = WC, the imaginary contribu- of SSE terms, meaning that a single PMR configura- tions cancel out. Expressed differently, the imag- tion represents very many standard SSE configura- inary portions of complex-valued weights do not tions and a single PMR weight sums up very many contribute to the partition function and may be standard SSE weights. This can be immediately seen disregarded altogether. We may therefore redefine h i in the derivation above, particularly Eq. (8), which Qq (ij ) D(z,S ) = Re dzj , obtaining strictly real- relates the standard SSE weight, which involves se- iq j=1 valued weights. quences of diagonal as well as off-diagonal bonds, to Before we move on, we note that hz|S |zi eval- the weights of the current approach that only involve iq uates either to 1 or to zero. Moreover, since the off-diagonal bonds. permutation matrices with the exception of P have It is worth noting that the cost of the massive 0 no fixed points, the condition hz|S |zi = 1 implies grouping of the off-diagonal series expansion is man- iq S = 1, i.e., S must evaluate to the identity el- ifested in the computational cost associated with iq iq ement P (note that the identity element does not calculating PMR terms (or ratios thereof). As we 0 appear in the sequences S ). The expansion can discuss in Sec.V in more detail, these can be cal- iq thus be more succinctly rewritten as culated (or more precisely, updated) using O(q) ba-

sic arithmetic operations (where q is the number of X X −β[Ez ,...,Ez ] Z = D(z,S )e 0 q . (19) operators in a sequence or equivalently the size of iq z S =1 the imaginary-time dimension), as opposed to SSE’s iq O(1). We refer the reader to Ref. [13] for a more de- tailed comparison between the off-diagonal expan- IV. NON-STOQUASTICITY AND sion and SSE. In Sec.VIA we provide a runtime EMERGENCE OF THE SIGN PROBLEM comparison of PMR vs SSE for simulations of TFIM instances. An attractive property of the formalism intro- Aside from SSE, it is also interesting to ob- duced above is that it allows us to identify the emer- serve that the exponent of divided differences also gence of the sign problem in QMC via inspection of has close relations to continuous-time QMC (e.g., the weights W , thereby making more apparent the Ref. [11]), via the Hermite-Genocchi formula [27]: C connection between the notion of non-stoquasticity Z −β[E0,...,Eq ] −β(E0t0+E1t1+...+Eq tq ) — the existence of positive or complex-valued off- e = dt0 ··· dtqe , Ω diagonal Hamiltonian matrix entries — which has (17) garnered increasing attention with the advent of quantum computers in recent years [28–30] and the where ti ≥ 0 and the area of integration Ω is onset of the sign problem. bounded by t0 + t1 + ... + tq from above. To interpret the real-valued weight terms WC as Having derived the expansion Eq. (15) for any actual weights (equivalently, un-normalized proba- Hamiltonian cast in the form Eq. (1), we are now bilities), they must be nonnegative. The occurrence in a position to interpret the partition function ex- P of negative weights marks the onset of the infamous pansion as a sum of weights, i.e., Z = {C} WC, sign problem. A weight is positive iff where the set of configurations {C} is all the distinct  q  pairs {|zi,Siq }. Because of the form of WC, q Y (ij ) (−1) D(z,S ) = Re (−d ) −β[E ,...,E ] iq  zj  W = D e z0 zq , (18) C (z,Siq ) j=1 we refer to it as a ‘generalized Boltzmann is positive, that is, a QMC algorithm will encounter weight’ (or, a GBW). It can be shown [13] that a sign problem, equivalently a negative weight, dur- q −β[E ,...,E ] (−1) e z0 zq is strictly positive. Another fea- ing a simulation if and only if there exists a closed ture of divided differences is that they are invariant walk on the hypercube of basis states along which h i under rearrangement of the input values. Qq (ij ) Re j=1(−dzj ) < 0. It is thus clear that it is We note that as written, the weights WC are complex-valued, despite the partition function be- not mere non-stoquasticity (equivalently, the sign of ing real (and positive). Since for every configu- off-diagonal entries) that creates the sign problem, but rather the sign of closed walks on the hypercube ration C = {|zi,Si } there is a conjugate config- q of basis states that determines its occurrence. uration C¯ = {|zi,S† }2 that produces the conju- iq

2 S† = P −1P −1 ...P −1. For Siq = Piq ...Pi2 Pi1 , the conjugate sequence is simply iq i1 i2 iq 6

A special class of models where the sign problem a product S of permutation operators that must h i iq Qq (ij ) evaluate to the identity element P = . To take full does not emerge, i.e., where Re (−dz ) ≥ 0 0 1 j=1 j advantage of our partition function decomposition for all configurations, is that of ‘stoquastic’ Hamilto- above, we treat the off-diagonal permutation terms (ij ) nians [28, 29] for which all dz are negative, which j {Pj} in our QMC algorithm as elements in a permu- is equivalent to having only nonpositive off-diagonal tation group G (with matrix product as the group elements in the matrix representation of the Hamil- operation). Since the elements {Pj} appearing in tonian. In this case, all products trivially yield the Hamiltonian may not form a complete group, positive-valued walks. we shall treat any additional element Pj0 required The existence of positive off-diagonal terms does to complete the set to form a group as appearing in not however immediately imply a sign problem for the Hamiltonian with an associated diagonal matrix QMC (the reader is referred to Ref. [31] for a more Dj0 = 0 [see Eq. (1)]. detailed discussion). Another example of a sign- The pair C induces a list of states (ij ) problem-free family of models is one where all dzj {|z0i = |zi, |z1i,..., |zqi = |zi}, which in turn also elements are positive but closed walks are all of even generates a corresponding multiset of diagonal en- length. One such model is the transverse-field Ising ergies EC = {Ez0 ,Ez1 ,...,Ezq } of not-necessarily-

Hamiltonian distinct values (recall that Ezi = hzi|Hc|zii). For X X X systems with discrete energy values, the multiset H = JijZiZj + hjZj + Γ Xj . (20) can be stored efficiently in a ‘multiplicity table’ i,j j j MC = {m0, m1, . . . , mj,...}, where mj is the

multiplicity of the energy Ezj in the multiset. for Γ > 0. A slightly less trivial example is the two- Given the multiset E , the evaluation of the GBW body model C WC follows from its definition as a function of di- X X vided differences (the reader is referred to Ref. [13] H = JijZiZj + Γ XiXj , (21) for a more detailed description). The calculation of i,j hi,ji a GBW consisting of q permutation operators re- quires the evaluation of a divided-differences expo- provided that the underlying connectivity hi, ji of nential with (q+1) energies. This calculation can be the two-body X terms is bi-partite (allowing only accomplished with at most O(q) operations [32, 33]. even cycles). It is also interesting to note that any single-qubit Hamiltonian is necessarily also sign-problem-free. In B. Initial state this case, the Hamiltonian is H = D0P0 + D1P1 as described in Sec.IIA. Since here the Siq are se- quences consisting of only one type of non-identity At this point we can consider a QMC algorithm based on the partition function expansion gener- permutation matrices, namely P1 = X, the ex- ating the weights WC, Eq. (18). The Markov pansion order q must be even for Siq to evaluate to the identity element. This in turn results in process would start with the initial configuration h i C0 = {|zi,S0 = 1} where |zi is a randomly generated Qq (ij ) 2 2 q/2 j=1(−dzj ) = (α1 + α2) being strictly non- initial classical state. The weight of this initial con- negative. The same is however not true for a single figuration is qutrit in which case a sign problem may arise. −β[Ez ] −βEz WC0 = e = e , (22)

V. THE QMC ALGORITHM i.e., the classical Boltzmann weight of the initial ran- dom state |zi. Having derived the series expansion of the parti- tion function for permutation-represented Hamilto- nians, we are now in a position to discuss a QMC C. Updates algorithm that can be associated with the above ex- pansion. We next describe the basic update moves for the algorithm. These are also succinctly summarized in Fig.2. A. QMC configurations and GBW calculation

As was discussed above, a configuration

C = {|zi,Siq } is a pair of a classical state and 7

(a) � � � ⟨�| !! … !" !# |�⟩ ⟨�′| �!! … �!" �!# |�′⟩ (b) � … � � … � � … � � � … � |�⟩ ⟨�| !! !$%" !$%# �!$ !# |�⟩ ⟨�| !! !$ !$%" !$%# !# (c) �’ �’’ �′′′′ �′′′ � … … � � … � � …� ⟨�| !! �!$%# �!$ !# |�⟩ ⟨�′| !$ !# !! !$%#|�′⟩ � (d) �′ � … � … � � … � … � ⟨�| !! !$%# �!$ !# |�⟩ ⟨�| !! !$ !# |�⟩

Figure 2. Basic update moves of the QMC algorithm. (a) Classical moves (e.g., a single bit flip), whereby only 0 the initial state z is changed to z leaving Siq unchanged. (b) Cyclic rotation, whereby two adjacent sequences of

group elements (in this case Pik and Pik+1 Pik+2 ) whose product is the identity operation are interchanged, changing

their internal classical states. (c) Block swap, whereby two partitions of the sequence Siq are interchanged. This also 0 changes the initial state from z to z as well as the ordering of Siq . (d) Cycle completion, whereby a sub-sequence of ˜ operators is replaced by an equivalent one (in this case, Pik Pik+1 is replaced by Pik . This is the only update where the number of group element (equivalently, the expansion order of the configuration) may change.

1. Classical moves reduces to a classical thermal algorithm keeping the size of the imaginary-time dimension at zero (q = 0) Classical moves are any moves that involve a ma- for the duration of the simulation.

nipulation of the classical state |zi while leaving Siq unchanged [see Fig.2(a)]. In a single bit-flip clas- sical move, a spin from the classical bit-string state 2. Cyclic rotations |zi of C is picked randomly and is flipped, generating a state |z0i and hence a new configuration C0. Cal- The ‘cyclic rotation’ move, [Fig.2(b)], consists of culating the weight of C0 requires recalculating the identifying short sub-sequences, or cycles, of consec- utive operators in the sequence S , whose product energies associated with the product Siq leading to a iq new energy multiset EC0 and can become computa- is the identity element, i.e., sub-sequences that obey tionally intensive if q is large. Classical moves should therefore be attempted with low probabilities if q is Pij ··· Pij+C = 1 . (24) large. Simply enough, the acceptance probability for a classical move is Depending on the nature of the operators, preparing a lookup table of short cycles that evaluate to the    −β[E 0 ]  W 0 e C identity may prove useful. Once a cycle is identified, p = min 1, C = min 1, , (23) −β[E ] a random cycle rotation is attempted. Here, a ran- WC e C dom internal insertion point within the sub-sequence −β[E ] −β[E ,E ,...,E ] where e C is a shorthand for e z0 z1 zq is picked and a rotation is attempted: of configuration C and likewise for C0. P ··· P P ··· P → P ··· P P ··· P . In the absence of a quantum part to the Hamil- ij ik ik+1 ij+C ik+1 ij+C ij ik (25) tonian (Dj = 0 for all j > 0), not only are classical moves the only moves necessary, they are also the The rotated sequence also evaluates to the identity. only moves that have nonzero acceptance probabili- Since the internal classical states between the ele- ties. Since the initial configuration of the QMC al- ments in the cycle may change by the rotation, the gorithm is a random classical configuration |zi and rotation involves adding new energies {Ez0 ...} and removing old ones {Ez0 ...} − {Ez ...} from the en- an empty operator sequence S0 = 1, for a purely classical Hamiltonian, the algorithm automatically ergy multiset. Short cycles should therefore be pre- 8

ferred. The acceptance probability for the move is by its complement, resulting in a new configuration 0 −1 −1 as in Eq. (23) with EC0 = EC + {Ez0 ...} − {Ez ...}. C . Because we can interpret P0 = P0 = PjPj (for any arbitrary index j) and so on, the cycle com- pletion move can grow and shrink the sequence. The 3. Block-swap acceptance probability is as in Eq. (23) with the new configuration. A block swap [Fig.2(c)] is an update that in- volves a change of the classical state z. Here, a D. Measurements random position k in the product Siq is picked such that the product is split into two (non-empty) sub-

sequences, Siq = S2S1, with S1 = Pik ··· Pi1 and Having reviewed the various update moves we 0 S2 = Piq ··· Pik+1 . The classical state |z i at posi- next turn to discuss measurements within the al- tion k in the product is given by gorithm. 0 |z i = S1|zi = Pik ··· Pi1 |zi , (26) where |zi is the classical state of the current config- 1. Diagonal measurements 0 uration. The state |z i has energy Ez0 , and the state |zi has energy Ez. The new block-swapped config- A diagonal operator Λ obeys Λ|zi = λ(z)|zi 0 0 uration is C = {|z i,S1S2}. The multiplicity table where λ(z) is a number that depends both on of this configuration differs from that of the current the operator and the state it acts on. Since configuration by having one fewer Ez state and one hz|ΛSiq |zi = λ(z)hz|Siq |zi, for any given configura- additional Ez0 state. The weight of the new config- tion C = (|zi,Siq ), there is a contribution λ = λ(z) −β[E 0 ] uration is then proportional to e C where the to the diagonal operator thermal average hΛi. To multiset EC0 = EC + {Ez0 } − {Ez}. The acceptance improve statistics, one may also consider rotations probability is as in Eq. (23) with the aforementioned in (the periodic) imaginary time. To do that, we may EC0 . consider ‘virtual’ block-swap moves (see Sec.VC3)

that rotate Siq and as a result also change the classi- cal configuration from |zi to |zii. The contribution 4. Cycle completion to the expectation value of a diagonal operator Λ thus becomes: The moves presented so far have left the number of group elements in the sequence, or expansion order, q−1 1 X −β[E ] λ = λ(z )e Ci . (29) namely q, unchanged. The cycle completion move Z i has the effect of changing the value of q. A lookup i=0 table of short cycles obeying where ECi is the energy multiset associated with con- figuration C whose multiset is E = E + {E } − Pij ··· Pij+C = 1 (27) i Ci C zi

{Ez} (recall that z0 ≡ z, so EC0 = EC). The nor- will be helpful in this case. The cycle completion malization factor Z above is the sum move identifies a sub-cycle in the sequence Siq , e.g., q−1 Pi ··· Pi and replaces it with its complement j k X −β[E ] X −β[E ] Z = e Cj = m e Cj (30) −1 j P ··· P  = P −1 ··· P −1 . (28) ik+1 ij+C ij+C ik+1 j=0 j Note that the inverses of permutation matrices are over all nonzero multiplicities mj. In the case where also permutation matrices and are therefore also Λ = Hc the above expression simplifies to: present in G. For concreteness, let us consider the case of sub- q−1 1 X −β[E ] 1 X −β[E ] sequences of length two. We randomly pick a point λ = E e Ci = m E e Cj . Z zi Z j zj k ∈ [0, q] in the sequence. With probability 1/4, the i=0 j subsequence is taken to be P0P0, P0Pik , Pik−1 P0, (31) 3 Pik−1 Pik . The identified subsequence is replaced

2. Off-diagonal measurements 3 Note that if k = 0 or q, namely, the edges of the sequence, then two of the choices correspond to non-starters as there We next consider the case of measuring the expec- is an operator only to one side of the insertion point and tation value of an off-diagonal operator Pk, namely, so either Pik−1 or Pik are not defined. 9

hPki. To do this, we interpret the instantaneous con- doing so, hPk1 Pk2 i becomes figuration as follows −β[EC0 ] −β[ECj ] j X δk1,ij δk2,ij−1 e e −β[EC ] ! Pk1,k2 = −β[E ] diq e (ij ) (ij−1) q−1 −β[EC ] −β[EC ] C P j0 e j WC = D(z,Si )e hz|Siq |zi = j dzj dzj−1 j0=0 e q e−β[EC0 ]

1 δk1,ij δk2,ij−1 −β[EC0 ] X j h −β[E 0 ] i = e , (36) C (ij ) (ij−1) × D(z,Si )e hz|Siq−1 Piq |zi , (32) Z q−1 j dzj dzj−1 0 where C is the configuration associated with the where E = E + {E } − {E }, E 0 = E − Ck C zk z Ci C multiset EC0 = EC − {Ez}. In the above form, we 00 0 0 {Ez,Ezq−1 } with |z i = Pk2 |z i and |z i is the clas- can reinterpret the weight WC as contributing sical state after the block swap. Similar to the single off-diagonal operator case, the sum P is over all ro- −β[EC0 ] j e 0 pk = δk,iq , (33) tated configurations C whose Siq ends with Pk1 Pk2 . (iq ) −β[EC ] dzq e Measurements of thermal averages of products of more than two off-diagonal operators can also be (iq ) to hPki where dzq is the ’hopping strength’ of Pk derived in a straightforward manner. Eq. (9). As in the case of the diagonal measurements, one can take advantage of the periodicity in the imag- 4. Improved measurements inary time direction to improve statistics by rotat- ing the sequence such that any of the elements of As will often happen, certain physical operators

Siq becomes the last element of the sequence (see will have more than one representation as group el- Sec.VC3), weighted accordingly by the block-swap ement. E.g., if P3 = P1P2, one could measure both probability. By doing so, Pk becomes the single operator hP3i and the operator product hP1P2i and combine the results. −βECj −β[EC0 ] X δk,ij e e pk = (ij ) q−1 −β[EC ] −β[EC ] P j0 e j j dzj j0=0 e VI. RESULTS 1 X δk,ij = e−β[EC0 ] , (34) Z (ij ) j dzj In this section we present some results that high- light some of the advantages that PMR has to of- where E = E +{E } −{E }, the sum P is over Ci C zi z j fer over existing methods. Specifically, we com- 0 all rotated configurations C . pare the performance of PMR over SSE on random 3-regular MAX2SAT instances augmented with a transverse field. This class of instances corresponds 3. Products of off-diagonal measurements to a particular choice of the Ising Hamiltonian given in Eq. (20) whereby each spin is coupled antiferro- The sampling of expectation values of the form magnetically (with strength Jij = 1) with exactly hPk1 Pk2 i proceeds very similarly to the single op- three other spins picked at random. This class of erator case except that now both operators must instances is known to exhibit a quantum spin-glass appear at the end of the sequence. The argument phase transition and is notoriously difficult to sim- proceeds similarly to the single off-diagonal mea- ulate by standard QMC techniques, making it suit- surement, and we have that the contribution to the able to illustrate the strengths of the PMR algorithm expectation value of hPk1 Pk2 i is (see Ref. [34] for more details). In a subsequent section, we demonstrate the ver- e−β[EC0 ] satility of our formalism by considering the perfor- pk1,k2 = δk1,iq δk2,iq−1 (35) (iq ) (iq−1) −β[E ] mance of PMR on a transverse-field Ising model dz dz e C q q−1 augmented with random two-body XX connections,

0 which to the authors’ knowledge cannot be readily with EC = EC − {Ez,Ezq−1 }. As in the single off- diagonal operator case, we can use the block-swap implemented using existing methods [12]. move to alter the elements at the end of the se- quence, and for each pair of adjacent operators in the sequence obtain an improved contribution. By 10

A. PMR vs SSE: Transverse-field Ising model instance is taken to be nm/2, where m ∈ {3, 4, 5} is simulations the average degree of the graph (we focus only on single component graphs for simplicity). To demonstrate the advantages of PMR over ex- Hamiltonians of the above general form appear isting state-of-the-art, we study random 3-regular widely in the context of quantum annealing pro- MAX2SAT instances augmented with a transverse cesses [38], where the system is evolved according field. We study the thermal properties by utilizing to the above Hamiltonian by varying the parameter a parallel tempering scheme [34–37] for both PMR s slowly in time from s = 0 to s = 1. The goal and SSE with 11 replicas at inverse temperatures, in quantum annealing is for the system to reach a β ∈ {0.1, 0.2, 0.5, 1, 2, 5, 10, 20, 30, 40, 50}. We carry state at the end of the anneal that has consider- out simulations for 50 random MAX2SAT instances able overlap with the ground state manifold of the at sizes N = 96, 128 and at transverse field strengths Z-dependent ‘problem’ Hamiltonian, which in this Γ = 0.1, 0.4. case is a MaxCut instance (or a random antiferro- To quantify the performance of the algorithms, we magnetic) [34]. While in standard quantum anneal- fix the total number of measurements of our observ- ing the two-body X terms are normally absent (i.e., ables, and we vary the total number of updates be- b = 0), one is often interested in understanding the tween measurements. Thus, we are able to measure effects of augmenting the Hamiltonian with a ‘cat- the dependence of the estimate of the thermal expec- alyst’ — an extra term that is hoped to reduce the tation value with the total time spent de-correlating amount of time required for the annealing process measurement samples. The performance compar- to take place (see, e.g., Refs. [39–43]). Setting b = 1 isons of the PMR algorithm against SSE are sum- can be viewed as an example of such a situation. marized in Figs.3 and4. Both figures depict the For H in Eq. (37), we have P thermal average of x-magnetization as a function of D0 = Hc = s hiji ZiZj as well as one-body simulation runtime (other observables exhibit simi- and two-body (in the b = 1 case) off-diagonal lar behavior). Pj operators: {Xi} ∪ {XiXj}hiji. The Dj oper- As is evident from the figures, PMR is four or ators are all of the form Dj = dj · 1 where for more orders of magnitude faster than SSE, converg- the one body operators dj = −(1 − s) and for ing on average after only a few seconds in all cases. the two-body operators dj = −s(1 − s). That On the other hand, for a large fraction of the in- dj ≤ 0 implies that the model is stoquastic and stances, the SSE simulations did not finish running hence sign-problem-free. Given a configuration over the 24 hour window allocated for each run. {|zi,Sq = Pi1 Pi2 ...Piq }, the QMC algorithm gen- As expected, the difference in performance is even erated new configuration by either changing basis more pronounced in the more ‘classical’ Γ = 0.1 case state |zi or subsequence of off-diagonal operator (Fig.4). Sq. For our updates, we restrict to subsequences of length two which is enough to ensure ergodicity. The possible completion moves are summarized in B. Ising model with random XX interactions TableI. For illustration, say Pi1 = XiXk then new configuration {|zi,S0 = P 0 P 0 P ...P } where q i0 i1 i2 iq P 0 = X X , P 0 = X X can be generated by We next illustrate the utility of our technique by i0 i j i1 j k studying a model that likely requires highly non- replacing XiXk → XiXj,XjXk and accepting the trivial implementations if studied by other QMC al- move with a probability satisfying detailed balance gorithms. We consider a transverse-field Ising model condition. with random XX interactions whose Hamiltonian is By inspecting the results of QMC simulations of given by the above Hamiltonian we are able to answer a num- ber of questions that are relevant to quantum an- X X H = s ZiZj − (1 − s) Xi nealing. We first examine the variance of H (denoted 2 hiji i σH ), which when close to 0 indicates that the ther- X mal state is close to being purely in the ground state − bs(1 − s) XiXj . (37) of the system (technically, any energy eigenstate of hiji 2 the system will give σH = 0). For sufficiently low Here, s is a parameter in the range (0, 1), and temperatures, this will always be the case, but the b ∈ {0, 1} determines whether the two-body X terms energy gap and the density of states determine how are absent (b = 0) or present (b = 1). We exam- low the temperature needs to be. 2 ine underlying connectivity graphs that are Erd˝os– We therefore study the dependence of σH on the R´enyi random, meaning we randomly pick a pair of instances’ tree-widths. This is shown in Fig.5. spins to connect. The total number of edges for each We see that while instances with different m val- 11

0.3

0.2

0.1 10 1 10 3 10 5 10 1 10 3 10 5 10 1 10 3 10 5 Total time (sec)

Figure 3. Thermal average of the off-diagonal Hamiltonian as obtained by parallel tempering simula- tions using PMR (blue) and SSE (red) for Γ = 0.4 as a function of simulation time. A subset of the results are shown for β = 2, 10 and 50. Each data point is the mean value over 50 random instances, with each simulation performing 500 measurements of the x-magnetization-per-spin in the reported runtime (horizontal axis). The error bars correspond to the 2σ confidence interval generated by 103 bootstraps performed over the instances.

0.1

0.05

0.01 10 1 10 3 10 5 10 1 10 3 10 5 10 1 10 3 10 5 Total time (sec)

Figure 4. Thermal average of the off-diagonal Hamiltonian as obtained by parallel tempering simula- tions using PMR (blue) and SSE (red) for Γ = 0.1 as a function of simulation time. A subset of the results are shown for β = 2, 10 and 50. Each data point is the mean value over 50 random instances, with each simulation performing 500 measurements of the x-magnetization-per-spin in the reported runtime (horizontal axis). The error bars correspond to the 2σ confidence interval generated by 103 bootstraps performed over the instances. ues may have the same tree-width, in the presence of Move Change in q

XX interactions there can be a significant difference i) 1 ↔ (Xi,Xi) ±2 in their σ2 values. We make several observations. H ii) 1 ↔ (XiXj ,XiXj ) ±2 First, we find that the Hamiltonian with XX inter- iii) X X ↔ (X ,X ) ±1 actions requires more sweeps in order to de-correlate, i j i j iv) (X ,X ) ↔ (X ,X ) no change i.e. thermalize. Second, the differences for different i j j i m are more substantial with XX interactions than v) (XiXj ,Xj Xk) ↔ XiXk ±1 without them, indicating that the XX interaction Table I. Cycle completion moves for the transverse field makes the spectrum much more susceptible to m. Ising model with two-body X interactions. The moves Third, larger m values with XX interactions tend to include insertions or removals of pairs of identical one- 2 correspond to lower σH values. This suggests that body and two-body X terms [i) and and ii)], the break- in the presence of XX interactions, larger m values ing up of a two-body term to its one-body constituents are effectively ‘colder’. Finally, we find that the tree- and the inverse operation [iii)], swapping [iv)] and the 2 width makes little difference to the σH values, with contraction of two operators into one [v)]. or without XX interactions. 12

0 1.2

1 -5

0.8 -10 0.6

0.4 -15 0.2

0 -20 3 4 5 6 7 0.3 0.4 0.5 0.6 0.7 Tree width

Figure 5. Variance of H, denoted σ2 , as a function of H Figure 6. Diagonal energy hHpi as a function of s with the tree-width of the underlying graph (here, n = 16, and without the XX catalyst for m = 3 and m = 5 (here, s = 0.5 and β = 2). For m = 3, 4, 5, we have 68, 89, n = 16, β = 2). Note that without the XX catalyst the and 99 instances. The tree-width of each instance is more connected graphs (m = 5) have significantly lower identified, and each bar corresponds to the median value diagonal energy than the m = 3 case. of σH over the instances of a fixed m and tree-width after 107 QMC sweeps. Error bars correspond to 95% confidence interval calculated using a bootstrap over the instances of a fixed m and tree-width. physical models. In addition, studying the thermal properties of transverse-field Ising models, we illus- trated the advantages of our technique over exist- Figure6 shows the average diagonal energy as a ing state-of-the-art, specifically the stochastic series function of the annealing parameter s for the b = 0 expansion (SSE) algorithm. We showed that one (no XX) and the b = 1 case and for two different of the features of the permutation matrix represen- values of average graph degree, namely, m = 3 and tation QMC that distinguishes it from SSE is that m = 5. We find that the presence of the XX cat- it bundles infinitely many SSE weights into a sin- alyst has two important consequences: it minimizes gle weight. We demonstrated that this translates the effect of the graph degree and significantly raises to orders-of-magnitude runtime advantages in most the average value. The former effect suggests that cases. the presence of an XX catalyst will minimize perfor- We further illustrated the flexibility of our method mance differences in solving random MaxCut prob- by studying models with variable localities of in- lems with different graph degrees. The latter effect teractions and underlying connectivities, namely, a is not surprising since the presence of both X and transverse-field Ising model augmented with ran- XX in the Hamiltonian means that we can expect domly placed two-body X interactions, which to the the eigenstates to remain disordered for a larger re- authors’ knowledge cannot be readily implemented gion of s. using existing methods. For these, we found that the presence of the XX interactions can mitigate differ- ences associated with connectivity graphs of differ- VII. SUMMARY AND DISCUSSION ent degree. We expect that our algorithm can be further improved by implementing updates that uti- We presented a parameter-free, Trotter-error free, lize more ‘global’ moves that, e.g., update longer universal quantum Monte Carlo scheme for the sim- cycles of operators. The extent to which this can ulation of a broad range of physical models under translate to further runtime advantages remains an a single unifying framework. Our technique enables open question that we leave for future work. the study of essentially any model on an equal foot- We believe that the generality and flexibility of ing. In our approach, the quantum dimension con- our algorithm will make it a useful tool in the study sists of products of elements of permutation groups, of physical models that have so far been inaccessible, allowing us to formulate update rules and measure- cumbersome or too large to implement with existing ment schemes independently of the model being techniques. studied. We used our approach to clarify the emergence of the sign problem in the simulation of non-stoquastic 13

ACKNOWLEDGMENTS No. de-sc0020280. Work by TA is supported by the Office of the Director of National Intelligence Computation for the work described here was (ODNI), Intelligence Advanced Research Projects supported by the University of Southern Cali- Activity (IARPA), via the U.S. Army Research Of- fornia’s Center for High-Performance Computing fice contract W911NF-17-C-0050. The U.S. Gov- (http://hpcc.usc.edu) and by ARO grant num- ernment is authorized to reproduce and distribute ber W911NF1810227. IH is supported by the reprints for Governmental purposes notwithstand- U.S. Department of Energy, Office of Science, Of- ing any copyright annotation thereon. The views fice of Advanced Scientific Computing Research and conclusions contained herein are those of the (ASCR) Quantum Computing Application Teams authors and should not be interpreted as necessarily (QCATS) program, under field work proposal num- representing the official policies or endorsements, ei- ber ERKJ347. Work by LG was supported by the ther expressed or implied, of the ODNI, IARPA, or U.S. Department of Energy (DOE), Office of Sci- the U.S. Government. ence, Basic Energy Sciences (BES) under Award

[1] D. Landau and K. Binder, A Guide to Monte Carlo [20] D. Joyner, Adventures in group theory. Rubik’s Simulations in Statistical Physics (Cambridge Uni- cube, Merlin’s machine, and other mathematical versity Press, New York, NY, USA, 2005). toys (Baltimore, MD: Johns Hopkins University [2] M. N. . G. Barkema, Monte Carlo Methods in Sta- Press, 2008). tistical Physics (Oxford Uinversity Press, 1999). [21] M. Lewenstein, A. Sanpera, and V. Ahufinger, Ul- [3] A. Aspuru-Guzik, A. D. Dutoi, P. J. Love, and tracold Atoms in Optical Lattices: Simulating quan- M. Head-Gordon, Science 309, 1704 (2005). tum many-body systems (OUP Oxford, 2012). [4] I. Kassal, S. P. Jordan, P. J. Love, M. Mohseni, [22] M. P. A. Fisher, P. B. Weichman, G. Grinstein, and and A. Aspuru-Guzik, Proceedings of the National D. S. Fisher, Phys. Rev. B 40, 546 (1989). Academy of Sciences 105, 18681 (2008). [23] D. Jaksch and P. Zoller, Annals of Physics 315, 52 [5] B. P. Lanyon, J. D. Whitfield, G. G. Gillett, M. E. (2005), special Issue. Goggin, M. P. Almeida, I. Kassal, J. D. Biamonte, [24] T. Giamarchi, C. R¨uegg,and O. Tchernyshyov, Na- M. Mohseni, B. J. Powell, M. Barbieri, A. Aspuru- ture Physics 4, 198 (2008), arXiv:0712.2250 [cond- Guzik, and A. G. White, Nature Chemistry 2, 106 mat.str-el]. EP (2010). [25] I. Hen, Phys. Rev. E 99, 033306 (2019). [6] D. Lonardoni, F. Pederiva, and S. Gandolfi, Journal [26] E. T. Whittaker and G. Robinson, in The Calculus of Physics: Conference Series 529, 012012 (2014). of Observations: A Treatise on Numerical Mathe- [7] S. Chandrasekharan and U.-J. Wiese, Phys. Rev. matics (New York: Dover, New York, 1967). Lett. 83, 3116 (1999). [27] C. de Boor, Surveys in Approximation Theory 1, 46 [8] T. D. Kieu and C. J. Griffin, Phys. Rev. E 49, 3855 (2005). (1994). [28] S. Bravyi, D. P. DiVincenzo, R. I. Oliveira, and [9] A. W. Sandvik, J. Phys, A 25, 3667 (1992). B. M. Terhal, Quant. Inf. Comp. 8, 0361 (2008). [10] A. W. Sandvik, Phys. Rev. B 59, R14157 (1999). [29] S. Bravyi and M. Hastings, arXiv:1410.0703 (2014). [11] N. V. Prokof’ev, B. V. Svistunov, and I. S. [30] M. Marvian, D. A. Lidar, and I. Hen, Nature Com- Tupitsyn, Journal of Experimental and Theoretical munications 10, 1571 (2019). Physics 87, 310 (1998). [31] L. Gupta and I. Hen, Advanced Quantum Technolo- [12] G. Mazzola and M. Troyer, Journal of Statistical gies 2, 1900108 (2019). Mechanics: Theory and Experiment 2017, 053105 [32] F. Zivcovich, Dolomites Research Notes on Approx- (2017). imation 12, 28 (2019). [13] T. Albash, G. Wagenbreth, and I. Hen, Phys. Rev. [33] L. Gupta, L. Barash, and I. Hen, Com- E 96, 063309 (2017). puter Physics Communications 254 (2020), [14] I. Hen, Journal of Statistical Mechanics: Theory and 10.1016/j.cpc.2020.107385. Experiment 2018, 053102 (2018). [34] E. Farhi, D. Gosset, I. Hen, A. W. Sandvik, P. Shor, [15] D. C. Handscomb, Proc. Camb. Phil. Soc. 58, A. P. Young, and F. Zamponi, Phys. Rev. A 86, 594598 (1962). 052334 (2012). [16] D. C. Handscomb, Proc. Camb. Phil. Soc. 60, [35] K. Hukushima and K. Nemoto, J. Phys. Soc. Japan 115122 (1964). 65, 1604 (1996), arXiv:cond-mat/9512035. [17] A. W. Sandvik and J. Kurkij¨arvi, Phys. Rev. B 43, [36] E. Marinari, G. Parisi, and J. J. Ruiz-Lorenzo, 5950 (1991). in Spin Glasses and Random Fields, edited by [18] A. W. Sandvik, Phys. Rev. E 68, 056701 (2003). A. P. Young (World Scientific, Singapore, Singa- [19] J. A. Barker, The Journal of Chemical Physics 70, pore, 1998) p. 59, (arXiv:cond-mat/9701016). 2914 (1979), https://doi.org/10.1063/1.437829. [37] I. Hen and A. P. Young, Phys. Rev. E 84, 061152 14

(2011). The identity operation can thus be written as [38] T. Kadowaki and H. Nishimori, Phys. Rev. E 58,   5355 (1998). 1 0 0 [39] G. A. Durkin, Phys. Rev. A 99, 032315 (2019). P = P 3 = 1 = 0 1 0 ≡ (1)(2)(3) . (A3) [40] T. Albash, Phys. Rev. A 99, 042334 (2019). 0   [41] E. Crosson, E. Farhi, C. Yen-Yu Lin, H.-H. Lin, 0 0 1 and P. Shor, ArXiv e-prints (2014), arXiv:1401.7320 [quant-ph]. We illustrate the evaluation of a product of two [42] L. Hormozi, E. W. Brown, G. Carleo, and cycles by computing P 2 in cycle notation. Since M. Troyer, Phys. Rev. B 95, 184416 (2017). 2 [43] H. Nishimori and K. Takada, Frontiers in ICT 4, 2 we defined P = (1, 2, 3) we have P2 = P = (2017). (1, 2, 3)(1, 2, 3). By the above definition [44] Rotman, An Introduction to the Theory of Groups P2(1) = (1, 2, 3)(1, 2, 3)(1) = (1, 2, 3)(2) = (3), (New York: Springer-Verlag, 1995). P2(2) = (1, 2, 3)(1, 2, 3)(2) = (1, 2, 3)(3) = (1), P2(3) = (1, 2, 3)(1, 2, 3)(3) = (1, 2, 3)(1) = (2). Therefore P2 = (1, 3, 2). If we enumerate the basis Appendix A: Finite dimensional permutation states as matrix representations       1 0 0       To show that any finite-dimensional matrix can be 1 ≡ |1i ≡ 0 , 2 ≡ |2i ≡ 1 , 3 ≡ |3i ≡ 0 , written in the form of Eq. (1), we will make use of 0 0 1 ‘cycle notation’ [44] — a compact representation of (A4) permutations — to represent the permutation ma- then with the action described above one can trices P . We start with some terminology. matrix notation j see that |ki = Pi |ji corresponds to A cycle is a string of integers that represents an cycle notation k = Pi (j). element of the symmetric permutation group Sn, which cyclically permutes these integers and fixes all In the above notation, the set of n × n other integers. For example cycle (a1, a2, . . . , am) is permutation matrices can be seen as groups the permutation that sends ai to ai+1, 1 ≤ i ≤ m−1 generated by an n-cycle. Any given n-cycle and sends am to a1. The cycle given in the above σ = (a0, a1, a2, a3, ..., an−1) ∈ Sn, where Sn is the example is an m-cycle. In general, any element symmetric permutation group, has order n. To σ ∈ Sn can be written as a product of k cy- k see why this is so, observe that σ (a0) = ak for cles as (a a . . . a )(a a . . . a ) ... 1 2 m1 m1+1 m1+2 m2 0 < k < n so its order cannot be less than n and (am +1 am +2 . . . am ). The order of a permu- n n k−1 k−1 k σ is the identity as σ (ai) = ai for i ∈ {0, n − 1}. tation σ is defined as the smallest positive integer Therefore the group generated by σ has n elements. p p such that σ is the identity element. In this k More generally we have: σ (aj) = a(j+k) mod n. This notation, the action of σ on any number from 1 implies σk1 (a ) 6= σk2 (a ) for k 6= k . to n can be determined as follows. If a appears j j 1 2 Let P be the permutation matrix corresponding at the right end of one of the k cycles, then σ(a) to σ. Then, P k1 |a i 6= P k2 |a i for k 6= k and is the integer at the start of the cycle to which a j j 1 2 basis vector |a i. Since any row (or column) of a belongs. If an integer a does not appear at the j permutation matrix has value 0 at all positions but right end of one of the k cycles, then σ(a) is the in- one, where it has the value 1, this implies that no teger to the right of a in the cycle to which a belongs. two permutation matrices generated from P have the same row otherwise that would mean P k1 |a i = For concreteness, let us write the 3×3 permutation j P k2 |a i for some k , k , a . matrices used in Sec.IIC in cycle notation. j 1 2 j Next, we will show that for any given matrix en-   try (i, j) there is at least one permutation matrix 0 0 1 P k having value 1 at (i, j), that is P k = 1 for some   i,j P1 = P = 1 0 0 ≡ (1, 2, 3) , (A1) k. Let |ii denote the basis vector which has value 0 1 0 1 at the i-th index and 0 at all others. Since these permutation matrices form a group there must be some matrix P k such that P k|ji = |ii. This how-   0 1 0 ever means that P k has value 1 at (i, j). Combining 2 P2 = P = 0 0 1 ≡ (1, 3, 2) . (A2) the two statements above, we find that for any en-   try (i, j) there is exactly one permutation matrix 1 0 0 generated from P that has value 1 at that entry. The diagonal matrices Dj may be used to convert 15

these 1’s to any desired value. We have thus shown ement s such that Dj(s,s) 6= 0. Let us denote that by choosing our Pj permutation matrices as by r the element that is sent to s in Pj, that is n-cycles, one can construct arbitrary Hamiltonians P = ... (. . . p, q . . . r, s . . .) .... Again since H is P j H = j DjPj. This proof also provides a prescrip- hermitian there must exist Pj0 = ... (. . . s, r . . .) ... tion as to how to explicitly choose the permutations in the decomposition of H. But then PjPj0 = Pj. ... (. . . s, s . . .) ... which has to be equal to 1. Thus Pj has an inverse Pj0 in the decomposition of H. Next, we prove that there can be no two permu- tation matrices in H with the same nonzero ele- Appendix B: Hermiticity of H ment. In cycle notation, this assertion translates to the assertion that there can be no two distinct permutations that send a basis state to the same Here we show that H = P D P is hermitian if j j basis state. We prove this by contradiction. Let and only if for every index j there is an associated 0 −1 ∗ Pj and Pk be two distinct permutations both of index j such that P = P 0 and D = D 0 (in j j j j which send a basis vector p to q. In ’cycle nota- general, j and j0 may correspond to the same index). tion’ this mean Pj = ... (. . . p, q . . .) ... and Pk = We first prove the if direction. Let H above be ... (. . . p, q . . .) ... with Pj 6= Pk. We showed above a . We show that this implies that Pj has an inverse Pj0 , that is PjPj0 = 1. But that for every index j there is an associated in- Pj0 Pk = ... (. . . p, p . . .) ... has fixed point and can- 0 −1 dex j such that Pj = Pj0 . Let Pj be the per- not be the identity due to the uniqueness of the in- mutation sending a basis vector p to another ba- verse. Thus we have reached a contradiction. sis vector q. Thus in ‘cycle notation’ (see Ap- We now prove the full if part. By definition, that P P ∗ −1 pendixA) Pj = ... (. . . p, q . . .) .... Let us assume H is hermitian implies DjPj = Dj Pj . Let that Dj is not the zero matrix (otherwise DjPj is Pj0 be the inverse of Pj which as we proved should trivially zero). Case I: Let Di(q,q) 6= 0. Since H is exist in the decomposition with a nonzero weight Dj. hermitian thus there exist Pj0 = ... (. . . q, p . . .) ... Equating the right-hand side and the left-hand side ∗ in the decomposition of H. But then PjPj0 = of the equality gives Dj = Dj0 . ... (. . . q, q . . .) ... which has to be equal to 1 (as Proving the other direction is simpler. We assume P otherwise this would imply the existence of a per- that in the summation H = DjPj there is for 0 mutation matrix that has a fixed point contradic- every index j an associated index j such that Pj = −1 ∗ † P ∗ −1 tory to our initial setup). Case II: Let Dj(q,q) = 0. Pj0 and Dj = Dj0 . Thus H = Dj0 Pj0 = P Since Dj is not identically zero, there exists an el- DjPj = H.