Non-uniform Mixing of Quantum Walks on the Symmetric Group
Abstract
It is well-known that classical random walks on regular graphs converge to the uniform distribution. Quantum walks, in their various forms, are quantizations of their corresponding classical random walk processes. Gerhardt and Watrous (2003) demonstrated that continuous-time quantum walks do not converge to the uniform distribution on certain Cayley graphs of the Symmetric group, which by definition are all regular. In this paper, we demonstrate that discrete-time quantum walks, in the sense of quantized Markov chains as introduced by Szegedy (2004), also do not converge to the uniform distribution. We analyze the spectra of the Szegedy walk operators using the representation theory of the symmetric group. In the discrete setting, the analysis is complicated by the fact that we work within a Hilbert space of a higher dimension than the continuous case, spanned by pairs of vertices. Our techniques are general, and we believe they can be applied to derive similar analytical results for other non-commutative groups using the characters of their irreducible representation.
Keywords: Quantum Walks, Symmetric Group, Non-commutative Fourier analysis
Mathematics Subject Classification: 81P68, 20C30.
1 Introduction
The phenomenon of random walks on graphs has been widely studied and holds significant applications across a myriad of problems in computational sciences. They have been instrumental in developing randomized and approximation algorithms [25]. Random walks can be characterized by Markov chains and they can be fully characterized using methods from spectral graph theory.
We look at the problem of sampling from the symmetric group via a quantization of random walk. The study of sampling from groups has a rich history [15, 16, 7]. Particularly, sampling an element from the symmetric group has been well-studied in the classical setting [16], especially with respect to functions over elements of groups. Sampling from group elements ties with certain random walks; in some cases, even if the original sampling problem does not involve sampling from a group element (for example, the famous Ehrenfest process). These random walks take place on the Cayley graphs of the groups, which are constructed using some generating set. Since Cayley graphs are regular, a uniform random walk on them converges to the uniform distribution. However, this does not seem to be the case in the quantum setting. We extend the analysis of Gerhardt and Watrous (2003) [20] to demonstrate the uniformity of the distribution, both instantaneous and average, arising from the quantization of a uniform Markov chain on the symmetric group.
Unlike classical random walks a quantum walk propagates using the principle of quantum mechanics. Few difference of note include - 1) Instead of real probabilities the state of the walk is specified by complex probability amplitudes111However, in some case if the amplitudes are constrained to be in , working with them becomes slightly simpler.. 2) The random (walk) coin is now replaced by a unitary transformation. The unitary evolution ensures the walk is reversible222For open systems the walk operator need not be unitary. Interspersing walking with measurements also leads to non-unitary dynamics[22]. 2) Propagation of the walk generates a superposition state overs all possible positions available to the walker. 3) Finally, we can sample the positions by applying suitable measurements on the state of the walker.
There are various (somewhat equivalent) models of quantum walks. Study of quantum walks has a long history, going back to the early works of Feynman, Meyer, Aharonov, Gutmann and others [3, 18, 27]. The hope is that quantum walk can emulate the success of random walk in the development of classical algorithms in developing quantum algorithms. Quantum or classical walk333Henceforth we will refer to classical random walk simply as classical walk. has been primarily used as a generative models for probability distributions. Hence, two of the most important properties to study are the kind of distributions they can generate and their converging behavior. In general, quantum walks do not converge to a stationary distribution. However, their time-averaged distribution (introduced later) does converge. Quantum walk has been shown to generalize Grover’s diffusion based search on graphs. It has been used to obtain currently best known quantum algorithms for certain problems. Most notable among them are element distinctness, triangle finding, faster simulation of Markov chains, expansion testing etc. [26, 5, 6].
1.1 Overview of our techniques and Results.
In this paper, we focus on a discrete-time model of quantum walk. The model we examine has its origins in the seminal paper by Aharonov et al.[2]. Since its introduction, numerous variants of discrete-time quantum walks (DTQW) have emerged. When it comes to accelerating randomized algorithms by harnessing the faster mixing properties of quantum walks, the most extensively studied framework involves the quantization of a classical Markov chain. This framework was first introduced by Szegedy [35], and has since been expanded upon and applied to a multitude of graph search algorithms within the black-box query model. In this paper, we employ the Szegedy quantum walk framework to analyze the distribution properties of a specific type of Cayley graph of the symmetric group. Our primary concern is not the mixing time or other measures of convergence, but rather illustrating how the probability distribution deviates from that of the corresponding classical Markov chain. This research implies that the probability of observing a group element is intricately tied to the “weight” of its various irreducible representations. In the case of abelian groups, given that their representations are 1-dimensional, they play a consistent role for all group elements, and, as demonstrated previously, such walks converge to the uniform distribution. This distinctive difference renders the study of such walks for symmetric and other simple non-abelian groups considerably more intricate. Further discussions on previous results can be found in Section 3.
Given a Markov chain with its transition matrix , we begin by constructing a bipartite walk on the combined state space . The transition matrix of this bipartite walk forms the basis for deriving a unitary operator in the quantum context. Informally, for each state , one constructs a vector that represents a superposition of the edges linking to its neighbors, with weights corresponding to the transition amplitudes. These vectors are utilized to define a reflection operator as well as a shift operator (which will be introduced later). The reflection operator allows the quantum walk to “propagate in superposition” along the edges adjacent to a vertex. The shift operator alternates the propagation direction from left-to-right and vice versa. The composition of these operators results in a unitary , which characterizes a step of the quantum walk. This construction facilitates a relatively straightforward determination of the spectral decomposition of in relation to the spectra of . The elements of can be interpreted as vertices of an edge-weighted directed graph, where the weights correspond to the transition probabilities. In our context, this graph is associated with a specific Cayley graph of the symmetric group. The quantization of the bipartite Markov chain gives rise to a quantum walk, which fundamentally occurs on the edges of the original graph. Due to the inverse closure of the generating set we utilized, this graph is undirected.
In the case of continuous-time quantum walks, which evolve based on the Hamiltonian where is the graph Laplacian, Gerhart and Watrous studied the walk on several Cayley graphs of the symmetric group. They utilized the spectral decomposition of the random walk operator in terms of the irreducible representations (irreps) of the symmetric group, initially derived by Diaconis [15], to determine the probability of observing an -cycle for the quantum walk. They demonstrated that this probability, , is exponentially smaller than in the case of the uniform distribution (which is ). We extend their technique in the setting of the Szegedy walk. We apply the spectral decomposition of in terms of the irreps of the group to construct a similar, albeit more technical, spectral decomposition of . This enables us to similarly upper bound the probability of observing an -cycle, in our case determined by all edges incident to -cycles. However, due to the difference in the Hilbert space of the quantized Markov chain compared to the continuous-time version, our analysis presents a considerably greater challenge.
The analysis highly depends on the tractability of working directly with irreps. Unfortunately, the irreps are matrices with no simple formulaic description. This restricts us to focusing our analysis on cases where we can use the characters of the group elements instead of the irreps. In light of this, we limit our study to generating sets that are conjugacy closed. Specifically, in this paper, we focus on the generating sets which consist of transpositions. Even with this limitation, the analysis is still influenced by the choice of the initial state and the form of the final state. Particularly, the support of the probabilities in the final state should also be over a conjugacy-invariant set. However, this isn’t an issue for us as, when testing the probability of observing an -cycle, we can choose the uniform distribution over all -cycles (more technically, all edges incident to -cycles) and determine its overlap with the final state. Much of the technical calculations in this paper are focused on determining this overlap using the characters of the group. This lead to our main theorem.
Theorem 1.
For any constant ,
where is the uniform superposition of all edges incident to -cycles, and is the uniform superposition of all edges incident to the identity permutation.
Ignoring the polynomial factor, which arises due to the technical limitations of our approach, we observe that the walk operator behaves roughly similarly to that of the continuous version.
1.2 Discussion
1.2.1 Classical Complexity of Generating Quantum Walk Distribution, Localization
Consider a unitary operator drawn from the unitary group according to the Haar measure. When acts on the state , it results in the state . It’s worth noting that can be replaced with any fixed initial state, not necessarily the all-zero state. Let represent the probability of observing the state when the system is in state , with being a computational basis state.
It is a well-established fact that when is viewed as a continuous random variable (determined by the Haar measure on ), the distribution of probability values, , conforms to the Porter-Thomas distribution as given by: . An intrinsic property of this distribution is that the probability amplitude distribution of is anti-concentrated; this means that no specific basis state is noticeably more or less probable than the uniformly distributed probability value of . Nevertheless, there isn’t a classical process available to effectively approximate such outputs, even though it’s feasible to devise a classical stochastic process that can mimic the Porter-Thomas distribution for these probabilities [9, 10]. Our result, which is consistent with those from continuous-time cases, reveals that the unitary operator of the Discrete-Time Quantum Walk (DTQW) on the symmetric group is considerably different from a Haar-random unitary. Additionally, its probability distribution deviates markedly from uniformity. We observe that the quantum walker has certain blind spots, which may hint at some localization phenomenon and consequently a lack of anti-concentration. It is not a priori evident that such distributions can be generated efficiently in the classical setting.
1.2.2 Beyond Class Functions
While the techniques presented here are farely general, their capacity to derive analytical results largely hinges on the ability to use characters of certain irreducible representations of the symmetric group. Generally, the spectrum of the DQTW operator will depend directly on the irreps, unless the discriminant matrix, derived from the Markov chain, possesses some special structure. As outlined in [15], the eigenvalue expression utilizing class functions is valid for any matrix whose entry can be expressed by a class function . More comprehensively, for any transition matrix , one can form a block-diagonal decomposition:
Here, the columns of the matrix form an orthonormal set of vectors spanning the vector space defined by elements of . Moreover, is a block-diagonal matrix, with blocks corresponding to components of the Fourier transformation of (defined as a function from to where for a certain probability distribution on , not necessarily a class function) in terms of the irreps of . In scenarios where is a class function, is strictly a diagonal matrix, and the column vectors of are the vectors , with denoting an irrep of . Nonetheless, within this broader framework, the spectral decomposition of the Szegedy walk operator is directly influenced by the matrix entries of the irreps, not solely the trace (commonly known as the characters of the irreps). Currently, there is no apparent method to broaden our analysis beyond class functions.
1.2.3 Discrete Heisenberg Group
The discrete Heisenberg group has found applications in physics [19, 4] as well in complexity theory [24]. It is one of the simplest non-abelian extension of the regular 2d lattice, for which random walks have been thoroughly studied. Furthermore, it has an elegant description with respect to its center. The 3-dimensional discrete Heisenberg group over is defined by the following multiplication rule : (modulo ). Dynamics of random walks over them are well understood, and known to converge to the uniform distribution in steps [11, 39]. Here, we are especially interested in the case where , a prime. This group is extra-special, in particular, is abelian and is cyclic where is the center of . For this case we can to consider the Schreier coset graph of with respect to the cosets of . As far as we are aware, quantum walks have not been studied on coset graphs in the discrete setting (for an example in the continuous case see [30]). The techniques presented here could be applied to this group, possibly taking advantage of its “nested” structure and more manageable representations.
2 Preliminaries
2.1 Symmetric Group, Cayley Graphs and Representation Theory
2.1.1 Cayley Graphs
Let be any finite group, and let be a generator of . We define and . The Cayley graph of the pair, denoted as , is a directed graph , defined as follows: The vertex set is , and the edge set is defined as
Henceforth, we omit the “” and simply write as for all . If is closed under taking inverses, i.e., , then is undirected. In this paper, we set , the symmetric group of permutations of elements. We use to denote the identity permutation, where . We will exclusively work with the generating set composed of all transpositions in , i.e., . Thus, is closed under conjugation, and is a -regular undirected graph. Throughout this paper, we use , , , , etc. to denote generic group elements. Occasionally, we use and to emphasize elements of the symmetric group.
The quantum walk studied in this paper does not take place directly on but on a bipartite extension, denoted as , of . Where
We further elaborate on this when introducing Szegedy walks in Section 2.2.
2.1.2 Representation of the Symmetric Group
Representation theory provides a framework to study abstract algebraic structures by representing their elements as linear transformations of vector spaces. In particular, the representation theory of the symmetric group, the group of all permutations of a set, holds profound mathematical importance and has ties to diverse areas. Here, we briefly introduce only the relevant definitions needed to present our analysis. A comprehensive introduction to representation theory in the context of symmetric groups, non-commutative Fourier analysis, and random walks can be found in the books and monographs by Sagan [32], A. Terras [36], and Diaconis [15], respectively, as well as in the references therein. Much of the following material has been taken from those sources.
A representation of a group on a vector space over a field is a homomorphism , where is the group of invertible linear transformations of . Specifically, for each element , there’s an associated matrix that respects the group operation: for all . In our setting we take , the field of complex numbers. The dimension of a representation corresponds to the dimension of its associated vector space , denoted as . The representative matrices are matrices, which can be made to be unitary. A representation is termed irreducible if no non-trivial invariant subspaces exist within it. This means the only subspaces of invariant under every transformation , are itself and the zero subspace. Henceforth we shall refer to irreducible representations as simply irreps for brevity. For a given representation , the character of this representation, , is a function from to the field defined by the trace of the representation’s matrix: . Following properties of will be useful:
-
1.
-
2.
(cyclic property)
-
3.
( is constant over the conjugacy classes)
Elements and in are termed conjugate if there’s an such that . All elements conjugate to form its conjugacy class. In symmetric groups, conjugacy classes are characterized by a permutation’s cycle type. A class function on group is a function (for some field ) that remains constant on conjugacy classes. Characters of representations are classic instances of class functions.
A Young diagram associated with a partition of the number (that is , denoted as ) consists of left-justified boxes in the top row, in the second, and so on. We will construct Young diagrams using the English convention, with row lengths decreasing or remaining constant from top to bottom. A Young tableau fills this diagram with numbers from to such that entries in each row and column are increasing. A rim hook is a set of boxes that can be removed from a Young diagram, leaving another Young diagram behind.
The following text about Young normal form is taken from [20] and has been modified to match the language of the present paper. A more comprehensive description can be found in [38, 32]. The symmetric group containing elements provides a unique method to link partitions of (which have a bijection to the conjugacy classes of ) to a full set of distinct, irreps of . These distinct irreps are identified as the Young normal forms. A notable feature of these irreps is that each matrix entry within them is an integer. For each such representation, we may associate another irrep up to an isomorphism, with all its matrices being unitary. The irreducible, unitary representation corresponding to a specific partition is denoted as , and its corresponding character is expressed as .
Finally, we briefly discuss the Murnaghan-Nakayama Rule, which is useful for computing the characters of certain irreps and conjugacy classes we used in this paper. This is a combinatorial method used to compute character values for symmetric group representations indexed by Young tableaux. The character of a permutation with cycle type in the representation corresponding to a Young tableau of shape is determined by iteratively removing rim-hooks and summing the associated contributions. We need to define a few more terms before we can proceed. Given a Young diagram of shape (partition) and a composition of 444A composition of is a partition where the order of the parts matters. For a given partition, the collection of compositions having the same parts corresponds to different permutations with the same cycle structure; hence, their characters are the same., a filling of using the content from is a labeling of the cells of such that cells are labeled with , cells are labeled with , and so on. Additionally, the labeling must satisfy the following two conditions: (1) the cells corresponding to the same label have the shape of a rim-hook, and (2) labels are non-decreasing along rows and columns. Such a filling of the shape is called a rim-hook tableau. The leg-length (denoted as ) of a rim-hook is defined as . The sign of a rim-hook tableau is given by:
Then, the Murnaghan-Nakayama Rule provides a way to compute :
where the sum is over all valid rim-hook tableaux for the pair .
2.1.3 Fourier Transform Over Non-Commutative Groups
The Fourier transform is a powerful tool in signal processing and applied mathematics, enabling the analysis of a signal’s frequency content. In the case of groups (), the Fourier transform of a function performs a basis change from to . Here, is the entry of the matrix presentation of for different group elements, thus constituting a function of the form . As observed, in the case of non-commutative groups, the Fourier transform takes a more complex form than in the commutative case, owing to the fact that the irreps are themselves linear operators. More formally, the Fourier transform over finite groups is defined as:
In the case where is a class function, this simplifies to:
where the sum is over all conjugacy classes in , and denotes the size of . We shall use the latter expression when computing certain projections with respect to the Szegedy walk operator.
2.2 Quantizing Markov Chains: Szegedy Walk
Szegedy developed the framework of quantizing a Markov chain [35], which was then used to derive a quantum speedup of random-walk-based search algorithms on graphs. Here, we borrow Szegedy’s terminology. Let and be two parts of a bipartite graph, and let and be probabilistic maps from to and from to , respectively. For any and let
and | ||
Further, let (resp. ) be matrix composed of the column vectors (resp. ). Define two reflection operators - and . Finally the quantum walk unitary is defined as
The formulation above generalizes coined quantum walks on regular graphs in the following sense. To provide some intuition about this definition, we briefly introduce coined quantum walks. A classical random walk on vertices cannot be directly quantized into a unitary operator on the Hilbert space spanned by the vertices. To create a unitary, one has to lift the space on which the quantum walk takes place to a product of two Hilbert spaces: 1) a “coin space,” which is used to propagate amplitudes from a vertex to its neighbors in superposition, and 2) a shift or move operator that transfers the walker from the current vertex to its neighbors. More formally, the state of such a particle at any moment is described by a vector in the Hilbert space , with a basis set (standard basis), for some -regular graph with vertex set . Thus, we can express as . The space describes the position of the particle over the vertices. is the coin space, which describes the state of the particle’s internal degrees of freedom (sometimes referred to as the particle’s chirality). One step of the walk consists of successively applying the two unitaries and , where . Here, denotes the neighbor of . The shift operator moves the walker to its neighboring vertex in superposition. The coin operator determines how the amplitudes spread to neighboring vertices, acting like a quantum analogue of a classical -sided die. For graphs with arbitrary vertex degree, is replaced by the reflection operator . It is easy to see that if we take , then , where shift operator is generalized as . The spectral properties of the walk operator are closely related to the discriminant matrix , whose entries are defined as . In the setting of quantized random walks, we shall take . Furthermore, the transition probabilities are assumed to be uniform, and as such, is symmetric. Thus, . Specifically,
(1) |
2.2.1 Instantaneous and Limiting Distribution
Given the initial state , the state after steps of the walk is represented as:
As previously discussed, the basis of the Hilbert space in which the walk takes place consists of pairs of permutations from , and we refer to this as the standard basis. From this point onward, we assume that all measurements are performed in this basis. The probability of sampling a permutation —more specifically, observing it in the first register—of after steps of the walk is given by:
Since is unitary, exhibits periodicity [2], provided that is not an eigenvector of . Generally, does not converge. However, the time-averaged distribution, defined below, does converge as :
can be interpreted as the expected value of the distribution when is selected uniformly at random from the set .
In this paper, we are interested in upper-bounding , where is the state representing the uniform superposition of pairs in which the first register is an -cycle. For the case of classical random walk this is analogous to determining the probability of sampling an -cycle. defines a distribution on , which depends on the initial state and the choice of the generating set . Since in our analysis both and are fixed we simply use to denote this time averaged distribution.
2.2.2 Continuous Time Quantum Walks and Average Mixing Matrix
Here we briefly introduce some notion related to continuous time walks that will be useful to compare this work with some previous and recent results in the domain of continuous time quantum walks. A comprehensive introduction to concepts presented here can be found in [13] and the references therein. Let be the adjacency matrix of a undirected regular graph with vertex set . is hermitian and as such defines a Hamiltonian evolution on the Hilbert space spanned by the vertices of the graph. Let . is known as the continuous time quantum walk operator. It is important to note here that acts directly on the vertices, which is not possible in the discrete setting. In the setting of continuous time quantum walk a there is another notion of time average distribution - the average mixing matrix [21]. Let denote the Schur product of and . Then , which is a doubly stochastic matrix. On the standard basis spanned by the vertices, the collection gives rise family of probability densities - which can be interpreted as the resulting distribution after evolving for time starting from the vertex . To define the time average distribution , we can work with the time average version of , denoted as , called the average mixing matrix. :
Recently, average mixing matrix has been extended in the discrete setting [34]. In the language of this paper, we can define an average mixing matrix for the Szegedy walk operator as follows:
Here, the matrix is the matrix we defined earlier when introducing the Szegedy walk. We can interpret as the average probability of observing in the first register, starting from the state . In this context, the initial state is a superposition over the outgoing555Even though the walk takes place on an undirected graph, we may assign an orientation to an edge with respect to the vertex where the walker is situated, treating it as the tail of the edge. Since we are only considering bipartite walks, the walker can be present at most at one of the endpoints. edges from , according to the amplitude distribution . The average mixing matrix can be useful for studying the average limiting behavior of the quantized Markov chain. In particular, if the Markov chain mixes to a uniform distribution, then an analogous notion can be considered in the quantum case, where we are interested in how close is to the matrix , with being the matrix whose all entries are 1. If equals in the limit, then we say the chain exhibits average uniform mixing.
3 Previous and Related Work
Discrete Time Quantum Walk on Groups.
In their seminal paper [2], Aharonov et al. presented several results on DTQWs. They characterized the convergence behavior of walks on abelian groups, showing that the time-averaged distribution converges to the uniform distribution whenever the eigenvalues of are all distinct. They also provided an upper bound on the mixing time for (the cycle graph), and proved some lower bounds in terms of the graph’s conductance. Following their introduction, DTQWs have been studied for several graph families. Nayak and Vishwanath [29] conducted a detailed analysis for the line using Fourier analysis, demonstrating that the Hadamard walk mixes almost uniformly in only steps, achieving a quadratic speedup over its classical counterpart. Moore and Russell [28] analyzed the Grover walk on the Cayley graph of (also known as the hypercube), showing an instantaneous mixing time of , which beats the classical bound. Acevedo and Gobron [1] studied quantum walks for certain Cayley graphs, providing several results for graphs generated by free groups in particular. D’Ariano et al. [17] investigated the case where the group is virtually abelian, a condition that allowed them to reduce the problem to an equivalent one on an abelian group with a larger chiral space dimension, employing the Fourier method introduced in [29]. More recently, DTQWs on the Dihedral group have been studied by Dai et al. [14] and Sarkar and Adhikari [33]. Since is isomorphic to the semi-direct product , the Fourier approach introduced in [29] is applicable once again. Using this method, the authors in [14] provided a spectral decomposition of for the Grover walk. In [33], the authors study the periodicity and localization properties of the walk using generalized Grover coins. A detailed survey of various types of quantum walks, including DTQW, can be found in [37], with additional references therein. A survey specifically addressing DTQWs on Cayley graphs is available in [23].
Quantum Walk on the Symmetric Group.
In a previous work by the author [8], DTQW on the symmetric group was studied using the coin-based model. Utilizing the Fourier transform (see Section 2.1.3), a recurrence relation was derived for the amplitudes of , from which a “sum-over-paths” type expression was determined for the amplitudes. It was also determined under which conditions the amplitudes are class functions. Prior to this, as indicated earlier, Gerhardt and Watrous [20] studied the continuous time quantum walk model on the symmetric group. They showed that when is the set of transpositions, the time-averaged distribution is far from the uniform distribution. They explicitly calculated the probability of reaching an -cycle starting from by expressing the eigenstates of using the characters of . They also considered the generating set (for ) consisting of all -cycles (where is odd) and derived similar results as in the transposition case.
Average limiting behavior.
Considerable research has been conducted on studying the limiting behavior of quantized Markov chains, as mentioned earlier. More recently, the properties of the average mixing matrix have been explored, especially for continuous-time chains. For example, in [34, 12], the entries of the average mixing matrix in the limit as were expressed using the projectors in the spectral decomposition of the walk operator. One of the interesting questions with respect to limiting behavior is whether a quantized Markov chain exhibits average uniform mixing. To this end, the authors in [34] have constructed a family of Markov chains whose quantized versions do exhibit such average mixing behavior, and have shown that average uniform mixing of the continuous-time quantum walk implies the same for its discretized version.
4 Eigendecomposition of over the irreps of
Gerhardt and Watrous used representation theory to express the eigenstates using the irreps of . This method is effective due to the fact that the walker’s Hamiltonian is completely specified by the adjacency matrix of . In the discrete case, as acts on a larger space, this decomposition becomes more complex for an arbitrary generating set. However, there is at least one special case where we can directly apply their Fourier method.
This special case occurs when the generating set forms a group itself. For the amplitudes to be uniform over the conjugacy class, which is a necessity for using Fourier analysis over the characters of the irreps., the generating set must be conjugate invariant. However, the only non-trivial subgroup of that is also conjugate invariant is the alternating group , which is the subgroup of all even permutations in . In this scenario, it becomes possible to factorize the space of irreps. for the walk on , and determine the spectral decomposition of using the characters of both and . To overcome this issue we use the Szegedy walk formalism, which considers an even larger coin-space, as introduced earlier.
As Szegedy showed, the dynamics of the walk operator can be determined from the discriminant matrix . Since is Hermitian (in fact, symmetric), the singular values of lie in the interval . We index the singular values of using the conjugacy classes of , which are the partitions of . For each , if the corresponding eigenvalue is also , then the left and right singular vectors are equal (and are equal to the corresponding eigenvector); otherwise, they differ by a minus sign. For the former case, we use to denote both the left and the right singular vectors. For the latter case, without loss of generality, we use to denote the left singular vector by appropriately choosing the sign of the corresponding eigenvector. Let (resp., ) denote the projector onto the column space of (resp., ), and let (resp., ) denote the projector onto the orthogonal complement of the column space of (resp., ). We can restate the spectral lemma from [35] in the language of this paper, which will be used in our subsequent analysis.
Lemma 2 (modified Lemma-1 from [35]).
Let (with multiplicity) are the sequence of singular values of in the interval and be the eigenvalue corresponding to . Then the eigenvalues and eigenvectors of the walk operator is and respectively (up to a normalization). Additionally, corresponding to the singular value , acts as the identity () on the space and corresponding to the singular value , acts as on the space .
From the definition of , as presented in [20], we may define a class function on such that , provided that is conjugate invariant (which holds true in our case). We will employ the spectral decomposition of in terms of the characters of , as given in [20], which takes advantage of the fact that the entries of behave as a class function. This is a special case of a more general result by Diaconis [15] presented earlier. It has also been shown that the basis of the irreps are the eigenvectors of . This can, in turn, be utilized to derive the spectra of the walker’s Hamiltonian by exponentiating the corresponding eigenvalues of . In the discrete case, the relationship between and the walk operator is somewhat more subtle, and the remainder of this section is devoted to elucidating it.
Recall that the conjugacy classes in have a one-to-one correspondence with the collection of non-isomorphic irreps of . These irreps can be expressed using the so-called Young normal form, whose matrix entries are all integers. Let denote the irrep corresponding to the conjugacy class (with size given by ), which is a partition of (). Define . The vectors (in the -module over the field of complex numbers) form an orthonormal basis corresponding to the irrep . For reference, we restate Lemma 6 from [20], which gives the expressions for the eigenvalues of in terms of the characters of .
Lemma 3 (modified lemma-6 from [20]).
Given as above, then are the eigenvectors of with the eigenvalues,
(2) |
Recall that are the singular values of . Let . Associated with each non-extremal singular value () of , there is a collection of eigenvectors , where . The component of is given by . The projectors onto the space spanned by are given by
where is a normalization factor, since we have . Let be the projector onto the column space of . Let the projectors corresponding to the subspaces and be and , respectively. Further, let . We can completely determine the spectral decomposition of using that of , as given by the following lemma. Let , and let .
Lemma 4.
Spectral decomposition of is given by,
Proof.
The proof directly follows from the preceding discussions and Lemma 2. The factor comes from normalizing the eigenvectors. ∎
The decomposition of can then be divided into two parts: one corresponding to the non-trivial eigenvalues, and the other corresponding to the two trivial ones, . Let be the term corresponding to . Then,
(3) |
where, by a slight abuse of notation, we assume the sum above is over all partitions except those corresponding to the trivial eigenvalues. It follows that , since the projectors are mutually orthogonal. Expanding we get,
(4) |
From the above, we see that for any pair of initial and final states and , respectively, an upper bound on the inner product depends on the projectors respective to the irreps. In particular, if the overlap is sufficiently low, then we can ignore the effect of the cosine terms (replacing them with 1 or -1, as appropriate) and still obtain a non-trivial upper bound. In the following, we use this approach to compute .
5 Divergence of from uniform mixing
In [20], the divergence of the instantaneous distribution from the uniform distribution was confirmed by upper bounding the probability of being at an -cycle of . This probability was shown to be exponentially smaller than in the classical case. Specifically, they showed that the probability of being at some -cycle is , as compared to in the classical case (uniform distribution). In the discrete case that we study here, the walk takes place on the edges of the graph. We show that the instantaneous probability of observing an -cycle in the first register is upper bounded away from by a function which is . Here, represents the soft-O notation, but in our case, we ignore any functions up to polynomial in (which are most likely due to artifacts from our techniques). Since the bound we provide is on the instantaneous probability, it also implies that the average mixing probabilities (entries of the average mixing matrix) are exponentially far from uniform.
In the following, we use “” to denote generic permutations from , aiming to avoid confusion arising from our slight notational abuse. Specifically, we use to denote a generic conjugacy class that contains the permutation . Recall,
Also, is the degree of . We will provide an upper bound for . The state is the uniform superposition over all outgoing edges of -cycles, analogous to the state in classical walks and continuous time quantum walks, which is the uniform superposition of all -cycles. In the discrete case, another possibility is to compute for some arbitrary -cycle . By then summing over , we find that gives the instantaneous probability of observing in the first register after steps. In both cases, we start from the state , which is the uniform superposition over all outgoing edges from the identity permutation. This choice of initial state is consistent with our definition of the average mixing matrix and is symmetric with respect to the generating set . Unfortunately, to compute the aforementioned quantity, we need to have direct access to the entries of the irreps, as the final state does not span all the basis vectors of a conjugacy class. Thus, we make do with computing the former expression. We discuss this issue briefly at the end of this section.
This section is organized as follows. First, in Section 5.1, we derive expressions for the ’s using the character theory of . In Section 5.2, we prove Theorem 1. Finally, we discuss the issue related to computing in Section 5.3.
5.1 Computing
Here, we derive expressions for , which will be used later for bounding the probabilities. Recall that our walk takes place on , derived from , where is the class of all transpositions. Furthermore, the entries of the discriminant matrix form a class function over (equation 1). Specifically we define:
In the expression for characters, we will use instead of to denote the class of transpositions going forward. The above definition of simplifies the expression for the eigenvalues given in equation 2 as follows:
(5) |
Next, we set out to compute the values of and the dimension of . Let , where . It is known that [31]:
(6) |
Fortunately, we only need to compute these quantities for a subset of partitions (’s) of , which will greatly simplify our analysis. Specifically, we assume , where is the collection of partitions of having the following property: There exist four non-negative integers , and such that . Here, we allow , but in that case, can simply be written as (that is, ), and we say . The following fact is a simple application of the Murnaghan-Nakayama rule (see for example Chapter 4 in [32]).
Fact 1.
Let set of all -cycles of . Then
Let (where ) be the conjugacy class of all permutations with one cycle of length and another of length . Then, the following two facts are immediately apparent:
Fact 2.
If and then .
Fact 3.
If for some and then .
Fact 4.
Size of the set is .
Again using the Murnaghan-Nakayama rule we get:
Fact 5.
Although we do not need to know in order to compute we will need this for our analysis in later sections.
Lemma 5.
Let then
Proof.
The proof of the lemma follows directly from the application of the hook-length formula, given by:
where represents the hook-length of the cell in the Young diagram of the partition . The case is straightforward, and we will focus on deriving the latter case. Since , we can express as , where . Below, we provide the explicit hook lengths for a cell , from which the lemma immediately follows.
∎
Lemma 6.
If and then .
Proof.
Given , it is necessary for the partition to be in in order for to be non-zero, as any valid rim-hook tableaux filled using the labels from (according to the composition ) must take the form of partitions in . As it turns out, for any such shape , there is at most one way to create a rim-hook tableau . In other words, if exists, and otherwise.
∎
Lemma 7.
If then
Proof.
Lemma 8.
If then for any constant we have
Proof.
Substituting in Equation 6 we get:
Hence,
(7) |
where . Suppose that for , and let , , with and , where . Note that . Here, we ignore the fact that the terms may not be integers, as it does not affect our analysis. Finally, by substituting Stirling’s approximation formula for factorials (), we obtain the claimed bound, as shown below.
In the above derivation, denotes that the quantity on the left-hand side is less than or approximately equal to the quantity on the right-hand side up to a multiplicative constant. We set . To justify the last inequality, consider the fact that , which implies . Hence, taking completes the derivation. Now, for the fraction on the right-hand side of Equation 7, we observe that the absolute value of the denominator is
and the numerator is
Hence, we conclude that , which proves the lemma. ∎
Lemma 9.
If and , then for any non-trivial singular value, we have .
5.2 Computing the divergence from the uniform distribution
Let be the normalized vector (in the norm) corresponding to the distribution over , which is uniformly supported over the -cycles. Let represent the uniform distribution over elements of . Their inner product is given by . For some normalized distribution vector , the quantity serves as a measure of the distance between and . In the quantum walk setting, we can use the state in place of . We aim to determine the norm of the projection of onto , as discussed earlier. In this section, we set out to explicitly determine this projection.
Lemma 10 (projection lemma).
Given and as defined earlier the following holds for the operator :
-
1.
-
2.
-
3.
-
4.
where .
where the last sum is over the conjugacy classes of .
Proof.
We only prove the relations as the first one is easy to check. In what follows we denote . Recall, , where . Then,
Hence,
Similarly,
and
Let, . Since is a class function, . Hence, is a class function. Thus,
where the last sum is taken over the conjugacy classes of , and denotes the Fourier transform of (as defined in Section 2.1.3). ∎
Let,
(9) |
Deriving the exact expression for is quite challenging. Instead, we will attempt to find asymptotic bounds for them, which will serve our intended purpose.
Lemma 11.
If is the class of transpositions and is as defined in Lemma 10,
Proof.
The proof follows directly from an application of Fact 2 and a straightforward counting argument. ∎
Let , and we identify the elements of with . In equation 9, we only need to consider the sum over . The following lemma is a consequence of Lemma 11.
Lemma 12.
For any , we have , where
Proof.
Note that except when is even and , in that case, . ∎
If we can be more specific.
Corollary 13.
If and then
In the next lemma, we show that the component can be ignored in the spectral decomposition of the walk operator. This greatly simplifies the analysis and allows us to focus on the action of on the space orthogonal to the trivial singular values of .
Lemma 14.
Given as defined earlier .
Proof.
We have . Since ,
In either case, we have . Intuitively, this is because has positive support only on the conjugacy class , while has support solely on in their first register, respectively. ∎
From the above lemma, it follows that . Finally, we are ready to prove our main theorem.
See 1
Proof.
Using the spectral idempotents in the decomposition of from Equation 4, we derive the following intermediate terms.
Similarly, we have
and
We can rewrite as:
If where then,
and | |||
thus | |||
where and
Else if then:
and | |||
Now,
Let and . To compute we only need to sum over . Thus,
(10) |
where,
Next, we use Lemmas 6 through 9:
-
1.
Since each (by Lemma 6), we have .
-
2.
From Lemma 8, we have (for any ).
-
3.
Additionally, Lemma 9 yields for and for any .
Thus for any ,
(11) |
We also have,
for . Hence,
(12) |
Using Fact 4 and substituting the expression for the left-hand side of Equations 11 and 12 into equation 5.2, we obtain:
Thus,
∎
5.3 Computing the probability of observing for some
Here, we provide a brief discussion on why the above analysis fails (at least when applied directly) if we consider determining the probability of detecting a basis state , where is an -cycle. The key to our analysis was the projection lemma (Lemma 10). However, in this case, we are not dealing solely with class functions, which implies explicit computation of the irreps that lack straightforward analytical expressions. More specifically, we wish to compute . Additionally, due to our choice of the starting state, the individual probabilities do not depend on . Now,
However, the function is not a class function, as can be easily seen. Thus, we cannot use the projection lemma in the manner we did earlier without knowing the irreps explicitly.
References
- [1] Acevedo, O.L., Gobron, T.: Quantum walks on cayley graphs. Journal of Physics A: Mathematical and General 39(3), 585 (2005)
- [2] Aharonov, D., Ambainis, A., Kempe, J., Vazirani, U.: Quantum walks on graphs. In: Proceedings of the thirty-third annual ACM symposium on Theory of computing. pp. 50–59 (2001)
- [3] Aharonov, Y., Davidovich, L., Zagury, N.: Quantum random walks. Physical Review A 48(2), 1687 (1993)
- [4] Aliferis, G., Leontaris, G., Vlachos, N.: Sl (2, 7) representations and their relevance to neutrino physics. The European Physical Journal C 77(6), 1–11 (2017)
- [5] Ambainis, A.: Quantum random walks–new method for designing quantum algorithms. In: International Conference on Current Trends in Theory and Practice of Computer Science. pp. 1–4. Springer (2008)
- [6] Apers, S.: Expansion testing using quantum fast-forwarding and seed sets. Quantum 4, 323 (2020)
- [7] Babai, L.: Local expansion of vertex-transitive graphs and random generation in finite groups. In: Proceedings of the twenty-third annual ACM symposium on Theory of computing. pp. 164–174 (1991)
- [8] Banerjee, A.: Discrete quantum walks on the symmetric group. arXiv preprint arXiv:2203.15148 (2022)
- [9] Barak, B., Chou, C.N., Gao, X.: Spoofing linear cross-entropy benchmarking in shallow quantum circuits. arXiv preprint arXiv:2005.02421 (2020)
- [10] Bouland, A., Fefferman, B., Nirkhe, C., Vazirani, U.: Quantum supremacy and the complexity of random circuit sampling. arXiv preprint arXiv:1803.04402 (2018)
- [11] Bump, D., Diaconis, P., Hicks, A., Miclo, L., Widom, H.: An exercise (?) in fourier analysis on the heisenberg group. In: Annales de la Faculté des sciences de Toulouse: Mathématiques. vol. 26, pp. 263–288 (2017)
- [12] Chan, A., Zhan, H.: Pretty good state transfer in discrete-time quantum walks. Journal of Physics A: Mathematical and Theoretical 56(16), 165305 (2023)
- [13] Coutinho, G., Godsil, C.: Graph spectra and continuous quantum walks. Unpublished notes (2021)
- [14] Dai, W., Yuan, J., Li, D.: Discrete-time quantum walk on the cayley graph of the dihedral group. Quantum Information Processing 17(12), 1–21 (2018)
- [15] Diaconis, P.: Group representations in probability and statistics. Lecture notes-monograph series 11, i–192 (1988)
- [16] Diaconis, P., Shahshahani, M.: Generating a random permutation with random transpositions. Zeitschrift für Wahrscheinlichkeitstheorie und verwandte Gebiete 57(2), 159–179 (1981)
- [17] D’Ariano, G.M., Erba, M., Perinotti, P., Tosini, A.: Virtually abelian quantum walks. Journal of Physics A: Mathematical and Theoretical 50(3), 035301 (2016)
- [18] Farhi, E., Gutmann, S.: Quantum computation and decision trees. Physical Review A 58(2), 915 (1998)
- [19] Floratos, E., Leontaris, G.: Discrete flavour symmetries from the heisenberg group. Physics Letters B 755, 155–161 (2016)
- [20] Gerhardt, H., Watrous, J.: Continuous-time quantum walks on the symmetric group. In: Approximation, Randomization, and Combinatorial Optimization.. Algorithms and Techniques, pp. 290–301 (2003)
- [21] Godsil, C.: Average mixing of continuous quantum walks. Journal of Combinatorial Theory, Series A 120(7), 1649–1662 (2013)
- [22] Kendon, V.: Decoherence in quantum walks–a review. Mathematical Structures in Computer Science 17(6), 1169–1220 (2007)
- [23] Knittel, M., Bassirian, R.: Quantum random walks on cayley graphs
- [24] Lee, J.R., Naor, A.: Lp metrics on the heisenberg group and the goemans-linial conjecture. In: 2006 47th Annual IEEE Symposium on Foundations of Computer Science (FOCS’06). pp. 99–108. IEEE (2006)
- [25] Lovász, L.: Random walks on graphs. Combinatorics, Paul erdos is eighty 2(1-46), 4 (1993)
- [26] Magniez, F., Nayak, A., Roland, J., Santha, M.: Search via quantum walk. SIAM journal on computing 40(1), 142–164 (2011)
- [27] Meyer, D.A.: From quantum cellular automata to quantum lattice gases. Journal of Statistical Physics 85(5), 551–574 (1996)
- [28] Moore, C., Russell, A.: Quantum walks on the hypercube. In: International Workshop on Randomization and Approximation Techniques in Computer Science. pp. 164–178 (2002)
- [29] Nayak, A., Vishwanath, A.: Quantum walk on the line. arXiv preprint quant-ph/0010117 (2000)
- [30] Osborne, T.J., Severini, S.: Quantum algorithms and covering spaces. arXiv preprint quant-ph/0403127 (2004)
- [31] RE Ingram, S.: Some characters of the symmetric group. Proceedings of the American Mathematical Society pp. 358–369 (1950)
- [32] Sagan, B.E.: The symmetric group: representations, combinatorial algorithms, and symmetric functions, vol. 203. Springer Science & Business Media (2013)
- [33] Sarkar, R.S., Adhikari, B.: Discrete-time quantum walks on cayley graphs of dihedral groups using generalized grover coins. arXiv preprint arXiv:2309.15194 (2023)
- [34] Sorci, J.: Average mixing in quantum walks of reversible markov chains. arXiv preprint arXiv:2211.02037 (2022)
- [35] Szegedy, M.: Quantum speed-up of markov chain based algorithms. In: 45th Annual IEEE symposium on foundations of computer science. pp. 32–41. IEEE (2004)
- [36] Terras, A.: Fourier analysis on finite groups and applications. No. 43, Cambridge University Press, Cambridge (1999)
- [37] Venegas-Andraca, S.E.: Quantum walks: a comprehensive review. Quantum Information Processing 11(5), 1015–1106 (2012)
- [38] Wallace, D.: G. james and a. kerber, the representation theory of the symmetric group (encyclopedia of mathematics and its applications, vol. 16, addison-wesley, reading, mass., 1981), pp. 510,£ 37.80. Proceedings of the Edinburgh Mathematical Society 27(1), 103–104 (1984)
- [39] Zack, M.: Measuring randomness and evaluating random number generators using the finite heisenberg group. Limit theorems in probability and statistics (Pécs, 1989) 57, 537–544 (1990)