Quantum Expander Mixing Lemma and its Converse
Abstract
Expander graphs are fundamental in both computer science and mathematics, with a wide array of applications. With quantum technology reshaping our world, quantum expanders have emerged, finding numerous uses in quantum information theory. The classical expander mixing lemma plays a central role in graph theory, offering essential insights into edge distribution within graphs and aiding in the analysis of diverse network properties and algorithms. This paper establishes the quantum analogue of the classical expander mixing lemma and its converse for quantum expanders.
1 Introduction
Expander graphs are foundational in both computer science and mathematics, offering diverse applications across these domains (Hoory et al.,, 2006; Lubotzky,, 2012). Leveraging their expansion property, these graphs contribute significantly to algorithmic innovations, cryptographic protocols, analysis of circuit and proof complexity, development of derandomization techniques and pseudorandom generators, as well as the theory of error-correcting codes. Additionally, expander graphs play pivotal roles in shaping structural insights within group theory, algebra, number theory, geometry, and combinatorics. As quantum technology is revolutionizing the world, quantum expanders were introduced and have foundd numerous applications in the field of quantum information theory (Ambainis and Smith,, 2004; Ben-Aroya et al.,, 2008; Hastings, 2007a, ; Hastings and Harrow,, 2009; Aharonov et al.,, 2014).
Quantum expanders are the quantum extension of expander graphs, described by means of operators that map quantum states to quantum states. A general quantum state is a density matrix, which is a trace-one and positive semidefinite operator, i.e.,
with being some orthonormal basis of a Hilbert space . Notice that . An admissible quantum transformation is any transformation that can be implemented by a quantum circuit, with unitary operators and measurements. Thus, admissible quantum operators map density matrices to density matrices; see, Nielsen and Chuang, (2001); Kitaev et al., (2002).
In this paper, we rigorously define quantum expanders following Hastings, 2007b and Pisier, (2014). In essence, we consider a quantum operator in the following normalized form (dividing by ):
(1.1) |
where stands for the space of complex matrices, is a -tuple of unitary matrices with being a positive integer that is independent of , and the superscript represents the conjugate transpose. We have the operator norm and , where is the identity matrix. In the quantum context, the spectral gap is delineated by the reduced spectral radius , where is restricted to the orthogonal complement of the identity matrix. A quantum expander sequence comprises those having reduced spectral radius smaller than uniformly for all , as tends to infinity with .
The aim of this paper is to establish the quantum counterpart of the classical expander mixing lemma and its converse. The expander mixing lemma intuitively states that the edges of certain -regular graphs are evenly distributed throughout the graph. Specifically, the normalized number of edges connecting two vertex subsets and is always close to the expected number of edges between them in a random -regular graph, which is . This well-known result is demonstrated, for example, in Corollary 9.2.5 of Alon and Spencer, (2000). The expander mixing lemma plays a pivotal role in graph theory, providing crucial insights into the distribution of edges within graphs and facilitating the analysis of various network properties and algorithms. Variants of the expander mixing lemma have been proposed. For instance, Chen et al., (2013) established an operator version of the expander mixing lemma, utilizing it iteratively for bias amplification. This approach was further employed in Jeronimo et al., (2022) to formulate the iterated operator expander mixing lemma.
We establish the expander mixing lemma for quantum expanders, termed the quantum expander mixing lemma, for the first time to the authors’ best knowledge. This lemma asserts that for any two orthogonal projections , the Hilbert–Schmidt inner product of and is always close to for all , for being in the normalized form given in (1.1). In quantum mechanics, the Hilbert-Schmidt inner product of two matrices (which symbolize quantum states) serves as a measure of the similarity between these states. A substantial inner product implies a significant degree of similarity or correlation between the states, whereas a small inner product indicates dissimilarity or orthogonality.
Studying the converse of the expander mixing lemma enhances the understanding of graph properties and their implications in diverse fields, leading to more robust theories, algorithms, and applications. The converse of the expander mixing lemma has been articulated in two distinct ways. The first way involves investigating the upper bound of the second-largest (in absolute value) eigenvalue of the -regular graph, when the normalized number of edges between two vertex subsets and is always close to , as demonstrated in Bilu and Linial, (2006). The second way examines whether there exist vertex subsets and such that the normalized number of edges in between, compared to , cannot fall below a certain threshold, as shown in Lev, (2015).
In this paper, we further establish the converse of the expander mixing lemma for quantum expanders, which is the first result to the best of the authors’ knowledge. Following the second way, our focus lies on exploring the bounds of that difference. Specifically, our quantum expander mixing lemma states that there exist two orthogonal projections such that the difference of the Hilbert–Schmidt inner product of and to cannot be small up to a certain threshold for all .
The rest of the paper proceeds as follows. In Section 2, we formally define expander graphs and then present the classical expander mixing lemma. In Section 3, we rigorously define quantum expander graphs and then state our quantum expander mixing lemma along with its converse. Technical proofs of this paper are provided in Section 4. Throughout the paper, denotes the Hilbert–Schmidt norm when applied to a matrix and the operator norm when applied to an operator, with the specific meaning determined by the context.
2 Expander mixing lemma
We start by defining formally the concept of an expander graph and then state the expander mixing lemma in Theorem 2.3. There exist several equivalent definitions of expander graphs, which can be characterized through vertex, edge, or spectral expansion. We adopt the spectral perspective towards expander graphs, defining them in terms of a certain spectral gap by means of Markov operators. We refer interested readers to Tao, (2015) for further details.
We consider an undirected graph , where is the set of vertices and is the set of edges. A graph is finite if is finite, and hence is finite. We consider as a -regular graph, where each vertex of is contained in exactly edges in ; we say that is the degree of the regular graph . We let be the finite-dimensional complex Hilbert space of functions with norm and inner product given respectively as
where the overline notation represents the complex conjugate. We then define the adjacency operator on functions as the sum of over all of the neighbours of , i.e.,
Clearly, is a linear operator and one can associate it with the adjacency matrix.
Since is -regular, a linear algebraic definition of expansion is possible based on the eigenvalues of the adjacency matrix . For being undirected, is real symmetric. It is known that for being a -regular graph having vertices, the spectral theorem implies that has real-valued eigenvalues and these eigenvalues are in . Because is regular, the uniform distribution , with entry for all , is its stationary distribution. Thus, is an eigenvector of with eigenvalue , i.e., . The spectral gap of is defined to be , which measures the spectral expansion.
The normalized variants of these definitions are commonly employed and offer greater convenience when displaying certain results. Consider the matrix , which is the Markov transition matrix of , and its eigenvalues are in . It is associated with the Markov operator defined by
(2.3) |
where is the Kronecker delta function on . Then, is self-adjoint (equal to its conjugate transpose ) and the operator norm . Denote for the constant function , and then . We can observe the reduction in the operator norm of when it is restricted to the orthogonal complement .
Definition 2.1.
Define the reduced spectral radius of as the restricted operator norm
In this paper, we assume graph is connected, and then the largest eigenvalue of has multiplicity . We assume is non-bipartite, and then is not an eigenvalue of . Therefore, the reduced spectral radius is always smaller than . We identify with the set of vertices. A sequence is a sequence of finite -regular simple connected graphs with the cardinality of the vertex set . For a given sequence of -regular graphs , the reduced spectral radius for each is always smaller than . The operator is sometimes known as the (combinatorial or graph) Laplacian. This is a positive semidefinite operator with at least one zero eigenvalue. To inquire whether a uniform gap exists between and , here is the definition of the expander sequence.
Definition 2.2.
A graph is called an expander if there is a spectral gap of size in , in the sense that the first eigenvalue of exceeds the second by at least , equivalently . A sequence of -regular graphs is called an expander sequence if there exists such that for all .
We have defined the notion of an expander sequence, which is a family of graphs instead of an individual graph. It is also commonly defined through the -graph, which is a -regular graph with vertices such that all of the eigenvalues of its adjacency matrix except one have absolute value at most , i.e., . If we fix and then -graphs form a family of expander graphs with a constant spectral gap. The expander mixing lemma states the number of edges between two vertex subsets and is always close to the expected number of edges between them in a random -regular graph, namely ; see, for instance, Corollary 9.2.5 of Alon and Spencer, (2000).
Theorem 2.3 (Expander mixing lemma, Alon and Spencer, (2000)).
Consider the expander sequence defined in accordance to Definition 2.2. For any two subsets , let
be the number of edges between and (counting edges contained in the intersection of and twice). Then
A converse to the above well-known expander mixing lemma is provided in Corollary 4 of Lev, (2015).
3 Main results
In this section, we define formally the concept of quantum expanders following Hastings, 2007b and Pisier, (2014). Before that, it is worth providing the motivation from classical expanders and explaining how quantum expanders appear as a non-commutative version of the classical ones.
The story could be started with the Cayley graph. Consider as a finite group generated by . Suppose is symmetric in the sense that whenever and does not contain the identity to avoid loops. The Cayley graph is defined as the graph with vertex set and edge set . Each vertex is connected to the elements for and hence is a -regular graph. A unitary representation of is the Hilbert space , together with a homomorphism from to the group of unitary transformations on . Define the (left) regular unitary representation of as such that
for and . Unitary operator is a unitary induced by the permutation of vertices of the graph.
To discuss expanders, let us assume that we are given a sequence of Cayley graphs , where for each with being a positive integer independent of . Then where
is an expander sequence if and only if the sequence of -tuples of unitaries
satisfies that with ,
where is the Hilbert–Schmidt norm. This is equivalent to the condition that there exists a sequence of -tuples of unitaries
such that for any , some , and any being a diagonal complex matrix,
(3.1) |
where stands for the set of unitary matrices.
The term “quantum expander” is just to designate unitaries to satisfy (3.1) for a general beyond being diagonal, where stands for the space of complex matrices. In this light, quantum expanders can be seen as a non-commutative version of the classical ones. Identifying with the space of bounded operators on the -dimensional Hilbert space , different to the Markov operator defined in (2.3), in the quantum setting we consider quantum operators act on of the following form for each :
where is the complex conjugate of the matrix . Identifying as usual as , the Hilbert space obtained by equipping with the corresponding scalar product and the Hilbert–Schmidt norm. We can write
Then we may consider as an operator acting on defined by
which satisfies that
Analogous to the classical setting, we have the operator norm and , where is the identity matrix. Then we ask how much the operator norm of the quantum operator would be decreased if it is restricted to the orthogonal complement of the identity matrix. For that purpose, we give the quantum version of Definition 2.1.
Definition 3.1.
Define the reduced spectral radius as the restricted operator norm
The definition of a quantum expander is defined in terms of . It says that there exists such that for all , which is in consistent with the classical expander sequence in Definition 2.2.
Definition 3.2 (Hastings, 2007b and Pisier, (2014)).
Fix a positive integer and a sequence of positive integers with as . We say a sequence of -tuples of unitaries
is a quantum expander sequence if their reduced spectral radius is smaller than uniformly for all .
The following two theorems are our main results of this paper.
Theorem 3.3 (Quantum expander mixing lemma).
Consider a quantum expander sequence defined in accordance to Definition 3.2. For any two orthogonal projections , we have that for all ,
where the Hilbert–Schmidt inner product of two matrices and is defined as usual
Theorem 3.4 (Converse of quantum expander mixing lemma).
Consider a quantum expander sequence defined in accordance to Definition 3.2. There exist two universal constants and two non-zero orthogonal projections , such that for all ,
4 Proofs of the paper
Proof of Theorem 3.3.
Let be the orthogonal projection onto the space . Then
Therefore,
We have by the Cauchy-Schwarz inequality and properties of the unitaries that
∎
To prove Theorem 3.4, we commence by introducing the Schatten norm. In mathematics, particularly in functional analysis, the Schatten norm, also known as the Schatten–von-Neumann norm, emerges as a generalization of p-integrability, akin to the trace class norm and the Hilbert–Schmidt norm. For , define the Schatten -norm of as
for , the convention is to define as the operator norm
The Schatten norms are unitarily invariant in the sense that for unitaries and ,
(4.1) |
Considering as the dual index to , the noncommutative Hölder’s inequality says that for all such that , we have that for any matrices ,
Its special case is the well-known inequality
(4.2) |
where is called the trace norm or the nuclear norm, is called the Frobenius norm or the Hilbert–Schmidt norm, and is called the spectral norm or the operator norm. Lev, (2015) defined height (Definition 4.1) of a matrix using these three important norms; by (4.2) the height of any matrix is at least one.
Definition 4.1 (Lev, (2015)).
Define the height of a non-zero matrix by
We define the Schatten norm induced operator norm as for operator . Then the three special cases and give operator norms , and , respectively. Then we define a linear map for unitaries as follows:
(4.3) |
where is the diagonal matrix . Here, we follow the convention that when the operator is applied to a vector, it yields the corresponding diagonal matrix, and when applied to a matrix, it produces the corresponding diagonal vector. Then we have
(4.4) |
The next lemma is a special case of Gillespie, (1991) on page 226 therein.
Lemma 4.2 (Gillespie, (1991)).
For any linear operator , we have
Similar to the definition of height for a matrix in Definition 4.1, we give definition of the height of a linear operator here; by Lemma 4.2 the height of any linear operator is at least one.
Definition 4.3.
For any linear operator , define its height by
Recall that in the classical setting, for an operator , we have
In the quantum setting, denote as the adjoint of with respect to the Hilbert–Schmidt inner product. For , we have
which yields that
Hence, we have
(4.5) | ||||
By Definition 4.3,
(4.6) |
In the classical setting, Lev, (2015) defined the logarithmic diameter for a non-zero vector by
(4.7) |
In the quantum setting, considering complex matrices, we define the logarithmic diameter as follows.
Definition 4.4.
For , let
be all the non-zero (strictly positive) eigenvalues of . We define the logarithmic diameter of by
The following lemma will serve as the cornerstone for the subsequent proofs.
Lemma 4.5.
If is a linear map with for a real number, then there exists a non-zero matrix such that
Proof.
By the definition of given in (4) where and are unitaries in , together with (4.4), we have
(4.8) |
which yields that
Then by the definition of the height of a matrix given in Definition 4.1, together with (4.4), we have
Therefore, by the definition of height of a linear operator given in Definition 4.2, we have
By Lemma 1 of Lev, (2015), there exists such that
(4.9) |
and then the logarithmic diameter for a non-zero vector defined in (4.7)
Taking the matrix
(4.10) |
with being the diagonal matrix , the logarithmic diameter for a non-zero matrix defined in Definition 4.4
Note that
Recall that by the definition of given in (4),
using the specific form of . By equation (4.9),
At last, using equations (4.1) and (4.8), we have by the specific form of that
which completes the proof. ∎
Lemma 4.6.
If is a matrix with for a real number, then there exists a non-zero orthogonal projection such that
Proof.
We proceed the proof in two steps. In the first step, we establish the desired result for a self-adjoint matrix. In the second step, we extend that to a general matrix.
Step 1. Suppose is a self-adjoint matrix with height for a real number, where is given in Definition 4.1. Since is self-adjoint, there exists a unitary such that
By the definition of height and the unitarily invariant property (4.1), we have
Next, we have
where the superscript represents the transpose, and the norms applied to vectors or matrices by their corresponding definitions. Hence,
Then by Lemma 4 of Lev, (2015), there exists an -dimensional binary vector such that
(4.11) |
where, for non-zero vectors ,
Note that
and given that the trace is invariant under circular shifts
Set
(4.12) |
which is an orthogonal projection in . Then by equation (4.11), we have that
(4.13) |
Step 2. According to the Toeplitz decomposition, each can be written uniquely as
in which are self-adjoint matrices; see Horn and Johnson, (2012). Since
we have
(4.14) |
Additionally, for the Hilbert-Schmidt norm, we have
Replacing by if necessarily, we may assume , which gives
Then, by the definition of height given in Definition 4.1, equation (4.14), and the condition that ,
By the result of Step 1, specifically (4.13), there exists an orthogonal projection such that
Combining this with
yields the desired estimate. ∎
Theorem 4.7.
For any linear map , there exists a non-zero orthogonal projection such that
Proof.
The second inequality clearly holds by the definition of , since being a non-zero projection is a special case of a complex matrix. Now, we prove the first inequality.
Suppose for a real number. By Lemma 4.5, there exists a non-zero matrix such that
(4.15) |
Furthermore, by equations (4.5) and (4.6), we have
(4.16) |
Consider the spectral decomposition of as
where is the set of eigenvalues of and is the corresponding spectral projection. Then, for , the Schatten -norm of can be written as
while for , it can be written as
Now, choose such that
By the definition of given in Definition 4.4 which is defined in terms of eigenvalues of (instead of ), together with equation (4.15), we obtain
Furthermore, by equation (4.15),
Therefore,
By Lemma 4.6, there exists an orthogonal projection such that
Therefore, by equation (4.16),
which yields that
as desired.
∎
Theorem 4.8.
There exist two universal constants and two non-zero orthogonal projections , such that for any linear map , we have
Proof.
The second inequality clearly holds by the definition of , because the non-zero projections and are specific instances of complex matrices. Now, we prove the first inequality.
Suppose for a real number. Define
Theorem 4.7 states that there exists an orthogonal projection such that
(4.17) |
which yields that
Note that, for an orthogonal projection , we have
Then we have
With Theorem 4.8, we are finally ready to establish the converse of quantum expander mixing lemma.
Proof of Theorem 3.4.
As in the proof of Theorem 3.3, we let be the orthogonal projection onto the space . Then is the restriction of onto .
According to the definition of a quantum expander sequence given in Definition 3.2, there exists such that
Here, we consider as the largest of those constants that suffice the above inquality. Then by the definition of , we have
Therefore,
By Theorem 4.8, there exist two universal constants and two non-zero orthogonal projections , such that
Noting that
we have and satisfies
where are two universal constants. ∎
Acknowledgments
This research project was partially supported by NIH grant 1R21AI180492-01 and the Individual Research Grant at Texas A&M University. The author gives special thanks to Ryo Toyota for conversations on this subject.
References
- Aharonov et al., (2014) Aharonov, D., Harrow, A. W., Landau, Z., Nagaj, D., Szegedy, M., and Vazirani, U. (2014). Local tests of global entanglement and a counterexample to the generalized area law. In 2014 IEEE 55th Annual Symposium on Foundations of Computer Science, pages 246–255. IEEE.
- Alon and Spencer, (2000) Alon, N. and Spencer, J. H. (2000). The Probabilistic Method, second edition. John Wiley & Sons.
- Ambainis and Smith, (2004) Ambainis, A. and Smith, A. (2004). Small pseudo-random families of matrices: Derandomizing approximate quantum encryption. In Approximation, Randomization, and Combinatorial Optimization. Algorithms and Techniques: 7th International Workshop on Approximation Algorithms for Combinatorial Optimization Problems, APPROX 2004, and 8th International Workshop on Randomization and Computation, RANDOM 2004, Cambridge, MA, USA, August 22-24, 2004. Proceedings, pages 249–260. Springer.
- Ben-Aroya et al., (2008) Ben-Aroya, A., Schwartz, O., and Ta-Shma, A. (2008). Quantum expanders: Motivation and constructions. In 2008 23rd Annual IEEE Conference on Computational Complexity, pages 292–303. IEEE.
- Bilu and Linial, (2006) Bilu, Y. and Linial, N. (2006). Lifts, discrepancy and nearly optimal spectral gap. Combinatorica, 26(5):495–519.
- Chen et al., (2013) Chen, S., Moore, C., and Russell, A. (2013). Small-bias sets for nonabelian groups: derandomizations of the Alon-Roichman theorem. In International Workshop on Approximation Algorithms for Combinatorial Optimization, pages 436–451. Springer.
- Gillespie, (1991) Gillespie, T. (1991). Noncommutative variations on theorems of Marcel Riesz and others. In PAUL HALMOS Celebrating 50 Years of Mathematics, pages 221–236. Springer.
- (8) Hastings, M. B. (2007a). Entropy and entanglement in quantum ground states. Physical Review B, 76(3):035–114.
- (9) Hastings, M. B. (2007b). Random unitaries give quantum expanders. Physical Review A, 76(3):032315.
- Hastings and Harrow, (2009) Hastings, M. B. and Harrow, A. W. (2009). Classical and quantum tensor product expanders. arXiv preprint arXiv:0804.0011, 9(3):336–360.
- Hoory et al., (2006) Hoory, S., Linial, N., and Wigderson, A. (2006). Expander graphs and their applications. Bulletin of the American Mathematical Society, 43(4):439–561.
- Horn and Johnson, (2012) Horn, R. A. and Johnson, C. R. (2012). Matrix analysis. Cambridge university press.
- Jeronimo et al., (2022) Jeronimo, F. G., Mittal, T., Roy, S., and Wigderson, A. (2022). Almost Ramanujan expanders from arbitrary expanders via operator amplification. In 2022 IEEE 63rd Annual Symposium on Foundations of Computer Science (FOCS), pages 378–388. IEEE.
- Kitaev et al., (2002) Kitaev, A. Y., Shen, A., and Vyalyi, M. N. (2002). Classical and quantum computation. Number 47. American Mathematical Soc.
- Lev, (2015) Lev, V. F. (2015). Discrete norms of a matrix and the converse to the expander mixing lemma. Linear Algebra and its Applications, 483:158–181.
- Lubotzky, (2012) Lubotzky, A. (2012). Expander graphs in pure and applied mathematics. Bulletin of the American Mathematical Society, 49(1):113–162.
- Nielsen and Chuang, (2001) Nielsen, M. A. and Chuang, I. L. (2001). Quantum computation and quantum information, volume 2. Cambridge university press Cambridge.
- Pisier, (2014) Pisier, G. (2014). Quantum expanders and geometry of operator spaces. Journal of the European Mathematical Society, 16(6):1183–1219.
- Tao, (2015) Tao, T. (2015). Expansion in finite simple groups of Lie type, volume 164. American Mathematical Soc.