This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Quantum Expander Mixing Lemma and its Converse

Ning Ninglabel=e1][email protected] [ Department of Statistics, Texas A&M University.
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.

\startlocaldefs\endlocaldefs

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.,

ρ=vpv|ψvψv|,0pv1,vpv=1,\rho=\sum\limits_{v}p_{v}\left|\psi_{v}\right\rangle\left\langle\psi_{v}\right|,\qquad 0\leq p_{v}\leq 1,\qquad\sum\limits_{v}p_{v}=1,

with {ψv}\left\{\psi_{v}\right\} being some orthonormal basis of a Hilbert space 𝒱\mathcal{V}. Notice that ρL(𝒱):=Hom(𝒱,𝒱)\rho\in L(\mathcal{V}):=\operatorname{Hom}(\mathcal{V},\mathcal{V}). An admissible quantum transformation F:L(𝒱)L(𝒱)F:L(\mathcal{V})\rightarrow L(\mathcal{V}) 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 Tn:M(Nn)M(Nn)T_{n}:M(N_{n})\rightarrow M(N_{n}) in the following normalized form (dividing by dd):

Tn(η)=1dj=1duj(n)ηuj(n),\displaystyle T_{n}(\eta)=\frac{1}{d}\sum_{j=1}^{d}u_{j}^{(n)}\eta u_{j}^{(n)*}, (1.1)

where M(Nn)M(N_{n}) stands for the space of Nn×NnN_{n}\times N_{n} complex matrices, (u1(n),,ud(n))\big{(}u_{1}^{(n)},\cdots,u_{d}^{(n)}\big{)} is a dd-tuple of unitary matrices with dd being a positive integer that is independent of nn, and the superscript ``"``\ast" represents the conjugate transpose. We have the operator norm Tn=1\|T_{n}\|=1 and Tn(INn)=INnT_{n}(I_{N_{n}})=I_{N_{n}}, where INnM(Nn)I_{N_{n}}\in M(N_{n}) is the identity matrix. In the quantum context, the spectral gap is delineated by the reduced spectral radius Tn|INn\big{\|}T_{n}|_{I_{N_{n}}^{\perp}}\big{\|}, where TnT_{n} is restricted to the orthogonal complement INnI_{N_{n}}^{\perp} of the identity matrix. A quantum expander sequence comprises those having reduced spectral radius smaller than 11 uniformly for all nn, as NnN_{n} tends to infinity with nn.

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 dd-regular graphs are evenly distributed throughout the graph. Specifically, the normalized number of edges connecting two vertex subsets S1S_{1} and S2S_{2} is always close to the expected number of edges between them in a random dd-regular graph, which is 1n|S1||S2|\frac{1}{n}|S_{1}||S_{2}|. 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 P1,P2M(Nn)P_{1},P_{2}\in M(N_{n}), the Hilbert–Schmidt inner product of P1P_{1} and TnP2T_{n}P_{2} is always close to 1Nntr(P1)tr(P2)\frac{1}{N_{n}}{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2}) for all nn, for TnT_{n} 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 dd-regular graph, when the normalized number of edges between two vertex subsets S1S_{1} and S2S_{2} is always close to 1n|S1||S2|\frac{1}{n}|S_{1}||S_{2}|, as demonstrated in Bilu and Linial, (2006). The second way examines whether there exist vertex subsets S1S_{1} and S2S_{2} such that the normalized number of edges in between, compared to 1n|S1||S2|\frac{1}{n}|S_{1}||S_{2}|, 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 P1,P2M(Nn)P_{1},P_{2}\in M(N_{n}) such that the difference of the Hilbert–Schmidt inner product of P1P_{1} and TnP2T_{n}P_{2} to 1Nntr(P1)tr(P2)\frac{1}{N_{n}}{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2}) cannot be small up to a certain threshold for all nn.

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, \|\cdot\| 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 𝒢=(V,E)\mathcal{G}=(V,E), where VV is the set of vertices and EE is the set of edges. A graph is finite if VV is finite, and hence EE is finite. We consider 𝒢\mathcal{G} as a dd-regular graph, where each vertex of VV is contained in exactly dd edges in EE; we say that dd is the degree of the regular graph 𝒢\mathcal{G}. We let 2(V)\ell^{2}(V) be the finite-dimensional complex Hilbert space of functions f:Vf:V\rightarrow\mathbb{C} with norm and inner product given respectively as

f2(V):=(vV|f(v)|2)1/2andf,g2(V):=vVf(v)g(v)¯,\|f\|_{\ell^{2}(V)}:=\Big{(}\sum_{v\in V}|f(v)|^{2}\Big{)}^{1/2}\quad\text{and}\quad\langle f,g\rangle_{\ell^{2}(V)}:=\sum_{v\in V}f(v)\overline{g(v)},

where the overline notation represents the complex conjugate. We then define the adjacency operator A:2(V)2(V)A:\ell^{2}(V)\rightarrow\ell^{2}(V) on functions f2(V)f\in\ell^{2}(V) as the sum of ff over all of the neighbours of vv, i.e.,

Af(v):=wV:{v,w}Ef(w).Af(v):=\sum_{w\in V:\{v,w\}\in E}f(w).

Clearly, AA is a linear operator and one can associate it with the adjacency matrix.

Since 𝒢\mathcal{G} is dd-regular, a linear algebraic definition of expansion is possible based on the eigenvalues of the adjacency matrix AA. For 𝒢\mathcal{G} being undirected, AA is real symmetric. It is known that for 𝒢\mathcal{G} being a dd-regular graph having nn vertices, the spectral theorem implies that AA has nn real-valued eigenvalues λ1λ2λn\lambda_{1}\geq\lambda_{2}\geq\ldots\geq\lambda_{n} and these eigenvalues are in [d,d][-d,d]. Because 𝒢\mathcal{G} is regular, the uniform distribution unu\in\mathbb{R}^{n}, with entry ui=1/nu_{i}=1/n for all i=1,,ni=1,…,n, is its stationary distribution. Thus, uu is an eigenvector of AA with eigenvalue λ1=d\lambda_{1}=d, i.e., Au=duAu=du. The spectral gap of 𝒢\mathcal{G} is defined to be dλ2d-\lambda_{2}, 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 1dA\frac{1}{d}A, which is the Markov transition matrix of 𝒢\mathcal{G}, and its eigenvalues are in [1,1][-1,1]. It is associated with the Markov operator T:2(V)2(V)T:\ell^{2}(V)\rightarrow\ell^{2}(V) defined by

Tδx,δy2(V)={1/d(x,y)E,0(x,y)E,\displaystyle\langle T\delta_{x},\delta_{y}\rangle_{\ell^{2}(V)}=\left\{\begin{array}[]{ll}1/d&(x,y)\in E,\\ 0&(x,y)\notin E,\end{array}\right. (2.3)

where δx\delta_{x} is the Kronecker delta function on xx. Then, TT is self-adjoint (equal to its conjugate transpose TT^{*}) and the operator norm T=1\|T\|=1. Denote 1V2(V)1_{V}\in\ell^{2}(V) for the constant function v1v\mapsto 1, and then T(1V)=1VT(1_{V})=1_{V}. We can observe the reduction in the operator norm of TT when it is restricted to the orthogonal complement 1V1_{V}^{\perp}.

Definition 2.1.

Define the reduced spectral radius ρ(𝒢)\rho(\mathcal{G}) of 𝒢\mathcal{G} as the restricted operator norm

ρ(𝒢):=T|1V.\displaystyle\rho(\mathcal{G}):=\big{\|}T|_{1_{V}^{\perp}}\big{\|}.

In this paper, we assume graph 𝒢\mathcal{G} is connected, and then the largest eigenvalue 11 of TT has multiplicity 11. We assume 𝒢\mathcal{G} is non-bipartite, and then 1-1 is not an eigenvalue of TT. Therefore, the reduced spectral radius ρ(𝒢)\rho(\mathcal{G}) is always smaller than 11. We identify 𝒢\mathcal{G} with the set of vertices. A sequence (𝒢n)n=1(\mathcal{G}_{n})_{n=1}^{\infty} is a sequence of finite dd-regular simple connected graphs with the cardinality of the vertex set |Vn|=n|V_{n}|=n\rightarrow\infty. For a given sequence of dd-regular graphs (𝒢n)n=1(\mathcal{G}_{n})_{n=1}^{\infty}, the reduced spectral radius ρ(𝒢n)\rho(\mathcal{G}_{n}) for each nn\in\mathbb{N} is always smaller than 11. The operator Δ:=1T\Delta:=1-T 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 ρ(𝒢n)\rho(\mathcal{G}_{n}) and 11, 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 ϵ>0\epsilon>0 in Δ\Delta, in the sense that the first eigenvalue of Δ\Delta exceeds the second by at least ϵ\epsilon, equivalently 1ρ(𝒢)ϵ1-\rho(\mathcal{G})\geq\epsilon. A sequence of dd-regular graphs (𝒢n)(\mathcal{G}_{n}) is called an expander sequence if there exists ϵ>0\epsilon>0 such that ρ(𝒢n)1ϵ\rho(\mathcal{G}_{n})\leq 1-\epsilon for all nn.

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 (n,d,λ)(n,d,\lambda)-graph, which is a dd-regular graph with nn vertices such that all of the eigenvalues of its adjacency matrix except one have absolute value at most λ\lambda, i.e., maxi1|λi|λ\max_{i\neq 1}|\lambda_{i}|\leq\lambda. If we fix dd and λ\lambda then (n,d,λ)(n,d,\lambda)-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 S1S_{1} and S2S_{2} is always close to the expected number of edges between them in a random dd-regular graph, namely dn|S1||S2|\frac{d}{n}|S_{1}||S_{2}|; 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 S1,S2VS_{1},S_{2}\subseteq V, let

e(S1,S2)=|{(x,y)S1×S2:xyE}|e(S_{1},S_{2})=\Big{|}\big{\{}(x,y)\in S_{1}\times S_{2}:xy\in E\big{\}}\Big{|}

be the number of edges between S1S_{1} and S2S_{2} (counting edges contained in the intersection of S1S_{1} and S2S_{2} twice). Then

|1de(S1,S2)1n|S1S2(1ϵ)|S1||S2|.\displaystyle\left|\frac{1}{d}e(S_{1},S_{2})-\frac{1}{n}|S_{1}||S_{2}|\right|\leq(1-\epsilon)\sqrt{|S_{1}||S_{2}|}.

A converse to the above well-known expander mixing lemma is provided in Corollary 4 of Lev, (2015).

Theorem 2.4 (Converse of expander mixing lemma, Lev, (2015)).

Consider the expander sequence defined in accordance to Definition 2.2. With e(S1,S2)e(S_{1},S_{2}) as in Theorem 2.3, there exist non-empty subsets S1,S2VS_{1},S_{2}\subseteq V such that

|1de(S1,S2)1n|S1S2(1ϵ)322(log(2/(1ϵ))+4)|S1||S2|.\displaystyle\left|\frac{1}{d}e(S_{1},S_{2})-\frac{1}{n}|S_{1}||S_{2}|\right|\geq\frac{(1-\epsilon)}{32\sqrt{2}(\log(2/(1-\epsilon))+4)}\sqrt{|S_{1}||S_{2}|}.

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 GG as a finite group generated by S={s1,,sd}S=\{s_{1},\ldots,s_{d}\}. Suppose SS is symmetric in the sense that sSs\in S whenever s1Ss^{-1}\in S and does not contain the identity 11 to avoid loops. The Cayley graph Cay(G,S)\operatorname{Cay}(G,S) is defined as the graph with vertex set GG and edge set {{sx,x}:xG,sS}\{\{sx,x\}:x\in G,s\in S\}. Each vertex xGx\in G is connected to the |S|=d|S|=d elements sxsx for sSs\in S and hence Cay(G,S)\operatorname{Cay}(G,S) is a dd-regular graph. A unitary representation of GG is the Hilbert space 2(G)\ell^{2}(G), together with a homomorphism ρ:GU(2(G))\rho:G\rightarrow U(\ell^{2}(G)) from GG to the group U(2(G))U(\ell^{2}(G)) of unitary transformations on 2(G)\ell^{2}(G). Define the (left) regular unitary representation of GG as πG:GU(2(G))\pi_{G}:G\rightarrow U(\ell^{2}(G)) such that

(πG(g)f)(x)=f(g1x)\displaystyle(\pi_{G}(g)f)(x)=f(g^{-1}x)

for fl2(G)f\in l^{2}(G) and x,gGx,g\in G. Unitary operator πG(s)\pi_{G}(s) 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 Cay(Gn,Sn)n=1{\operatorname{Cay}(G_{n},S_{n})}_{n=1}^{\infty}, where Sn={s1(n),sd(n)}GnS_{n}=\{s_{1}(n),\cdots s_{d}(n)\}\subset G_{n} for each nn with dd being a positive integer independent of nn. Then {Cay(Gn,Sn)}n=1\{\operatorname{Cay}(G_{n},S_{n})\}_{n=1}^{\infty} where

Nn:=|Gn|asn,N_{n}:=|G_{n}|\rightarrow\infty\quad\text{as}\quad n\rightarrow\infty,

is an expander sequence if and only if the sequence of dd-tuples of unitaries

{(πGn(s1(n)),,πGn(sd(n)))U(l2(Gn))d}n=1\displaystyle\Big{\{}\big{(}\pi_{G_{n}}(s_{1}(n)),\cdots,\pi_{G_{n}}(s_{d}(n))\big{)}\in U(l^{2}(G_{n}))^{d}\Big{\}}_{n=1}^{\infty}

satisfies that with ϵ>0\epsilon>0,

supn(1di=1dπGn(si(n)))|1Gn1ϵ,\displaystyle\sup_{n}\left\|\Bigg{(}\frac{1}{d}\sum_{i=1}^{d}\pi_{G_{n}}(s_{i}(n))\Bigg{)}\Bigg{|}_{1_{G_{n}}^{\perp}}\right\|\leq 1-\epsilon,

where \|\cdot\| is the Hilbert–Schmidt norm. This is equivalent to the condition that there exists a sequence of dd-tuples of unitaries

{(u1(n),,ud(n))U(Nn)d}n=1withuj(n)=πGn(sj(n)),\displaystyle\Big{\{}\big{(}u_{1}^{(n)},\cdots,u_{d}^{(n)}\big{)}\in U(N_{n})^{d}\Big{\}}_{n=1}^{\infty}\quad\text{with}\quad u_{j}^{(n)}=\pi_{G_{n}}(s_{j}(n)),

such that for any n1n\geq 1, some ϵ>0\epsilon>0, and any xx being a Nn×NnN_{n}\times N_{n} diagonal complex matrix,

1dj=1duj(n)(x1Nntr(x))uj(n)(1ϵ)x1Nntr(x),\displaystyle\left\|\frac{1}{d}\sum_{j=1}^{d}u_{j}^{(n)}\Big{(}x-\frac{1}{N_{n}}{\operatorname{tr}}(x)\Big{)}u_{j}^{(n)*}\right\|\leq(1-\epsilon)\Bigg{\|}x-\frac{1}{N_{n}}{\operatorname{tr}}(x)\Bigg{\|}, (3.1)

where U(Nn)U(N_{n}) stands for the set of Nn×NnN_{n}\times N_{n} unitary matrices.

The term “quantum expander” is just to designate unitaries (u1(n),,ud(n))\big{(}u_{1}^{(n)},\cdots,u_{d}^{(n)}\big{)} to satisfy (3.1) for a general xM(Nn)x\in M(N_{n}) beyond being diagonal, where M(Nn)M(N_{n}) stands for the space of Nn×NnN_{n}\times N_{n} complex matrices. In this light, quantum expanders can be seen as a non-commutative version of the classical ones. Identifying M(Nn)M(N_{n}) with the space B(2Nn)B(\ell_{2}^{N_{n}}) of bounded operators on the NnN_{n}-dimensional Hilbert space 2Nn\ell_{2}^{N_{n}}, different to the Markov operator TT defined in (2.3), in the quantum setting we consider quantum operators act on 2Nn2Nn¯\ell_{2}^{N_{n}}\otimes\overline{\ell_{2}^{N_{n}}} of the following form for each nn:

Tn=1dj=1duj(n)u¯j(n),uj(n)U(Nn),T_{n}=\frac{1}{d}\sum_{j=1}^{d}u_{j}^{(n)}\otimes\overline{u}_{j}^{(n)},\qquad u_{j}^{(n)}\in U(N_{n}),

where u¯j(n)\overline{u}_{j}^{(n)} is the complex conjugate of the matrix uj(n)u_{j}^{(n)}. Identifying as usual 2Nn2Nn¯\ell_{2}^{N_{n}}\otimes\overline{\ell_{2}^{N_{n}}} as S2NnS_{2}^{N_{n}}, the Hilbert space obtained by equipping M(Nn)M(N_{n}) with the corresponding scalar product and the Hilbert–Schmidt norm. We can write

1dj=1duj(n)u¯j(n):2Nn2Nn¯2Nn2Nn¯=Tn:S2NnS2Nn.\displaystyle\Bigg{\|}\frac{1}{d}\sum_{j=1}^{d}u_{j}^{(n)}\otimes\overline{u}_{j}^{(n)}:\ell_{2}^{N_{n}}\otimes\overline{\ell_{2}^{N_{n}}}\rightarrow\ell_{2}^{N_{n}}\otimes\overline{\ell_{2}^{N_{n}}}\Bigg{\|}=\left\|T_{n}:S_{2}^{N_{n}}\rightarrow S_{2}^{N_{n}}\right\|.

Then we may consider TnT_{n} as an operator acting on M(Nn)M(N_{n}) defined by

Tn(η)=1dj=1duj(n)ηuj(n),ηM(Nn),T_{n}(\eta)=\frac{1}{d}\sum_{j=1}^{d}u_{j}^{(n)}\eta u_{j}^{(n)*},\qquad\forall\eta\in M(N_{n}),

which satisfies that

1dj=1duj(n)u¯j(n)\displaystyle\Bigg{\|}\frac{1}{d}\sum_{j=1}^{d}u_{j}^{(n)}\otimes\overline{u}_{j}^{(n)}\Bigg{\|}
=\displaystyle= sup{1dj=1duj(n)ηuj(n)|ηM(Nn),η1}\displaystyle\sup\Bigg{\{}\Bigg{\|}\frac{1}{d}\sum_{j=1}^{d}u_{j}^{(n)}\eta u_{j}^{(n)*}\Bigg{\|}\;\Bigg{|}\;\eta\in M(N_{n}),\,\|\eta\|\leq 1\Bigg{\}}
=\displaystyle= sup{1d|j=1dtr(uj(n)ηuj(n)ζ)||η,ζM(Nn),η1,ζ1}.\displaystyle\sup\Bigg{\{}\frac{1}{d}\Bigg{|}\sum_{j=1}^{d}{\operatorname{tr}}(u_{j}^{(n)}\eta u_{j}^{(n)*}\zeta^{*})\Bigg{|}\;\Bigg{|}\;\eta,\zeta\in M(N_{n}),\,\|\eta\|\leq 1,\,\|\zeta\|\leq 1\Bigg{\}}.

Analogous to the classical setting, we have the operator norm Tn=1\|T_{n}\|=1 and Tn(INn)=INnT_{n}(I_{N_{n}})=I_{N_{n}}, where INnM(Nn)I_{N_{n}}\in M(N_{n}) is the identity matrix. Then we ask how much the operator norm of the quantum operator TnT_{n} would be decreased if it is restricted to the orthogonal complement INnI_{N_{n}}^{\perp} of the identity matrix. For that purpose, we give the quantum version of Definition 2.1.

Definition 3.1.

Define the reduced spectral radius ρ(n)\rho(n) as the restricted operator norm

ρ(n):=Tn|INn.\displaystyle\rho(n):=\big{\|}T_{n}|_{I_{N_{n}}^{\perp}}\big{\|}.

The definition of a quantum expander is defined in terms of ρ(n)\rho(n). It says that there exists ϵ>0\epsilon>0 such that ρ(n)1ϵ\rho(n)\leq 1-\epsilon for all nn, 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 dd and a sequence of positive integers {Nn}\{N_{n}\} with NnN_{n}\rightarrow\infty as nn\rightarrow\infty. We say a sequence of dd-tuples of unitaries

{(u1(n),,ud(n))U(Nn)d}n=1\Big{\{}\big{(}u_{1}^{(n)},\cdots,u_{d}^{(n)}\big{)}\in U(N_{n})^{d}\Big{\}}_{n=1}^{\infty}

is a quantum expander sequence if their reduced spectral radius ρ(n)\rho(n) is smaller than 11 uniformly for all nn.

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 P1,P2M(Nn)P_{1},P_{2}\in M(N_{n}), we have that for all nn,

|P1,TnP2HS1Nntr(P1)tr(P2)|(1ϵ)tr(P1)tr(P2),\displaystyle\left|\langle P_{1},T_{n}P_{2}\rangle_{{\operatorname{HS}}}-\frac{1}{N_{n}}{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2})\right|\leq(1-\epsilon)\sqrt{{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2})},

where the Hilbert–Schmidt inner product of two matrices AA and BB is defined as usual

A,BHS=tr(AB).\langle A,B\rangle_{{\operatorname{HS}}}={\operatorname{tr}}(A^{*}B).
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 C1,C2>0C_{1},C_{2}>0 and two non-zero orthogonal projections P1,P2M(Nn)P_{1},P_{2}\in M(N_{n}), such that for all nn,

|P1,TnP2HS1Nntr(P1)tr(P2)|1ϵC1(log(1ϵ)+C2)tr(P1)tr(P2).\displaystyle\left|\langle P_{1},T_{n}P_{2}\rangle_{{\operatorname{HS}}}-\frac{1}{N_{n}}{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2})\right|\geq\frac{1-\epsilon}{C_{1}(-\log(1-\epsilon)+C_{2})}\sqrt{{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2})}.

4 Proofs of the paper

Proof of Theorem 3.3.

Let E:M(Nn)M(Nn)E:M(N_{n})\rightarrow M(N_{n}) be the orthogonal projection onto the space INn=ker(1Tn)\langle I_{N_{n}}\rangle=\ker(1-T_{n}). Then

E(P2)=tr(P2)NnINn.E(P_{2})=\frac{{\operatorname{tr}}(P_{2})}{N_{n}}I_{N_{n}}.

Therefore,

tr(P1)tr(P2)=P1,NnE(P2)HS.\displaystyle{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2})=\langle P_{1},N_{n}E(P_{2})\rangle_{{\operatorname{HS}}}.

We have by the Cauchy-Schwarz inequality and properties of the unitaries that

|P1,TnP2HS1Nntr(P1)tr(P2)|2\displaystyle\left|\langle P_{1},T_{n}P_{2}\rangle_{{\operatorname{HS}}}-\frac{1}{N_{n}}{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2})\right|^{2} =|P1,(TE)P2HS|2\displaystyle=\Big{|}\langle P_{1},(T-E)P_{2}\rangle_{{\operatorname{HS}}}\Big{|}^{2}
tr(P1P1)tr(((TE)P2)((TE)P2))\displaystyle\leq{\operatorname{tr}}(P_{1}^{*}P_{1}){\operatorname{tr}}\Big{(}\big{(}(T-E)P_{2}\big{)}^{*}\big{(}(T-E)P_{2}\big{)}\Big{)}
tr(P1)T|INn2tr(P2)\displaystyle\leq{\operatorname{tr}}(P_{1})\big{\|}T|_{I_{N_{n}}^{\perp}}\big{\|}^{2}{\operatorname{tr}}(P_{2})
(1ϵ)2tr(P1)tr(P2).\displaystyle\leq(1-\epsilon)^{2}{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2}).

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 p[1,)p\in[1,\infty), define the Schatten pp-norm of AM(Nn)A\in M(N_{n}) as

Ap=[tr((AA)p2)]1p;\|A\|_{p}=\Big{[}{\operatorname{tr}}\big{(}(A^{*}A)^{\frac{p}{2}}\big{)}\Big{]}^{\frac{1}{p}};

for p=p=\infty, the convention is to define \|\cdot\|_{\infty} as the operator norm

A=supxn,x=1Ax.\|A\|_{\infty}=\sup_{x\in\mathbb{C}^{n},\|x\|=1}\|Ax\|.

The Schatten norms are unitarily invariant in the sense that for unitaries U,VU,V and p[1,]p\in[1,\infty],

UAVp=Ap.\displaystyle\|UAV\|_{p}=\|A\|_{p}. (4.1)

Considering q=1q=1 as the dual index to p=p=\infty, the noncommutative Hölder’s inequality says that for all p,q,r[1,]p,q,r\in[1,\infty] such that 1p+1q=1r\frac{1}{p}+\frac{1}{q}=\frac{1}{r}, we have that for any matrices A,BM(Nn)A,B\in M(N_{n}),

ABrApBq.\|AB\|_{r}\leq\|A\|_{p}\|B\|_{q}.

Its special case is the well-known inequality

A2=A22A1A,\displaystyle\|A\|^{2}=\|A\|_{2}^{2}\leq\|A\|_{1}\|A\|_{\infty}, (4.2)

where 1\|\cdot\|_{1} is called the trace norm or the nuclear norm, 2\|\cdot\|_{2} is called the Frobenius norm or the Hilbert–Schmidt norm, and \|\cdot\|_{\infty} 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 AM(Nn)A\in M(N_{n}) by

h(A):=A1AA.\displaystyle h(A):=\frac{\sqrt{\|A\|_{1}\|A\|_{\infty}}}{\|A\|}.

We define the Schatten norm induced operator norm as Tnpp\|T_{n}\|_{p\rightarrow p} for operator Tn:M(Nn)M(Nn)T_{n}:M(N_{n})\rightarrow M(N_{n}). Then the three special cases p=1,2p=1,2 and \infty give operator norms T11\|T\|_{1\rightarrow 1}, T22\|T\|_{2\rightarrow 2} and T\|T\|_{\infty\rightarrow\infty}, respectively. Then we define a linear map for unitaries U1,U2,V1,V2M(Nn)U_{1},U_{2},V_{1},V_{2}\in M(N_{n}) as follows:

Tn(U1,U2,V1,V2):n\displaystyle T_{n}(U_{1},U_{2},V_{1},V_{2}):\mathbb{C}^{n} n\displaystyle\rightarrow\mathbb{C}^{n}
(λ1,λn)\displaystyle(\lambda_{1},\cdots\lambda_{n}) diag(U2Tn(U1DλV1)V2),\displaystyle\mapsto{\operatorname{diag}}(U_{2}T_{n}(U_{1}D_{\lambda}V_{1})V_{2}), (4.3)

where DλD_{\lambda} is the diagonal matrix diag(λ1,λn){\operatorname{diag}}(\lambda_{1},\cdots\lambda_{n}). Here, we follow the convention that when the diag{\operatorname{diag}} 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

Tnpp=supU1,U2,V1,V2Tn(U1,U2,V1,V2)pp.\displaystyle\|T_{n}\|_{p\rightarrow p}=\sup_{U_{1},U_{2},V_{1},V_{2}}\|T_{n}(U_{1},U_{2},V_{1},V_{2})\|_{p\rightarrow p}. (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 Tn:M(Nn)M(Nn)T_{n}:M(N_{n})\rightarrow M(N_{n}), we have

Tn222Tn11Tn.\displaystyle\|T_{n}\|_{2\rightarrow 2}^{2}\leq\|T_{n}\|_{1\rightarrow 1}\|T_{n}\|_{\infty\rightarrow\infty}.

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 Tn:M(Nn)M(Nn)T_{n}:M(N_{n})\rightarrow M(N_{n}), define its height by

h(Tn):=Tn11TnTn22.\displaystyle h(T_{n}):=\frac{\sqrt{\|T_{n}\|_{1\rightarrow 1}\|T_{n}\|_{\infty\rightarrow\infty}}}{\|T_{n}\|_{2\rightarrow 2}}.

Recall that in the classical setting, for an operator T:nnT:\mathbb{C}^{n}\rightarrow\mathbb{C}^{n}, we have

T11=T.\displaystyle\|T\|_{\ell^{1}\rightarrow\ell^{1}}=\|T^{*}\|_{\ell^{\infty}\rightarrow\ell^{\infty}}.

In the quantum setting, denote TnT_{n}^{*} as the adjoint of Tn:M(Nn)M(Nn)T_{n}:M(N_{n})\rightarrow M(N_{n}) with respect to the Hilbert–Schmidt inner product. For λ,μn\lambda,\mu\in\mathbb{C}^{n}, we have

μ,diag(U2Tn(U1DλV1)V2)=\displaystyle\Big{\langle}\mu,\,{\operatorname{diag}}\big{(}U_{2}T_{n}(U_{1}D_{\lambda}V_{1})V_{2}\big{)}\Big{\rangle}= Dμ,U2Tn(U1DλV1)V2HS\displaystyle\Big{\langle}D_{\mu},\,U_{2}T_{n}(U_{1}D_{\lambda}V_{1})V_{2}\Big{\rangle}_{{\operatorname{HS}}}
=\displaystyle= U2DμV2,Tn(U1DλV1)HS\displaystyle\Big{\langle}U_{2}^{*}D_{\mu}V_{2}^{*},\,T_{n}(U_{1}D_{\lambda}V_{1})\Big{\rangle}_{{\operatorname{HS}}}
=\displaystyle= U1Tn(U2DμV2)V1,DλHS\displaystyle\Big{\langle}U_{1}^{*}T_{n}^{*}(U_{2}^{*}D_{\mu}V_{2}^{*})V_{1}^{*},\,D_{\lambda}\Big{\rangle}_{{\operatorname{HS}}}
=\displaystyle= diag(U1Tn(U2DμV2)V1),λ,\displaystyle\Big{\langle}{\operatorname{diag}}\big{(}U_{1}^{*}T_{n}^{*}(U_{2}^{*}D_{\mu}V_{2}^{*})V_{1}^{*}\big{)},\,\lambda\Big{\rangle},

which yields that

Tn(U1,U2,V1,V2)=T(U2,U1,V2,V1).T_{n}(U_{1},U_{2},V_{1},V_{2})^{*}=T^{*}(U_{2}^{*},U_{1}^{*},V_{2}^{*},V_{1}^{*}).

Hence, we have

Tn22\displaystyle\|T_{n}\|_{2\rightarrow 2} =Tn22,\displaystyle=\|T_{n}^{*}\|_{2\rightarrow 2}, (4.5)
Tn11=Tn\displaystyle\|T_{n}\|_{1\rightarrow 1}=\|T_{n}^{*}\|_{\infty\rightarrow\infty}\quad andTn=Tn11.\displaystyle\text{and}\quad\|T_{n}\|_{\infty\rightarrow\infty}=\|T_{n}^{*}\|_{1\rightarrow 1}.

By Definition 4.3,

h(Tn)=h(Tn)\displaystyle h(T_{n})=h(T_{n}^{*}) (4.6)

In the classical setting, Lev, (2015) defined the logarithmic diameter for a non-zero vector z=(z1,,zn)nz=(z_{1},\ldots,z_{n})\in\mathbb{C}^{n} by

(z):=max{|zi|:i{1,,n}}max{|zi|:i{1,,n},zi0}.\displaystyle\ell(z):=\frac{\max\{|z_{i}|:i\in\{1,\ldots,n\}\}}{\max\{|z_{i}|:i\in\{1,\ldots,n\},z_{i}\neq 0\}}. (4.7)

In the quantum setting, considering complex matrices, we define the logarithmic diameter as follows.

Definition 4.4.

For AM(Nn)A\in M(N_{n}), let

0<λ1λ2λk\displaystyle 0<\lambda_{1}\leq\lambda_{2}\leq\cdots\leq\lambda_{k}

be all the non-zero (strictly positive) eigenvalues of AAA^{*}A. We define the logarithmic diameter of AA by

(A):=λk/λ1.\displaystyle\ell(A):=\sqrt{\lambda_{k}/\lambda_{1}}.

The following lemma will serve as the cornerstone for the subsequent proofs.

Lemma 4.5.

If T:M(Nn)M(Nn)T:M(N_{n})\rightarrow M(N_{n}) is a linear map with h(T)Kh(T)\leq K for K1K\geq 1 a real number, then there exists a non-zero matrix AM(Nn)A\in M(N_{n}) such that

Tn(A)>14T22Aand(A)<32K2+1.\|T_{n}(A)\|>\frac{1}{4}\|T\|_{2\rightarrow 2}\|A\|\quad\text{and}\quad\ell(A)<32K^{2}+1.
Proof.

By the definition of Tn(U1,U2,V1,V2)T_{n}(U_{1},U_{2},V_{1},V_{2}) given in (4) where U1,U2,V1U_{1},U_{2},V_{1} and V2V_{2} are unitaries in M(Nn)M(N_{n}), together with (4.4), we have

Tn22=supU1,U2,V1,V2Tn(U1,U2,V1,V2)222Tn(U1,U2,V1,V2),\displaystyle\|T_{n}\|_{2\rightarrow 2}=\sup_{U_{1},U_{2},V_{1},V_{2}}\|T_{n}(U_{1},U_{2},V_{1},V_{2})\|_{2\rightarrow 2}\leq 2\|T_{n}(U_{1},U_{2},V_{1},V_{2})\|, (4.8)

which yields that

1Tn2212Tn(U1,U2,V1,V2).\displaystyle\frac{1}{\|T_{n}\|_{2\rightarrow 2}}\geq\frac{1}{2\|T_{n}(U_{1},U_{2},V_{1},V_{2})\|}.

Then by the definition of the height of a matrix given in Definition 4.1, together with (4.4), we have

h(Tn(U1,U2,V1,V2))2\displaystyle h(T_{n}(U_{1},U_{2},V_{1},V_{2}))^{2} =Tn(U1,U2,V1,V2)1Tn(U1,U2,V1,V2)Tn(U1,U2,V1,V2)22\displaystyle=\frac{\|T_{n}(U_{1},U_{2},V_{1},V_{2})\|_{1}\|T_{n}(U_{1},U_{2},V_{1},V_{2})\|_{\infty}}{\|T_{n}(U_{1},U_{2},V_{1},V_{2})\|_{2}^{2}}
4Tn(U1,U2,V1,V2)1Tn(U1,U2,V1,V2)Tn22\displaystyle\leq 4\frac{\|T_{n}(U_{1},U_{2},V_{1},V_{2})\|_{1}\|T_{n}(U_{1},U_{2},V_{1},V_{2})\|_{\infty}}{\|T_{n}\|_{2\rightarrow 2}}
4Tn11TnTn22.\displaystyle\leq 4\frac{\|T_{n}\|_{1\rightarrow 1}\|T_{n}\|_{\infty\rightarrow\infty}}{\|T_{n}\|_{2\rightarrow 2}}.

Therefore, by the definition of height of a linear operator given in Definition 4.2, we have

h(Tn(U1,U2,V1,V2))2h(T)2K.h(T_{n}(U_{1},U_{2},V_{1},V_{2}))\leq 2h(T)\leq 2K.

By Lemma 1 of Lev, (2015), there exists znz\in\mathbb{C}^{n} such that

Tn(U1,U2,V1,V2)z>12Tn(U1,U2,V1,V2)z,\displaystyle\|T_{n}(U_{1},U_{2},V_{1},V_{2})z\|>\frac{1}{2}\|T_{n}(U_{1},U_{2},V_{1},V_{2})\|\cdot\|z\|, (4.9)

and then the logarithmic diameter for a non-zero vector defined in (4.7)

(z)<8(2K)2+1.\ell(z)<8(2K)^{2}+1.

Taking the matrix

A=U1DzV1,\displaystyle A=U_{1}D_{z}V_{1}, (4.10)

with DzD_{z} being the diagonal matrix diag(z){\operatorname{diag}}(z), the logarithmic diameter for a non-zero matrix defined in Definition 4.4

(A)=(z)<32K2+1.\ell(A)=\ell(z)<32K^{2}+1.

Note that

Tn(A),Tn(A)HS\displaystyle\langle T_{n}(A),T_{n}(A)\rangle_{{\operatorname{HS}}} =U2Tn(A)V2,U2Tn(A)V2HS\displaystyle=\Big{\langle}U_{2}T_{n}(A)V_{2},\,U_{2}T_{n}(A)V_{2}\Big{\rangle}_{{\operatorname{HS}}}
diag(U2Tn(A)V2),diag(U2Tn(A)V2).\displaystyle\geq\Big{\langle}{\operatorname{diag}}(U_{2}T_{n}(A)V_{2}),\,{\operatorname{diag}}(U_{2}T_{n}(A)V_{2})\Big{\rangle}.

Recall that by the definition of Tn(U1,U2,V1,V2)T_{n}(U_{1},U_{2},V_{1},V_{2}) given in (4),

diag(U2Tn(A)V2)=Tn(U1,U2,V1,V2)z,{\operatorname{diag}}(U_{2}T_{n}(A)V_{2})=T_{n}(U_{1},U_{2},V_{1},V_{2})z,

using the specific form of AA. By equation (4.9),

Tn(A),Tn(A)HS=Tn(U1,U2,V1,V2)z2\displaystyle\langle T_{n}(A),T_{n}(A)\rangle_{{\operatorname{HS}}}=\|T_{n}(U_{1},U_{2},V_{1},V_{2})z\|^{2} >14Tn(U1,U2,V1,V2)2z2.\displaystyle>\frac{1}{4}\|T_{n}(U_{1},U_{2},V_{1},V_{2})\|^{2}\|z\|^{2}.

At last, using equations (4.1) and (4.8), we have by the specific form of AA that

Tn(A),Tn(A)HS>116T222A2,\displaystyle\langle T_{n}(A),T_{n}(A)\rangle_{{\operatorname{HS}}}>\frac{1}{16}\|T\|_{2\rightarrow 2}^{2}\|A\|^{2},

which completes the proof. ∎

Lemma 4.6.

If XM(Nn)X\in M(N_{n}) is a matrix with h(X)Kh(X)\leq K for K1K\geq 1 a real number, then there exists a non-zero orthogonal projection PM(Nn)P\in M(N_{n}) such that

|X,PHS|124log(2K)+2XP.\displaystyle|\langle X,P\rangle_{{\operatorname{HS}}}|\geq\frac{1}{2\sqrt{4\log(2K)+2}}\|X\|\|P\|.
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 AM(Nn)A\in M(N_{n}) is a self-adjoint matrix with height h(A)Kh(A)\leq K for K1K\geq 1 a real number, where h(A)h(A) is given in Definition 4.1. Since AA is self-adjoint, there exists a unitary UM(Nn)U\in M(N_{n}) such that

UAU=diag(λ1,,λn)M(Nn).\displaystyle UAU^{*}={\operatorname{diag}}(\lambda_{1},\cdots,\lambda_{n})\in M(N_{n}).

By the definition of height and the unitarily invariant property (4.1), we have

h(A)2=A1AA2=UAU1UAUUAU2=h(UAU)2.\displaystyle h(A)^{2}=\frac{\|A\|_{1}\|A\|_{\infty}}{\|A\|^{2}}=\frac{\|UAU^{*}\|_{1}\|UAU^{*}\|_{\infty}}{\|UAU^{*}\|^{2}}=h(UAU^{*})^{2}.

Next, we have

h((λ1,,λn)T)2=\displaystyle h((\lambda_{1},\cdots,\lambda_{n})^{T})^{2}= (λ1,,λn)T1(λ1,,λn)T(λ1,,λn)T2\displaystyle\frac{\|(\lambda_{1},\cdots,\lambda_{n})^{T}\|_{1}\|(\lambda_{1},\cdots,\lambda_{n})^{T}\|_{\infty}}{\|(\lambda_{1},\cdots,\lambda_{n})^{T}\|^{2}}
=\displaystyle= diag(λ1,,λn)1diag(λ1,,λn)diag(λ1,,λn)2\displaystyle\frac{\|{\operatorname{diag}}(\lambda_{1},\cdots,\lambda_{n})\|_{1}\|{\operatorname{diag}}(\lambda_{1},\cdots,\lambda_{n})\|_{\infty}}{\|{\operatorname{diag}}(\lambda_{1},\cdots,\lambda_{n})\|^{2}}
=\displaystyle= h(UAU)2,\displaystyle h(UAU^{*})^{2},

where the superscript ``T"``T" represents the transpose, and the norms applied to vectors or matrices by their corresponding definitions. Hence,

h((λ1,,λn)T)=h(A)K.h((\lambda_{1},\cdots,\lambda_{n})^{T})=h(A)\leq K.

Then by Lemma 4 of Lev, (2015), there exists an nn-dimensional binary vector ξ{0,1}n\xi\in\{0,1\}^{n} such that

|cos((λ1,,λn)T,ξ)|12log(2K2)+1,\displaystyle|\cos{((\lambda_{1},\cdots,\lambda_{n})^{T},\xi)}|\geq\frac{1}{2\sqrt{\log(2K^{2})+1}}, (4.11)

where, for non-zero vectors u,vnu,v\in\mathbb{C}^{n},

cos(u,v)=u,v/uv.\cos(u,v)=\langle u,v\rangle/\|u\|\|v\|.

Note that

A=(λ1,,λn)T,\displaystyle\|A\|=\|(\lambda_{1},\cdots,\lambda_{n})^{T}\|,
Udiag(ξ1,,ξn)U=(ξ1,,ξn)T,\displaystyle\|U^{*}{\operatorname{diag}}(\xi_{1},\cdots,\xi_{n})U\|=\|(\xi_{1},\cdots,\xi_{n})^{T}\|,

and given that the trace is invariant under circular shifts

A,Udiag(ξ1,,ξn)UHS=\displaystyle\Big{\langle}A,\,U^{*}{\operatorname{diag}}(\xi_{1},\cdots,\xi_{n})U\Big{\rangle}_{{\operatorname{HS}}}= tr(AUdiag(ξ1,,ξn)U)\displaystyle{\operatorname{tr}}\Big{(}A^{*}U^{*}{\operatorname{diag}}(\xi_{1},\cdots,\xi_{n})U\Big{)}
=\displaystyle= tr(UAUdiag(ξ1,,ξn))\displaystyle{\operatorname{tr}}\Big{(}UA^{*}U^{*}{\operatorname{diag}}(\xi_{1},\cdots,\xi_{n})\Big{)}
=\displaystyle= tr(diag(λ1,,λn)diag(ξ1,,ξn))\displaystyle{\operatorname{tr}}\Big{(}{\operatorname{diag}}(\lambda_{1},\cdots,\lambda_{n})\,{\operatorname{diag}}(\xi_{1},\cdots,\xi_{n})\Big{)}
=\displaystyle= (λ1,,λn)T,(ξ1,,ξn)T.\displaystyle\Big{\langle}(\lambda_{1},\cdots,\lambda_{n})^{T},\,(\xi_{1},\cdots,\xi_{n})^{T}\Big{\rangle}.

Set

P=Udiag(ξ1,,ξn)U\displaystyle P=U^{*}{\operatorname{diag}}(\xi_{1},\cdots,\xi_{n})U (4.12)

which is an orthogonal projection in M(Nn)M(N_{n}). Then by equation (4.11), we have that

|A,PHS|12log(2K2)+1AP.\displaystyle|\langle A,P\rangle_{{\operatorname{HS}}}|\geq\frac{1}{2\sqrt{\log(2K^{2})+1}}\|A\|\|P\|. (4.13)

Step 2. According to the Toeplitz decomposition, each XM(Nn)X\in M(N_{n}) can be written uniquely as

X=A+iB,X=A+iB,

in which A,BM(Nn)A,B\in M(N_{n}) are self-adjoint matrices; see Horn and Johnson, (2012). Since

X1=X1andX=X,\|X\|_{1}=\|X^{*}\|_{1}\quad\text{and}\quad\|X\|_{\infty}=\|X^{*}\|_{\infty},

we have

A1=X+X21X1+X12=X1andAX.\displaystyle\|A\|_{1}=\left\|\frac{X+X^{*}}{2}\right\|_{1}\leq\frac{\|X\|_{1}+\|X^{*}\|_{1}}{2}=\|X\|_{1}\quad\text{and}\quad\|A\|_{\infty}\leq\|X\|_{\infty}. (4.14)

Additionally, for the Hilbert-Schmidt norm, we have

X2=tr((A+iB)(AiB))=tr(A2)+tr(B2)=A2+B2.\displaystyle\|X\|^{2}={\operatorname{tr}}((A+iB)(A-iB))={\operatorname{tr}}(A^{2})+{\operatorname{tr}}(B^{2})=\|A\|^{2}+\|B\|^{2}.

Replacing XX by iXiX if necessarily, we may assume A2B2\|A\|^{2}\geq\|B\|^{2}, which gives

A12X.\|A\|\geq\frac{1}{\sqrt{2}}\|X\|.

Then, by the definition of height given in Definition 4.1, equation (4.14), and the condition that h(X)Kh(X)\leq K,

h(A)=A1AAX1XX/22K.\displaystyle h(A)=\frac{\sqrt{\|A\|_{1}\|A\|_{\infty}}}{\|A\|}\leq\frac{\sqrt{\|X\|_{1}\|X\|_{\infty}}}{\|X\|/\sqrt{2}}\leq\sqrt{2}K.

By the result of Step 1, specifically (4.13), there exists an orthogonal projection PM(Nn)P\in M(N_{n}) such that

|A,PHS|12log(2(2K)2)+1AP\displaystyle|\langle A,P\rangle_{{\operatorname{HS}}}|\geq\frac{1}{2\sqrt{\log(2(\sqrt{2}K)^{2})+1}}\|A\|\|P\| 12log(4K2)+1X2P\displaystyle\geq\frac{1}{2\sqrt{\log(4K^{2})+1}}\frac{\|X\|}{\sqrt{2}}\|P\|
=124log(2K)+2XP.\displaystyle=\frac{1}{2\sqrt{4\log(2K)+2}}\|X\|\|P\|.

Combining this with

|X,PHS|=|A,PHS+iB,PHS||A,PHS|\displaystyle|\langle X,P\rangle_{{\operatorname{HS}}}|=|\langle A,P\rangle_{{\operatorname{HS}}}+i\langle B,P\rangle_{{\operatorname{HS}}}|\geq|\langle A,P\rangle_{{\operatorname{HS}}}|

yields the desired estimate. ∎

Theorem 4.7.

For any linear map Tn:M(Nn)M(Nn)T_{n}:M(N_{n})\rightarrow M(N_{n}), there exists a non-zero orthogonal projection PM(Nn)P\in M(N_{n}) such that

Tn2284(log(48h2(Tn))+2)<TnPPTn22.\displaystyle\frac{\|T_{n}\|_{2\rightarrow 2}}{8\sqrt{4(\log(48h^{2}(T_{n}))+2)}}<\frac{\|T_{n}P\|}{\|P\|}\leq\|T_{n}\|_{2\rightarrow 2}.
Proof.

The second inequality clearly holds by the definition of Tn22\|T_{n}\|_{2\rightarrow 2}, since PP being a non-zero projection is a special case of a complex matrix. Now, we prove the first inequality.

Suppose h(Tn)=Kh(T_{n})=K for K1K\geq 1 a real number. By Lemma 4.5, there exists a non-zero matrix AM(Nn)A\in M(N_{n}) such that

Tn(A)>14Tn22Aand(A)<32K2+1.\displaystyle\|T_{n}(A)\|>\frac{1}{4}\|T_{n}\|_{2\rightarrow 2}\|A\|\quad\text{and}\quad\ell(A)<32K^{2}+1. (4.15)

Furthermore, by equations (4.5) and (4.6), we have

Tn(A)14Tn22A=14Tn22A.\displaystyle\|T_{n}^{*}(A)\|\geq\frac{1}{4}\|T_{n}^{*}\|_{2\rightarrow 2}\|A\|=\frac{1}{4}\|T_{n}\|_{2\rightarrow 2}\|A\|. (4.16)

Consider the spectral decomposition of (AA)12(A^{*}A)^{\frac{1}{2}} as

(AA)12=iIσiPi,(A^{*}A)^{\frac{1}{2}}=\sum\limits_{i\in I}\sigma_{i}P_{i},

where {σi}iI\{\sigma_{i}\}_{i\in I} is the set of eigenvalues of (AA)12(A^{*}A)^{\frac{1}{2}} and PiP_{i} is the corresponding spectral projection. Then, for p[1,)p\in[1,\infty), the Schatten pp-norm of AM(Nn)A\in M(N_{n}) can be written as

Ap=[tr((AA)p2)]1p=(iIσip)1p;\|A\|_{p}=\Big{[}{\operatorname{tr}}\big{(}(A^{*}A)^{\frac{p}{2}}\big{)}\Big{]}^{\frac{1}{p}}=\left(\sum_{i\in I}\sigma_{i}^{p}\right)^{\frac{1}{p}};

while for p=p=\infty, it can be written as

A=supxn,x=1Ax=maxiIσi.\|A\|_{\infty}=\sup_{x\in\mathbb{C}^{n},\|x\|=1}\|Ax\|=\max_{i\in I}\sigma_{i}.

Now, choose jIj\in I such that

σj=miniI{σiσi0}.\displaystyle\sigma_{j}=\min_{i\in I}\{\sigma_{i}\mid\sigma_{i}\neq 0\}.

By the definition of (A)\ell(A) given in Definition 4.4 which is defined in terms of eigenvalues of AAA^{*}A (instead of (AA)12(A^{*}A)^{\frac{1}{2}}), together with equation (4.15), we obtain

h2(A)=A1AA2(σi)l(A)σjσi2(A)<36K2.\displaystyle h^{2}(A)=\frac{\|A\|_{1}\|A\|_{\infty}}{\|A\|^{2}}\leq\frac{(\sum\limits\sigma_{i})l(A)\sigma_{j}}{\sum\limits\sigma_{i}^{2}}\leq\ell(A)<36K^{2}.

Furthermore, by equation (4.15),

h2(Tn(A))=Tn(A)1Tn(A)Tn(A)2\displaystyle h^{2}(T_{n}(A))=\frac{\|T_{n}(A)\|_{1}\|T_{n}(A)\|_{\infty}}{\|T_{n}(A)\|^{2}} <Tn11A1TnA(14Tn22A)2\displaystyle<\frac{\|T_{n}\|_{1\rightarrow 1}\|A\|_{1}\|T_{n}\|_{\infty\rightarrow\infty}\|A\|_{\infty}}{(\frac{1}{4}\|T_{n}\|_{2\rightarrow 2}\|A\|)^{2}}
=16h2(Tn)h2(A).\displaystyle=16h^{2}(T_{n})h^{2}(A).

Therefore,

h(Tn(A))<4h(Tn)h(A)<24K2.h(T_{n}(A))<4h(T_{n})h(A)<24K^{2}.

By Lemma 4.6, there exists an orthogonal projection PM(Nn)P\in M(N_{n}) such that

|TnA,PHS|>124log(48K2)+2TnAP.\displaystyle|\langle T_{n}^{*}A,P\rangle_{{\operatorname{HS}}}|>\frac{1}{2\sqrt{4\log(48K^{2})+2}}\|T_{n}^{*}A\|\|P\|.

Therefore, by equation (4.16),

|A,TnPHS|=|TnA,PHS|>184(log(48K2)+2)Tn22AP,\displaystyle|\langle A,T_{n}P\rangle_{{\operatorname{HS}}}|=|\langle T_{n}^{*}A,P\rangle_{{\operatorname{HS}}}|>\frac{1}{8\sqrt{4(\log(48K^{2})+2)}}\|T_{n}\|_{2\rightarrow 2}\|A\|\|P\|,

which yields that

TnP>\displaystyle\|T_{n}P\|> 184(log(48K2)+2)Tn22P,\displaystyle\frac{1}{8\sqrt{4(\log(48K^{2})+2)}}\|T_{n}\|_{2\rightarrow 2}\|P\|,

as desired.

Theorem 4.8.

There exist two universal constants C1,C2>0C_{1},C_{2}>0 and two non-zero orthogonal projections P,QM(Nn)P,Q\in M(N_{n}), such that for any linear map Tn:M(Nn)M(Nn)T_{n}:M(N_{n})\rightarrow M(N_{n}), we have

Tn22C1(log(h(TnQ))+C2)<P,TnQHStr(P)tr(Q)Tn22\displaystyle\frac{\|T_{n}\|_{2\rightarrow 2}}{C_{1}(\log(h(T_{n}Q))+C_{2})}<\frac{\langle P,\,T_{n}Q\rangle_{{\operatorname{HS}}}}{\sqrt{{\operatorname{tr}}(P){\operatorname{tr}}(Q)}}\leq\|T_{n}\|_{2\rightarrow 2}
Proof.

The second inequality clearly holds by the definition of Tn22\|T_{n}\|_{2\rightarrow 2}, because the non-zero projections PP and QQ are specific instances of complex matrices. Now, we prove the first inequality.

Suppose h(Tn)=Kh(T_{n})=K for K1K\geq 1 a real number. Define

f(K):=84(log(48K2)+2).\displaystyle f(K):={8\sqrt{4(\log(48K^{2})+2)}}.

Theorem 4.7 states that there exists an orthogonal projection PP such that

Tn(P)>Tn22P/f(K),\displaystyle\|T_{n}(P)\|>\|T_{n}\|_{2\rightarrow 2}\|P\|/f(K), (4.17)

which yields that

h(Tn(P))=Tn(P)1Tn(P)Tn(P)\displaystyle h(T_{n}(P))=\frac{\sqrt{\|T_{n}(P)\|_{1}\|T_{n}(P)\|_{\infty}}}{\|T_{n}(P)\|} <Tn11P1TnPTn22P/f(K)\displaystyle<\frac{\sqrt{\|T_{n}\|_{1\rightarrow 1}{\|P\|_{1}{\|T_{n}\|}_{\infty\rightarrow\infty}{\|P\|}_{\infty}}}}{\|T_{n}\|_{2\rightarrow 2}\|P\|/f(K)}
=h(Tn)P1PPf(K).\displaystyle=h(T_{n})\frac{\sqrt{{\|P\|_{1}{\|P\|}_{\infty}}}}{\|P\|}f(K).

Note that, for an orthogonal projection PP, we have

P2=P1P.\|P\|^{2}=\|P\|_{1}\|P\|_{\infty}.

Then we have

h(Tn(P))<Kf(K).\displaystyle h(T_{n}(P))<Kf(K).

By Lemma 4.6, there exists an orthogonal projection QM(Nn)Q\in M(N_{n}) such that

|Tn(P),QHS|\displaystyle|\langle T_{n}(P),Q\rangle_{{\operatorname{HS}}}| 124log(2Kf(K))+2Tn(P)Q\displaystyle\geq\frac{1}{2\sqrt{4\log(2Kf(K))+2}}\|T_{n}(P)\|\|Q\|
>12f(K)4log(2Kf(K))+2Tn22PQ,\displaystyle>\frac{1}{2f(K)\sqrt{4\log(2Kf(K))+2}}\|T_{n}\|_{2\rightarrow 2}\|P\|\|Q\|,

where we used (4.17) in the last inequality. Note that

2f(K)4log(2Kf(K))+2C1(logK+C2)\displaystyle 2f(K)\sqrt{4\log(2Kf(K))+2}\leq C_{1}(\log K+C_{2})

for some C1>0C_{1}>0 and C2>0C_{2}>0. The proof is complete. ∎

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 E:M(Nn)M(Nn)E:M(N_{n})\rightarrow M(N_{n}) be the orthogonal projection onto the space INn=ker(1Tn)\langle I_{N_{n}}\rangle=\ker(1-T_{n}). Then Tn|INnT_{n}|_{I_{N_{n}}^{\perp}} is the restriction of TnT_{n} onto (1E)(1-E).

According to the definition of a quantum expander sequence given in Definition 3.2, there exists ϵ>0\epsilon>0 such that

Tn|INn1ϵ.\big{\|}T_{n}|_{I_{N_{n}}^{\perp}}\big{\|}\leq 1-\epsilon.

Here, we consider ϵ\epsilon as the largest of those constants that suffice the above inquality. Then by the definition of 22\|\cdot\|_{2\rightarrow 2}, we have

Tn|INn221ϵ.\displaystyle\big{\|}T_{n}|_{I_{N_{n}}^{\perp}}\big{\|}_{2\rightarrow 2}\geq 1-\epsilon.

Therefore,

h(Tn|INn)Tn|INn11Tn|INn1ϵ21ϵ.h\big{(}T_{n}|_{I_{N_{n}}^{\perp}}\big{)}\leq\frac{\sqrt{\big{\|}T_{n}|_{I_{N_{n}}^{\perp}}\big{\|}_{1\rightarrow 1}\big{\|}T_{n}|_{I_{N_{n}}^{\perp}}\big{\|}_{\infty\rightarrow\infty}}}{1-\epsilon}\leq\frac{2}{1-\epsilon}.

By Theorem 4.8, there exist two universal constants C1,C2>0C_{1}^{\prime},C_{2}^{\prime}>0 and two non-zero orthogonal projections P1,P2M(Nn)P_{1},P_{2}\in M(N_{n}), such that

P1,(TnE)P2HS1ϵC1(log(21ϵ)+C2)tr(P1)tr(P2).\displaystyle{\langle P_{1},\,(T_{n}-E)P_{2}\rangle_{{\operatorname{HS}}}}\geq\frac{1-\epsilon}{C_{1}^{\prime}(\log(\frac{2}{1-\epsilon})+C_{2}^{\prime})}\sqrt{{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2})}.

Noting that

P1,EP2HS=tr(P1tr(P2)NnINn)=1Nntr(P1)tr(P2),\displaystyle\langle P_{1},\,EP_{2}\rangle_{{\operatorname{HS}}}={\operatorname{tr}}\left(P_{1}^{*}\frac{{\operatorname{tr}}(P_{2})}{N_{n}}I_{N_{n}}\right)=\frac{1}{N_{n}}{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2}),

we have P1P_{1} and P2P_{2} satisfies

|P1,Tn(P2)HS1Nntr(P1)tr(P2)|1ϵC1(log(1ϵ)+C2)tr(P1)tr(P2),\displaystyle\left|{\langle P_{1},\,T_{n}(P_{2})\rangle_{{\operatorname{HS}}}}-\frac{1}{N_{n}}{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2})\right|\geq\frac{1-\epsilon}{C_{1}(-\log(1-\epsilon)+C_{2})}\sqrt{{\operatorname{tr}}(P_{1}){\operatorname{tr}}(P_{2})},

where C1,C2>0C_{1},C_{2}>0 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.