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

Almost Ramanujan Expanders from Arbitrary Expanders via Operator Amplification

Fernando Granha Jeronimo University of Illinois, Urbana-Champaign. This material is based upon work supported by the NSF grant CCF-1900460.    Tushant Mittal Stanford University.    Sourya Roy University of Iowa.    Avi Wigderson Institute for Advanced Study, Princeton. This work was partially supported by NSF grant CCF-1900460.
Any opinions, findings and conclusions or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of the NSF.

We give an efficient algorithm that transforms any bounded degree expander graph into another that achieves almost optimal (namely, near-quadratic, d1/λ2+o(1)d\leq 1/\lambda^{2+o(1)}) trade-off between (any desired) spectral expansion λ\lambda and degree dd. Furthermore, the algorithm is local: every vertex in the new graph can compute its new neighbors as a subset of its original neighborhood of radius O(log(1/λ))O(\log(1/\lambda)). The optimal quadratic trade-off is known as the Ramanujan bound, so our construction gives almost Ramanujan expanders from arbitrary expanders.

The locality of the transformation preserves structural properties of the original graph, and thus has many consequences. Applied to Cayley graphs, our transformation shows that any expanding finite group has almost Ramanujan expanding generators. Similarly, one can obtain almost optimal explicit constructions of quantum expanders, dimension expanders, monotone expanders, etc., from existing (suboptimal) constructions of such objects. Another consequence is a "derandomized" random walk on the original (suboptimal) expander with almost optimal convergence rate. Our transformation also applies when the degree is not bounded or the expansion is not constant.

We obtain our results by a generalization of Ta-Shma’s technique in his breakthrough paper [STOC 2017], used to obtain explicit almost optimal binary codes. Specifically, our spectral amplification extends Ta-Shma’s analysis of bias amplification from scalars to matrices of arbitrary dimension in a very natural way. Curiously, while Ta-Shma’s explicit bias amplification derandomizes a well-known probabilistic argument (underlying the Gilbert–Varshamov bound), there seems to be no known probabilistic (or other existential) way of achieving our explicit operator-valued spectral amplification.

1 Introduction

1.1 Background

Expander graphs are fundamental objects in computer science and mathematics, possessing a variety of applications in both fields [HLW06, Lub12]. Indeed, expanders (and expansion) play a central role in numerous algorithmic advances, cryptographic schemes, circuit and proof complexity lower bounds, derandomization and pseudorandom generators, error correcting codes, … and are central to structural results in group theory, algebra, number theory, geometry, combinatorics.

A central quality measure of expansion of an infinite family of dd-regular graphs {Xi}i\{X_{i}\}_{i\in\mathbb{N}} is the second largest singular value of its normalized adjacency matrix, which we denote by λ(Xi)[0,1]\lambda(X_{i})\in[0,1]. We say that a family {Xi}i\{X_{i}\}_{i\in\mathbb{N}} is λ\lambda-expanding, for some fixed λ<1\lambda<1, if λ(Xi)λ\lambda(X_{i})\leq\lambda for every member XiX_{i} of the family. The smaller is the expansion parameter λ\lambda, the more spectrally expanding is the family. (For simplicity, we will sometimes discuss single graphs rather than families, and say that XX is a (d,λ)(d,\lambda)-expander if it is dd-regular and satisfies λ(X)λ\lambda(X)\leq\lambda.)

A random dd-regular graph with d3d\geq 3 is easily shown [Pin73] to be .99.99-expanding with high probability, giving rise to the existence of expanding families. The quest to explicitly111See Definition 2.2 for a formal definition. construct bounded degree expanders started with Margulis’ paper [Mar73], and has been an extremely active research area in the past half-century. Today, we have a large arsenal of constructions and tools to establish expansion which are quite different in nature—algebraic, analytic, combinatorial, and mixtures of these (for a short survey of this wealth see [Wig18, Sec 8.7])—and we will discuss a few of them below.

All the different constructions above yield dd-regular λ\lambda-expanding families with some specific constants dd and λ\lambda. Now, a large variety of structural and algorithmic applications call for optimizing both parameters, and understanding the best trade-off between them. One example which is directly related to this paper is the study of random walks on expanders, sometimes used for randomness-efficient error-reduction of probabilistic algorithms, and also in the construction of randomness extractors. The surprising expander Chernoff bound of Gillman [Gil98] informally says that a sequence of highly correlated kk vertices along a random walk in a (d,λ)(d,\lambda)-expander, is almost as good a sampler as a sequence of kk independent vertices, with respect to empirical sums. Saving randomness calls for minimizing the degree dd while improving the quality of the sample requires minimizing the expansion parameter λ\lambda.

However, for any choice of degree dd, the spectral expansion λ\lambda cannot be made arbitrarily small. The Alon–Boppana bound [Nil91] shows that λ(Xi)2d1/don(1)\lambda(X_{i})\geq 2\sqrt{d-1}/d-{o_{n}}(1). It intuitively says that the infinite dd-regular tree is the best possible spectral expander, raising the challenge of achieving it by finite graphs. This challenge was first met by the (independent) seminal papers of [LPS88, Mar88]; they constructed optimal spectrally expanding families, dubbed Ramanujan graphs, satisfying the (Ramanujan bound) λ(Xi)2d1/d\lambda(X_{i})\leq 2\sqrt{d-1}/d. The investigation of expanding families near or achieving the optimal Ramanujan bound has received much attention. However, since then, only one essentially different construction of Ramanujan graphs was found 30 years later by [MSS15].

A study of almost Ramanujan expanders, in which the bound above is nearly matched, has received much attention as well. Friedman [Fri03] greatly strengthened Pinsker’s bound above [Pin73], showing that with high probability, a random dd-regular graph XX satisfies λ(X)2d1/d+on(1)\lambda(X)\leq 2\sqrt{d-1}/d+{o_{n}}(1). For explicit constructions, one successful approach was the lifting method of Bilu–Linial [BL06], which achieves dO~(1/λ2)d\leq\widetilde{O}(1/\lambda^{2}), and famously led to the (exact) Ramanujan expanders of [MSS15] mentioned above.

A different paradigm to explicitly construct almost optimal expanders, which is central for this paper, is to start with a weak expander and amplify it to a strong one via combinatorially defined graph products. This originated with the zig-zag product of [RVW00]. They showed that their basic zig-zag construction achieves an explicit family of expanders with d1/λ4d\leq 1/\lambda^{4}. They further derandomize the basic zig-zag product to achieve d1/λ3d\leq 1/\lambda^{3}. Ben-Aroya and Ta-Shma [BATS08] in their ingenious “s-wide zig-zag product", nearly matched the optimal quadratic bound222We call any such bound near-optimal or almost Ramanujan. Of course, reducing the o(1)o(1) slack in the exponent is clearly of much interest., achieving d1/λ2+oλ(1)d\leq 1/\lambda^{2+{o_{\lambda}}(1)}. Their “higher-order” version of zig-zag [BATS08] will be central in our work. Note that these results yield specific constructions but do not provide a generic technique to amplify arbitrary expanders.

While for some applications and structural results, one is free to start with any family of expanders, for many others, the initial graphs are externally given to us (as e.g. is the case for understanding the expansion of Cayley graphs of groups). So it is highly desirable to have generic algorithms for expander construction that can operate within these constraints. Moreover, seeking different constructions and analysis tools has led to surprising applications beyond those intended (e.g., the resolution of the Kadison–Singer conjecture by [MSS14] and the proof of SL=L\mathrm{SL}=\mathrm{L} by Reingold [Rei05]). Motivated by this, we study the following question:

When can a family of weak expanders be amplified to a strong (almost-Ramanujan) one?

We show that this is always possible: any expander family can be locally and efficiently converted into an almost Ramanujan family. More precisely, starting from any family of bounded degree expanders, it is possible to obtain, for any desired target expansion λ>0\lambda>0, a new family of λ\lambda-expanders close to the Ramanujan bound.

1.2 Main Results

Our main result for general families of expander graphs is as follows.

Theorem 1.1 (Main I - Informal).

Let {Xi}i\{X_{i}\}_{i\in\mathbb{N}} be a family of (d0,λ0)(d_{0},\lambda_{0})-expanders where λ0<1\lambda_{0}<1 is a constant. For any (target) λ(0,1)\lambda\in(0,1) and XiX_{i}, we can explicitly333See Definition 2.2 construct a (d,λ)(d,\lambda)-expander, XiX_{i}^{\prime}, on the same vertex set, where d=O(d0/λ2+oλ(1))d=O(d_{0}/\lambda^{2+{o_{\lambda}}(1)}), Moreover, the construction is local in the sense that edges in XiX_{i}^{\prime} correspond to short walks in XiX_{i}.

The key insight is to use a result of König that says that the adjacency matrix, say AXA_{X}, of an arbitrary regular graph, can be written as a sum of permutation matrices. Thus we have, AX=𝔼sS[ρdef(s)]A_{X}=\mathchoice{\underset{s\sim S}{\mathbb{E}}\left[\rho_{\mathrm{def}}(s)\right]}{{\mathbb{E}}_{s\sim S}[\rho_{\mathrm{def}}(s)]}{{\mathbb{E}}_{s\sim S}[\rho_{\mathrm{def}}(s)]}{{\mathbb{E}}_{s\sim S}[\rho_{\mathrm{def}}(s)]} where SS is a set of permutations, and ρdef\rho_{\mathrm{def}} is the defining representation that maps a permutation to the matrix defining it. The task is to construct a set SS^{\prime}, with 𝔼sS[ρdef(s)]opλ\left\lVert\mathchoice{\underset{s\sim S^{\prime}}{\mathbb{E}}\left[\rho_{\mathrm{def}}(s)\right]}{{\mathbb{E}}_{s\sim S^{\prime}}[\rho_{\mathrm{def}}(s)]}{{\mathbb{E}}_{s\sim S^{\prime}}[\rho_{\mathrm{def}}(s)]}{{\mathbb{E}}_{s\sim S^{\prime}}[\rho_{\mathrm{def}}(s)]}\right\rVert_{\textup{op}}\leq\lambda. We study a general version of this question with an arbitrary group representation of a finite group (See Definition 2.3) . This is directly connected to the expansion of Cayley graphs.

Cayley Graphs

A Cayley graph Cay(G,S)\operatorname{Cay}(G,S) on a finite group GG is specified by a symmetric set of generators SGS\subseteq G, where vertices are elements of GG and g,hGg,h\in G are adjacent if and only if hg1hg^{-1} belongs to SS.

While many groups admit Cayley expanders, most of these are far from the Ramanujan bound. This is true, in particular, in the case of non-Abelian finite simple groups which includes the symmetric group. Breuillard and Lubotzky [BL18] ask whether it is possible to have near-Ramanujan expanders for all families of finite simple groups. More generally,

Which (family of) groups admit expanding Cayley graphs close to the Ramanujan bound?

Our key result is that any group that admits a Cayley expander also admits one that is almost Ramanujan.

Theorem 1.2 (Main II).

Let GG be a finite group and SS be such that Cay(G,S)\operatorname{Cay}(G,S) is a λ0\lambda_{0}-expander, for some constant λ0(0,1)\lambda_{0}\in(0,1). For every λ(0,1)\lambda\in(0,1), there exists SS^{\prime} such that

  1. \cdot

    Cay(G,S)\operatorname{Cay}(G,S^{\prime}) is a λ\lambda-expander.

  2. \cdot

    |S|=O(|S|/λ2+oλ(1))\left\lvert S^{\prime}\right\rvert=O\left(\left\lvert S\right\rvert/\lambda^{2+{o_{\lambda}}(1)}\right), and

  3. \cdot

    SS^{\prime} can be computed deterministically in poly(|S|/λ){\mathrm{poly}}(\left\lvert S\right\rvert/\lambda)-time assuming an oracle for group operations.

Furthermore, if Cay(G,S)\operatorname{Cay}(G,S) is strongly explicit444Neighbors of a vertex can be computed in time polynomial in the description length of a vertex., then so is Cay(G,S)\operatorname{Cay}(G,S^{\prime}).

Since expanding families of Cayley graphs are known for non-Abelian finite simple groups [BL18, Theorem 3.1], this result makes substantial progress towards the question asked therein (the o(1)o(1) term needs to be removed to resolve it completely). Moreover, these are strongly explicit (except for the Suzuki group). Thus, our result yields strongly explicit almost Ramanujan Cayley graphs for these these groups, which notably includes the symmetric group!

Corollary 1.3 (Explicit almost Ramanujan Cayley Expanders).

For every non-Abelian finite simple555This holds for other groups as well, as long as they have expanding generators. One non-simple example is the Cayley expanders of Rozenman, Shalev and Wigderson [RSW06]. group GG and λ>0\lambda>0, we can explicitly construct almost-Ramanujan (d,λ)(d,\lambda)-Cayley multigraphs on GG where dO(1/λ2+oλ(1))d\leq O(1/\lambda^{2+{o_{\lambda}}(1)}).

Remark 1.4 (Connection with Codes).

A linear λ0\lambda_{0}-balanced code over 𝔽2n0{\mathbb{F}}_{2}^{n_{0}} of dimension kk is equivalent to a Cayley λ0\lambda_{0}-expander over G=𝔽2kG={\mathbb{F}}_{2}^{k} of degree n0n_{0}. Let SGS\subseteq G be the rows of a generator matrix of a good λ0\lambda_{0}-balanced code (good means k/n0k/n_{0} and λ0<1\lambda_{0}<1 are constants). Thus, the breakthrough construction of explicit almost optimal binary codes of Ta-Shma [TS17] close to the Gilbert–Varshamov [Gil52, Var57] bound can be viewed as a particular case of Theorem 1.2 applied to the group G=𝔽2kG={\mathbb{F}}_{2}^{k}. Applying Theorem 1.2 above to SS with final expansion parameter λ>0\lambda>0, we obtain a generating set SGS^{\prime}\subseteq G of a Cayley λ\lambda-expander with degree O(k/λ2+oλ(1))O(k/\lambda^{2+{o_{\lambda}}(1)}), or equivalently, we obtain a λ\lambda-balanced code of rate Θ(λ2+oλ(1))\Theta(\lambda^{2+{o_{\lambda}}(1)}).

1.3 Applications

We will now discuss some applications of this operator amplification technique which allows us to improve other pseudorandom objects. All the "pseudorandom" objects below are expanders (with various structural properties), and for each of these, we amplify their spectral bound to almost Ramanujan. We stress that our amplification preserves the underlying structure and therefore, produces another object with the same properties. Precise definitions of these objects will be given in Section 5.

Quantum Expanders

Roughly speaking, a quantum expander is an operator defined by dd complex matrices, whose (linear) action on quantum states has a constant spectral gap. Quantum expanders were defined in [AS04, BASTS08, Has07a], and Hastings [Has07c] showed that the Ramanujan bound also applies to them. Existing explicit constructions are far from the Ramanujan bound. In [Har07], Harrow gave a generic construction using expanding Cayley graphs which is explicit if the group has a large irreducible representation and admits efficient Quantum Fourier Transform (QFT). Both these conditions are satisfied by the symmetric group 𝖲𝗒𝗆n\mathsf{Sym}_{n} using the generating family by Kassabov [Kas07] and the QFT algorithm by Beals [Bea97].

By amplifying the expansion of the generators of [Kas07], we give the first explicit family of almost Ramanujan quantum expanders.

Corollary 1.5 (Explicit Almost Ramanujan Quantum Expanders).

For every λ(0,1)\lambda\in(0,1), there is an explicit infinite family of (efficient) (O(1/λ2+oλ(1)),λ)(O(1/\lambda^{2+{o_{\lambda}}(1)}),\lambda)-quantum expanders.

Monotone Expanders

Monotone expanders are expanders, whose edge set can be decomposed into a constant number of monotone partial maps on [n][n]. Bourgain and Yehudayoff [BY13] gave the only known explicit construction of monotone expanders with constant degree. There are two natural notions of degree for a monotone expander. The usual vertex degree and the number of monotone maps. Our almost Ramanujan trade-off is with respect to the vertex degree (and the monotone degree is polynomial in the vertex degree).

Corollary 1.6 (Almost Ramanujan Monotone Expanders).

For every λ>0\lambda>0, there is an explicit family {Xi}i\{X_{i}\}_{i\in\mathbb{N}} of (vertex) dd-regular dO(1)d^{O(1)}-monotone expanders with d=O(1/λ2+oλ(1))d=O(1/\lambda^{2+{o_{\lambda}}(1)}) and λ(Xi)λ\lambda(X_{i})\leq\lambda.

The approach is similar to that used for Theorem 1.1; we express it as a sum of permutation matrices and amplify their expansion obtaining the following result. It would be really interesting to obtain an almost Ramanujan trade-off with respect to the monotone degree.

Dimension Expanders

Loosely speaking, dimension expanders (over any field 𝔽{\mathbb{F}}) are a linear algebraic extension of expanders: a collection of dd linear maps on 𝔽n{\mathbb{F}}^{n}, which significantly expands (the span of) any vector space of dimension below n/2n/2. They were defined by Barak et al. in [BISW01]. Over complex numbers, any quantum expander is a dimension expander. More generally, Dvir and Shpilka [DS09] proved that a monotone expander directly yields a dimension expander over every field. We give spectral almost Ramanujan expanders that have the additional property of being dimension expanders. Additionally, if the starting dimension is small enough then we achieve almost doubling of the starting dimension. See Corollary 5.18 for a precise statement.

Kazhdan Constant

We can also amplify operators in infinite dimensional Hilbert spaces. This allows us to obtain improved (average) Kazhdan constants of groups with “Property (T)”, which is an analog of expansion for discrete groups. This implies better bounds for the product replacement algorithm [LP00] to sample group elements (See Section 5 for details).

Corollary 1.7 (Amplifying Average Kazhdan Constant).

Let GG be a finitely generated group and SS a finite set of generators such that the average Kazhdan constant 𝒦¯(G,S)\overline{\mathcal{K}}(G,S) is equal to 2(1λ0)2\cdot(1-\lambda_{0}) for some constant λ0(0,1)\lambda_{0}\in(0,1). For every λ(0,1)\lambda\in(0,1), there is a set SGS^{\prime}\subseteq G such that

  1. 1.

    𝒦¯(G,S)2(1λ)\overline{\mathcal{K}}(G,S^{\prime})\geq 2\cdot(1-\lambda), and thus, 𝒦(G,S)2(1λ)\mathcal{K}(G,S^{\prime})\geq 2\cdot(1-\lambda).

  2. 2.

    |S|=Oλ0(|S|/λ2+oλ(1))\left\lvert S^{\prime}\right\rvert=O_{\lambda_{0}}(\left\lvert S\right\rvert/\lambda^{2+{o_{\lambda}}(1)}), and

  3. 3.

    SS^{\prime} can be found in time poly(|S|/λ){\mathrm{poly}}(\left\lvert S\right\rvert/\lambda) assuming an oracle for group operations on GG.

1.4 Techniques

We first rephrase the technical result of our paper from the perspective of “bias amplification”. This makes it easy to compare to prior work, which we crucially build upon. Let f:SM()f:S\rightarrow\symsf{M}_{\ell}({\mathbb{C}}) be a matrix-valued function. The quantity 𝔼sS[f(s)]op\left\lVert\mathchoice{\underset{s\sim S}{\mathbb{E}}\left[\symsf{f}(s)\right]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}\right\rVert_{\textup{op}} is known as the bias of the operator-valued function ff with respect to SS. The key idea in the prior works (and ours) is to establish an “bias amplification” result of the form,

Theorem 1.8 (Template Amplification Result).

Let SS be a finite set and λ0(0,1)\lambda_{0}\in(0,1) be a constant. For every λ>0\lambda>0, there exists a deterministic polynomial time algorithm to construct WStW\subseteq S^{t} of size |W|poly(|S|,1/λ)\left\lvert W\right\rvert\leq{\mathrm{poly}}(|S|,1/\lambda) such that

  • -

    Scalar Amplification For every function f:S\symsf{f}:S\rightarrow{\mathbb{C}} such that f1\left\lVert f\right\rVert_{\infty}\leq 1,
    if |𝔼sS[f(s)]|λ0\left\lvert\mathchoice{\underset{s\sim S}{\mathbb{E}}\left[\symsf{f}(s)\right]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}\right\rvert\leq\lambda_{0} then we have |𝔼(s1,,st)W[f(st)f(s1)]|λ\left\lvert\mathchoice{\underset{(s_{1},\cdots,s_{t})\sim W}{\mathbb{E}}\left[\symsf{f}(s_{t})\cdots f(s_{1})\right]}{{\mathbb{E}}_{(s_{1},\cdots,s_{t})\sim W}[\symsf{f}(s_{t})\cdots f(s_{1})]}{{\mathbb{E}}_{(s_{1},\cdots,s_{t})\sim W}[\symsf{f}(s_{t})\cdots f(s_{1})]}{{\mathbb{E}}_{(s_{1},\cdots,s_{t})\sim W}[\symsf{f}(s_{t})\cdots f(s_{1})]}\right\rvert\leq\lambda.

  • -

    Operator Amplification For every function f:SM()\symsf{f}:S\rightarrow\symsf{M}_{\ell}({\mathbb{C}}) such that maxsf(s)op1\max_{s}\left\lVert\symsf{f}(s)\right\rVert_{\textup{op}}\leq 1, if 𝔼sS[f(s)]opλ0\left\lVert\mathchoice{\underset{s\sim S}{\mathbb{E}}\left[\symsf{f}(s)\right]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}\right\rVert_{\textup{op}}\leq\lambda_{0} then we have 𝔼(s1,,st)W[f(st)f(s1)]opλ\left\lVert\mathchoice{\underset{(s_{1},\cdots,s_{t})\sim W}{\mathbb{E}}\left[\symsf{f}(s_{t})\cdots f(s_{1})\right]}{{\mathbb{E}}_{(s_{1},\cdots,s_{t})\sim W}[\symsf{f}(s_{t})\cdots f(s_{1})]}{{\mathbb{E}}_{(s_{1},\cdots,s_{t})\sim W}[\symsf{f}(s_{t})\cdots f(s_{1})]}{{\mathbb{E}}_{(s_{1},\cdots,s_{t})\sim W}[\symsf{f}(s_{t})\cdots f(s_{1})]}\right\rVert_{\textup{op}}\leq\lambda.

Such a result is closely related to expansion of Cayley graphs in the following way. For a group GG, a set SGS\subseteq G is called an ε\varepsilon-biased set if for every non-trivial irreducible representation, ρ\rho, of GG, we have 𝔼sS[ρ(s)]opε\left\lVert\mathchoice{\underset{s\sim S}{\mathbb{E}}\left[\rho(s)\right]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}\right\rVert_{\textup{op}}\leq\varepsilon. Here, a group representation is a homomorphism from GG to the general linear group of operators, M()\symsf{M}_{\ell}({\mathbb{C}}). The notion of ε\varepsilon-biased set over G=2kG={\mathbb{Z}}^{k}_{2} was introduced in the pioneering work of Naor and Naor [NN90]. These small-bias sets have found numerous applications (e.g., [ABN+92, Vad12, TS17]).

A symmetric subset SGS\subseteq G is an ε\varepsilon-biased set if and only if the Cay(G,S)\operatorname{Cay}(G,S) is an ε\varepsilon-expander. Therefore if Cay(G,S)\operatorname{Cay}(G,S) is a given λ0\lambda_{0}-expander, we can apply Theorem 1.8 for each irreducible representation ρ\rho. This yields that Cay(G,S)\operatorname{Cay}(G,S^{\prime}) is a λ\lambda-expander where S={sts1(s1,,st)W}S^{\prime}=\{s_{t}\cdots s_{1}\mid(s_{1},\cdots,s_{t})\in W\}. The degree of the new graph is |S|=|W||S^{\prime}|=|W|. Note that over Abelian groups, the scalar amplification suffices as the irreducible representations of Abelian groups are 11-dimensional, i.e., scalar-valued functions called characters. However, for general finite groups, one needs the amplification result for operator-valued functions.

A first approach

One can take W=StW=S^{t} with tlogλ0(λ)t\approx\log_{\lambda_{0}}(\lambda). However, now the degree, |S|t=O(1/λ100)|S|^{t}=O(1/\lambda^{100}), has also increased and the trade-off remains the same. Thus, we want to efficiently compute a sparse subset of StS^{t} that retains the expansion. Since we know what degree we are aiming for, we could try to take a sparse random sample SStS^{\prime}\subseteq S^{t} of size d=O(1/λ2.001)d=O(1/\lambda^{2.001}) and hope that some form of matrix concentration ensures that Cay(G,S)\operatorname{Cay}(G,S^{\prime}) is λ\lambda^{\prime}-spectral expander with λλ\lambda^{\prime}\approx\lambda. Unfortunately, it is not clear how to show even the existence of a single sparse subset SS^{\prime} that achieves the required expansion.

For example, one could use matrix Chernoff plus union bound to select a subset SS^{\prime} such that 𝔼sS[ρ(s)]op\left\lVert\mathchoice{\underset{s\in S^{\prime}}{\mathbb{E}}\left[\rho(s)\right]}{{\mathbb{E}}_{s\in S^{\prime}}[\rho(s)]}{{\mathbb{E}}_{s\in S^{\prime}}[\rho(s)]}{{\mathbb{E}}_{s\in S^{\prime}}[\rho(s)]}\right\rVert_{\textup{op}} is small for any irreducible representation ρ\rho. However, this naive approach requires |S|Ω(logdim(ρ))|S^{\prime}|\geq\Omega(\log\dim(\rho)). For non-abelian groups, we have irreducible representations such that dim(ρ)=poly(|G|)\dim(\rho)={\mathrm{poly}}(|G|), and therefore, this cannot deduce the existence of constant-sized subsets that achieve expansion. This difficulty is also present in the proof of the Alon–Roichman theorem [AR94] and the reason why even for non-Abelian groups the only generic upper bound known uses Ω(log(|G|))\Omega(\log(\left\lvert G\right\rvert)) random generators to obtain an expander. Nevertheless, we will show that, when applied correctly, the equivalence between expansion and small bias distribution can be used to build almost-Ramanujan expanders.

Prior Results on Bias Amplification

Scalar amplification

Much of the earlier work has focused on the case of Abelian groups, especially G=2nG={\mathbb{Z}}_{2}^{n}, as this has connections to other objects like error-correcting codes.

Rozenman and Wigderson introduced the approach of using expanders for “scalar amplification” via an iterated application of the expander mixing lemma (see Section 6 for details). Alon (in an unpublished email666We thank the anonymous reviewer for pointing out this reference.) introduced the idea of using walks on an (auxiliary) expander graph XX, whose vertices are identified with elements of SS. The set WStW\subseteq S^{t} is chosen to be the collection of all walks of length (t1)(t-1) on XX. This technique gives a λ\lambda-biased set WW, of size |W|O(|S|/λ4+o(1))|W|\leq O(|S|/\lambda^{4+o(1)}) (cf., [TS17]), which is quite good but still sub-optimal (λ4\lambda^{-4} instead of λ2\lambda^{-2} ).

Ta-Shma [TS17] managed to close the gap almost optimally (λ2o(1)\lambda^{-2-o(1)}) using the ss-wide replacement product to derandomize the above amplification. The ss-wide replacement product of Ben-Aroya and Ta-Shma [BATS08] is a higher-order version of the zig-zag product [RVW00]. Using the collection of walks on the ss-wide replacement product allows for a much smaller collection WStW\subseteq S^{t} with nearly optimal size. This scalar technique was later applied to the more general case of arbitrary Abelian groups by Jalan and Moshkowitz [JM21].

Operator amplification

To extend Ta-Shma’s approach to non-Abelian groups, it is necessary to work with operator-valued functions, f:SM()\symsf{f}\colon S\rightarrow\symsf{M}_{\ell}({\mathbb{C}}), as the irreducible representations are no longer of dimension one. To the best of our knowledge, only one general result was known for general groups. Chen, Moore and Russell [CMR13] analyzed the above expander walk construction using a matrix version of the expander mixing lemma. This gives an amplification procedure for Cayley graphs of general groups, but the resulting degree O(|S|/λ11)O(\left\lvert S\right\rvert/\lambda^{11}) to achieve final expansion λ\lambda is sub-optimal.

Our Results

We view our main contribution as the identification of appropriate natural linear algebraic extensions to Ta-Shma’s amplification framework [TS17]. This gives an almost-optimal dimension independent generalization of the scalar amplification result to operator-valued functions. Our result sharpens that of Chen, Moore and Russell [CMR13] be reducing the degree from O(|S|/λ11)O(\left\lvert S\right\rvert/\lambda^{11}) to O(|S|/λ2+o(1))O(\left\lvert S\right\rvert/\lambda^{2+o(1)})

Theorem 1.9 (Operator Amplification (this work)).

Let SS be a finite set and λ0(0,1)\lambda_{0}\in(0,1) be a constant. For every λ>0\lambda>0, there exists a deterministic polynomial time algorithm to construct WStW\subseteq S^{t} of size |W|O(|S|/λ2+o(1))\left\lvert W\right\rvert\leq O(\left\lvert S\right\rvert/\lambda^{2+o(1)}) such that for every function f:SM()\symsf{f}:S\rightarrow\symsf{M}_{\ell}({\mathbb{C}}) with 𝔼sS[f(s)]opλ0\left\lVert\mathchoice{\underset{s\sim S}{\mathbb{E}}\left[\symsf{f}(s)\right]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}{{\mathbb{E}}_{s\sim S}[\symsf{f}(s)]}\right\rVert_{\textup{op}}\leq\lambda_{0} and maxsf(s)op1\max_{s}\left\lVert\symsf{f}(s)\right\rVert_{\textup{op}}\leq 1, we have 𝔼wW[f(w)]opλ\left\lVert\mathchoice{\underset{w\sim W}{\mathbb{E}}\left[\symsf{f}(w)\right]}{{\mathbb{E}}_{w\sim W}[\symsf{f}(w)]}{{\mathbb{E}}_{w\sim W}[\symsf{f}(w)]}{{\mathbb{E}}_{w\sim W}[\symsf{f}(w)]}\right\rVert_{\textup{op}}\leq\lambda.

The key extension is a simple and yet extremely useful change in the bias operator (Πf\symsf{\Pi}_{\symsf{f}}) defined by Ta-Shma which is a central object in the analysis of both [TS17] and [JM21]. In both these cases, f\symsf{f} is scalar, and they define,

Πf:[S][S] whereΠfs=f(s)s.\displaystyle\symsf{{\symsf{\Pi}}_{\symsf{f}}}:{\mathbb{C}}[S]\rightarrow{\mathbb{C}}[S]\;\text{ where}\;\symsf{{\symsf{\Pi}}_{\symsf{f}}}\cdot s=\symsf{f}(s)\cdot s\,.

However, this approach is not readily generalizable to operators and the view we take is that if f:SM()\symsf{f}:S\rightarrow\symsf{M}_{\ell}({\mathbb{C}}), then, Πf{\symsf{\Pi}}_{\symsf{f}} is actually an operator on [S]{\mathbb{C}}^{\ell}\otimes{\mathbb{C}}[S] defined as

Πf:[S][S] whereΠf(vs)=f(s)vs.\displaystyle\symsf{{\symsf{\Pi}}_{\symsf{f}}}:{\mathbb{C}}^{\ell}\otimes{\mathbb{C}}[S]\rightarrow{\mathbb{C}}^{\ell}\otimes{\mathbb{C}}[S]\;\text{ where}\;\symsf{{\symsf{\Pi}}_{\symsf{f}}}\,(v\otimes s)=\symsf{f}(s)\,v\otimes s\,.

Clearly, in the Abelian case, we have =1\ell=1 and this is isomorphic to the setup by Ta-Shma. This generalization is very natural and we show that not only does the older machinery gel well with this, but the proof remains intuitive with the different spaces neatly delineated.

More precisely, we first establish an operator version of the expander walk amplification, and then we derandomize it using (a suitable version of) the ss-wide replacement product. Furthermore, since the result does not depend on the dimension, \ell, we can use it even for functions f:S()\symsf{f}:S\rightarrow\mathcal{L}(\mathcal{H}) where ()\mathcal{L}(\mathcal{H}) is the space of bounded linear operators on an arbitrary Hilbert space, \mathcal{H}, possibly infinite dimensional. This is useful if the underlying group is not finite but finitely generated by SS.

1.5 Discussion

The results of this paper have some curious features, which we would like to elaborate on. For most of them, we will use the following "bare bones" description of our main spectral amplification result. Namely, let SS be a finite set and {\mathcal{H}} a Hilbert space. Let f\symsf{f} be a function mapping elements of SS to operators on {\mathcal{H}} of unit norm, such that 𝔼sS[f(s)]opλ0\left\lVert{\mathbb{E}}_{s\in S}[\symsf{f}(s)]\right\rVert_{\textup{op}}\leq\lambda_{0}. For any λ>0\lambda>0 take t=clog(1/λ)t=c\log(1/\lambda) (for appropriate cc). We extend f\symsf{f} from SS to StS^{t} by defining f(s1,,st)=f(s1)f(st)\symsf{f}(s_{1},\ldots,s_{t})=\symsf{f}(s_{1})\cdots\symsf{f}(s_{t}). Clearly, 𝔼rSt[f(r)]op(𝔼sS[f(s)]op)tλ\left\lVert{\mathbb{E}}_{r\in S^{t}}[\symsf{f}(r)]\right\rVert_{\textup{op}}\leq(\left\lVert{\mathbb{E}}_{s\in S}[\symsf{f}(s)]\right\rVert_{\textup{op}})^{t}\leq\lambda. Our main result is an explicit construction of a (pseudorandom) subset SStS^{\prime}\subseteq S^{t}, of size only |S|=O(|S|/λ2+o(1))|S^{\prime}|=O(|S|/\lambda^{2+o(1)}), with a similar guarantee, namely 𝔼sS[f(s)]opλ\left\lVert{\mathbb{E}}_{s^{\prime}\in S^{\prime}}[\symsf{f}(s^{\prime})]\right\rVert_{\textup{op}}\leq\lambda.

Dimension Independence

Note that if the operators in SS are 1-dimensional, namely scalars, then the existence of a set SS^{\prime} of this size (which is best possible even in this 1-dimensional case) follows directly from the Chernoff bound. Indeed, Ta-Shma’s construction [TS17] may be viewed as derandomizing this result, producing an explicit such SS^{\prime}.

One may try to do the same for operators in a higher dimension, say \ell, by appealing to the Matrix Chernoff bounds of Ahlswede–Winter [AW02] (see also Tropp [Tro15]). However, these concentration inequalities pay a factor of \ell in the tail bound, resulting in a set SS^{\prime} of size Ω(log())\Omega(\log(\ell)). As the dimension \ell is arbitrary (indeed, may be infinite), such a bound is useless.

Thus, our explicit construction has no known probabilistic (or other existential) analog! What is curious is that our dimension-independent analysis follows very closely that of Ta-Shma for 1-dimension, roughly speaking, replacing scalar absolute values by the operator norm in any dimension. We feel that it would be extremely interesting to find a matrix concentration inequality for sampling product sets like StS^{t}, which is dimension independent.

Algebraic vs.  Combinatorial Expander Constructions

Our explicit construction of the pseudorandom set SS^{\prime} above uses expanders obtained from the ss-wide zig-zag product of [BATS08]. This is a combinatorial construction, a refinement of the original zig-zag product construction of [RVW00]. Nonetheless, it has significant consequences to algebraic expander constructions which use group theory, namely to the expansion of Cayley graphs over the same group. We can take SS to be an expanding generating set of a group and f\symsf{f} to be some non-trivial irreducible representation ρ\rho. Instead of using the tt-product of elements of StS^{t} to obtain a new amplified generating set SS^{\prime}, a much sparser subset is chosen using the ss-wide zig-zag construction. The analysis of the norm amplification discussed above yields the required expansion bound, in a way that has no dependence on the group or the representation. The flexibility in mapping element of SS to operators underlies the versatility of our spectral amplification. It allows us to preserve some of the structure of the expanders whose expansion are being amplified. In this case, both the starting expander and the amplified expander are Cayley graph over the same group.

It is interesting to note that this is a recurring phenomenon. In [ALW01], it was discovered that the zig-zag product may be viewed as a combinatorial generalization of the algebraic semi-direct product of groups. This connection made possible the construction of new expanding Cayley graphs in groups that are far from being simple, e.g., in [MW04, RSW06]. It is rewarding to see again how new combinatorial constructions, sometimes inferior in certain parameters to some algebraic ones, yield new results in group theory.

Iterated Pseudorandomness

Another interesting aspect of our result is the following. Recall that expanders are pseudorandom objects for many purposes. One important purpose is sampling - rather than sampling tt independent random elements in some set SS, one may sample tt points along a random walk on an expander on the vertex set SS and a Chernoff type bound still holds (a nontrivial result of [Gil98])- this affords significant savings in the number of random bits spent. For this result, any expander would do. What happens in this paper is an iterated use of expanders as samplers as follows. We first choose a sparse pseudorandom set of tt-walks inside StS^{t} using expanders walks. Then, we choose a yet sparser pseudorandom set inside it, again using walks on an additional expander. This repeated use of expanders improves the trade-off between the quality of spectral amplification and the size of the final pseudorandom set to near-optimal. The construction of this iterated selection of walks seems critical, and (as in Ta-Shma’s paper) is chosen to come from the ss-wide zig-zag product of two expanders [BATS08].

Groups and Expansion

For us, the most surprising consequence of our results is that “weak” simple groups, especially the symmetric group,777See also the groups in [RSW06], which are iterated wreath products of symmetric groups. can have near-Ramanujan generators. The question of which groups are expanding, and just how expanding they are, is an old quest of group theory. One dichotomy is whether every finite set of generators of the group is expanding (these are “strongly expanding” groups), or if some are and some aren’t (these are “weakly expanding” groups). For the symmetric group, many finite non-expanding generating sets of constant size were long known, and Kassabov’s breakthrough construction [Kas07] designed a constant size expanding generating set. The symmetric group is then a weakly expanding group, while, e.g., simple groups of Lie type (namely, matrix groups) are believed, and in some cases known, to be strongly expanding. Nonetheless, our construction works equally well for all, and we have almost Ramanujan generators for all expanding groups.

Closeness to the Ramanujan Bound

As mentioned above, a family of dd-regular graphs is called Ramanujan if its spectral expansion parameter λ\lambda is at most 2d1/d2\sqrt{d-1}/d. This terminology was introduced in the seminal work of Lubotzky, Phillips and Sarnak [LPS88], and it designates the optimal degree versus expansion trade-off that a family of bounded degree expanders can achieve. Several notions of closeness to the Ramanujan bound were investigated, e.g., (2d1+ε)/d(2\sqrt{d-1}+\varepsilon)/d (with ε>0\varepsilon>0 small or vanishing) in [Fri03, MOP20, JMO+22], O(1/d)O(1/\sqrt{d}) in [ACKM19], polylog(d)/d{\mathrm{polylog}}(d)/\sqrt{d} in [BL06, JMO+22] and more generally dod(1)/d1/2d^{o_{d}(1)}/d^{1/2}.

In this work, we obtain a bound of the form λO(2log(d)c/d1/2)\lambda\leq O(2^{\log(d)^{c}}/d^{1/2}) for some constant c(0,1)c\in(0,1), which we refer to as an almost Ramanujan bound. Rephrasing in terms of the expansion parameter, we achieve expansion λ\lambda with degree O(1/λ2+β)O(1/\lambda^{2+\beta}), where β=O(1/log(1/λ))c\beta=O(1/\log(1/\lambda))^{c^{\prime}} for some c(0,1)c^{\prime}\in(0,1). We stress that the nomenclatures almost Ramanujan, near Ramanujan and etc, may vary depending on the author. Improving the results in this work to achieve trade-offs even closer to the Ramanujan bound (if possible) is of great interest. We suspect that new ideas may be required to substantially improve the bound to, say, expansion λ\lambda versus degree O(polylog(1/λ)/λ2)O({\mathrm{polylog}}(1/\lambda)/\lambda^{2}).

1.6 Outline

We start in Section 2 by summarizing basic definitions and the notation used throughout the paper. In Section 3, we generalize the simpler construction of Ta-Shma based on expander walks. Apart from serving as a nice warm-up to the more-involved construction, it will be used as a bootstrap for the more involved construction based on ss-wide replacement product which is the subject of Section 4. Here, we prove the main amplification result (a formal version of Theorem 1.9 above) and instantiate using known constructions and those obtained from Section 3 which establishes Theorem 1.2. Section 5 discusses the permutation amplification trick and formally completes the proof of Theorem 1.1. It also discusses the other applications in more detail. Finally, Section 6 gives an operator version of the expander mixing lemma which improves the analysis of [CMR13].

2 Preliminaries

Let X=(V,E)X=(V,E) be an nn-vertex dd-regular undirected multigraph for some d1d\geq 1. We denote by AX\symsf{A}_{X} the normalized adjacency matrix of XX, i.e., the adjacency matrix divided by dd.

Definition 2.1 (λ\lambda-spectral Expander).

Let the eigenvalues of the symmetric matrix AX\symsf{A}_{X}, denoted as Spec(AX)\operatorname{Spec}(A_{X}), be {1=λ1λn}\{1=\lambda_{1}\geq\cdots\geq\lambda_{n}\} and define λ(X)=max{|λ2|,|λn|}\lambda(X)=\max\{\left\lvert\lambda_{2}\right\rvert,\left\lvert\lambda_{n}\right\rvert\}. We say that XX is a λ\lambda-spectral expander if λ(X)λ\lambda(X)\leq\lambda.

We denote by GG a finite group (except in Section 5.5 where we only need it to be finitely generated). For a multiset SGS\subseteq G, Cay(G,S)\operatorname{Cay}(G,S) denotes a multigraph with the vertex set being GG and edges {(g,sg)|gG,sS}\{(g,sg)\;|\;g\in G,\;s\in S\}. We will assume throughout that S=S1S=S^{-1} and therefore, the graph is undirected.

Definition 2.2 (Explicit graph).

A family of graphs {Xi}i\{X_{i}\}_{i\in{\mathbb{N}}} is said to be explicit if the adjacency matrix of XiX_{i} can be computed deterministically in poly(|Xi|){\mathrm{poly}}(|X_{i}|)-time. Moreover, it is said to be strongly explicit if the list of its neighbors of any vertex in XiX_{i} can be computed poly(log|Xi|){\mathrm{poly}}(\log|X_{i}|)-time.

Group Representations

In order to study the expansion of a Cayley graph, we will use the notion of a group representation888Additional background on representation theory of finite groups can be found in [SS96].. Weyl’s unitary trick, says that for a large family of groups (which includes all finite groups), every representation can be made unitary and thus, we can restrict to studying these.

Let \mathcal{H} be a complex Hilbert space and denote by ()\mathcal{L}(\mathcal{H}) the algebra of bounded linear operators999For most applications, one can think of =n\mathcal{H}={\mathbb{C}}^{n} for some nn, and ()=Mn()\mathcal{L}(\mathcal{H})=\symsf{M}_{n}({\mathbb{C}}), the space of n×nn\times n complex matrices. However, we will need the generality in Section 5.5. on \mathcal{H}. We denote by U\textup{U}_{\mathcal{H}} the unitary group of operators acting on \mathcal{H}.

Definition 2.3 (Group Representation).

For a group GG, a representation is a pair (ρ,)(\rho,\mathcal{H}) where ρ:GU\rho:G\rightarrow\textup{U}_{\mathcal{H}} is a group homomorphism, i.e., for every g1,g2Gg_{1},g_{2}\in G, we have ρ(g1g2)=ρ(g1)ρ(g2)\rho(g_{1}g_{2})=\rho(g_{1})\,\rho(g_{2}). A representation is irreducible if the only subspaces of \mathcal{H} that are invariant under the action of ρ(G)\rho(G) are the empty space, {0}\{0\}, and the entire space, \mathcal{H}.

Every group has two special representations, which are,

  1. 1.

    (Trivial representation ) – (ρ,)(\rho,{\mathbb{C}}) where for every gg, ρ(g)=1\rho(g)=1.

  2. 2.

    ((left) Regular representation ) – (ρreg,𝒱reg)(\rho_{\mathrm{reg}},\mathcal{V}_{\mathrm{reg}}) where, 𝒱reg=[G]\mathcal{V}_{\mathrm{reg}}={\mathbb{C}}[G] is a vector space with the elements of GG being an orthonormal basis, and ρreg(g):hgh\rho_{\mathrm{reg}}(g):h\mapsto g\cdot h.

Fact 2.4.

Let GG be a finite group and let 𝒱reg\mathcal{V}_{\mathrm{reg}} be the regular representation over \mathbb{C}. We have

𝒱reg(ρ,Vρ)Irrep(G)dim(ρ)𝒱ρ,\displaystyle\mathcal{V}_{\mathrm{reg}}\cong\bigoplus_{(\rho,V_{\rho})\in\textup{Irrep}(G)}\textup{dim}(\rho)\cdot\mathcal{V}_{\rho},

where Irrep(G)\textup{Irrep}(G) denotes the set of irreducible representations of GG.

Expanders and Biased Distributions

It follows from definitions that the normalized adjacency matrix of Cay(G,S)\operatorname{Cay}(G,S) is given by A=𝔼sS[ρreg(s)]\symsf{A}=\mathchoice{\underset{s\sim S}{\mathbb{E}}\left[\rho_{\mathrm{reg}}(s)\right]}{{\mathbb{E}}_{s\sim S}[\rho_{\mathrm{reg}}(s)]}{{\mathbb{E}}_{s\sim S}[\rho_{\mathrm{reg}}(s)]}{{\mathbb{E}}_{s\sim S}[\rho_{\mathrm{reg}}(s)]}. Moreover, the copy of the trivial representation is the space spanned by the all-ones vector. 2.4 implies that this can be block diagonalized and therefore,

Spec(A)\displaystyle\operatorname{Spec}(\symsf{A}) =ρIrrep(G)Spec(𝔼sS[ρ(s)]), and thus,\displaystyle=\bigcup_{\rho\in\textup{Irrep}(G)}\operatorname{Spec}(\mathchoice{\underset{s\sim S}{\mathbb{E}}\left[\rho(s)\right]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}),\;\;\;\text{ and thus, }
λ(Cay(G,S))\displaystyle\lambda(\operatorname{Cay}(G,S)) =max \Let@\restore@math@cr\default@tag ρIrrep(G) ρ is non-trivial 𝔼sS[ρ(s)]op.\displaystyle=\max_{\vbox{ \Let@\restore@math@cr\default@tag\halign{\hfil$\m@th\scriptstyle#$&$\m@th\scriptstyle{}#$\cr\rho\in\textup{Irrep}(G)\\ \rho\text{ is non-trivial }\crcr} }}\left\lVert\mathchoice{\underset{s\sim S}{\mathbb{E}}\left[\rho(s)\right]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}\right\rVert_{\textup{op}}\,.

For any bounded linear operator, T:\symsf{T}\colon\mathcal{H}\rightarrow\mathcal{H}^{\prime}, between Hilbert spaces, we have

Top=supv:v=1Tv=supv,w:v=w=1|Tv,w|,\left\lVert\symsf{T}\right\rVert_{\textup{op}}=\sup_{v\in\mathcal{H}:\left\lVert v\right\rVert=1}\left\lVert\symsf{T}v\right\rVert=\sup_{v\in\mathcal{H},w\in\mathcal{H}^{\prime}:\left\lVert v\right\rVert=\left\lVert w\right\rVert=1}\left\lvert\left\langle\symsf{T}v,w\right\rangle\right\rvert\,,

where v=v,v\left\lVert v\right\rVert=\sqrt{\left\langle v,v\right\rangle_{\mathcal{H}}} and w=w,w\left\lVert w\right\rVert=\sqrt{\left\langle w,w\right\rangle_{\mathcal{H}^{\prime}}}. Given this equivalence, we will find it convenient to work with the operator norm version referred to as bias in the literature [CMR13].

Definition 2.5 (Biased Set on GG).

Let ε(0,1)\varepsilon\in(0,1). We say that a multiset SS of elements of a group GG is ε\varepsilon-biased if for every non-trivial irreducible representation ρ\rho, we have 𝔼sS[ρ(s)]opε\left\lVert\mathchoice{\underset{s\sim S}{\mathbb{E}}\left[\rho(s)\right]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}\right\rVert_{\textup{op}}\leq\varepsilon. We sometimes use the shorthand 𝖻𝗂𝖺𝗌(S)ε\mathsf{bias}(S)\leq\varepsilon, where 𝖻𝗂𝖺𝗌(S)=λ(Cay(G,S))\mathsf{bias}(S)=\lambda(\operatorname{Cay}(G,S)).

Irreducible representations of Abelian groups, called characters, have dimension 11. Thus, this definition coincides with the usual one of ε\varepsilon-biased distribution fooling non-trivial characters [NN90, AGHP92]. These pseudorandom distributions were introduced in the pioneering work of Naor and Naor where several applications to derandomization were given [NN90].

Notation

Since we deal with various vector spaces and graphs, we will find it useful to establish some convenient notation. While we recall these in the relevant section, the following is a summary for ready reference.

  1. \cdot

    The main multigraphs we study will be XX and YY with vertices VX,VYV_{X},V_{Y} and normalized adjacency operators AX,AY\symsf{A}_{X},\symsf{A}_{Y}.

  2. \cdot

    We denote vertices of X,YX,Y by x,yx,y and an ordered tuple of vertices by x=(x0,,xt)\vec{x}=(x_{0},\cdots,x_{t}).

  3. \cdot

    We use u,v,wu,v,w to denote arbitrary vectors in \mathcal{H} and x,yx,y for basis vectors of [VX],[VY]{\mathbb{C}}[V_{X}],{\mathbb{C}}[V_{Y}] where [VX]{\mathbb{C}}[V_{X}] is the complex vector space with the elements of VXV_{X} being a orthonormal basis.

  4. \cdot

    The tensored vector spaces have an induced inner product. For 𝒳[VX]\mathcal{X}_{\mathcal{H}}\coloneqq\mathcal{H}\otimes{\mathbb{C}}[V_{X}], it is vx,wx=v,wx,x\left\langle v\otimes x,w\otimes x^{\prime}\right\rangle=\left\langle v,w\right\rangle_{\mathcal{H}}\left\langle x,x^{\prime}\right\rangle. Similarly, we have one on 𝒳𝒴𝒳[VY]\mathcal{XY}_{\mathcal{H}}\coloneqq\mathcal{X}_{\mathcal{H}}\otimes{\mathbb{C}}[V_{Y}].

  5. \cdot

    Orthogonal decomposition: 𝒳=𝒳𝒳\mathcal{X}_{\mathcal{H}}=\mathcal{X}_{\mathcal{H}}^{\parallel}\oplus\mathcal{X}_{\mathcal{H}}^{\perp} where 𝒳span{v1v}\mathcal{X}_{\mathcal{H}}^{\parallel}\coloneqq\textup{span}\{v\otimes\vec{1}\mid v\in\mathcal{H}\}. Here, 1\vec{1} denotes the un-normalized all-ones vector. Similarly, 𝒳𝒴=𝒳𝒴𝒳𝒴\mathcal{XY}_{\mathcal{H}}=\mathcal{XY}_{\mathcal{H}}^{\parallel}\oplus\mathcal{XY}_{\mathcal{H}}^{\perp}, where 𝒳𝒴span{z1z𝒳}\mathcal{XY}_{\mathcal{H}}^{\parallel}\coloneqq\textup{span}\{z\otimes\vec{1}\mid z\in\mathcal{X}_{\mathcal{H}}\}.

  6. \cdot

    The operator A\smash{\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}} denotes the extension of operator A\symsf{A} to a tensor product of spaces where it acts as identity on the other spaces. For example, AX\symsf{A}_{X} acts on [VX]{\mathbb{C}}[V_{X}] and its extension to 𝒳\mathcal{X}_{\mathcal{H}} is AX=IAX\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}=\symsf{I}_{\mathcal{H}}\otimes\symsf{A}_{X}. However, if we were working on 𝒳𝒴\mathcal{XY}_{\mathcal{H}}, it would be AX=IAXIY\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}=\symsf{I}_{\mathcal{H}}\otimes\symsf{A}_{X}\otimes\symsf{I}_{Y} instead101010The spaces will be self-evident and the use of the same notation should not be confusing..

  7. \cdot

    Given an operator valued function f:VX()\symsf{f}:V_{X}\rightarrow\mathcal{L}(\mathcal{H}), the generalized bias operator is defined on the basis as111111An equivalent matrix definition is ΠfxVXf(x)Ex,x\symsf{\Pi}_{\symsf{f}}\coloneqq\sum_{x\in V_{X}}\symsf{f}(x)\,\otimes\symsf{E}_{x,x} where Ex,xVX×VX\symsf{E}_{x,x}\in\mathbb{C}^{V_{X}\times V_{X}} is the diagonal matrix with exactly one non-zero entry of value 11 in the row and column indexed by the vertex xx.,

    Πf:𝒳𝒳,vxf(x)vx.{\symsf{\Pi}}_{\symsf{f}}\colon\mathcal{X}_{\mathcal{H}}\rightarrow\mathcal{X}_{\mathcal{H}},\;\ v\otimes x\mapsto\symsf{f}(x)\,v\otimes x.

3 Operator Bias Reduction via Expander Walks

In this section, we establish a new operator analog of the expander walk–based bias amplification procedure for scalars. An analysis of this scalar amplification was given by Ta-Shma in [TS17]. We prove an operator analog that can amplify from any bias (Theorem 3.6) which implies the main result below (Theorem 3.1), that amplifies from constant bias.

Theorem 3.1 (Operator Amplification via Expander Walks).

Let XX be a λ(X)\lambda(X)-spectral expander, and let 𝒲t\mathcal{W}_{t} be the collection of walks obtained from walks of length tt on XX. Then for any operator valued function f\symsf{f} such that 𝔼xVX[f(x)]opλ0\left\lVert\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right\rVert_{\textup{op}}\leq\lambda_{0} and maxxVXf(x)op1\max_{x\in V_{X}}\left\lVert\symsf{f}(x)\,\right\rVert_{\textup{op}}\leq 1, we have

𝔼(x0,xt)𝒲t[f(xt)f(x0)]op(2λ(X)+λ0)t/2.\displaystyle\left\lVert\mathchoice{\underset{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}{\mathbb{E}}\left[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})\right]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}\right\rVert_{\textup{op}}~{}\leq~{}\left(2\lambda(X)+\lambda_{0}\right)^{\lfloor t/2\rfloor}\,.

We remark that a precursor of these techniques, in the simpler setting of Abelian groups appears in the pioneering work of Naor and Naor introducing ε\varepsilon-biased distributions over the group 2m{\mathbb{Z}}_{2}^{m} using expanders [NN90].

This simpler amplification of Theorem 3.1 will be crucially used to bootstrap the almost-optimal amplification. Moreover, it yields a construction of expanding Cayley graphs close to any desired size, which will be required later.

This bias reduction procedure uses walks on an auxiliary expander graph. Here, we only use its expansion property (as opposed to later when we rely on its structure for the ss-wide construction). With this it is already possible to obtain 1/λ4+o(1)1/\lambda^{4+o(1)} dependence on the final degree of an λ\lambda-expander.

Theorem 3.2.

Let SGS\subseteq G such that λ(Cay(G,S))=λ0<1\lambda(\operatorname{Cay}(G,S))=\lambda_{0}<1. For every λ(0,1)\lambda\in(0,1) and constant β(0,1)\beta\in(0,1), we can find SGS^{\prime}\subseteq G in time poly(|S|,1/λ0,1/λ){\mathrm{poly}}(\left\lvert S\right\rvert,1/\lambda_{0},1/\lambda) such that λ(Cay(G,S))λ\lambda(\operatorname{Cay}(G,S^{\prime}))\leq\lambda and |S|=Oλ0(|S|/λ4+β)\left\lvert S^{\prime}\right\rvert=O_{\lambda_{0}}(\left\lvert S\right\rvert/\lambda^{4+\beta}).

Notation Recall

Let SS be any finite set and let XX be a graph on the vertex set Vx=SV_{x}=S with AX\symsf{A}_{X} being its normalized adjacency matrix. Let {\mathcal{H}} be a complex Hilbert space and ()\mathcal{L}({\mathcal{H}}) be the (bounded) operators on {\mathcal{H}}; an important example will be ()=M()\mathcal{L}({\mathcal{H}})=\symsf{M}_{\ell}({\mathbb{C}}). For any operator valued function, f:S()\symsf{f}\colon S\rightarrow\mathcal{L}({\mathcal{H}}), we define the generalized bias operator as

Πf:[VX][VX],Πf(vx)=f(x)vx.\displaystyle{\symsf{\Pi}}_{\symsf{f}}:\mathcal{H}\otimes{\mathbb{C}}[V_{X}]\mapsto\mathcal{H}\otimes{\mathbb{C}}[V_{X}],\;\;{\symsf{\Pi}}_{\symsf{f}}(v\otimes x)=\symsf{f}(x)\,v\otimes x\,.

In the scalar case, since =\mathcal{H}={\mathbb{C}}, earlier works [TS17, JM21] used the implicit identification [VX][VX]{\mathbb{C}}\otimes{\mathbb{C}}[V_{X}]\cong{\mathbb{C}}[V_{X}] and defined Πf{\symsf{\Pi}}_{\symsf{f}} as a diagonal matrix. This identification is no longer is suitable when f\symsf{f} is operator valued in dimension >1>1. However, a simple yet crucial observation is that merely decoupling the spaces allows us to collect the terms as we proceed along the walk.

3.1 Operator Norm Decay

Lemma 3.3.

Let 𝒲tVXt+1\mathcal{W}_{t}\subseteq V_{X}^{t+1} be the collection of all length tt walks on the graph XX and we define AX=IAX\smash{\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}=\symsf{I}_{\mathcal{H}}\otimes\symsf{A}_{X}}. Then, we have

𝔼(x0,xt)𝒲t[f(xt)f(x0)]opΠf(AXΠf)top(AXΠf)2opt/2.\left\lVert\mathchoice{\underset{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}{\mathbb{E}}\left[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})\right]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}\right\rVert_{\textup{op}}~{}\leq~{}\left\lVert{\symsf{\Pi}}_{\symsf{f}}\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{t}\right\rVert_{\textup{op}}\leq\left\lVert\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{2}\right\rVert_{\textup{op}}^{\lfloor t/2\rfloor}\,.
  • Proof:  

    Πf(AXΠf)t𝔼xVX[vx]=𝔼(xt,,x0)𝒲t[f(xt)f(x0)]vxt.{\symsf{\Pi}}_{\symsf{f}}\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{t}\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[v\otimes x\right]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}~{}=~{}\mathchoice{\underset{(x_{t},\cdots,x_{0})\in\mathcal{W}_{t}}{\mathbb{E}}\left[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})\right]}{{\mathbb{E}}_{(x_{t},\cdots,x_{0})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{t},\cdots,x_{0})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{t},\cdots,x_{0})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}v\otimes x_{t}\,. (1)

    This can be shown easily via induction on tt, and we refer to Lemma 4.7 for a formal proof of a more general statement. We use projection and lifting maps to move between the spaces 𝒳\mathcal{X}_{\mathcal{H}} and \mathcal{H}. Define P:𝒳\symsf{P}_{\mathcal{H}}:\mathcal{X}_{\mathcal{H}}\rightarrow\mathcal{H} and L:𝒳\symsf{L}_{\mathcal{H}}:\mathcal{H}\rightarrow\mathcal{X}_{\mathcal{H}}, as,

    P(wx)=w,L(v)=𝔼xVX[vx]=1|VX|v1.\displaystyle\symsf{P}_{\mathcal{H}}(w\otimes x)=w,\;\;\symsf{L}_{\mathcal{H}}(v)=\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[v\otimes x\right]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}=\frac{1}{\left\lvert V_{X}\right\rvert}\,v\otimes\vec{1}\,.

    From the definition, L(v)=v1|VX|=v|VX|\left\lVert\symsf{L}_{\mathcal{H}}(v)\right\rVert=\left\lVert v\right\rVert\frac{\left\lVert\vec{1}\right\rVert}{\left\lvert V_{X}\right\rvert}=\frac{\left\lVert v\right\rVert}{\sqrt{\left\lvert V_{X}\right\rvert}} and thus, Lop=1/|VX|\smash{\left\lVert\symsf{L}_{\mathcal{H}}\right\rVert_{\textup{op}}=1/\sqrt{\left\lvert V_{X}\right\rvert}}. We can use Cauchy-Schwarz to get that Pop=|VX|\smash{\left\lVert\symsf{P}_{\mathcal{H}}\right\rVert_{\textup{op}}=\sqrt{\left\lvert V_{X}\right\rvert}}. Now, we put this together to obtain a simple expression on the quantity we need to bound

    𝔼(x0,xt)𝒲t[f(xt)f(x0)]op\displaystyle\left\lVert\mathchoice{\underset{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}{\mathbb{E}}\left[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})\right]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}\right\rVert_{\textup{op}} =supv=1𝔼(x0,,xt)𝒲t[f(xt)f(x0)]v2\displaystyle~{}=~{}\sup_{\left\lVert v\right\rVert=1}\left\lVert\mathchoice{\underset{(x_{0},\cdots,x_{t})\in\mathcal{W}_{t}}{\mathbb{E}}\left[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})\right]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}v\right\rVert_{2}
    =supv=1P(𝔼(x0,,st)𝒲t[f(xt)f(x0)]vxt)2\displaystyle~{}=~{}\sup_{\left\lVert v\right\rVert=1}\left\lVert\symsf{P}_{\mathcal{H}}\,\left(\mathchoice{\underset{(x_{0},\cdots,s_{t})\in\mathcal{W}_{t}}{\mathbb{E}}\left[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})\right]}{{\mathbb{E}}_{(x_{0},\cdots,s_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\cdots,s_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\cdots,s_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}v\otimes x_{t}\right)\right\rVert_{2}
    =supv=1PΠf(AXΠf)t𝔼xVX[vx]2\displaystyle~{}=~{}\sup_{\left\lVert v\right\rVert=1}\left\lVert\symsf{P}_{\mathcal{H}}{\symsf{\Pi}}_{\symsf{f}}\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{t}\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[v\otimes x\right]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}\right\rVert_{2}
    =supv=1PΠf(AXΠf)tLv2\displaystyle~{}=~{}\sup_{\left\lVert v\right\rVert=1}\left\lVert\symsf{P}_{\mathcal{H}}{\symsf{\Pi}}_{\symsf{f}}\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{t}\symsf{L}_{\mathcal{H}}\,v\right\rVert_{2}
    Πf(AXΠf)topPopLop\displaystyle~{}\leq~{}\left\lVert{\symsf{\Pi}}_{\symsf{f}}\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{t}\right\rVert_{\textup{op}}\left\lVert\symsf{P}_{\mathcal{H}}\right\rVert_{\textup{op}}\left\lVert\symsf{L}_{\mathcal{H}}\right\rVert_{\textup{op}}
    Πf(AXΠf)top.\displaystyle~{}\leq~{}\left\lVert{\symsf{\Pi}}_{\symsf{f}}\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{t}\right\rVert_{\textup{op}}\,.

    The last inequality follows from submultiplicativity of the operator norm and the observation that Πfop=f1\left\lVert{\symsf{\Pi}}_{\symsf{f}}\right\rVert_{\textup{op}}=\left\lVert f\right\rVert_{\infty}\leq 1.      

Now that we have reduced the problem to studying the operator norm, we will study how the norm decays as we take walks. We use the decomposition, 𝒳=𝒳𝒳\mathcal{X}_{\mathcal{H}}=\mathcal{X}_{\mathcal{H}}^{\parallel}\oplus\mathcal{X}_{\mathcal{H}}^{\perp} where 𝒳span{v1v}\mathcal{X}_{\mathcal{H}}^{\parallel}\coloneqq\textup{span}\{v\otimes\vec{1}\mid v\in\mathcal{H}\}. The decay comes from two sources. For z𝒳z\in\mathcal{X}_{\mathcal{H}}^{\perp}, we get a decay by λ(X)\lambda(X) by the definition of XX being an expander. 3.4 shows that for z𝒳z\in\mathcal{X}_{\mathcal{H}}^{\parallel}, we get a decay from Πf{\symsf{\Pi}}_{\symsf{f}}, equal to the initial bias. We put this together in Theorem 3.1 to obtain the desired exponential decay.

Claim 3.4.

For z𝒳z\in\mathcal{X}_{\mathcal{H}}^{\parallel}, we have

(Πfz)2𝔼xVX[f(x)]opz2.\displaystyle\left\lVert\left({\symsf{\Pi}}_{\symsf{f}}\,z\right)^{\parallel}\right\rVert_{2}~{}\leq~{}\left\lVert\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right\rVert_{\textup{op}}\cdot\left\lVert z\right\rVert_{2}\,.
  • Proof:   From definition of 𝒳\mathcal{X}_{\mathcal{H}}^{\parallel}, we can assume that z=u1z=u\otimes\vec{1}. Computing we have,

    (Πf(u1))2\displaystyle\left\lVert\left({\symsf{\Pi}}_{\symsf{f}}\left(u\otimes\vec{1}\right)\right)^{\parallel}\right\rVert_{2}~{} =supw:w12=1|w1,Πf(u1)|\displaystyle=~{}\sup_{w\in\mathcal{H}\colon\left\lVert w\otimes\vec{1}\right\rVert_{2}=1}\left\lvert\left\langle w\otimes\vec{1},{\symsf{\Pi}}_{\symsf{f}}\left(u\otimes\vec{1}\right)\right\rangle\right\rvert
    =supw:w12=1|w1,Πf(uxVXx)|\displaystyle=~{}\sup_{w\in\mathcal{H}\colon\left\lVert w\otimes\vec{1}\right\rVert_{2}=1}\left\lvert\left\langle w\otimes\vec{1},{\symsf{\Pi}}_{\symsf{f}}\left(u\otimes\sum_{x\in V_{X}}x\right)\right\rangle\right\rvert
    =supw:w12=1|w1,xVX(f(x)ux)|\displaystyle=~{}\sup_{w\in\mathcal{H}\colon\left\lVert w\otimes\vec{1}\right\rVert_{2}=1}\left\lvert\left\langle w\otimes\vec{1},\sum_{x\in V_{X}}\left(\symsf{f}(x)\,u\otimes x\right)\right\rangle\right\rvert
    =supw:w12=1|xVXw,f(x)u1,x|\displaystyle=~{}\sup_{w\in\mathcal{H}\colon\left\lVert w\otimes\vec{1}\right\rVert_{2}=1}\left\lvert\sum_{x\in V_{X}}\left\langle w,\symsf{f}(x)\,u\right\rangle\left\langle\vec{1},x\right\rangle\right\rvert
    =supw:w12=1|w,|VX|(𝔼xVX[f(x)])u|\displaystyle=~{}\sup_{w\in\mathcal{H}\colon\left\lVert w\otimes\vec{1}\right\rVert_{2}=1}\left\lvert\left\langle w,\left\lvert V_{X}\right\rvert\left(\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right)u\right\rangle\right\rvert
    𝔼xVX[f(x)]op|VX|wu=𝔼xVX[f(x)]opz2.\displaystyle~{}\leq~{}\left\lVert\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right\rVert_{\textup{op}}\left\lvert V_{X}\right\rvert\left\lVert w\right\rVert\left\lVert u\right\rVert=\left\lVert\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right\rVert_{\textup{op}}\left\lVert z\right\rVert_{2}\,.

    The last line follows as z2=u2|Vx|\left\lVert z\right\rVert_{2}=\left\lVert u\right\rVert_{2}\sqrt{\left\lvert V_{x}\right\rvert} and w=1|Vx|\left\lVert w\right\rVert=\frac{1}{\sqrt{\left\lvert V_{x}\right\rvert}}.      

We show that for every two121212This is the source of loss of a factor of 22 in the exponent (which leads to degree O(|S|/λ4+o(1))O(\left\lvert S\right\rvert/\lambda^{4+o(1)}) rather than the desired degree of O(|S|/λ2+o(1))O(\left\lvert S\right\rvert/\lambda^{2+o(1)}) we will later achieve. Note that the same loss occurs in the original zig-zag analysis of [RVW00], which was later remedied by the ss-wise zig-zag of [BATS08]. steps of the walk, the norm of the (associated) operator decays as follows.

Lemma 3.5.

Let XX be a λ(X)\lambda(X)-spectral expander and let f\symsf{f} be such that 𝔼xVX[f(x)]opλ0\left\lVert\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right\rVert_{\textup{op}}\leq\lambda_{0} and maxxVXf(x)op1\max_{x\in V_{X}}\left\lVert\symsf{f}(x)\,\right\rVert_{\textup{op}}\leq 1. Then,

(AXΠf)2op1(1λ(X))2(1λ0).\displaystyle\left\lVert\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{2}\right\rVert_{\textup{op}}~{}\leq~{}1-(1-\lambda(X))^{2}(1-\lambda_{0})\,.
  • Proof:   Let AJ=J/|V(X)|\symsf{A}_{J}=\symsf{J}/\left\lvert V(X)\right\rvert, where J\symsf{J} is the |V(X)|×|V(X)|\left\lvert V(X)\right\rvert\times\left\lvert V(X)\right\rvert all ones matrix. We can write AX=(1λ)AJ+λE\symsf{A}_{X}=(1-\lambda)\symsf{A}_{J}+\lambda\symsf{E}, where λ=λ(X)\lambda=\lambda(X) and Eop1\left\lVert\symsf{E}\right\rVert_{\textup{op}}\leq 1. Then

    AXΠfAXop\displaystyle\left\lVert\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}\right\rVert_{\textup{op}}~{}\leq~{} (1λ)2AJΠfAJop+λ(1λ)EΠfAJop\displaystyle(1-\lambda)^{2}\left\lVert\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{J}{\symsf{\Pi}}_{\symsf{f}}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{J}\right\rVert_{\textup{op}}~{}+~{}\lambda(1-\lambda)\left\lVert\mathrel{\mathop{\symsf{E}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}{\symsf{\Pi}}_{\symsf{f}}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{J}\right\rVert_{\textup{op}}
    +(1λ)λAJΠfEop+λ2EΠfEop.\displaystyle+~{}(1-\lambda)\lambda\left\lVert\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{J}{\symsf{\Pi}}_{\symsf{f}}\mathrel{\mathop{\symsf{E}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}\right\rVert_{\textup{op}}~{}+~{}\lambda^{2}\left\lVert\mathrel{\mathop{\symsf{E}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}{\symsf{\Pi}}_{\symsf{f}}\mathrel{\mathop{\symsf{E}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}\right\rVert_{\textup{op}}\,.

    To analyze AJΠfAJop\left\lVert{\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{J}{\symsf{\Pi}}_{\symsf{f}}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{J}}\right\rVert_{\textup{op}} , let z𝒳z\in\mathcal{X}_{\mathcal{H}} be a unit vector which is decomposed as z=z+zz=z^{\parallel}+z^{\perp}. We have,

    (AJΠfAJ)(z+z)2\displaystyle\left\lVert\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{J}{\symsf{\Pi}}_{\symsf{f}}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{J}\right)\left(z^{\perp}+z^{\parallel}\right)\right\rVert_{2} =(AXΠfAX)z2\displaystyle~{}=~{}\left\lVert\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}\right)z^{\parallel}\right\rVert_{2} (As λ(AJ)=0)\displaystyle(\text{As }\lambda(\symsf{A}_{J})=0)
    =AX((Πfz)+(Πfz))2\displaystyle~{}=~{}\left\lVert\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}\left(\left({\symsf{\Pi}}_{\symsf{f}}z^{\parallel}\right)^{\perp}+\left({\symsf{\Pi}}_{\symsf{f}}z^{\parallel}\right)^{\parallel}\right)\right\rVert_{2}
    =(Πfz)2\displaystyle~{}=~{}\left\lVert\left({\symsf{\Pi}}_{\symsf{f}}z^{\parallel}\right)^{\parallel}\right\rVert_{2}
    λ0.\displaystyle~{}\leq~{}\lambda_{0}. (By 3.4)

    Thus, AJΠfAJopλ0\left\lVert\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{J}{\symsf{\Pi}}_{\symsf{f}}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{J}\right\rVert_{\textup{op}}\leq\lambda_{0}. Recall that Πfop1\left\lVert{\symsf{\Pi}}_{\symsf{f}}\right\rVert_{\textup{op}}\leq 1 since maxxf(x)op1\max_{x}\left\lVert\symsf{f}(x)\,\right\rVert_{\textup{op}}\leq 1, and we also have Eop,AJop1\left\lVert\symsf{E}\right\rVert_{\textup{op}},\left\lVert\symsf{A}_{J}\right\rVert_{\textup{op}}\leq 1. Then,

    AXΠfAXop\displaystyle\left\lVert\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}\right\rVert_{\textup{op}}~{}\leq~{} (1λ)2λ0+2λ(1λ)+λ2\displaystyle(1-\lambda)^{2}\lambda_{0}~{}+~{}2\lambda(1-\lambda)~{}+~{}\lambda^{2}
    =\displaystyle~{}=~{} (1λ)2λ0+1(1λ)2,\displaystyle(1-\lambda)^{2}\lambda_{0}~{}+~{}1-(1-\lambda)^{2},
    =\displaystyle~{}=~{} 1(1λ)2(1λ0),\displaystyle 1-(1-\lambda)^{2}(1-\lambda_{0})\,,

    concluding the proof.      

The amplification guarantee of Theorem 3.1 trivializes if 2λ(X)+λ012\lambda(X)+\lambda_{0}\geq 1. Nonetheless, we now show that amplification does occur under much weaker conditions, namely, whenever 𝔼xVX[f(x)]op<1\left\lVert\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right\rVert_{\textup{op}}<1 and the auxiliary graph XX has expansion λ(X)<1\lambda(X)<1. This regime of bias amplification was instrumental in the breakthrough 𝖲𝖫=𝖫\mathsf{SL}=\mathsf{L} result of Reingold [Rei05].

Theorem 3.6 (Operator Amplification via Expander Walks (strengthening of Theorem 3.1)).

Let XX be a λ(X)\lambda(X)-spectral expander and let 𝒲t\mathcal{W}_{t} be the collection of walks obtained from walks of length tt on XX. Then for any operator valued function f\symsf{f} such that 𝔼xVX[f(x)]opλ0\left\lVert\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right\rVert_{\textup{op}}\leq\lambda_{0} and maxxVXf(x)op1\max_{x\in V_{X}}\left\lVert\symsf{f}(x)\,\right\rVert_{\textup{op}}\leq 1, we have

𝔼(x0,xt)𝒲t[f(xt)f(x0)]op[1(1λ(X))2(1λ0)]t/2.\displaystyle\left\lVert\mathchoice{\underset{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}{\mathbb{E}}\left[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})\right]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\ldots x_{t})\in\mathcal{W}_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}\right\rVert_{\textup{op}}~{}\leq~{}\left[1-(1-\lambda(X))^{2}(1-\lambda_{0})\right]^{\lfloor t/2\rfloor}\,.

This establishes that expander walks can be used to derandomize powers of an operator, itself given by an average of bounded operators, in the general case. In this derandomization, we still have an exponential norm decay, but we only “pay additional randomness” proportional to the degree of the auxiliary expander regardless of the number of operators.

3.2 Cayley Graphs and the Construction of Amplified Biased Sets

In this section, we will prove the following,

See 3.2

Towards this, we first formalize the connection between bias of a special subset of a group and the operator norm of a certain operator. The subset is obtained by taking random walks over an expander graph as mentioned above. We then proceed to bound this operator norm. Finally, we instantiate our construction with an explicit expander graph due to [Alo21].

The particular case of SGS\subseteq G (for some group GG) and the function f\symsf{f} being a representation ρ\rho on \mathcal{H} leads to the amplification of biased sets. We will construct a new multiset SGS^{\prime}\subseteq G such if 𝔼sS[ρ(s)]opλ0\left\lVert\mathchoice{\underset{s\sim S}{\mathbb{E}}\left[\rho(s)\right]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}{{\mathbb{E}}_{s\sim S}[\rho(s)]}\right\rVert_{\textup{op}}\leq\lambda_{0}, then we have 𝔼sS[ρ(s)]opλλ0\left\lVert\mathchoice{\underset{s\sim S^{\prime}}{\mathbb{E}}\left[\rho(s)\right]}{{\mathbb{E}}_{s\sim S^{\prime}}[\rho(s)]}{{\mathbb{E}}_{s\sim S^{\prime}}[\rho(s)]}{{\mathbb{E}}_{s\sim S^{\prime}}[\rho(s)]}\right\rVert_{\textup{op}}\leq\lambda\ll\lambda_{0}. Note here that the construction of SS^{\prime} is agnostic to ρ\rho, and thus we can reduce the bias of all irreducible representations simultaneously! Assume that we have a graph XX on the vertex set SS. For sSs\in S, we have f(s)=ρ(s)\symsf{f}(s)=\rho(s) in this case. Let

S={stst1s0|(s0,s1,st)𝒲t},\displaystyle S^{\prime}~{}=~{}\{s_{t}s_{t-1}\cdots s_{0}\;|\;(s_{0},s_{1},\cdots s_{t})\in\mathcal{W}_{t}\}\,,

which will be our new amplified biased set. Using the homomorphism property of ρ\rho, we have the following simplification

𝔼w=(s0,st)𝒲t[f(st)f(s0)]\displaystyle\mathchoice{\underset{w=(s_{0},\ldots s_{t})\in\mathcal{W}_{t}}{\mathbb{E}}\left[\symsf{f}(s_{t})\cdots\symsf{f}(s_{0})\right]}{{\mathbb{E}}_{w=(s_{0},\ldots s_{t})\in\mathcal{W}_{t}}[\symsf{f}(s_{t})\cdots\symsf{f}(s_{0})]}{{\mathbb{E}}_{w=(s_{0},\ldots s_{t})\in\mathcal{W}_{t}}[\symsf{f}(s_{t})\cdots\symsf{f}(s_{0})]}{{\mathbb{E}}_{w=(s_{0},\ldots s_{t})\in\mathcal{W}_{t}}[\symsf{f}(s_{t})\cdots\symsf{f}(s_{0})]}~{} =𝔼(s0,,st)𝒲t[ρ(st)ρ(s0)]=𝔼sS[ρ(s)],\displaystyle=~{}\mathchoice{\underset{(s_{0},\ldots,s_{t})\in\mathcal{W}_{t}}{\mathbb{E}}\left[\rho(s_{t})\cdots\rho(s_{0})\right]}{{\mathbb{E}}_{(s_{0},\ldots,s_{t})\in\mathcal{W}_{t}}[\rho(s_{t})\cdots\rho(s_{0})]}{{\mathbb{E}}_{(s_{0},\ldots,s_{t})\in\mathcal{W}_{t}}[\rho(s_{t})\cdots\rho(s_{0})]}{{\mathbb{E}}_{(s_{0},\ldots,s_{t})\in\mathcal{W}_{t}}[\rho(s_{t})\cdots\rho(s_{0})]}~{}=~{}\mathchoice{\underset{s^{\prime}\in S^{\prime}}{\mathbb{E}}\left[\rho(s^{\prime})\right]}{{\mathbb{E}}_{s^{\prime}\in S^{\prime}}[\rho(s^{\prime})]}{{\mathbb{E}}_{s^{\prime}\in S^{\prime}}[\rho(s^{\prime})]}{{\mathbb{E}}_{s^{\prime}\in S^{\prime}}[\rho(s^{\prime})]}\,, (2)
and thus, bias(S)\displaystyle\text{and thus, }\;\mathrm{bias}(S^{\prime})~{} Πf(AXΠf)top\displaystyle\leq~{}\left\lVert{\symsf{\Pi}}_{\symsf{f}}\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{t}\right\rVert_{\textup{op}} (3)

where SS^{\prime} is the new biased multiset of the construction and the second inequality follows from the preceding calculation when 𝒲t\mathcal{W}_{t} is a collection of walks on XX.

3.2.1 Instantiating the Construction

To construct SS^{\prime}, our construction requires an auxiliary expander graph XX to perform walks on. One convenient source (among several) is a recent construction of Alon.

Theorem 3.7 (Corollary of [Alo21, Thm. 1.3] ).

For every λ(0,1)\lambda\in(0,1), there exists a positive integer mλm_{\lambda} such that for every nn\in{\mathbb{N}}, there is an explicit construction of a graph XX on mλnm_{\lambda}n vertices with degree at most 9/λ29/\lambda^{2} and λ(X)λ\lambda(X)\leq\lambda.

We now establish the key amplification lemma.

Lemma 3.8.

Let SGS\subseteq G such that 𝖻𝗂𝖺𝗌(S)=λ0<1\mathsf{bias}(S)=\lambda_{0}<1 and let ε0\varepsilon_{0} be a constant such that (1+2ε0)λ0<1(1+2\varepsilon_{0})\lambda_{0}<1. Then, for any λ>0\lambda>0, we can explicitly compute SS^{\prime} such that 𝖻𝗂𝖺𝗌(S)λ\mathsf{bias}(S^{\prime})\leq\lambda and |S|=Oε0λ0(|S|λ4+4δ(λ0,ε0))|S^{\prime}|=O_{\varepsilon_{0}\lambda_{0}}\left(\frac{|S|}{\lambda^{4+4\delta(\lambda_{0},\varepsilon_{0})}}\right) where δ(λ0,ε0)=log(3+6ε0)+log(1/ε0)log(1/λ0)\delta(\lambda_{0},\varepsilon_{0})=\frac{\log(3+6\varepsilon_{0})+\log(1/\varepsilon_{0})}{\log(1/\lambda_{0})}.

  • Proof:   Pick a constant ε0\varepsilon_{0} such that λ1(1+2ε0)λ0<1\lambda_{1}\coloneqq(1+2\varepsilon_{0})\lambda_{0}<1 and use Theorem 3.7 to obtain an explicit (m|S|,d,ε0λ0)(m|S|,d,\varepsilon_{0}\lambda_{0})-graph XX where d9ε02λ02d\leq\frac{9}{\varepsilon_{0}^{2}\lambda_{0}^{2}}. Let S1S_{1} be the multiset consisting of mm copies of SS. The bias remains the same and now, |V(X)|=|S1||V(X)|=|S_{1}|. We construct SS^{\prime} by multiplying elements of tt-length walks on XX where t=2(1+logλ1(λ))t=\lceil 2(1+\log_{\lambda_{1}}(\lambda))\rceil. The size of SS^{\prime} is

    |S|=(m|S|)dt\displaystyle|S^{\prime}|=(m|S|)\cdot d^{t}~{} =m|S|(3ε0λ0)4+4logλ1λ\displaystyle=~{}m\cdot|S|\cdot\left(\frac{3}{\varepsilon_{0}\lambda_{0}}\right)^{4+4\log_{\lambda_{1}}\lambda}
    =m(3ε0λ0)4|S|λ4log(3ε0λ0)log(1/λ1)\displaystyle=~{}m\left(\frac{3}{\varepsilon_{0}\lambda_{0}}\right)^{4}|S|\cdot\lambda^{\frac{-4\log\left(\frac{3}{\varepsilon_{0}\lambda_{0}}\right)}{\log(1/\lambda_{1})}}
    Oε0λ0(|S|)λ4(1+log(3+6ε0ε0)log(1/λ0)).\displaystyle\leq~{}O_{\varepsilon_{0}\lambda_{0}}(|S|)\cdot\lambda^{-4\left(1+\frac{\log\left(\frac{3+6\varepsilon_{0}}{\varepsilon_{0}}\right)}{\log(1/\lambda_{0})}\right)}\,.

    Let ρ\rho be any irreducible representation of GG. From  Eq. 3 and Theorem 3.1, we get,

    𝔼s0stS[ρ(sts0)]op(2λ(X)+𝖻𝗂𝖺𝗌(S))t/21(λ1)t/21λ.\displaystyle\left\lVert\mathchoice{\underset{s_{0}\cdots s_{t}\in S^{\prime}}{\mathbb{E}}\left[\rho(s_{t}\cdots s_{0})\right]}{{\mathbb{E}}_{s_{0}\cdots s_{t}\in S^{\prime}}[\rho(s_{t}\cdots s_{0})]}{{\mathbb{E}}_{s_{0}\cdots s_{t}\in S^{\prime}}[\rho(s_{t}\cdots s_{0})]}{{\mathbb{E}}_{s_{0}\cdots s_{t}\in S^{\prime}}[\rho(s_{t}\cdots s_{0})]}\right\rVert_{\textup{op}}\leq\left(2\lambda(X)+\mathsf{bias}(S)\right)^{t/2-1}\leq\left(\lambda_{1}\right)^{t/2-1}\leq\lambda\,.

     

Using the amplification above, we now derive our first simplified explicit construction. See 3.2

  • Proof:   Pick a constant λ<min(12,(34)4β)\lambda^{\prime}<\min\left(\frac{1}{2},\left(\frac{3}{4}\right)^{4\beta}\right). This choice ensures that δ(λ,1/2)β\delta(\lambda^{\prime},1/2)\leq\beta. Use Lemma 3.8 with target expansion λ\lambda^{\prime} to obtain a set S1S_{1} with size |S1|=Oλ0,β(|S|)|S_{1}|=O_{\lambda_{0},\beta}(|S|) as λ\lambda^{\prime} is a constant. Now use Lemma 3.8 again with S1S_{1} as the initial set, ε0=12\varepsilon_{0}=\frac{1}{2} and the final expansion as λ\lambda to obtain SS^{\prime}. Thus, the final size is |S|Oλ(|S1|λ4+δ(λ,1/2))Oλ0,β(|S|λ4+β)|S^{\prime}|\leq O_{\lambda^{\prime}}\left(\frac{|S_{1}|}{\lambda^{4+\delta(\lambda^{\prime},1/2)}}\right)\leq O_{\lambda_{0},\beta}\left(\frac{|S|}{\lambda^{4+\beta}}\right).      

3.3 Explicit Expanders close to any desired size

As an application of Theorem 3.2, we demonstrate an explicit construction of Cayley expanders of size n+o(n)n+o(n) vertices for every (large enough) nn.

Such a construction will be crucial for us to prove Theorem 5.4. We cannot use existing results like the recent work of Alon [Alo21] or the construction in [TS17]. This is because Alon’s construction does not have a Cayley graph structure (which our proof utilizes). On the other hand, the construction in [TS17] is a Cayley graph based on [LPS88], but it only guarantees a graph of size O(n)O(n) rather than n+o(n)n+o(n).

Recall that SL2(p)\textrm{{SL}}_{2}(p) is the group of 2×22\times 2 invertible matrices over 𝔽p{\mathbb{F}}_{p} with determinant 11. We obtain a base generating set for SL2(p)\textrm{{SL}}_{2}(p) via the following result.

Theorem 3.9 ([Lub11]).

There exists an absolute constant λ0<1\lambda_{0}<1 such that for every p>17p>17, there exists an explicit generating set SS (of constant size independent of pp) for SL2(p)\textrm{{SL}}_{2}(p), such that λ(Cay(SL2(p),S))λ0\lambda\left(Cay(\textrm{{SL}}_{2}(p),S)\right)\leq\lambda_{0} .

Theorem 3.10 ([Che10]).

For every n23215n\geq 2^{3\cdot 2^{15}}, there exists a prime in [n,n+4n2/3][n,n+4n^{2/3}].

Corollary 3.11.

For any n>29215,λ>0n>2^{9\cdot 2^{15}},\lambda>0, there is a deterministic polynomial time algorithm to construct an (n,d,λ)(n^{\prime},d,\lambda)-graph Cay(SL2(p),S)Cay(\textrm{{SL}}_{2}(p),S), where n=n+O(n8/9)n^{\prime}=n+O(n^{8/9}) and d=O(λ4.1)d=O(\lambda^{-4.1}).

  • Proof:   Find a prime p[n1/3+1,n1/3+O(n2/9)]p\in[n^{1/3}+1,n^{1/3}+O(n^{2/9})], which exists due to Theorem 3.10, via brute-force search. Since, SL2(p)\textrm{{SL}}_{2}(p) is a group of order (p21)p(p^{2}-1)p, we have n|SL2(p)|n+O(n8/9)n\leq|\textrm{{SL}}_{2}(p)|\leq n+O(n^{8/9}). We use the constant-sized generating set SS from Theorem 3.9 and amplify using Theorem 3.2.      

4 Operator Bias Reduction via the ss-wide Replacement Walk

We have seen in Section 3 that bias reduction via random walks on an expander XX is sub-optimal (by a factor of 22 in the exponent). We will derandomize this random walk construction to achieve an almost optimal bias reduction. The idea is to introduce a new graph YY which has a much smaller degree, and to “simulate" a random walk on XX via a walk on YY. This is realized by a higher-order version of the zig-zag product [RVW00] called the ss-wide replacement product defined by Ben-Aroya and Ta-Shma [BATS08] (see Definition 4.5).

This section establishes our key technical result which states that given any initial operator valued function of constant bias <1<1, we amplify the bias in an almost optimal way. This generalizes the analysis of Ta-Shma [TS17] from scalar valued functions to operator valued functions.

Theorem 4.1 (Operator Generalization of Theorem 24 [TS17]).

Fix integers ts1t\geq s\geq 1. Let XX be any d1d_{1}-regular graph (with d1d_{1} a power of 22), and YY be any d2d_{2}-regular Cayley graph on 𝔽2slogd1\mathbb{F}_{2}^{s\log d_{1}}. Let s𝒲t\mathrm{s}\mathcal{W}_{t} be the collection of length tt walks on the ss-wide replacement product of XX and YY. Let {\mathcal{H}} be a Hilbert space. For any operator valued function f:VX()\symsf{f}\colon V_{X}\rightarrow\mathcal{L}(\mathcal{H}), satisfying maxxVXf(x)op1\max_{x\in V_{X}}\left\lVert\symsf{f}(x)\,\right\rVert_{\textup{op}}\leq 1 and 𝔼xVX[f(x)]op:=λ0λ(Y)22λ(X)\left\lVert\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right\rVert_{\textup{op}}:=\lambda_{0}\leq~{}\lambda(Y)^{2}-2\lambda(X), we have

𝔼(x0,,xt)s𝒲t[f(xt)f(x0)]op(λ(Y)s+sλ(Y)s1+s2λ(Y)s3)t/sOs(λ(Y))(1os(1))t.\displaystyle\left\lVert\mathchoice{\underset{(x_{0},\cdots,x_{t})\in\mathrm{s}\mathcal{W}_{t}}{\mathbb{E}}\left[\symsf{\symsf{}}f(x_{t})\cdots\symsf{f}(x_{0})\right]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in\mathrm{s}\mathcal{W}_{t}}[\symsf{\symsf{}}f(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in\mathrm{s}\mathcal{W}_{t}}[\symsf{\symsf{}}f(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in\mathrm{s}\mathcal{W}_{t}}[\symsf{\symsf{}}f(x_{t})\cdots\symsf{f}(x_{0})]}\right\rVert_{\textup{op}}\leq\left(\lambda(Y)^{s}+s\cdot\lambda(Y)^{s-1}+s^{2}\cdot\lambda(Y)^{s-3}\right)^{\lfloor t/s\rfloor}\leq O_{s}\left(\lambda(Y)\right)^{(1-o_{s}(1))t}\,.

Furthermore, the size of the collection is |s𝒲t|=|X|d1sd2t|\mathrm{s}\mathcal{W}_{t}|=|X|\cdot d_{1}^{s}\cdot d_{2}^{t}.

Remark 4.2.

Note that there is an inherent trade-off between the spectral bound amplification (on the operator norm), and the degree bound (on the number of walks), which causes the suboptimality in how close this technique lets us approach the Ramanujan bound. As in [TS17], the oλ(1)o_{\lambda}(1) term we obtain from the bound above is (1/log(1/λ))c(1/\log(1/\lambda))^{c} for some c>0c>0 (see Theorem 4.19 for the precise computation).

Section Outline

In Section 4.1, we recall the ss-wide replacement product and describe random walks on it. Then, in Section 4.2, we formalize the distributions we work with and reprove the result that if YY is a Cayley graph over any product group of appropriate size, then it enables the transfer of pseudorandomness from YY to XX. The key generalization to operator valued functions is established in Lemma 4.7 which is identical in spirit to Eq. 1. In Section 4.3, we then finish the amplification analysis in a manner similar to [TS17]. In Section 4.4, we provide details about instantiating the setup by explicitly constructing the graphs we need.

4.1 The ss-wide Replacement Product and its Walks

Let XX be a d1d_{1}-regular graph. For each xVXx\in V_{X} and j[d1]j\in[d_{1}], let x[j]x[j] be the jj-th neighbor of xx.

Definition 4.3 (Locally Invertible Rotation Map).

XX admits a locally invertible rotation map if there exists a bijection φ:[d1][d1]\varphi\colon[d_{1}]\rightarrow[d_{1}] such that for every (x,j)VX×[d1](x,j)\in V_{X}\times[d_{1}],

if x=x[j], then, x[φ(j)]=x.\displaystyle\text{if }\;x^{\prime}=x[j],\;\text{ then, }\;x^{\prime}[\varphi(j)]=x\,.
Example 4.4 (Cayley Graphs are Locally Invertible).

Let GG be a group and AGA\subseteq G where the set AA is closed under inversion. Label the neighbors of vertices in Cay(G,A)\textup{Cay}(G,A), by elements of AA such that g[a]=agg[a]=a\cdot g. Then, Cay(G,A)\textup{Cay}(G,A) is locally invertible as the map φ:AA\varphi\colon A\rightarrow A defined as φ(a)=a1\varphi(a)=a^{-1} clearly satisfies the criteria,

if g=g[a]=ag, then, g[φ(a)]=a1g=g,\displaystyle\text{if }\;g^{\prime}=g[a]=a\cdot g,\;\text{ then, }\;g^{\prime}[\varphi(a)]=a^{-1}\cdot g^{\prime}=g\,,

for every gGg\in G, aAa\in A.

We now define the ss-wide replacement product which is a generalization of standard replacement product of graphs which can be seen as the special case when s=1s=1.

Definition 4.5 (ss-wide Replacement Product).

Suppose we are given the following:

  • -

    A d1d_{1}-regular graph XX with a bijection φ:[d1][d1]\varphi:[d_{1}]\rightarrow[d_{1}] which defines a locally invertible rotation map.

  • -

    A d2d_{2}-regular graph YY on the vertex set [d1]s[d_{1}]^{s}.

And we define:

  • -

    For i{0,1,,s1}i\in\{0,1,\dots,s-1\}, we define Roti:VX×VYVX×VY\textup{Rot}_{i}\colon V_{X}\times V_{Y}\rightarrow V_{X}\times V_{Y} such that,

    Roti((x,(a0,,as1)))(x[ai],(a0,,ai1,φ(ai),ai+1,,as1)),\displaystyle\textup{Rot}_{i}((x,(a_{0},\dots,a_{s-1})))\coloneqq(x[a_{i}],(a_{0},\dots,a_{i-1},\varphi(a_{i}),a_{i+1},\dots,a_{s-1}))\,,

    for every xVXx\in V_{X} and (a0,,as1)VY=[d1]s(a_{0},\dots,a_{s-1})\in V_{Y}=[d_{1}]^{s}. (Note that the YY component of the rotation map depends only on a vertex’s YY component, not its XX component.)

  • -

    Denote by Xi\symsf{X}_{i}, the operator on [VX×VY]{\mathbb{C}}[V_{X}\times V_{Y}] which acts on the natural basis via the permutation Roti\textup{Rot}_{i}, and let AY\symsf{A}_{Y} be the normalized random walk operator of YY.

Then tt steps of the ss-wide replacement product are given by the operator

Xt1modsAYX1modsAYX0modsAY.\displaystyle\symsf{X}_{t-1\bmod s}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y}\,\cdots\,\symsf{X}_{1\bmod s}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y}\symsf{X}_{0\bmod s}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y}\,.

Observe that a walk on the ss-wide replacement product yields a walk on the outer graph XX by recording the XX component after each step of the walk. Since a walk is completely determined by its intra-cloud steps, the number of tt-step walks on the ss-wide replacement product is,

|VX||VY|d2t=nd1sd2tnd1t,\displaystyle\left\lvert V_{X}\right\rvert\cdot\left\lvert V_{Y}\right\rvert\cdot d_{2}^{t}=n\cdot d_{1}^{s}\cdot d_{2}^{t}\ll n\cdot d_{1}^{t}\,,

which therefore gives us a very sparse subset of all tt-walks on XX. Thus the ss-wide replacement product will be used to simulate random walks on XX while requiring a reduced amount of randomness (as we shall see this simulation is only possible under special conditions, namely, when we are uniformly distributed on each cloud).

4.2 The Collection of Derandomized Walks

We now describe the distribution obtained by the walks on the ss-wide replacement product using the language of operators.

Definition 4.6 (Operators and Distributions).

Given a tuple of random walk operators131313Markov chain operators on VX×VYV_{X}\times V_{Y}. B=(B0,,Bt1)\symsf{B}=(\symsf{B}_{0},\cdots,\symsf{B}_{t-1}) on [VX][VY]{\mathbb{C}}[V_{X}]\otimes\,{\mathbb{C}}[V_{Y}] and a starting vertex x0VXx_{0}\in V_{X}, we can define a distribution induced by the walk using these operators. More precisely, 𝒟(B,x0)\mathcal{D}(\symsf{B},x_{0}) is the distribution on (VX×VY)t+1(V_{X}\times V_{Y})^{t+1} such that for every 1t1\leq\ell\leq t,

(B1B0)(x01|VY|1)=𝔼(x,y)𝒟(B,x0)xy.\left(\symsf{B}_{\ell-1}\cdots\symsf{B}_{0}\right)\left(x_{0}\otimes\frac{1}{\left\lvert V_{Y}\right\rvert}\vec{1}\right)={\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B},x_{0})}x_{\ell}\otimes y_{\ell}. (4)

We typically suppress x0x_{0} as it will not matter and denote 𝒟X(B)\mathcal{D}_{X}(\symsf{B}), and 𝒟Y(B)\mathcal{D}_{Y}(\symsf{B}) to specify the projections to VXV_{X}, and VYV_{Y} respectively.

The next lemma is a generalization of Eq. 1 which we need for the ss-wide replacement walk. This can also be specialized to prove Eq. 1 by letting YY be a graph with one vertex (and thus 𝒳𝒳𝒴\mathcal{X}_{\mathcal{H}}\cong\mathcal{XY}_{\mathcal{H}}). Recall that Πf(vxy)=f(x)vxy\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\,(v\otimes x\otimes y)=\symsf{f}(x)\,v\otimes x\otimes y.

Lemma 4.7 (Operator Generalization).

For any tuple of random walk operators B\symsf{B}, any operator valued f\symsf{f}, and any vv\in\mathcal{H}, x0VXx_{0}\in V_{X}, we have

(Bt1ΠfB0Πf)(vx01|VY|1)=𝔼(x,y)𝒟(B)[f(xt1)f(x0)vxtyt].\displaystyle\left(\mathrel{\mathop{\symsf{B}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{t-1}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\cdots\mathrel{\mathop{\symsf{B}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{0}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)\left(v\otimes x_{0}\otimes\frac{1}{\left\lvert V_{Y}\right\rvert}\vec{1}\right)=\mathchoice{\underset{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}{\mathbb{E}}\left[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}\right]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}]}\,.
  • Proof:   We prove the computation via induction on tt. The base case is when t=1t=1

    (B0Πf)(vx01|VY|1)\displaystyle\left(\mathrel{\mathop{\symsf{B}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{0}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)\left(v\otimes x_{0}\otimes\frac{1}{\left\lvert V_{Y}\right\rvert}\vec{1}\right)~{} =B0(f(x0)vx01|VY|1)\displaystyle=~{}\mathrel{\mathop{\symsf{B}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{0}\left(\symsf{f}(x_{0})\,v\otimes x_{0}\otimes\frac{1}{\left\lvert V_{Y}\right\rvert}\vec{1}\right)
    =𝔼(x,y)𝒟(B)[f(x0)vx1y1]\displaystyle=~{}\mathchoice{\underset{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}{\mathbb{E}}\left[\symsf{f}(x_{0})\,v\otimes x_{1}\otimes y_{1}\right]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{0})\,v\otimes x_{1}\otimes y_{1}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{0})\,v\otimes x_{1}\otimes y_{1}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{0})\,v\otimes x_{1}\otimes y_{1}]} (Using Eq. 4 for =1)\displaystyle(\text{Using \lx@cref{creftype~refnum}{eqn:distrib} for }\ell=1)

    Let y0=1|VY|1y_{0}=\frac{1}{\left\lvert V_{Y}\right\rvert}\vec{1} and assume the statement holds for t1t-1. Then,

    (Bt1ΠfB0Πf)(vx0y0)\displaystyle\left(\mathrel{\mathop{\symsf{B}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{t-1}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\cdots\mathrel{\mathop{\symsf{B}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{0}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)\left(v\otimes x_{0}\otimes y_{0}\right)~{} =Bt1Πfi=t20(BiΠf)(vx0y0)\displaystyle=~{}\mathrel{\mathop{\symsf{B}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{t-1}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\cdot\prod_{i=t-2}^{0}\left(\mathrel{\mathop{\symsf{B}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{i}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)\left(v\otimes x_{0}\otimes y_{0}\right)
    =Bt1Πf𝔼(x,y)𝒟(B)[f(xt2)f(x0)vxt1yt1]\displaystyle=~{}\mathrel{\mathop{\symsf{B}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{t-1}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\mathchoice{\underset{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}{\mathbb{E}}\left[\symsf{f}(x_{t-2})\cdots\symsf{f}(x_{0})\,v\otimes x_{t-1}\otimes y_{t-1}\right]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-2})\cdots\symsf{f}(x_{0})\,v\otimes x_{t-1}\otimes y_{t-1}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-2})\cdots\symsf{f}(x_{0})\,v\otimes x_{t-1}\otimes y_{t-1}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-2})\cdots\symsf{f}(x_{0})\,v\otimes x_{t-1}\otimes y_{t-1}]}
    =Bt1𝔼(x,y)𝒟(B)[f(xt1)f(xt2)f(x0)vxt1yt1]\displaystyle=~{}\mathrel{\mathop{\symsf{B}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{t-1}\mathchoice{\underset{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}{\mathbb{E}}\left[\symsf{f}(x_{t-1})\symsf{f}(x_{t-2})\cdots\symsf{f}(x_{0})\,v\otimes x_{t-1}\otimes y_{t-1}\right]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-1})\symsf{f}(x_{t-2})\cdots\symsf{f}(x_{0})\,v\otimes x_{t-1}\otimes y_{t-1}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-1})\symsf{f}(x_{t-2})\cdots\symsf{f}(x_{0})\,v\otimes x_{t-1}\otimes y_{t-1}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-1})\symsf{f}(x_{t-2})\cdots\symsf{f}(x_{0})\,v\otimes x_{t-1}\otimes y_{t-1}]}
    =𝔼(x,y)𝒟(B)[f(xt1)f(x0)vxtyt].\displaystyle=~{}\mathchoice{\underset{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}{\mathbb{E}}\left[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}\right]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim\mathcal{D}(\symsf{B})}[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}]}\,.

    The second equality uses the inductive hypothesis, and the last two equalities use Eq. 4 for =t1\ell=t-1 and =t\ell=t respectively.      

Using Definition 4.6, we further define the operators for the distributions we wish to study.

Uniform Distribution

Let us first capture, using this notation, the uniform distribution on walks on XX starting from x0Vxx_{0}\in V_{x}. We define BU\symsf{B}_{U} where for each ii, Bi=AXIY\symsf{B}_{i}=\symsf{A}_{X}\otimes I_{Y} for every ii. Then, for any ,(AXIY)=AXIY\ell,\;(\symsf{A}_{X}\otimes I_{Y})^{\ell}=\symsf{A}_{X}^{\ell}\otimes I_{Y}. Therefore, we obtain that 𝒟X(BU)\mathcal{D}_{X}(\symsf{B}_{U}) is the tt-step random walk distribution on XX i.e., xiAXix0x_{i}\sim A_{X}^{i}x_{0}.

The ss-wide Distribution

This is the distribution obtained by the ss-wide walks as described in the earlier section. For 0ab<s0\leq a\leq b<s, we define

B[a,b]=(XaAY,Xa+1AY,,XbAY).\displaystyle\symsf{B}[a,b]=\left(\symsf{X}_{a}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y},\symsf{X}_{a+1}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y},\cdots,X_{b}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y}\right)\,.

We can view this random walk as occurring in two steps. The first being picking an initial vertex y0Yy_{0}\in Y and then, picking the sequence of neighbors according to which we will perform the walk in YY. To formalize this, let AY=(1/d2)j=1d2Pj\symsf{A}_{Y}=(1/d_{2})\sum_{j=1}^{d_{2}}\symsf{P}_{j} where Pj\symsf{P}_{j} are permutation matrices and let J=(j0,,jba)[d2]ba+1J=(j_{0},\cdots,j_{b-a})\in[d_{2}]^{b-a+1}. For a fixed sequence of permutations JJ, the conditional distribution (conditioned on JJ) is defined by,

B[a,b,J]=(XaPj0,Xa+1Pj1,,XbPjba).\displaystyle\symsf{B}[a,b,J]=\left(\symsf{X}_{a}\mathrel{\mathop{\symsf{P}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{j_{0}},\,\symsf{X}_{a+1}\mathrel{\mathop{\symsf{P}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{j_{1}},\cdots,\symsf{X}_{b}\mathrel{\mathop{\symsf{P}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{j_{b-a}}\right)\,.

We would like these two distributions, i.e., the uniform distribution and the ss-wide distribution , to be the same and a graph YY is said to be compatible with respect to (X,φ)(X,\varphi), if for any fixed sequence, JJ, of a walk of length s\ell\leq s, the distribution obtained on XX via the uniform sampling of y0y_{0}, is the same as the usual \ell-length walk on XX from any fixed initial vertex, x0x_{0}. Thus, the randomness of sampling a vertex from YY is effectively transferred to a random walk on XX.

Definition 4.8 (Compatible).

A graph YY is compatible with respect to (X,φ)(X,\varphi) if for every 0ab<s0\leq a\leq b<s, J[d2]ba+1J\in[d_{2}]^{b-a+1} and x0VXx_{0}\in V_{X}, we have141414It is important to note that 𝒟Y(B[a,b,J])𝒟Y(BU)\mathcal{D}_{Y}(\symsf{B}[a,b,J])\neq\mathcal{D}_{Y}(\symsf{B}_{U}).

𝒟X(B[a,b,J],x0)=𝒟X(BU,x0)=AXba+1x0.\displaystyle\mathcal{D}_{X}(\symsf{B}[a,b,J],x_{0})=\mathcal{D}_{X}(\symsf{B}_{U},x_{0})=\symsf{A}_{X}^{b-a+1}x_{0}\,.

This compatible property is the same as the 0-pseudorandom property in [TS17]. We rename it as it is more of a structural compatibility property than a pseudorandomness one. We now prove, for the sake of completeness, that Cayley graphs are compatible with every locally invertible graph.

Lemma 4.9 ([TS17, Lemma 29]).

Let Y=Cay(Gs,T)Y=\operatorname{Cay}(G^{s},T) where |G|=d1\left\lvert G\right\rvert=d_{1}. Then, YY is compatible with respect to any X,φX,\varphi.

  • Proof:   Let x0VXx_{0}\in V_{X} be arbitrary and let y0=(r1,,rs)Gsy_{0}=(r_{1},\cdots,r_{s})\sim G^{s} be sampled uniformly. Since YY is a Cayley graph, the sequence of permutations, JJ, is equivalent to a sequence of generators (t,1,,ts)Ts(t,^{1},\cdots,t^{s})\in T^{s}, and the permutation is group multiplication, ytkyy\mapsto t^{k}\cdot y.

    For each 1is1\leq i\leq s, let tiyi1=(r1i,,rsi)t^{i}\cdot y_{i-1}=(r_{1}^{i},\cdots,r_{s}^{i}). The walk evolves as follows,

    xi\displaystyle x_{i} =Xi(xi1,tiyi1)=xi1[rii]\displaystyle=X_{i}(x_{i-1},t^{i}\cdot y_{i-1})=x_{i-1}[r_{i}^{i}]
    yi\displaystyle y_{i} =φ(tiyi1)=(r1i,,ri1i,φ(rii),ri+1i,,rsi).\displaystyle=\varphi(t^{i}\cdot y_{i-1})=\left(r_{1}^{i},\cdots,r_{i-1}^{i},\varphi(r_{i}^{i}),r_{i+1}^{i},\cdots,r_{s}^{i}\right)\,.

Compatibility requires that xiAXix0x_{i}\sim A_{X}^{i}x_{0}, which is inductively implied if for every ii, riir_{i}^{i} is uniform over GG and independent of rjjr_{j}^{j} for jij\neq i.

This clam is true initially as y0y_{0} is uniform over GsG^{s}. Assume it is true for yi1y_{i-1}. Since, ti,φt^{i},\varphi are fixed permutations, they do not affect the uniformity of the distribution of rkir^{i}_{k} for any kk. Since, φ\varphi acts only on the ithi^{th} component, the independence of rki and riir_{k}^{i}\text{ and }r_{i}^{i}, guaranteed by inductive claim, implies the independence of rki and φ(rii)r_{k}^{i}\text{ and }\varphi(r_{i}^{i}).      

For a fixed x0x_{0} and JJ, we say that y0y_{0} gives rise to a sequence z=(z1,,zt)\vec{z}=(z_{1},\cdots,z_{t}) if zi=xiz_{i}=x_{i} where xix_{i} is as defined in the above proof.

Observation 4.10.

Fix x0VXx_{0}\in V_{X} and a sequence JJ. For any z=(z1,,zk)\vec{z}=(z_{1},\cdots,z_{k}), the number of y0y_{0} that gives rise to z\vec{z}\, is d2skd_{2}^{s-k}.

  • Proof:   From the proof of Lemma 4.9, we see that enforcing zi=xiz_{i}=x_{i} for each ii starting from i=1i=1, forces exactly rir_{i} to be fixed. Thus, the remaining (rk+1,,rs)(r_{k+1},\cdots,r_{s}) are free.      

4.3 The ss-wide Operator Norm Decay

We are now ready to establish the key technical lemma in the analysis of the ss-wide replacement which is an operator-valued generalization of the scalar version of present in [TS17].

Lemma 4.11 (Simulation Lemma (generalization of Lemma 26 from [TS17])).

Let 0s1s2<s0\leq s_{1}\leq s_{2}<s. For every pair of vectors z,z𝒳z,z^{\prime}\in\mathcal{X}_{\mathcal{H}}, we have,

i=s1s2(XiAYΠf)(z1|VY|1),z1=(AXΠf)s2s1+1z,z.\displaystyle\left\langle\prod_{i=s_{1}}^{s_{2}}\left(\mathrel{\mathop{\symsf{X}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{i}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)\left(z\otimes\frac{1}{\left\lvert V_{Y}\right\rvert}\vec{1}\right),z^{\prime}\otimes\vec{1}\right\rangle~{}=~{}\left\langle\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{s_{2}-s_{1}+1}z,z^{\prime}\right\rangle.
  • Proof:   Let z=xvxxz=\sum_{x}v_{x}\otimes x and z=xwxxz^{\prime}=\sum_{x}w_{x}\otimes x. Since the expression is bilinear, it suffices to prove the equation for vx0,wxv\otimes x_{0},\,w\otimes x^{\prime} for an arbitrary pair (x0,x)(x_{0},x^{\prime}). Let t=s2s1+1t=s_{2}-s_{1}+1.

    i=s1s2(XiAYΠf)=𝔼(js1,,js2)[d2]t[i=s1s2(XiPjiΠf)]\displaystyle\prod_{i=s_{1}}^{s_{2}}\left(\mathrel{\mathop{\symsf{X}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{i}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)~{}=~{}\mathchoice{\underset{(j_{s_{1}},\cdots,j_{s_{2}})\sim[d_{2}]^{t}}{\mathbb{E}}\left[\prod_{i=s_{1}}^{s_{2}}\left(\mathrel{\mathop{\symsf{X}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{i}\mathrel{\mathop{\symsf{P}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{j_{i}}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)\right]}{{\mathbb{E}}_{(j_{s_{1}},\cdots,j_{s_{2}})\sim[d_{2}]^{t}}[\prod_{i=s_{1}}^{s_{2}}\left(\mathrel{\mathop{\symsf{X}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{i}\mathrel{\mathop{\symsf{P}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{j_{i}}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)]}{{\mathbb{E}}_{(j_{s_{1}},\cdots,j_{s_{2}})\sim[d_{2}]^{t}}[\prod_{i=s_{1}}^{s_{2}}\left(\mathrel{\mathop{\symsf{X}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{i}\mathrel{\mathop{\symsf{P}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{j_{i}}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)]}{{\mathbb{E}}_{(j_{s_{1}},\cdots,j_{s_{2}})\sim[d_{2}]^{t}}[\prod_{i=s_{1}}^{s_{2}}\left(\mathrel{\mathop{\symsf{X}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{i}\mathrel{\mathop{\symsf{P}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{j_{i}}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)]}

    Therefore, we can fix J=(js1,,js2)[d2]tJ=(j_{s_{1}},\cdots,j_{s_{2}})\in[d_{2}]^{t} and prove it for that. Recall the notation 𝒲t\mathcal{W}_{t} that denotes the set of tt-length walks on the graph XX. Applying Lemma 4.7 to B[s1,s2,J]\symsf{B}[s_{1},s_{2},J], we get,

    i=s1s2(XiPjiΠf)(vx01|VY|1)\displaystyle\prod_{i=s_{1}}^{s_{2}}\left(\mathrel{\mathop{\symsf{X}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{i}\mathrel{\mathop{\symsf{P}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{j_{i}}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)\left(v\otimes x_{0}\otimes\frac{1}{\left\lvert V_{Y}\right\rvert}\vec{1}\right) =𝔼(x,y)D(B[s1,s2,J])[f(xt1)f(x0)vxtyt]\displaystyle~{}=~{}\mathchoice{\underset{(\vec{x},\vec{y})\sim D(\symsf{B}[s_{1},s_{2},J])}{\mathbb{E}}\left[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}\right]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim D(\symsf{B}[s_{1},s_{2},J])}[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim D(\symsf{B}[s_{1},s_{2},J])}[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}]}{{\mathbb{E}}_{(\vec{x},\vec{y})\sim D(\symsf{B}[s_{1},s_{2},J])}[\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v\otimes x_{t}\otimes y_{t}]}
    =x𝒲t𝔼y0VY[f(x)vxtyt]𝕀[y0 gives rise to x],\displaystyle~{}=~{}\sum_{\vec{x}\in\mathcal{W}_{t}}\mathchoice{\underset{y_{0}\sim V_{Y}}{\mathbb{E}}\left[\symsf{f}(\vec{x})\,v\otimes x_{t}\otimes y_{t}\right]}{{\mathbb{E}}_{y_{0}\sim V_{Y}}[\symsf{f}(\vec{x})\,v\otimes x_{t}\otimes y_{t}]}{{\mathbb{E}}_{y_{0}\sim V_{Y}}[\symsf{f}(\vec{x})\,v\otimes x_{t}\otimes y_{t}]}{{\mathbb{E}}_{y_{0}\sim V_{Y}}[\symsf{f}(\vec{x})\,v\otimes x_{t}\otimes y_{t}]}\mathbb{I}[y_{0}\text{ gives rise to }\vec{x}],
    =d1std1sx𝒲tf(x)vxtyt,\displaystyle~{}=~{}\frac{d_{1}^{s-t}}{d_{1}^{s}}\sum_{\vec{x}\in\mathcal{W}_{t}}\symsf{f}(\vec{x})\,v\otimes x_{t}\otimes y_{t},

    where f(x)=f(xt1)f(x0)\symsf{f}(\vec{x})=\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0}). The last equality uses 4.10. Therefore, the conditioning on yy does not change the distribution 𝒟X\mathcal{D}_{X} and when we take inner products, we obtain,

    i=s1s2(XiPjiΠf)(vx01|VY|1),wx1\displaystyle\left\langle\prod_{i=s_{1}}^{s_{2}}\left(\mathrel{\mathop{\symsf{X}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{i}\mathrel{\mathop{\symsf{P}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{j_{i}}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)\left(v\otimes x_{0}\otimes\frac{1}{\left\lvert V_{Y}\right\rvert}\vec{1}\right),w\otimes x^{\prime}\otimes\vec{1}\right\rangle =d1std1sx𝒲txt,xf(xt1)f(x0)v,w\displaystyle~{}=~{}\frac{d_{1}^{s-t}}{d_{1}^{s}}\sum_{\vec{x}\in\mathcal{W}_{t}}{\left\langle x_{t},x^{\prime}\right\rangle\left\langle\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v,w\right\rangle}
    =𝔼xDX(B[s1,s2,J])[xt,xf(xt1)f(x0)v,w].\displaystyle~{}=~{}\mathchoice{\underset{\vec{x}\sim D_{X}(\symsf{B}[s_{1},s_{2},J])}{\mathbb{E}}\left[\left\langle x_{t},x^{\prime}\right\rangle\left\langle\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v,w\right\rangle\right]}{{\mathbb{E}}_{\vec{x}\sim D_{X}(\symsf{B}[s_{1},s_{2},J])}[\left\langle x_{t},x^{\prime}\right\rangle\left\langle\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v,w\right\rangle]}{{\mathbb{E}}_{\vec{x}\sim D_{X}(\symsf{B}[s_{1},s_{2},J])}[\left\langle x_{t},x^{\prime}\right\rangle\left\langle\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v,w\right\rangle]}{{\mathbb{E}}_{\vec{x}\sim D_{X}(\symsf{B}[s_{1},s_{2},J])}[\left\langle x_{t},x^{\prime}\right\rangle\left\langle\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v,w\right\rangle]}\,.

    We now use151515 As we only want to work with the space 𝒳\mathcal{X}_{\mathcal{H}} here, we can assume in the application of the lemma that |VY|=1\left\lvert V_{Y}\right\rvert=1. Else, one could directly apply Eq. 1 and use the observation that 𝒟X(BU)\mathcal{D}_{X}(\symsf{B}_{U}) is the same as the random walk distribution on XX. Lemma 4.7 for BU\symsf{B}_{U} and take inner product to get,

    (AXΠf)s2s1+1(vx0),wx\displaystyle\left\langle\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{s_{2}-s_{1}+1}\left(v\otimes x_{0}\right),w\otimes x^{\prime}\right\rangle =𝔼xDX(BU)[xt,xf(xt1)f(x0)v,w].\displaystyle~{}=~{}\mathchoice{\underset{\vec{x}\sim D_{X}(\symsf{B}_{U})}{\mathbb{E}}\left[\left\langle x_{t},x^{\prime}\right\rangle\left\langle\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v,w\right\rangle\right]}{{\mathbb{E}}_{\vec{x}\sim D_{X}(\symsf{B}_{U})}[\left\langle x_{t},x^{\prime}\right\rangle\left\langle\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v,w\right\rangle]}{{\mathbb{E}}_{\vec{x}\sim D_{X}(\symsf{B}_{U})}[\left\langle x_{t},x^{\prime}\right\rangle\left\langle\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v,w\right\rangle]}{{\mathbb{E}}_{\vec{x}\sim D_{X}(\symsf{B}_{U})}[\left\langle x_{t},x^{\prime}\right\rangle\left\langle\symsf{f}(x_{t-1})\cdots\symsf{f}(x_{0})\,v,w\right\rangle]}\,.

    From Lemma 4.9, we know that YY is compatible and thus, 𝒟X(B[s1,s2,J])=𝒟X(BU)\mathcal{D}_{X}(\symsf{B}[s_{1},s_{2},J])=\mathcal{D}_{X}(\symsf{B}_{U}). Thus, the right hand side (and therefore left hand sides) of these two equations above are equal.      

The s-step Decay

Just like the amplification in Section 3 was analyzed by studying the norm decay obtained in every two steps (cf.,Lemma 3.8), this amplification via the ss-wide walks will be analyzed by bounding the norm decay for steps of length ss using Lemma 4.11 similarly to [BATS08, TS17]. We will use the shorthand LiXiΠfAY\symsf{L}_{i}\coloneqq\mathrel{\mathop{\symsf{X}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{i}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y}.

The goal is to bound Ls1L0op\left\lVert\symsf{L}_{s-1}\cdots\symsf{L}_{0}\right\rVert_{\textup{op}} which controls the bias of the set obtained by ss-long, ss-wide walks (cf.,proof of Section 3.1). Equivalently, we will bound (iLi)v0,ws\left\langle(\prod_{i}\symsf{L}_{i})v_{0},w_{s}\right\rangle for any unit vectors161616Here we deviate from our notation and use v,wv,w for vectors in 𝒳𝒴\mathcal{XY}_{\mathcal{H}}. v0,ws𝒳𝒴{v}_{0},{w}_{s}\in\mathcal{XY}_{\mathcal{H}}. We will use the orthogonal decomposition,

𝒳𝒴:=𝒳[VY]=𝒳𝒴𝒳𝒴 where 𝒳𝒴span{z1z𝒳}.\displaystyle\mathcal{XY}_{\mathcal{H}}:=\mathcal{X}_{\mathcal{H}}\otimes{\mathbb{C}}[V_{Y}]=\mathcal{XY}_{\mathcal{H}}^{\parallel}\oplus\mathcal{XY}_{\mathcal{H}}^{\perp}\;\text{ where }\;\mathcal{XY}_{\mathcal{H}}^{\parallel}\coloneqq\textup{span}\{z\otimes\vec{1}\mid z\in\mathcal{X}_{\mathcal{H}}\}\,.

For i1i\geq 1, we inductively define the vectors vi,wi,ziv_{i},{w}_{i},{z}_{i} and bound their norms171717By definition viAYvi1λ(Y)vi1.\left\lVert v_{i}\right\rVert\leq\left\lVert\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y}v_{i-1}^{\perp}\right\rVert\leq\lambda(Y)\left\lVert v_{i-1}\right\rVert.The computation is similar for w and zw\text{ and }z.,

vi=Li1vi1,\displaystyle v_{i}=\symsf{L}_{i-1}v_{i-1}^{\perp},\;\;\;\;\; zsi=(XsiΠf)wsi+1,\displaystyle{z}_{s-i}=\left(\mathrel{\mathop{\symsf{X}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{s-i}\mathrel{\mathop{\symsf{\Pi}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{\symsf{f}}\right)^{*}{w}_{s-i+1},\;\;\;\; wsi=(AY)zsi\displaystyle{w}_{s-i}=\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y}\right)^{*}{z}_{s-i}^{\perp} (5)
viλ(Y)i,\displaystyle\left\lVert v_{i}\right\rVert\leq\lambda(Y)^{i},\;\;\;\;\; zsiλ(Y)i1,\displaystyle\left\lVert{z}_{s-i}\right\rVert\leq\lambda(Y)^{i-1},\;\;\;\; wsiλ(Y)i.\displaystyle\left\lVert{w}_{s-i}\right\rVert\leq\lambda(Y)^{i}\,. (6)

The following lemma follows readily from a calculation and we omit its proof.

Lemma 4.12.

For any v0,wsv_{0},{w}_{s} and 0rs20\leq r\leq s-2 we have,

Ls1L0v0\displaystyle\symsf{L}_{s-1}\cdots\symsf{L}_{0}v_{0} =vs+i=0s1Ls1Livi\displaystyle~{}=~{}v_{s}~{}+~{}\sum_{i=0}^{s-1}\symsf{L}_{s-1}\cdots\symsf{L}_{i}v_{i}^{\parallel}
Ls1ws\displaystyle\symsf{L}^{*}_{s-1}{w}_{s} =ws1+zs1\displaystyle={w}_{s-1}+{z}_{s-1}^{\parallel}
LrLs1ws\displaystyle\symsf{L}^{*}_{r}\cdots\symsf{L}^{*}_{s-1}{w}_{s} =wr+zr+i=r+1s1LrLi1zi\displaystyle~{}=~{}{w}_{r}~{}+~{}{z}_{r}^{\parallel}~{}+~{}\sum_{i=r+1}^{s-1}\symsf{L}^{*}_{r}\cdots\symsf{L}^{*}_{i-1}{z}_{i}^{\parallel}
Theorem 4.13 (Operator Generalization of Theorem 24 [TS17]).

Let XX be any d1d_{1}-regular graph and YY be a Cayley graph on 𝔽2slogd1\mathbb{F}_{2}^{s\log d_{1}}. Let WtW_{t} be the collection of tt-length ss-wide walks, on the s-wide replacement product on XX and YY. For any operator valued function f\symsf{f} on VXV_{X}, such that maxxVXf(x)op1\max_{x\in V_{X}}\left\lVert\symsf{f}(x)\,\right\rVert_{\textup{op}}\leq 1 and 𝔼xVX[f(x)]op:=λ0λ(Y)22λ(X)\left\lVert\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right\rVert_{\textup{op}}:=\lambda_{0}\leq~{}\lambda(Y)^{2}-2\lambda(X),

𝔼(x0,,xt)Wt[f(xt)f(x0)]op(λ(Y)s+sλ(Y)s1+s2λ(Y)s3)t/s.\displaystyle\left\lVert\mathchoice{\underset{(x_{0},\cdots,x_{t})\in W_{t}}{\mathbb{E}}\left[\symsf{\symsf{}}f(x_{t})\cdots\symsf{f}(x_{0})\right]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in W_{t}}[\symsf{\symsf{}}f(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in W_{t}}[\symsf{\symsf{}}f(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in W_{t}}[\symsf{\symsf{}}f(x_{t})\cdots\symsf{f}(x_{0})]}\right\rVert_{\textup{op}}~{}\leq~{}\left(\lambda(Y)^{s}+s\cdot\lambda(Y)^{s-1}+s^{2}\cdot\lambda(Y)^{s-3}\right)^{\lfloor t/s\rfloor}\,.
  • Proof:   Using Lemma 4.7, we can repeat the proof of Section 3.1 to see that,

    𝔼(x0,,xt)Wt[f(xt)f(x0)]opLtL0opLs1L0opt/s.\displaystyle\left\lVert\mathchoice{\underset{(x_{0},\cdots,x_{t})\in W_{t}}{\mathbb{E}}\left[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})\right]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in W_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in W_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}{{\mathbb{E}}_{(x_{0},\cdots,x_{t})\in W_{t}}[\symsf{f}(x_{t})\cdots\symsf{f}(x_{0})]}\,\right\rVert_{\textup{op}}\leq\left\lVert\symsf{L}_{t}\cdots\symsf{L}_{0}\right\rVert_{\textup{op}}\leq\left\lVert\symsf{L}_{s-1}\cdots\symsf{L}_{0}\right\rVert_{\textup{op}}^{\lfloor t/s\rfloor}\,.
    Ls1L0v0,ws\displaystyle\left\langle\symsf{L}_{s-1}\cdots\symsf{L}_{0}v_{0},{w}_{s}\right\rangle~{} =vs,w0+r=0s1Ls1Lrvr,ws\displaystyle=~{}\left\langle v_{s},{w}_{0}\right\rangle~{}+~{}\sum_{r=0}^{s-1}\left\langle\symsf{L}_{s-1}\cdots\symsf{L}_{r}v_{r}^{\parallel},{w}_{s}\right\rangle
    =vs,ws+r=0s1vr,LrLs1ws\displaystyle=~{}\left\langle v_{s},{w}_{s}\right\rangle~{}+~{}\sum_{r=0}^{s-1}\left\langle v_{r}^{\parallel},\symsf{L}^{*}_{r}\cdots\symsf{L}^{*}_{s-1}{w}_{s}\right\rangle
    =vs,ws+i=0s1vr,wr+zr+r=0s2i=r+1s1vr,LrLi1zi\displaystyle=~{}\left\langle v_{s},{w}_{s}\right\rangle~{}+~{}\sum_{i=0}^{s-1}\left\langle v_{r}^{\parallel},{w}_{r}+{z}_{r}^{\parallel}\right\rangle~{}+~{}\sum_{r=0}^{s-2}\sum_{i=r+1}^{s-1}\left\langle v_{r}^{\parallel},\symsf{L}^{*}_{r}\cdots\symsf{L}^{*}_{i-1}{z}_{i}^{\parallel}\right\rangle
    =vs,ws+i=0s1vr,zr+r=0s2i=r+1s1vr,LrLi1zi.\displaystyle=~{}\left\langle v_{s},{w}_{s}\right\rangle~{}+~{}\sum_{i=0}^{s-1}\left\langle v_{r}^{\parallel},{z}_{r}^{\parallel}\right\rangle~{}+~{}\sum_{r=0}^{s-2}\sum_{i=r+1}^{s-1}\left\langle v_{r}^{\parallel},\symsf{L}^{*}_{r}\cdots\symsf{L}^{*}_{i-1}{z}_{i}^{\parallel}\right\rangle\,.

    The last step uses vr,wr=AYvr,zr=0\left\langle v_{r}^{\parallel},{w}_{r}\right\rangle=\left\langle\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{Y}v_{r}^{\parallel},{z}_{r}^{\perp}\right\rangle=0. Using Eq. 6, we get vr,zrλ(Y)s1\left\langle v_{r}^{\parallel},{z}_{r}^{\parallel}\right\rangle\leq\lambda(Y)^{s-1}. To bound the last term, we finally use Lemma 4.11. Let vr=vr1v_{r}^{\parallel}=v^{\prime}_{r}\otimes\vec{1}, and zi=zi1|VY|1{z}_{i}^{\parallel}=z^{\prime}_{i}\otimes\frac{1}{\left\lvert V_{Y}\right\rvert}\vec{1}. Then,

    vr,LrLi1zi\displaystyle\left\langle v_{r}^{\parallel},\symsf{L}^{*}_{r}\cdots\symsf{L}^{*}_{i-1}{z}_{i}^{\parallel}\right\rangle =vr,(AXΠf)irzi\displaystyle~{}=~{}\left\langle v^{\prime}_{r},\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{i-r}{z}^{\prime}_{i}\right\rangle (Using Lemma 4.11)
    (AXΠf)iropzivr\displaystyle~{}\leq~{}\left\lVert\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}{\symsf{\Pi}}_{\symsf{f}}\right)^{i-r}\right\rVert_{\textup{op}}\left\lVert z^{\prime}_{i}\right\rVert\left\lVert v^{\prime}_{r}\right\rVert
    λ(Y)2ir2λ(Y)r+si1λ(Y)s3,\displaystyle~{}\leq~{}\lambda(Y)^{2\lfloor\frac{i-r}{2}\rfloor}\lambda(Y)^{r+s-i-1}\leq\lambda(Y)^{s-3},

    where the penultimate inequality uses Theorem 3.1 and plugs in the assumption that 2λ(X)+𝔼xVX[f(x)]opλ(Y)22\lambda(X)+\left\lVert\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right\rVert_{\textup{op}}\leq\lambda(Y)^{2}. Substituting this back in our expression above gives us the result.      

4.4 Instantiating the ss-wide Replacement Product

The goal of this section is to prove the following result that implies Theorem 1.2.

Theorem 4.14 (Almost Ramanujan Expanders I).

Let Cay(G,S)\operatorname{Cay}(G,S) be λ0\lambda_{0}-expander with constant λ0(0,1)\lambda_{0}\in(0,1). For every function β(λ)>0\beta(\lambda)>0, and for any λ>0\lambda>0, sufficiently small such that

32β(λ)(log(1/λ)4loglog(1/λ))1/3,\displaystyle\frac{32}{\beta(\lambda)}~{}\leq~{}\left(\frac{\log(1/\lambda)}{4\log\log(1/\lambda)}\right)^{1/3}\,,

there exists a deterministic polynomial time algorithm to construct SS^{\prime} such that Cay(G,S)\operatorname{Cay}(G,S^{\prime}) is a λ\lambda-expander with degree |S|=Oλ0(|S|/λ2+β)\left\lvert S^{\prime}\right\rvert=O_{\lambda_{0}}(\left\lvert S\right\rvert/\lambda^{2+\beta}).

Furthermore, each element in SS^{\prime} is the product of O(log(1/λ))O(\log(1/\lambda)) elements of SS.

Overview

We will explicitly construct the graphs XX and YY as needed for the ss-wide product. Once we obtain the graphs, we identify the vertices of XX, i.e., VXV_{X} with the initial generating set SS, or perhaps a slightly modified set SS^{\prime}, obtained by duplicating and adding identities. The final set is obtained by multiplying elements along each (t1)(t-1)-length walk on the ss-wide replacement product of XX and YY. The way we choose parameters and objects for it borrows heavily from Ta-Shma’s arguments in [TS17]. The analysis follows an analogous structure of [JQST20] for binary codes, which in turn builds on the original analysis of Ta-Shma [TS17]. We will also use the following result from that work,

Lemma 4.15 (Based on Lemma 6 [TS17]).

For every m+m\in\mathbb{N}^{+} and d=22k2md=2^{2k}\leq 2^{m}, there exists a fully explicit set A2mA\subseteq\mathbb{Z}_{2}^{m} such that the graph Cay(2m,A)\textup{Cay}(\mathbb{Z}_{2}^{m},A) is a (2m,d,λ=md)(2^{m},d,\lambda=\frac{m}{\sqrt{d}})-expander graph.

The construction

Given as input n:=|S|,λn:=|S|,\lambda and a slowly growing function β(λ)\beta(\lambda), we construct the graphs X,YX,Y as described below with the parameters (s,d1,d2,λ1,λ2)(s,d_{1},d_{2},\lambda_{1},\lambda_{2}) which are all functions of λ\lambda and β(λ)\beta(\lambda). These are summarized in a table below for reference181818The choice of parameters is similar but not identical to Ta-Shma’s choice.. Recall that a (n,d,λ)(n,d,\lambda)-graph has nn vertices, is dd-regular, and has the second largest singular value of its normalized adjacency matrix at most λ\lambda.

  • -

    Outer Graph — The outer graph XX will be an (n,d1,λ1)(n^{\prime},d_{1},\lambda_{1})-graph which is a Cayley graph on SL2(p)\textrm{{SL}}_{2}(p) constructed using Corollary 3.11 with (n,λ1)(n,\lambda_{1}) as input. By Example 4.4, we obtain a locally invertible graph on nnn^{\prime}\approx n. The condition on the size is satisfied as n=2|S|d25d252217n=2|S|d_{2}^{5}\geq d_{2}^{5}\geq 2^{2^{17}} by the assumption that s210s\geq 2^{10}. Moreover, the degree is cλ124.1cd24.1b28.2d25\frac{c}{\lambda_{1}^{2*4.1}}\leq\frac{cd_{2}^{4.1}}{b_{2}^{8.2}}\leq d_{2}^{5}. We increase its degree to d25d_{2}^{5} by taking multiple copies of the generating set which does not change bias191919This is wasteful, but we do it to ensure that V(Y)=d1sV(Y)=d_{1}^{s} and that d1sd_{1}^{s} is a power of 22. . Thus, we obtain a (n,d1,λ1)(n^{\prime},d_{1},\lambda_{1})-graph where n=n+O(n8/9)n^{\prime}=n+O(n^{8/9}).

  • -

    Inner Graph — The inner graph YY will be a (d1s,d2,λ2)(d_{1}^{s},d_{2},\lambda_{2})-graph which is a Cayley graph on 2m{\mathbb{Z}}_{2}^{m} and therefore by Lemma 4.9, it is compatible. For this, we use the construction of Alon et al. [AGHP92] and the analysis of Ta-Shma (Lemma 4.15).

We summarize the construction and the choice of parameters n,d1,d2,λ1,λ2n^{\prime},d_{1},d_{2},\lambda_{1},\lambda_{2} and ss, which are chosen as follows for a fixed β(λ)\beta(\lambda) here -

s is the smallest power of 2 such that 32βs(log(1/λ)4loglog(1/λ))1/3s\text{ is the smallest power of 2 such that }\frac{32}{\beta}\leq s\leq\left(\frac{\log(1/\lambda)}{4\log\log(1/\lambda)}\right)^{1/3} Every other parameter is a function of ss. Y:(n2,d2,λ2),n2=d25s,d2=s4s,λ2b2d2,b2=5slogd2Y:(n_{2},d_{2},\lambda_{2}),\quad n_{2}=d_{2}^{5s},\quad d_{2}=s^{4s},\quad\lambda_{2}\leq\frac{b_{2}}{\sqrt{d_{2}}},\quad b_{2}=5s\log d_{2} X:(n,d1,λ1),nn=O(|S|d25),d1=d25,λ1=λ2210X:(n^{\prime},d_{1},\lambda_{1}),\quad n^{\prime}\approx n=O(\left\lvert S\right\rvert d_{2}^{5}),\quad d_{1}=d_{2}^{5},\quad\lambda_{1}=\frac{\lambda_{2}^{2}}{10} t: smallest integer such that (λ2)(15α)(1α)(t1)λ,; where α=1/st:\text{ smallest integer such that }(\lambda_{2})^{(1-5\alpha)(1-\alpha)(t-1)}\leq\lambda,\,;\;\text{ where }\alpha=1/s

Note: We assume that s210s\geq 2^{10} since otherwise λ\lambda is a constant and we can use Theorem 3.2.

Now, we mention the central claim that we need from our choice of parameters.

Claim 4.16.

The selection of the parameters above implies the following bounds on tt,

  1. (i).

    t12s2t-1\geq 2s^{2}

  2. (ii).

    (d2)(t1)λ2(1+10α)(d_{2})^{(t-1)}\leq\lambda^{-2(1+10\alpha)},

  • Proof:   Proof of (i) : Using d2=s4sd_{2}=s^{4s} and the upper bound on ss, we have

    (1λ2)(15α)(1α)2s2\displaystyle\left(\frac{1}{\lambda_{2}}\right)^{(1-5\alpha)(1-\alpha)2s^{2}} (1λ2)2s2=(d2b22)s2(d2)s2=s4s3\displaystyle~{}\leq~{}\left(\frac{1}{\lambda_{2}}\right)^{2s^{2}}~{}=~{}\left(\frac{d_{2}}{b_{2}^{2}}\right)^{s^{2}}~{}\leq~{}\left(d_{2}\right)^{s^{2}}=s^{4s^{3}}
    =24s3log2(s)2log2(1/λ)=1λ.\displaystyle~{}=~{}2^{4s^{3}\log_{2}(s)}~{}\leq~{}2^{\log_{2}(1/\lambda)}~{}=~{}\frac{1}{\lambda}\,.

    Hence, (λ2)(15α)(1α)s/αλ(\lambda_{2})^{(1-5\alpha)(1-\alpha)s/\alpha}\geq\lambda and thus t1t-1 must be at least 2s22s^{2}. Also observe that,

    λ2(15α)(1α)2(t1)\displaystyle\lambda_{2}^{(1-5\alpha)(1-\alpha)^{2}(t-1)} =λ2(15α)(1α)(t2)((1α)11/(t1))\displaystyle~{}=~{}\lambda_{2}^{(1-5\alpha)(1-\alpha)(t-2)\left(\frac{(1-\alpha)}{1-1/(t-1)}\right)} (7)
    λ2(15α)(1α)(t2)\displaystyle~{}\geq~{}\lambda_{2}^{(1-5\alpha)(1-\alpha)(t-2)} (t1s=1/αt-1\geq s=1/\alpha) (8)
    λ\displaystyle~{}\geq~{}\lambda (From the choice of minimal tt) (9)

    Since b2=5slog2(d2)=20s2log2(s)s4b_{2}=5s\log_{2}(d_{2})=20s^{2}\log_{2}(s)\leq s^{4} (recall that s=1/α210s=1/\alpha\geq 2^{10}),

    d212α=d2d22α=d2s8d2b22=1λ2.\displaystyle d_{2}^{1-2\alpha}~{}=~{}\frac{d_{2}}{d_{2}^{2\alpha}}~{}=~{}\frac{d_{2}}{s^{8}}~{}\leq~{}\frac{d_{2}}{b_{2}^{2}}~{}=~{}\frac{1}{\lambda_{2}}\,.

    We obtain claim (ii) by the following computation,

    (d2)(t1)\displaystyle(d_{2})^{(t-1)} λ2(t1)12α\displaystyle~{}\leq~{}{\lambda_{2}}^{\frac{-(t-1)}{1-2\alpha}}
    λ2(12α)(15α)(1α)2\displaystyle~{}\leq~{}\lambda^{\frac{-2}{(1-2\alpha)(1-5\alpha)(1-\alpha)^{2}}} (Using Eq. 9)
    λ2(1+10α).\displaystyle~{}\leq~{}\lambda^{-2(1+10\alpha)}\,.\hfill

     

Lemma 4.17.

The number of walks of length t1t-1 on the ss-wide replacement product of XX and YY is O(|S|/λ2+β)O(\left\lvert S\right\rvert/\lambda^{2+\beta}).

  • Proof:   Since each step of the walk has d2d_{2} options, the number of walks is

    |V(X)||V(Y)|d2(t1)=nd1sd2(t1)\displaystyle\left\lvert V(X)\right\rvert\left\lvert V(Y)\right\rvert\cdot d_{2}^{(t-1)}~{}=~{}n^{\prime}\cdot d_{1}^{s}\cdot d_{2}^{(t-1)} =nd2(t1)+5s\displaystyle=n^{\prime}\cdot d_{2}^{(t-1)+5s}
    =Θ(|S|d2(t1)+5s+5)\displaystyle~{}=~{}\Theta\left(\left\lvert S\right\rvert\cdot d_{2}^{(t-1)+5s+5}\right)
    =O(|S|d2(1+5α)(t1)).\displaystyle~{}=~{}O\left(\left\lvert S\right\rvert\cdot d_{2}^{(1+5\alpha)(t-1)}\right)\,.

    which from 4.16 (ii), implies a size of

    O(|S|d2(1+5α)(t1))=O(|S|λ2(1+10α)(1+5α))=O(|S|λ2+32α)=O(|S|λ2+β).\displaystyle O\left(\left\lvert S\right\rvert\cdot d_{2}^{(1+5\alpha)(t-1)}\right)~{}=~{}O\left(\frac{\left\lvert S\right\rvert}{\lambda^{2(1+10\alpha)(1+5\alpha)}}\right)~{}=~{}O\left(\frac{\left\lvert S\right\rvert}{\lambda^{2+32\alpha}}\right)~{}=~{}O\left(\frac{\left\lvert S\right\rvert}{\lambda^{2+\beta}}\right)\,.

     

Before we prove the main result, we need the following simple observation that will be used to construct a modified (ε+o(1))(\varepsilon+o(1))-biased set starting from an ε\varepsilon-biased set, SS. This is needed because the graph obtained from Corollary 3.11 does not have size exactly |S||S| but is only guaranteed to be at most (1+o(1))|S|(1+o(1))|S|.

Lemma 4.18.

Let SS be an ε\varepsilon-biased set of a group GG. And let SS^{\prime} be obtained by adding θ|S|\theta\left\lvert S\right\rvert many identity elements. Then, SS^{\prime} is an (ε+θ)(\varepsilon+\theta)-biased set.

  • Proof:   Denote by ee the identity element of GG. Let ρ\rho be any non-trivial irreducible representation of a group GG. From the computation we have,

    𝔼sSρ(s)op\displaystyle\left\lVert{\mathbb{E}}_{s\in S^{\prime}}\rho(s)\right\rVert_{\textup{op}} =11+θ𝔼sSρ(s)+θ𝔼sSSρ(1)op\displaystyle~{}=~{}\frac{1}{1+\theta}\left\lVert{\mathbb{E}}_{s\in S}\rho(s)+\theta\cdot{\mathbb{E}}_{s\in S\setminus S^{\prime}}\rho(1)\right\rVert_{\textup{op}}
    𝔼sSρ(s)op+θ\displaystyle~{}\leq~{}\left\lVert{\mathbb{E}}_{s\in S}\rho(s)\right\rVert_{\textup{op}}+\theta (ρ(1)op=1\left\lVert\rho(1)\right\rVert_{\textup{op}}=1)
    ε+θ\displaystyle~{}\leq~{}\varepsilon+\theta (S is ε- biased).\displaystyle\text{($S\text{ is }\;\varepsilon\text{- biased}$)}\,.

     

Theorem 4.19 (Almost Ramanujan Expanders I).

Let Cay(G,S)\operatorname{Cay}(G,S) be λ0\lambda_{0}-expander with constant λ0(0,1)\lambda_{0}\in(0,1). For every function β(λ)>0\beta(\lambda)>0, and for any λ>0\lambda>0, sufficiently small such that

32β(λ)(log(1/λ)4loglog(1/λ))1/3,\displaystyle\frac{32}{\beta(\lambda)}~{}\leq~{}\left(\frac{\log(1/\lambda)}{4\log\log(1/\lambda)}\right)^{1/3}\,,

there exists a deterministic polynomial time algorithm to construct SS^{\prime} such that Cay(G,S)\operatorname{Cay}(G,S^{\prime}) is a λ\lambda-expander with degree |S|=Oλ0(|S|/λ2+β)\left\lvert S^{\prime}\right\rvert=O_{\lambda_{0}}(\left\lvert S\right\rvert/\lambda^{2+\beta}).

Furthermore, each element in SS^{\prime} is the product of O(log(1/λ))O(\log(1/\lambda)) elements of SS.

  • Proof:   We can assume that s210s\geq 2^{10} since otherwise λ\lambda is a constant and we can just use Theorem 3.2.

Initial Boost  We first boost the expansion from λ0\lambda_{0} to 1/d2λ22/31/d_{2}\leq\lambda_{2}^{2}/3. Using Theorem 3.2 (with its parameter β\beta equal to 11), we can find a new set of generators, S1S_{1}, such that Cay(G,S1)\operatorname{Cay}(G,S_{1}) is 1/d21/d_{2}-spectral expander and |S1|=O(|S|d25)\left\lvert S_{1}\right\rvert=O(\left\lvert S\right\rvert d_{2}^{5}). Moreover, we also know that each element in S1S_{1} is a multiple of at most log(d25)\log\left(d_{2}^{5}\right) elements in SS. We add multiple copies of the entire set to make the size |S|d25\left\lvert S\right\rvert d_{2}^{5}.

The ss-wide walk

Obtain an (n,d1,λ1)(n^{\prime},d_{1},\lambda_{1}) Cayley graph XX from Corollary 3.11 as explained before. We add nn=O(n8/9)n^{\prime}-n=O(n^{8/9}) copies of the identity to S1S_{1} to obtain S2S_{2}. By Lemma 4.18 and the assumption that s210s\geq 2^{10}, S2S_{2} is a λ22/3+O(n1/9)2λ22/3\lambda_{2}^{2}/3+O(n^{-1/9})\leq 2\lambda_{2}^{2}/3-biased set. We denote by SS^{\prime} the final set of generators obtained by tt steps of the ss-wide replacement product of XX and YY. By definition, each element in SS^{\prime} is a product of tt elements in S2S_{2} which has the same elements as S1S_{1}. Thus, each element in SS^{\prime} is a product of at most

O(tlog(d2))\displaystyle O(t\log(d_{2})) O((1+10α)log(1/λ))\displaystyle~{}\leq~{}O((1+10\alpha)\log(1/\lambda)) (Using 4.16 [ii])
O(log(1/λ))\displaystyle~{}\leq~{}O(\log(1/\lambda)) (By the assumption that α1/128\alpha\leq 1/128)

elements of SS. The only thing that remains is to prove expansion of Cay(G,S)\operatorname{Cay}(G,S^{\prime}). We pick any irreducible representation ρ\rho and apply Theorem 4.13 to the function ρ\rho on S2V(X)S_{2}\leftrightarrow V(X). The condition that 2λ(X)+𝔼gS2[ρ(g)]opλ(Y)22\lambda(X)+\left\lVert\mathchoice{\underset{g\sim S_{2}}{\mathbb{E}}\left[\rho(g)\right]}{{\mathbb{E}}_{g\sim S_{2}}[\rho(g)]}{{\mathbb{E}}_{g\sim S_{2}}[\rho(g)]}{{\mathbb{E}}_{g\sim S_{2}}[\rho(g)]}\right\rVert_{\textup{op}}\leq\lambda(Y)^{2} translates to λ1λ22/6\lambda_{1}\leq\lambda_{2}^{2}/6 which is satisfied by our choice of λ1\lambda_{1}. Thus, the final expansion is given by,

𝔼gS[ρ(g)]op\displaystyle\left\lVert\mathchoice{\underset{g\in S^{\prime}}{\mathbb{E}}\left[\rho(g)\right]}{{\mathbb{E}}_{g\in S^{\prime}}[\rho(g)]}{{\mathbb{E}}_{g\in S^{\prime}}[\rho(g)]}{{\mathbb{E}}_{g\in S^{\prime}}[\rho(g)]}\right\rVert_{\textup{op}} (λ2s+sλ2s1+s2λ2s3)(t1)/s\displaystyle\coloneqq\left(\lambda_{2}^{s}+s\cdot\lambda_{2}^{s-1}+s^{2}\cdot\lambda_{2}^{s-3}\right)^{\lfloor(t-1)/s\rfloor}
(3s2λ2s3)((t1)/s)1\displaystyle~{}\leq~{}\left(3s^{2}\lambda_{2}^{s-3}\right)^{((t-1)/s)-1} (Using λ220s2logss2s213s2)\displaystyle\left(\text{Using $\lambda_{2}\leq\frac{20s^{2}\log s}{s^{2s^{2}}}\leq\frac{1}{3s^{2}}$}\right)
(λ2s4)(t1s)/s\displaystyle~{}\leq~{}\left(\lambda_{2}^{s-4}\right)^{(t-1-s)/s}
λ2(15/s)(1s/(t1))(t1)\displaystyle~{}\leq~{}\lambda_{2}^{(1-5/s)(1-s/(t-1))(t-1)}
λ2(15α)(1α)(t1)\displaystyle~{}\leq~{}\lambda_{2}^{(1-5\alpha)(1-\alpha)(t-1)} (Using 4.16 [i])\displaystyle\left(\text{Using ~{}\lx@cref{creftype~refnum}{claim:t_bound} [i]}\right)
=λ2(15α)(1α)(t1)λ,\displaystyle~{}=~{}\lambda_{2}^{(1-5\alpha)(1-\alpha)(t-1)}\leq\lambda, (From the choice of t).\displaystyle\text{(From the choice of $t$)}\,.

 

5 Some Applications

Our operator amplification leads to almost optimal explicit constructions of many pseudorandom objects (from existing suboptimal ones): transforming arbitrary regular expander graphs into almost-Ramanujan expanders (Section 5.2), quantum expanders (Section 5.3), monotone expanders (Section 5.4), to generating sets with improved (average) Kazhdan constants (Section 5.5) and to dimension expanders (Section 5.6). These pseudorandom objects embody various notions of expansion.

Permutation Amplification

The key to these applications is observing that the adjacency matrix of an arbitrary graph and that of a monotone expander can be written as a sum of permutation matrices which can be interpreted as Pσ=ρdef(σ)\symsf{P}_{\sigma}=\rho_{\text{def}}(\sigma) for the defining (or natural) representation ρdef\rho_{\text{def}}. We plug in the collection of these permutations {σ}\{\sigma\} in our amplification machinery to obtain almost optimal spectral expanders and monotone expanders.

Almost Ramanujan Expanders for the Symmetric Group

Constructing constant size expanding generating sets for the symmetric group was quite challenging (even non-explicitly). In a breakthrough work [Kas07], Kassabov provided the first family of such expanding generators which was also explicit. However, this family was not close to the Ramanujan bound and no such generating set was known. Theorem 1.2 lets us amplify Kassabov’s generating set to one close to optimum bound showing that the symmetric group has explicit almost Ramanujan Cayley expanders. The same obviously holds for every expanding group.

Quantum Expanders

A quantum expander is a generalization of an expander graph having many applications in quantum information theory [AS04, BASTS08, Has07b, Has07a, HH09, AHL+14]. Harrow [Har07] showed that Cayley graphs can be used to construct quantum expanders inheriting the expansion of the starting Cayley graph. However, the construction is only explicit if the group admits an efficient quantum Fourier transform (QFT). Since we can now obtain almost Ramanujan Cayley graphs for the symmetric group which has a known efficient QFT [Bea97], we obtain the first explicit almost Ramanujan quantum expanders.

Improving the Kazhdan Constant

The Kazhdan constant 𝒦(G,S)\mathcal{K}(G,S) of a finitely generated group GG, with respect to a generating set SS, is a quantitative version of Property (T)(\mathrm{T}) which has been used to construct explicit expanders (e.g., Margulis [Mar88]). We show that this can be amplified by considering a slightly different version called the average Kazhdan constant which directly relates to the bias of the set SS. This is interesting as typically the bound on the Kazhdan constant is used to construct expanders but here we construct expanding generating sets to improve the constant! The improved constants and the generating sets have algorithmic implications and we mention two of them.

  1. \cdot

    Dimension expanders - Lubotzky and Zelmanov [LZ08] showed that the image of a generating set of a group under an irreducible representation gives a dimension expander and its expansion is controlled by its Kazhdan constant.

  2. \cdot

    Product replacement algorithm - uses random walks on kk-tuples of groups elements. Lubotzky and Pak [LP00] showed that the mixing time of the algorithm relates to the Kazhdan constant of certain lattice groups like SLn()\mathrm{SL}_{n}({\mathbb{Z}}), assuming Property (T)(\textup{T}). This crucial assumption was proven in complete generality202020[KKN21] prove that Aut(Fn)\mathrm{Aut}(\mathrm{F}_{n}), the automorphism group of the free group generated by nn elements, has Property (T)(\textup{T}). This implies Property (T)(\textup{T}) for quotients of Aut(Fn)\mathrm{Aut}(\mathrm{F}_{n}), which includes SLn()\mathrm{SL}_{n}({\mathbb{Z}}). recently by Kaluba, Kielak and Nowak [KKN21]. In particular, we have a mixing time bound of 4log|G|𝒦(G,S)2\frac{4\log{\left\lvert G\right\rvert}}{\mathcal{K}(G,S)^{2}}.

Using our amplified generating set (Corollary 1.7), we can improve both these results.

5.1 Permutation Amplification

The defining representation - (ρdef(σ),n\rho_{\text{def}}(\sigma),{\mathbb{C}}^{n}) for Symn\mathrm{Sym}_{n} is defined as the representation that maps a permutation to the matrix defining it. More formally, ρdef(σ)ei=eσ(i)\rho_{\text{def}}(\sigma)e_{i}=e_{\sigma(i)} for every unit basis vector eie_{i} of n{\mathbb{C}}^{n}. It is a fact that 𝒱def=𝒱triv𝒱standard\mathcal{V}_{\text{def}}=\mathcal{V}_{\text{triv}}\oplus\mathcal{V}_{\text{standard}} where 𝒱standard\mathcal{V}_{\text{standard}} is an irreducible non-trivial representation. Note that if we are given a set {P1,,Pr}\{\symsf{P}_{1},\cdots,\symsf{P}_{r}\} of permutation matrices acting on n{\mathbb{C}}^{n}, we can identify a set S={σ1,,σr}SymnS=\{\sigma_{1},\cdots,\sigma_{r}\}\subseteq\mathrm{Sym}_{n} such that ρdef(σi)=Pi\rho_{\text{def}}\,(\sigma_{i})=\symsf{P}_{i}.

Corollary 5.1 (Permutation Amplification).

Let P={P1,,Pr}\symsf{P}=\{\symsf{P}_{1},\cdots,\symsf{P}_{r}\} be a collection of permutation matrices such that λ(𝔼i[r][Pi])λ0\lambda(\mathchoice{\underset{i\sim[r]}{\mathbb{E}}\left[\symsf{P}_{i}\right]}{{\mathbb{E}}_{i\sim[r]}[\symsf{P}_{i}]}{{\mathbb{E}}_{i\sim[r]}[\symsf{P}_{i}]}{{\mathbb{E}}_{i\sim[r]}[\symsf{P}_{i}]})\leq\lambda_{0}. Then, for any λ(0,1)\lambda\in(0,1), we can explicitly construct a collection P\symsf{P}^{\prime} such that

  1. 1.

    λ(𝔼MP[M])λ\lambda\left(\mathchoice{\underset{\symsf{M}\sim\symsf{P}^{\prime}}{\mathbb{E}}\left[\symsf{M}\right]}{{\mathbb{E}}_{\symsf{M}\sim\symsf{P}^{\prime}}[\symsf{M}]}{{\mathbb{E}}_{\symsf{M}\sim\symsf{P}^{\prime}}[\symsf{M}]}{{\mathbb{E}}_{\symsf{M}\sim\symsf{P}^{\prime}}[\symsf{M}]}\right)\leq\lambda,

  2. 2.

    |P|O(|P|/λ2+o(1))\left\lvert\symsf{P}^{\prime}\right\rvert\leq O\left(\left\lvert\symsf{P}\right\rvert/\lambda^{2+o(1)}\right) and

  3. 3.

    each PiP\symsf{P}^{\prime}_{i}\in\symsf{P}^{\prime} is a product of at most Oλ0(log(1/λ))O_{\lambda_{0}}\left(\log(1/\lambda)\right) many matrices from P\symsf{P}.

  • Proof:   Let Pi=σi\symsf{P}_{i}=\sigma_{i}. Applying Theorem 4.13 to the set S={σi}S=\{\sigma_{i}\} we get a larger set of permutations, SS^{\prime} of the form σ=σi1σik\sigma^{\prime}=\sigma_{i_{1}}\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}\cdots\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}\sigma_{i_{k}} where k=Oλ0(log(1/λ))k=O_{\lambda_{0}}\left(\log(1/\lambda)\right). By the decomposition of the defining representation, we have that

    Spec(𝔼MP[M])\displaystyle\operatorname{Spec}\left(\mathchoice{\underset{\symsf{M}\sim\symsf{P}^{\prime}}{\mathbb{E}}\left[\symsf{M}\right]}{{\mathbb{E}}_{\symsf{M}\sim\symsf{P}^{\prime}}[\symsf{M}]}{{\mathbb{E}}_{\symsf{M}\sim\symsf{P}^{\prime}}[\symsf{M}]}{{\mathbb{E}}_{\symsf{M}\sim\symsf{P}^{\prime}}[\symsf{M}]}\right) =Spec(𝔼σS[ρdef(σ)])\displaystyle=\operatorname{Spec}\left(\mathchoice{\underset{\sigma^{\prime}\sim S^{\prime}}{\mathbb{E}}\left[\rho_{\text{def}}(\sigma^{\prime})\right]}{{\mathbb{E}}_{\sigma^{\prime}\sim S^{\prime}}[\rho_{\text{def}}(\sigma^{\prime})]}{{\mathbb{E}}_{\sigma^{\prime}\sim S^{\prime}}[\rho_{\text{def}}(\sigma^{\prime})]}{{\mathbb{E}}_{\sigma^{\prime}\sim S^{\prime}}[\rho_{\text{def}}(\sigma^{\prime})]}\right)
    ={1}Spec(𝔼σS[ρstandard(σ)]).\displaystyle=\{1\}\cup\operatorname{Spec}\left(\mathchoice{\underset{\sigma^{\prime}\sim\symsf{S}^{\prime}}{\mathbb{E}}\left[\rho_{\text{standard}}(\sigma^{\prime})\right]}{{\mathbb{E}}_{\sigma^{\prime}\sim\symsf{S}^{\prime}}[\rho_{\text{standard}}(\sigma^{\prime})]}{{\mathbb{E}}_{\sigma^{\prime}\sim\symsf{S}^{\prime}}[\rho_{\text{standard}}(\sigma^{\prime})]}{{\mathbb{E}}_{\sigma^{\prime}\sim\symsf{S}^{\prime}}[\rho_{\text{standard}}(\sigma^{\prime})]}\right)\,.

    where the 11 corresponds to the eigenvalue from the trivial representation. Since the operator amplification reduces the bias of every non-trivial irreducible representation, it also does so for 𝒱standard\mathcal{V}_{\text{standard}}.      

5.2 Arbitrary Expanders via Permutation Amplification

We can make any family of bounded degree expander graphs into an almost Ramanujan family while preserving their adjacency structure. First, we recall König’s theorem that says that the adjacency matrix of a dd-regular graph can be expressed in terms of permutation matrices.

Theorem 5.2 (König).

Let AX\symsf{A}_{X} be normalized adjacency matrix of a dd-regular nn-vertex simple graph XX. Then, there exists dd permutation matrices P1,,Pdn×n\symsf{P}_{1},\ldots,\symsf{P}_{d}\in\mathbb{R}^{n\times n} such that

AX=1dj=1dPj.\displaystyle\symsf{A}_{X}=\frac{1}{d}\sum_{j=1}^{d}\symsf{P}_{j}.

It is also efficient to find permutation matrices as above.

Claim 5.3.

The permutations in Theorem 5.2 can be found in time poly(n){\mathrm{poly}}(n).

  • Proof:   We view AXA_{X} as encoding the adjacency relation of a bipartite graph with vertex bipartition (A=V(X),B=V(X))(A=V(X),B=V(X)). This bipartite graph is dd-regular so it has at least one perfect matching MM, which can be found in poly(n){\mathrm{poly}}(n) time. We remove this matching MM obtaining a (d1)(d-1)-regular graph and we repeat till the resulting graph is empty.      

Our general transformation into an almost Ramanujan bound follows by using 5.3 to obtain an initial set of permutation matrices which are amplified using Corollary 5.1.

Theorem 5.4 (Main I (Formal version of Theorem 1.1)).

Let {Xi}i\{X_{i}\}_{i\in\mathbb{N}} be a family of d0d_{0}-regular λ0\lambda_{0}-expanders with constant λ0<1\lambda_{0}<1. For any λ(0,1)\lambda\in(0,1) and any expander XiX_{i}, we can deterministically compute a dd-regular λ\lambda-expander XiX_{i}^{\prime} with d=Oλ0(d0/λ2+o(1))d=O_{\lambda_{0}}(d_{0}/\lambda^{2+o(1)}) in time poly(|V(Xi)|){\mathrm{poly}}(\left\lvert V(X_{i})\right\rvert). Moreover, the construction is local in the sense that edges in XiX_{i}^{\prime} correspond to short walks in XiX_{i}. More precisely, if the adjacency matrix of XiX_{i} is

AXi=1d0j=1d0Pj,\symsf{A}_{X_{i}}=\frac{1}{d_{0}}\sum_{j=1}^{d_{0}}\symsf{P}_{j},

where P1,,Pd0\symsf{P}_{1},\ldots,\symsf{P}_{d_{0}} are permutation matrices, then the adjacency matrix of XiX_{i}^{\prime} is

AXi=1dj=1dPj,\symsf{A}_{X_{i}^{\prime}}=\frac{1}{d}\sum_{j=1}^{d}\symsf{P}_{j}^{\prime},

where each Pj\symsf{P}_{j}^{\prime} is the product of at most k=Oλ0(log(1/λ))k=O_{\lambda_{0}}(\log(1/\lambda)) permutation matrices among P1,,Pd0\symsf{P}_{1},\ldots,\symsf{P}_{d_{0}}.

5.3 Explicit Almost Ramanujan Quantum Expanders

Quantum expanders were defined in [AS04, BASTS08, Has07a] and have found many applications in quantum information theory. For instance, they can be used in the construction of designs and gates sets [HH09], in quantum statistical zero knowledge (QSZK) [BASTS08], in detecting EPR pairs [AHL+14] and in the study of entanglement [Has07b].

Roughly speaking, a quantum expander is a generalization of an expander graph (see Definition 5.5 for precise details). While a usual degree-dd expander graph X=(V,E)X=(V,E) is given by dd permutation matrices acting on a vector space [V]{\mathbb{C}}[V], a quantum expander is given by dd (suitable) linear operators acting on quantum states (i.e., PSD matrices of trace 11). The normalized adjacency matrix of a λ\lambda-expander shrinks the 2\ell_{2}-norm of vectors orthogonal the all ones function by a factor of λ\lambda. Similarly, a quantum expander shrinks the Frobenius norm of PSD matrices orthogonal 212121With respect to the Hilbert–Schmidt inner product. to the identity matrix (the quantum analogue of the all ones function) by a factor of λ\lambda (the quantum expansion parameter).

Definition 5.5 (Quantum Expander [AHL+14]).

The (super) operator Φ:N×NN×N\Phi:{\mathbb{C}}^{N\times N}\rightarrow{\mathbb{C}}^{N\times N} is an (N,d,λ)(N,d,\lambda) quantum expander if

  1. \cdot

    “Degree” – The operator Φ\Phi can be expressed as a sum of dd linear operators as follows, Φ(ρ)=i=1dBiρBi\Phi(\rho)=\sum_{i=1}^{d}B_{i}\rho B_{i}^{\dagger} where222222A useful special case is when each BiB_{i} is a (normalized) unitary. i=1dBiBi=IN\sum_{i=1}^{d}B_{i}^{\dagger}B_{i}=\symsf{I}_{N}.

  2. \cdot

    “Expansion” – The second largest eigenvalue232323If ρ\rho satisfies Tr(ρ)=0\operatorname{Tr}(\rho)=0, then Φ(ρ)2λρ2\left\lVert\Phi(\rho)\right\rVert_{2}\leq\lambda\left\lVert\rho\right\rVert_{2}, where ρ2Tr(ρρ)\left\lVert\rho\right\rVert_{2}\coloneqq\sqrt{\operatorname{Tr}(\rho^{\dagger}\rho)}. of Φ\Phi as a linear map is λ\leq\lambda.

In [Has07c], Hastings showed that the Ramanujan bound also applies to quantum expanders and that dd random unitaries get arbitrarily close to the bound. However, such a construction cannot be efficiently implemented and thus used in applications like [AHL+14] which rely on existing explicit constructions (e.g., [BASTS08, Har07]) that are far from the Ramanujan bound and thus give sub-optimal results.

Harrow [Har07] proved that one can construct a quantum expander using an expander Cayley graph over a group for which efficient Quantum Fourier Transform (QFT) is known [Bea97].

Theorem 5.6 (Harrow [Har07]).

Let GG be a group and SGS\subseteq G be a multiset such that Cay(G,S)\operatorname{Cay}(G,S) is a λ\lambda-spectral expander. Let VμV^{\mu} be an irreducible representation of GG of dimension NN. Then, there exists an (N,|S|,λ)(N,\left\lvert S\right\rvert,\lambda)-quantum expander. Furthermore, if the group GG admits an efficient QFT and logN=Ω(log|G|)\log N=\Omega(\log\left\lvert G\right\rvert), then the quantum expander is explicit.

Until now, we did not have almost-Ramanujan expanders over such a group. Since the symmetric group admits such a QFT algorithm, we deduce the existence of explicit families of almost Ramanujan quantum expanders by applying our amplification to the Cayley graphs over the symmetric group due to Kassabov [Kas07].

See 1.5

5.4 Explicit Almost Ramanujan Monotone Expander

We now show how to obtain almost Ramanujan monotone expanders starting from the explicit construction in Bourgain and Yehudayoff [BY13]. First, we recall the definition of a monotone graph. All graphs we consider are undirected.

Definition 5.7 (Monotone partial map).

A partial map f:[n][n]f:[n]\rightarrow[n] is monotone if for every pair {x,y}\{x,y\} for which ff is defined, if x<yx<y, we have f(x)<f(y)f(x)<f(y).

Definition 5.8 (Monotone Graph).

A bipartite graph X=([n]A[n]B,E)X=([n]_{A}\sqcup[n]_{B},E) is a dd-monotone graph if there are dd monotone partial maps f1,,fd:[n][n]f_{1},\ldots,f_{d}:[n]\rightarrow[n], such that the edges set EE is the following disjoint union,

E=i=1d{(vA,fi(v)B)vDomain(fi)}.E=\bigsqcup_{i=1}^{d}\{(v_{A},f_{i}(v)_{B})\mid v\in\textup{Domain}(f_{i})\}.

We observe that there are two notions of degree of a monotone graph: the usual vertex degree and the number of monotone functions. Clearly, if a graph is dd-monotone, all vertex degrees are at most dd. The converse is not necessarily true; for example, the complete bipartite graph on 22 vertices on each side, K2,2K_{2,2}, has vertex degree 22, but the graph is not 22-monotone. We stress that our almost Ramanujan bound is with respect to the usual notion of vertex degree (and keeps the number of monotone maps polynomial in the vertex degree).

Definition 5.9 (Monotone Vertex Expander).

We say that X=(A=[n]AB=[n]B,E)X=(A=[n]_{A}\sqcup B=[n]_{B},E) is a dd-monotone expander if it is a dd-monotone graph and there exists δ>0\delta>0 such that for all AAA^{\prime}\subseteq A with |A|n/2\left\lvert A\right\rvert\leq n/2, we have |(A)|(1+δ)|A|\left\lvert\partial(A^{\prime})\right\rvert\geq(1+\delta)\left\lvert A^{\prime}\right\rvert, where (A)\partial(A^{\prime}) is the set of vertices in BB adjacent to AA^{\prime}.

Theorem 5.10 (Bourgain and Yehudayoff [BY13]).

There is an explicit family {Xn}n\{X_{n}\}_{n\in\mathbb{N}} of dd-monotone vertex expanders with d=Θ(1)d=\Theta(1).

We will work with a spectral definition of monotone expander. For a bipartite graph XX, we define its biadjacency matrix, BXB_{X} such that the adjacency matrix AX=(0BXBXT0)A_{X}=\left(\begin{matrix}0&B_{X}\\ B_{X}^{T}&0\end{matrix}\right). Precisely, for a monotone graph X=([n]A[n]B,E)X=([n]_{A}\sqcup[n]_{B},E), we have (BX)ij=𝕀[(iA,jB)E](B_{X})_{ij}=\mathbb{I}[(i_{A},j_{B})\in E]. Note that if XX is dd-regular, then BXB_{X} is dd-regular. We will define the graph XX via BXB_{X} throughout.

Definition 5.11 (Spectral Monotone Expander).

We say that a dd-monotone graph, XX, is a λ\lambda-spectral monotone expander if λ(X)=max{|λ2(BX)|,|λn(BX)|}<λ\lambda(X)=\max\{\left\lvert\lambda_{2}(\symsf{B}_{X})\right\rvert,\left\lvert\lambda_{n}(\symsf{B}_{X})\right\rvert\}<\lambda.

It is well-known that starting from a monotone expander (not necessarily a vertex regular graph), we can add monotone partial functions to obtain a monotone graph of regular (vertex) degree that is still expanding. We use this to establish the following,

Corollary 5.12.

There is explicit family {Xn}n\{X_{n}\}_{n\in\mathbb{N}} of d0d_{0}-regular 2d02d_{0}-monotone expanders with λ(Xn)λ0<1\lambda(X_{n})\leq\lambda_{0}<1 and d0=Θ(1)d_{0}=\Theta(1). Furthermore, the unnormalized adjacency matrix of XnX_{n} can be written as a sum of d0d_{0} permutation matrices each corresponding to two monotone maps.

  • Proof Sketch:   Let {Xn}n\{X_{n}^{\prime}\}_{n\in\mathbb{N}} be the family in Theorem 5.10. Let X=XnX=X_{n}^{\prime} be a fixed d0d_{0}-regfular graph that is also d0d_{0}-monotone expander with the maps {fi}\{f_{i}\}.

    For each monotone function fif_{i}, we define its “complement”, f¯i\overline{f}_{i}, as the (unique) monotone partial function f¯i\overline{f}_{i} such that fif¯if_{i}\cup\overline{f}_{i} is a total function. Let YY be the 2d02d_{0}-monotone graph corresponding to the maps {fi,fi¯}\{f_{i},\overline{f_{i}}\}. Then, we have

    BY\displaystyle\symsf{B}_{Y} =i=1d0Pi,\displaystyle=\sum_{i=1}^{d_{0}}\symsf{P}_{i}\,,

    where Pi=Mfi+Mfi¯\symsf{P}_{i}=\symsf{M}_{f_{i}}+\symsf{M}_{\overline{f_{i}}} and (Mfi)x,y=𝟙[fi(x)=y](\symsf{M}_{f_{i}})_{x,y}=\mathbb{1}\left[f_{i}(x)=y\right].

    Each matrix Pi\symsf{P}_{i} is a permutation matrix as fif¯if_{i}\cup\overline{f}_{i} is a total function. Adding more maps preserves the constant vertex expansion parameter which (together with having constant vertex degree) implies constant spectral expansion bounded away from 11 (see [Vad12, Theorem 4.9]). Thus, {Yn}n\{Y_{n}\}_{n\in{\mathbb{N}}} is the required family.     \Box

In the amplification process, we will be multiplying permutation matrices rather than just composing monotone maps since the latter operation can result in a map with empty domain. We now establish the derandomized spectral amplification of monotone expanders.

See 1.6

  • Proof:   Let {Xn}n\{X_{n}^{\prime}\}_{n\in\mathbb{N}} be the family in Corollary 5.12. Fix X=XnX=X_{n}^{\prime} and let P1,,Pd0n×n\symsf{P}_{1},\ldots,\symsf{P}_{d_{0}}\in\mathbb{R}^{n\times n} be the permutation matrices guaranteed by Corollary 5.12, where each Pi=Mfi+Mf¯i\symsf{P}_{i}=\symsf{M}_{f_{i}}+\symsf{M}_{\overline{f}_{i}}. Use Corollary 5.1 to obtain a collection PP^{\prime} of size |P|=dO(1/λ2+β)|P^{\prime}|=d\coloneqq O(1/\lambda^{2+\beta}) such that,

    P={σσ=Pi1Pik for some ii,,ik[d0]}.P^{\prime}=\big{\{}\sigma\mid\sigma=P_{i_{1}}\cdots P_{i_{k}}\text{ for some }i_{i},\cdots,i_{k}\in[d_{0}]\big{\}}.

    Our final bipartite monotone graph will be YY given by BY=σPσB_{Y}=\sum_{\sigma\in P^{\prime}}\sigma. To compute its monotone degree, we observe that,

    Pi1Pik=\displaystyle\symsf{P}_{i_{1}}\cdots\symsf{P}_{i_{k}}~{}= gi{fi,fi¯}Mgi1Mgik\displaystyle~{}\sum_{g_{i}\in\{f_{i},\overline{f_{i}}\}}\symsf{M}_{g_{i_{1}}}\cdots\symsf{M}_{g_{i_{k}}}
    =\displaystyle~{}= gi{fi,fi¯}Mgi1gi2gik,\displaystyle~{}\sum_{g_{i}\in\{f_{i},\overline{f_{i}}\}}\symsf{M}_{g_{i_{1}}\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}g_{i_{2}}\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}\,\cdots\;\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}g_{i_{k}}}\,,

    where gi1gi2gikg_{i_{1}}\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}g_{i_{2}}\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}\cdots g_{i_{k}} is the composed map which is monotone (possibly with an empty domain). This means that we can have at most 2k2^{k} monotone maps (and at least one since Pi1Pik0\symsf{P}_{i_{1}}\cdots\symsf{P}_{i_{k}}\neq 0). Therefore, the total number of maps is at most d2k=dO(1)d\cdot 2^{k}=d^{O(1)} as k=Oλ0(log(1/λ))k=O_{\lambda_{0}}(\log(1/\lambda)).      

5.5 Amplifying the Average Kazhdan Constant

The Kazhdan constant is a notion of “spectral gap” (and so it is related to bias) for discrete groups which predates, and was central to the study of expansion in finite groups and graphs. In particular, we can work with finitely generated groups that can have infinitely many irreducible representations on more general Hilbert spaces, possibly of infinite dimension. Nonetheless, we can still apply our operator version of Ta-Shma’s amplification procedure as it is independent of dimension and works for any unitary representation ρ\rho. Therefore, we amplify the average Kazhdan constant which also amplifies the Kazhdan constant. We now define these two parameters formally.

Let GG be a group generated by a finite set SS of generators. The Kazhdan constant of GG with respect to generators SS is defined as

𝒦(G,S)\displaystyle\mathcal{K}(G,S) inf{𝒦(G,S,ρ)(ρ,) irreducible and non-trivial unitary representation},\displaystyle\coloneqq\inf\{\mathcal{K}(G,S,\rho)\mid(\rho,\mathcal{H})\textup{ irreducible and non-trivial {unitary representation}}\}\,,

where 𝒦(G,S,ρ)infv:v2=1maxgSρ(g)vv22\mathcal{K}(G,S,\rho)\coloneqq\inf_{v\in\mathcal{H}\colon\left\lVert v\right\rVert_{2}=1}\max_{g\in S}\left\lVert\rho(g)\,v-v\right\rVert_{2}^{2}.

Analogously, an average version of the Kazhdan constant, as in the work of Pak and Zuk [PZ01], can be defined as

𝒦¯(G,S)\displaystyle\overline{\mathcal{K}}(G,S) inf{𝒦¯(G,S,ρ)(ρ,) irreducible and non-trivial unitary representation}\displaystyle\coloneqq\inf\{\overline{\mathcal{K}}(G,S,\rho)\mid(\rho,\mathcal{H})\textup{ irreducible and non-trivial {unitary representation}}\}
𝒦¯(G,S,ρ)\displaystyle\overline{\mathcal{K}}(G,S,\rho) infv:v2=11|S|gSρ(g)vv22\displaystyle\coloneqq\inf_{v\in\mathcal{H}\colon\left\lVert v\right\rVert_{2}=1}\frac{1}{\left\lvert S\right\rvert}\sum_{g\in S}\left\lVert\rho(g)\,v-v\right\rVert_{2}^{2}
=infv:v2=11|S|gS22ρ(g)v,v\displaystyle=\inf_{v\in\mathcal{H}\colon\left\lVert v\right\rVert_{2}=1}\frac{1}{\left\lvert S\right\rvert}\sum_{g\in S}2-2\left\langle\rho(g)\,v,v\right\rangle
=infv:v2=122𝔼gS[ρ(g)]v,v\displaystyle=\inf_{v\in\mathcal{H}\colon\left\lVert v\right\rVert_{2}=1}2-2\;\Big{\langle}{\mathchoice{\underset{g\sim S}{\mathbb{E}}\left[\rho(g)\right]}{{\mathbb{E}}_{g\sim S}[\rho(g)]}{{\mathbb{E}}_{g\sim S}[\rho(g)]}{{\mathbb{E}}_{g\sim S}[\rho(g)]}v},\,{v}\Big{\rangle}
=2(1𝔼gS[ρ(g)]op).\displaystyle=2\left(1-\Big{\|}{\mathchoice{\underset{g\sim S}{\mathbb{E}}\left[\rho(g)\right]}{{\mathbb{E}}_{g\sim S}[\rho(g)]}{{\mathbb{E}}_{g\sim S}[\rho(g)]}{{\mathbb{E}}_{g\sim S}[\rho(g)]}}\Big{\|}_{\mathrm{op}}\right)\,.

Theorem 1.2 gives an improved generating set in this more general setting.

See 1.7

Remark 5.13.

Note that the above amplification for 𝒦¯\overline{\mathcal{K}} immediately implies the same amplification for 𝒦\mathcal{K} (since the maximum is at least the average, K¯(G,S)K(G,S)\overline{K}(G,S)\leq K(G,S)) . Moreover, we remark that the above amplification can also similarly improve the constant of Lubotzky’s property (τ)(\tau) (the latter being a weaker version of property (T)(\textup{T})), so it is more general and applies to expansion in many more discrete groups [RL10].

In Section 5.6, we will apply this corollary to a specific family of representations which will give a simple improvement to the bounds on the dimension expander constructed in [LZ08].

5.6 Explicit Almost Ramanujan Dimension Expanders

Dimension expanders were defined in [BISW01] motivated by applications in theoretical computer science. A conjectured construction based on irreducible representations was suggested by Wigderson to hold over every field. The conjecture was subsequently established by Lubotzky and Zelmanov [LZ08] for fields of characteristic zero. We now define dimension expanders, explain the [LZ08] proof, and our amplification in this setting.

Definition 5.14 ((ε,γ)(\varepsilon,\gamma) Dimension Expander).

Let 𝔽{\mathbb{F}} be a field, dd\in\mathbb{N}, ε>0\varepsilon>0, V{V} be a vector space of dimension nn and T1,,Td:VVT_{1},\ldots,T_{d}\colon V\rightarrow V be linear transformations. We say that (V,{Ti}i[d])({V},\{T_{i}\}_{i\in[d]}) is an (ε,γ)(\varepsilon,\gamma)-dimension expander if for every subspace WV{W}\subseteq{V} of dimension at most γn\gamma n, we have dim(W+i=1dTi(W))(1+ε)dim(W)\dim({W}+\sum_{i=1}^{d}T_{i}({W}))\geq(1+\varepsilon)\cdot\dim({W}).

Remark 5.15.

Observe that if the maps TiT_{i} are restricted to being permutation matrices, and the expansion condition is restricted only to subspaces WW generated by elementary basis vectors, then one obtains the usual definition of vertex expansion of graphs. Thus dimension expanders may be viewed as a linear-algebraic extension of expander graphs.

For an irreducible unitary representation ρ\rho, there exists an associated representation242424Let 𝔰𝔩n()={tr(A)=0|AMn()}\mathfrak{sl}_{n}({\mathbb{C}})=\{\mathrm{tr}(\symsf{A})=0\;|\;\symsf{A}\in\symsf{M}_{n}({\mathbb{C}})\}. Equip the space with the Frobenius inner product defined as A,B=tr(AB)\left\langle\symsf{A},\symsf{B}\right\rangle=\mathrm{tr}(\symsf{A}^{\dagger}\symsf{B}) where A\symsf{A}^{\dagger} is the conjugate transpose. For any finite-dimensional unitary representation ρ:G𝕌n\rho:G\rightarrow\mathbb{U}_{n}, we have an adjoint representation (adjρ,𝔰𝔩n)(\mathrm{adj}_{\rho},\mathfrak{sl}_{n}) where the action is by conjugation adjρ(g)A=ρ(g)Aρ(g)1\mathrm{adj}_{\rho}(g)\cdot\symsf{A}=\rho(g)\cdot\symsf{A}\cdot\rho(g)^{-1}. Since conjugation by unitary matrices preserves the trace, 𝔰𝔩n\mathfrak{sl}_{n} is closed under the representation. Moreover, it is unitary as adjρ(g)A,adjρ(g)B=tr(ρ(g)Aρ(g)ρ(g)Bρ(g)1)=A,B.\left\langle\mathrm{adj}_{\rho}(g)\symsf{A},\mathrm{adj}_{\rho}(g)\symsf{B}\right\rangle=\mathrm{tr}\left(\rho(g)\symsf{A}^{\dagger}\rho(g)^{\dagger}\rho(g)\symsf{B}\rho(g)^{-1}\right)=\left\langle\symsf{A},\symsf{B}\right\rangle. adjρ\mathrm{adj}_{\rho}. The construction in [LZ08] relates dimension expansion with the Kazhdan constant. Their result gives a dimension expander, which gives expansion for all subspaces WW, such that dim(W)n/2\dim(W)\leq n/2, but their expansion guarantee is significantly stronger when dim(W)\dim(W) is smaller. To obtain this, we first state a simple improvement to a computation in [LZ08].

Claim 5.16.

Let W,Wd{W},{W}^{\prime}\subseteq\mathbb{C}^{d} be two vector spaces. Let P,P\symsf{P},\symsf{P}^{\prime} be orthogonal projectors onto W,W{W},{W}^{\prime}, respectively. Then,

Re(Tr(PP))=Tr(PP)dim(WW).\mathrm{Re}\left(\operatorname{Tr}(\symsf{P}\symsf{P}^{\prime})\right)=\operatorname{Tr}(\symsf{P}\symsf{P}^{\prime})\geq\dim({W}\cap{W}^{\prime}).
  • Proof:   Let 𝒰0=𝒲𝒲\mathcal{U}_{0}=\mathcal{W}\cap\mathcal{W}^{\prime}, 𝒰1=𝒲𝒰0\mathcal{U}_{1}=\mathcal{W}\cap\mathcal{U}_{0}^{\perp} and 𝒰2=𝒲𝒰0\mathcal{U}_{2}=\mathcal{W}^{\prime}\cap\mathcal{U}_{0}^{\perp}, with orthonormal bases {u1,,uk}\{u_{1},\ldots,u_{k}\}, {a1,,a}\{a_{1},\ldots,{a_{\ell}}\} and {b1,,bm}\{{b_{1}},\ldots,{b_{m}}\}, respectively. We can write PP and PP^{\prime} as

    P=i=1kuiuiT+i=1aiaiT and P=i=1kuiuiT+i=1mbibiT.\displaystyle P=\sum_{i=1}^{k}u_{i}u_{i}^{T}+\sum_{i=1}^{\ell}a_{i}a_{i}^{T}\textup{ and }P^{\prime}=\sum_{i=1}^{k}u_{i}u_{i}^{T}+\sum_{i=1}^{m}b_{i}b_{i}^{T}.

    Using linearity and orthogonality, we obtain

    Tr(PP)\displaystyle\operatorname{Tr}(PP^{\prime}) =i=1kui2Tr(uiuiT)+i=1j=1mai,bjTr(aibjT)\displaystyle=\sum_{i=1}^{k}\left\lVert u_{i}\right\rVert^{2}\operatorname{Tr}(u_{i}u_{i}^{T})+\sum_{i=1}^{\ell}\sum_{j=1}^{m}\left\langle a_{i},b_{j}\right\rangle\operatorname{Tr}(a_{i}b_{j}^{T})
    =k+i=1j=1mai,bj2dim(𝒰0),\displaystyle=k+\sum_{i=1}^{\ell}\sum_{j=1}^{m}\left\langle a_{i},b_{j}\right\rangle^{2}\geq\dim(\mathcal{U}_{0}),

    here in the last step we used that ui2=Tr(uiuiT)=1\left\lVert u_{i}\right\rVert^{2}=\operatorname{Tr}(u_{i}u_{i}^{T})=1.      

The above claim is a variant of the one used in  [LZ08] to prove their main result. By plugging in 5.16 in their proof we obtain,

Proposition 5.17 (Adapted from [LZ08] using 5.16).

Let ρ:G𝕌n\rho\colon G\rightarrow\mathbb{U}_{\mathbb{C}^{n}} be a unitary irreducible representation. Then (n,{ρ(g)}gS)(\mathbb{C}^{n},\{\rho(g)\}_{g\in S}) is (1λon(1),1/2O(λ))\left(1-\lambda-o_{n}(1),1/2-O(\lambda)\right)-dimension expander, where 2(1λ):=𝒦(G,S,adjρ)2(1-\lambda):=\mathcal{K}(G,S,\text{adj}_{\rho}).

Corollary 5.18.

Let λ>0\lambda>0 be any fixed constant. Then, there exists an explicit infinite family of (1λon(1),12O(λ))\left(1-\lambda-o_{n}(1),\tfrac{1}{2}-O(\lambda)\right)-dimension expanders.

  • Proof:   Pick a family of groups {Gn}n\{G_{n}\}_{n} such that each GnG_{n} satisfies the condition of Corollary 1.7; for example, one can take any non-abelian finite simple group. By definition, for any such GG, we have 𝒦(G,S,adjρ)𝒦(G,S)\mathcal{K}(G,S,\mathrm{adj}_{\rho})\geq\mathcal{K}(G,S) and therefore we obtain a set SS^{\prime} such that 𝒦(G,S,adjρ)2(1λ)\mathcal{K}(G,S^{\prime},\mathrm{adj}_{\rho})\geq 2(1-\lambda) for the given λ\lambda. We can now apply Proposition 5.17.      

Remark 5.19.

Forbes and Guruswami [FG15] point out that the quantum expander construction of Harrow [Har07] also yields a dimension expander (with a similar construction of the dimension expanders from [LZ08]). As mentioned earlier, monotone expanders are dimension expanders over any field [DS09, DW10]. Moreover, the Bourgain and Yehudayoff [BY13] construction of monotone expanders with constant generating set yields such dimension expanders with constant generating set!

5.7 Diameter of Finite Groups

The study of the diameter of Cayley graphs can take many forms, e.g., it can be with respect to every generating set (as in the celebrated Babai–Seress conjecture [BS88]) or with respect to some constant size generating set as in [BKL89]. Here, we explore the latter case.

First, recall that any nn-vertex degree-dd graph has diameter at least logd1(n)\log_{d-1}(n). On the other hand, it is well-known that expansion directly implies diameter at most Clogd1(n)C\cdot\log_{d-1}(n) for some constant C1C\geq 1 (depending on the expansion). The best upper bound a spectral proof can provide is 22 due to the Alon–Bopanna bound for spectral expansion. Using our amplification to almost-optimal spectral expansion, deduce that any expanding group GG has a constant degree-dd Cayley expander of diameter 2+od(1)2+o_{d}(1).

More precisely, we have the following.

Lemma 5.20.

Suppose {Cay(Gi,Si)}i\{\operatorname{Cay}(G_{i},S_{i})\}_{i\in\mathbb{N}} is a family of bounded degree Cayley expanders. Then, there exists a family {Cay(Gi,Si)}i\{\operatorname{Cay}(G_{i},S_{i}^{\prime})\}_{i\in\mathbb{N}} of constant degree-dd Cayley expanders with diameter at most (2+od(1))logd1(Gi)(2+o_{d}(1))\cdot\log_{d-1}(G_{i}).

  • Proof:   We apply Theorem 1.2 to the family {Cay(Gi,Si)}i\{\operatorname{Cay}(G_{i},S_{i})\}_{i\in\mathbb{N}} obtaining a new family of {Cay(Gi,Si)}i\{\operatorname{Cay}(G_{i},S_{i}^{\prime})\}_{i\in\mathbb{N}} of (d,λ)(d,\lambda)-expanders with d=1/λ2+βd=1/\lambda^{2+\beta} for some sufficiently small constants λ,β>0\lambda,\beta>0. Let Ai\symsf{A}_{i} be the normalized adjacency matrix of Cay(Gi,Si)\operatorname{Cay}(G_{i},S_{i}^{\prime}) and ni=|Gi|n_{i}=\left\lvert G_{i}\right\rvert. Let ege_{g} be the indicator vector of some fixed gGig\in G_{i}. Note that

    (AiJ/ni)teg2λt=dt/(2+β)<1/|Gi|,\displaystyle\left\lVert(\symsf{A}_{i}-\symsf{J}/n_{i})^{t}e_{g}\right\rVert_{2}\leq\lambda^{t}=d^{-t/(2+\beta)}<1/\left\lvert G_{i}\right\rvert,
    for t=(2+2β)logd(|Gi|)=(2+od,β(1))logd1(|Gi|).\displaystyle\text{for }t=(2+2\beta)\cdot\log_{d}(\left\lvert G_{i}\right\rvert)=(2+o_{d,\beta}(1))\cdot\log_{d-1}(\left\lvert G_{i}\right\rvert).

    This implies that Aiteg\symsf{A}_{i}^{t}e_{g} is supported on all elements of GiG_{i}, and thus the diameter of GiG_{i} is at most tt.      

6 Operator Expander Mixing Lemma

In Section 3, we showed an operator amplification based on walks on an auxiliary expander. The same construction can be analyzed differently by using the expander mixing lemma (EML) iteratively.

For the scalar case, this is a classic result due to Rozenman–Wigderson (see also [Bog12]). An operator version of this approach was proved by Chen, Moore, and Russell [CMR13] but they obtain a much larger degree, λ11\lambda^{-11}, rather than the λ4\lambda^{-4} similar to the expander walk approach (Theorem 3.2). Our proof obtains the bound of λ4\lambda^{-4}.

Theorem 6.1 (Iterated Operator EML).

Let SGS\subseteq G. Suppose λ(Cay(G,S))=λ0<1\lambda(\operatorname{Cay}(G,S))=\lambda_{0}<1, where λ0(0,1)\lambda_{0}\in(0,1). For every λ(0,1)\lambda\in(0,1), we can find SGS^{\prime}\subseteq G such that,

  1. 1.

    λ(Cay(G,S))λ\lambda(\operatorname{Cay}(G,S^{\prime}))\leq\lambda and |S|=Oλ0(|S|/λ4+o(1))\left\lvert S^{\prime}\right\rvert=O_{\lambda_{0}}(\left\lvert S\right\rvert/\lambda^{4+o(1)}), and

  2. 2.

    the running time is poly(|S|,1/λ0,1/λ){\mathrm{poly}}(\left\lvert S\right\rvert,1/\lambda_{0},1/\lambda).

We now show an operator version of the expander mixing lemma for completeness. As we mentioned above, a similar result was first derived in [CMR13]. While a simple generalization of EML, it is of the same nature of the generalizations of this paper and is of independent interest.

Lemma 6.2 (Matrix EML [CMR13]).

Let X=(V,E)X=(V,E) be a λ(X)\lambda(X)-spectral expander and let f:V()\symsf{f}\colon V\rightarrow\mathcal{L}(\mathcal{H}). Then,

𝔼(x,x)E[f(x)f(x)](𝔼xVX[f(x)])2opλ(X)maxxVXf(x)op2.\displaystyle\left\lVert\mathchoice{\underset{(x^{\prime},x)\in E}{\mathbb{E}}\left[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,\right]}{{\mathbb{E}}_{(x^{\prime},x)\in E}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]}{{\mathbb{E}}_{(x^{\prime},x)\in E}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]}{{\mathbb{E}}_{(x^{\prime},x)\in E}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]}-\left(\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}{{\mathbb{E}}_{x\in V_{X}}[\symsf{f}(x)\,]}\right)^{2}\right\rVert_{\textup{op}}~{}\leq~{}\lambda(X)\cdot\max_{x\in V_{X}}\left\lVert\symsf{f}(x)\,\right\rVert_{\textup{op}}^{2}.

We start with a simple claim describing an operator form the process of sampling according to the edges of an expander and sampling according to pairs of vertices. Recall the following maps from Section 3, P:𝒳\symsf{P}_{\mathcal{H}}:\mathcal{X}_{\mathcal{H}}\rightarrow\mathcal{H} and L:𝒳\symsf{L}_{\mathcal{H}}:\mathcal{H}\rightarrow\mathcal{X}_{\mathcal{H}},

P(wx)=w,L(v)=𝔼xVX[vx].\symsf{P}_{\mathcal{H}}(w\otimes x)=w,\;\;\symsf{L}_{\mathcal{H}}(v)=\mathchoice{\underset{x\in V_{X}}{\mathbb{E}}\left[v\otimes x\right]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}{{\mathbb{E}}_{x\in V_{X}}[v\otimes x]}\,.

We will need again that PopLop=1\left\lVert\symsf{P}_{\mathcal{H}}\right\rVert_{\textup{op}}\left\lVert\symsf{L}_{\mathcal{H}}\right\rVert_{\textup{op}}=1.

Claim 6.3.

Let AX\symsf{A}_{X} be the normalized adjacency matrix of a dd-regular graph XX and let JX\symsf{J}_{X} be the normalized |VX|×|VX|\left\lvert V_{X}\right\rvert\times\left\lvert V_{X}\right\rvert all-ones matrix.

𝔼(x,x)E[f(x)f(x)]\displaystyle\mathchoice{\underset{(x,x^{\prime})\in E}{\mathbb{E}}\left[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,\right]}{{\mathbb{E}}_{(x,x^{\prime})\in E}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]}{{\mathbb{E}}_{(x,x^{\prime})\in E}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]}{{\mathbb{E}}_{(x,x^{\prime})\in E}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]} =PΠzAXΠzL.\displaystyle~{}=~{}\symsf{P}_{\mathcal{H}}\symsf{\Pi}_{z}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}\symsf{\Pi}_{z}\symsf{L}_{\mathcal{H}}.
𝔼x,xV[f(x)f(x)]\displaystyle\mathchoice{\underset{x,x^{\prime}\in V}{\mathbb{E}}\left[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,\right]}{{\mathbb{E}}_{x,x^{\prime}\in V}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]}{{\mathbb{E}}_{x,x^{\prime}\in V}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]}{{\mathbb{E}}_{x,x^{\prime}\in V}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]} =PΠzJXΠzL.\displaystyle~{}=~{}\symsf{P}_{\mathcal{H}}\symsf{\Pi}_{z}\mathrel{\mathop{\symsf{J}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}\symsf{\Pi}_{z}\symsf{L}_{\mathcal{H}}.
  • Proof:   The proof is identical for both so we prove just the first one. For any ww\in\mathcal{H}, we have

    PΠzAXΠzLw\displaystyle\symsf{P}_{\mathcal{H}}\symsf{\Pi}_{z}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}\symsf{\Pi}_{z}\symsf{L}_{\mathcal{H}}w =1|VX|PΠzAXΠz(xVXxw)\displaystyle~{}=~{}\frac{1}{|V_{X}|}\symsf{P}_{\mathcal{H}}\symsf{\Pi}_{z}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}\symsf{\Pi}_{z}\left(\sum_{x\in V_{X}}x\otimes\symsf{w}\right)
    =1|VX|PΠzAX(xVXxf(x)w).\displaystyle~{}=~{}\frac{1}{|V_{X}|}\symsf{P}_{\mathcal{H}}\symsf{\Pi}_{z}\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}\left(\sum_{x\in V_{X}}x\otimes\symsf{f}(x)\,w\right).
    =1d|VX|PΠz(xVXxxxf(x)w).\displaystyle~{}=~{}\frac{1}{d|V_{X}|}\symsf{P}_{\mathcal{H}}\symsf{\Pi}_{z}\left(\sum_{x\in V_{X}}\sum_{x^{\prime}\sim x}x^{\prime}\otimes\symsf{f}(x)\,w\right).
    =1|E|P(xVXxxxf(x)f(x)w).\displaystyle~{}=~{}\frac{1}{|E|}\symsf{P}_{\mathcal{H}}\left(\sum_{x\in V_{X}}\sum_{x^{\prime}\sim x}x^{\prime}\otimes\symsf{f}(x^{\prime})\,\symsf{f}(x)\,w\right).
    =1|E|xxf(x)f(x)w=𝔼(x,x)E[f(x)f(x)]w.\displaystyle~{}=~{}\frac{1}{|E|}\sum_{x\sim x^{\prime}}\symsf{f}(x^{\prime})\,\symsf{f}(x)\,w=\mathchoice{\underset{(x^{\prime},x)\in E}{\mathbb{E}}\left[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,\right]}{{\mathbb{E}}_{(x^{\prime},x)\in E}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]}{{\mathbb{E}}_{(x^{\prime},x)\in E}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]}{{\mathbb{E}}_{(x^{\prime},x)\in E}[\symsf{f}(x^{\prime})\,\cdot\symsf{f}(x)\,]}w.

    as claimed.      

We now prove the operator mixing lemma above.

  • Proof:  [Proof of Lemma 6.2] By 6.3, it is enough to bound the operator norm

    PΠz(AXJX)ΠzLop\displaystyle\left\lVert\symsf{P}_{\mathcal{H}}\symsf{\Pi}_{z}\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}-\mathrel{\mathop{\symsf{J}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}\right)\symsf{\Pi}_{z}\symsf{L}_{\mathcal{H}}\right\rVert_{\textup{op}} PopΠzop2(AXJX)opLop\displaystyle~{}\leq~{}\left\lVert\symsf{P}_{\mathcal{H}}\right\rVert_{\textup{op}}\left\lVert\symsf{\Pi}_{z}\right\rVert_{\textup{op}}^{2}\left\lVert\left(\mathrel{\mathop{\symsf{A}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}-\mathrel{\mathop{\symsf{J}}\limits^{\vbox to0.0pt{\kern-2.0pt\hbox{$\scriptstyle\mathbin{\mathchoice{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}{\vbox{\hbox{$\scriptscriptstyle\circ$}}}}$}\vss}}}_{X}\right)\right\rVert_{\textup{op}}\left\lVert\symsf{L}_{\mathcal{H}}\right\rVert_{\textup{op}}
    λ(X)Πzop2=λ(X)maxxVXf(x)op2,\displaystyle~{}\leq~{}\lambda(X)\cdot\left\lVert\symsf{\Pi}_{z}\right\rVert_{\textup{op}}^{2}=\lambda(X)\cdot\max_{x\in V_{X}}\left\lVert\symsf{f}(x)\,\right\rVert_{\textup{op}}^{2},

    concluding the proof.      

Corollary 6.4 (Non-abelian EML).

Let X=(V,E)X=(V,E) be a λ(X)\lambda(X)-spectral expander, ρ:GU\rho\colon G\rightarrow\textup{U}_{\mathcal{H}} be an unitary representation and (gv)vVGV(g_{v})_{v\in V}\in G^{V}. Then

𝔼(u,v)E[ρ(gu)ρ(gv)](𝔼uV[ρ(gu)])2opλ(X).\displaystyle\left\lVert\mathchoice{\underset{(u,v)\in E}{\mathbb{E}}\left[\rho(g_{u})\cdot\rho(g_{v})\right]}{{\mathbb{E}}_{(u,v)\in E}[\rho(g_{u})\cdot\rho(g_{v})]}{{\mathbb{E}}_{(u,v)\in E}[\rho(g_{u})\cdot\rho(g_{v})]}{{\mathbb{E}}_{(u,v)\in E}[\rho(g_{u})\cdot\rho(g_{v})]}-\left(\mathchoice{\underset{u\in V}{\mathbb{E}}\left[\rho(g_{u})\right]}{{\mathbb{E}}_{u\in V}[\rho(g_{u})]}{{\mathbb{E}}_{u\in V}[\rho(g_{u})]}{{\mathbb{E}}_{u\in V}[\rho(g_{u})]}\right)^{2}\right\rVert_{\textup{op}}~{}\leq~{}\lambda(X).
  • Proof:   Follows immediately from Lemma 6.2 and the fact that unitary operators have operator norm bounded by 11.      

We now prove the main result of this section. This iterated amplification also appears in the derandomized squaring of Rozenman and Vadhan [RV05] used to give an alternative proof of the SL=L\textup{SL}=\textup{L} result of Reingold [Rei04].

  • Proof:  [Proof of Theorem 6.1] We amplify the expansion in two phases. The first phase amplifies the initial expansion of SS from λ0\lambda_{0} to a constant expansion λ0′′=1/4\lambda_{0}^{\prime\prime}=1/4. This phase increases the size of the generator set by a constant factor.

    (First Phase) Let ε0,γ0\varepsilon_{0},\gamma_{0} be constants such that

    ε0=λ0(1λ0)/2,    0<γ0(1λ0)/2<1\varepsilon_{0}=\lambda_{0}(1-\lambda_{0})/2,\;\;\;\;0<\gamma_{0}\leq(1-\lambda_{0})/2<1

    Let X0=(V0,E0)X_{0}=(V_{0},E_{0}) be an explicit expander via Theorem 3.7, with λ(X0)ε0\lambda(X_{0})\leq\varepsilon_{0}, degree O(1/ε02)O(1/\varepsilon_{0}^{2}) and with the number of vertices |V0|=m|S|\left\lvert V_{0}\right\rvert=m\left\lvert S\right\rvert with m=O(1)m=O(1). Replicate each element of SS mm times and still call the resulting multiset SS (observe that expansion remains λ0\lambda_{0}). For every edge (u,v)E0(u,v)\in E_{0}, add gugvg_{u}g_{v} to S0S_{0}. By Corollary 6.4,

    λ(G,S0)λ02+ε0λ0(1γ0),|S0|=9|S|/ε02=O(|S|)\lambda(G,S_{0})\leq\lambda_{0}^{2}+\varepsilon_{0}\leq\lambda_{0}(1-\gamma_{0}),\;\;\;\left\lvert S_{0}\right\rvert=9\left\lvert S\right\rvert/\varepsilon_{0}^{2}=O(\left\lvert S\right\rvert)

    Repeat this procedure log1γ01/4λ0\log_{1-\gamma_{0}}{1/4\lambda_{0}} times which ensures that the expansion is λ0′′=1/4\lambda_{0}^{\prime\prime}=1/4. Let S0S_{0} be this final set.

    (Second Phase) We will amplify the bias inductively using a stronger (i.e., more expanding) auxiliary expander graph XiX_{i} at each step. As mentioned, this inductive amplification is similar to the derandomized squaring of Rozenman and Vadhan [RV05]. We start with S0S_{0} and expansion λ0′′=22\lambda_{0}^{\prime\prime}=2^{-2} as in the first phase. At each step assume that you have a set Si1S_{i-1} with expansion λi1\lambda_{i-1}. Use Theorem 3.7, to construct Xi1X_{i-1} to have expansion λi12\lambda_{i-1}^{2} and degree at most 9/λi149/\lambda_{i-1}^{4}. Then, SiS_{i} is obtained via edges of XiX_{i} as before and we have λi2λi12\lambda_{i}\leq 2\lambda_{i-1}^{2}. It is easy to check that the recurrence yields λi2(2i)\lambda_{i}\leq 2^{-(2^{i})} for i1i\geq 1. Assume for convenience that logλ=2r\log\lambda=-2^{r}. Clearly, then we need to iterate this rr times. In each iteration, the size grows by a factor of the degree which is 9/λi149/\lambda_{i-1}^{4} and thus the final size of SS^{\prime} can be bounded as,

    |S|=|S0|i=0r19λi4|S0|9r24+4(20++2r1)=|S0|λ4(1logλ)log9Oλ0(|S|λ4+o(1)).\displaystyle\left\lvert S^{\prime}\right\rvert~{}=~{}\left\lvert S_{0}\right\rvert\prod_{i=0}^{r-1}\frac{9}{\lambda_{i}^{4}}~{}\leq~{}\left\lvert S_{0}\right\rvert\cdot 9^{r}2^{4+4\left(2^{0}+\cdots+2^{r-1}\right)}=\frac{\left\lvert S_{0}\right\rvert}{\lambda^{4}}\cdot\left(\frac{1}{\log\lambda}\right)^{\log 9}\leq O_{\lambda_{0}}\left(\frac{\left\lvert S\right\rvert}{\lambda^{4+o(1)}}\right)\,.

    concluding the proof.      

Acknowledgement

We thank Alexander Lubotzky for stimulating and enlightening discussions in the initial phase of this project. We are grateful to the anonymous reviewers for their valuable suggestions that improved the quality of this paper.

References

  • [ABN+92] N. Alon, J. Bruck, J. Naor, M. Naor, and R. Roth. Construction of asymptotically good, low-rate error-correcting codes through pseudo-random graphs. IEEE Transactions on Information Theory, 28:509–516, 1992. doi:10.1109/18.119713.
  • [ACKM19] Naman Agarwal, Karthekeyan Chandrasekaran, Alexandra Kolla, and Vivek Madan. On the expansion of group-based lifts. SIAM J. Discret. Math., 33(3):1338–1373, 2019. doi:10.1137/17M1141047.
  • [AGHP92] N. Alon, O. Goldreich, J. Håstad, and R. Peralta. Simple constructions of almost kk-wise independent random variables. Random Structures and Algorithms, 3(3):289–304, 1992.
  • [AHL+14] Dorit Aharonov, Aram W. Harrow, Zeph Landau, Daniel Nagaj, Mario Szegedy, and Umesh V. Vazirani. Local tests of global entanglement and a counterexample to the generalized area law. In Proceedings of the 55th IEEE Symposium on Foundations of Computer Science, 2014. arXiv:1410.0951, doi:10.1109/FOCS.2014.34.
  • [Alo21] Noga Alon. Explicit expanders of every degree and size. Combinatorica, February 2021. doi:10.1007/s00493-020-4429-x.
  • [ALW01] N. Alon, A. Lubotzky, and A. Wigderson. Semi-direct product in groups and zig-zag product in graphs: connections and applications. In Proceedings of the 42nd IEEE Symposium on Foundations of Computer Science, 2001.
  • [AR94] Noga Alon and Yuval Roichman. Random cayley graphs and expanders. Random Struct. Algorithms, 5(2):271–285, 1994. doi:10.1002/rsa.3240050203.
  • [AS04] Andris Ambainis and Adam D. Smith. Small pseudo-random families of matrices: Derandomizing approximate quantum encryption. In APPROX-RANDOM 2004 Proceedings, volume 3122 of Lecture Notes in Computer Science, pages 249–260. Springer, 2004. arXiv:0404075, doi:10.1007/978-3-540-27821-4\_23.
  • [AW02] R. Ahlswede and A. Winter. Strong converse for identification via quantum channels. IEEE Transactions on Information Theory, 48(3), 2002.
  • [BASTS08] Avraham Ben-Aroya, Oded Schwartz, and Amnon Ta-Shma. Quantum expanders: Motivation and constructions. In Proceedings of the 23rd IEEE Conference on Computational Complexity, pages 292–303. IEEE Computer Society, 2008. doi:10.1109/CCC.2008.23.
  • [BATS08] Avraham Ben-Aroya and Amnon Ta-Shma. A combinatorial construction of almost-ramanujan graphs using the zig-zag product. In Proceedings of the 40th ACM Symposium on Theory of Computing, pages 325–334, 2008.
  • [Bea97] Robert Beals. Quantum computation of fourier transforms over symmetric groups. In Proceedings of the 29th ACM Symposium on Theory of Computing, STOC ’97, pages 48–53, 1997.
  • [BISW01] B. Barak, R. Impagliazzo, A. Shpilka, and A. Wigderson. Dimension expanders. unpublished, 2001.
  • [BKL89] László Babai, William M. Kantor, and A. Lubotzky. Small-diameter cayley graphs for finite simple groups. Eur. J. Comb., 10, 1989.
  • [BL06] Yonatan Bilu and Nathan Linial. Lifts, discrepancy and nearly optimal spectral gap. Combinatorica, 26(5):495–519, October 2006.
  • [BL18] Emmanuel Breuillard and Alexander Lubotzky. Expansion in simple groups, 2018. arXiv:1807.03879.
  • [Bog12] Andrej Bogdanov. Techniques in the theory of computing: Lecture notes. Online, 2012. Accessed March 3, 2024. URL: https://www.andrejb.net/csc5060/notes/12L12.pdf.
  • [BS88] László Babai and Akos Seress. On the diameter of cayley graphs of the symmetric group. Journal of Combinatorial Theory, Series A, 49(1), 1988.
  • [BY13] Jean Bourgain and Amir Yehudayoff. Expansion in 𝖲𝖫2()\mathsf{SL}_{2}{(\mathbb{R})} and monotone expanders. Geometric and Functional Analysis, 23(1), 2013. doi:10.1007/s00039-012-0200-9.
  • [Che10] Yuan-You Fu-Rui Cheng. Explicit estimate on primes between consecutive cubes. Rocky Mountain Journal of Mathematics, 40(1), February 2010. arXiv:0810.2113, doi:10.1216/rmj-2010-40-1-117.
  • [CMR13] Sixia Chen, Cristopher Moore, and Alexander Russell. Small-bias sets for nonabelian groups - derandomizations of the Alon–Roichman theorem. In APPROX-RANDOM, volume 8096 of Lecture Notes in Computer Science, pages 436–451, 2013.
  • [DS09] Zeev Dvir and Amir Shpilka. Towards dimension expanders over finite fields. Combinatorica, 31(3), sep 2009. doi:10.1007/s00493-011-2540-8.
  • [DW10] Zeev Dvir and Avi Wigderson. Monotone expanders: Constructions and applications. Theory of Computing, 6(12), 2010. doi:10.4086/toc.2010.v006a012.
  • [FG15] Michael A. Forbes and Venkatesan Guruswami. Dimension Expanders via Rank Condensers. In Approximation, Randomization, and Combinatorial Optimization. Algorithms and Techniques (APPROX/RANDOM 2015), volume 40, pages 800–814, 2015. arXiv:1411.7455, doi:10.4230/LIPIcs.APPROX-RANDOM.2015.800.
  • [Fri03] Joel Friedman. A proof of Alon’s second eigenvalue conjecture. In Proceedings of the 35th ACM Symposium on Theory of Computing, 2003. arXiv:cs/0405020, doi:10.1145/780542.780646.
  • [Gil52] E.N. Gilbert. A comparison of signalling alphabets. Bell System Technical Journal, 31:504–522, 1952. doi:10.1002/j.1538-7305.1952.tb01393.x.
  • [Gil98] D. Gillman. A Chernoff bound for random walks on expander graphs. SIAM Journal on Computing, 27(4):1203–1220, 1998. doi:10.1137/S0097539794268765.
  • [Har07] Aram W. Harrow. Quantum expanders from any classical cayley graph expander. Quantum Information & Computation, 2007.
  • [Has07a] M. B. Hastings. Entropy and entanglement in quantum ground states. Physical Review B, 76(3), jul 2007. arXiv:0701055, doi:10.1103/physrevb.76.035114.
  • [Has07b] M. B. Hastings. Entropy and entanglement in quantum ground states. Phys. Rev. B, 2007.
  • [Has07c] M. B. Hastings. Random unitaries give quantum expanders. Phys. Rev. A, 76:032315, Sep 2007. arXiv:0706.0556, doi:10.1103/PhysRevA.76.032315.
  • [HH09] M. B. Hastings and A. W. Harrow. Classical and quantum tensor product expanders. Quantum Info. Comput., 2009.
  • [HLW06] Shlomo Hoory, Nathan Linial, and Avi Wigderson. Expander graphs and their applications. Bull. Amer. Math. Soc., 43(04):439–562, August 2006.
  • [JM21] Akhil Jalan and Dana Moshkovitz. Near-optimal cayley expanders for abelian groups, 2021. arXiv:2105.01149.
  • [JMO+22] Fernando Granha Jeronimo, Tushant Mittal, Ryan O’Donnell, Pedro Paredes, and Madhur Tulsiani. Explicit abelian lifts and quantum ldpc codes. In ITCS 2022, 2022.
  • [JQST20] Fernando Granha Jeronimo, Dylan Quintana, Shashank Srivastava, and Madhur Tulsiani. Unique decoding of explicit ε\varepsilon-balanced codes near the Gilbert–Varshamov bound. In Proceedings of the 61st IEEE Symposium on Foundations of Computer Science, 2020.
  • [Kas07] Martin Kassabov. Symmetric groups and expander graphs. Inventiones mathematicae, 170(2):327–354, August 2007. arXiv:0505624, doi:10.1007/s00222-007-0065-y.
  • [KKN21] Marek Kaluba, Dawid Kielak, and Piotr W. Nowak. On property (T) for Aut(Fn)\mathrm{Aut}(F_{n}) and SLn()\mathrm{SL}_{n}(\mathbb{Z}). Annals of Mathematics, 193(2):539 – 562, 2021. doi:10.4007/annals.2021.193.2.3.
  • [LP00] Alexander Lubotzky and Igor Pak. The product replacement algorithm and kazhdan’s property (t). Journal of the American Mathematical Society, 14(2):347–363, October 2000. doi:10.1090/s0894-0347-00-00356-8.
  • [LPS88] Alexander Lubotzky, R. Phillips, and Peter Sarnak. Ramanujan graphs. Combinatorica, 8:261–277, 1988.
  • [Lub11] Alexander Lubotzky. Finite simple groups of Lie type as expanders. Journal of the European Mathematical Society, pages 1331–1341, 2011. arXiv:0904.3411, doi:10.4171/JEMS/282.
  • [Lub12] Alexander Lubotzky. Expander graphs in pure and applied mathematics. Bull. Amer. Math. Soc., 2012.
  • [LZ08] Alexander Lubotzky and Efim Zelmanov. Dimension expanders. Journal of Algebra, 319(2):730–738, 2008.
  • [Mar73] G. A. Margulis. Explicit constructions of concentrators. Probl. Peredachi Inf., 9, 1973.
  • [Mar88] G. A. Margulis. Explicit group-theoretical constructions of combinatorial schemes and their application to the design of expanders and concentrators. 1988.
  • [MOP20] Sidhanth Mohanty, Ryan O’Donnell, and Pedro Paredes. Explicit near-ramanujan graphs of every degree. In Proceedings of the 52nd ACM Symposium on Theory of Computing, pages 510–523. ACM, 2020.
  • [MSS14] Adam Marcus, Daniel Spielman, and Nikhil Srivastava. Interlacing families ii: Mixed characteristic polynomials and the kadison–singer problem. Annals of Mathematics, 2014.
  • [MSS15] Adam Marcus, Daniel Spielman, and Nikhil Srivastava. Interlacing families i: Bipartite Ramanujan graphs of all degrees. Annals of Mathematics, 2015.
  • [MW04] Roy Meshulam and Avi Wigderson. Expanders in group algebras. Combinatorica, 24, 2004.
  • [Nil91] Alon Nilli. On the second eigenvalue of a graph. Discrete Mathematics, 91(2):207–210, 1991. doi:10.1016/0012-365X(91)90112-F.
  • [NN90] J. Naor and M. Naor. Small-bias probability spaces: efficient constructions and applications. In Proceedings of the 22nd ACM Symposium on Theory of Computing, pages 213–223, 1990.
  • [Pin73] Mark S. Pinsker. On the complexity of a concentrator. In 7th International Teletraffic Conference, 1973.
  • [PZ01] Igor Pak and Andrzej Zuk. Two Kazhdan constants and mixing of random walks. Technical report, Int. Math. Res. Not. 2002, 2001.
  • [Rei04] Omer Reingold. Undirected st-connectivity in log-space. Technical Report TR04-094, Electronic Colloquium on Computational Complexity, 2004.
  • [Rei05] Omer Reingold. Undirected ST-connectivity in log-space. In Proceedings of the 37th ACM Symposium on Theory of Computing, pages 376–385, 2005.
  • [RL10] J.D. Rogawski and A. Lubotzky. Discrete Groups, Expanding Graphs and Invariant Measures. Modern Birkhäuser Classics. Birkhäuser Basel, 2010.
  • [RSW06] Eyal Rozenman, Aner Shalev, and Avi Wigderson. Iterative construction of cayley expander graphs. Theory of Computing, 2(5):91–120, 2006.
  • [RV05] Eyal Rozenman and Salil Vadhan. Derandomized squaring of graphs. In Proceedings of RANDOM’05, pages 436–447. Springer-Verlag, 2005.
  • [RVW00] O. Reingold, S. Vadhan, and A. Wigderson. Entropy waves, the zig-zag graph product, and new constant-degree expanders and extractors. In Proceedings of the 41st IEEE Symposium on Foundations of Computer Science, 2000.
  • [SS96] L. L. Scott and J. P. Serre. Linear Representations of Finite Groups. Graduate Texts in Mathematics. Springer New York, 1996.
  • [Tro15] Joel A. Tropp. An introduction to matrix concentration inequalities. Found. Trends Mach. Learn., 2015.
  • [TS17] Amnon Ta-Shma. Explicit, almost optimal, epsilon-balanced codes. In Proceedings of the 49th ACM Symposium on Theory of Computing, STOC 2017, pages 238–251, New York, NY, USA, 2017. ACM.
  • [Vad12] Salil P. Vadhan. Pseudorandomness. Now Publishers Inc., 2012.
  • [Var57] R.R. Varshamov. Estimate of the number of signals in error correcting codes. Doklady Akademii Nauk SSSR, 117:739–741, 1957.
  • [Wig18] Avi Wigderson. Mathematics and computation. Book draft at https://www.math.ias.edu/files/mathandcomp.pdf, 2018.