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

The Computational Complexity of Quantum Determinants

Shih-Han Hung111Department of Computer Science, University of Texas at Austin. Email: [email protected].    En-Jui Kuo222Joint Center for Quantum Information and Computer Science, NIST and University of Maryland, College Park, Maryland 20742, USA. Email:[email protected]
Abstract

In this work, we study the computational complexity of quantum determinants, a qq-deformation of matrix permanents: Given a complex number qq on the unit circle in the complex plane and an n×nn\times n matrix XX, the qq-permanent of XX is defined as

Perq(X)=σSnq(σ)X1,σ(1)Xn,σ(n),\mathrm{Per}_{q}(X)=\sum_{\sigma\in S_{n}}q^{\ell(\sigma)}X_{1,\sigma(1)}\ldots X_{n,\sigma(n)},

where (σ)\ell(\sigma) is the inversion number of permutation σ\sigma in the symmetric group SnS_{n} on nn elements. The function family generalizes determinant and permanent, which correspond to the cases q=1q=-1 and q=1q=1 respectively.

For worst-case hardness, by Liouville’s approximation theorem and facts from algebraic number theory, we show that for primitive mm-th root of unity qq for odd prime power m=pkm=p^{k}, exactly computing qq-permanent is 𝖬𝗈𝖽p𝖯\mathsf{Mod}_{p}\mathsf{P}-hard. This implies that an efficient algorithm for computing qq-permanent results in a collapse of the polynomial hierarchy. Next, we show that computing qq-permanent can be achieved using an oracle that approximates to within a polynomial multiplicative error and a membership oracle for a finite set of algebraic integers. From this, an efficient approximation algorithm would also imply a collapse of the polynomial hierarchy. By random self-reducibility, computing qq-permanent remains to be hard for a wide range of distributions satisfying a property called the strong autocorrelation property. Specifically, this is proved via a reduction from 11-permanent to qq-permanent for O(1/n2)O(1/n^{2}) points zz on the unit circle. Since the family of permanent functions shares common algebraic structure, various techniques developed for the hardness of permanent can be generalized to qq-permanents.

1 Introduction

The study of matrix permanents has played an essential role in theoretical computer science and combinatorics. In a seminal work, Valiant proved that computing matrix permanent is #𝖯{\#\mathsf{P}}-hard, even if the matrix is over integers, non-negative integers or the set {0,1}\{0,1\} [Val79a]. This is considered to be one of the most influential results in computational complexity theory. These results have been used to justify of the hardness of other computational tasks in graph theory [Bro86, CX21] and matrix analysis [HJ12]. By contrast, matrix determinant is known to be polynomial-time computable from standard tools in elementary linear algebra [M+97].

An interesting observation is that both permanent and determinant of a matrix XX can be written as a polynomial in matrix elements with phase coefficients, i.e., they are elements on the unit circle in the complex plane. More specifically, both functions can be written as a summation of q(σ)X1,σ(1)Xn,σ(n)q^{\ell(\sigma)}X_{1,\sigma(1)}\ldots X_{n,\sigma(n)} over permutations σ\sigma on nn elements for a phase qq. Here (σ)\ell(\sigma) is the inversion number of the permutation σ\sigma, defined as the number of transpositions on adjacent elements applied to restore σ(1),,σ(n)\sigma(1),\ldots,\sigma(n) to its natural order 1,2,,n1,2,\ldots,n [Yan91]. In the case of permanent, q=1q=1, whereas in the case of determinant, q=1q=-1. However, given the expression of both functions in a unified form, what makes their computational complexities so different remains elusive.

A natural idea to explore this question is to study a qq-deformation of the matrix polynomial, connecting the cases q{1,1}q\in\{-1,1\}. In mathematics literature, the qq-deformation on the unit circle or in [1,1][-1,1] has been studied, and was given different names including μ\mu-permanent, qq-permanent, qq-determinant, or quantum determinant [Yan91, BL94, Lal98, dF18, DF10, dS18, TAG93]. However, to our knowledge, none of these results discussed the complexity of computing the analogue. In this work, we focus on the unit circle, and study the computational complexity of the function defined as

Perz(X):=σSnz(σ)iXi,σ(i).\displaystyle\mathrm{Per}_{z}(X):=\sum_{\sigma\in S_{n}}z^{\ell(\sigma)}\prod_{i}X_{i,\sigma(i)}. (1)

for the symmetric group SnS_{n} on nn elements and a complex number zz such that |z|=1|z|=1. In the rest of the paper, for matrix XX, we will call Perz(X)\mathrm{Per}_{z}(X) the zz-permanent of XX.

The zz-permanent is not the only qq-analogue of matrix permanents. In particular, immanants [Har85, Bür00, MM11, dRA13, Spi20, Cur21] and fermionants [MM11, BKW19, RA13, dRA13] are two families of functions that have been widely studied. They are both defined with class functions, i.e., for some function ff invariant under conjugation on the symmetric group, these function can be written in the form

σSnf(σ)X1,σ(1)Xn,σ(n).\displaystyle\sum_{\sigma\in S_{n}}f(\sigma)X_{1,\sigma(1)}\ldots X_{n,\sigma(n)}. (2)

For immanants, f(σ)=χλ(σ)f(\sigma)=\chi_{\lambda}(\sigma) where χλ\chi_{\lambda} is the character of the group representation λ\lambda of SnS_{n}. The permanent corresponds to the special case that χλ\chi_{\lambda} is the trivial representation, and the determinant corresponds to the sign representation. In a recent breakthrough, Curticapean classified the complexity of computing immanants for partition families. The complexity of immanants is settled down by proving that under plausible assumptions, for every partition family Λ\Lambda, matrix immanants are efficiently computable if and only if a quantity b(Λ)b(\Lambda), which counts the boxes to the right of the first column in the Young diagram of λΛ\lambda\in\Lambda, is unbounded [Cur21]. For fermionants, f(σ)=(1)n(k)c(σ)f(\sigma)=(-1)^{n}(-k)^{c(\sigma)}, where c(σ)c(\sigma) denotes the number of cycles in σ\sigma. On the hardness side, Martins and Moore showed that fermionants are #𝖯{\#\mathsf{P}}-hard for k>2k>2 under Turing reduction, and for k=2k=2, the problem is 𝖯\oplus\mathsf{P}-hard [MM11]. On the other hand, Björklund, Kaski, and Williams gave an algorithm in time 2mΩ(m/loglogq))O(M(q))2^{m-\Omega(m/\log\log q))}\cdot O(\mathrm{M}(q)) for m×mm\times m matrices over a finite field 𝔽q\mathbb{F}_{q} of qq elements, where M(q)\mathrm{M}(q) is the time complexity of multiplication and division over 𝔽q\mathbb{F}_{q} [BKW19]. These results justify that these functions are computationally intractable in general, except in special cases.

However, to our knowledge, no classification of the complexity of zz-permanents has been given for any z{1,+1}z\notin\{-1,+1\}, and none of the aforementioned results seems to obviously imply the hardness for this family, partly because the inversion number is not a class function. In fact, no known complexity result for qq-deformations of matrix permanents has even been given beyond class functions.

In addition to the fact that the coefficients are not class functions, one may wonder why zz-permanent (also called quantum determinant) is interesting object to study? First, in representation theory, the standard Lie algebra can be generalized into the so-called qq-deformed algebra (also known as quantum group [Maj00]). Quantum determinants show up naturally in such contexts. More specifically, consider the zz-Grassmann algebra which is the associative algebra K[z]K[z] generated by x1,x2,,xnx_{1},x_{2},...,x_{n} satisfying the xi2=0x_{i}^{2}=0 for every i[n]i\in[n], and xixj=zxjxix_{i}x_{j}=zx_{j}x_{i} for i<ji<j. It is straightforward to see that

(j=1na1jxj)(j=1na2jxj)(j=1nanjxj)=Perz(A)x1x2xn.\mathopen{}\mathclose{{}\left(\sum_{j=1}^{n}a_{1j}x_{j}}\right)\mathopen{}\mathclose{{}\left(\sum_{j=1}^{n}a_{2j}x_{j}}\right)\ldots\mathopen{}\mathclose{{}\left(\sum_{j=1}^{n}a_{nj}x_{j}}\right)=\mathrm{Per}_{z}(A)x_{1}x_{2}\ldots x_{n}. (3)

which AA is nn by nn matrix such that Aij=aijA_{ij}=a_{ij}. Quantum determinants also serve as an important object in the representation of such qq-deformed algebra [ER98], and are closely related to the representation theory over different characteristics. Second, it is known that there are many applications of quantum determinants in various areas of physics, including string theory [FK06] and conformal field theory [GRASRA96, PS90]. In quantum information theory, quantum determinants also play an essential role in the Fock representations of qq-deformed commutation relations [BLW17] or anyonic relation [GM04], generalizing the standard bosonic/fermionic commutation relation [BS91].

Motivated by the above observations, we initiate the study on the computational complexity of zz-permanents, by giving general hardness results for complex number zz on the unit circle. In the following sections, we give an overview of our contributions.

1.1 Our Contributions

In this paper, we study the computational complexity of zz-permanent for an mm-th primitive root of unity zz for integer m>2m>2. Our contribution is built upon tools from algebraic number theory. We summarize our result in Table 1. From Section 3 to Section 5, we prove one theorem in each section.

z=ζpk,𝕋{±1},|Perz(X)|2z=\zeta_{p^{k}},\in{\displaystyle\mathbb{T}\setminus\{\pm 1\},}|\mathrm{Per}_{z}(X)|^{2} Exact Approximate within multiplicative error poly(n)\operatorname{poly}(n)
Worst case Theorem 3.10 Theorem 4.2
Average case for XX\sim\mathcal{H} constant fraction Theorem 5.3 Theorem 6.4
Table 1: We list hardness results into common four categories and specify the four main theorems shown in this paper.

1.1.1 Worst-case Hardness

First, we study the hardness of computing ζm\zeta_{m}-permanent of a n×nn\times n binary matrix XX for an mm-th primitive root of unity ζm\zeta_{m}. By (1), we can write the zz-permanent as a polynomial in zz whose coefficients can be written as multivariate polynomials in the elements of XX. For z{+1,1}z\notin\{+1,-1\}, zz-permanent is in general a complex number which can be written as an integral polynomial in zz. Since complex number arithmetic can only be computed using a finite precision, we first clarify what it means by an algorithm computing Perz(X)\mathrm{Per}_{z}(X) exactly.

While we do not know how to give a sensible definition for every zz on the unit circle, applying facts of cyclotomic fields yields a natural definition for a primitive root of unity and a binary matrix. More specifically, ζm\zeta_{m}-permanent of a binary matrix can be written as an integral polynomial in ζm\zeta_{m} of degree d=degΦm1d=\deg\Phi_{m}-1, where Φm\Phi_{m} is the mm-th cyclotomic polynomial. This implies that representing Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) as a degree-dd polynomial, the tuple of coefficients is unique. Thus we define an algorithm which exactly computes Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) to be one that outputs a set of coefficients a0,,ada_{0},\ldots,a_{d} such that Perζm(X)=i=0daiζmi\mathrm{Per}_{\zeta_{m}}(X)=\sum_{i=0}^{d}a_{i}\zeta_{m}^{i}.

Given this definition, we show that for odd prime power m=pkm=p^{k}, any oracle 𝒪\mathcal{O} that computes ζm\zeta_{m}-permanent implies that 𝖡𝖯𝖯𝖬𝗈𝖽p𝖯𝖡𝖯𝖯𝒪\mathsf{BPP}^{\mathsf{Mod}_{p}\mathsf{P}}\subseteq\mathsf{BPP}^{\mathcal{O}}.

Theorem 1.1 (Hardness of zz-permanent, informal).

For odd prime pp, prime power m=pkm=p^{k} and a primitive mm-th root of unity ζm\zeta_{m}, it is 𝖬𝗈𝖽p𝖯\mathsf{Mod}_{p}\mathsf{P}-hard to compute ζm\zeta_{m}-permanent of {0,1}\{0,1\}-matrices. More explicitly, for every oracle 𝒪\mathcal{O} that on input a binary n×nn\times n matrix XX outputs the coefficients of Perζm(X)\mathrm{Per}_{\zeta_{m}}(X), it holds that

𝖯𝖧𝖡𝖯𝖯𝖬𝗈𝖽p𝖯𝖡𝖯𝖯𝒪.\displaystyle\mathsf{PH}\subseteq\mathsf{BPP}^{\mathsf{Mod}_{p}\mathsf{P}}\subseteq\mathsf{BPP}^{\mathcal{O}}. (4)

The first inequality in (4) follows from the generalized Toda’s theorem [TO92, Tod91]. To prove the second, we rely on the fact that for prime power m=pkm=p^{k}, the cyclotomic polynomial evaluated on 11 is pp. This means that the summation of the coefficients is congruent to Per(X)\mathrm{Per}(X) modulo pp. Then the second inequality follows from the fact that computing Per(X)\mathrm{Per}(X) modulo an odd prime is known to be 𝖬𝗈𝖽p𝖯\mathsf{Mod}_{p}\mathsf{P}-hard by Valiant [Val79a].

While our definition of exact computation seems to be natural, one may wonder whether the requirement to output all the coefficients correctly would be too strong to allow the existence of efficient classical algorithms. We answer this question by studying hardness of approximating Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) in 2\ell_{2}-norm. In particular, we show that finding a close enough complex number to Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) already suffices to extract all the coefficients using an 𝖭𝖯\mathsf{NP} oracle.

1.1.2 Hardness of Approximation

Next, we show that an oracle that approximates Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) to within polynomial multiplicative error is also 𝖬𝗈𝖽p𝖯\mathsf{Mod}_{p}\mathsf{P}-hard.

Theorem 1.2 (Hardness of approximation, informal).

For odd prime pp, prime power m=pkm=p^{k} and a primitive mm-th root of unity ζm\zeta_{m}, it is 𝖬𝗈𝖽p𝖯\mathsf{Mod}_{p}\mathsf{P}-hard to approximate ζm\zeta_{m}-permanent of {0,1}\{0,1\}-matrices. More explicitly, for every algorithm 𝒪\mathcal{O} that on input an n×nn\times n binary matrix XX, outputs an approximation to within multiplicative error g=poly(n)g=\operatorname{poly}(n), it holds that

𝖯𝖧𝖡𝖯𝖯𝖬𝗈𝖽p𝖯𝖡𝖯𝖯𝖭𝖯𝒪.\displaystyle\mathsf{PH}\subseteq\mathsf{BPP}^{\mathsf{Mod}_{p}\mathsf{P}}\subseteq\mathsf{BPP}^{\mathsf{NP}^{\mathcal{O}}}. (5)

Theorem 1.2 implies that an efficient simulation of 𝒪\mathcal{O} leads to a collapse of the polynomial hierarchy. The first inequality, again, follows from generalized Toda’s theorem [TO92, Tod91]. We then prove the second inequality in the following two steps. First, we show that there is a deterministic algorithm using polynomially many queries to 𝒪\mathcal{O} to decrease the error to δ=2poly(n)\delta=2^{-\operatorname{poly}(n)} in 2\ell_{2}-norm. Our algorithm always output a complex number δ\delta-close to Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) successfully. However, this step alone does not yield a tuple of coefficients that satisfies our definition of exact computation. We overcome this issue by invoking a variant of Liouville’s approximation theorem. From the theorem, we show that two arbitrary algebraic integers, when represented as the minimal polynomial in ζm\zeta_{m} with coefficients in [A,B][-A,B] for A,B=2poly(n)A,B=2^{\operatorname{poly}(n)} must be 2poly(n)2^{-\operatorname{poly}(n)}-far in 2\ell_{2}-distance. Using an oracle that determines whether there is a δ\delta-close algebraic integer, we can perform a binary search to identify all the coefficients. Finally, we show determining whether there is an algebraic integer that is close enough to our approximation, and whose coefficients are bounded in [A,B][-A,B] that can be solved using a non-deterministic Turing machine.

As a bonus, in Theorem 3.11, we give evidence that there is no polynomial algorithm for computing most of the roots of unity.

1.1.3 Average-Case Hardness of Exactly Computing zz-Permanent

For the average-case hardness of 1-permanent, Aaronson and Arkhipov [AA11] showed that under reasonable assumptions, computing Per(X)\mathrm{Per}(X) exactly for most of the Gaussian matrices is #𝖯{\#\mathsf{P}}-hard. Haferkamp, Hangleiter, Eisert, and Gluza [HHEG20] proved the hardness for most of the truncated uniform distribution.

In Section 5, we adapt their techniques to show that for every zz, if computing zz-permanent is hard in the worst case, via random self-reducibility, computing zz-permanent of Gaussian matrices is hard. This proof utilizes the Berlekamp-Welch algorithm and the fact that zz-permanent is a low-degree polynomial of its matrix elements.

Moreover, we extend the hardness results to a family of probability distributions that satisfy a property, called the strong autocorrelation property. Specifically, a random matrix is said to satisfy the strong autocorrelation property if the total variational distance between shifted or scaled original distribution and the original one is small (see Definition 5.2 for a detailed definition). We show that computing zz-permanent of matrices drawn from distributions satisfying this property remains to be #𝖯{\#\mathsf{P}}-hard. The hardness result can thus be applied to, for example, a uniform distribution centered around zero and truncated at some chosen threshold.

1.1.4 Average-Case Hardness of Approximating zz-Permanent

In Section 6, we show that if one can compute particular Ω(n2)\Omega(n^{2}) points of Perz(X)\mathrm{Per}_{z}(X) in time poly(n,1/ϵ,1/δ)\operatorname{poly}(n,1/\epsilon,1/\delta), then one can determine Per(X)\mathrm{Per}(X) for XX sampled from a distribution \mathcal{H} that satisfies the strong autocorrelation property. This implies that given a distribution \mathcal{H} and XX\sim\mathcal{H}, approximating Per(X)\mathrm{Per}(X) is hard. Our results work for multiplicative and additive errors.

In addition, we also show that approximating Perz(X)\mathrm{Per}_{z}(X) is as hard as approximating Perz(X)\mathrm{Per}_{z^{*}}(X) for any distribution XX such that X,XX,X^{*} are identically distributed. The proof works for |Perz(X)|2|\mathrm{Per}_{z}(X)|^{2} as well.

1.2 Related Work

As we described in the introduction, permanent and determinant have completely different computational complexity. Any path that connects both cases is interesting from the point of computational complexity. More generally, one can consider such

νSnf(ν)i=1nAi,ν(i)\sum_{\nu\in S_{n}}f(\nu)\prod_{i=1}^{n}A_{i,\nu(i)} (6)

function and ask the complexity of functions in the form of (6). People have already investigated a few function families including immanants [MM11, Har85, Bür00, dRA13, Spi20, Bür00] and fermionants [MM11, BKW19, RA13, dRA13]. However, our work is completely different from them in the following perspective:

  • Unlike immanants and fermionants, zz-permanent is not a class function. To our best knowledge, the complexity of such matrix function (6) has only been investigated when f(ν)f(\nu) is a class function. So one can not use the standard immanant results.

  • Since zz-permanent is not a class function, we provide completely new techniques involving algebraic number theory.

  • To our best knowledge, All results of approximate hardness in the worst case and average case are investigated for 1-permanent. For f(ν)1f(\nu)\neq 1, no approximate hardness results are given previously in the literature. In Section 4, Section 5, and Section 6, we provide technical tools which lead to the results of approximate hardness for the general function ff.

1.3 Organization

The rest of the paper is structured as follows. A preliminary background is presented in Section 2. In Section 3, for worst-case hardness, we show that for primitive mm-th root of unity for odd prime power m=pkm=p^{k}, exactly computing zz-permanent is 𝖬𝗈𝖽p𝖯\mathsf{Mod}_{p}\mathsf{P}-hard. In Section 4, we show that efficiently approximating the zz-permanent of binary matrices implies a collapse of the polynomial hierarchy. In Section 5, based on the worst-case hardness, we give average-case hardness of exactly computing zz-permanent for a family of distributions satisfying the strong autocorrelation property. In Section 6, we move to the average-case hardness. Finally, in Section 7, we summarize our results and discuss a few open questions related to zz-permanent.

2 Preliminaries

2.1 Introduction to Algebraic Number Theory

Here we introduce some facts about field extension [Ben87] and algebraic number theory [Lan13, Neu13, Wei98] relevant to our work, although in this paper, we will only use facts about cyclotomic fields.

First, we recall basic definitions in abstract algebra and algebraic number theory. An abelian group is a set of elements closed under its group operation often denoted ++ (for an additive group) or \cdot (for a multiplicative group) and every element has an inverse. A commutative ring is an additive group that also has an associative multiplication operation and a multiplicative identity 11. A field is a ring, and its non-zero elements form an abelian multiplicative group. For example, the set \mathbb{Q} of rational numbers is a field and the set \mathbb{Z} of integers forms a ring.

A field KK is said to be an extension field (or simply extension), denoted K/FK/F, of field FF if KK contains FF as a subfield. The extension field KK can be viewed as a vector space over FF. The degree of the field extension K/FK/F, denoted [K:F][K:F], is the dimension of KK as a vector space of FF. Given a field FF and some value α\alpha, let F(α)F(\alpha), called FF adjoined α\alpha, be the smallest field containing FF and α\alpha. The operation of adjoining an element to a field is a field extension. For example, K=(ζ3)={a+bζ3:a,b}K=\mathbb{Q}(\zeta_{3})=\{a+b\zeta_{3}:a,b\in\mathbb{Q}\} is an extension field of \mathbb{Q} of degree two for ζ3=ei2π/3\zeta_{3}=e^{i2\pi/3} since every element of KK can be written as a vector of two components in \mathbb{Q}. We also denote [x]\mathbb{Z}[x] for indeterminant xx the ring of polynomials with integer coefficients. For complex number α\alpha, the set [α]={f(α):f(x)[x]}\mathbb{Z}[\alpha]=\{f(\alpha):f(x)\in\mathbb{Z}[x]\} forms a ring.

A number field is an extension field of \mathbb{Q} of finite degree. We recall the definition of number fields and algebraic numbers.

Definition 2.1.

An algebraic number field (or simply number field) is an extension field KK of the field of rational numbers \mathbb{Q} such that the field extension K/K/\mathbb{Q} has finite degree (hence is an algebraic field extension).

Definition 2.2.

A complex number xx\in\mathbb{C} is an algebraic number if there exists a nonzero polynomial ff with rational coefficients (equivalently, integral) such that xx is a root of ff. The set of algebraic numbers ¯\bar{\mathbb{Q}} forms a subfield of .\mathbb{C}.

Definition 2.3.

A complex number xx\in\mathbb{C} is an algebraic integer if there exists a nonzero monic polynomial ff (whose leading coefficient equals one) with integral coefficients such that xx is a root of ff.

The set of algebraic integer forms a subring of ¯\bar{\mathbb{Q}}. We also have αK\alpha\in K is an algebraic integer if the minimal monic polynomial of α\alpha over \mathbb{Q} is in [x]\mathbb{Z}[x] [Lan13]. We have a very simple fact about rational field and algebraic number using Gauss’s lemma.

Theorem 2.4 ([Lan13]).

Let 𝒪\mathcal{O} be the set of algebraic integers. The intersection of algebraic integers and \mathbb{Q} is \mathbb{Z}, i.e., 𝒪=.\mathbb{Q}\cap\mathcal{O}=\mathbb{Z}.

Definition 2.5 (Ring of integers).

The ring of integers 𝒪K\mathcal{O}_{K} of an algebraic number field KK is the ring of all algebraic integers contained in KK, i.e., 𝒪K=𝒪K\mathcal{O}_{K}=\mathcal{O}\cap K.

The cyclotomic field is a number field obtained by adjoining a primitive root of unity to \mathbb{Q}. In this paper, we denote ζn\zeta_{n} an nnth primitive root of unity for integer n>1n>1. Note that a choice of the primitive roots of unity is not unique. Indeed, for n>1n>1, e2πia/ne^{2\pi ia/n} is a primitive root of unity if gcd(a,n)=1\gcd(a,n)=1.

Definition 2.6 (Cyclotomic fields).

The nnth cyclotomic field is the extension (ζn)\mathbb{Q}(\zeta_{n}) of \mathbb{Q} generated by ζn\zeta_{n}. More explicitly:

(ζn)={i=0n1aiζni|ai}.\mathbb{Q}(\zeta_{n})=\mathopen{}\mathclose{{}\left\{\mathopen{}\mathclose{{}\left.\sum_{i=0}^{n-1}a_{i}\zeta_{n}^{i}}\right|a_{i}\in\mathbb{Q}}\right\}. (7)

While there can be multiple choices of ζn\zeta_{n}, the field (ζn)\mathbb{Q}(\zeta_{n}) is unique. This follows from a simple observation: Let ζ,ξ\zeta,\xi be two distinct nnth primitive roots of unity. Since there exists a{0,1,,n1}a\in\{0,1,\ldots,n-1\} and gcd(a,n)=1\gcd(a,n)=1 such that ξ=ζa\xi=\zeta^{a}, (ξ)(ζ)\mathbb{Q}(\xi)\subseteq\mathbb{Q}(\zeta) and thus they are equal. Similarly, we consider the ring

[ζn]={i=0n1aiζni|ai}.\mathbb{Z}[\zeta_{n}]=\mathopen{}\mathclose{{}\left\{\mathopen{}\mathclose{{}\left.\sum_{i=0}^{n-1}a_{i}\zeta_{n}^{i}}\right|a_{i}\in\mathbb{Z}}\right\}. (8)

For cyclotomic field (ζn)\mathbb{Q}(\zeta_{n}), the ring of integers is [ζn]\mathbb{Z}[\zeta_{n}].

Theorem 2.7 ([Lan13]).

For integer n>1n>1 and K=(ζn)K=\mathbb{Q}(\zeta_{n}), 𝒪K=K𝒪=[ζn]\mathcal{O}_{K}=K\cap\mathcal{O}=\mathbb{Z}[\zeta_{n}].

We will need some useful properties about Euler’s totient function [Lan13].

Definition 2.8 (Euler’s totient function [Lan13]).

Euler’s totient function ϕ\phi counts the number of positive integers up to a given integer nn that are relatively prime to nn. More explicitly, it can be computed through the following formula:

ϕ(n)=npn(11p),{\displaystyle\phi(n)=n\prod_{p\mid n}\mathopen{}\mathclose{{}\left(1-{\frac{1}{p}}}\right),} (9)

where the product is over the distinct prime numbers dividing n.n.

We will need the following lower bound of ϕ(n)\phi(n) [RS62].

Theorem 2.9 ([RS62]).

For n>2n>2,

ϕ(n)>neγloglogn+3loglogn.\phi(n)>{\frac{n}{e^{\gamma}\;\log\log n+{\frac{3}{\log\log n}}}}. (10)

simply speaking: the order of ϕ(n)\phi(n) is nearly nn. So this implies that we always have ϕ(Ω(nc))>n\phi(\Omega(n^{c}))>n for any c>1c>1 for large enough n.n. Here γ\gamma is Euler–Mascheroni constant.

For an nnth primitive root of unity ζn\zeta_{n}, the nnth cyclotomic polynomial Φn\Phi_{n} is the minimal polynomial of ζn\zeta_{n} over \mathbb{Q}.

Definition 2.10 ([Lan12]).

For integer n1n\geq 1, the nnth cyclotomic polynomial is defined as

Φn(x)=gcd(k,n)=11kn(xe2iπkn).{\displaystyle\Phi_{n}(x)=\prod_{\stackrel{{\scriptstyle 1\leq k\leq n}}{{\gcd(k,n)=1}}}\mathopen{}\mathclose{{}\left(x-e^{2i\pi{\frac{k}{n}}}}\right)}. (11)

The polynomial Φn(x)\Phi_{n}(x) has integer coefficients, i.e., Φn(x)[x]\Phi_{n}(x)\in\mathbb{Z}[x].

We include some basic properties of cyclotomic polynomials useful in our work. Since Φn(x)\Phi_{n}(x) is irreducible over \mathbb{Q}, so it is the minimal polynomial of ζn\zeta_{n} over \mathbb{Q}, the degree [(ζn):]=deg Φn=ϕ(n)[\mathbb{Q}(\zeta_{n}):\mathbb{Q}]=\text{deg }\Phi_{n}=\phi(n). The minimal polynomial of every element in (ζn)\mathbb{Q}(\zeta_{n}) over \mathbb{Q} has degree at most ϕ(n)\phi(n). The equivalent statement is that the following set {e2πaij/q:0j<ϕ(q)}\{e^{2\pi aij/q}:0\leq j<\phi(q)\} is linearly independent over \mathbb{Q}. For n2n\geq 2,

Φn(0)\displaystyle\Phi_{n}(0) =1,\displaystyle=1,
Φn(1)\displaystyle\Phi_{n}(1) =1, if n is not a prime power,\displaystyle=1,\quad\text{ if $n$ is not a prime power},
Φn(1)\displaystyle\Phi_{n}(1) =p, if n=pk is a prime power with k1.\displaystyle=p,\quad\text{ if $n=p^{k}$ is a prime power with $k\geq 1$.} (12)
Theorem 2.11 (Bounds for the coefficients of cyclotomic polynomials [Vau75]).

Let Φn(x):=i=0ϕ(n)aixi\Phi_{n}(x):=\sum_{i=0}^{\phi(n)}a_{i}x^{i} be the nnth cyclotomic polynomial. The maximum absolute value of coefficient aia_{i} of Φn(x)\Phi_{n}(x) satisfies

maxi,0iϕ(n)|ai|e12d(n)logn,\max_{i\in\mathbb{N},0\leq i\leq\phi(n)}|a_{i}|\leq e^{\frac{1}{2}d(n)\log n}, (13)

where d(n)d(n) is the divisor function which counts the number of divisors of nn (including 11 and nn) for integer n1n\geq 1.333For instance, the divisors of 66 are 11, 22, 33, and 66, and thus d(6)=4d(6)=4.

For n=p1a1,pkakn=p_{1}^{a_{1}}\ldots,p_{k}^{a_{k}}, d(n)=(a1+1)(ak+1)d(n)=(a_{1}+1)\ldots(a_{k}+1). Since for prime p2p\geq 2 and integer a1a\geq 1, a+1paa+1\leq p^{a}, we can get the trivial bound d(n)nd(n)\leq n. By Theorem 2.11, we can represent Φn(x)\Phi_{n}(x) using at most only O(nlogn)O(n\log n) bits.

The following theorem will be useful.

Theorem 2.12 (Liouville’s approximation theorem [LW12]).

Let α\alpha\in\mathbb{C} be an algebraic number with minimal polynomial f(x)=a0xd+a1xd1+..+adf(x)=a_{0}x^{d}+a_{1}x^{d-1}+..+a_{d} of degree dd. Then for any a,ba,b\in\mathbb{Z} that are relatively prime,

|αab|1bda0(2|α¯|+1)d1\mathopen{}\mathclose{{}\left|\alpha-\frac{a}{b}}\right|\geq\frac{1}{b^{d}a_{0}(2|\bar{\alpha}|+1)^{d-1}} (14)

where f(x)=a0σ(xσ(α))f(x)=a_{0}\prod_{\sigma}(x-\sigma(\alpha)), and |α¯||\bar{\alpha}| the maximum complex modulus of the algebraic conjuguates of α\alpha in \mathbb{C}, i.e., |α¯|=max|σ(α)||\bar{\alpha}|=\max|\sigma(\alpha)|.

In particular, since 0 and 11 are relatively prime, applying Theorem 2.12, for every algebraic number α\alpha whose minimal polynomial is f(x)=a0xd++adf(x)=a_{0}x^{d}+\ldots+a_{d}, |α|1a0(|α¯|+1)d1|\alpha|\geq\frac{1}{a_{0}(|\bar{\alpha}|+1)^{d-1}}. For two algebraic integers α,β[ζn]\alpha,\beta\in\mathbb{Z}[\zeta_{n}], αβ[ζn]\alpha-\beta\in\mathbb{Z}[\zeta_{n}], and |αβ|1(|αβ¯|+1)ϕ(n)1|\alpha-\beta|\geq\frac{1}{(|\overline{\alpha-\beta}|+1)^{\phi(n)-1}}.

Theorem 2.13 (Fermat’s little theorem [Wei04]).

Let pp be a prime number. Then for any integer aa, the number apaa^{p}-a is an integer multiple of pp. In the notation of modular arithmetic, this is expressed as

apa(modp).a^{p}\equiv a\pmod{p}. (15)

Theorem 2.13 implies

apka(modp).a^{p^{k}}\equiv a\pmod{p}. (16)

for k1k\geq 1 since by Theorem 2.13, apkapk1(modp)a^{p^{k}}\equiv a^{p^{k-1}}\pmod{p}.

Theorem 2.14.

For field FF of characteristic pp and a,bFa,b\in F,

Proof.

By the binomial theorem,

(a+b)p=ap+(p1)ap1b+(p2)ap2b2++(pp1)abp1+bp.(a+b)^{p}=a^{p}+\binom{p}{1}a^{p-1}b+\binom{p}{2}a^{p-2}b^{2}+...+\binom{p}{p-1}ab^{p-1}+b^{p}. (17)

We just need the following Lemma: The prime pp divides each binomial coefficient (pr)\binom{p}{r} for 1rp1.1\leq r\leq p-1. The proof of lemma is simple by definition:

(pr)=p(p1)(pr+1)12r.\binom{p}{r}=\frac{p\cdot(p-1)\cdot...\cdot(p-r+1)}{1\cdot 2\cdot...\cdot r}. (18)

The result follows since pp divides the top but not the bottom. ∎

2.2 Properties of zz-permanent

Let 𝕋:={z:|z|=1}\mathbb{T}:=\{z\in\mathbb{C}:|z|=1\} be the unit circle in the complex plane. For matrix Xn×nX\in\mathbb{C}^{n\times n}, the zz-permanent of XX is defined as

Perz(X):=σSnz(σ)iXi,σ(i),\displaystyle\mathrm{Per}_{z}(X):=\sum_{\sigma\in S_{n}}z^{\ell(\sigma)}\prod_{i}X_{i,\sigma(i)}, (19)

where SnS_{n} is the symmetric group on nn elements and (σ)\ell(\sigma) is the inversion number of σ\sigma. The inversion number is defined as the number of pairs (i,j)(i,j) such that i<ji<j and σ(i)>σ(j)\sigma(i)>\sigma(j). It is clear that (σ)(n2)=12n(n1)\ell(\sigma)\leq\binom{n}{2}=\frac{1}{2}n(n-1). This implies that we can write a zz-permanent as a polynomial in zz:

Perz(X)==0(n2)z(σ:(σ)=i=1nXi,σ(i)).\displaystyle\mathrm{Per}_{z}(X)=\sum_{\ell=0}^{\binom{n}{2}}z^{\ell}\mathopen{}\mathclose{{}\left(\sum_{\sigma:\ell(\sigma)=\ell}\prod_{i=1}^{n}X_{i,\sigma(i)}}\right). (20)

Note that by (20), zz-permanent is a polynomial with degree (n2){n\choose 2} in terms of z.z.

Here we list basic properties of zz-permanent [Yan91, DS] for every matrix Xn×nX\in\mathbb{C}^{n\times n}. Most of them are exact the same as permanent and determinant.

  1. 1.

    Perz(X)\mathrm{Per}_{z}(X) is a multilinear function of the rows and columns.

  2. 2.

    For block triangular XX, Perz(X)\mathrm{Per}_{z}(X) is the product of the zz-permanent of its diagonal blocks.

  3. 3.

    Perz(X)=Perz(XT)\mathrm{Per}_{z}(X)=\mathrm{Per}_{z}(X^{T}) where XTX^{T} is the transpose of XX.

  4. 4.

    (Expansion Theorem) Let XijX^{ij} denote the (i,j)(i,j) minor of XX and XijX_{ij} denote the element of X.X. Then

    Perz(X)=k=1nX1kzk1Perz(X1k)=k=1nXjkzk1Perz(Xjk)\mathrm{Per}_{z}(X)=\sum_{k=1}^{n}X_{1k}z^{k-1}\mathrm{Per}_{z}(X^{1k})=\sum_{k=1}^{n}X_{jk}z^{k-1}\mathrm{Per}_{z}(X^{jk}) (21)

    for j=1,2,,nj=1,2,\ldots,n.

  5. 5.

    For integer matrix XX and nn-th root of unity zz, Perz(X)[ζn]\mathrm{Per}_{z}(X)\in\mathbb{Z}[\zeta_{n}]. Thus, the minimal polynomial of Perz(X)\mathrm{Per}_{z}(X) over \mathbb{Q} has leading coefficient 11 and its degree is at most ϕ(n)\phi(n).

We give the explicit formula for the simplest case n=2n=2 and n=3n=3: For n=2n=2,

Perz(X)=(X11X22+X21X12z).\displaystyle\mathrm{Per}_{z}(X)=(X_{11}X_{22}+X_{21}X_{12}z).

For n=3n=3,

Perz(X)=X13X22X31z3+(X12X23X31+X13X21X32)z2+\displaystyle\mathrm{Per}_{z}(X)=X_{13}X_{22}X_{31}z^{3}+\mathopen{}\mathclose{{}\left(X_{12}X_{23}X_{31}+X_{13}X_{21}X_{32}}\right)z^{2}+
(X11X23X32+X12X21X33)z+X11X22X33.\displaystyle\mathopen{}\mathclose{{}\left(X_{11}X_{23}X_{32}+X_{12}X_{21}X_{33}}\right)z+X_{11}X_{22}X_{33}.

2.3 Common Tasks Related to Permanent

We recall the following task defined by Aaronson and Arkhipov [AA11]. We will discuss some of their generalizations in our paper.

Problem 1 (GPE×\mathrm{GPE}_{\times}).

Given as input X𝒩(0,1)n×nX\sim\mathcal{N}_{\mathbb{C}}(0,1)^{n\times n} of i.i.d Gaussians (Complex normal distribution) and ϵ,δ>0\epsilon,\delta>0, estimate Per(X)\mathrm{Per}(X) to within error ±ϵ|Per(X)|\pm\epsilon|\mathrm{Per}(X)| to probability at least 1δ1-\delta over XX.

Conjecture 1.

GPE×\mathrm{GPE}_{\times} is #𝖯{\#\mathsf{P}}-hard. In other words, if 𝒪\mathcal{O} is any oracle that solves GPE×\mathrm{GPE}_{\times}, then 𝖯#𝖯𝖡𝖯𝖯𝒪\mathsf{P}^{\#\mathsf{P}}\subseteq\mathsf{BPP}^{\mathcal{O}}.

Conjecture 2.

There exists a polynomial pp such that for positive integer nn, real number δ>0\delta>0,

PrX[|Per(X)|2n!p(n,1/δ)]1δ.\displaystyle\Pr_{X}\mathopen{}\mathclose{{}\left[|\mathrm{Per}(X)|^{2}\geq\frac{\sqrt{n!}}{p(n,1/\delta)}}\right]\geq 1-\delta. (22)

3 Hardness of zz-permanent for {0,1}\{0,1\}-matrices

In this section, we investigate the hardness of zz-permanent. We only consider the case that zz is root of a unity, i.e., there is positive integer mm such that zm=1z^{m}=1, for the following two reasons. First, the set {z=e2πia/b:a,b}\{z=e^{2\pi ia/b}:a,b\in\mathbb{N}\} is dense on the unit circle, so for every z𝕋z\in\mathbb{T}, there exists an arbitrarily close element in this set. Second, from the computational point of view, as we will see, it is more relevant and convenient to give a definition of exact computation with a rational multiple of 2π2\pi. Thus, in this paper, we always assume that zz is root of a unity, and write z=ζmz=\zeta_{m} for primitive mm-th root for unity for integer m2m\geq 2.

We start with the cubic root of unity, showing that computing zz-permanant for zz satisfying z3=1z^{3}=1 is 𝖬𝗈𝖽3𝖯\mathsf{Mod}_{3}\mathsf{P}-hard. Thus the existence of an algorithm for computing the zz-permanent implies the collapse of the polynomial hierarchy by Toda’s theorem.

3.1 Cube Roots of Unity

Recall that from algebraic number theory, for z=ζpz=\zeta_{p} and binary matrix X{0,1}n×nX\in\{0,1\}^{n\times n}, Perz(X)\mathrm{Per}_{z}(X) can be written as a polynomial in zz:

Perz(X)=kAkzk,\displaystyle\mathrm{Per}_{z}(X)=\sum_{k}A_{k}z^{k}, (23)

where Ak:=σ:(σ)k(modp)i=1nXi,σ(i)A_{k}:=\sum_{\sigma:\ell(\sigma)\equiv k\pmod{p}}\prod_{i=1}^{n}X_{i,\sigma(i)}. Furthermore, for z=ζpz=\zeta_{p}, Perz(X)[ζp]\mathrm{Per}_{z}(X)\in\mathbb{Z}[\zeta_{p}], and thus is an algebraic integer in the number field [ζp]\mathbb{Q}[\zeta_{p}]. We will link the observation to hardness of computing 11-permanent and the complexity 𝖬𝗈𝖽p𝖯\mathsf{Mod}_{p}\mathsf{P}, defined as follows.

Definition 3.1 (𝖬𝗈𝖽k𝖯\mathsf{Mod}_{k}\mathsf{P}).

For any integer k2k\geq 2, let 𝖬𝗈𝖽k𝖯\mathsf{Mod}_{k}\mathsf{P} be the class of decision problems solvable by a polynomial time non-deterministic machine which rejects if the number of accepting paths is divisible by kk, and accepts otherwise. When k=2k=2, 𝖬𝗈𝖽k𝖯\mathsf{Mod}_{k}\mathsf{P} is also known as parity 𝖯\mathsf{P}, and denoted 𝖯\oplus\mathsf{P}.

It is well known due to Valiant that computing Per(X)modp\mathrm{Per}(X)\bmod p is 𝖬𝗈𝖽p𝖯\mathsf{Mod}_{p}\mathsf{P}-complete for prime p>2p>2 [Val79a]. Interestingly, we can reduce the problem of computing Per(X)modp\mathrm{Per}(X)\bmod p to the problem of computing Perζp(X)\mathrm{Per}_{\zeta_{p}}(X) in the special case p=3p=3.

Theorem 3.2 (Valiant [Val79a]).

The problem of computing Per(M)modp\mathrm{Per}(M)\bmod p for a square matrix M𝔽pn×nM\in\mathbb{F}_{p}^{n\times n} is 𝖬𝗈𝖽p𝖯\mathsf{Mod}_{p}\mathsf{P}-complete for any prime p>2p>2.

Now we state the theorem for this section.

Theorem 3.3.

For every cube roots of unity ζ\zeta, computing Perζ(X)\mathrm{Per}_{\zeta}(X) and |Perζ(X)|2|\mathrm{Per}_{\zeta}(X)|^{2} are both 𝖬𝗈𝖽3𝖯\mathsf{Mod}_{3}\mathsf{P}-hard.

Proof.

For ζ=1\zeta=1, the theorem holds due to Valiant [Val79a], and thus it suffices to prove the theorem for ζ{ei2π/3,ei4π/3}\zeta\in\{e^{i2\pi/3},e^{i4\pi/3}\}. For ζ=ei2π/3\zeta=e^{i2\pi/3}, we can write Perζ(X)\mathrm{Per}_{\zeta}(X) as

Perζ(X)\displaystyle\mathrm{Per}_{\zeta}(X) =A0+ζA1+ζ2A2\displaystyle=A_{0}+\zeta A_{1}+\zeta^{2}A_{2}
=A0+(12+3i2)A1+(123i2)A2\displaystyle=A_{0}+\mathopen{}\mathclose{{}\left(-\frac{1}{2}+\frac{\sqrt{3}i}{2}}\right)A_{1}+\mathopen{}\mathclose{{}\left(-\frac{1}{2}-\frac{\sqrt{3}i}{2}}\right)A_{2}
=2A0A1A22+i3A1A22\displaystyle=\frac{2A_{0}-A_{1}-A_{2}}{2}+i\sqrt{3}\frac{A_{1}-A_{2}}{2} (24)

If one can compute Perζ(X)\mathrm{Per}_{\zeta}(X), then one can compute Per(X)mod3\mathrm{Per}(X)\bmod 3 by taking the real part of Perζ(X)\mathrm{Per}_{\zeta}(X) since

2A0A1A22Per(X)(mod3),\displaystyle 2A_{0}-A_{1}-A_{2}\equiv 2\mathrm{Per}(X)\pmod{3}, (25)

and 22 is a unit in the ring 3\mathbb{Z}_{3}. The same reasoning holds for ζ=ei4π/3\zeta=e^{i4\pi/3} by interchanging ii and i-i.

It was shown in Valiant’s original paper [Val79b] that the permanent modulo m=2km=2^{k} can be computed in time nO(k)n^{O(k)}, but for all m2km\neq 2^{k}, it is 𝖭𝖯\mathsf{NP}-hard under randomized reductions [CX15, DJ21].

For |Perζ(X)|2|\mathrm{Per}_{\zeta}(X)|^{2}, we observe that

|Perζ(X)|2\displaystyle|\mathrm{Per}_{\zeta}(X)|^{2} |Per(X)|2(mod3).\displaystyle\equiv|\mathrm{Per}(X)|^{2}\pmod{3}.

Thus we can obtain |Per(X)|2mod3|\mathrm{Per}(X)|^{2}\bmod 3 if |Perζ(X)|2|\mathrm{Per}_{\zeta}(X)|^{2} can be exactly computed. ∎

The generalized Toda’s theorem [TO92] states that a few counting classes are as hard as the polynomial hierarchy.

Theorem 3.4 (Toda theorem [Tod91, TO92]).

For prime k>2k>2, let AA be one of the counting classes 𝖬𝗈𝖽k𝖯,#𝖯,𝖯𝖯,\mathsf{Mod}_{k}\mathsf{P},\mathsf{\#P},\mathsf{PP}, or 𝖦𝖺𝗉𝖯\mathsf{GapP}. Then 𝖯𝖧𝖡𝖯𝖯A\mathsf{PH}\subseteq\mathsf{BPP}^{A}.

By Toda’s theorem, we immediately have the following theorem.

Theorem 3.5.

For ζ{1,ei2π/3,ei4π/3}\zeta\in\{1,e^{i2\pi/3},e^{i4\pi/3}\}, let 𝒪\mathcal{O} be any oracle oracle that on input binary matrix X{0,1}n×nX\in\{0,1\}^{n\times n}, outputs Perζ(X)\mathrm{Per}_{\zeta}(X). Then

𝖯𝖧𝖡𝖯𝖯𝖬𝗈𝖽3𝖯𝖡𝖯𝖯𝒪.\mathsf{PH}\subseteq\mathsf{BPP}^{\mathsf{Mod}_{3}\mathsf{P}}\subseteq\mathsf{BPP}^{\mathcal{O}}. (26)

In other words, if there is an efficient classical simulation of 𝒪\mathcal{O}, then the polynomial hierarchy collapses to the second level.

If the polynomial hierarchy is infinite, Theorem 3.5 excludes the existence of an efficient classical algorithm for exactly computing ζ\zeta-permanent.

3.2 Primitive mmth roots of unity for prime power mm

Next, we turn our attention to any prime number p=nO(1)p=n^{O(1)}. Since in general Perζp(X)\mathrm{Per}_{\zeta_{p}}(X) can be an irrational number in the complex plane, any truncation in finite precision only results in an approximation. In this section, we focus the hardness of computing Perζp(X)\mathrm{Per}_{\zeta_{p}}(X) exactly, and thus the first question we must clarify is how we define “computing Perζp(X)\mathrm{Per}_{\zeta_{p}}(X) exactly” for binary XX.

For binary matrix XX, AkA_{k}\in\mathbb{Z} (see (23)). Thus Perζp(X)\mathrm{Per}_{\zeta_{p}}(X) can be written as a polynomial in ζp\zeta_{p} of degree at most p1p-1 with integer coefficients. We can define an exact representation of Perζp(X)\mathrm{Per}_{\zeta_{p}}(X) as a set of coefficients.

Definition 3.6 (Representation of an algebraic integer).

A representation of algebraic integer α[ζp]\alpha\in\mathbb{Z}[\zeta_{p}] is a tuple of integers (b0,b1,,bp1)(b_{0},b_{1},\ldots,b_{p-1}) satisfying

α=b0+b1ζp+bp1ζpp1.\displaystyle\alpha=b_{0}+b_{1}\zeta_{p}+\ldots b_{p-1}\zeta_{p}^{p-1}. (27)

Note that the representation is not unique since ζp\zeta_{p} is a root of f(x)=1+x++xp1f(x)=1+x+\ldots+x^{p-1}, and thus α+cf(ζp)=α\alpha+cf(\zeta_{p})=\alpha for any integer cc\in\mathbb{Z}.

In general, one can define a representation of algebraic integers α[ζn]\alpha\in\mathbb{Z}[\zeta_{n}] for general nn in Definition 4.5. However, in this section, the Definition 3.6 is sufficient for showing the hardness. Notice that one can also define such representation by a tuple of integers (b0,b1,,bp2)(b_{0},b_{1},\ldots,b_{p-2}) satisfying α=b0+b1ζp+bp2ζpp2\alpha=b_{0}+b_{1}\zeta_{p}+\ldots b_{p-2}\zeta_{p}^{p-2} since one has ζpp1=1ζpζp2ζpp2.\zeta_{p}^{p-1}=-1-\zeta_{p}-\zeta_{p}^{2}-...-\zeta_{p}^{p-2}.

An algorithm which computes Perζp(X)\mathrm{Per}_{\zeta_{p}}(X) exactly outputs a tuple of coefficients (b0,,bp1)p(b_{0},\ldots,b_{p-1})\in\mathbb{Z}^{p} such that (27) holds. This definition is without loss of generality for our purpose. In Section 4.2, we will construct an algorithm that maps an approximation in the complex plane to an exact representation using an 𝖭𝖯\mathsf{NP} oracle.

Theorem 3.7.

For prime pp, let ζp\zeta_{p} be a primitive pp-th root of unity and 𝒪\mathcal{O} be any oracle that on input a binary matrix X{0,1}n×nX\in\{0,1\}^{n\times n}, outputs a representation of Perζp(X)\mathrm{Per}_{\zeta_{p}}(X). Then 𝖡𝖯𝖯𝖬𝗈𝖽p𝖯𝖡𝖯𝖯𝒪\mathsf{BPP}^{\mathsf{Mod}_{p}\mathsf{P}}\subseteq\mathsf{BPP}^{\mathcal{O}}.

Proof.

For a binary matrix XX, let the output of 𝒪(X)\mathcal{O}(X) be (a0,,ap1)(a_{0},\ldots,a_{p-1}). Also, let AiA_{i} be defined as in (23). By the correctness of 𝒪\mathcal{O},

Perζp(X)=i=0p1Aiζpi=i=0p1aiζpi.\mathrm{Per}_{\zeta_{p}}(X)=\sum_{i=0}^{p-1}A_{i}\zeta_{p}^{i}=\sum_{i=0}^{p-1}a_{i}\zeta_{p}^{i}. (28)

Note that it does not necessarily hold that for every ii, Ai=aiA_{i}=a_{i}, but the polynomials are identical congruent to the minimal polynomial of ζp\zeta_{p}. In particular, for prime pp, since 0=i(Aiai)ζpi0=\sum_{i}(A_{i}-a_{i})\zeta_{p}^{i},

(i=0p1(Aiai)ζpi)p=i=0p1(Aiai)p+ph=0\mathopen{}\mathclose{{}\left(\sum_{i=0}^{p-1}(A_{i}-a_{i})\zeta_{p}^{i}}\right)^{p}=\sum_{i=0}^{p-1}(A_{i}-a_{i})^{p}+ph=0 (29)

where h[ω]h\in\mathbb{Z}[\omega]. Since the left side of the equality

i=0p1(Aiai)p=ph\displaystyle\sum_{i=0}^{p-1}(A_{i}-a_{i})^{p}=-ph (30)

is an integer, h[ζp]h\in\mathbb{Z}[\zeta_{p}]\cap\mathbb{Q}\subseteq\mathbb{Z}. This means that pp divides i(Aiai)p\sum_{i}(A_{i}-a_{i})^{p}, and

0\displaystyle 0 i=0p1(Aiai)p\displaystyle\equiv\sum_{i=0}^{p-1}(A_{i}-a_{i})^{p}
i=0p1Aipi=0p1aip\displaystyle\equiv\sum_{i=0}^{p-1}A_{i}^{p}-\sum_{i=0}^{p-1}a_{i}^{p}
i=0p1Aiai(modp).\displaystyle\equiv\sum_{i=0}^{p-1}A_{i}-a_{i}\pmod{p}. (31)

The last equality holds from Fermat’s little theorem (Theorem 2.13).

In summary, given access to an oracle 𝒪\mathcal{O} which exactly computes Perζp(X)\mathrm{Per}_{\zeta_{p}}(X) for binary X{0,1}n×nX\in\{0,1\}^{n\times n}, the following algorithm computes Per(X)modp\mathrm{Per}(X)\bmod p:

  • Run (a0,,ap1)𝒪(X)(a_{0},\ldots,a_{p-1})\leftarrow\mathcal{O}(X).

  • Output i=0p1aimodp\sum_{i=0}^{p-1}a_{i}\bmod p.

By Toda’s theorem Theorem 3.4, we have the following corollary.

Corollary 3.8.

For prime p>3p>3, let 𝒪\mathcal{O} denote any oracle which computes Perζp(X)\mathrm{Per}_{\zeta_{p}}(X) exactly for binary matrix X{0,1}n×nX\in\{0,1\}^{n\times n}. Then

𝖯𝖧𝖡𝖯𝖯𝖬𝗈𝖽p𝖯𝖡𝖯𝖯𝒪.\mathsf{PH}\subseteq\mathsf{BPP}^{\mathsf{Mod}_{p}\mathsf{P}}\subseteq\mathsf{BPP}^{\mathcal{O}}. (32)

We can easily extend our result to a prime power q=pkq=p^{k} for prime pp.

Theorem 3.9.

For prime pp and prime power q=pkq=p^{k}, let ζq\zeta_{q} be a primitive qq-th root of unity and 𝒪\mathcal{O} be any oracle that on input a binary matrix X{0,1}n×nX\in\{0,1\}^{n\times n}, outputs an exact representation of Perζq(X)\mathrm{Per}_{\zeta_{q}}(X). Then 𝖡𝖯𝖯𝖬𝗈𝖽p𝖯𝖡𝖯𝖯𝒪\mathsf{BPP}^{\mathsf{Mod}_{p}\mathsf{P}}\subseteq\mathsf{BPP}^{\mathcal{O}}.

Proof.

The idea basically follows from the proof of Theorem 3.9. For ζq=e2πia/q\zeta_{q}=e^{2\pi ia/q} and gcd(a,p)=1\gcd(a,p)=1, we have

(i=0q1(Aiai)ζqi)q=i=0q1(Aiai)q+ph=0\mathopen{}\mathclose{{}\left(\sum_{i=0}^{q-1}(A_{i}-a_{i})\zeta_{q}^{i}}\right)^{q}=\sum_{i=0}^{q-1}(A_{i}-a_{i})^{q}+ph=0 (33)

where h[ω]h\in\mathbb{Z}[\omega]\cap\mathbb{Q}\subseteq\mathbb{Z}. We then use Fermat’s little theorem (Theorem 2.13) which imples that aqa(modp)a^{q}\equiv a(\bmod p), and

i=0q1(Aiai)+ph=0.\sum_{i=0}^{q-1}(A_{i}-a_{i})+ph=0. (34)

So then we have (using finite field property Theorem 2.14)

Per(X)i=0q1Aii=0q1ai(modp),\mathrm{Per}(X)\equiv\sum_{i=0}^{q-1}A_{i}\equiv\sum_{i=0}^{q-1}a_{i}\pmod{p}, (35)

which finish the proof. ∎

We end this section by summarizing the classical hardness of Perz(X)\mathrm{Per}_{z}(X) for such particular zz.

Theorem 3.10.

For odd prime pp and prime power q=pkq=p^{k}, let ζq\zeta_{q} be a primitive qq-th root of unity. Exactly computing ζq\zeta_{q}-permanent for binary matrices is not in 𝖯𝖧\mathsf{PH}; otherwise, the polynomial hierarchy collapses. More precisely, for any oracles 𝒪\mathcal{O} which computes the ζq\zeta_{q}-permanent of binary matrices, the following containments hold:

𝖯𝖧𝖡𝖯𝖯𝖬𝗈𝖽p𝖯𝖡𝖯𝖯𝒪.\displaystyle\mathsf{PH}\subseteq\mathsf{BPP}^{\mathsf{Mod}_{p}\mathsf{P}}\subseteq\mathsf{BPP}^{\mathcal{O}}.

Unfortunately, our method does not generalize to a primitive mm-th root of unity for mm divisible by more than one prime. While our result does not rule out the existence of an efficient solver for non-prime powers, it is interesting to see whether there exists a polynomial-time algorithm for any of these points. Since it is not so relevant to our result, we leave the complexity of computing general composite numbers as future work.

For p=2p=2, we notice that our method with oracle which computes ζ2k\zeta_{2^{k}}-permanent implies one can compute Per(X)(mod2)\mathrm{Per}(X)\pmod{2}. However, since Per(X)det(X)(mod2),\mathrm{Per}(X)\equiv\det(X)(\bmod 2), we do not have the hardness results for computing ζ2k\zeta_{2^{k}}-permanet Indeed, it is known that permanent modulo 2k2^{k} admits an efficient algorithm [DJ21, CX15]. Any hardness results for ζ2k\zeta_{2^{k}}-permanent cannot be implied by the hardness of permanent modulo 2k2^{k}.

Unfortunately, our method cannot be used to establish hardness for rational numbers with a non-prime-power denominator.

3.3 Hardness of Rational Phases

Recall that for every z𝕋z\in\mathbb{T}, the zz-permanent of a binary matrix is always an integral polynomial in zz of degree at most (n2)\binom{n}{2}. For a primitive mm-th root of unity z=ζmz=\zeta_{m}, we can further reduce the degree by taking the congruence to the minimal polynomial of the degree of ζm\zeta_{m}, when ϕ(m)<(n2)\phi(m)<\binom{n}{2}. In the previous sections, we have applied these facts to reduce the problem of computing 11-permanent modulo mm to ζm\zeta_{m}-permanent for prime power mm.

While we do not have a general result for every non-prime-power rational phases, in this section, we consider a special case where ϕ(m)(n2)\phi(m)\geq\binom{n}{2}. In this case, since the monomials 1,ζm,ζm2,,ζm(n2)1,\zeta_{m},\zeta_{m}^{2},\ldots,\zeta_{m}^{\binom{n}{2}} are linearly indepenent, 11-permanent can be directly obtained from ζm\zeta_{m}-permanent.

Theorem 3.11.

Let 𝒪\mathcal{O} denote any oracle which input are binary matrix XX and m:m:\mathbb{N}\to\mathbb{N} such that ϕ(m(n))(n2)+1\phi(m(n))\geq\binom{n}{2}+1, it outputs Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) exactly. Then 𝖯#𝖯𝖯𝒪\mathsf{P}^{\#\mathsf{P}}\subseteq\mathsf{P}^{\mathcal{O}}.

Proof.

Let d:=(n2)d:=\binom{n}{2}. Computing 𝒪(m,X)\mathcal{O}(m,X) yields a tuple of coefficients (a0,,ad)(a_{0},\ldots,a_{d}) satisfying

Perζm(X)=i=0dAiζmi=i=0daiζmi.\mathrm{Per}_{\zeta_{m}}(X)=\sum_{i=0}^{d}A_{i}\zeta_{m}^{i}=\sum_{i=0}^{d}a_{i}\zeta_{m}^{i}. (36)

where Ai=σ𝕊N,(σ)=i(j=1nXj,σ(j))A_{i}=\sum_{\sigma\in\mathbb{S}_{N},\ell(\sigma)=i}\mathopen{}\mathclose{{}\left(\prod_{j=1}^{n}X_{j,\sigma(j)}}\right). Since the monimials 1,ζm,,ζmd1,\zeta_{m},\ldots,\zeta_{m}^{d} are linearly dependent for ϕ(m)d+1\phi(m)\geq d+1, Ai=aiA_{i}=a_{i} for every 0id0\leq i\leq d. This implies that i=0dai=Perm(X)\sum_{i=0}^{d}a_{i}=\mathrm{Per}_{m}(X). ∎

However, for cases with non-prime-power m=O(1)m=O(1), neither of the methods seems to work. We leave the hardness of primitive mm-th roots of unity for fix non-prime-power mm as an open question.

4 Hardness of Approximating ζm\zeta_{m}-permanent

In Section 3, we have shown that for prime power m=pkm=p^{k} for odd prime pp, exactly computing Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) is as hard as computing Per(X)modp\mathrm{Per}(X)\bmod p. The task is known to be 𝖬𝗈𝖽p𝖯\mathsf{Mod}_{p}\mathsf{P}-hard. In this section, we show that approximating Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) for integer mm and binary matrices is as hard as exactly computing it.

In Section 4.1, we show that approximating |Perζm(X)|2|\mathrm{Per}_{\zeta_{m}}(X)|^{2} to multiplicative error within g=poly(n)g=\operatorname{poly}(n) is as hard as approximating Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) to multiplicative error 2poly(n)2^{-\operatorname{poly}(n)}, up to a polynomial multiplicative overhead. In Section 4.2, we show that with an 𝖭𝖯\mathsf{NP} oracle, exactly computing Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) is as hard as approximating it to error 2poly(n)2^{-\operatorname{poly}(n)}. Combining the results in these two sections, we conclude that an efficient approximation algorithm would also imply a collapse of the polynomial hierarchy.

4.1 Reducing the Error

In this section, we show that for a prime number mm and an mm-th primitive root of unity ζm\zeta_{m}, approximating |Perζm(X)|2|\mathrm{Per}_{\zeta_{m}}(X)|^{2} to multiplicative error g=nO(1)g=n^{O(1)} is as hard as approximating Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) to multiplicative error 1+2nO(1)1+2^{-n^{O(1)}}.

First we prove the following lemma which will be useful later.

Lemma 4.1.

For integer n1n\geq 1 and δ12n\delta\leq\frac{1}{2n}, let α1,,αn\alpha_{1},\ldots,\alpha_{n} be complex numbers and α~1,,α~n\tilde{\alpha}_{1},\ldots,\tilde{\alpha}_{n} satisfying for i[n]i\in[n],

|αiα~i|δ|αi|.\displaystyle|\alpha_{i}-\tilde{\alpha}_{i}|\leq\delta|\alpha_{i}|. (37)

Then it holds that

|α1αnα~1α~n|2nδ|α1αn|.\displaystyle|\alpha_{1}\ldots\alpha_{n}-\tilde{\alpha}_{1}\ldots\tilde{\alpha}_{n}|\leq 2n\delta|\alpha_{1}\ldots\alpha_{n}|. (38)
Proof.

We prove the lemma by induction. For n=1n=1, the lemma obviously hold. Suppose that the lemma holds for n=kn=k. Then for n=k+1n=k+1, let A:=α1αkA:=\alpha_{1}\ldots\alpha_{k} and A~:=α~1α~k\tilde{A}:=\tilde{\alpha}_{1}\ldots\tilde{\alpha}_{k}. By the inductive hypothesis, |A~A|2kδ|A||\tilde{A}-A|\leq 2k\delta|A|. Then

|αk+1Aα~k+1A~||αk+1A|\displaystyle\frac{|\alpha_{k+1}A-\tilde{\alpha}_{k+1}\tilde{A}|}{|\alpha_{k+1}A|} |αk+1α~k+1||αk+1|+|α~k+1||αk||AA~||A|\displaystyle\leq\frac{|\alpha_{k+1}-\tilde{\alpha}_{k+1}|}{|\alpha_{k+1}|}+\frac{|\tilde{\alpha}_{k+1}|}{|\alpha_{k}|}\frac{|A-\tilde{A}|}{|A|}
δ+(1+δ)2kδ\displaystyle\leq\delta+(1+\delta)2k\delta
(2k+1)δ+2kδ2\displaystyle\leq(2k+1)\delta+2k\delta^{2}
(2k+2)δ.\displaystyle\leq(2k+2)\delta. (39)

The first inequality holds by (37) and |α~k+1||αk+1α~k+1|+|αk+1|(1+δ)|αk+1||\tilde{\alpha}_{k+1}|\leq|\alpha_{k+1}-\tilde{\alpha}_{k+1}|+|\alpha_{k+1}|\leq(1+\delta)|\alpha_{k+1}|. The last inequality holds since 2kδ22nδ2δ2k\delta^{2}\leq 2n\delta^{2}\leq\delta. ∎

We can decrease the error of our approximation algorithm from a multiplicative factor g=nO(1)g=n^{O(1)} to 1+2nO(1)1+2^{-n^{O(1)}} using a polynomial overhead.

Theorem 4.2.

For g[1,poly(n)]g\in[1,\operatorname{poly}(n)], prime number mm, and a primitive mm-th root of unity ζm\zeta_{m}, if there is an algorithm 𝒪\mathcal{O} which on input X{0,1}n×nX\in\{0,1\}^{n\times n}, approximates |Perζm(X)|2|\mathrm{Per}_{\zeta_{m}}(X)|^{2} to within a multiplicative factor of gg, then there exists an algorithm which outputs an esatimate P~\tilde{P} of Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) satisfying

|Perζm(X)P~||Perζm(X)|ϵ\displaystyle\frac{|\mathrm{Per}_{\zeta_{m}}(X)-\tilde{P}|}{|\mathrm{Per}_{\zeta_{m}}(X)|}\leq\epsilon (40)

using O(mn2g4logn+ng4logg+ng4log(1/ϵ))O(mn^{2}g^{4}\log n+ng^{4}\log g+ng^{4}\log(1/\epsilon)) queries to 𝒪\mathcal{O}.

Proof.

Let z=ζmz=\zeta_{m}. We describe the algorithm 𝒜𝒪\mathcal{A}^{\mathcal{O}} as follows. First 𝒜𝒪\mathcal{A}^{\mathcal{O}} decides if |Perz(X)|2=0|\mathrm{Per}_{z}(X)|^{2}=0. If |Perz(X)|20|\mathrm{Per}_{z}(X)|^{2}\neq 0, then 𝒪\mathcal{O} returns a value

p[g1|Perz(X)|2,g|Perz(X)|2]\displaystyle p\in[g^{-1}|\mathrm{Per}_{z}(X)|^{2},g|\mathrm{Per}_{z}(X)|^{2}] (41)

for g>1g>1. This implies that p>0p>0. Otherwise, 𝒪(X)=0\mathcal{O}(X)=0. This implies that there is an algorithm which decides if Perz(X)=0\mathrm{Per}_{z}(X)=0. Furthermore, if p=0p=0, 𝒜\mathcal{A} outputs 0.

It suffices to consider the case |Perz(X)|20|\mathrm{Per}_{z}(X)|^{2}\neq 0. Since XX is binary-valued, we can write

|Perz(X)|2=a0+a1z++am1zm1,\displaystyle|\mathrm{Per}_{z}(X)|^{2}=a_{0}+a_{1}z+\ldots+a_{m-1}z^{m-1}, (42)

where each aia_{i} is a non-negative integer, provided that zz is a primitive mm-th root of unity.

The algorithm 𝒜𝒪\mathcal{A}^{\mathcal{O}} finds a sequence of matrices X0=X,X1,,Xn1X_{0}=X,X_{1},\ldots,X_{n-1} such that each Xi(ni)×(ni)X_{i}\in\mathbb{C}^{(n-i)\times(n-i)} and Xi+1X_{i+1} is a submatrix of XiX_{i} obtained by deleting the first column and one of the rows kk determined as follows: starting from k=1k=1, repeat incrementing kk until the submatrix WW obtained by deleting the kk-th row and the first column satisfies that |Perz(W)|20|\mathrm{Per}_{z}(W)|^{2}\neq 0. Finally set Xi+1=WX_{i+1}=W. Since Perz(X)0\mathrm{Per}_{z}(X)\neq 0, it is guarantee that there exists WW such that |Perz(W)|0|\mathrm{Per}_{z}(W)|\neq 0. Computing the sequence of matrices takes O(n2)O(n^{2}) queries.

The algorithm inductively approximates Perz(Xk1)Perz(Xk)\frac{\mathrm{Per}_{z}(X_{k-1})}{\mathrm{Per}_{z}(X_{k})} to error O(ϵ/n)O(\epsilon/n) using O(mng4logn+g4logg+g4log(1/ϵ))O(mng^{4}\log n+g^{4}\log g+g^{4}\log(1/\epsilon)) queries to 𝒪\mathcal{O} from k=n1k=n-1 to 0 and the telescoping product

Perz(X)=Perz(X0)Perz(X1)Perz(X1)Perz(X2)Perz(Xn2)Perz(Xn1)Perz(Xn1).\displaystyle\mathrm{Per}_{z}(X)=\frac{\mathrm{Per}_{z}(X_{0})}{\mathrm{Per}_{z}(X_{1})}\cdot\frac{\mathrm{Per}_{z}(X_{1})}{\mathrm{Per}_{z}(X_{2})}\cdots\frac{\mathrm{Per}_{z}(X_{n-2})}{\mathrm{Per}_{z}(X_{n-1})}\cdot\mathrm{Per}_{z}(X_{n-1}). (43)

Computing the base case Perz(Xn1)\mathrm{Per}_{z}(X_{n-1}) is trivial. Applying Lemma 4.1, this implies that an approximation to error ϵ\epsilon takes O(mn2g4logn+g4nlogg)O(mn^{2}g^{4}\log n+g^{4}n\log g) queries.

In the following proof, we will only consider the case from Y:=X1Y:=X_{1} to X0X_{0}, and the same argument applies to k=2,,n1k=2,\ldots,n-1. For simplicity, we only prove the case where YY is obtained from XX by deleting the first row and the first column. Other cases (i.e., deleting the ii-th row of XX for i>1i>1) can be applied using the same argument.

We consider the quantity Perz(X)αPerz(Y)\mathrm{Per}_{z}(X)-\alpha\mathrm{Per}_{z}(Y). Note that for a primitive root of unity zz, both Perz(X)\mathrm{Per}_{z}(X) and Perz(Y)\mathrm{Per}_{z}(Y) are algebraic integers. This implies that the ratio α=Perz(X)Perz(Y)\alpha=\frac{\mathrm{Per}_{z}(X)}{\mathrm{Per}_{z}(Y)} is an algebraic number. To see there is a lower bound between two algebraic numbers, we apply a variant of the Liouville’s approximation theorem, stated in Theorem 2.12.

Since Perz(X)\mathrm{Per}_{z}(X) and Perz(Y)\mathrm{Per}_{z}(Y) are both algebraic integers, their minimal polynomials are monic. Furthermore, for Perz(X)=μ\mathrm{Per}_{z}(X)=\mu, max|σ(μ)|Per(X)n!\max|\sigma(\mu)|\leq\mathrm{Per}(X)\leq n!, and similarly the same upper bound holds for YY. Setting a=0a=0 and b=1b=1, |Perz(Y)|1/M|\mathrm{Per}_{z}(Y)|\geq 1/M and Perz(X)1/M\mathrm{Per}_{z}(X)\geq 1/M, where M:=(2n!+1)m2M:=(2n!+1)^{m-2}, and

1n!M|Perz(X)||Perz(Y)|n!M.\displaystyle\frac{1}{n!M}\leq\frac{|\mathrm{Per}_{z}(X)|}{|\mathrm{Per}_{z}(Y)|}\leq n!M. (44)

The quantity Perz(X)tPerz(Y)=Perz(X[t])\mathrm{Per}_{z}(X)-t\mathrm{Per}_{z}(Y)=\mathrm{Per}_{z}(X^{[t]}), where Xij[t]=Xijt𝟙[i=1]𝟙[j=1]X_{ij}^{[t]}=X_{ij}-t\cdot\mathbbm{1}[i=1]\mathbbm{1}[j=1], so its norm square can be approximated to multiplicative error gg using one call to 𝒪\mathcal{O}. To find an estimate of α=Perz(X)Perz(Y)\alpha=\frac{\mathrm{Per}_{z}(X)}{\mathrm{Per}_{z}(Y)}, we recursively find complex numbers α0,,α\alpha_{0},\ldots,\alpha_{\ell} such that the following invariant holds:

𝒪(X[αi+1])12𝒪(X[αi])\displaystyle\mathcal{O}(X^{[\alpha_{i+1}]})\leq\frac{1}{2}\mathcal{O}(X^{[\alpha_{i}]}) (45)

for every ii. To see how close αi\alpha_{i} is to α\alpha,

|ααi|\displaystyle|\alpha-\alpha_{i}| =|Perz(X)αiPerz(Y)||Perz(Y)|\displaystyle=\frac{|\mathrm{Per}_{z}(X)-\alpha_{i}\mathrm{Per}_{z}(Y)|}{|\mathrm{Per}_{z}(Y)|}
g𝒪(X[αi])|Perz(Y)|\displaystyle\leq\frac{\sqrt{g\mathcal{O}(X^{[\alpha_{i}]})}}{|\mathrm{Per}_{z}(Y)|}
g𝒪(X[αi])𝒪(Y)\displaystyle\leq g\sqrt{\frac{\mathcal{O}(X^{[\alpha_{i}]})}{\mathcal{O}(Y)}} (46)

Also let

βi:=g𝒪(X[αi])𝒪(Y).\displaystyle\beta_{i}:=g\sqrt{\frac{\mathcal{O}(X^{[\alpha_{i}]})}{\mathcal{O}(Y)}}. (47)

Setting the initial guess α0=0\alpha_{0}=0,

𝒪(X[α0])\displaystyle\mathcal{O}(X^{[\alpha_{0}]}) g|Perz(X)|2\displaystyle\leq g|\mathrm{Per}_{z}(X)|^{2}
g|Per(X)|2\displaystyle\leq g|\mathrm{Per}(X)|^{2}
=g(n!)2.\displaystyle=g(n!)^{2}. (48)

Recall that 𝒪(Y)g1|Perz(Y)|21gM2\mathcal{O}(Y)\geq g^{-1}|\mathrm{Per}_{z}(Y)|^{2}\geq\frac{1}{gM^{2}}. Then we have

|ααi|\displaystyle|\alpha-\alpha_{i}| g𝒪(X[αi])𝒪(Y)\displaystyle\leq g\sqrt{\frac{\mathcal{O}(X^{[\alpha_{i}]})}{\mathcal{O}(Y)}}
g2i𝒪(X[α0])𝒪(Y)\displaystyle\leq g\sqrt{\frac{2^{-i}\mathcal{O}(X^{[\alpha_{0}]})}{\mathcal{O}(Y)}}
g2ig(n!)2gM2\displaystyle\leq g\sqrt{2^{-i}\cdot g(n!)^{2}\cdot gM^{2}}
g2n!M2i/2.\displaystyle\leq g^{2}\cdot n!\cdot M\cdot 2^{-i/2}. (49)

Our goal is to get |αα|ϵ2nn!M|\alpha-\alpha_{\ell}|\leq\frac{\epsilon}{2n\cdot n!M} since it implies that |αα|ϵ|α|2n|\alpha-\alpha_{\ell}|\leq\frac{\epsilon|\alpha|}{2n}. Thus it suffices to set 2/2>g2(n!)2nM2(1/ϵ)2^{\ell/2}>g^{2}\cdot(n!)^{2}\cdot n\cdot M^{2}\cdot(1/\epsilon) and =O(mnlogn+logg+log(1/ϵ))\ell=O(mn\log n+\log g+\log(1/\epsilon)).

Now we give the algorithm that finds an αi+1\alpha_{i+1} that satisfies (45) from αi\alpha_{i}. By definition, |ααi|βi|\alpha-\alpha_{i}|\leq\beta_{i}. We consider the disk

Si:={t:|tαi|βi},\displaystyle S_{i}:=\{t:|t-\alpha_{i}|\leq\beta_{i}\}, (50)

and a grid of O(L2)O(L^{2}) points inside the disk for integer LL determined later:

Li:={αi+(j1)βiL+1(k1)βiLSi:|j|,|k|[L]}.\displaystyle L_{i}:=\mathopen{}\mathclose{{}\left\{\alpha_{i}+(j-1)\frac{\beta_{i}}{L}+\sqrt{-1}(k-1)\frac{\beta_{i}}{L}\in S_{i}:|j|,|k|\in[L]}\right\}. (51)

The algorithm computes 𝒪(X[t])\mathcal{O}(X^{[t]}) for every tLiSit\in L_{i}\cap S_{i} and outputs tt for which 𝒪(X[t])\mathcal{O}(X^{[t]}) is minimized (with ties broken arbitrarily).

By definition, αSi\alpha\in S_{i}, and thus there exists tLit\in L_{i} such that |tα|βi2L|t-\alpha|\leq\frac{\beta_{i}}{\sqrt{2}L}. For such tt,

𝒪(X[t])\displaystyle\mathcal{O}(X^{[t]}) g|Perz(X)tPerz(Y)|2\displaystyle\leq g|\mathrm{Per}_{z}(X)-t\mathrm{Per}_{z}(Y)|^{2}
g|αt|2|Perz(Y)|2\displaystyle\leq g|\alpha-t|^{2}|\mathrm{Per}_{z}(Y)|^{2}
gβi22L2|Perz(Y)|2\displaystyle\leq g\cdot\frac{\beta_{i}^{2}}{2L^{2}}\cdot|\mathrm{Per}_{z}(Y)|^{2}
gβi22L2g𝒪(Y)\displaystyle\leq g\cdot\frac{\beta_{i}^{2}}{2L^{2}}\cdot g\mathcal{O}(Y)
=g42L2𝒪(X[αi]).\displaystyle=\frac{g^{4}}{2L^{2}}\mathcal{O}(X^{[\alpha_{i}]}). (52)

Thus for L=g2L=g^{2}, the ratio 𝒪(X[t])𝒪(X[αi])\frac{\mathcal{O}(X^{[t]})}{\mathcal{O}(X^{[\alpha_{i}]})} is at most 1/21/2. ∎

As a corollary, we recover a result of Aaronson and Arkhipov for 11-permanent (with slightly worse query complexity) by Theorem 4.2, but the same theorem can apply to a primitive root of unity z𝕋z\in\mathbb{T}.

Corollary 4.3 (cf. [AA11, Theorem 28]).

For g[1,poly(n)]g\in[1,\operatorname{poly}(n)], if there is an algorithm 𝒪\mathcal{O} which on input X{0,1}n×nX\in\{0,1\}^{n\times n} approximates |Per(X)|2|\mathrm{Per}(X)|^{2} to within multiplicative factor of gg, then there is an algorithm which exactly computes Per(X)\mathrm{Per}(X) using O(mn2g4logn+ng4logg)O(mn^{2}g^{4}\log n+ng^{4}\log g) queries to 𝒪\mathcal{O}.

Proof.

Since Per(X)\mathrm{Per}(X) is an integer, we apply Theorem 4.2 with ϵ=1/n!\epsilon=1/n!. The algorithm takes O(mn2g4logn+ng4logg)O(mn^{2}g^{4}\log n+ng^{4}\log g) queries. ∎

Furthermore, obtaining an approximation to additive error within 2n22^{-n^{2}} takes poly(n)\operatorname{poly}(n) queries.

Corollary 4.4.

For g[1,poly(n)]g\in[1,\operatorname{poly}(n)], if there is an algorithm 𝒪\mathcal{O} which on input X{0,1}n×nX\in\{0,1\}^{n\times n}, approximates |Perζm(X)|2|\mathrm{Per}_{\zeta_{m}}(X)|^{2} to within a multiplicative factor of gg, then there exists an algorithm which outputs an esatimate P~\tilde{P} of Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) satisfying

|Perζm(X)P~|2n2\displaystyle|\mathrm{Per}_{\zeta_{m}}(X)-\tilde{P}|\leq 2^{-n^{2}} (53)

using O(mn2g4logn+ng4logn+g4n3)O(mn^{2}g^{4}\log n+ng^{4}\log n+g^{4}n^{3}) queries to 𝒪\mathcal{O}.

Proof.

To approximate to additive error within 2n22^{-n^{2}}, it suffices to approximate to multiplicative error ϵ=1n!2n2\epsilon=\frac{1}{n!2^{n^{2}}} since the |Perζm(X)||\mathrm{Per}_{\zeta_{m}}(X)| norm is at most n!n!, and then by the Theorem 4.2 and it takes O(mn2g4logn+ng4logn+ng4log(1ϵ))=O(mn2g4logn+ng4logn+g4n3)O(mn^{2}g^{4}\log n+ng^{4}\log n+ng^{4}\log(\frac{1}{\epsilon}))=O(mn^{2}g^{4}\log n+ng^{4}\log n+g^{4}n^{3}) queries. ∎

Here we see that the proof of approximate hardness relies on the fact that zz-permanent of given matrix XX can be computed from the zz-permanent of its submatrix (see equation (21)). Other matrix functions such as immanants do not have this property.

4.2 A Search-to-Decision Reduction

In the previous section, we have shown how to approximate Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) to multiplicative error 2nO(1)2^{-n^{O(1)}} for an mm-th primitive root of unity ζm\zeta_{m} using an oracle that approximates |Perζm(X)|2|\mathrm{Per}_{\zeta_{m}}(X)|^{2} to multiplicative error g=nO(1)g=n^{O(1)}. In this section, we show how to search for a representation of Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) from the approximation (see Definition 3.6).

Recall that by (27), for binary matrix XX, Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) is an element in the ring [ζm]\mathbb{Z}[\zeta_{m}] and can be written as a polynomial in ζm\zeta_{m}. Since Φm(ζm)=0\Phi_{m}(\zeta_{m})=0 and degΦm=ϕ(m)\deg\Phi_{m}=\phi(m), we can represent any algebraic integer in [ζm]\mathbb{Z}[\zeta_{m}] as a polynomial of degree at most ϕ(m)1\phi(m)-1. In fact, such a representation is unique. A more detailed background can be seen Definition 2.10.

Definition 4.5 (Representation of [ζm]\mathbb{Z}[\zeta_{m}]).

For integer m1m\geq 1, the representation of an algebraic integer α[ζm]\alpha\in\mathbb{Z}[\zeta_{m}] is a tuple (a0,,aϕ(m)1)(a_{0},\ldots,a_{\phi(m)-1}) such that

α=i=0ϕ(m)1aiζmi.\displaystyle\alpha=\sum_{i=0}^{\phi(m)-1}a_{i}\zeta_{m}^{i}. (54)

As a polynomial in ζm\zeta_{m}, the representation is the remainder of α\alpha divided by the cyclotomic polynomial Φm\Phi_{m}. A representation is said to be bounded in [A,B][A,B] for integers A,BA,B if for every 0iϕ(m)10\leq i\leq\phi(m)-1, AaiBA\leq a_{i}\leq B.

To extract a representation from an approximation, we rely on the following lemma which states that the coefficients aia_{i} can be represented using nO(1)n^{O(1)} bits.

Theorem 4.6 (Bounds on the coefficients of cyclotomic polynomials [Vau75]).

For integer m1m\geq 1, let Φm(x)\Phi_{m}(x) be the mm-th cyclotomic polynomial. The maximum absolute value of the coefficients of Φm(x)\Phi_{m}(x) is at most e12d(m)log(m)e^{\frac{1}{2}d(m)\log(m)}, where d(n)d(n) is the divisor function which count the number of divisors of mm (including 11 and mm).444For instance d(6)=4d(6)=4.

For m=p1a1,pkakm=p_{1}^{a_{1}}\ldots,p_{k}^{a_{k}}, d(m)=(a1+1)(ak+1)d(m)=(a_{1}+1)\ldots(a_{k}+1). Thus we can get the trivial bound d(m)md(m)\leq m. By Theorem 4.6, we can represent ϕm(x)\phi_{m}(x) using at most only O(mlogm)O(m\log m) bits. Then the following lemma shows that every element of the representation of Perζm(X)\mathrm{Per}_{\zeta_{m}}(X) can be represented using poly(n)\operatorname{poly}(n) bits.

Lemma 4.7 ([BP86, Lemma 2]).

Let s(x)=i=0msixi,t(x)=i=0ntixi,s(x)=\sum_{i=0}^{m}s_{i}x^{i},t(x)=\sum_{i=0}^{n}t_{i}x^{i}, be two polynomials of degrees mnm\geq n. Also let the quotient q(x)=i=0mnqixiq(x)=\sum_{i=0}^{m-n}q_{i}x^{i} and the remainder rr be such that s(x)=q(x)t(x)+r(x)s(x)=q(x)t(x)+r(x). Then

i=0mn|qi|(1+Ntn)mni=nm|si|tn,\sum_{i=0}^{m-n}|q_{i}|\leq\mathopen{}\mathclose{{}\left(1+\frac{N}{t_{n}}}\right)^{m-n}\sum_{i=n}^{m}\frac{|s_{i}|}{t_{n}}, (55)

where N=max(|tn|,|tn1|,,|t2nm|)N=\max(|t_{n}|,|t_{n-1}|,...,|t_{2n-m}|) and tg=0t_{g}=0 if g<0.g<0.

For our purpose, since each AiA_{i} satisfies

A=σ:(σ)=i=1nXi,σ(i)n!,\displaystyle A_{\ell}=\sum_{\sigma:\ell(\sigma)=\ell}\prod_{i=1}^{n}X_{i,\sigma(i)}\leq n!, (56)

by Lemma 4.7, we have the following corollary.

Corollary 4.8.

For integer m1m\geq 1 and binary XX, let (a0,,aϕ(m)1)(a_{0},\ldots,a_{\phi(m)-1}) be the representation of Perζm(X)\mathrm{Per}_{\zeta_{m}}(X). Then

max0iϕ(m)1log|ai|O(m2logm+nlogn).\displaystyle\max_{0\leq i\leq\phi(m)-1}\log|a_{i}|\leq O(m^{2}\log m+n\log n). (57)
Proof.

Let

s(x):==0m1Aζm.\displaystyle s(x):=\sum_{\ell=0}^{m-1}A_{\ell}\zeta_{m}^{\ell}. (58)

By Theorem 4.6, the coefficients of the cyclotomic polynomial is at most M=2O(mlogm)M=2^{O(m\log m)}. By Lemma 4.7 and the fact that Φm\Phi_{m} is monic, the summation of the coefficients of the quotient satisfies that

i=0m1ϕ(m)|qi|\displaystyle\sum_{i=0}^{m-1-\phi(m)}|q_{i}| (1+M)mϕ(m)1n!(mϕ(m)1)\displaystyle\leq\mathopen{}\mathclose{{}\left(1+M}\right)^{m-\phi(m)-1}\cdot n!\cdot(m-\phi(m)-1)
2O((mϕ(m))mlogm+nlogn).\displaystyle\leq 2^{O((m-\phi(m))\cdot m\log m+n\log n)}. (59)

The representation is the remainder of ss devided by Φm\Phi_{m}. Let r(x)=i=0ϕ(m)1aixi=s(x)q(x)Φm(x)r(x)=\sum_{i=0}^{\phi(m)-1}a_{i}x^{i}=s(x)-q(x)\Phi_{m}(x). We have

|ai|\displaystyle|a_{i}| |sij=0iqjtij|\displaystyle\leq\mathopen{}\mathclose{{}\left|s_{i}-\sum_{j=0}^{i}q_{j}t_{i-j}}\right|
maxi|si|+(ϕ(m)1)maxi|qi|maxi|ti|\displaystyle\leq\max_{i}|s_{i}|+(\phi(m)-1)\cdot\max_{i}|q_{i}|\cdot\max_{i}|t_{i}|
n!+(ϕ(m)1)2O((mϕ(m))mlogm+nlogn)M.\displaystyle\leq n!+(\phi(m)-1)\cdot 2^{O((m-\phi(m))\cdot m\log m+n\log n)}\cdot M. (60)

Thus

maxilog|ai|O((mϕ(m))mlogm+nlogn).\displaystyle\max_{i}\log|a_{i}|\leq O((m-\phi(m))\cdot m\log m+n\log n). (61)

Since ϕ(m)=Ω(mloglogm)\phi(m)=\Omega(\frac{m}{\log\log m}), mϕ(m)=Θ(m)m-\phi(m)=\Theta(m). ∎

Let AA denote the maximum of the coefficients in the representation of Perζm(X)\mathrm{Per}_{\zeta_{m}}(X). By Corollary 4.8, logA=O(m2logm+nlogn)\log A=O(m^{2}\log m+n\log n), and thus the coefficients can be represented using poly(n)\operatorname{poly}(n) bits, provided m=poly(n)m=\operatorname{poly}(n).

By Liouville’s approximation theorem (Theorem 2.12), the distance of two arbitrary algebraic integers whose representation is bounded in [A,A][-A,A] are at least δ=2nO(1)\delta=2^{-n^{O(1)}} in the complex plane.

Corollary 4.9.

For integer AA, let α,β\alpha,\beta be two algebraic integers in [ζm]\mathbb{Z}[\zeta_{m}] whose representation is bounded in [A,A][-A,A]. Then it holds that |αβ|1(2Aϕ(m)+1)ϕ(m)1|\alpha-\beta|\geq\frac{1}{(2A\cdot\phi(m)+1)^{\phi(m)-1}}.

Proof.

Since [ζm]\mathbb{Z}[\zeta_{m}] is an additive group, αβ[ζm]\alpha-\beta\in\mathbb{Z}[\zeta_{m}] whose minimal polynomial is monic and whose representation is bounded in [2A,2A][-2A,2A]. By Theorem 2.12 with (a,b)=(0,1)(a,b)=(0,1), d[(ζm):]=ϕ(m)d\leq[\mathbb{Q}(\zeta_{m}):\mathbb{Q}]=\phi(m), and |αβ¯|Ad|\overline{\alpha-\beta}|\leq Ad. These facts give

|αβ|1(2Aϕ(m)+1)ϕ(m)1.\displaystyle|\alpha-\beta|\geq\frac{1}{(2A\cdot\phi(m)+1)^{\phi(m)-1}}. (62)

Now let M:=(2Aϕ(m)+1)ϕ(m)1M:=(2A\phi(m)+1)^{\phi(m)-1}. If we already know a complex number α~\tilde{\alpha} that is δ\delta-close to an algebraic integer α\alpha for δ14M\delta\leq\frac{1}{4M}, then for every other algebraic integer βα\beta\neq\alpha whose representation is also bounded in [A,A][-A,A], |βα~|>34M|\beta-\tilde{\alpha}|>\frac{3}{4M}. Given the observation, we define the following problem.

Problem 2 (Close algebraic integer problem (caip)).

The close algebraic integer problem (caip)(\textsc{caip}) is defined as follows: given (α,L0,,Lϕ(m)1,T)(\alpha,L_{0},\ldots,L_{\phi(m)-1},T), where α\alpha\in\mathbb{C}, L0,,Lϕ(m)1,ML_{0},\ldots,L_{\phi(m)-1},M are non-negative integers (all represented using nO(1)n^{O(1)} bits), determine if

  • there is an algebraic integer αS\alpha^{\prime}\in S such that |αα|1/T|\alpha-\alpha^{\prime}|\leq 1/T, or

  • every αS\alpha^{\prime}\in S satisfies |αα|2/T|\alpha-\alpha^{\prime}|\geq 2/T,

where the set SS consists of algebraic integers in [ζm]\mathbb{Z}[\zeta_{m}] whose representation (a0,,aϕ(m)1)(a_{0},\ldots,a_{\phi(m)-1}) satisfies 0aiLi0\leq a_{i}\leq L_{i} for 0iϕ(m)10\leq i\leq\phi(m)-1. It is promised that the given instance is in one of the cases.

Indeed, caip can be solved using a non-deterministic Turing machine (NTM).

Theorem 4.10.

caip𝖭𝖯\textsc{caip}\in\mathsf{NP}.

Proof.

We can use sufficiently many bits (but still polynomial) to represent ζm\zeta_{m} as a complex number ξ\xi such that

|ζmξ|=o(1Tϕ(m)2maxiLi).\displaystyle|\zeta_{m}-\xi|=o\mathopen{}\mathclose{{}\left(\frac{1}{T\cdot\phi(m)^{2}\cdot\max_{i}L_{i}}}\right). (63)

For every β=b0+b1ζm++bϕ(m)1ζmϕ(m)1\beta=b_{0}+b_{1}\zeta_{m}+\ldots+b_{\phi(m)-1}\zeta_{m}^{\phi(m)-1} approximated by β~=b0+b1ξ++bϕ(m)1ξϕ(m)1\tilde{\beta}=b_{0}+b_{1}\xi+\ldots+b_{\phi(m)-1}\xi^{\phi(m)-1} satisfying |ββ~|=o(1/T)|\beta-\tilde{\beta}|=o(1/T) for every ii and integer bi[Li]b_{i}\in[L_{i}].

If the NTM non-deterministically finds b0,,bϕ(m)1b_{0},\ldots,b_{\phi(m)-1} satisfying |αβ~|1.5/T|\alpha-\tilde{\beta}|\leq 1.5/T, then |αβ|1/T|\alpha-\beta|\leq 1/T. Otherwise, every β\beta satisfies |αβ|2/T|\alpha-\beta|\geq 2/T. ∎

The problem caip only consider the case where each coefficient bi[0,Li]b_{i}\in[0,L_{i}]. To extend it to general cases where the ii-th element is in [Ui,Li][U_{i},L_{i}], we can shift α\alpha to β=αUiζmi\beta=\alpha-U_{i}\zeta_{m}^{i} and solves caip with β\beta and the ii-th paremeter being LiUiL_{i}-U_{i}. Thus it suffices to only consider the former setting.

Given an oracle for caip, we can search for a satisfying algebraic integer.

Theorem 4.11.

Let 𝒪\mathcal{O} be an oracle which solves caip. Then there exists an algorithm which makes O(log(L1Lϕ(m)1))O(\log(L_{1}\ldots L_{\phi(m)-1})) queries to 𝒪\mathcal{O} and finds a satisfying algebraic integer, or returns \bot if there is not.

Proof.

The algorithm 𝒜\mathcal{A} is basically a binary search for each coefficient.

First 𝒜\mathcal{A} runs b𝒪(α,L0,,Lϕ(m)1,T)b\leftarrow\mathcal{O}(\alpha,L_{0},\ldots,L_{\phi(m)-1},T); if the output b=0b=0, then 𝒜\mathcal{A} returns \bot. Otherwise, there must be a satisfying algebraic integer α\alpha^{\prime}. Then 𝒜\mathcal{A} runs the following steps:

  1. 1.

    Set βα\beta\leftarrow\alpha.

  2. 2.

    For j=ϕ(m)1,,0j=\phi(m)-1,\ldots,0,

    1. (a)

      Set W0W\leftarrow 0 and RLj/2R\leftarrow\lfloor L_{j}/2\rfloor.

    2. (b)

      Repeat the following steps until R=0R=0:

      1. i.

        Compute bi𝒪(β,L0,,Lj1,R,0,,0,T)b_{i}\leftarrow\mathcal{O}(\beta,L_{0},\ldots,L_{j-1},R,0,\ldots,0,T).

      2. ii.

        If b=0b=0, set ββ(R+1)ξj\beta\leftarrow\beta-(R+1)\xi^{j} and WW+R+1W\leftarrow W+R+1.

      3. iii.

        Set RR/2R\leftarrow\lfloor R/2\rfloor.

    3. (c)

      Set ajWa_{j}\leftarrow W.

  3. 3.

    Output (a0,,aϕ(m)1)(a_{0},\ldots,a_{\phi(m)-1}).

We prove the algorithm is correct for j=ϕ(m)1j=\phi(m)-1; for other coefficients, the correctness analysis applies similarly. First, the algorithm must terminate since in each iteration RR gets decreased by a factor of two. For the correctness, we note that β+Wzϕ(m)1=α\beta+Wz^{\phi(m)-1}=\alpha holds after each iteration. In each step, let a satisfying algebraic integer denote β\beta^{\prime}, written as

β=cϕ(m)1zϕ(m)1+aϕ(m)2zϕ(m)2++a0\displaystyle\beta^{\prime}=c_{\phi(m)-1}z^{\phi(m)-1}+a_{\phi(m)-2}z^{\phi(m)-2}+\ldots+a_{0} (64)

before Step 2(b)ii is run. It suffices to show that after Step 2(b)ii, the leading coefficient is no more than RR by induction. For the base case, since the leading coefficient of α\alpha^{\prime} is no more than Lϕ(m)1L_{\phi(m)-1}; the assertion holds. If b=0b=0, then R<cϕ(m)12R+1R<c_{\phi(m)-1}\leq 2R+1, and cϕ(m)1(R+1)0c_{\phi(m)-1}-(R+1)\geq 0; otherwise, there exists a satisfying β\beta^{\prime} such that cϕ(m)1Rc_{\phi(m)-1}\leq R. Thus after Step 2(b)ii, there is a satisfying algebraic integer whose leading coefficient is no more than RR. ∎

5 Average-case Hardness

In the context of average-case hardness, Aaronson and Arkhipov [AA11] considered the hardness of 11-permanent for i.i.d. Gaussian matrices. More recently, Haferkamp, Hangleiter, Eisert, and Gluza [HHEG20] considered the average-case hardness of the truncated uniform distribution.

In this section, we extend their results both to zz-permanent and to a wide range of distributions, provided that the distribution satisfies a property called the strong autocorrelation property.

Our idea follows from Aaronson and Arkhipov [AA11]: If one can compute a low-degree polynomial for sufficiently many points, then there is an algorithm that recovers the whole polynomial with high probability. More precisely, we use the following algorithm for noisy polynomial interpolation.

Theorem 5.1 (Berlekamp-Welch algorithm [BV]).

Let qq be a univariate polynomial of degree dd over any field 𝔽\mathbb{F}. Suppose we are given m pairs of 𝔽\mathbb{F} elements (x1,y1),,(xm,ym)(x_{1},y_{1}),...,(x_{m},y_{m}) (with the xisx_{i}^{\prime}s all distinct), and are promised that yi=q(xi)y_{i}=q(x_{i}) for more than (m+d)/2(m+d)/2 values of i. Then there is a deterministic algorithm to reconstruct qq using poly(n,d)\operatorname{poly}(n,d) field operations.

Now we define the following property, under which a random self-reducibility using the Berlekamp-Welch algorithm can be applied.

Definition 5.2 (Strong autocorrelation property).

Consider a random variable XX with a real distribution \mathcal{F} with mean 0 and variance 1.1. We denote it as X(0,1).X\sim\mathcal{F}_{\mathbb{R}}(0,1). We define (μ,σ2)\mathcal{F}_{\mathbb{R}}(\mu,\sigma^{2}) with mean μ\mu and variance σ2\sigma^{2} by shifting and scale the random variable Xμ+σ2X.X\to\mu+\sigma^{2}X. We pick ϵ<12.\epsilon<\frac{1}{2}. We can define complex random variable Z=X+iYZ=X+iY,X(μ1,1/2),Y(μ2,1/2)X\sim\mathcal{F}_{\mathbb{R}}(\mu_{1},1/2),Y\sim\mathcal{F}_{\mathbb{R}}(\mu_{2},1/2) and X,YX,Y are independent. We write Z(μ1+iμ2,1).Z\sim\mathcal{F}_{\mathbb{C}}(\mu_{1}+i\mu_{2},1). Now we consider the following three distributions

X\displaystyle X =(0,1)n\displaystyle=\mathcal{F}_{\mathbb{C}}(0,1)^{n} (65)
𝒟1\displaystyle\mathcal{D}_{1} =(0,(1ε)2)n,\displaystyle=\mathcal{F}_{\mathbb{C}}\mathopen{}\mathclose{{}\left(0,\mathopen{}\mathclose{{}\left(1-\varepsilon}\right)^{2}}\right)^{n}, (66)
𝒟2\displaystyle\mathcal{D}_{2} =i=1n(vi,1)\displaystyle=\prod_{i=1}^{n}\mathcal{F}_{\mathbb{C}}\mathopen{}\mathclose{{}\left(v_{i},1}\right) (67)

for some vector vnv\in\mathbb{C}^{n}. A distribution is said to satisfy the strong autocorrelation property Definition 5.2 if there exists constant M1,M2M_{1},M_{2} independent of n,ϵn,\epsilon, such that

𝒟1X\displaystyle\mathopen{}\mathclose{{}\left\|\mathcal{D}_{1}-X}\right\| 2nM1ε,\displaystyle\leq 2nM_{1}\varepsilon, (68)
𝒟2X\displaystyle\mathopen{}\mathclose{{}\left\|\mathcal{D}_{2}-X}\right\| 2nM2v2.\displaystyle\leq\sqrt{2n}M_{2}\mathopen{}\mathclose{{}\left\|v}\right\|_{2}\,. (69)

When \mathcal{F} is complex Gaussian or truncated uniform distribution both satisfy this criterion from direct calculation [AA11, HHEG20]. In Appendix B, we give a sufficient condition under which almost all well-behaved \mathcal{F} satisfies the such criterion.

Now we are ready to show that, as permanent, zz-permanent also has random self-reducibility.

Theorem 5.3 (Random self-reducibility of zz-permanent).

Assume that there is no classical efficient algorithm for computing zz-permanent and for all {0,1}\{0,1\}-matrices. For all δ1/poly(n)\delta\geq 1/\operatorname{poly}(n) the following problem is classically hard: Given an n×nn\times n matrix XX drawn from (0,1)n×n\mathcal{F}_{\mathbb{C}}(0,1)^{n\times n} which satisfies the strong autocorrelation property, output Perz(X)\mathrm{Per}_{z}\mathopen{}\mathclose{{}\left(X}\right) with probability at least 3/4+δ{3}/{4}+\delta over (0,1)n×n\mathcal{F}_{\mathbb{C}}(0,1)^{n\times n}.

Proof.

Let M=(xij){0,1}n×nM=\mathopen{}\mathclose{{}\left(x_{ij}}\right)\in\mathopen{}\mathclose{{}\left\{0,1}\right\}^{n\times n} be an arbitrary {0,1}\{0,1\}-matrix. We show how to compute Perz(M)\mathrm{Per}_{z}\mathopen{}\mathclose{{}\left(M}\right) in probabilistic polynomial time, given access to an oracle 𝒪\mathcal{O} such that

PrX(0,1)n×n[𝒪(X)=Perz(X)]34+δ.\Pr_{X\sim\mathcal{F}_{\mathbb{C}}(0,1)^{n\times n}}\mathopen{}\mathclose{{}\left[\mathcal{O}\mathopen{}\mathclose{{}\left(X}\right)=\mathrm{Per}_{z}\mathopen{}\mathclose{{}\left(X}\right)}\right]\geq\frac{3}{4}+\delta\,. (70)

We choose a matrix Yn×nY\in\mathbb{C}^{n\times n} from the distribution (0,1)n×n\mathcal{F}_{\mathbb{C}}(0,1)^{n\times n} and define

X(t):=(1t)Y+tM,q(t):=Perz(X(t))X\mathopen{}\mathclose{{}\left(t}\right):=\mathopen{}\mathclose{{}\left(1-t}\right)Y+tM,q\mathopen{}\mathclose{{}\left(t}\right):=\mathrm{Per}_{z}\mathopen{}\mathclose{{}\left(X\mathopen{}\mathclose{{}\left(t}\right)}\right) (71)

so that X(0)=YX\mathopen{}\mathclose{{}\left(0}\right)=Y and X(1)=MX\mathopen{}\mathclose{{}\left(1}\right)=M, and q(t)q\mathopen{}\mathclose{{}\left(t}\right) is a univariate polynomial in tt of degree at most nn and q(1)=Perz(X(1))=Perz(M)q\mathopen{}\mathclose{{}\left(1}\right)=\mathrm{Per}_{z}\mathopen{}\mathclose{{}\left(X\mathopen{}\mathclose{{}\left(1}\right)}\right)=\mathrm{Per}_{z}\mathopen{}\mathclose{{}\left(M}\right). Now let

L:=nδandε:=δ(4n2M1+4n2M2)L.L:=\mathopen{}\mathclose{{}\left\lceil\frac{n}{\delta}}\right\rceil\quad\text{and}\quad\varepsilon:=\frac{\delta}{\mathopen{}\mathclose{{}\left(4n^{2}M_{1}+4n^{2}M_{2}}\right)L}\,.

For each [L]\ell\in\mathopen{}\mathclose{{}\left[L}\right] call the oracle 𝒪\mathcal{O} on input matrix X(ε)X\mathopen{}\mathclose{{}\left(\varepsilon\ell}\right). Then, using the Berlekamp-Welch algorithm (Theorem 5.1), attempt to find a degree-nn polynomial q:q^{\prime}:\mathbb{C}\rightarrow\mathbb{C} such that

q(ε)=𝒪(X(ε))q^{\prime}\mathopen{}\mathclose{{}\left(\varepsilon\ell}\right)=\mathcal{O}\mathopen{}\mathclose{{}\left(X\mathopen{}\mathclose{{}\left(\varepsilon\ell}\right)}\right) (72)

for at least a 3/4+δ{3}/{4}+\delta fraction of [L]\ell\in\mathopen{}\mathclose{{}\left[L}\right].  If no such qq^{\prime} is found, then fail; otherwise, output q(1)q^{\prime}\mathopen{}\mathclose{{}\left(1}\right) as the guessed value of Perz(M)\mathrm{Per}_{z}\mathopen{}\mathclose{{}\left(M}\right). Now we show that the above algorithm succeeds (that is, outputs q(1)=Perz(M)q^{\prime}\mathopen{}\mathclose{{}\left(1}\right)=\mathrm{Per}_{z}\mathopen{}\mathclose{{}\left(M}\right)) with probability at least 1/2+δ/2{1}/{2}+{\delta}/{2} over YY. Provided that holds, it is clear that the success probability can be boosted to (say) 2/32/3 by simply repeating the algorithm O(1/δ2)O\mathopen{}\mathclose{{}\left(1/\delta^{2}}\right) times with different choices of YY and then outputting the majority result. To prove the claim, note that for each [L]\ell\in\mathopen{}\mathclose{{}\left[L}\right] one can think of the matrix X(ε)X\mathopen{}\mathclose{{}\left(\varepsilon\ell}\right) as having been drawn from the distribution

𝒟:=i,j=1n(εaij,(1ε)2).\mathcal{D}_{\ell}:=\prod_{i,j=1}^{n}\mathcal{F}_{\mathbb{C}}\mathopen{}\mathclose{{}\left(\varepsilon\ell a_{ij},\mathopen{}\mathclose{{}\left(1-\varepsilon\ell}\right)^{2}}\right)\,. (73)

Let

𝒟:=i,j=1n(εaij,1).\mathcal{D}_{\ell}^{\prime}:=\prod_{i,j=1}^{n}\mathcal{F}_{\mathbb{C}}\mathopen{}\mathclose{{}\left(\varepsilon\ell a_{ij},1}\right). (74)

Then by the triangle inequality together with Definition 5.2 (X(0,1)n×nX\sim\mathcal{F}_{\mathbb{C}}(0,1)^{n\times n}),

𝒟X\displaystyle\mathopen{}\mathclose{{}\left\|\mathcal{D}_{\ell}-X}\right\| 𝒟𝒟+𝒟X\displaystyle\leq\mathopen{}\mathclose{{}\left\|\mathcal{D}_{\ell}-\mathcal{D}_{\ell}^{\prime}}\right\|+\mathopen{}\mathclose{{}\left\|\mathcal{D}_{\ell}^{\prime}-X}\right\|
2n2M1ε+2n2M2n2(ε)2\displaystyle\leq 2n^{2}M_{1}\varepsilon\ell+\sqrt{2n^{2}}M_{2}\sqrt{n^{2}\mathopen{}\mathclose{{}\left(\varepsilon\ell}\right)^{2}}
(2n2M1+2n2M2)εl(2n2M1+2n2M2)εL\displaystyle\leq\mathopen{}\mathclose{{}\left(2n^{2}M_{1}+2n^{2}M_{2}}\right)\varepsilon l\leq(2n^{2}M_{1}+2n^{2}M_{2})\varepsilon L
δ2.\displaystyle\leq\frac{\delta}{2}\,. (75)

We have

Pr[𝒪(X(ε))=q(ε)]\displaystyle\Pr\mathopen{}\mathclose{{}\left[\mathcal{O}\mathopen{}\mathclose{{}\left(X\mathopen{}\mathclose{{}\left(\varepsilon\ell}\right)}\right)=q\mathopen{}\mathclose{{}\left(\varepsilon\ell}\right)}\right] 34+δ𝒟X\displaystyle\geq\frac{3}{4}+\delta-\mathopen{}\mathclose{{}\left\|\mathcal{D}_{\ell}-X}\right\|
34+δ2.\displaystyle\geq\frac{3}{4}+\frac{\delta}{2}\,. (76)

Now let SS be the set of all [L]\ell\in\mathopen{}\mathclose{{}\left[L}\right] such that 𝒪(X(ε))=q(ε)\mathcal{O}\mathopen{}\mathclose{{}\left(X\mathopen{}\mathclose{{}\left(\varepsilon\ell}\right)}\right)=q\mathopen{}\mathclose{{}\left(\varepsilon\ell}\right). Then by the reverse Markov’s inequality [EG01],

Pr[|S|(12+δ2)L]114δ212δ212+δ2.\Pr\mathopen{}\mathclose{{}\left[\mathopen{}\mathclose{{}\left|S}\right|\geq\mathopen{}\mathclose{{}\left(\frac{1}{2}+\frac{\delta}{2}}\right)L}\right]\geq 1-\frac{\frac{1}{4}-\frac{\delta}{2}}{\frac{1}{2}-\frac{\delta}{2}}\geq\frac{1}{2}+\frac{\delta}{2}\,. (77)

So we then just run the above algorithm O(1δ2)O(\frac{1}{\delta^{2}}) times to amplify the success probability. After that, with high probability, we have

|S|(12+δ2)L.\mathopen{}\mathclose{{}\left|S}\right|\geq\mathopen{}\mathclose{{}\left(\frac{1}{2}+\frac{\delta}{2}}\right)L\,.

Then by Theorem 5.1, the Berlekamp-Welch algorithm succeeds; It will output polynomial qq^{\prime} will be equal to qq. This proves the claim. ∎

6 Average-case Hardness of Approximation

Since Aaronson and Arkhipov’s seminal work on Boson Sampling [AA11] appeared, approximating 11-permanent in the average case has been investigated for particular distributions [AA11, HHEG20]. In particular, Aaronson and Arkhipov [AA11] conjectured a problem called Gaussian Permanent Estimation (also denoted GPE×\mathrm{GPE}_{\times}) is #𝖯{\#\mathsf{P}}-hard. The problem GPE×\mathrm{GPE}_{\times} asks to compute Per(X)\mathrm{Per}(X) to multiplicative error ϵ\epsilon with probability at least 1δ1-\delta over matrix XX whose each element is sampled independently from the normalized Gaussian distribution. For quantum supremacy from Boson Sampling, the ultimate goal was to prove the additive version of GPE×\mathrm{GPE}_{\times}, denoted GPE±\mathrm{GPE}_{\pm}, is also hard. Based on another conjecture, called the Permanent Anti-Concentration Conjecture (PACC), one can deduce that GPE±\mathrm{GPE}_{\pm} is as hard as GPE×\mathrm{GPE}_{\times}. In the following, we denote 𝒢=𝒩(0,1)n×n\mathcal{G}=\mathcal{N}_{\mathbb{C}}(0,1)^{n\times n} the distribution of an n×nn\times n i.i.d. Gaussian matrix with zero mean and unit variance.

Problem 3 (GPE×\mathrm{GPE}_{\times}).

Given as input X𝒩(0,1)n×nX\sim\mathcal{N}_{\mathbb{C}}(0,1)^{n\times n} of i.i.d. Gaussians (complex normal distribution) and ϵ,δ>0\epsilon,\delta>0, estimate Per(X)\mathrm{Per}(X) to within error ±ϵ|Per(X)|\pm\epsilon|\mathrm{Per}(X)| to probability at least 1δ1-\delta over XX.

Problem 4 (GPE±\mathrm{GPE}_{\pm}).

Given as input X𝒩(0,1)n×nX\sim\mathcal{N}_{\mathbb{C}}(0,1)^{n\times n} of i.i.d. Gaussians (complex normal distribution) and ϵ,δ>0\epsilon,\delta>0, estimate Per(X)\mathrm{Per}(X) additively (i.e., to within error ±ϵn!\pm\epsilon\sqrt{n!}), with probability at least 1δ1-\delta over XX

In this section, we consider the hardness of zz-permanent for matrices sampled from a general distribution \mathcal{H} whose each entry is i.i.d. sampled. We first define the multiplicative and the additive versions of Permanent Estimation (PE) for distribution \mathcal{H}.

Problem 5 (PE,z,×\mathrm{PE}_{\mathcal{H},z,\times}).

Given as input XX\sim\mathcal{H} and ϵ,δ>0\epsilon,\delta>0, estimate Per(X)\mathrm{Per}(X) to within error ±ϵ|Per(X)|\pm\epsilon|\mathrm{Per}(X)| to probability at least 1δ1-\delta over XX.

For convenience, PE,1,×\mathrm{PE}_{\mathcal{H},1,\times} is also denoted PE,×\mathrm{PE}_{\mathcal{H},\times}. Similarly, we can define an additive version of the above problem.

Problem 6 (PE,z,±\mathrm{PE}_{\mathcal{H},z,\pm}).

Given as input XX\sim\mathcal{H} and ϵ,δ>0\epsilon,\delta>0, estimate Perz(X)\mathrm{Per}_{z}(X) to within error ±ϵg(n)\pm\epsilon g(n) to probability at least 1δ1-\delta over XX. For convenience, PE,1,±\mathrm{PE}_{\mathcal{H},1,\pm} is also denoted PE,±\mathrm{PE}_{\mathcal{H},\pm}.

From the definition, we have GPE×=PE𝒢,×\mathrm{GPE}_{\times}=\mathrm{PE}_{\mathcal{G},\times} and GPE±=PE𝒢,±\mathrm{GPE}_{\pm}=\mathrm{PE}_{\mathcal{G},\pm} with g=n!g=\sqrt{n!}. Let d=(n2)d=\binom{n}{2} For r(0,1d+1)r\in(0,\frac{1}{d+1}), let Sr:={z=ei2πrζd+1i1:i[d+1]}S_{r}:=\{z=e^{i2\pi r}\zeta_{d+1}^{i-1}:i\in[d+1]\}, where ζk:=ei2π/k\zeta_{k}:=e^{i2\pi/k} is a primitive kk-th root of unity. Let 𝒪\mathcal{O} be an oracle, given XX\sim\mathcal{H}, outputs the approximations of zz-permanent of d+1d+1 points, i.e., Perz1(X),,Perzd+1(X)\mathrm{Per}_{z_{1}}(X),\ldots,\mathrm{Per}_{z_{d+1}}(X) for z1,,zd+1Srz_{1},\ldots,z_{d+1}\in S_{r}. More formally, we define 𝒪×\mathcal{O}_{\times} and 𝒪±\mathcal{O}_{\pm} to be oracles that approximate to multiplicative and additive errors, respectively:

Definition 6.1 (𝒪×,𝒪±\mathcal{O}_{\times},\mathcal{O}_{\pm}).

The oracles 𝒪×\mathcal{O}_{\times} and 𝒪±\mathcal{O}_{\pm} are defined as follows:

  • Given X,X\sim\mathcal{H}, 𝒪×(X,ϵ,δ)\mathcal{O}_{\times}(X,\epsilon,\delta) is the oracle that outputs a vector y=(y1,,yd+1)y=(y_{1},\ldots,y_{d+1})^{\top} such that for every i[d+1]i\in[d+1], |yiPerzi(X)|ϵ|Perzi(X)||y_{i}-\mathrm{Per}_{z_{i}}(X)|\leq\epsilon|\mathrm{Per}_{z_{i}}(X)| with probability at least 1δ1-\delta over choices of XX.

  • Given XX\sim\mathcal{H} and efficiently computable function gg, 𝒪±(X,ϵ,δ)\mathcal{O}_{\pm}(X,\epsilon,\delta) is the oracle that outputs a vector y=(y1,,yd+1)y=(y_{1},\ldots,y_{d+1})^{\top} such that for every i[d+1],i\in[d+1], we have: |yiPerzi(X)|ϵg(n)|y_{i}-\mathrm{Per}_{z_{i}}(X)|\leq\epsilon g(n) with probability at least 1δ1-\delta over choices of XX.

The rest of this section is devoted to showing that if 𝒪×\mathcal{O}_{\times} can be classically simulated, then PE,×\mathrm{PE}_{\mathcal{H},\times} can be solved efficiently, and similarly holds for 𝒪±\mathcal{O}_{\pm} and PE,±\mathrm{PE}_{\mathcal{H},\pm} in place of 𝒪×\mathcal{O}_{\times} and PE,±\mathrm{PE}_{\mathcal{H},\pm}. The proof is similar to Lipton’s proof [Lip91] for the average hardness of permanent over a finite field 𝔽p\mathbb{F}_{p} for p>np>n. In our case, we can approximate Per(X)\mathrm{Per}(X) given an approximation of Ω(n2)\Omega(n^{2}) points on the unit circle.

First, we show the following two lemmas for polynomial interpolation. One is for multiplicative error and one is for additive error.

Lemma 6.2.

For integer d1d\geq 1, let ff be a degree-dd complex-valued polynomial and 𝒜\mathcal{A} be an algorithm that approximates ff to additive error ϵ\epsilon on points x1,,xd+1x_{1},\ldots,x_{d+1}. There exists an algorithm which, on input x𝕋x\in\mathbb{T}, makes d+1d+1 calls to 𝒜\mathcal{A} and outputs an estimate of f(x)f(x) to additive error (d+1)V1ϵ(d+1)\|V^{-1}\|\epsilon, where V=V(x1,,xd+1)V=V(x_{1},\ldots,x_{d+1}) is the Vandermonde matrix with nodes x1,,xd+1x_{1},\ldots,x_{d+1}.

Proof.

Let y~1,,y~d+1\tilde{y}_{1},\ldots,\tilde{y}_{d+1} be the estimate of 𝒜\mathcal{A} on x1,,xd+1x_{1},\ldots,x_{d+1} respectively. Let y~=(y~1,,y~d+1)\tilde{y}=(\tilde{y}_{1},\ldots,\tilde{y}_{d+1})^{\top} and x=(x1,,xd+1)x=(x_{1},\ldots,x_{d+1})^{\top} be the associated column vectors. Since for each i[d+1]i\in[d+1], |y~iyi|ϵ|\tilde{y}_{i}-y_{i}|\leq\epsilon, y~yϵd+1\|\tilde{y}-y\|\leq\epsilon\sqrt{d+1}.

Let c=(c0,,cd)c=(c_{0},\ldots,c_{d})^{\top} be the vector of the coefficients of f(x)=i=0dcixif(x)=\sum_{i=0}^{d}c_{i}x^{i}. Solving the linear equation Vc~=y~V\tilde{c}=\tilde{y} yields an estimate of the coefficients. For x𝕋x\in\mathbb{T}, let v=(1,x,x2,,xd)v=(1,x,x^{2},\ldots,x^{d}) be the vector of monomials evaluated on xx. Our goal is to bound the distance

|v(c~c)|(V1)vV(cc)vV1y~y(d+1)V1ϵ.\displaystyle|v\cdot(\tilde{c}-c)|\leq\|(V^{-1})^{\top}v\|\|V(c-c^{\prime})\|\leq\|v\|\|V^{-1}\|\|\tilde{y}-y\|\leq(d+1)\|V^{-1}\|\epsilon. (78)

The first inequality holds by Cauchy-Schwarz inequality. The second holds by the definition of matrix norm. The last holds because y~yϵ\|\tilde{y}-y\|\leq\epsilon and v=d+1\|v\|=\sqrt{d+1}. ∎

If we restrict our problem to approximating f(1)f(1) for polynomials with non-negative coefficients, we can do better. In particular, we show that if 𝒜\mathcal{A} approximates x1,,xd+1x_{1},\ldots,x_{d+1} for polynomial ff with non-negative coefficients to multiplicative error ϵ\epsilon, then there is an algorithm that approximates f(1)f(1) to the multiplicative error as above.

Lemma 6.3.

For integer d1d\geq 1, let ff be a degree-dd polynomial with non-negative coefficients and 𝒜\mathcal{A} be an algorithm that approximates ff to multiplicative error ϵ\epsilon on x1,,xd+1𝕋x_{1},\ldots,x_{d+1}\in\mathbb{T}. Then there exists an algorithm which approximates f(1)f(1) to multiplicative error (d+1)V1ϵ(d+1)\|V^{-1}\|\epsilon, where V=V(x1,,xd+1)V=V(x_{1},\ldots,x_{d+1}) is the Vandermonde matrix with nodes x1,,xd+1x_{1},\ldots,x_{d+1}.

Proof.

Let y~1,,y~d+1\tilde{y}_{1},\ldots,\tilde{y}_{d+1} be the estimate of 𝒜\mathcal{A} on x1,,xd+1x_{1},\ldots,x_{d+1} and y=(y~1,,y~d+1)y=(\tilde{y}_{1},\ldots,\tilde{y}_{d+1})^{\top}. Solving the linear equation Vc~=y~V\tilde{c}=\tilde{y} yields c~=V1y~\tilde{c}=V^{-1}\tilde{y}. Also let f~(z):=i=0dc~izi\tilde{f}(z):=\sum_{i=0}^{d}\tilde{c}_{i}z^{i} be the approximated polynomial obtained by solving the linear system.

Since ff has non-negative coefficients, |f(x)|f(1)=1d+1,c|f(x)|\leq f(1)=\langle 1^{d+1},c\rangle for every x𝕋x\in\mathbb{T}. Since y~yϵyϵd+1f(1)\|\tilde{y}-y\|\leq\epsilon\|y\|\leq\epsilon\sqrt{d+1}\cdot f(1),

|f~(1)f(1)|=|1d+1,c~c|d+1c~c(d+1)V1ϵ|f(1)|.\displaystyle|\tilde{f}(1)-f(1)|=|\langle 1^{d+1},\tilde{c}-c\rangle|\leq\sqrt{d+1}\|\tilde{c}-c\|\leq(d+1)\|V^{-1}\|\cdot\epsilon\cdot|f(1)|. (79)

The first inequality holds by Cauchy-Schwarz inequality. Thus an algorithm which outputs |c~|1|\tilde{c}|_{1} as the approximation with multiplicative error (d+1)ϵV1(d+1)\epsilon\|V^{-1}\|. ∎

Now we are ready to prove Theorem 6.4. By Lemma 6.2 and Lemma 6.3, we separate them into two cases.

Theorem 6.4.

The following statements hold:

  1. 1.

    Suppose PE,×\mathrm{PE}_{\mathcal{H},\times} is #𝖯{\#\mathsf{P}}-hard and \mathcal{H} is non-negative distribution, and the polynomial hierarchy is infinite, then 𝒪×\mathcal{O}_{\times} does not admit a classical efficient algorithm, i.e., there is no simulation of 𝒪×\mathcal{O}_{\times} running in time poly(n,1ϵ,1δ)\operatorname{poly}(n,\frac{1}{\epsilon},\frac{1}{\delta}).

  2. 2.

    Suppose (PE,±,g)(\mathrm{PE}_{\mathcal{H},\pm},g) is #𝖯{\#\mathsf{P}}-hard and the polynomial hierarchy is infinite, then 𝒪±\mathcal{O}_{\pm} does not admit a classical efficient algorithm, i.e., there is no classical simulation of 𝒪±\mathcal{O}_{\pm} running in time poly(n,1ϵ,1δ)\operatorname{poly}(n,\frac{1}{\epsilon},\frac{1}{\delta})).

Proof.

For (1), we give non-constructive proof. Let d=(n2)d=\binom{n}{2}. For r(0,1d+1)r\in(0,\frac{1}{d+1}), let Sr:={z=ei2πrζd+1i1:i[d+1]}S_{r}:=\{z=e^{i2\pi r}\zeta_{d+1}^{i-1}:i\in[d+1]\}, where ζk:=ei2π/k\zeta_{k}:=e^{i2\pi/k} is a primitive kk-th root of unity. We show that the number of points zSrz\in S_{r} such that Perz(X)\mathrm{Per}_{z}(X) is efficiently computable cannot be more than dd. Suppose toward contradiction this is not the case. Then let 𝒪×\mathcal{O}_{\times} be algorithms that runs in time poly(n,1/ϵ,1/δ)\operatorname{poly}(n,1/\epsilon,1/\delta) and approximates Perz1(X),,Perzd+1(X)\mathrm{Per}_{z_{1}}(X),\ldots,\mathrm{Per}_{z_{d+1}}(X) to precision ϵ\epsilon with probability 1δ1-\delta, for Sr={z1,,zd+1}S_{r}=\{z_{1},\ldots,z_{d+1}\}. Applying the algorithm in the proof of Lemma 6.3, we apply the oracle 𝒪×(X,ϵ,δ)\mathcal{O}_{\times}(X,\epsilon,\delta) for ×\mathcal{H}_{\times} by the following algorithm:

  1. 1.

    Get y=(y1,,yd+1)𝒪×(X,ϵ(d+1)1/2,δ/(d+1))y=(y_{1},\ldots,y_{d+1})^{\top}\leftarrow\mathcal{O}_{\times}(X,\epsilon(d+1)^{-1/2},\delta/(d+1));

  2. 2.

    Solving the linear system Vc=yVc=y, where V=V(z1,,zd+1)V=V(z_{1},\ldots,z_{d+1}).

  3. 3.

    Output the summation of elements of cc as the approximation of Per(X)\mathrm{Per}(X).

By the assumption, for each i[d+1],i\in[d+1],with probability at least 1δd+11-\frac{\delta}{d+1}, |yiPerzi(X)|ϵd+1|Perzi(X)||y_{i}-\mathrm{Per}_{z_{i}}(X)|\leq\frac{\epsilon}{\sqrt{d+1}}|\mathrm{Per}_{z_{i}}(X)|. Since points in SrS_{r} is equally spaced, 1d+1V\frac{1}{\sqrt{d+1}}V is a unitary matrix, and so is d+1V1\sqrt{d+1}V^{-1}. This implies that the matrix norm V1=(d+1)1/2\|V^{-1}\|=(d+1)^{-1/2}. By Lemma 6.3 and union bound, the algorithm outputs an approximation of Per(X)\mathrm{Per}(X) to multiplicative error ϵ\epsilon with probability at least 1δ1-\delta. By the assumption PE,×\mathrm{PE}_{\mathcal{H},\times} is hard, we conclude 𝖯#𝖯𝖡𝖯𝖯𝒪×𝖡𝖯𝖯\mathsf{P}^{{\#\mathsf{P}}}\subseteq\mathsf{BPP}^{\mathcal{O}_{\times}}\subseteq\mathsf{BPP} and a collapse of the polynomial hierarchy.

For (2), we use the same SrS_{r}, then let 𝒪±\mathcal{O}_{\pm} be algorithms that runs in time poly(n,1/ϵ,1/δ)\operatorname{poly}(n,1/\epsilon,1/\delta) and approximates Perz1(X),,Perzd+1(X)\mathrm{Per}_{z_{1}}(X),\ldots,\mathrm{Per}_{z_{d+1}}(X) to precision ϵ\epsilon with probability 1δ1-\delta, respectively for Sr={z1,,zd+1}S_{r}=\{z_{1},\ldots,z_{d+1}\}. Applying the algorithm in the proof of Lemma 6.2, we apply the oracle 𝒪±(X,ϵ,δ)\mathcal{O}_{\pm}(X,\epsilon,\delta) for ±\mathcal{H}_{\pm} by the following algorithm:

  1. 1.

    Get y=(y1,,yd+1)𝒪±(X,ϵ(d+1)1/2,δ/(d+1))y=(y_{1},\ldots,y_{d+1})^{\top}\leftarrow\mathcal{O}_{\pm}(X,\epsilon(d+1)^{-1/2},\delta/(d+1)).

  2. 2.

    Solving the linear system Vc=yVc=y, where V=V(z1,,zd+1)V=V(z_{1},\ldots,z_{d+1}).

  3. 3.

    Output the summation of elements of cc as the approximation of Per(X)\mathrm{Per}(X).

By the assumption, for each i[d+1],i\in[d+1],with probability at least 1δd+11-\frac{\delta}{d+1}, |yiPerzi(X)|ϵd+1g(n)|y_{i}-\mathrm{Per}_{z_{i}}(X)|\leq\frac{\epsilon}{\sqrt{d+1}}g(n). Since points in SrS_{r} is equally spaced, 1d+1V\frac{1}{\sqrt{d+1}}V is a unitary matrix, and so is d+1V1\sqrt{d+1}V^{-1}. This implies that the matrix norm V1=(d+1)1/2\|V^{-1}\|=(d+1)^{-1/2}. By Lemma 6.2 and a union bound, the algorithm outputs an approximation of Per(X)\mathrm{Per}(X) to additive error ϵg(n)\epsilon g(n) with probability at least 1δ1-\delta. By our assumption (PE,×,g)(\mathrm{PE}_{\mathcal{H},\times},g) #𝖯{\#\mathsf{P}}-hard, we conclude that 𝖯#𝖯𝖡𝖯𝖯𝒪±𝖡𝖯𝖯\mathsf{P}^{{\#\mathsf{P}}}\subseteq\mathsf{BPP}^{\mathcal{O}_{\pm}}\subseteq\mathsf{BPP} and the collapse of the polynomial hierarchy. ∎

At first glance, it seems like our result does not work for an i.i.d. Gaussian matrix since it is not positive except with exponentially small probability. However, similarly as in [AA11], we prove that PE𝒢,z,±\mathrm{PE}_{\mathcal{G},z,\pm} with g(n)=n!g(n)=\sqrt{n!} can be reduced to PE𝒢,z,×\mathrm{PE}_{\mathcal{G},z,\times}.

Lemma 6.5.

(PE𝒢,z,±)(\mathrm{PE}_{\mathcal{G},z,\pm}) with g=n!g=\sqrt{n!} is polynomial-time reducible to PE𝒢,z,×\mathrm{PE}_{\mathcal{G},z,\times}.

Proof.

Suppose we have a polynomial-time algorithm MM outputs a good multiplicative approximation to Perz(X)\mathrm{Per}_{z}(X)—that is, a ww such that:

|wPerz(X)|ϵ|Perz(X)|.|w-\mathrm{Per}_{z}(X)|\leq\epsilon|\mathrm{Per}_{z}(X)|. (80)

with probability at least 1δ1-\delta over XX by Markov’s inequality (𝔼X𝒢[|Perz(X)|2]=n!,\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{2}]=n!, see Lemma A.1):

PrX𝒢[|wPerz(X)|>kn!]<1k2\Pr_{X\sim\mathcal{G}}\mathopen{}\mathclose{{}\left[|w-\mathrm{Per}_{z}(X)|>k\sqrt{n!}}\right]<\frac{1}{k^{2}} (81)

by the union bound,

PrX𝒢[|wPerz(X)|>ϵkn!]\displaystyle\Pr_{X\sim\mathcal{G}}\mathopen{}\mathclose{{}\left[|w-\mathrm{Per}_{z}(X)|>\epsilon k\sqrt{n!}}\right] PrX𝒢[|wPerz(X)|>ϵ|Perz(X)|]+PrX𝒢[ϵ|Perz(X)|>ϵkn!]\displaystyle\leq\Pr_{X\sim\mathcal{G}}\Big{[}|w-\mathrm{Per}_{z}(X)|>\epsilon|\mathrm{Per}_{z}(X)|\Big{]}+\Pr_{X\sim\mathcal{G}}\Big{[}\epsilon|\mathrm{Per}_{z}(X)|>\epsilon k\sqrt{n!}\Big{]}
δ+1k2.\displaystyle\leq\delta+\frac{1}{k^{2}}.

Letting δ=δ2,k=2δ\delta=\frac{\delta^{\prime}}{2},k=\sqrt{\frac{2}{\delta}} and ϵ=ϵk\epsilon=\frac{\epsilon^{\prime}}{k}, we have δ+1k2=δ\delta+\frac{1}{k^{2}}=\delta^{\prime} which finishes the proof. ∎

We now introduce the conjecture that generalize the original permanent anti-concentration conjecture.

Conjecture 3 (zz-permanent anti-concentration conjecture (zz-PACC)).

There exists a polynomial pp such that for positive integer nn, real number δ>0\delta>0 and z𝕋z\in\mathbb{T},

PrX𝒢[|Perz(X)|2n!p(n,1/δ)]1δ.\displaystyle\Pr_{X\sim\mathcal{G}}\mathopen{}\mathclose{{}\left[|\mathrm{Per}_{z}(X)|^{2}\geq\frac{\sqrt{n!}}{p(n,1/\delta)}}\right]\geq 1-\delta. (82)

In particular, z=1z=-1 is proved by Aaronson [AA11] and z=1z=1 is standard conjecture that widely believed to be True.

Corollary 6.6.

Assume that the 11-PACC holds. Suppose GPE×\mathrm{GPE}_{\times} is #𝖯{\#\mathsf{P}}-hard and the polynomial hierarchy is infinite, then 𝒪×\mathcal{O}_{\times} does not admit a classical efficient algorithm, i.e., there is no simulation of 𝒪×\mathcal{O}_{\times} running in time poly(n,1ϵ,1δ)\operatorname{poly}(n,\frac{1}{\epsilon},\frac{1}{\delta}).

Proof.

Suppose PE𝒢,z,×\mathrm{PE}_{\mathcal{G},z,\times} admit a classical efficient algorithm in poly(n,1ϵ,1δ)\operatorname{poly}(n,\frac{1}{\epsilon},\frac{1}{\delta}). Then one can solve PE𝒢,z,×\mathrm{PE}_{\mathcal{G},z,\times} then solving GPEz,×\mathrm{GPE}_{z,\times} for d+1d+1 points. By Lemma 6.5, it solves (PE𝒢,z,±)(\mathrm{PE}_{\mathcal{G},z,\pm}) with n!\sqrt{n!} for d+1d+1 points which solves GPE±\mathrm{GPE}_{\pm}. Assuming 11-PACC [AA11], GPE±\mathrm{GPE}_{\pm} is polynomial equivalent to GPE×.\mathrm{GPE}_{\times}. Hence there is classical polynomial solves GPE×\mathrm{GPE}_{\times} which contradicts our assumption that polynomial hierarchy is infinite. ∎

Finally, we show that approximating Perz(X)\mathrm{Per}_{z}(X) is as hard as approximating Perz(X)\mathrm{Per}_{z^{*}}(X) for any distribution XX such that X,XX,X^{*} are identically distributed. A similar proof idea applies to |Perz(X)|2|\mathrm{Per}_{z}(X)|^{2} as well.

Lemma 6.7.

For any distribution XX such that X,XX,X^{*} are identically distributed, if there is an algorithm running in time poly(n,1/δ,1/ϵ)\operatorname{poly}(n,1/\delta,1/\epsilon) for approximating Perz(X)\mathrm{Per}_{z}(X) to within multiplicative (resp. additive) error ϵ\epsilon with probability 1δ1-\delta, then there is an algorithm which approximates Perz(X)\mathrm{Per}_{z^{*}}(X) to within multiplicative (resp. additive) error ϵ\epsilon with probability 1δ1-\delta.

Proof.

By the assumption, let 𝒜(X,01/ϵ,01/δ)\mathcal{A}(X,0^{1/\epsilon},0^{1/\delta}) be an algorithm that approximates Perz(X)\mathrm{Per}_{z}(X) to error ϵ\epsilon with probability 1δ1-\delta.

Since Perz(X)=Perz(X)\mathrm{Per}_{z^{*}}(X)^{*}=\mathrm{Per}_{z}(X^{*}) and X,XX,X^{*} are identically distributed, for computing Perz(X)\mathrm{Per}_{z^{*}}(X), the algorithm \mathcal{B} runs 𝒜(X,01/ϵ,01/δ)\mathcal{A}(X^{*},0^{1/\epsilon},0^{1/\delta}) to obtain an approximation P~\tilde{P} of Perz(X)\mathrm{Per}_{z}(X^{*}) with probability 1δ1-\delta over XX^{*}. Then since the absolute value of a complex number does not change by taking complex conjugate,

PrX[|P~Perz(X)|α(z,X)]\displaystyle\Pr_{X}[|\tilde{P}-\mathrm{Per}_{z}(X^{*})|\leq\alpha(z,X^{*})] =PrX[|P~Perz(X)|α(z,X)]\displaystyle=\Pr_{X}[|\tilde{P}^{*}-\mathrm{Per}_{z^{*}}(X)|\leq\alpha(z^{*},X)]
1δ.\displaystyle\geq 1-\delta. (83)

Here for multiplicative error

α(z,X)=ϵ|Perz(X)|=ϵ|Perz(X)|=α(z,X).\displaystyle\alpha(z,X^{*})=\epsilon|\mathrm{Per}_{z}(X^{*})|=\epsilon|\mathrm{Per}_{z^{*}}(X)|=\alpha(z^{*},X). (84)

For additive error,

α(z,X)=ϵn!=α(z,X).\displaystyle\alpha(z,X^{*})=\epsilon\sqrt{n!}=\alpha(z^{*},X). (85)

7 Open Problems and Discussion

In this paper, we studied the computational complexity of quantum determinants, a qq-deformation of matrix permanents. The qq-permanent of a matrix XX is defined and it generalizes both determinant and permanent. It is shown that computing the qq-permanent is 𝖬𝗈𝖽p𝖯\mathsf{Mod}_{p}\mathsf{P}-hard for a primitive mm-th root of unity qq for odd prime power m=pkm=p^{k}. This result implies that an efficient algorithm for computing qq-permanent would result in a collapse of the polynomial hierarchy. We also showed that an efficient approximation algorithm for qq-permanent would also imply a collapse of the polynomial hierarchy. The hardness of computing qq-permanent remains to hold for a wide range of distributions by random self-reducibility. The techniques developed for the hardness of permanent can be generalized to zz-permanents for both worst and average cases.

Here we include a few interesting questions that are not solved in this paper.

  • It is known that there is a polynomial-time algorithm for determining whether Per(X)0\mathrm{Per}(X)\neq 0 for {0,1}\{0,1\}-matrix XX by determining if there is a perfect matching in a bipartite graph whose adjacent matrix is XX. Is the property about zz-permanent efficiently computable for z1z\neq 1, i.e., is there a polynomial-time algorithm for determining whether Perz(X)0\mathrm{Per}_{z}(X)\neq 0?

  • We have shown that Problem 2 is in 𝖭𝖯\mathsf{NP}. Does Problem 2 have polynomial-time algorithm?

  • In this work, we have discussed the computational complexity of zz-permanent for z𝕋z\in\mathbb{T}. It is natural to consider other subsets of the complex numbers. For example, what is the computational complexity of zz-permanent for real zz?

  • As explained in Section 3.2, the techniques we used for proving the hardness of ζm\zeta_{m}-permanent can only work for the special case where mm is a prime power. Thus our result does not rule out the existence of an efficient algorithm for non-prime-power mm. Is there a polynomial-time algorithm computing ζm\zeta_{m}-permanent where mm is not prime power?

  • First proposed by Aaronson and Arkhipov, the permanent anti-concentration conjecture is crucial for showing the equivalence of additive and multiplicative approximation of 11-permanent for Gaussian matrices.

    Relation between the permanent anti-concentration conjecture and the zz-permanent anti-concentration Conjecture 3. Can one prove one from another or use polynomial interpolation to give multiple-to-one reduction?

  • Higher moment of |Perz(X)|2|\mathrm{Per}_{z}(X)|^{2} when XX is drawn from complex Gaussian 𝒢\mathcal{G}.

Acknowledgement

We thank Scott Aaronson, Mohammad Hafezi, and Dominik Hangleiter for useful discussions. EJ was supported by ARO W911NF-15-1-0397, National Science Foundation QLCI grant OMA-2120757, AFOSRMURI FA9550-19-1-0399, Department of Energy QSA program, and the Simons Foundation. SHH acknowledges the support from Simons Investigator in Computer Science award, award number 510817.

References

  • [AA11] Scott Aaronson and Alex Arkhipov. The computational complexity of linear optics. In Proceedings of the forty-third annual ACM symposium on Theory of computing, pages 333–342, 2011.
  • [Ben87] Georgia Benkart. Abstract algebra, by in herstein. The American Mathematical Monthly, 94(8):804–806, 1987.
  • [BKW19] Andreas Björklund, Petteri Kaski, and Ryan Williams. Generalized kakeya sets for polynomial evaluation and faster computation of fermionants. Algorithmica, 81(10):4010–4028, 2019.
  • [BL94] RB Bapat and AK Lal. Inequalities for the q-permanent. Linear Algebra and its Applications, 197:397–409, 1994.
  • [BLW17] Marek Bożejko, Eugene Lytvynov, and Janusz Wysoczański. Fock representations of q-deformed commutation relations. Journal of Mathematical Physics, 58(7):073501, 2017.
  • [BP86] Dario Bini and Victor Pan. Polynomial division and its computational complexity. Journal of Complexity, 2(3):179–203, 1986.
  • [Bro86] Andrei Z Broder. How hard is it to marry at random?(on the approximation of the permanent). In Proceedings of the eighteenth annual ACM symposium on Theory of computing, pages 50–58, 1986.
  • [BS91] Marek Bożejko and Roland Speicher. An example of a generalized brownian motion. Communications in Mathematical Physics, 137(3):519–531, 1991.
  • [Bür00] Peter Bürgisser. The computational complexity of immanants. SIAM Journal on Computing, 30(3):1023–1040, 2000.
  • [BV] ODYSSEAS BAKAS and MARCO VITTURI. The polynomial method lecture 4 applications in technology.
  • [Cur21] Radu Curticapean. A full complexity dichotomy for immanant families. In Proceedings of the 53rd Annual ACM SIGACT Symposium on Theory of Computing, pages 1770–1783, 2021.
  • [CX15] Radu Curticapean and Mingji Xia. Parameterizing the permanent: Genus, apices, minors, evaluation mod 2k. In 2015 IEEE 56th Annual Symposium on Foundations of Computer Science, pages 994–1009. IEEE, 2015.
  • [CX21] Radu Curticapean and Mingji Xia. Parameterizing the permanent: Hardness for k_8k\_8-minor-free graphs. arXiv preprint arXiv:2108.12879, 2021.
  • [DF10] CM Da Fonseca. The μ\mu-permanent of a tridiagonal matrix, orthogonal polynomials, and chain sequences. Linear algebra and its applications, 432(5):1258–1266, 2010.
  • [dF18] Carlos M da Fonseca. The μ\mu -permanent revisited. arXiv preprint arXiv:1804.02231, 2018.
  • [DJ21] Samir Datta and Kishlaya Jaiswal. Parallel polynomial permanent mod powers of 2 and shortest disjoint cycles. arXiv preprint arXiv:2106.00714, 2021.
  • [dRA13] Nicolas de Rugy-Altherre. Determinant versus permanent: salvation via generalization? the algebraic complexity of the fermionant and the immanant. arXiv preprint arXiv:1309.2156, 2013.
  • [DS] EDUARDO MARQUES DE SA. On q-permanent expansions and a theorem on cycle surgery.
  • [dS18] Eduardo Marques de Sá. Noncrossing partitions, noncrossing graphs, and q-permanental equations. Linear Algebra and its Applications, 541:36–53, 2018.
  • [EG01] Bennett Eisenberg and BK Ghosh. A generalization of markov’s inequality. Statistics & probability letters, 53(1):59–65, 2001.
  • [EM18] Lior Eldar and Saeed Mehraban. Approximating the permanent of a random matrix with vanishing mean, 2018.
  • [ER98] Pavel Etingof and Vladimir Retakh. Quantum determinants and quasideterminants. arXiv preprint math/9808065, 1998.
  • [FK06] Jürg Fröhlich and Thomas Kerler. Quantum groups, quantum categories and quantum field theory. Springer, 2006.
  • [GM04] Gerald A Goldin and Shahn Majid. On the fock space for nonrelativistic anyon fields and braided tensor products. Journal of mathematical physics, 45(10):3770–3787, 2004.
  • [GRASRA96] César Gómez, Martí Ruiz-Altaba, Germán Sierra, and Marti Ruiz-Altaba. Quantum groups in two-dimensional physics, volume 139. Cambridge University Press Cambridge, 1996.
  • [Har85] Werner Hartmann. On the complexity of immanants. Linear and Multilinear Algebra, 18(2):127–140, 1985.
  • [HHEG20] Jonas Haferkamp, Dominik Hangleiter, Jens Eisert, and Marek Gluza. Contracting projected entangled pair states is average-case hard. Physical Review Research, 2(1):013010, 2020.
  • [HJ12] Roger A Horn and Charles R Johnson. Matrix analysis. Cambridge university press, 2012.
  • [Lal98] AK Lal. Inequalities for the q-permanent. ii. Linear algebra and its applications, 274(1-3):1–16, 1998.
  • [Lan12] Serge Lang. Cyclotomic fields I and II, volume 121. Springer Science & Business Media, 2012.
  • [Lan13] Serge Lang. Algebraic number theory, volume 110. Springer Science & Business Media, 2013.
  • [Lip91] Richard Lipton. New directions in testing. Distributed computing and cryptography, 2:191–202, 1991.
  • [LW12] Claude Levesque and Michel Waldschmidt. Approximation of an algebraic number by products of rational numbers and units. Journal of the Australian Mathematical Society, 93(1-2):121–131, 2012.
  • [M+97] Meena Mahajan et al. Determinant: Combinatorics, algorithms, and complexity. Chicago Journal of Theoretical Computer Science, 1997(5), 1997.
  • [Maj00] Shahn Majid. Foundations of quantum group theory. Cambridge university press, 2000.
  • [MM11] Stephan Mertens and Cristopher Moore. The complexity of the fermionant, and immanants of constant width. arXiv preprint arXiv:1110.1821, 2011.
  • [Neu13] Jürgen Neukirch. Algebraic number theory, volume 322. Springer Science & Business Media, 2013.
  • [Nez21] Sepehr Nezami. Permanent of random matrices from representation theory: moments, numerics, concentration, and comments on hardness of boson-sampling. arXiv preprint arXiv:2104.06423, 2021.
  • [PS90] V Pasquier and H Saleur. Common structures between finite systems and conformal field theories through quantum groups. Nuclear Physics B, 330(2-3):523–556, 1990.
  • [R+76] Walter Rudin et al. Principles of mathematical analysis, volume 3. McGraw-hill New York, 1976.
  • [RA13] Nicolas de Rugy-Altherre. Determinant versus permanent: salvation via generalization? In Conference on Computability in Europe, pages 87–96. Springer, 2013.
  • [RS62] J Barkley Rosser and Lowell Schoenfeld. Approximate formulas for some functions of prime numbers. Illinois Journal of Mathematics, 6(1):64–94, 1962.
  • [Spi20] Dylan Spivak. Immanants and their applications in quantum optics. PhD thesis, 2020.
  • [TAG93] Hiroyuki TAGAWA. q-analogues of determinants and symmetric chain decompositions. Tokyo Journal of Mathematics, 16(2):311–320, 1993.
  • [TO92] Seinosuke Toda and Mitsunori Ogiwara. Counting classes are at least as hard as the polynomial-time hierarchy. SIAM Journal on Computing, 21(2):316–328, 1992.
  • [Tod91] Seinosuke Toda. Pp is as hard as the polynomial-time hierarchy. SIAM Journal on Computing, 20(5):865–877, 1991.
  • [Val79a] Leslie G Valiant. Completeness classes in algebra. In Proceedings of the eleventh annual ACM symposium on Theory of computing, pages 249–261, 1979.
  • [Val79b] Leslie G Valiant. The complexity of computing the permanent. Theoretical computer science, 8(2):189–201, 1979.
  • [Vau75] RC Vaughan. Bounds for the coefficients of cyclotomic polynomials. Michigan Mathematical Journal, 21(4):289–295, 1975.
  • [Wei98] Edwin Weiss. Algebraic number theory. Courier Corporation, 1998.
  • [Wei04] Eric W Weisstein. Fermat’s little theorem. https://mathworld. wolfram. com/, 2004.
  • [Yan91] Kung-Wei Yang. q–determinants and permutations. The Fibonacci Quarterly, 29:160–163, 1991.

Appendix A Moments of zz-permanents and Anti-Concentration Conjecture

Anti-concentration not only plays an important rule in the Boson Sampling [AA11] but also has its own interests in random matrix theory [EM18, Nez21]. In this section, we discuss the moments of zz-permanents drawn from i.i.d complex Gaussian and anti-concentration conjecture of zz-permanent.

First, we explicitly calculate the first moment of |Perz(X)|2|\mathrm{Per}_{z}(X)|^{2}. Interestingly, the result does not depend on zz.

Lemma A.1.

For z𝕋z\in\mathbb{T}, 𝔼X𝒢[|Perz(X)|2]=n!\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{2}]=n!

Proof.

By direct calculation,

𝔼X𝒢[|Perz(X)|2]\displaystyle\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{2}] =σ,αz(σ)(α)𝔼X𝒢[iXi,σ(i)Xi,α(i)]\displaystyle=\sum_{\sigma,\alpha}z^{\ell(\sigma)-\ell(\alpha)}\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}\mathopen{}\mathclose{{}\left[\prod_{i}X_{i,\sigma(i)}X^{*}_{i,\alpha(i)}}\right]
=σ,αz(σ)(α)i𝟙[σ(i)=α(i)]\displaystyle=\sum_{\sigma,\alpha}z^{\ell(\sigma)-\ell(\alpha)}\prod_{i}\mathbbm{1}[\sigma(i)=\alpha(i)]
=σ,αz(σ)(α)𝟙[σ=α]\displaystyle=\sum_{\sigma,\alpha}z^{\ell(\sigma)-\ell(\alpha)}\cdot\mathbbm{1}[\sigma=\alpha]
=σ1=n!.\displaystyle=\sum_{\sigma}1=n!. (86)

In fact, if we replace 𝒢\mathcal{G} with a random matrix whose each entry is independently sampled from any distribution of zero mean and unit variance, the second moment is n!n!. Now we want to calculate 𝔼X𝒢[|Perz(X)|4]\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{4}]. We first recall the computation of 𝔼X𝒢[|Per(X)|4]\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}(X)|^{4}] in [AA11].

Lemma A.2.

For z=1z=1, 𝔼X𝒢[|Perz(X)|4]=(n!)2(n+1)\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{4}]=(n!)^{2}(n+1).

Proof.

By direction calculation,

𝔼X𝒢[|Perz(X)|4]\displaystyle\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{4}] =σ,τ,α,βz(σ)+(τ)(α)(β)M(σ,τ,α,β),\displaystyle=\sum_{\sigma,\tau,\alpha,\beta}z^{\ell(\sigma)+\ell(\tau)-\ell(\alpha)-\ell(\beta)}\cdot M(\sigma,\tau,\alpha,\beta), (87)

where

M(σ,τ,α,β):=i𝔼X𝒢[Xi,σ(i)Xi,τ(i)Xi,α(i)Xi,β(i)].\displaystyle M(\sigma,\tau,\alpha,\beta):=\prod_{i}\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[X_{i,\sigma(i)}X_{i,\tau(i)}X_{i,\alpha(i)}^{*}X_{i,\beta(i)}^{*}]. (88)

If σταβ\sigma\cup\tau\neq\alpha\cup\beta, M(σ,τ,α,β)=0M(\sigma,\tau,\alpha,\beta)=0.

To compute M(σ,τ,α,β)M(\sigma,\tau,\alpha,\beta), we first provide some intuition: if τ=1\tau=1 and σ=γ1γm\sigma=\gamma_{1}\ldots\gamma_{m}, where each γi\gamma_{i} is a cycle of length greater than 1. Then either α\alpha has γi\gamma_{i} and β\beta has the identity, or α\alpha has the identity and β\beta has γi\gamma_{i}. This implies that βα=σ\beta\alpha=\sigma and there are 2m2^{m} choices of α\alpha. For arbitrary τ\tau, we can “shift” each permutation by multiplying τ1\tau^{-1} to the right and reducing to the above simpler case. Since στ1\sigma\tau^{-1} and τσ1\tau\sigma^{-1} have the same cycle type, the above argument is symmetric with respect to swapping the roles of σ\sigma and τ\tau (i.e., we can start with σ=1\sigma=1 and get the same result).

More formally, if στ1\sigma\tau^{-1} is of cycle type λ=(λ1,,λk)\lambda=(\lambda_{1},\ldots,\lambda_{k}) with λ1λ2λm>λm+1==λk=1\lambda_{1}\geq\lambda_{2}\geq\ldots\geq\lambda_{m}>\lambda_{m+1}=\ldots=\lambda_{k}=1, we may write in the form

στ1=γ1γ2γm.\displaystyle\sigma\tau^{-1}=\gamma_{1}\gamma_{2}\ldots\gamma_{m}. (89)

Then if στ=αβ\sigma\cup\tau=\alpha\cup\beta, it holds that

(α,β)A(σ,τ)\displaystyle(\alpha,\beta)\in A(\sigma,\tau) (90)

where

A(σ,τ)={(α,β):α=γ1b1γmbmτ,β=τα1σ,b1,,bm{0,1}}.\displaystyle A(\sigma,\tau)=\mathopen{}\mathclose{{}\left\{(\alpha,\beta):\alpha=\gamma_{1}^{b_{1}}\ldots\gamma_{m}^{b_{m}}\tau,\beta=\tau\alpha^{-1}\sigma,b_{1},\ldots,b_{m}\in\{0,1\}}\right\}. (91)

Clearly for every (σ,τ)(\sigma,\tau) such that στ1\sigma\tau^{-1} is of type λ\lambda, |A(σ,τ)|=2m|A(\sigma,\tau)|=2^{m} since there are 2m2^{m} choice of α\alpha and β\beta is uniquely determined once α\alpha is fixed.

Furthermore, for each fixed point ii, σ(i)=τ(i)=α(i)=β(i)\sigma(i)=\tau(i)=\alpha(i)=\beta(i), and thus i:σ(i)=τ(i)𝔼X𝒢[|Xi,σ(i)|4]=2km\prod_{i:\sigma(i)=\tau(i)}\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|X_{i,\sigma(i)}|^{4}]=2^{k-m}. Therefore, we have shown that

M(σ,τ,α,β)=2F(στ1)𝟙[(α,β)A(σ,τ)]\displaystyle M(\sigma,\tau,\alpha,\beta)=2^{F(\sigma\tau^{-1})}\cdot\mathbbm{1}[(\alpha,\beta)\in A(\sigma,\tau)] (92)

and

α,βM(σ,τ,α,β)=2C(στ1),\displaystyle\sum_{\alpha,\beta}M(\sigma,\tau,\alpha,\beta)=2^{C(\sigma\tau^{-1})}, (93)

where F(ξ)F(\xi) is the number of fixed points, and C(ξ)C(\xi) is the number of cycles in ξSn\xi\in S_{n}. For z=1z=1,

𝔼X𝒢[|Per(X)|4]\displaystyle\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}(X)|^{4}] =σ,τ,α,βM(σ,τ,α,β)=σ,τ2C(στ1)=(n!)2(n+1).\displaystyle=\sum_{\sigma,\tau,\alpha,\beta}M(\sigma,\tau,\alpha,\beta)=\sum_{\sigma,\tau}2^{C(\sigma\tau^{-1})}=(n!)^{2}(n+1). (94)

Here we have applied the fact that ξSn2C(ξ)=(n+1)!\sum_{\xi\in S_{n}}2^{C(\xi)}=(n+1)!. ∎

However, we can not get the explicit formula for the 𝔼X𝒢[|Perz(X)|4]\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{4}] because when στ=αβ\sigma\cup\tau=\alpha\cup\beta, the inversion number can be nonzero: namely (σ)+(τ)(α)(β)0\ell(\sigma)+\ell(\tau)-\ell(\alpha)-\ell(\beta)\neq 0. But we can still get the following result:

Lemma A.3.

For z𝕋z\in\mathbb{T}, 𝔼X𝒢[|Perz(X)|4]𝔼X𝒢[|Per(X)|4].\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{4}]\leq\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}(X)|^{4}].

The proof is simple since M(σ,τ,α,β)0M(\sigma,\tau,\alpha,\beta)\geq 0 and 𝔼X𝒢[|Perz(X)|4]\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{4}] must be a positive number. So we have

𝔼X𝒢[|Perz(X)|4]\displaystyle\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{4}] =σ,τ,α,βz(σ)+(τ)(α)(β)M(σ,τ,α,β)\displaystyle=\sum_{\sigma,\tau,\alpha,\beta}z^{\ell(\sigma)+\ell(\tau)-\ell(\alpha)-\ell(\beta)}\cdot M(\sigma,\tau,\alpha,\beta)
σ,τ,α,βM(σ,τ,α,β)\displaystyle\leq\sum_{\sigma,\tau,\alpha,\beta}M(\sigma,\tau,\alpha,\beta)
=𝔼X𝒢[|Per(X)|4].\displaystyle=\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}(X)|^{4}]. (95)

More generally, we have the following inequality.

Lemma A.4.

For z𝕋z\in\mathbb{T} and k2k\geq 2, 𝔼X𝒢[|Perz(X)|2k]𝔼X𝒢[|Per(X)|2k].\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{2k}]\leq\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}(X)|^{2k}].

Due to the lack of efficient formula for computing permanent, showing anti-concentration conjecture of permanent is hard. As we show in the paper, There is no efficient formula for general zz-permanent. So we believe that showing the anti-concentration conjecture of zz-permanent is also a hard problem. However, the higher moment of zz-permanent is less than permanent (but with the same first moment). This gives us a hint that anti-concentration is also true for zz-permanent. We have the following similar as z=1z=1.

Theorem A.5 (Weak Anti-Concentration of zz-Permanent).

For α<1,\alpha<1,

PrX𝒢[|Perz(X)|2>αn!]>(1α)2n+1.\Pr_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{2}>\alpha\cdot n!]>\frac{(1-\alpha)^{2}}{n+1}. (96)
Proof.

By Chebyshev’s inequality using 𝔼X𝒢[|Perz(X)|2]=n!\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{2}]=n! and 𝔼X𝒢[|Perz(X)|4]n!2(n+1)\operatorname*{\mathbb{E}}_{X\sim\mathcal{G}}[|\mathrm{Per}_{z}(X)|^{4}]\leq n!^{2}(n+1), the bound immediately follows. ∎

Appendix B Strong Autocorrelation Property

Consider a random variable XX with a real distribution \mathcal{F} with mean 0 and variance 1.1. We denote it as X(0,1).X\sim\mathcal{F}_{\mathbb{R}}(0,1). We define (μ,σ2)\mathcal{F}_{\mathbb{R}}(\mu,\sigma^{2}) with mean μ\mu and variance σ2\sigma^{2} by shifting and scale the random variable Xμ+σ2X.X\to\mu+\sigma^{2}X. We denote such distribution (μ,σ2)\mathcal{F}_{\mathbb{R}}(\mu,\sigma^{2}). We pick ϵ\epsilon such that ϵ<12.\epsilon<\frac{1}{2}.

X\displaystyle X =(0,1)n\displaystyle=\mathcal{F}_{\mathbb{R}}(0,1)^{n} (97)
𝒟1\displaystyle\mathcal{D}_{1} =(0,(1ε)2)n,\displaystyle=\mathcal{F}_{\mathbb{R}}\mathopen{}\mathclose{{}\left(0,\mathopen{}\mathclose{{}\left(1-\varepsilon}\right)^{2}}\right)^{n}, (98)
𝒟2\displaystyle\mathcal{D}_{2} =i=1n(vi,1)\displaystyle=\prod_{i=1}^{n}\mathcal{F}_{\mathbb{R}}\mathopen{}\mathclose{{}\left(v_{i},1}\right) (99)

We define

G(ϵ)=(0,1)(0,(1ϵ)2),H(vi)=(0,1)(vi,1).G(\epsilon)=\mathopen{}\mathclose{{}\left\|\mathcal{F}_{\mathbb{R}}(0,1)-\mathcal{F}_{\mathbb{R}}(0,(1-\epsilon)^{2})}\right\|,H(v_{i})=\mathopen{}\mathclose{{}\left\|\mathcal{F}_{\mathbb{R}}(0,1)-\mathcal{F}_{\mathbb{R}}(v_{i},1)}\right\|. (100)
Lemma B.1.

If G(ϵ),dG(ϵ)dϵG(\epsilon),\frac{dG(\epsilon)}{d\epsilon} are both continuous when ϵ[0,12]\epsilon\in[0,\frac{1}{2}] and H(x),dH(x)dxH(x),\frac{dH(x)}{dx} are both continuous when x[0,1]x\in[0,1]. Then,

𝒟1X\displaystyle\mathopen{}\mathclose{{}\left\|\mathcal{D}_{1}-X}\right\| nM1ϵ,ϵ[0,1/2]\displaystyle\leq nM_{1}\epsilon,\forall\epsilon\in[0,1/2]
𝒟2X\displaystyle\mathopen{}\mathclose{{}\left\|\mathcal{D}_{2}-X}\right\| M2v1,\displaystyle\leq M_{2}\mathopen{}\mathclose{{}\left\|v}\right\|_{1},\,

where M1,M2M_{1},M_{2} are independent of ϵ,vi\epsilon,v_{i} for every ii.

Proof.

We first prove the first inequality:

𝒟1Xn(0,1)(0,(1ϵ)2)=n|G(ϵ)|nϵM1.\displaystyle\mathopen{}\mathclose{{}\left\|\mathcal{D}_{1}-X}\right\|\leq n\mathopen{}\mathclose{{}\left\|\mathcal{F}_{\mathbb{R}}(0,1)-\mathcal{F}_{\mathbb{R}}(0,(1-\epsilon)^{2})}\right\|=n|G(\epsilon)|\leq n\epsilon M_{1}.

In order to bound the difference, we use the mean value theorem on variable ϵ\epsilon. Notice that G(0)=0G(0)=0. For all ϵ[0,12]\epsilon\in[0,\frac{1}{2}] we have:

|G(ϵ)G(0)|ϵmaxϵ[0,12]|dG(ϵ)dϵ|ϵM1.|G(\epsilon)-G(0)|\leq\epsilon\max_{\epsilon\in[0,\frac{1}{2}]}\mathopen{}\mathclose{{}\left|\frac{dG(\epsilon)}{d\epsilon}}\right|\leq\epsilon M_{1}. (101)

Here M1M_{1} is defined to be |dG(ϵ)dϵ|\mathopen{}\mathclose{{}\left|\frac{dG(\epsilon)}{d\epsilon}}\right|. The existence is guaranteed due to the continuity of |dG(ϵ)dϵ|\mathopen{}\mathclose{{}\left|\frac{dG(\epsilon)}{d\epsilon}}\right| and the following lemma.

Lemma B.2 (Extreme value theorem [R+76]).

A continuous function from a non-empty compact space to a subset of the real numbers attains a maximum and a minimum.

So finish the proof of the first inequality.

The second inequality follows using again the triangle inequality:

𝒟2Xi=1n(0,1)(vi,1)i=1n|vi|M2M2v1\displaystyle\mathopen{}\mathclose{{}\left\|\mathcal{D}_{2}-X}\right\|\leq\sum_{i=1}^{n}\mathopen{}\mathclose{{}\left\|\mathcal{F}_{\mathbb{R}}(0,1)-\mathcal{F}_{\mathbb{R}}(v_{i},1)}\right\|\leq\sum_{i=1}^{n}|v_{i}|M_{2}\leq M_{2}\mathopen{}\mathclose{{}\left\|v}\right\|_{1}

Here we use the same trick (notice H(0)=0H(0)=0, without loss of generality, we can assume vi0.v_{i}\geq 0.)

|H(vi)H(0)||vi|maxx[0,1]|dH(x)dx||vi|M2.|H(v_{i})-H(0)|\leq|v_{i}|\max_{x\in[0,1]}\mathopen{}\mathclose{{}\left|\frac{dH(x)}{dx}}\right|\leq|v_{i}|M_{2}. (102)

Here M2M_{2} is defined to be |dH(x)dx|\mathopen{}\mathclose{{}\left|\frac{dH(x)}{dx}}\right|. ∎

We can define complex random variable Z=X+iYZ=X+iY,X(μ1,1/2),Y(μ2,1/2)X\sim\mathcal{F}_{\mathbb{R}}(\mu_{1},1/2),Y\sim\mathcal{F}_{\mathbb{R}}(\mu_{2},1/2) and X,YX,Y are independent. So then Z(μ1+iμ2,1).Z\sim\mathcal{F}_{\mathbb{C}}(\mu_{1}+i\mu_{2},1). It will be helpful to think of each complex coordinate as two real coordinates and vv is a vector in 2N\mathbb{R}^{2N}.

Corollary B.3.

If we consider the complex version composed by real and with same condition:

X\displaystyle X =(0,1)n=(0,1/2)2n\displaystyle=\mathcal{F}_{\mathbb{C}}(0,1)^{n}=\mathcal{F}_{\mathbb{R}}(0,1/2)^{2n} (103)
𝒟1\displaystyle\mathcal{D}_{1} =(0,(1ε)2)n=(0,(1ϵ)2/2)2n,\displaystyle=\mathcal{F}_{\mathbb{C}}\mathopen{}\mathclose{{}\left(0,\mathopen{}\mathclose{{}\left(1-\varepsilon}\right)^{2}}\right)^{n}=\mathcal{F}_{\mathbb{R}}(0,(1-\epsilon)^{2}/2)^{2n}, (104)
𝒟2\displaystyle\mathcal{D}_{2} =i=1n(vi,1)\displaystyle=\prod_{i=1}^{n}\mathcal{F}_{\mathbb{C}}\mathopen{}\mathclose{{}\left(v_{i},1}\right) (105)

We have:

𝒟1X\displaystyle\mathopen{}\mathclose{{}\left\|\mathcal{D}_{1}-X}\right\| 2nM1ϵ,ϵ[0,12]\displaystyle\leq 2nM_{1}\epsilon,\forall\epsilon\in[0,\frac{1}{2}]
𝒟2X\displaystyle\mathopen{}\mathclose{{}\left\|\mathcal{D}_{2}-X}\right\| M2v12nM2v2.\displaystyle\leq M_{2}\mathopen{}\mathclose{{}\left\|v}\right\|_{1}\leq\sqrt{2n}M_{2}\mathopen{}\mathclose{{}\left\|v}\right\|_{2}.

In the final line, we use 11- and 22-norm inequalities. Notice that for the 11-norm, we imagine ν\nu as an 2n\mathbb{R}^{2n} vectors but for 22-norm, we pack them into n\mathbb{C}^{n}.

Notice that in [AA11], one can use rotational invariance to get the tighter bound for Gaussian matrices. But it is enough for our purpose.