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

Quantum Implications of Huang’s Sensitivity Theorem

Scott Aaronson111Department of Computer Science, University of Texas at Austin. [email protected]    Shalev Ben-David222University of Waterloo. [email protected]    Robin Kothari333Microsoft Quantum and Microsoft Research. [email protected]    Avishay Tal444Department of Electrical Engineering and Computer Sciences, University of California at Berkeley. [email protected]
Abstract

Based on the recent breakthrough of Huang (2019), we show that for any total Boolean function ff, the deterministic query complexity, 𝖣(f)\operatorname{\mathsf{D}}(f), is at most quartic in the quantum query complexity, 𝖰(f)\operatorname{\mathsf{Q}}(f): 𝖣(f)=O(𝖰(f)4)\operatorname{\mathsf{D}}(f)=O(\operatorname{\mathsf{Q}}(f)^{4}). This matches the known separation (up to log factors) due to Ambainis, Balodis, Belovs, Lee, Santha, and Smotrovs (2017). We also use the result to resolve the quantum analogue of the Aanderaa–Karp–Rosenberg conjecture. We show that if ff is a nontrivial monotone graph property of an nn-vertex graph specified by its adjacency matrix, then 𝖰(f)=Ω(n)\operatorname{\mathsf{Q}}(f)=\Omega(n), which is also optimal.

1 Introduction

Last year, Huang resolved a major open problem in the analysis of Boolean functions called the sensitivity conjecture [Hua19], which was open for nearly 30 years [NS94]. Surprisingly, Huang’s elegant proof takes less than 2 pages—truly a “proof from the book.” Specifically, Huang showed that for any total Boolean function, which is a function f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\}, we have

𝖽𝖾𝗀(f)𝗌(f)2,\operatorname{\mathsf{deg}}(f)\leq\operatorname{\mathsf{s}}(f)^{2}, (1)

where 𝖽𝖾𝗀(f)\operatorname{\mathsf{deg}}(f) is the real degree of ff and 𝗌(f)\operatorname{\mathsf{s}}(f) is the (maximum) sensitivity of ff. These measures and other measures appearing in this introduction are defined in Section 2.

In this note, we describe some implications of Huang’s resolution of the sensitivity conjecture to quantum query complexity. We observe that Huang actually proves a stronger claim, in which 𝗌(f)\operatorname{\mathsf{s}}(f) in Eq. 1 can be replaced by λ(f)\operatorname{\lambda}(f), a spectral relaxation of sensitivity that we define later. This observation has several implications for quantum query complexity.

We use this observation to settle the optimal relation between the deterministic query complexity, 𝖣(f)\operatorname{\mathsf{D}}(f), and quantum query complexity, 𝖰(f)\operatorname{\mathsf{Q}}(f), for total functions. We know from the seminal results of Nisan [Nis91], Nisan and Szegedy [NS94] and Beals et al. [BBC+01] that any total Boolean function ff satisfies555This means that for total functions, quantum query algorithms can only outperform classical query algorithms by a polynomial factor. On the other hand, for partial functions, which are defined on a subset of {0,1}n\{0,1\}^{n}, exponential and even larger speedups are possible.

𝖣(f)=O(𝖰(f)6).\operatorname{\mathsf{D}}(f)=O(\operatorname{\mathsf{Q}}(f)^{6}). (2)

Grover’s algorithm [Gro96] shows that for the or function, a quadratic separation between 𝖣\operatorname{\mathsf{D}} and 𝖰\operatorname{\mathsf{Q}} is possible. This was the best known quantum speedup for total functions until the work of Ambainis et al. [ABB+17], who constructed a total function ff with

𝖣(f)=Ω~(𝖰(f)4).\operatorname{\mathsf{D}}(f)=\widetilde{\Omega}(\operatorname{\mathsf{Q}}(f)^{4}). (3)

In this note, we show that the quartic separation (up to log factors) in Eq. 3 is actually the best possible:

Theorem 1.

For all Boolean functions f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\}, we have 𝖣(f)=O(𝖰(f)4)\operatorname{\mathsf{D}}(f)=O(\operatorname{\mathsf{Q}}(f)^{4}).

We deduce Theorem 1 as a corollary of a new tight quadratic relationship between 𝖽𝖾𝗀(f)\operatorname{\mathsf{deg}}(f) and 𝖰(f)\operatorname{\mathsf{Q}}(f):

Theorem 2.

For all Boolean functions f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\}, we have 𝖽𝖾𝗀(f)=O(𝖰(f)2)\operatorname{\mathsf{deg}}(f)=O(\operatorname{\mathsf{Q}}(f)^{2}).

Observe that Theorem 2 is tight for the or function on nn variables, whose degree is nn and whose quantum query complexity is Θ(n)\Theta(\sqrt{n}) [Gro96, BBBV97]. Prior to this work, the best relation between 𝖽𝖾𝗀(f)\operatorname{\mathsf{deg}}(f) and 𝖰(f)\operatorname{\mathsf{Q}}(f) was a sixth power relation, 𝖽𝖾𝗀(f)=O(𝖰(f)6)\operatorname{\mathsf{deg}}(f)=O(\operatorname{\mathsf{Q}}(f)^{6}), which follows from Eq. 2.

As discussed earlier, our proof relies on the restatement of Huang’s result (Theorem 5), showing that 𝖽𝖾𝗀(f)λ(f)2\operatorname{\mathsf{deg}}(f)\leq\operatorname{\lambda}(f)^{2}, where λ(f)\operatorname{\lambda}(f) is the spectral relaxation of sensitivity defined in Section 3. We then show that the measure λ(f)\operatorname{\lambda}(f) lower bounds the original quantum adversary method of Ambainis [Amb02], which in turn lower bounds 𝖰(f)\operatorname{\mathsf{Q}}(f).

We now show how Theorem 1 straightforwardly follows from Theorem 2 using two previously known connections between complexity measures of Boolean functions.

Proof of Theorem 1 assuming Theorem 2.

In [Mid04], Midrijanis showed that for all total functions ff, we have

𝖣(f)𝖻𝗌(f)𝖽𝖾𝗀(f),\operatorname{\mathsf{D}}(f)\leq\operatorname{\mathsf{bs}}(f)\operatorname{\mathsf{deg}}(f), (4)

where 𝖻𝗌(f)\operatorname{\mathsf{bs}}(f) is the block sensitivity of ff.

Theorem 2 shows that 𝖽𝖾𝗀(f)=O(𝖰(f)2)\operatorname{\mathsf{deg}}(f)=O(\operatorname{\mathsf{Q}}(f)^{2}). Combining the relationship between block sensitivity and approximate degree from [NS94] with the results of [BBC+01], we get that 𝖻𝗌(f)=O(𝖰(f)2)\operatorname{\mathsf{bs}}(f)=O(\operatorname{\mathsf{Q}}(f)^{2}). (This can also be proved directly using the lower bound method in [BBBV97].)

Combining these three inequalities yields 𝖣(f)=O(𝖰(f)4)\operatorname{\mathsf{D}}(f)=O(\operatorname{\mathsf{Q}}(f)^{4}) for all total Boolean functions ff. ∎

It remains to show the main result, Theorem 2, which we do in Section 3 using the proof of the sensitivity conjecture by Huang [Hua19] and the spectral adversary method in quantum query complexity [BSS03].

In Section 4, we also use Theorem 2 to prove the quantum analogue of the famous Aanderaa–Karp–Rosenberg conjecture. Briefly, this conjecture is about the minimum possible query complexity of a nontrivial monotone graph property, for graphs specified by their adjacency matrices.

There are variants of the conjecture for different models of computation. For example, the randomized variant of the Aanderaa–Karp–Rosenberg conjecture, attributed to Karp [SW86, Conjecture 1.2] and Yao [Yao77, Remark (2)], states that for all nontrivial monotone graph properties ff, we have 𝖱(f)=Ω(n2)\operatorname{\mathsf{R}}(f)=\Omega(n^{2}). Following a long line of work, the current best lower bound is 𝖱(f)=Ω(n4/3log1/3n)\operatorname{\mathsf{R}}(f)=\Omega(n^{4/3}\log^{1/3}n) due to Chakrabarti and Khot [CK01].

The quantum version of the conjecture was raised by Buhrman, Cleve, de Wolf, and Zalka [BCdWZ99], who observed that the best one could hope for is 𝖰(f)=Ω(n)\operatorname{\mathsf{Q}}(f)=\Omega(n), because the nontrivial monotone graph property “contains at least one edge” can be decided with O(n)O(n) queries using Grover’s algorithm [Gro96]. Buhrman et al. [BCdWZ99] also showed that all nontrivial monotone graph properties ff satisfy 𝖰(f)=Ω(n)\operatorname{\mathsf{Q}}(f)=\Omega(\sqrt{n}). The current best lower bound is 𝖰(f)=Ω(n2/3log1/6n)\operatorname{\mathsf{Q}}(f)=\Omega(n^{2/3}\log^{1/6}n), which was credited to Yao in [MSS07]. We resolve this conjecture by showing an optimal Ω(n)\Omega(n) lower bound.

Theorem 3.

Let f:{0,1}(n2){0,1}f:\{0,1\}^{\binom{n}{2}}\to\{0,1\} be a nontrivial monotone graph property. Then 𝖰(f)=Ω(n)\operatorname{\mathsf{Q}}(f)=\Omega(n).

Theorem 3 follows by combining Theorem 2 with a known quadratic lower bound on the degree of monotone graph properties.

1.1 Known relations and separations

𝖣\operatorname{\mathsf{D}}𝖱0\operatorname{\mathsf{R}}_{0}𝖰E\operatorname{\mathsf{Q}}_{E}𝖢\operatorname{\mathsf{C}}𝖱\operatorname{\mathsf{R}}𝖱𝖢\operatorname{\mathsf{RC}}𝖻𝗌\operatorname{\mathsf{bs}}𝗌\operatorname{\mathsf{s}}λ\operatorname{\lambda}𝖽𝖾𝗀\operatorname{\mathsf{deg}}𝖰\operatorname{\mathsf{Q}}𝖽𝖾𝗀~\operatorname{\mathsf{\widetilde{deg}}}
Figure 1: Relations between complexity measures. An upward line from a measure M1(f)M_{1}(f) to M2(f)M_{2}(f) denotes M1(f)=O(M2(f))M_{1}(f)=O(M_{2}(f)) for all total functions ff.

Table 1 summarizes the known relations and separations between complexity measures studied in this paper (and more). This is an update to a similar table that appears in [ABK16] with the addition of 𝗌(f)\operatorname{\mathsf{s}}(f) and λ(f)\operatorname{\lambda}(f). Definitions and additional details about interpreting the table can be found in [ABK16].

For all the separations claimed in the table, we provide either an example of a separating function or a citation to a result that constructs such a function. All the relationships in the table follow by combining the relationships depicted in Figure 1 and the following inequalities that hold for all total Boolean functions:

  • 𝖢(f)𝖻𝗌(f)𝗌(f)\operatorname{\mathsf{C}}(f)\leq\operatorname{\mathsf{bs}}(f)\operatorname{\mathsf{s}}(f) [Nis91]

  • 𝖣(f)𝖻𝗌(f)𝖢(f)\operatorname{\mathsf{D}}(f)\leq\operatorname{\mathsf{bs}}(f)\operatorname{\mathsf{C}}(f) [BBC+01]

  • 𝖣(f)𝖻𝗌(f)𝖽𝖾𝗀(f)\operatorname{\mathsf{D}}(f)\leq\operatorname{\mathsf{bs}}(f)\operatorname{\mathsf{deg}}(f) [Mid04]

  • 𝖱𝖢(f)=O(𝖽𝖾𝗀~(f)2)\operatorname{\mathsf{RC}}(f)=O(\operatorname{\mathsf{\widetilde{deg}}}(f)^{2}) [KT16]

  • 𝖱0(f)=O(𝖱(f)𝗌(f)log𝖱𝖢(f))\operatorname{\mathsf{R}}_{0}(f)=O(\operatorname{\mathsf{R}}(f)\operatorname{\mathsf{s}}(f)\log\operatorname{\mathsf{RC}}(f)) [KT16]

  • 𝖽𝖾𝗀(f)λ(f)2\operatorname{\mathsf{deg}}(f)\leq\operatorname{\lambda}(f)^{2} [Hua19]

  • 𝗌(f)λ(f)2\operatorname{\mathsf{s}}(f)\leq\operatorname{\lambda}(f)^{2} (Lemma 15)

Table 1: Best known separations between complexity measures
𝖣\operatorname{\mathsf{D}} 𝖱0\operatorname{\mathsf{R}}_{0} 𝖱\operatorname{\mathsf{R}} 𝖢\operatorname{\mathsf{C}} 𝖱𝖢\operatorname{\mathsf{RC}} 𝖻𝗌\operatorname{\mathsf{bs}} 𝗌\operatorname{\mathsf{s}} λ\operatorname{\lambda} 𝖰E\operatorname{\mathsf{Q}}_{E} 𝖽𝖾𝗀\operatorname{\mathsf{deg}} 𝖰\operatorname{\mathsf{Q}} 𝖽𝖾𝗀~\operatorname{\mathsf{\widetilde{deg}}}
𝖣\operatorname{\mathsf{D}}
2, 2
[ABB+17]
2, 3
[ABB+17]
2, 2
\wedge\circ\vee
2, 3
\wedge\circ\vee
2, 3
\wedge\circ\vee
3,6
[BHT17]
4,6
[ABB+17]
2, 3
[ABB+17]
2, 3
[GPW18]
4,4
[ABB+17]
4, 6
[ABB+17]
𝖱0\operatorname{\mathsf{R}}_{0}
1, 1
\oplus
2, 2
[ABB+17]
2, 2
\wedge\circ\vee
2, 3
\wedge\circ\vee
2, 3
\wedge\circ\vee
3,6
[BHT17]
3,6
[BHT17]
2, 3
[ABB+17]
2, 3
[GJPW18]
3,4
[ABB+17]
4, 6
[ABB+17]
𝖱\operatorname{\mathsf{R}}
1, 1
\oplus
1, 1
\oplus
2, 2
\wedge\circ\vee
2, 3
\wedge\circ\vee
2, 3
\wedge\circ\vee
3,6
[BHT17]
3,6
[BHT17]
32\frac{3}{2}, 3
[ABB+17]
2, 3
[GJPW18]
83\frac{8}{3},4
[Tal19]
4, 6
[ABB+17]
𝖢\operatorname{\mathsf{C}}
1, 1
\oplus
1, 1
\oplus
1, 2
\oplus
2, 2
[GSS13]
2, 2
[GSS13]
2.22,5
[BHT17]
2.22,6
[BHT17]
1.15, 3
[Amb13]
1.63, 3
[NW95]
2, 4
\wedge
2, 4
\wedge
𝖱𝖢\operatorname{\mathsf{RC}}
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
32\frac{3}{2}, 2
[GSS13]
2,4
[Rub95]
2,4
\wedge
1.15, 2
[Amb13]
1.63, 2
[NW95]
2, 2
\wedge
2, 2
\wedge
𝖻𝗌\operatorname{\mathsf{bs}}
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
2,4
[Rub95]
2,4
\wedge
1.15, 2
[Amb13]
1.63, 2
[NW95]
2, 2
\wedge
2, 2
\wedge
𝗌\operatorname{\mathsf{s}}
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
2,2
\wedge
1.15, 2
[Amb13]
1.63, 2
[NW95]
2, 2
\wedge
2, 2
\wedge
λ\operatorname{\lambda}
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
1, 2
\oplus
1, 1
\oplus
1, 2
\oplus
𝖰E\operatorname{\mathsf{Q}}_{E}
1, 1
\oplus
1.33, 2
¯\bar{\wedge}-tree
1.33, 3
¯\bar{\wedge}-tree
2, 2
\wedge\circ\vee
2, 3
\wedge\circ\vee
2, 3
\wedge\circ\vee
3,6
[BHT17]
3,6
[BHT17]
2, 3
[ABK16]
2,4
\wedge
4, 6
[ABK16]
𝖽𝖾𝗀\operatorname{\mathsf{deg}}
1, 1
\oplus
1.33, 2
¯\bar{\wedge}-tree
1.33,2
¯\bar{\wedge}-tree
2, 2
\wedge\circ\vee
2,2
\wedge\circ\vee
2,2
\wedge\circ\vee
2,2
\wedge\circ\vee
2,2
\wedge
1, 1
\oplus
2,2
\wedge
2,4
\wedge
𝖰\operatorname{\mathsf{Q}}
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
2, 2
[ABK16]
2, 3
[ABK16]
2, 3
[ABK16]
3,6
[BHT17]
3,6
[BHT17]
1, 1
\oplus
2, 3
[ABK16]
4, 6
[ABK16]
𝖽𝖾𝗀~\operatorname{\mathsf{\widetilde{deg}}}
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
22, 2
[BT17]
22,2
[BT17]
22,2
[BT17]
2,2
[BT17]
2,2
[BT17]
1, 1
\oplus
1, 1
\oplus
1, 1
\oplus
  • An entry a,ba,b in the row M1M_{1} and column M2M_{2} roughly means that for all total functions ff, M1(f)M2(f)b+o(1)M_{1}(f)\leq M_{2}(f)^{b+o(1)} and there exists a function gg with M1(g)M2(g)ao(1)M_{1}(g)\geq M_{2}(g)^{a-o(1)} (see [ABK16] for a precise definition).

  • The second row of each cell contains an example of a function that achieves the separation (or a citation to an example), where =parity\oplus=\textsc{parity}, =and\wedge=\textsc{and}, =or\vee=\textsc{or}, =and-or\wedge\circ\vee=\textsc{and-or}, and ¯\bar{\wedge}-tree is the balanced nand-tree function.

  • Cells have a white background if the relationship is optimal and a gray background otherwise.

  • Entries with a green background follow from Huang’s result. Entries with a red background follow from this work.

1.2 Paper organization

Section 2 contains some preliminaries required to understand the proof of Theorem 2, which is proved in Section 3. Section 4 gives some background and motivation for the Aanderaa–Karp–Rosenberg conjecture and proves Theorem 3. We end with some open problems in Section 5.

Appendix A describes some properties of λ(f)\operatorname{\lambda}(f), its many equivalent formulations, and its relationship with other complexity measures.

2 Preliminaries

2.1 Query complexity

Let f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\} be a Boolean function. Let AA be a deterministic algorithm that computes f(x)f(x) on input x{0,1}nx\in\{0,1\}^{n} by making queries to the bits of xx. The worst-case number of queries AA makes (over choices of xx) is the query complexity of AA. The minimum query complexity of any deterministic algorithm computing ff is the deterministic query complexity of ff, denoted by 𝖣(f)\operatorname{\mathsf{D}}(f).

We define the bounded-error randomized (respectively quantum) query complexity of ff, denoted by 𝖱(f)\operatorname{\mathsf{R}}(f) (respectively 𝖰(f)\operatorname{\mathsf{Q}}(f)), in an analogous way. We say an algorithm AA computes ff with bounded error if 𝐏𝐫[A(x)=f(x)]2/3\mathop{\bf Pr\/}[A(x)=f(x)]\geq 2/3 for all x{0,1}nx\in\{0,1\}^{n}, where the probability is over the internal randomness of AA. Then 𝖱(f)\operatorname{\mathsf{R}}(f) (respectively 𝖰(f)\operatorname{\mathsf{Q}}(f)) is the minimum number of queries required by any randomized (respectively quantum) algorithm that computes ff with bounded error. It is clear that 𝖰(f)𝖱(f)𝖣(f)\operatorname{\mathsf{Q}}(f)\leq\operatorname{\mathsf{R}}(f)\leq\operatorname{\mathsf{D}}(f). For more details on these measures, see the survey by Buhrman and de Wolf [BDW02].

2.2 Sensitivity and block sensitivity

Let f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\} be a Boolean function, and let x{0,1}nx\in\{0,1\}^{n} be a string. A block is a subset of [n][n]. We say that a block B[n]B\in[n] is sensitive for xx (with respect to ff) if f(x𝟙𝔹)𝕗(𝕩)f(x\oplus\mathbbold{1}_{B})\neq f(x), where 𝟙𝔹\mathbbold{1}_{B} is the nn-bit string that is 11 on bits in BB and 0 otherwise. We say a bit ii is sensitive for xx if the block {i}\{i\} is sensitive for xx. The maximum number of disjoint blocks that are all sensitive for xx is called the block sensitivity of xx (with respect to ff), denoted by 𝖻𝗌x(f)\operatorname{\mathsf{bs}}_{x}(f). The number of sensitive bits for xx is called the sensitivity of xx, denoted by 𝗌x(f)\operatorname{\mathsf{s}}_{x}(f). Clearly, 𝖻𝗌x(f)𝗌x(f)\operatorname{\mathsf{bs}}_{x}(f)\geq\operatorname{\mathsf{s}}_{x}(f), since 𝗌x(f)\operatorname{\mathsf{s}}_{x}(f) is has the same definition as 𝖻𝗌x(f)\operatorname{\mathsf{bs}}_{x}(f) except that the size of the blocks is restricted to 11. We define 𝗌(f)=maxx{0,1}n𝗌x(f)\operatorname{\mathsf{s}}(f)=\max_{x\in\{0,1\}^{n}}{\operatorname{\mathsf{s}}_{x}(f)} and 𝖻𝗌(f)=maxx{0,1}n𝖻𝗌x(f)\operatorname{\mathsf{bs}}(f)=\max_{x\in\{0,1\}^{n}}{\operatorname{\mathsf{bs}}_{x}(f)}.

2.3 Degree measures

A polynomial q[x1,,xn]q\in{\mathbb{R}}[x_{1},\ldots,x_{n}] is said to represent the function f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\} if q(x)=f(x)q(x)=f(x) for all x{0,1}nx\in\{0,1\}^{n}. A polynomial qq is said to ε\varepsilon-approximate ff if q(x)[0,ε]q(x)\in[0,\varepsilon] for all xf1(0)x\in f^{-1}(0) and q(x)[1ε,1]q(x)\in[1-\varepsilon,1] for all xf1(1)x\in f^{-1}(1). The degree of ff, denoted by 𝖽𝖾𝗀(f)\operatorname{\mathsf{deg}}(f), is the minimum degree of a polynomial representing ff. The ε\varepsilon-approximate degree, denoted by 𝖽𝖾𝗀~ε(f)\widetilde{\operatorname{\mathsf{deg}}}_{\varepsilon}(f), is the minimum degree of a polynomial ε\varepsilon-approximating ff. We will omit ε\varepsilon when ε=1/3\varepsilon=1/3. We know that 𝖣(f)𝖽𝖾𝗀(f)\operatorname{\mathsf{D}}(f)\geq\operatorname{\mathsf{deg}}(f), 𝖱(f)𝖽𝖾𝗀~(f)\operatorname{\mathsf{R}}(f)\geq\widetilde{\operatorname{\mathsf{deg}}}(f), and 𝖰(f)𝖽𝖾𝗀~(f)/2\operatorname{\mathsf{Q}}(f)\geq\widetilde{\operatorname{\mathsf{deg}}}(f)/2.

The degree of ff as a polynomial is also called the Fourier-degree of ff, which equals max{|S|:|f^(S)|0}\max\{|S|:|\widehat{f}(S)|\neq 0\} where f^(S):=𝐄x[f(x)(1)iSxi]\widehat{f}(S):=\mathop{\bf E\/}_{x}[f(x)\cdot(-1)^{\sum_{i\in S}x_{i}}]. In particular, 𝖽𝖾𝗀(f)<n\operatorname{\mathsf{deg}}(f)<n if and only if ff agrees with the Parity function, parityn(x)=i=1nxi\textsc{parity}_{n}(x)=\oplus_{i=1}^{n}x_{i}, on exactly half of the inputs.

3 Proof of main result (Theorem 2)

Before proving Theorem 2, which is based on Huang’s proof, we reinterpret his result in terms of a new complexity measure of Boolean functions that we call λ(f)\operatorname{\lambda}(f): the spectral norm of the sensitivity graph of ff.

Definition 4 (Sensitivity Graph GfG_{f}, Spectral Sensitivity λ(f)\operatorname{\lambda}(f)).

Let f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\} be a Boolean function. The sensitivity graph of ff, Gf=(V,E)G_{f}=(V,E) is a subgraph of the Boolean hypercube, where V={0,1}nV=\{0,1\}^{n}, and E={(x,xei)V×V:i[n],f(x)f(xei)}E=\{(x,x\oplus e_{i})\in V\times V:i\in[n],f(x)\neq f(x\oplus e_{i})\}. That is, EE is the set of edges between neighbors on the hypercube that have different ff-value. Let AfA_{f} be the adjacency matrix of the graph GfG_{f}. We define the spectral sensitivity of ff as λ(f)=Af\operatorname{\lambda}(f)=\left\lVert A_{f}\right\rVert.

Note that because AfA_{f} is a real symmetric matrix, λ(f)\operatorname{\lambda}(f) is also the largest eigenvalue of AfA_{f}. Since GfG_{f} is bipartite, the largest and smallest eigenvalues of AfA_{f} are equal in magnitude.

Huang’s proof of the sensitivity conjecture can be divided into two steps:

  1. 1.

    f:𝖽𝖾𝗀(f)λ(f)2\forall{f}:\operatorname{\mathsf{deg}}(f)\leq\operatorname{\lambda}(f)^{2}

  2. 2.

    f:λ(f)𝗌(f)\forall{f}:\operatorname{\lambda}(f)\leq\operatorname{\mathsf{s}}(f)

The second step is the simple fact that the spectral norm of an adjacency matrix is at most the maximum degree of any vertex in the graph, which equals 𝗌(f)\operatorname{\mathsf{s}}(f) in this case.

We reprove the first claim, i.e., 𝖽𝖾𝗀(f)λ(f)2\operatorname{\mathsf{deg}}(f)\leq\operatorname{\lambda}(f)^{2}, for completeness.

Theorem 5 ([Hua19]).

For all Boolean functions f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\}, we have 𝖽𝖾𝗀(f)λ(f)2\operatorname{\mathsf{deg}}(f)\leq\operatorname{\lambda}(f)^{2}.

Proof.

Without loss of generality we can assume that 𝖽𝖾𝗀(f)=n\operatorname{\mathsf{deg}}(f)=n since otherwise we can restrict our attention to a subcube of dimension 𝖽𝖾𝗀(f)\operatorname{\mathsf{deg}}(f) in which the degree remains the same and the top eigenvalue is at most λ(f)\operatorname{\lambda}(f). Specifically, we can choose any monomial in the polynomial representing ff of degree 𝖽𝖾𝗀(f)\operatorname{\mathsf{deg}}(f) and set all the variables not appearing in this monomial to 0.

For ff with 𝖽𝖾𝗀(f)=n\operatorname{\mathsf{deg}}(f)=n, let V0={x{0,1}n:f(x)=parityn(x)}V_{0}=\{x\in\{0,1\}^{n}:f(x)=\textsc{parity}_{n}(x)\} and V1={x{0,1}n:f(x)parityn(x)}V_{1}=\{x\in\{0,1\}^{n}:f(x)\neq\textsc{parity}_{n}(x)\}. By the fact that 𝖽𝖾𝗀(f)=n\operatorname{\mathsf{deg}}(f)=n we know that |V0||V1||V_{0}|\neq|V_{1}| as otherwise ff would have 0 correlation with the nn-variate parity function, implying that ff’s top Fourier coefficient is 0.

We also note that any edge in the hypercube that goes between V0V_{0} and V0V_{0} is an edge in GfG_{f} since it changes the value of ff. This holds since for such an edge, (x,xei)(x,x\oplus e_{i}), we have f(x)=parityn(x)parityn(xei)=f(xei)f(x)=\textsc{parity}_{n}(x)\neq\textsc{parity}_{n}(x\oplus e_{i})=f(x\oplus e_{i}). Similarly, any edge in the hypercube that goes between V1V_{1} and V1V_{1} is an edge in GfG_{f}.

Assume without loss of generality that |V0|>|V1||V_{0}|>|V_{1}|. Thus, |V0|2n1+1|V_{0}|\geq 2^{n-1}+1. We will show that there exists a nonzero vector vv^{\prime} supported only on the entries of V0V_{0}, such that Afvnv\|A_{f}\cdot v^{\prime}\|\geq\sqrt{n}\cdot\|v^{\prime}\|.

Let G=(V,E)G=(V,E) be the complete nn-dimensional Boolean Hypercube. That is, V={0,1}nV=\{0,1\}^{n} and E={(x,xei):x{0,1}n,i[n]}E=\{(x,x\oplus e_{i})\;:\;x\in\{0,1\}^{n},i\in[n]\}. Take the following signing of the edges of the Boolean hypercube, defined recursively.

B1=(0110)andBi=(Bi1IIBi1)for i{2,,n}.B_{1}=\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\ \mathrm{and}\ B_{i}=\begin{pmatrix}B_{i-1}&I\\ I&-B_{i-1}\end{pmatrix}\ \text{for $i\in\{2,\ldots,n\}$}. (5)

This gives a new matrix Bn{1,0,1}V×VB_{n}\in\{-1,0,1\}^{V\times V} where Bn(x,y)=0B_{n}(x,y)=0 if and only if xx is not a neighbor of yy in the hypercube.

Huang showed that BnB_{n} has 2n/22^{n}/2 eigenvalues that equal n-\sqrt{n} and 2n/22^{n}/2 eigenvalues that equal +n+\sqrt{n}. To show this, he showed that Bn2=nIB_{n}^{2}=n\cdot I by induction on nn and thus all eigenvalues of BnB_{n} must be either +n+\sqrt{n} or n-\sqrt{n}. Then, observing that the trace of BnB_{n} is 0, as all diagonal entries equal 0, we see that we must have an equal number of +n+\sqrt{n} and n-\sqrt{n} eigenvalues.

Thus, the subspace of eigenvectors for BnB_{n} with eigenvalue n\sqrt{n} is of dimension 2n/22^{n}/2. Using |V1|<2n/2|V_{1}|<2^{n}/2, there must exists a nonzero eigenvector for BnB_{n} with eigenvalue n\sqrt{n} that vanishes on V1V_{1}. Fix vv to be any such vector.

Let vv^{\prime} be the vector whose entries are the absolute values of the entries of vv. We claim that Afv2nv2\|A_{f}\cdot v^{\prime}\|_{2}\geq\sqrt{n}\cdot\|v^{\prime}\|_{2}. To see so, note that for every xV0x\in V_{0} we have

(Afv)x\displaystyle(A_{f}\cdot v^{\prime})_{x} =yx:f(y)f(x)vy=yx:yV0vy=yxvy\displaystyle=\sum_{y\sim x:f(y)\neq f(x)}v^{\prime}_{y}=\sum_{y\sim x:y\in V_{0}}v^{\prime}_{y}=\sum_{y\sim x}v^{\prime}_{y}
y{0,1}n|Bx,yvy||y{0,1}nBx,yvy|=n|vx|=nvx.\displaystyle\geq\sum_{y\in\{0,1\}^{n}}|B_{x,y}v_{y}|\geq\left|\sum_{y\in\{0,1\}^{n}}B_{x,y}v_{y}\right|=\sqrt{n}\cdot|v_{x}|=\sqrt{n}\cdot v^{\prime}_{x}\;. (6)

On the other hand, for xV1x\in V_{1} we have (Afv)x=0=vx(A_{f}\cdot v^{\prime})_{x}=0=v^{\prime}_{x}. Thus the norm of AfvA_{f}\cdot v^{\prime} is at least n\sqrt{n} times the norm of vv^{\prime}, and hence λ(f)=Afn=𝖽𝖾𝗀(f)\operatorname{\lambda}(f)=\|A_{f}\|\geq\sqrt{n}=\sqrt{\operatorname{\mathsf{deg}}(f)}. ∎

Finally, we prove that λ(f)=O(𝖰(f))\operatorname{\lambda}(f)=O(\operatorname{\mathsf{Q}}(f)). We rely on a variant of the adversary method introduced by Barnum, Saks, and Szegedy [BSS03] (see also [SS06]).

Definition 6 (Spectral Adversary method).

Let {Di}i[n]\{D_{i}\}_{i\in[n]} and FF be matrices of size {0,1}n×{0,1}n\{0,1\}^{n}\times\{0,1\}^{n} with entries in {0,1}\{0,1\} satisfying Di[x,y]=1D_{i}[x,y]=1 if and only if xiyix_{i}\neq y_{i}, and F[x,y]=1F[x,y]=1 if and only if f(x)f(y)f(x)\neq f(y). Let Γ\Gamma denote a {0,1}n×{0,1}n\{0,1\}^{n}\times\{0,1\}^{n} nonnegative symmetric matrix such that ΓF=Γ\Gamma\circ F=\Gamma (i.e., the nonzero entries of Γ\Gamma are a subset of the the nonzero entries of FF). Then 𝖲𝖠(f)=maxΓΓmaxi[n]ΓDi\operatorname{\mathsf{SA}}(f)=\max_{\Gamma}\frac{\|\Gamma\|}{\max_{i\in[n]}{\|\Gamma\circ D_{i}\|}}.

Barnum, Saks, and Szegedy [BSS03] proved that 𝖰(f)=Ω(𝖲𝖠(f))\operatorname{\mathsf{Q}}(f)=\Omega(\operatorname{\mathsf{SA}}(f)).

Lemma 7.

For all Boolean functions 𝖰(f)=Ω(𝖲𝖠(f))=Ω(λ(f))\operatorname{\mathsf{Q}}(f)=\Omega(\operatorname{\mathsf{SA}}(f))=\Omega(\operatorname{\lambda}(f)).

Proof.

We prove that 𝖲𝖠(f)λ(f)\operatorname{\mathsf{SA}}(f)\geq\operatorname{\lambda}(f). Indeed, one can take Γ\Gamma to be simply the adjacency matrix of GfG_{f}. That is, for any x,y{0,1}nx,y\in\{0,1\}^{n} put Γ[x,y]=1\Gamma[x,y]=1 if and only if yxy\sim x in the hypercube and f(x)f(y)f(x)\neq f(y). We observe that Γ=λ(f)\|\Gamma\|=\operatorname{\lambda}(f). On the other hand, for any i[n]i\in[n], ΓDi\Gamma\circ D_{i} is the restriction of the sensitive edges in direction ii. The maximum degree in the graph represented by ΓDi\Gamma\circ D_{i} is 11 hence ΓDi\|\Gamma\circ D_{i}\| is at most 11. Thus we have

𝖲𝖠(f)Γmaxi[n]ΓDiλ(f).\operatorname{\mathsf{SA}}(f)\geq\frac{\|\Gamma\|}{\max_{i\in[n]}{\|\Gamma\circ D_{i}\|}}\geq\operatorname{\lambda}(f). (7)

Combining this with 𝖰(f)=Ω(𝖲𝖠(f))\operatorname{\mathsf{Q}}(f)=\Omega(\operatorname{\mathsf{SA}}(f)) [BSS03], we get 𝖰(f)=Ω(𝖲𝖠(f))=Ω(λ(f))\operatorname{\mathsf{Q}}(f)=\Omega(\operatorname{\mathsf{SA}}(f))=\Omega(\operatorname{\lambda}(f)). ∎

From Theorem 5 and Lemma 7 we immediately get Theorem 2.

4 Monotone graph properties

The Aanderaa–Karp–Rosenberg conjectures are a collection of conjectures related to the query complexity of deciding whether an input graph specified by its adjacency matrix satisfies a given property in various models of computation.

Specifically, let the input be an nn-vertex undirected simple graph specified by its adjacency matrix. This means we can query any unordered pair {i,j}\{i,j\}, where i,j[n]i,j\in[n], and learn whether there is an edge between vertex ii and jj. Note that the input size is (n2)=Θ(n2)\binom{n}{2}=\Theta(n^{2}).

A function ff on (n2)\binom{n}{2} variables is a graph property if it treats the input as a graph and not merely a string of length (n2)\binom{n}{2}. Specifically, the function must be invariant under permuting vertices of the graph. In other words, the function can only depend on the isomorphism class of the graph, not the specific labels of the vertices. A function ff is monotone (increasing) if for all x,y{0,1}nx,y\in\{0,1\}^{n}, xyf(x)f(y)x\leq y\implies f(x)\leq f(y), where xyx\leq y means xiyix_{i}\leq y_{i} for all i[n]i\in[n]. For a monotone function, negating a 0 in the input cannot change the function value from 11 to 0. In the context of graph properties, if the input graph has a certain monotone graph property, then adding more edges cannot destroy the property.

Examples of monotone graph properties include “GG is connected,” “GG contains a clique of size kk,” “GG contains a Hamiltonian cycle,” “GG has chromatic number greater than kk,” “GG is not planar”, and “GG has diameter at most kk.” Many commonly encountered graph properties (or their negation) are monotone graph properties. Finally, we say a function f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\} is nontrivial if there exist inputs xx and yy such that f(x)f(y)f(x)\neq f(y).

The deterministic Aanderaa–Karp–Rosenberg conjecture, also called the evasiveness conjecture,666A function ff is called evasive if its deterministic query complexity equals its input size. states that for all nontrivial monotone graph properties ff, 𝖣(f)=(n2)\operatorname{\mathsf{D}}(f)=\binom{n}{2}. This conjecture remains open to this day, although the weaker claim that 𝖣(f)=Ω(n2)\operatorname{\mathsf{D}}(f)=\Omega(n^{2}) was proved over 40 years ago by Rivest and Vuillemin [RV76]. Several works have improved on the constant in their lower bound, and the best current result is due to Scheidweiler and Triesch [ST13], who prove a lower bound of 𝖣(f)(1/3o(1))n2\operatorname{\mathsf{D}}(f)\geq(1/3-o(1))\cdot n^{2}. The evasiveness conjecture has been established in several special cases including when nn is prime [KSS84] and when restricted to bipartite graphs [Yao88].

The randomized Aanderaa–Karp–Rosenberg conjecture asserts that all nontrivial monotone graph properties ff satisfy 𝖱(f)=Ω(n2)\operatorname{\mathsf{R}}(f)=\Omega(n^{2}). A sequence of increasingly stronger lower bounds, starting with a lower bound of Ω(nlog1/12n)\Omega(n\log^{1/12}n) due to Yao [Yao91], a lower bound of Ω(n5/4)\Omega(n^{5/4}) due to King [Kin88], and a lower bound of Ω(n4/3)\Omega(n^{4/3}) due to Hajnal [Haj91], has led to the current best lower bound of Ω(n4/3log1/3n)\Omega(n^{4/3}\log^{1/3}n) due to Chakrabarti and Khot [CK01]. There are also two lower bounds due to Friedgut, Kahn, and Wigderson [FKW02] and O’Donnell, Saks, Schramm, and Servedio [OSSS05] that are better than this bound for some graph properties.

The quantum Aanderaa–Karp–Rosenberg conjecture states that all nontrivial monotone graph properties ff satisfy 𝖰(f)=Ω(n)\operatorname{\mathsf{Q}}(f)=\Omega(n). This is the best lower bound one could hope to prove since there exist properties with 𝖰(f)=O(n)\operatorname{\mathsf{Q}}(f)=O(n), such as the property of containing at least one edge. In fact, for any α[1,2]\alpha\in[1,2] it is possible to construct a graph property with quantum query complexity Θ(nα)\Theta(n^{\alpha}) using known lower bounds for the threshold function [BBC+01].

As stated in the introduction, the question was first raised by Buhrman, Cleve, de Wolf, and Zalka [BCdWZ99], who showed a lower bound of Ω(n)\Omega(\sqrt{n}). This was improved by Yao to Ω(n2/3log1/6n)\Omega(n^{2/3}\log^{1/6}n) using the technique in [CK01] and Ambainis’ adversary bound [Amb02]. Better lower bounds are known in some special cases, such as when the property is a subgraph isomorphism property, where we know a lower bound of Ω(n3/4)\Omega(n^{3/4}) due to Kulkarni and Podder [KP16].

As stated in Theorem 3, we resolve the quantum Aanderaa–Karp–Rosenberg conjecture and show an optimal Ω(n)\Omega(n) lower bound. The proof combines Theorem 2 with a quadratic lower bound on the degree of nontrivial monotone graph properties. With some work, the original quadratic lower bound on the deterministic query complexity of nontrivial monotone graph properties by Rivest and Vuillemin [RV76] can be modified to prove a similar lower bound for degree. We were not able to find such a proof in the literature, and instead combine the following two claims to obtain the desired claim.

First, we use the result of Dodis and Khanna [DK99, Theorem 2]:

Theorem 8.

For all nontrivial monotone graph properties, 𝖽𝖾𝗀2(f)=Ω(n2)\operatorname{\mathsf{deg}}_{2}(f)=\Omega(n^{2}).

Here 𝖽𝖾𝗀2(f)\operatorname{\mathsf{deg}}_{2}(f) is the minimum degree of a Boolean function when represented as a polynomial over the finite field with two elements, 𝔽2\mathbb{F}_{2}. We combine this with a standard lemma that shows that this measure lower bounds 𝖽𝖾𝗀(f)\operatorname{\mathsf{deg}}(f). A proof can be found in [O’D09, Proposition 6.23]:

Lemma 9.

For all Boolean functions f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\}, we have 𝖽𝖾𝗀2(f)𝖽𝖾𝗀(f)\operatorname{\mathsf{deg}}_{2}(f)\leq\operatorname{\mathsf{deg}}(f).

Combining these with Theorem 2, we get that all nontrivial monotone graph properties ff satisfy 𝖰(f)=Ω(n)\operatorname{\mathsf{Q}}(f)=\Omega(n), which is the statement of Theorem 3.

5 Open questions

We saw that λ(f)\operatorname{\lambda}(f) lower-bounds both 𝖠𝖽𝗏(f)\operatorname{\mathsf{Adv}}(f), and thus 𝖰(f)\operatorname{\mathsf{Q}}(f), and also the sensitivity 𝗌(f)\operatorname{\mathsf{s}}(f). One might conjecture that λ(f)\operatorname{\lambda}(f) lower-bounds all the complexity measures in Figure 1, including 𝖽𝖾𝗀~(f)\widetilde{\operatorname{\mathsf{deg}}}(f).

Conjecture 1.

For all Boolean functions f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\}, we have λ(f)=O(𝖽𝖾𝗀~(f))\operatorname{\lambda}(f)=O(\widetilde{\operatorname{\mathsf{deg}}}(f)).

If 1 we true, Theorem 5 would imply that 𝖽𝖾𝗀(f)=O(𝖽𝖾𝗀~(f)2)\operatorname{\mathsf{deg}}(f)=O(\widetilde{\operatorname{\mathsf{deg}}}(f)^{2}), settling a longstanding conjecture posed by Nisan and Szegedy [NS94]. The current best relation between the two measures is 𝖽𝖾𝗀(f)=O(𝖽𝖾𝗀~(f)6)\operatorname{\mathsf{deg}}(f)=O(\widetilde{\operatorname{\mathsf{deg}}}(f)^{6}). The following conjecture is weaker, and might be easier to tackle first.

Conjecture 2.

For all Boolean functions f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\}, we have λ(f)=O(𝖽𝖾𝗀(f))\operatorname{\lambda}(f)=O(\operatorname{\mathsf{deg}}(f)).

Another longstanding open problem is to show a quadratic relation between deterministic query complexity and block sensitivity:

Conjecture 3.

For all Boolean functions f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\}, we have 𝖣(f)=O(𝖻𝗌(f)2)\operatorname{\mathsf{D}}(f)=O(\operatorname{\mathsf{bs}}(f)^{2}).

If this conjecture were true, it would optimally resolve several relationships in Table 1, and would imply, for example, 𝖣(f)=O(𝖱(f)2))\operatorname{\mathsf{D}}(f)=O(\operatorname{\mathsf{R}}(f)^{2})) and 𝖣(f)=O(𝖽𝖾𝗀~(f)4)\operatorname{\mathsf{D}}(f)=O(\operatorname{\mathsf{\widetilde{deg}}}(f)^{4}).

After settling the best relation between 𝖣(f)\operatorname{\mathsf{D}}(f) and 𝖰(f)\operatorname{\mathsf{Q}}(f), the next pressing question is to settle the best relation between 𝖱(f)\operatorname{\mathsf{R}}(f) and 𝖰(f)\operatorname{\mathsf{Q}}(f). Recently, the fourth author [Tal19] showed a power 8/38/3 separation between 𝖱(f)\operatorname{\mathsf{R}}(f) and 𝖰(f)\operatorname{\mathsf{Q}}(f), while the best known relationship is a power 44 relationship (this work). We conjecture that both these bounds can be improved.

Conjecture 4.

For all Boolean functions f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\}, we have 𝖱(f)=O(𝖰(f)3)\operatorname{\mathsf{R}}(f)=O(\operatorname{\mathsf{Q}}(f)^{3}).

Conjecture 5.

There exists a Boolean function f:{0,1}n{0,1}f:\{0,1\}^{n}\to\{0,1\} such that 𝖱(f)=Ω(𝖰(f)3)\operatorname{\mathsf{R}}(f)=\Omega(\operatorname{\mathsf{Q}}(f)^{3}).

We note that there are candidate constructions based on the work of [AA18, ABK16, Tal19] that are conjectured to satisfy 𝖰(f)𝖱(f)3o(1)\operatorname{\mathsf{Q}}(f)\geq\operatorname{\mathsf{R}}(f)^{3-o(1)}. In particular, it suffices to prove a conjectured bound on the Fourier spectrum of deterministic decision trees [Tal19] to prove 5.

Finally, for the special case of monotone total Boolean functions ff, Beals et al. [BBC+01] already showed in 1998 that 𝖣(f)=O(𝖰(f)4)\operatorname{\mathsf{D}}(f)=O(\operatorname{\mathsf{Q}}(f)^{4}). It would be interesting to know whether this can be improved, perhaps all the way to 𝖣(f)=O(𝖰(f)2)\operatorname{\mathsf{D}}(f)=O(\operatorname{\mathsf{Q}}(f)^{2}).

References

  • [AA18] Scott Aaronson and Andris Ambainis. Forrelation: A problem that optimally separates quantum from classical computing. SIAM J. Comput., 47(3):982–1038, 2018. doi:10.1137/15M1050902.
  • [ABB+17] Andris Ambainis, Kaspars Balodis, Aleksandrs Belovs, Troy Lee, Miklos Santha, and Juris Smotrovs. Separations in query complexity based on pointer functions. Journal of the ACM, 64(5):1–24, September 2017. doi:10.1145/3106234.
  • [ABK16] Scott Aaronson, Shalev Ben-David, and Robin Kothari. Separations in query complexity using cheat sheets. In Proceedings of the 48th Annual ACM SIGACT Symposium on Theory of Computing, STOC 2016, pages 863–876, 2016. doi:10.1145/2897518.2897644.
  • [Amb02] Andris Ambainis. Quantum lower bounds by quantum arguments. Journal of Computer and System Sciences, 64(4):750–767, jun 2002. doi:10.1006/jcss.2002.1826.
  • [Amb13] Andris Ambainis. Superlinear advantage for exact quantum algorithms. In Proceedings of the 45th ACM Symposium on Theory of Computing (STOC 2013), pages 891–900, 2013. doi:10.1145/2488608.2488721.
  • [BBBV97] Charles H. Bennett, Ethan Bernstein, Gilles Brassard, and Umesh Vazirani. Strengths and weaknesses of quantum computing. SIAM Journal on Computing, 26(5):1510–1523, October 1997. doi:10.1137/S0097539796300933.
  • [BBC+01] Robert Beals, Harry Buhrman, Richard Cleve, Michele Mosca, and Ronald de Wolf. Quantum lower bounds by polynomials. Journal of the ACM, 48(4):778–797, 2001. doi:10.1145/502090.502097.
  • [BCdWZ99] Harry Buhrman, Richard Cleve, Ronald de Wolf, and Christof Zalka. Bounds for small-error and zero-error quantum algorithms. In 40th Annual Symposium on Foundations of Computer Science, pages 358–368, 1999. doi:10.1109/SFFCS.1999.814607.
  • [BDW02] Harry Buhrman and Ronald De Wolf. Complexity measures and decision tree complexity: a survey. Theoretical Computer Science, 288(1):21–43, 2002. doi:10.1016/S0304-3975(01)00144-X.
  • [BHT17] Shalev Ben-David, Pooya Hatami, and Avishay Tal. Low-sensitivity functions from unambiguous certificates. In 8th Innovations in Theoretical Computer Science Conference, ITCS 2017, volume 67 of LIPIcs, pages 28:1–28:23. Schloss Dagstuhl - Leibniz-Zentrum für Informatik, 2017. doi:10.4230/LIPIcs.ITCS.2017.28.
  • [BSS03] Howard Barnum, Michael E. Saks, and Mario Szegedy. Quantum query complexity and semi-definite programming. In 18th Annual IEEE Conference on Computational Complexity (Complexity 2003), pages 179–193, 2003. doi:10.1109/CCC.2003.1214419.
  • [BT17] Mark Bun and Justin Thaler. A nearly optimal lower bound on the approximate degree of AC0. In IEEE 58th Annual Symposium on Foundations of Computer Science (FOCS 2017), pages 1–12, 2017. doi:10.1109/FOCS.2017.10.
  • [CK01] Amit Chakrabarti and Subhash Khot. Improved lower bounds on the randomized complexity of graph properties. In Automata, Languages and Programming, pages 285–296, 2001. doi:10.1007/3-540-48224-5_24.
  • [DK99] Yevgeniy Dodis and Sanjeev Khanna. Space-time tradeoffs for graph properties. In Automata, Languages and Programming, pages 291–300, 1999. doi:10.1007/3-540-48523-6_26.
  • [FKW02] Ehud Friedgut, Jeff Kahn, and Avi Wigderson. Computing graph properties by randomized subcube partitions. In Randomization and Approximation Techniques in Computer Science, pages 105–113, 2002. doi:10.1007/3-540-45726-7_9.
  • [GJPW18] Mika Göös, T. S. Jayram, Toniann Pitassi, and Thomas Watson. Randomized communication versus partition number. ACM Transactions on Computation Theory, 10(1):1–20, January 2018. doi:10.1145/3170711.
  • [GL13] Gene H. Golub and Charles F. Van Loan. Matrix computations. Johns Hopkins University Press, Baltimore, 2013. URL: https://jhupbooks.press.jhu.edu/title/matrix-computations.
  • [GPW18] Mika Göös, Toniann Pitassi, and Thomas Watson. Deterministic communication vs. partition number. SIAM Journal on Computing, 47(6):2435–2450, 2018. doi:10.1137/16M1059369.
  • [Gro96] Lov K. Grover. A fast quantum mechanical algorithm for database search. In Proceedings of the Twenty-Eighth Annual ACM Symposium on the Theory of Computing, STOC 1996, pages 212–219, 1996. doi:10.1145/237814.237866.
  • [GSS13] Justin Gilmer, Michael Saks, and Srikanth Srinivasan. Composition limits and separating examples for some Boolean function complexity measures. In Proceedings of 2013 IEEE Conference on Computational Complexity (CCC 2013), pages 185–196, June 2013. doi:10.1109/CCC.2013.27.
  • [Haj91] Péter Hajnal. An Ω(n4/3)\Omega(n^{4/3}) lower bound on the randomized complexity of graph properties. Combinatorica, 11(2):131–143, jun 1991. doi:10.1007/bf01206357.
  • [Hua19] Hao Huang. Induced subgraphs of hypercubes and a proof of the sensitivity conjecture. Annals of Mathematics, 190(3):949–955, 2019. doi:10.4007/annals.2019.190.3.6.
  • [Khr71] V. M. Khrapchenko. Method of determining lower bounds for the complexity of P-schemes. Mathematical Notes of the Academy of Sciences of the USSR, 10(1):474–479, 1971. doi:10.1007/bf01747074.
  • [Kin88] Valerie King. Lower bounds on the complexity of graph properties. In Proceedings of the Twentieth Annual ACM Symposium on Theory of Computing, STOC ’88, page 468–476. Association for Computing Machinery, 1988. doi:10.1145/62212.62258.
  • [Kou93] Elias Koutsoupias. Improvements on Khrapchenko’s theorem. Theoretical Computer Science, 116(2):399–403, aug 1993. doi:10.1016/0304-3975(93)90330-V.
  • [KP16] Raghav Kulkarni and Supartha Podder. Quantum Query Complexity of Subgraph Isomorphism and Homomorphism. In 33rd Symposium on Theoretical Aspects of Computer Science (STACS 2016), volume 47 of Leibniz International Proceedings in Informatics (LIPIcs), pages 48:1–48:13. Schloss Dagstuhl–Leibniz-Zentrum fuer Informatik, 2016. doi:10.4230/LIPIcs.STACS.2016.48.
  • [KSS84] Jeff Kahn, Michael Saks, and Dean Sturtevant. A topological approach to evasiveness. Combinatorica, 4(4):297–306, dec 1984. doi:10.1007/bf02579140.
  • [KT16] Raghav Kulkarni and Avishay Tal. On fractional block sensitivity. Chicago Journal of Theoretical Computer Science, 2016(8), July 2016. doi:10.4086/cjtcs.2016.008.
  • [LLS06] Sophie Laplante, Troy Lee, and Mario Szegedy. The quantum adversary method and classical formula size lower bounds. computational complexity, 15(2):163–196, jun 2006. doi:10.1007/s00037-006-0212-7.
  • [LNS20] Sophie Laplante, Reza Naserasr, and Anupa Sunny. Sensitivity lower bounds from linear dependencies. Technical Report TR20-002, Electronic Colloquium on Computational Complexity (ECCC), January 2020. URL: https://eccc.weizmann.ac.il/report/2020/002/.
  • [Mid04] Gatis Midrijanis. Exact quantum query complexity for total Boolean functions. arXiv preprint quant-ph/0403168, 2004. arXiv:quant-ph/0403168.
  • [MSS07] Frédéric Magniez, Miklos Santha, and Mario Szegedy. Quantum algorithms for the triangle problem. SIAM Journal on Computing, 37(2):413–424, 2007. doi:10.1137/050643684.
  • [Nis91] Noam Nisan. CREW prams and decision trees. SIAM J. Comput., 20(6):999–1007, 1991. doi:10.1137/0220062.
  • [NS94] Noam Nisan and Mario Szegedy. On the degree of Boolean functions as real polynomials. Computational Complexity, 4(4):301–313, dec 1994. doi:10.1007/BF01263419.
  • [NW95] Noam Nisan and Avi Wigderson. On rank vs. communication complexity. Combinatorica, 15(4):557–565, 1995. doi:10.1007/BF01192527.
  • [O’D09] Ryan O’Donnell. Analysis of Boolean Functions. Cambridge University Press, 2009. doi:10.1017/cbo9781139814782.
  • [OSSS05] Ryan O’Donnell, Michael Saks, Oded Schramm, and Rocco A. Servedio. Every decision tree has an influential variable. In 46th Annual IEEE Symposium on Foundations of Computer Science (FOCS’05), pages 31–39, 2005. doi:10.1109/SFCS.2005.34.
  • [Rub95] David Rubinstein. Sensitivity vs. block sensitivity of Boolean functions. Combinatorica, 15(2):297–299, 1995. doi:10.1007/BF01200762.
  • [RV76] Ronald L. Rivest and Jean Vuillemin. On recognizing graph properties from adjacency matrices. Theoretical Computer Science, 3(3):371 – 384, 1976. doi:10.1016/0304-3975(76)90053-0.
  • [SS06] Robert Spalek and Mario Szegedy. All quantum adversary methods are equivalent. Theory of Computing, 2(1):1–18, 2006. doi:10.4086/toc.2006.v002a001.
  • [ST13] Robert Scheidweiler and Eberhard Triesch. A lower bound for the complexity of monotone graph properties. SIAM Journal on Discrete Mathematics, 27(1):257–265, 2013. doi:10.1137/120888703.
  • [SW86] Michael Saks and Avi Wigderson. Probabilistic Boolean decision trees and the complexity of evaluating game trees. In Proceedings of the 27th Annual Symposium on Foundations of Computer Science, SFCS ’86, page 29–38, 1986. doi:10.1109/SFCS.1986.44.
  • [Tal13] Avishay Tal. Properties and applications of Boolean function composition. In Proceedings of the 4th Conference on Innovations in Theoretical Computer Science, ITCS ’13, pages 441–454, 2013. doi:10.1145/2422436.2422485.
  • [Tal19] Avishay Tal. Towards optimal separations between quantum and randomized query complexities. arXiv, 2019. arXiv:1912.12561.
  • [Yao77] Andrew Chi-Chih Yao. Probabilistic computations: Toward a unified measure of complexity. In 18th Annual Symposium on Foundations of Computer Science (SFCS 1977), pages 222–227, 1977. doi:10.1109/SFCS.1977.24.
  • [Yao88] Andrew Chi-Chih Yao. Monotone bipartite graph properties are evasive. SIAM Journal on Computing, 17(3):517–520, 1988. doi:10.1137/0217031.
  • [Yao91] Andrew Chi-Chih Yao. Lower bounds to randomized algorithms for graph properties. Journal of Computer and System Sciences, 42(3):267 – 287, 1991. doi:10.1016/0022-0000(91)90003-N.

Appendix A Properties of the measure λ(f)\operatorname{\lambda}(f)

We show that the measure λ(f)\operatorname{\lambda}(f) satisfies various elegant properties. First, it can be defined in multiple ways, one of which was introduced by Koutsoupias back in 1993 [Kou93]. It also has a formulation as a special case of the quantum adversary bound and hence can be expressed as as a semidefinite program closely related to that of the quantum adversary bound. Due to this characterization, λ(f)\operatorname{\lambda}(f) can be viewed as both a maximization problem and a minimization problem. These equivalent formulations are described in Section A.1.

Second, we show that λ(f)𝗌0(f)𝗌1(f)\operatorname{\lambda}(f)\leq\sqrt{\operatorname{\mathsf{s}}_{0}(f)\operatorname{\mathsf{s}}_{1}(f)}, which was already observed by Laplante, Lee, and Szegedy [LLS06] (though we give a slightly different proof). Finally, we show lower bounds on λ(f)\operatorname{\lambda}(f) and an optimal quadratic separation between λ(f)\operatorname{\lambda}(f) and 𝗌(f)\operatorname{\mathsf{s}}(f).

A.1 Equivalent formulations

Theorem 10.

For all Boolean functions f:{0,1}n{0,1}f\colon\{0,1\}^{n}\to\{0,1\}, we have

λ(f)=𝖪(f)=𝖠𝖽𝗏1(f)=𝖠𝖽𝗏1±(f),\operatorname{\lambda}(f)=\operatorname{\mathsf{K}}(f)=\operatorname{\mathsf{Adv}}_{1}(f)=\operatorname{\mathsf{Adv}}_{1}^{\pm}(f), (8)

where the measures 𝖪(f)\operatorname{\mathsf{K}}(f), 𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}(f), and 𝖠𝖽𝗏1±(f)\operatorname{\mathsf{Adv}}_{1}^{\pm}(f) are defined below. Furthermore, 𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}(f) itself has several equivalent formulations: 𝖠𝖽𝗏1(f)𝖲𝖠1(f)=𝖲𝖶𝖠1(f)=𝖬𝖬1(f)=𝖦𝖲𝖠1(f)\operatorname{\mathsf{Adv}}_{1}(f)\coloneqq\operatorname{\mathsf{SA}}_{1}(f)=\operatorname{\mathsf{SWA}}_{1}(f)=\operatorname{\mathsf{MM}}_{1}(f)=\operatorname{\mathsf{GSA}}_{1}(f).

We now define all these measures before proving this theorem.

Koutsoupias complexity 𝖪(f)\operatorname{\mathsf{K}}(f).

For a Boolean function ff, let Af1(0)A\subseteq f^{-1}(0), and let Bf1(1)B\subseteq f^{-1}(1). Let QQ be the matrix with rows and columns labeled by AA and BB respectively, with Q[x,y]=1Q[x,y]=1 if the Hamming distance of xx and yy is 11, and Q[x,y]=0Q[x,y]=0 otherwise. Koutsoupias [Kou93] observed that Q2\|Q\|^{2} is a lower bound on formula size, for every such choice of AA and BB. We define 𝖪(f)\operatorname{\mathsf{K}}(f) to be the maximum value of Q\|Q\| over choices of AA and BB. Thus 𝖪(f)2\operatorname{\mathsf{K}}(f)^{2} is a lower bound on the formula size of ff.

Single-bit positive adversary 𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}(f).

We define 𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}(f) as a version of the adversary bound where we are only allowed to put nonzero weight on input pairs (x,y)(x,y) where f(x)f(y)f(x)\neq f(y) and the Hamming distance between xx and yy is exactly 11. We will define 𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}(f) in terms of the spectral adversary version, which we also denote by 𝖲𝖠1(f)\operatorname{\mathsf{SA}}_{1}(f). 𝖠𝖽𝗏1(f)=𝖲𝖠1(f)\operatorname{\mathsf{Adv}}_{1}(f)=\operatorname{\mathsf{SA}}_{1}(f) is defined as the maximum of

Γmaxi[n]ΓDi\frac{\|\Gamma\|}{\max_{i\in[n]}\|\Gamma\circ D_{i}\|} (9)

over matrices Γ\Gamma of a special form. We require Γ\Gamma satisfy the following: (1) its entries are nonnegative reals; (2) its rows and columns are indexed by Dom(f)\operatorname{Dom}(f); (3) Γ[x,y]=0\Gamma[x,y]=0 whenever f(x)=f(y)f(x)=f(y); (4) Γ[x,y]=0\Gamma[x,y]=0 whenever the Hamming distance of xx and yy is not 11; and (5) Γ\Gamma is not all 0. In the above expression, \circ refers to the Hadamard (entrywise) product, Dom(f)\operatorname{Dom}(f) is the domain of ff, and DiD_{i} is the {0,1}\{0,1\}-valued matrix with Di[x,y]=1D_{i}[x,y]=1 if and only if xiyix_{i}\neq y_{i}.

Single-bit negative adversary 𝖠𝖽𝗏1±(f)\operatorname{\mathsf{Adv}}_{1}^{\pm}(f).

We define 𝖠𝖽𝗏1±(f)\operatorname{\mathsf{Adv}}_{1}^{\pm}(f) using the same definition as 𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}(f) above, except that the matrix Γ\Gamma is allowed to have negative entries. Note that since this is a relaxation of the conditions on Γ\Gamma, we clearly have 𝖠𝖽𝗏1±(f)𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}^{\pm}(f)\geq\operatorname{\mathsf{Adv}}_{1}(f).

Single-bit strong weighted adversary 𝖲𝖶𝖠1(f)\operatorname{\mathsf{SWA}}_{1}(f).

We define 𝖲𝖶𝖠1(f)\operatorname{\mathsf{SWA}}_{1}(f) as a single-bit version of the strong weighted adversary method 𝖲𝖶𝖠(f)\operatorname{\mathsf{SWA}}(f) from [SS06]. For this definition, we say a weight function w:Dom(f)×Dom(f)[0,)w\colon\operatorname{Dom}(f)\times\operatorname{Dom}(f)\to[0,\infty) is feasible if it is symmetric (i.e., w(x,y)=w(y,x)w(x,y)=w(y,x)) and if it satisfies the conditions on Γ\Gamma above (i.e., it places weight 0 on a pair (x,y)(x,y) unless both f(x)f(y)f(x)\neq f(y) and the Hamming distance between xx and yy is 11). We view such a feasible weight scheme ww as the weights on a weighted bipartite graph, where the left vertex set is f1(0)f^{-1}(0) and the right vertex set is f1(1)f^{-1}(1). We let wt(x)yw(x,y)wt(x)\coloneqq\sum_{y}w(x,y) denote the weighted degree of xx in this graph, i.e., the sum of the weights of its incident edges. Then 𝖲𝖶𝖠1(f)\operatorname{\mathsf{SWA}}_{1}(f) is defined as the maximum, over such feasible weight schemes ww, of

minx,i:w(x,xi)>0wt(x)wt(xi)w(x,xi).\min_{x,i:w(x,x^{i})>0}\frac{\sqrt{wt(x)wt(x^{i})}}{w(x,x^{i})}. (10)

Here xx ranges over Dom(f)\operatorname{Dom}(f), ii ranges over [n][n], and xix^{i} denotes the string xx with bit ii flipped.777Readers familiar with the adversary bound should note that this definition is analogous a weighted version of Ambainis’s original adversary method; in the original method, the denominator was the geometric mean of (a) the weight of the neighbors of xx with disagree with xx at ii, and (b) the weight of the neighbors of xix^{i} which disagree with xix^{i} at ii; but in our case, both (a) and (b) are simply w(x,xi)w(x,x^{i}), since xix^{i} is the only string that disagrees with xx on bit ii and is connected to xx in the bipartite graph.

Single-bit minimax adversary 𝖬𝖬1(f)\operatorname{\mathsf{MM}}_{1}(f).

Unlike the other forms, we define 𝖬𝖬1(f)\operatorname{\mathsf{MM}}_{1}(f) as a minimization problem rather than a maximization problem. We say a weight function w:Dom(f)×[n][0,)w\colon\operatorname{Dom}(f)\times[n]\to[0,\infty) is feasible if for all x,yDom(f)x,y\in\operatorname{Dom}(f) with f(x)f(y)f(x)\neq f(y) and Hamming distance 11, we have w(x,i)w(y,i)1w(x,i)w(y,i)\geq 1, where ii is the bit on which xx and yy disagree. 𝖬𝖬1(f)\operatorname{\mathsf{MM}}_{1}(f) is defined as the minimum, over such feasible weight schemes ww, of

maxxDom(f)i[n]w(x,i).\max_{x\in\operatorname{Dom}(f)}\sum_{i\in[n]}w(x,i). (11)

Semidefinite program version 𝖦𝖲𝖠1(f)\operatorname{\mathsf{GSA}}_{1}(f).

We define 𝖦𝖲𝖠1(f)\operatorname{\mathsf{GSA}}_{1}(f) to be the optimal value of the following semidefinite program.

maximizeZ,Afsubject toΔ is diagonaltrΔ=1ΔZDi0i[n]Z0\begin{array}[]{lll}\text{maximize}&\langle Z,A_{f}\rangle&\\ \text{subject to}&\Delta\mbox{ is diagonal}&\\ &\operatorname{tr}\Delta=1&\\ &\Delta-Z\circ D_{i}\succeq 0&\forall i\in[n]\\ &Z\geq 0&\end{array} (12)

Here ZZ and Δ\Delta are variable matrices with rows and columns indexed by Dom(f)\operatorname{Dom}(f), AfA_{f} is the {0,1}\{0,1\}-matrix with Af[x,y]=1A_{f}[x,y]=1 if and only if both f(x)f(y)f(x)\neq f(y) and (x,y)(x,y) have Hamming distance 11, and DiD_{i} is the {0,1}\{0,1\}-matrix with Di[x,i]=1D_{i}[x,i]=1 if and only if xiyix_{i}\neq y_{i}.

We now prove Theorem 10.

Proof.

Recall that in the definition of 𝖪(f)\operatorname{\mathsf{K}}(f), we picked Af1(0)A\subseteq f^{-1}(0) and Bf1(1)B\subseteq f^{-1}(1) and defined the resulting matrix QQ. Since the spectral norm of a submatrix is always smaller than or equal to the spectral norm of the original matrix, we can always assume without loss of generality that A=f1(0)A=f^{-1}(0) and B=f1(1)B=f^{-1}(1). Then 𝖪(f)=Q\operatorname{\mathsf{K}}(f)=\|Q\| for the resulting matrix QQ with rows and columns indexed by f1(1)f^{-1}(1) and f1(0)f^{-1}(0) respectively. Now, recall that AfA_{f} was the adjacency matrix of the graph GfG_{f}, which has an edge between xx and yy if f(x)f(y)f(x)\neq f(y) and the Hamming weight between xx and yy is 11. The rows and columns of AfA_{f} are each indexed by Dom(f)\operatorname{Dom}(f). By rearranging them, we can make AfA_{f} be block diagonal with blocks equal to QQ and QQ^{\dagger}. From there it follows that Af=Q\|A_{f}\|=\|Q\|, so λ(f)=𝖪(f)\operatorname{\lambda}(f)=\operatorname{\mathsf{K}}(f).

Next, recall that 𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}(f) is defined as the maximum ratio Γ/maxiΓDi\|\Gamma\|/\max_{i}\|\Gamma\circ D_{i}\| over valid choices of Γ\Gamma. Note that since Γ[x,y]\Gamma[x,y] can only be nonzero if xx and yy disagree on one bit, ΓDi\Gamma\circ D_{i} is nonzero only on pairs (x,y)(x,y) which disagree exactly on bit ii. In other words, if PiP_{i} denotes the {0,1}\{0,1\}-valued matrix with Pi[x,y]=1P_{i}[x,y]=1 if and only if xx and yy disagree on bit ii and only on ii, then ΓDi\Gamma\circ D_{i} is nonzero only in entries where PiP_{i} is 11. Now, note that PiP_{i} is a permutation matrix. Hence, by rearranging the rows and columns of ΓDi\Gamma\circ D_{i}, we can get it to be diagonal. This means ΓDi\|\Gamma\circ D_{i}\| is the maximum entry of ΓDi\Gamma\circ D_{i}, and hence maxiΓDi\max_{i}\|\Gamma\circ D_{i}\| is the maximum entry of Γ\Gamma. It follows that 𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}(f) is the maximum of Γ\|\Gamma\| over feasible matrices Γ\Gamma with max(Γ)1\max(\Gamma)\leq 1, where max(Γ)=maxij|Γij|\max(\Gamma)=\max_{ij}|\Gamma_{ij}|. This argument also holds for 𝖠𝖽𝗏1±(f)\operatorname{\mathsf{Adv}}_{1}^{\pm}(f), which is the maximum of Γ\|\Gamma\| over feasible (possibly negative) matrices Γ\Gamma with max(Γ)1\max(\Gamma)\leq 1.

Next, observe that negative weights never help for maximizing Γ\|\Gamma\|: indeed, if we had Γ\Gamma with negative entries maximizing Γ\|\Gamma\|, then we would have vectors uu and vv with u2=v2=1\|u\|_{2}=\|v\|_{2}=1 and uTΓv=Γu^{T}\Gamma v=\|\Gamma\|; but then replacing uu and vv with their entry-wise absolute values, and replacing Γ\Gamma with its entry-wise absolute value Γ\Gamma^{\prime}, we clearly get that ΓΓ\|\Gamma^{\prime}\|\geq\|\Gamma\|. However, max(Γ)=max(Γ)\max(\Gamma^{\prime})=\max(\Gamma), so Γ\Gamma^{\prime} remains feasible. This means we can always take the maximizing matrix Γ\Gamma to be nonnegative, so 𝖠𝖽𝗏1±(f)=𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}^{\pm}(f)=\operatorname{\mathsf{Adv}}_{1}(f). We can similarly assume that the unit vectors uu and vv maximizing uTΓvu^{T}\Gamma v are nonnegative.

Finally, consider the maximizing matrix Γ\Gamma and the maximizing unit vectors uu and vv, all nonnegative, and satisfying max(Γ)1\max(\Gamma)\leq 1. Note that the expression uTΓvu^{T}\Gamma v is nondecreasing in the entries of Γ\Gamma, since everything is nonnegative. Hence to maximize uTΓvu^{T}\Gamma v, we can always take every nonzero entry of Γ\Gamma to be 11, since this maintains max(Γ)1\max(\Gamma)\leq 1. In other words, the matrix maximizing Γ\|\Gamma\| will always simply be AfA_{f}, and hence 𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}(f) is always exactly equal to λ(f)\operatorname{\lambda}(f).

It remains to show that 𝖲𝖠1(f)=𝖲𝖶𝖠1(f)=𝖬𝖬1(f)=𝖦𝖲𝖠1(f)\operatorname{\mathsf{SA}}_{1}(f)=\operatorname{\mathsf{SWA}}_{1}(f)=\operatorname{\mathsf{MM}}_{1}(f)=\operatorname{\mathsf{GSA}}_{1}(f). The proof of this essentially follows the arguments in [SS06] for the regular positive adversary, though some steps are a little simpler. To start, we’ve seen that 𝖲𝖠1(f)=λ(f)\operatorname{\mathsf{SA}}_{1}(f)=\operatorname{\lambda}(f). Since AfA_{f} is symmetric, we have λ(f)=vTAfv\operatorname{\lambda}(f)=v^{T}A_{f}v for some unit vector vv, which we’ve established is nonnegative; this vector is also an eigenvector, so Afv=λ(f)vA_{f}v=\operatorname{\lambda}(f)v. Consider the weight scheme w(x,y)=v[x]v[y]Af[x,y]w(x,y)=v[x]v[y]A_{f}[x,y]. Then wt(x)=yv[x]v[y]Af[x,y]=v[x](Afv)[x]=λ(f)v[x]2wt(x)=\sum_{y}v[x]v[y]A_{f}[x,y]=v[x](A_{f}v)[x]=\operatorname{\lambda}(f)v[x]^{2}. Hence if w(x,xi)>0w(x,x^{i})>0, we have

wt(x)wt(xi)w(x,xi)=λ(f)v[x]v[xi]v[x]v[xi]Af[x,xi]=λ(f).\frac{\sqrt{wt(x)wt(x^{i})}}{w(x,x^{i})}=\frac{\operatorname{\lambda}(f)v[x]v[x^{i}]}{v[x]v[x^{i}]A_{f}[x,x^{i}]}=\operatorname{\lambda}(f). (13)

This means 𝖲𝖶𝖠1(f)𝖲𝖠1(f)\operatorname{\mathsf{SWA}}_{1}(f)\geq\operatorname{\mathsf{SA}}_{1}(f). In the other direction, let ww be a feasible weight scheme for 𝖲𝖶𝖠1(f)\operatorname{\mathsf{SWA}}_{1}(f), let Γ[x,y]=w(x,y)/wt(x)wt(y)\Gamma[x,y]=w(x,y)/\sqrt{wt(x)wt(y)}, and let v[x]=wt(x)/Wv[x]=\sqrt{wt(x)/W}, where W=xwt(x)W=\sum_{x}wt(x). Then v22=xwt(x)/W=1\|v\|_{2}^{2}=\sum_{x}wt(x)/W=1, and

vTΓv=x,ywt(x)wt(y)w(x,y)/Wwt(x)wt(y)=(1/W)x,yw(x,y)=1.v^{T}\Gamma v=\sum_{x,y}\sqrt{wt(x)wt(y)}w(x,y)/W\sqrt{wt(x)wt(y)}=(1/W)\sum_{x,y}w(x,y)=1. (14)

Hence Γ1\|\Gamma\|\geq 1. On the other hand, we have max(Γ)=maxx,yw(x,y)/wt(x)wt(y)\max(\Gamma)=\max_{x,y}w(x,y)/\sqrt{wt(x)wt(y)}. This means that the ratio Γ/max(Γ)\|\Gamma\|/\max(\Gamma) equals minx,y:w(x,y)>0wt(x)wt(y)/w(x,y)\min_{x,y:w(x,y)>0}\sqrt{wt(x)wt(y)}/w(x,y), which is 𝖲𝖶𝖠1(f)\operatorname{\mathsf{SWA}}_{1}(f); thus 𝖲𝖠1(f)𝖲𝖶𝖠1(f)\operatorname{\mathsf{SA}}_{1}(f)\geq\operatorname{\mathsf{SWA}}_{1}(f).

Next we examine 𝖦𝖲𝖠1(f)\operatorname{\mathsf{GSA}}_{1}(f). Consider a solution (Z,Δ)(Z,\Delta) to this semidefinite program and define Γ=ZMAf\Gamma=Z\circ M\circ A_{f}, where MM is defined as M=uuTM=uu^{T} and uu is defined by u[x]=1/Δ[x,x]u[x]=1/\sqrt{\Delta[x,x]} when Δ[x,x]>0\Delta[x,x]>0 and u[x]=0u[x]=0 otherwise. Recall that Δ\Delta is diagonal and that ΔZDi0\Delta-Z\circ D_{i}\succeq 0 for all ii. Since positive semidefinite matrices are symmetric, ZDiZ\circ D_{i} must be symmetric for all ii, so ZZ is symmetric. Moreover, the diagonal of ZDiZ\circ D_{i} is all zeros, so we must have Δ0\Delta\geq 0. Further, if Δ[x,x]=0\Delta[x,x]=0 for some xx, we must have the corresponding row and column of ZZ be all zeros. If we let Δ\Delta^{\prime} and ZZ^{\prime} be Δ\Delta and ZZ with the all-zero rows and columns deleted, then it is clear that ΔZDi0\Delta-Z\circ D_{i}\succeq 0 if and only if ΔZDi0\Delta^{\prime}-Z^{\prime}\circ D_{i}\succeq 0. Defining MM^{\prime} as MM with those rows and columns deleted and uu^{\prime} as uu with those entries deleted, we have M=u(u)T>0M^{\prime}=u^{\prime}(u^{\prime})^{T}>0. Observe that ΔZDi0\Delta^{\prime}-Z^{\prime}\circ D_{i}\succeq 0 if and only if vT(ΔZDi)v0v^{T}(\Delta^{\prime}-Z^{\prime}\circ D_{i})v\geq 0 for all vectors vv, which is if and only if (vu)T(ΔZDi)(vu)0(v\circ u^{\prime})^{T}(\Delta^{\prime}-Z^{\prime}\circ D_{i})(v\circ u^{\prime})\geq 0 for all vectors vv (since we have u>0u^{\prime}>0). This, in turn, is equivalent to M(ΔZDi)0M^{\prime}\circ(\Delta^{\prime}-Z^{\prime}\circ D_{i})\succeq 0. Since MΔ=IM^{\prime}\circ\Delta^{\prime}=I, this is equivalent to IMZDi0I-M^{\prime}\circ Z^{\prime}\circ D_{i}\succeq 0, which is in turn equivalent to IMZDi0I-M\circ Z\circ D_{i}\succeq 0. Since Z0Z\geq 0 and we are maximizing Z,Af\langle Z,A_{f}\rangle, it never helps for ZZ to have nonzero entries in places where AfA_{f} is 0. Hence we can assume without loss of generality that Z=ZAfZ=Z\circ A_{f}, which means the constraint becomes IΓDi0I-\Gamma\circ D_{i}\succeq 0, where we defined Γ=MZAf\Gamma=M\circ Z\circ A_{f}. We thus have ΓDi1\|\Gamma\circ D_{i}\|\leq 1. On the other hand, letting v[x]=Δ[x,x]v[x]=\sqrt{\Delta[x,x]}, we have

vTΓv=x,yv[x]v[y]M[x,y]Z[x,y]Af[x,y]=x,y:Δ[x,x],Δ[y,y]>0Z[x,y]Af[x,y]=Z,Af.v^{T}\Gamma v=\sum_{x,y}v[x]v[y]M[x,y]Z[x,y]A_{f}[x,y]=\sum_{x,y:\Delta[x,x],\Delta[y,y]>0}Z[x,y]A_{f}[x,y]=\langle Z,A_{f}\rangle. (15)

Hence 𝖲𝖠1(f)𝖦𝖲𝖠1(f)\operatorname{\mathsf{SA}}_{1}(f)\geq\operatorname{\mathsf{GSA}}_{1}(f). The reduction in the other direction works similarly: start with an adversary matrix Γ\Gamma with max(Γ)1\max(\Gamma)\leq 1, and let vv be its principle eigenvector. Then set Z=Γ(vvT)Z=\Gamma\circ(vv^{T}) and Δ=I(vvT)\Delta=I\circ(vv^{T}). Then IΓDi0I-\Gamma\circ D_{i}\succeq 0, which implies that ΔZDi0\Delta-Z\circ D_{i}\succeq 0. We also have trΔ=1\operatorname{tr}\Delta=1, Z0Z\geq 0, and Z,Af=Γ\langle Z,A_{f}\rangle=\|\Gamma\|.

Finally, we handle 𝖬𝖬1(f)\operatorname{\mathsf{MM}}_{1}(f). To do so, we first take the dual of the semidefinite program for 𝖦𝖲𝖠1(f)\operatorname{\mathsf{GSA}}_{1}(f). This dual has the form

minimizeαsubject toiRiIαIiRiDiAfRi0i[n]\begin{array}[]{lll}\text{minimize}&\alpha&\\ \text{subject to}&\sum_{i}R_{i}\circ I\leq\alpha I&\\ &\sum_{i}R_{i}\circ D_{i}\geq A_{f}&\\ &R_{i}\succeq 0&\forall i\in[n]\end{array} (16)

where the variables are α\alpha (a scalar) and matrices RiR_{i}, each with rows and columns indexed by Dom(f)\operatorname{Dom}(f). Strong duality follows since when AfA_{f} is not all zeros, and the semidefinite program in 𝖦𝖲𝖠1(f)\operatorname{\mathsf{GSA}}_{1}(f) has a strictly feasible solution (just take ZZ to equal ϵAf\epsilon A_{f} for a small enough positive constant ϵ\epsilon, and take Δ=I/|Dom(f)|\Delta=I/|\operatorname{Dom}(f)|). This means the optimal solution of the minimization problem above equals 𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}(f). It remains to show that this optimal solution TT also equals 𝖬𝖬1(f)\operatorname{\mathsf{MM}}_{1}(f).

Let α\alpha and {Ri}i\{R_{i}\}_{i} be a feasible solution to the semidefinite program above. Since Ri0R_{i}\succeq 0, we have Ri=XiXiTR_{i}=X_{i}X_{i}^{T} for some matrix XiX_{i}. Define w(x,i)=Ri[x,x]w(x,i)=R_{i}[x,x]. Note that we also have w(x,i)=aXi[x,a]2w(x,i)=\sum_{a}X_{i}[x,a]^{2}. Then by Cauchy–Schwarz, w(x,i)w(y,i)(aXi[x,a]Xi[y,a])2=(XiXiT)[x,y]2=Ri[x,y]2w(x,i)w(y,i)\geq\left(\sum_{a}X_{i}[x,a]X_{i}[y,a]\right)^{2}=(X_{i}X_{i}^{T})[x,y]^{2}=R_{i}[x,y]^{2}. If xx and yy are such that Af[x,y]=1A_{f}[x,y]=1, then they disagree in only one bit ii, and hence Di[x,y]=1D_{i}[x,y]=1 for that ii and Dj[x,y]=0D_{j}[x,y]=0 for all jij\neq i. Since we have iRiDiAf\sum_{i}R_{i}\circ D_{i}\geq A_{f}, we conclude that for all such pairs (x,y)(x,y), we have w(x,i)w(y,i)Ri[x,y]2Af[x,y]2=1w(x,i)w(y,i)\geq R_{i}[x,y]^{2}\geq A_{f}[x,y]^{2}=1 on the bit ii where xx and yy differ; hence the weight scheme ww is feasible. Furthermore, for any xx, iw(x,i)=iRi[x,x]αI[x,x]=α\sum_{i}w(x,i)=\sum_{i}R_{i}[x,x]\leq\alpha I[x,x]=\alpha. Hence 𝖬𝖬1(f)\operatorname{\mathsf{MM}}_{1}(f) is at most the optimal value of this semidefinite program.

In the other direction, consider a feasible weight scheme ww, and define Ri[x,y]=w(x,i)w(y,i)R_{i}[x,y]=\sqrt{w(x,i)w(y,i)}. Then Ri=w(,i)w(,i)TR_{i}=w(\cdot,i)w(\cdot,i)^{T}, where we treat w(,i)w(\cdot,i) as a vector; hence Ri0R_{i}\succeq 0. Moreover, Ri0R_{i}\geq 0, and for a pair (x,y)(x,y) with Af[x,y]=1A_{f}[x,y]=1, there is some ii which is the unique bit they disagree on, and hence w(x,i)w(y,i)1w(x,i)w(y,i)\geq 1; but this means that Ri[x,y]1R_{i}[x,y]\geq 1, and so (RiDi)[x,y]1=Af[x,y](R_{i}\cdot D_{i})[x,y]\geq 1=A_{f}[x,y]. Finally, iRi[x,x]=iw(x,i)\sum_{i}R_{i}[x,x]=\sum_{i}w(x,i), which means that iRiI𝖬𝖬1(f)I\sum_{i}R_{i}\circ I\leq\operatorname{\mathsf{MM}}_{1}(f)\cdot I, as desired. ∎

A.2 Upper bounds

We now show a slightly better upper bound on λ(f)\operatorname{\lambda}(f), that it is upper bounded by the geometric mean of the 0-sensitivity and 1-sensitivity, which can be a better upper bound than 𝗌(f)\operatorname{\mathsf{s}}(f).

We provide two proofs of this. The first uses the λ(f)\operatorname{\lambda}(f) formulation and uses a linear algebra argument about norms. This proof is due to Laplante, Lee, and Szegedy [LLS06], who observed this about the measure 𝖪(f)\operatorname{\mathsf{K}}(f).

To describe this proof, we briefly need to describe some matrix norms. For a vector vnv\in{\mathbb{R}}^{n}, the pp-norm for a positive integer pp is defined as vp=(i[n]|vi|p)1/p\left\lVert v\right\rVert_{p}=(\sum_{i\in[n]}|v_{i}|^{p})^{1/p}. We also define v=maxi[n]|vi|\left\lVert v\right\rVert_{\infty}=\max_{i\in[n]}|v_{i}|. Note that v1\left\lVert v\right\rVert_{1} is simply the sum of the absolute values of all the entries of the vector.

Similarly, for a matrix An×mA\in{\mathbb{R}}^{n\times m}, we define the induced pp-norm of AA to be

Ap=max{Axp:xp=1}.\left\lVert A\right\rVert_{p}=\max\{\left\lVert Ax\right\rVert_{p}:\left\lVert x\right\rVert_{p}=1\}. (17)

The spectral norm A\left\lVert A\right\rVert is the induced 22-norm A2\left\lVert A\right\rVert_{2}. The 1-norm A1\left\lVert A\right\rVert_{1} is simply the maximum sum of absolute values of entries in any column of the matrix. The \infty-norm A\left\lVert A\right\rVert_{\infty} is the maximum sum of absolute values of entries in any row of the matrix.

Lastly, we need a useful relationship between these norms sometimes called Hölder’s inequality for induced matrix norms (see [GL13, Corollary 2.3.2] for a proof):

Proposition 11.

For all matrices An×mA\in{\mathbb{R}}^{n\times m}, we have AA1A\left\lVert A\right\rVert\leq\sqrt{\left\lVert A\right\rVert_{1}\left\lVert A\right\rVert_{\infty}}.

We can now prove the upper bound:

Lemma 12.

For all (possibly partial) functions ff, we have λ(f)𝗌0(f)𝗌1(f)\operatorname{\lambda}(f)\leq\sqrt{\operatorname{\mathsf{s}}_{0}(f)\operatorname{\mathsf{s}}_{1}(f)}.

Proof.

We know that λ(f)=Af\operatorname{\lambda}(f)=\left\lVert A_{f}\right\rVert and AfA_{f} is a matrix of the form (0BBT0)\Bigl{(}\begin{smallmatrix}0&B\\ B^{T}&0\end{smallmatrix}\Bigr{)} if we rearrange the rows and columns so that all 0-inputs come first and are followed by 11-inputs, since AfA_{f} only connects inputs with different ff-values. Thus we have

λ(f)=Af=BB1B=𝗌0(f)𝗌1(f),\operatorname{\lambda}(f)=\left\lVert A_{f}\right\rVert=\left\lVert B\right\rVert\leq\sqrt{\left\lVert B\right\rVert_{1}\left\lVert B_{\infty}\right\rVert}=\sqrt{\operatorname{\mathsf{s}}_{0}(f)\operatorname{\mathsf{s}}_{1}(f)}, (18)

where we used Hölder’s inequality (Proposition 11) and the fact that the maximum row and column sum of BB are precisely 𝗌0(f)\operatorname{\mathsf{s}}_{0}(f) and 𝗌1(f)\operatorname{\mathsf{s}}_{1}(f), respectively. ∎

Our second proof of this claim uses the 𝖬𝖬1(f)\operatorname{\mathsf{MM}}_{1}(f) formulation which yields an arguably simpler proof.

Lemma 13.

For all (possibly partial) functions ff, we have 𝖠𝖽𝗏1(f)𝗌0(f)𝗌1(f)\operatorname{\mathsf{Adv}}_{1}(f)\leq\sqrt{\operatorname{\mathsf{s}}_{0}(f)\operatorname{\mathsf{s}}_{1}(f)}.

Proof.

Using the 𝖬𝖬1(f)\operatorname{\mathsf{MM}}_{1}(f) version of 𝖠𝖽𝗏1(f)\operatorname{\mathsf{Adv}}_{1}(f), set w(x,i)=𝗌0(f)/𝗌1(f)w(x,i)=\sqrt{\operatorname{\mathsf{s}}_{0}(f)}/\sqrt{\operatorname{\mathsf{s}}_{1}(f)} if f(x)=1f(x)=1, and set w(x,i)=𝗌1(f)/𝗌0(f)w(x,i)=\sqrt{\operatorname{\mathsf{s}}_{1}(f)}/\sqrt{\operatorname{\mathsf{s}}_{0}(f)} if f(x)=0f(x)=0. Then if xx and yy differ in a single bit ii, we clearly have w(x,i)w(y,i)=1w(x,i)w(y,i)=1. On the other hand, iw(x,i)𝗌1(f)𝗌0(f)/𝗌1(f)=𝗌0(f)𝗌1(f)\sum_{i}w(x,i)\leq\operatorname{\mathsf{s}}_{1}(f)\cdot\sqrt{\operatorname{\mathsf{s}}_{0}(f)}/\sqrt{\operatorname{\mathsf{s}}_{1}(f)}=\sqrt{\operatorname{\mathsf{s}}_{0}(f)\operatorname{\mathsf{s}}_{1}(f)} for 11-inputs xx, and analogously iw(y,i)𝗌0(f)𝗌1(f)\sum_{i}w(y,i)\leq\sqrt{\operatorname{\mathsf{s}}_{0}(f)\operatorname{\mathsf{s}}_{1}(f)} for 0-inputs yy. ∎

Using this better bound on λ(f)\operatorname{\lambda}(f) and Huang’s result, we also get that for all total Boolean functions ff,

𝖽𝖾𝗀(f)𝗌0(f)𝗌1(f).\operatorname{\mathsf{deg}}(f)\leq\operatorname{\mathsf{s}}_{0}(f)\operatorname{\mathsf{s}}_{1}(f). (19)

This result was also recently observed by Laplante, Naserasr, and Sunny [LNS20]. Unlike their proof, the following uses Huang’s theorem in a completely black-box way.

Proposition 14.

Assume that 𝖽𝖾𝗀(f)𝗌(f)2\operatorname{\mathsf{deg}}(f)\leq\operatorname{\mathsf{s}}(f)^{2} for all total Boolean functions ff. Then we also have 𝖽𝖾𝗀(f)𝗌0(f)𝗌1(f)\operatorname{\mathsf{deg}}(f)\leq\operatorname{\mathsf{s}}_{0}(f)\operatorname{\mathsf{s}}_{1}(f).

Proof.

Let 𝗌0(f)=k\operatorname{\mathsf{s}}_{0}(f)=k and 𝗌1(f)=\operatorname{\mathsf{s}}_{1}(f)=\ell. We know that 𝖽𝖾𝗀(f)max{k,}\operatorname{\mathsf{deg}}(f)\leq\max\{k,\ell\} by assumption. Let andkor\textsc{and}_{k}\circ\textsc{or}_{\ell} be the AND function on kk bits composed with the OR function on \ell bits. Clearly 𝗌0(andkor)=\operatorname{\mathsf{s}}_{0}(\textsc{and}_{k}\circ\textsc{or}_{\ell})=\ell and 𝗌1(andkor)=k\operatorname{\mathsf{s}}_{1}(\textsc{and}_{k}\circ\textsc{or}_{\ell})=k. Furthermore, because the function is monotone, the sensitive bits for a 0-input are bits set to 0, and the sensitive bits for a 11-input are bits set to 11. This means that composing this function with ff with yield a function where the one-sided sensitivity will be upper bounded by the product of one-sided sensitivity of the individual functions. Hence for all b{0,1}b\in\{0,1\}, we have

𝗌b(andkorf)𝗌b(andkor)𝗌b(f)k.\operatorname{\mathsf{s}}_{b}(\textsc{and}_{k}\circ\textsc{or}_{\ell}\circ f)\leq\operatorname{\mathsf{s}}_{b}(\textsc{and}_{k}\circ\textsc{or}_{\ell})\operatorname{\mathsf{s}}_{b}(f)\leq k\ell. (20)

Using the assumption on the function andkorf\textsc{and}_{k}\circ\textsc{or}_{\ell}\circ f, we get

𝖽𝖾𝗀(andkorf)(𝗌(andkorf))2(k)2.\operatorname{\mathsf{deg}}(\textsc{and}_{k}\circ\textsc{or}_{\ell}\circ f)\leq(\operatorname{\mathsf{s}}(\textsc{and}_{k}\circ\textsc{or}_{\ell}\circ f))^{2}\leq(k\ell)^{2}. (21)

Finally, it is well known that 𝖽𝖾𝗀(fg)=𝖽𝖾𝗀(f)𝖽𝖾𝗀(g)\operatorname{\mathsf{deg}}(f\circ g)=\operatorname{\mathsf{deg}}(f)\operatorname{\mathsf{deg}}(g) (see, e.g., [Tal13]), and hence 𝖽𝖾𝗀(andkorf)=k𝖽𝖾𝗀(f)\operatorname{\mathsf{deg}}(\textsc{and}_{k}\circ\textsc{or}_{\ell}\circ f)=k\ell\operatorname{\mathsf{deg}}(f), which implies 𝖽𝖾𝗀(f)k\operatorname{\mathsf{deg}}(f)\leq k\ell. ∎

A.3 Lower bounds

Finally, we prove some lower bounds on λ(f)\operatorname{\lambda}(f).

Lemma 15.

For all (possibly partial) functions ff, 𝗌(f)λ(f)2\operatorname{\mathsf{s}}(f)\leq\operatorname{\lambda}(f)^{2}.

Proof.

Consider any input xx with sensitivity 𝗌(f)\operatorname{\mathsf{s}}(f). This means xx has 𝗌(f)\operatorname{\mathsf{s}}(f) neighbors on the hypercube with different ff value. The sensitivity graph restricted to these 𝗌(f)+1\operatorname{\mathsf{s}}(f)+1 inputs is a star graph centered at xx. The spectral norm of the adjacency matrix of the star graph on k+1k+1 vertices is k\sqrt{k}. Since the spectral norm of AfA_{f} is lower bounded by that of a submatrix, we have λ(f)𝗌(f)\operatorname{\lambda}(f)\geq\sqrt{\operatorname{\mathsf{s}}(f)}. ∎

This relationship is tight for the orn\textsc{or}_{n} function which has 𝗌(orn)=n\operatorname{\mathsf{s}}(\textsc{or}_{n})=n and λ(orn)=n\operatorname{\lambda}(\textsc{or}_{n})=\sqrt{n}. Although orn\textsc{or}_{n} has unbalanced sensitivities, with 𝗌0(orn)=n\operatorname{\mathsf{s}}_{0}(\textsc{or}_{n})=n and 𝗌1(orn)=1\operatorname{\mathsf{s}}_{1}(\textsc{or}_{n})=1, there are functions ff with 𝗌(f)=𝗌0(f)=𝗌1(f)=n\operatorname{\mathsf{s}}(f)=\operatorname{\mathsf{s}}_{0}(f)=\operatorname{\mathsf{s}}_{1}(f)=n and λ(f)=n\operatorname{\lambda}(f)=\sqrt{n}. One example of such a function is x1or(x2,,xn)x_{1}\oplus\textsc{or}(x_{2},\ldots,x_{n}). Another example of such a function with a quadratic gap between 𝗌(f)\operatorname{\mathsf{s}}(f) and λ(f)\operatorname{\lambda}(f) is the function that is 11 if and only if the input string has Hamming weight 11. This function has 𝗌0(f)=n\operatorname{\mathsf{s}}_{0}(f)=n since the all zeros string is fully sensitive and 𝗌1(f)=n\operatorname{\mathsf{s}}_{1}(f)=n since every Hamming weight 11 string is also fully sensitive. But we know that this problem can be solved by Grover’s algorithm with O(n)O(\sqrt{n}) queries, and hence λ(f)=O(𝖰(f))=O(n)\operatorname{\lambda}(f)=O(\operatorname{\mathsf{Q}}(f))=O(\sqrt{n}).

We can also lower bound Af\left\lVert A_{f}\right\rVert using the relationship between spectral norm and Frobenius norm. We have for all N×NN\times N matrices AA that A1NAF\left\lVert A\right\rVert\geq\frac{1}{\sqrt{N}}\left\lVert A\right\rVert_{F} [GL13, Eq. (2.3.7)], where AF2=i,j|Aij|2\left\lVert A\right\rVert^{2}_{F}=\sum_{i,j}|A_{ij}|^{2}. For the sensitivity graph of ff, 1NAfF\frac{1}{\sqrt{N}}\left\lVert A_{f}\right\rVert_{F} is just the average sensitivity.

Lemma 16.

For all (possibly partial) functions ff, λ(f)𝐄x[𝗌x(f)]\operatorname{\lambda}(f)\geq\mathop{\bf E\/}_{x}[\operatorname{\mathsf{s}}_{x}(f)].

This can be improved by only taking the expectation on the right over a subset of the inputs of ff, which then equals another complexity measure originally defined by Khrapchenko [Khr71]. See [Kou93] for more details.