This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Paper The following article is Open access

Optimal bounds for parity-oblivious random access codes

, , and

Published 7 April 2016 © 2016 IOP Publishing Ltd and Deutsche Physikalische Gesellschaft
, , Focus on Device Independent Quantum Information Citation André Chailloux et al 2016 New J. Phys. 18 045003 DOI 10.1088/1367-2630/18/4/045003

1367-2630/18/4/045003

Abstract

Random access coding is an information task that has been extensively studied and found many applications in quantum information. In this scenario, Alice receives an n-bit string x, and wishes to encode x into a quantum state ${\rho }_{x}$, such that Bob, when receiving the state ${\rho }_{x}$, can choose any bit $i\in [n]$ and recover the input bit xi with high probability. Here we study two variants: parity-oblivious random access codes (RACs), where we impose the cryptographic property that Bob cannot infer any information about the parity of any subset of bits of the input apart from the single bits xi; and even-parity-oblivious RACs, where Bob cannot infer any information about the parity of any even-size subset of bits of the input. In this paper, we provide the optimal bounds for parity-oblivious quantum RACs and show that they are asymptotically better than the optimal classical ones. Our results provide a large non-contextuality inequality violation and resolve the main open problem in a work of Spekkens et al (2009 Phys. Rev. Lett.102 010401). Second, we provide the optimal bounds for even-parity-oblivious RACs by proving their equivalence to a non-local game and by providing tight bounds for the success probability of the non-local game via semidefinite programming. In the case of even-parity-oblivious RACs, the cryptographic property holds also in the device independent model.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Quantum information theory studies how information is encoded in quantum mechanical systems and how it can be transmitted through quantum channels. A main question is whether quantum information is more powerful than classical information. A celebrated result by Holevo [Hol73] shows that quantum information cannot be used to compress classical information. In high level, in order to transmit n uniformly random classical bits, one needs to transmit no less than n quantum bits. This might imply that quantum information is no more powerful than classical information. This however is wrong in many situations. In the model of communication complexity, one can show that transmitting quantum information may result in exponential savings on the communication needed to solve specific problems [BCWdW01, BJK04, GKK+08, Raz99, RK11].

One specific information task that has been extensively studied in quantum information is the notion of random access codes (RACs) [ANTV02, ANTV99, Nay99]. In this scenario, Alice receives an n-bit string x, drawn from the uniform distribution, and wishes to encode x into a quantum state ${\rho }_{x}$, such that Bob, when receiving the state ${\rho }_{x}$, can choose any bit $i\in [n]$ and recover the input bit xi with high probability by performing some general quantum operation on ${\rho }_{x}$.

${\rm{RAC}}{\rm{s}}$ have been used in various situations in quantum information and computation, including in communication complexity, non-locality, extractors and device-independent cryptography [BARdW08, DV10, INRY07, LPY+, PZ10]. Even though this task seems easier than transmitting the entire input string x, it is known that the length of quantum ${\rm{RAC}}{\rm{s}}$ must be at least ${\rm{\Omega }}(n)$ [Nay99]. In fact, the length of a classical ${\rm{RAC}}$ can be within a logarithmic additive factor of a quantum ${\rm{RAC}}$ [ANTV99].

On the other hand, a well-known example shows the advantages of quantum ${\rm{RAC}}{\rm{s}}$ by using a single qubit to encode two uniformly random classical bits. In this case, the success of correctly decoding either bit is ${\mathrm{cos}}^{2}(\pi /8)$ [ANTV99, BBBW83] while the optimal classical encoding can achieve an average success probability of 3/4. An advantage can also be proven for the case of encoding three classical bits into one qubit as shown by Chuang (see [ANTV02] for details), but not for $n\geqslant 4$ [HIN+06].

Nevertheless, a question remained of whether there are variants of RACs, for which we can have an asymptotically significant advantage in the quantum case. We show that this is indeed the case for the so-called parity-oblivious${\rm{RAC}}{\rm{s}}$ (denoted here as PO-RACs). These are the usual ${\rm{RAC}}{\rm{s}}$ with the extra cryptographic property that the receiver cannot infer any information about the parity of any subset of bits of the input, apart from the single bits.

This cryptographic property means, in particular, that once some information about a bit is learned, then no other information can be extracted about any of the other bits. Such a notion has applications in various areas of cryptography. For example, this is a requirement for a class of classical or quantum protocols known as symmetric-private information retrieval schemes (PIR) [GIKM98, KdW04], where one or more servers have a database x, a user chooses an index i and at the end, the user learns xi but no other bit of x, and i remains hidden. A parity-oblivious ${\rm{RAC}}$ satisfies the security conditions of a PIR scheme since the index i remains hidden (the ${\rm{RAC}}$ is non-interactive) and the user cannot learn more than one bit of the database.

RACs that are parity-oblivious have been considered before. For example, the previously mentioned ${\rm{RAC}}{\rm{s}}$ for encoding two or three classical bits in one qubit have this property. It is not hard to check that for any subset of the inputs of size 2 or greater, Bob's reduced density matrix is exactly the same for the cases, where the parity is 0 or 1. In other words, Bob has no information about the parity. These ${\rm{RAC}}{\rm{s}}$ violate a non-contextuality inequality developed by Spekkens et al [SBK+09]. This inequality is discussed further in section 1.1.3.

We will also define a weaker variant called even-parity-oblivious${\rm{RAC}}{\rm{s}}$ (denoted as EPO-RACs), where the receiver can infer no information about the parity of any even-size subset of the input. These codes are interesting for two reasons: first, they will let us prove tight upper bounds for PO-RACs and second, due to their equivalence with a non-local game, their cryptographic property holds in the device-independent setting.

1.1. Our results

We split our results into four sections. We first present the optimal bounds for parity-oblivious quantum ${\rm{RAC}}{\rm{s}}$ and even-parity-oblivious RACs. We then contrast this to the classical case (section 1.1.2), discuss a violation of a non-contextuality inequality (section 1.1.3), then discuss the security of our optimal quantum ${\rm{RAC}}{\rm{s}}$ in the device-independent model (section 1.1.4).

1.1.1. Quantum RACs and cryptographic security definitions

Formally, a quantum ${\rm{RAC}}$ of n classical bits is simply a set of quantum states $\{{\rho }_{x}\;:x\in \{0,1\}{}^{n}\}$. We suppose Alice chooses $x\in \{0,1\}{}^{n}$ uniformly at random, prepares the state ${\rho }_{x}$, and sends ${\rho }_{x}$ to Bob who has a POVM $\{{M}_{0}^{t},{M}_{1}^{t}\}$, where the subscript b of ${M}_{b}^{t}$ serves as his guess for the tth bit of x, denoted xt, for each index $t\in [n]:= \{1,\ldots ,n\}$. Suppose that xt can be decoded with success probability $\tfrac{1}{2}(1+{\alpha }_{t})$. Then we say that the bias, or worst-case bias, of the ${\rm{RAC}}$ is

and the average-case bias is

where μ is the uniform probability distribution.

We consider two cryptographic variants of quantum ${\rm{RAC}}{\rm{s}}$ in this paper. We are concerned with designing quantum ${\rm{RAC}}{\rm{s}}$ which hide some information about the encoded string x from a potentially cheating Bob. By information being hidden, we mean that there exists no measurement which yields a correct guess with probability greater than that of randomly guessing. In particular, we consider the case where Bob cannot learn the value of

for certain choices of subset $S\subseteq [n]$, of Alice's encoded string x. We call the value xS the S-parity of the string x.

(Parity-oblivious and even-parity-oblivious quantum RACs).

Definition 1 We say that a quantum ${\rm{RAC}}$ $\{{\rho }_{x}\;:x\in \{0,1\}{}^{n}\}$ is parity-oblivious, denoted PO-RACn, if the receiver can infer no information about xS for any subset $S\subseteq [n]$ of size 2 or greater, when x is chosen uniformly at random. In other words, for all $S\subseteq [n]$ of size 2 or greater, we have

We say that a quantum ${\rm{RAC}}$ is even-parity-oblivious, denoted EPO-RACn, if the receiver can infer no information about xS for any subset $S\subseteq [n]$ of even size, 2 or greater.

Note that the usual treatment of ${\rm{RAC}}{\rm{s}}$ is to analyze the relationships between the number of encoded bits n, the bias α, and the encoding dimension of ${\rho }_{x}$. Here, we are not concerned with the encoding dimension, but rather the ability to achieve cryptographic security in terms of parity-obliviousness.

In this paper, we present the optimal bias for a quantum PO-RACn and show that they perform asymptotically better than the optimal classical version.

(Optimal quantum parity-oblivious random access codes).

Theorem 1 For any integer $n\geqslant 2$, a quantum parity-oblivious random access code of n bits has worst-case bias at most $1/\sqrt{n}$. Moreover, this bound can be achieved using $\lfloor n/2\rfloor $ qubits.

This is in contrast to the classical setting where the optimal average-case bias is provably $1/n$ [SBK+09] (discussed further in section 1.1.2).

The main idea of the proof of the upper bound is that quantum encodings can be studied through their close relationship to non-local games. Such connections were noted in [OW10] and in [CKS14] it was shown that certain non-local games are equivalent to quantum encodings in the sense that the optimal average decoding probability is equal to the success probability of the non-local game.

In a non-local game, two non-communicating parties, Alice and Bob, receive some inputs s and t, respectively, according to some probability distribution known to Alice and Bob, and must output a and b, respectively, such that $(s,t,a,b)$ satisfy some specific condition. For example, in the CHSH game [CHSH69], the condition is $a\oplus b=s\cdot t$. The goal is to find the optimal quantum (resp. classical) success probability of satisfying the condition when Alice and Bob are allowed to share some initial quantum state (resp. shared randomness).

We now define a very natural non-local game called the ${\rm{INDEX}}$ game which we use in the analysis in this paper.

Definition 2. [(INDEX game)] The INDEXn game, parameterized by n, is the following non-local game:

  • Alice's input: Alice receives a random s from the set $S:= \{0,1\}{}^{n}$.
  • Bob's input: Bob receives a random index t from the set $T:= [n]$.
  • Winning condition: they win if Alice's output bit a and Bob's output bit b satisfy $a\oplus b={s}_{t}$.

The choice of initial resource state and local measurement operators (that depend on the respective inputs) comprise a strategy. We say that a strategy for the ${{\rm{INDEX}}}^{n}$ game has bias α if

We show that even-parity-oblivious ${\rm{RAC}}{\rm{s}}$ with average-case bias are equivalent to the ${\rm{INDEX}}$ game. In other words, any ${\rm{INDEX}}$ game strategy with bias α yields an even-parity-oblivious ${\rm{RAC}}$ with average-case bias α and vice versa.

(Equivalence).

Theorem 2 For any $n\in {\mathbb{N}}$, there exists a quantum even-parity-oblivious RAC of n bits with average-case bias $\alpha $ if and only if there exists a quantum INDEXn strategy with bias $\alpha $.

Noting that the ${\rm{INDEX}}$ game is an XOR game, i.e., the winning condition depends only on the XOR of Alice and Bob's one-bit answers, we use a tight semidefinite programming characterization [CSUU08] to provide the exact optimal quantum bias.

(Optimal quantum INDEX game bias).

Theorem 3 For any $n\in {\mathbb{N}}$, the optimal quantum bias of an INDEXn strategy is $1/\sqrt{n}$.

The above two theorems imply the optimal bounds for even-parity-oblivious RAC.

(Optimal quantum even-parity-oblivious random access codes).

Corollary 1 For any integer $n\geqslant 2$, a quantum even-parity-oblivious random access code of n bits has average-case bias at most $1/\sqrt{n}$. Moreover, this bound can be achieved using $\lfloor n/2\rfloor $ qubits.

Since the worst-case bias of a quantum PO-RAC is obviously upper bounded by the optimal average-case bias of a quantum ${\rm{RAC}}$ hiding only the even parities, theorems 1 and 2 show that every PO-RACn has bias at most $1/\sqrt{n}$.

To prove this upper bound is tight, we give an explicit construction of a quantum PO-RAC of n bits with bias $1/\sqrt{n}$ that uses $\lfloor n/2\rfloor $ qubits and 1 classical bit. This ${\rm{RAC}}$ is based on the notion of hyperbits[PW12] and a proof of Tsirelson's theorem [Tsi87]. We then discuss how to remove the classical bit and make it device-independent (see section 1.1.4 for more details about the device-independent model).

We remark that parity-oblivious and even-parity-oblivious quantum RACs both share the same worst-case and average-case bias of $1/\sqrt{n}$. However, the same is not true if we consider odd-parity-oblivious${\rm{RAC}}{\rm{s}}$, where the parities are hidden for only odd-size subsets (greater or equal to 3). Consider encoding a six-bit string $({x}_{1},\ldots ,{x}_{6})$, where the first three bits are encoded using Chuang's PO-RAC and similarly for the last three bits. It is a straightforward exercise to verify that this is odd-parity-oblivious and that any bit can be decoded with bias $1/\sqrt{3}\gt 1/\sqrt{6}$. We leave finding the optimal bounds for odd-parity-oblivious ${\rm{RAC}}{\rm{s}}$ an open problem.

1.1.2. Parity-oblivious classical RACs

We also study classical RACs, defined below, for which both variants of bias and both variants of parity-obliviousness are defined analogously.

(Classical RAC s with worst-case and average-case biases).

Definition 3 A classical${\rm{RAC}}$ is a set of strings $\{e(x,r)\;:x\in \{0,1\}{}^{n},r\in \{0,1\}{}^{m}\}$, where r corresponds to private randomness. After choosing $x\in \{0,1\}{}^{n}$ uniformly at random, Alice samples r from the private randomness, sends to Bob the string $e(x,r)$, and Bob has a decoding procedure given as function ft, for each $t\in [n]$, for learning the t'th bit of x.

We note that the equivalence stated in theorem 2 holds in the classical case as well (remarked in section 2). To find the optimal average-case bias of even-parity-oblivious classical RACs, we provide the following theorem.

(Optimal classical INDEX game bias).

Theorem 4 For any $n\in {\mathbb{N}}$, the optimal classical bias of an INDEXn strategy is $\sqrt{\tfrac{2}{\pi n}}(1+O(1/n))$.

This theorem, together with the classical version of the equivalence shows that classical ${\rm{RAC}}{\rm{s}}$ that are even-parity-oblivious have an optimal average-case bias of $\sqrt{\tfrac{2}{\pi n}}(1+O(1/n))$. Note that, asymptotically, this value is the same as the quantum value, that is, having a bias of $O(1/\sqrt{n})$. However, differences arise when one considers ${\rm{RAC}}{\rm{s}}$ that also hide the odd parities. Consider the following proposition of Spekkens, Buzacott, Keehn, Toner, and Pryde.

(Optimal parity-oblivious classical RACs [SBK+09]).

Proposition 1 For any $n\in {\mathbb{N}}$, a parity-oblivious classical RAC of n bits has average-case bias at most $1/n$. Moreover, this bound can be achieved using 1 classical bit.

Thus, there is a difference between the optimal average-case biases of parity-oblivious and even-parity-oblivious ${\rm{RAC}}{\rm{s}}$ in the classical setting, in contrast to the quantum setting.

1.1.3. Large non-contextuality inequality violations

The basic primitives in an operational theory are preparations and measurements which can be thought of as instructions for the laboratory apparatus. For example, the operational theory can be given in terms of hidden variables which are probability distributions characterizing the outcomes of the preparations and measurements. That is, a preparation creates a physical state (each occurring with some probability) and a measurement acts upon a physical state and outputs a prediction or simply an outcome (each occurring with some probability). Thus, the probability distributions characterizing these actions are how they are represented in this model.

A hidden variable model is preparation non-contextual if whenever two preparations yield the same statistics for all possible measurements then they are represented equivalently in the model and a hidden variable model is measurement non-contextual if whenever two measurements have the same statistics for all preparations then they are represented equivalently in the model (see [SBK+09] and references therein for a more thorough discussion). Similar to non-locality, a non-contextuality inequality is any inequality on probability distributions that follows from the assumption that there exists a hidden variable model that is preparation or measurement non-contextual.

Spekkens et al [SBK+09] proved the following non-contextuality inequality (or NC inequality, for short).

(Non-contextuality inequality [SBK+09]).

Proposition 2 In any operational theory that admits a preparation non-contextual hidden variable model, the average-case bias for any parity-oblivious RAC is at most $1/n$.

Then, they discussed that quantum mechanics violates this NC inequality for $n\in \{2,3\}$, by noting the previously mentioned parity-oblivious quantum ${\rm{RAC}}{\rm{s}}$ of two and three classical bits into one qubit with respective average-case biases of $\tfrac{1}{\sqrt{2}}$ and $\tfrac{1}{\sqrt{3}}$. It was left as an open question whether quantum mechanics violates this NC inequality for $n\geqslant 4$.

Through our analysis, we have shown that the optimal average-case bias for quantum parity-oblivious ${\rm{RAC}}{\rm{s}}$ is $1/\sqrt{n}$, thus resolving their main open question. This provides a family of NC inequality violations that grow with the input size n.

Note, that if there exists a game for which the winning probability of any classical strategy cannot deviate from 1/2 by more than ${\delta }_{1}$ and, moreover, there is a quantum strategy with winning probability at least $1/2+{\delta }_{2}$, then we can obtain a violation of order ${\delta }_{2}/{\delta }_{1}$ (see [BRSdW12] for details). Hence, to quantify the violation of this NC inequality, we consider the ratio of the optimal average-case bias of quantum parity-oblivious ${\rm{RAC}}{\rm{s}}$ and that of any operational theory admitting a preparation non-contextual hidden variable model. More precisely, we show an explicit non-contextuality inequality violation of order $\sqrt{n}$.

Theorem 5. For any integer $n\geqslant 2$, there exists an explicit non-contextuality inequality that provides a violation of order $\sqrt{n}$.

Note that other large non-contextuality inequality violations have been found, see for example the work of Vidick and Wehner [VW11].

1.1.4. Device-independent quantum RACs

Until this point, we have discussed the bias and the parity-obliviousness of a quantum ${\rm{RAC}}$ which are functions of the encoding states $\{{\rho }_{x}\;:x\in \{0,1\}{}^{n}\}$ only. However, much of the cryptographic analysis in this paper is concerned with how the states ${\rho }_{x}$ are prepared. In this section, we discuss how the security of the ${\rm{RAC}}$ is affected if one cannot trust the quantum apparatus used in the preparation of ${\rho }_{x}$.

The device-independent model of cryptographic security deals with the setting when the devices used in the protocol are not trusted, or are even malicious, being created by the cheaters/eavesdroppers themselves. Many security proofs in this setting are based on quantum non-locality or the no-signalling principle, each having their own limitations which ultimately limits the cheating capabilities for anyone controlling the preparation and/or execution of the quantum devices in the protocol. Recall that the no-signalling principle, which is satisfied by the laws of quantum mechanics, roughly states that it is impossible to send information arbitrarily fast, in particular faster than the speed of light. For example, in a quantum setting, Alice cannot convey information to a distant Bob by simply measuring her half of a shared quantum state.

Obviously, if the preparation of the encoding ${\rho }_{x}$ is as simple as Alice having a quantum device which outputs ${\rho }_{x}$ on input x, then certainly device-independence is not feasible since Bob may control the quantum device and just have it prepare ${\rho }_{x}=| x\rangle \langle x| $ (or some other function of x according to what he wishes to learn). However, the preparation need not be so simple. We now sketch the preparation of the quantum ${\rm{RAC}}{\rm{s}}$ presented in this work to give an idea of how they can be device-independent.

First, Alice creates a bipartite quantum state $| \psi \rangle $ and sends a subsystem to Bob. Afterwards she chooses a string $s\in \{0,1\}{}^{n}$ uniformly at random and measures her half of the state to get an outcome $a\in \{0,1\}$. She then defines ${x}_{t}:= {s}_{t}\oplus a$, for all $t\in [n]$, and Bob's post-measured state is now his encoding of the string x. Since there is no communication from Alice to Bob after Alice chooses s, he must not be able to infer any information about s from his encoding of x. Thus, Bob has limited information of any function of x which contains information about s. For example, Bob cannot learn ${x}_{1}\oplus {x}_{2}$ since

which is hidden by the no-signalling principle. Therefore, even if Bob created the entire state which Alice shares at the beginning, and Alice's measurement, he cannot infer any information about ${x}_{1}\oplus {x}_{2}$, promised only by the no-signalling principle.

Theorem 6. There exists a preparation of an optimal PO-RACn with bias $1/\sqrt{n}$ which is even-parity-oblivious in the device-independent model against a no-signalling Bob.

We prove the above theorem using a small modification of our optimal PO-RACn in section 4. See section 4.3 for more details. Note also that the above theorem implies that there exist optimal even-parity-oblivious RAC which retain their cryptographic property in the device independent model.

1.2. Organization of the paper

In section 2, we prove the equivalence of even-parity-oblivious ${\rm{RAC}}{\rm{s}}$ and ${{\rm{INDEX}}}^{n}$ strategies. In section 3 we discuss the optimal quantum and classical bias of the ${{\rm{INDEX}}}^{n}$ game for any n. We conclude in section 4 by presenting an optimal parity-oblivious quantum ${\rm{RAC}}$ and prove the security for the even-parity-obliviousness in the device-independent model.

2. Equivalence of EPO-RACs and ${{\rm{INDEX}}}^{n}$ strategies

In this section we prove the equivalence in theorem 2, reproduced below.

(Equivalence).

Theorem 2 For any $n\in {\mathbb{N}}$, there exists a quantum even-parity-oblivious RAC of n uniformly random classical bits with average-case bias $\alpha $ if and only if there exists a quantum INDEX strategy with bias $\alpha $.

For this reason, even-parity-obliviousness of a ${\rm{RAC}}$ is a very natural property. In particular, in the simple reduction from ${\rm{INDEX}}$ strategies to ${\rm{RAC}}{\rm{s}}$ (section 2.2), we see how even-parity-obliviousness appears and how the ${\rm{RAC}}$ may not hide the odd parities.

2.1. From ${\rm{RAC}}{\rm{s}}$ to ${\rm{INDEX}}$ strategies

Let us fix an even-parity-oblivious ${\rm{RAC}}$ $\{{\rho }_{x}\;:x\in \{0,1\}{}^{n}\}$ with average-case bias α. Let ${ \mathcal B }$ be the Hilbert space used for the encoding. Our goal is to construct a strategy for ${{\rm{INDEX}}}^{n}$ with bias α. For each ${\rho }_{x}$, we fix a purification $| {\psi }_{x}\rangle $ of ${\rho }_{x}$ in the space ${ \mathcal A }\otimes { \mathcal B }$. For $a\in \{0,1\}$, let ${\boldsymbol{a}}$ be the n-bit string $(a,...,a)$ and $\bar{s}$ be the bit-wise complement of a string s. For $s\in {\{\mathrm{0,1}\}}^{n}$, define the following state

where ${ \mathcal O }$ is a qubit register containing the value of a. We would like to show that if Bob has the register ${ \mathcal B }$ of the above state, then he has no information about s. Note that his reduced state is ${\sigma }_{s}:= \tfrac{1}{2}{\rho }_{s}+\tfrac{1}{2}{\rho }_{\bar{s}}$.

The first step is to see that Bob has no information about any parity of s (not even of the values of the singleton bits). Fix an arbitrary, non-empty subset S. For fixed $b\in \{0,1\}$, Bob's reduced state, averaged over all $s\in \{0,1\}{}^{n}$ such that ${s}_{S}=b$, is given by

Note that ${s}_{S}={\bar{s}}_{S}$ when $| S| $ is even and ${s}_{S}={\bar{s}}_{S}\oplus 1$ when $| S| $ is odd. Thus, by defining ${\rho }_{S}^{b}$ in the similar way

we can easily verify that ${\sigma }_{S}^{b}={\rho }_{S}^{b}$ when $| S| $ is even and ${\sigma }_{S}^{b}=\displaystyle \frac{1}{2}{\rho }_{S}^{0}+\displaystyle \frac{1}{2}{\rho }_{S}^{1}$ when $| S| $ is odd. Note that since $\{{\rho }_{x}\;:x\in \{0,1\}{}^{n}\}$ is an even-parity-oblivious ${\rm{RAC}}$, we have by definition that ${\rho }_{S}^{0}={\rho }_{S}^{1}$ for $| S| $ even (otherwise, Bob could measure to learn some information about the even parity). Thus, we have that ${\sigma }_{S}^{0}={\sigma }_{S}^{1}$ for all non-empty subsets S and therefore all the parities are hidden from Bob when given ${\sigma }_{s}$ (when s is chosen uniformly at random). This means that for any non-empty subset S and measurement M, Bob has a maximum probability of 1/2 of successfully guessing sS from the ${\rm{RAC}}$ $\{{\sigma }_{s}\;:s\in \{0,1\}{}^{n}\}$.

In the following lemma, we prove that if an encoding reveals no information about the parity of any subset, then the encoding reveals no information about the string. This is intuitively an obvious statement that we rigorously prove below.

Lemma 1. If an encoding $\{{\sigma }_{s}\;:s\in \{0,1\}{}^{n}\}$ satisfies ${{\mathbb{E}}}_{s\sim \mu (\{\mathrm{0,1}\}{}^{n})}\mathrm{Pr}[{\rm{learn}}\;{\rm{}}{s}_{S}]=\displaystyle \frac{1}{2}$, for every subset $S\subseteq [n]\setminus \varnothing $, then ${\sigma }_{s}={\sigma }_{{s}^{\prime }}$ for all $s,{s}^{\prime }\in \{0,1\}{}^{n}$.

Proof. Suppose for a contradiction that there exists $s,{s}^{\prime }\in \{0,1\}{}^{n}$ such that ${\sigma }_{s}\ne {\sigma }_{s^{\prime} }$. Then there exists a subset $T\subseteq \{0,1\}{}^{n}$ of size ${2}^{n-1}$ such that ${\sigma }_{T}=\tfrac{1}{{2}^{n-1}}{\sum }_{s\in T}{\sigma }_{s}$ is not equal to ${\sigma }_{\bar{T}}=\tfrac{1}{{2}^{n-1}}{\sum }_{s\in \bar{T}}{\sigma }_{s}$, where $\bar{T}$ denotes the complement of the set T. To see this, take any subset $T\subseteq \{0,1\}{}^{n}$ of size ${2}^{n-1};$ if ${\sigma }_{T}={\sigma }_{\bar{T}}$, then we can find $s\in T$ and ${s}^{\prime }\in \bar{T}$ such that ${\sigma }_{s}\ne {\sigma }_{s^{\prime} }$, since all the ${\sigma }_{i}$ are not equal. We consider the subset ${T}^{\prime }$, where we add $\{{s}^{\prime }\}$ and remove $\{s\}$ from T to obtain ${\sigma }_{{T}^{\prime }}\ne {\sigma }_{\bar{{T}^{\prime }}}$.

This means that there exists a two-outcome measurement $\{{M}_{T},{M}_{\bar{T}}\}$ that outputs 1 if $s\in T$ and −1 otherwise, with positive bias. We now show for a contradiction that this measurement must also output a parity of some non-empty subset with positive bias. Define the function $f\;:\{0,1\}{}^{n}\to \{-1,+1\}$ as the indicator function of T and let b be the expectation over the measurement outcomes when measuring ${\sigma }_{s}$ with $\{{M}_{T},{M}_{\bar{T}}\}$, so $b(s):= \mathrm{Tr}({\sigma }_{s}{M}_{T})-\mathrm{Tr}({\sigma }_{s}{M}_{\bar{T}})$. Then

By taking the Fourier representation of the function, we have

Note that $\hat{f}(\varnothing )={\mathbb{E}}[f(s)]=0$, because $| T| =| \bar{T}| $, implying that there exists a non-empty subset S for which

which is a contradiction.□

The above statement means that for each s, we have ${\mathrm{Tr}}_{{ \mathcal O }{ \mathcal A }}| {{\rm{\Omega }}}_{s}\rangle \langle {{\rm{\Omega }}}_{s}| ={\mathrm{Tr}}_{{ \mathcal O }{ \mathcal A }}| {{\rm{\Omega }}}_{0}\rangle \langle {{\rm{\Omega }}}_{0}| $. In particular, for any $s\in \{0,1\}{}^{n}$ there exists a unitary Us acting on ${ \mathcal O }{ \mathcal A }$ such that $({U}_{s}\otimes I)| {{\rm{\Omega }}}_{0}\rangle =| {{\rm{\Omega }}}_{s}\rangle $. We use the state $| {{\rm{\Omega }}}_{0}\rangle $ to define the ${{\rm{INDEX}}}^{n}$ strategy:

  • Alice and Bob share the state $| {{\rm{\Omega }}}_{0}\rangle \in { \mathcal A }\otimes { \mathcal B }$.
  • Upon receiving $s\in \{0,1\}{}^{n}$, Alice applies Us on ${ \mathcal O }{ \mathcal A }$ such that Alice and Bob share $| {{\rm{\Omega }}}_{s}\rangle $. Alice measures register ${ \mathcal O }$ in the computational basis and outputs the measurement outcome a.
  • For Alice's input s and output a, Bob has an encoding ${\rho }_{x}$ where $x:= s\oplus {\boldsymbol{a}}$ occurs uniformly at random. Upon receiving $t\in [n]$, Bob measures ${ \mathcal B }$ just as in the ${\rm{RAC}}$ to learn xt. He outputs b equal to his guess.
  • Alice and Bob win the game if $b={s}_{t}\oplus a={x}_{t}$ meaning that they win the game if and only if Bob correctly guesses xt.

Since the ${\rm{RAC}}$ has average-case bias α, we see that with this ${{\rm{INDEX}}}^{n}$ strategy, they succeed with probability

as desired.

2.2. From ${\rm{INDEX}}$ strategies to ${\rm{RAC}}$

Suppose Alice and Bob have a strategy to win the ${{\rm{INDEX}}}^{n}$ game with bias α with starting state $| \psi \rangle \in { \mathcal A }\otimes { \mathcal B }$. On input $s\in \{0,1\}{}^{n}$, Alice performs on her side the corresponding measurement which generates her outcome a. We assume that a is uniformly random and independent of s (which can be guaranteed by taking the XOR with an independently and uniformly random bit that is shared with Bob). Let ${\rho }_{s,a}$ be the state that Bob has when Alice has input s and outputs a and define the ${\rm{RAC}}$ $\{{\sigma }_{x}\;:x\in \{0,1\}{}^{n}\}$, where ${\sigma }_{x}:= \tfrac{1}{2}{\rho }_{x,0}+\tfrac{1}{2}{\rho }_{\bar{x},1}$ for each $x\in \{0,1\}{}^{n}$. We now show that $\{{\sigma }_{x}\;:x\in \{0,1\}{}^{n}\}$ is an even-parity-oblivious ${\rm{RAC}}$ with average-case bias α. Note that $\tfrac{1}{2}{\rho }_{s,0}+\tfrac{1}{2}{\rho }_{s,1}$ is independent of s by the no-signalling principle. For convenience, define $\rho := \tfrac{1}{2}{\rho }_{s,0}+\tfrac{1}{2}{\rho }_{s,1}$ for any $s\in \{0,1\}{}^{n}$.

  • (1)  
    It hides the even parities: let $S\subseteq \{0,1\}{}^{n}$ be a subset of even size and $b\in \{0,1\}$ be an arbitrary bit. Then we have ${x}_{S}={\bar{x}}_{S}$ for any $x\in \{0,1\}{}^{n}$, since $| S| $ is even. Bob's reduced state, averaged over all $x\in \{0,1\}{}^{n}$ such that ${x}_{S}=b$, is given by
    which is independent of b, thus proving the ${\rm{RAC}}$ is even-parity-oblivious.
  • (2)  
    Since Alice and Bob win the ${{\rm{INDEX}}}^{n}$ game with average-case bias α, we know that
    By defining $x:= s\oplus {\boldsymbol{a}}$, we can write the above as
    as desired.

Note that in the proof above, we are treating x, the string Alice wishes to encode, as $s\oplus {\boldsymbol{a}}$. We now remark that some of the odd parities of x may not be hidden from Bob. For example, if in the ${\rm{INDEX}}$ game Alice simply outputs $a={s}_{1}\oplus {s}_{2}\oplus {s}_{3}\oplus d$, where d is the uniformly random bit Alice and Bob share to make a independent of s, then we have ${x}_{1}\oplus {x}_{2}\oplus {x}_{3}=d$ and Bob would know this odd parity exactly. However, the even parities of x are equal to those of s which are hidden by the no-signalling principle.

Remark 1. The above equivalence also holds in the classical setting.

3. On the structure of optimal ${\rm{INDEX}}$ game strategies

In this section, we prove theorems 3 and 4, that the optimal quantum bias of an ${{\rm{INDEX}}}^{n}$ strategy is $1/\sqrt{n}$ and the optimal classical bias of an ${{\rm{INDEX}}}^{n}$ strategy is $\sqrt{\tfrac{2}{\pi n}}(1+O(1/n))$.

3.1. The quantum bias

The quantum bias of any XOR game can be found efficiently by solving a semidefinite program (SDP) [CSUU08]. The optimization takes place over a matrix indexed by $s\in S$ and $t\in T$ with each entry corresponding to the expectation of the measurement outcome of a fixed game strategy. Such a matrix of inner products can be written as a positive semidefinite matrix and the expectation (or bias) of the game strategy is then an inner product of this matrix and one containing the information of the XOR game.

Specifically, the quantum bias of the ${{\rm{INDEX}}}^{n}$ game can be calculated as the optimal value of either SDP below

where

  • $\mathrm{diag}(X)$ is the vector on the diagonal of the square matrix X,
  • e is the vector of all ones,
  • $\mathrm{Diag}(y)$ is the diagonal matrix with the vector y on the diagonal,
  • $B:= \displaystyle \frac{1}{2}\left[\begin{array}{cc}0 & A\\ {A}^{\top } & 0\end{array}\right]$, where ${A}_{s,t}:= \displaystyle \frac{{(-1)}^{{s}_{t}}}{n{2}^{n}}$.

For (P), consider the positive semidefinite matrix $X:= {{YY}}^{\top }$, where

To show X is feasible in (P), one can check that each diagonal entry of X is equal to 1 from the definition of A above. Note that $\langle B,X\rangle := \sqrt{n}\;{2}^{n}\langle A,A\rangle =1/\sqrt{n}$ proving that the quantum bias is at least $1/\sqrt{n}$ (since the quantum bias is the maximum of $\langle B,X\rangle $ over all feasible X).

For (D), let $y:= \left[\begin{array}{c}u\;{e}_{S}\\ v\;{e}_{T}\end{array}\right]$, where $u,v\gt 0$ (determined later) and eS and eT are the vectors of all ones indexed by entries in S and T, respectively. Then

From above, if we set $v:= \displaystyle \frac{1}{2n\sqrt{n}}$ and $u:= \displaystyle \frac{1}{2\sqrt{n}\;{2}^{n}}$, then $y$ is feasible in (D). Since

we know the quantum bias is at most $1/\sqrt{n}$ (since the quantum bias is equal to the minimum of $\langle e,y\rangle $ over all feasible y). Therefore, the quantum bias is exactly $1/\sqrt{n}$, as required.

The ${\rm{INDEX}}$ game turns out to be equivalent to the retrieval game studied in [OW10] which is defined similarly except the first bit of Alice's input is always 0 and the other $n-1$ bits are chosen independently and uniformly at random. To see the equivalence, notice that in the ${\rm{INDEX}}$ game Alice can take her input $s\in \{0,1\}{}^{n}$, define ${s}^{\prime }={\boldsymbol{m}}\oplus s$, where m fixes the specific bit to a specific value, play the retrieval game strategy with input ${s}^{\prime }$ to generate ${a}^{\prime }$, and then output $a:= {a}^{\prime }\oplus m$ (Bob plays the same strategy). Thus, any strategy for the retrieval game with bias α yields a strategy for the ${\rm{INDEX}}$ game with bias α as well. We further remark that the quantum bias of the retrieval game is shown to be $1/\sqrt{n}$ in [OW10] through the use of uncertainty relations. Using this result, and the equivalence to the ${\rm{INDEX}}$ game, we have another proof that the quantum bias of the INDEX game is $1/\sqrt{n}$.

3.2. The classical bias

We can assume without loss of generality that Alice and Bob's strategies are deterministic. Define $b\in \{0,1\}{}^{n}$ as the string of potential answers Bob gives, where bt is the bit that Bob outputs on input $t\in [n]$. Now let us examine Alice's strategy. For a fixed input s, if she outputs 1, they win the game with probability

where $| x{| }_{{\rm{H}}}$ denotes the Hamming weight of a string $x\in \{0,1\}{}^{n}$. If she outputs 0, they win the game with probability

Since their strategies are deterministic, Alice should output the maximum of these two, so

Therefore, the classical bias is precisely$\tfrac{2}{n}{{\mathbb{E}}}_{s\sim \mu (\{\mathrm{0,1}\}{}^{n})}\left|\tfrac{n}{2}-| b\oplus s{| }_{{\rm{H}}}\right|$. Note that this quantity is independent of b, thus we could assume Bob always outputs 0 for every input. The quantity

corresponds to the mean deviation of the uniform binomial distribution. This is a well studied quantity [Fra45] and we know that

Therefore, the classical bias is $\tfrac{2}{n}\sqrt{\tfrac{n}{2\pi }}\left(1+O\left(\tfrac{1}{n}\right)\right)=\sqrt{\tfrac{2}{\pi n}}\left(1+O\left(\tfrac{1}{n}\right)\right)$, as desired.

4. A construction of a quantum PO-RACn with optimal bias

In this section, we give an explicit construction of a quantum PO-RACn with optimal bias.

(Optimal PO-RACn).

Lemma 2 For any integer $n\geqslant 2$, there exists a PO-RACn with bias $1/\sqrt{n}$ that uses $\lfloor n/2\rfloor $ qubits.

Our construction builds upon the previously mentioned ${\rm{RAC}}{\rm{s}}$ for sending 2 (resp. 3) classical bits with bias $1/\sqrt{2}$ (resp. $1/\sqrt{3}$). These are the vertices from the corners of a square inscribed in an equatorial plane in the Bloch sphere, and the corners of a cube inscribed in the Bloch sphere, respectively. To generalize this idea to an n-cube inscribed in an n-dimensional sphere, we use the intuition of hyperbits, which are a way to visualize such unit vectors in a quantum mechanical setting. A full discussion of hyperbits and their equivalence to certain quantum protocols is beyond the scope of this paper, but we refer the interested reader to the work of Pawlowski and Winter [PW12].

We note that, after the publication of this paper, we became aware that a similar ${\rm{RAC}}$ had been previously discovered by Wehner [Weh08], but remained unpublished.

4.1. The construction

Our construction is very similar to a proof of Tsirelson's theorem [Tsi87]. We start by recursively defining the observables ${G}_{n,1},\ldots ,{G}_{n,n}$, for $n\geqslant 2$, which are used to define the actions of Alice and Bob in the PO-RACn. For n = 2 and n = 3, we define

We use the n = 3 observables as a base case for a recursive formula:

Note that these act on $\lfloor n/2\rfloor $ qubits7 , have eigenvalues ±1, and satisfy the anti-commutation relation

Define the following operators for $x\in \{0,1\}{}^{n}$ and $t\in [n]$:

Note that ${A}_{x}^{2}=I$, for all $x\in \{0,1\}{}^{n}$, and ${B}_{t}^{2}=I$, for all $t\in [n]$, so each have ±1 eigenvalues.

The PO-RACn protocol is defined below.

  • Encoding states: Alice chooses a uniformly random $x\in \{0,1\}{}^{n}$, creates $\lfloor n/2\rfloor $ EPR pairs, and measures the first 'halves' with the observable Ax to get an outcome $a\in \{-1,+1\}$. The second 'halves' now contain the post-measurement state ${\tau }_{x,a}$ and since $\mathrm{Tr}({A}_{x})=0$, each a occurs with 1/2 probability. She sends ${\tau }_{x,a}$ and a to Bob who now has the mixed state
    encoding the string x.
  • Decoding procedure: if Bob wishes to learn xt, he measures his EPR halves with the observable Bt to get an outcome $b\in \{-1,+1\}$ and also measures to learn a. He computes $c={ab}$ and outputs 0 if $c=+1$, and 1 otherwise.

In the next two lemmas, we show that the worst-case bias of this ${\rm{RAC}}$ is $\tfrac{1}{\sqrt{n}}$ and that it is parity-oblivious, thereby proving lemma 2.

Lemma 3. The quantum RAC $\{{\rho }_{x}\;:x\in \{0,1\}{}^{n}\}$ has worst-case bias at least $1/\sqrt{n}$.

Proof. We can assume at the beginning of the protocol, Alice and Bob share the maximally entangled state

The expectation value of the observable $C={A}_{x}\otimes {B}_{t}$ in this state is given by:

where the third equality is derived from the anti-commutation relation. We can write

implying

This proves that

as desired. □

Lemma 4. The quantum RAC $\{{\rho }_{x}\;:x\in \{0,1\}{}^{n}\}$ is parity-oblivious.

Proof. Protocols involving shared entanglement and sending one classical bit have limited guessing probabilities for functions such as parity [PW12]. In particular, it can be shown that the biases ${\alpha }_{S}$ of learning xS satisfy

In the ${\rm{RAC}}$ above, we have

implying ${\alpha }_{S}=0$ for all S of size 2 or greater, implying it is parity-oblivious. □

This concludes a construction of a quantum PO-RACn with optimal bias. However, we have not yet proved theorem 1 since the encoding dimension is too high. We now discuss a small modification to simultaneously reduce the dimension of the ${\rm{RAC}}$ and to increase its device-independence.

4.2. Removing the classical message

Reducing the dimension of the PO-RACn$\{{\rho }_{x}\;:x\in \{0,1\}{}^{n}\}$ is straightforward. First notice that Bob simply takes his measurement outcome and changes it if $a=-1$ to obtain his guess for xt. We can remove the need for this message if Alice simply changes the value of x for which Bob has the encoding. In other words, instead of sending a to Bob, she just switches every bit of x if $a=-1$. Then Bob's guess b is just as accurate in guessing ${x}_{t}^{\prime }$ where ${x}_{t}^{\prime }={x}_{t}$, if $a=+1$, and $x{^{\prime} }_{t}=\bar{{x}_{t}}$, if $a=-1$. (Note that this is similar to what was done in section 2.2). Therefore, Bob now has an encoding of $x^{\prime} $, which we denote by $\{{\sigma }_{x^{\prime} }\;:x^{\prime} \in \{0,1\}{}^{n}\}$. Notice that ${\sigma }_{x^{\prime} }$ is a state on only $\lfloor n/2\rfloor $ qubits which coincides with the optimal PO-RACn for n = 2 and n = 3 previously discussed.

It is easy to see that this new ${\rm{RAC}}$ has bias at least $1/\sqrt{n}$ by construction. We now argue that this quantum ${\rm{RAC}}$ is still parity-oblivious.

Lemma 5. The quantum RAC $\{{\sigma }_{{x}^{\prime }}\;:x\in \{0,1\}{}^{n}\}$ is parity-oblivious.

Proof. We prove that any S-parity hidden from Bob in the quantum ${\rm{RAC}}$ $\{{\rho }_{x}\;:x\in \{0,1\}{}^{n}\}$ is still hidden from Bob in the quantum ${\rm{RAC}}$ $\{{\sigma }_{{x}^{\prime }}:{x}^{\prime }\in \{0,1\}\}$. Let ${\tau }_{x,a}$ be Bob's post-measured state immediately after Alice used measurement Ax and received output a. We can write

Fix a subset $S\subseteq [n]$ of size at least 2. Since xS is hidden from Bob in the ${\rm{RAC}}$ $\{{\rho }_{x}\;:x\in \{0,1\}{}^{n}\}$, we have ${\sum }_{x:{x}_{S}=0}{\rho }_{x}={\sum }_{x:{x}_{S}=1}{\rho }_{x}$. Thus,

implying that

Equation (1)

Now we can write

using equation (1) implying that the quantum ${\rm{RAC}}$ $\{{\sigma }_{x}\;:x\in \{0,1\}{}^{n}\}$ also hides xS from Bob. □

Since the quantum PO-RACn$\{{\sigma }_{x}\;:x\in \{0,1\}{}^{n}\}$ has optimal bias and uses only $\lfloor n/2\rfloor $ qubits, this concludes the proof of our main result, theorem 1.

4.3. Making our optimal PO-RACn device-independent

We first point out that our preparation of the PO-RACn$\{{\rho }_{x}\;:x\in \{0,1\}{}^{n}\}$ is not secure in the device-independent model for the following reason. Suppose the measurements are not trusted, in the sense that Bob controls them. Consider the case when Alice's measurement, which depends on x, simply outputs $a:= {x}_{1}\oplus {x}_{2}$. Then, when a is sent to Bob, he now knows some information about the parities of x.

Note that in the preparation of the ${\rm{RAC}}$ $\{{\sigma }_{x^{\prime} }\;:{x}^{\prime }\in \{0,1\}{}^{n}\}$ the even parities of $x^{\prime} $ are hidden from Bob by the no-signalling principle since they are equal to those for x and there is no communication from Alice to Bob in the preparation of ${\sigma }_{x^{\prime} }$ after Alice chooses x. This is the basis for our preparation of a quantum ${\rm{RAC}}$ secure in the device-independent security model.

There is one small caveat however that does not affect the security, but may change how $x^{\prime} $ is generated. That is, it may not be generated uniformly now. Consider the case when Bob controls the measurement and decides to set $a:= {x}_{1}$. Then, ${\sigma }_{x^{\prime} }$ will never be prepared if $x{^{\prime} }_{1}=1$. However, there is an easy fix at the price of adding one classical bit of communication.

Before sending Bob's part of the (supposed) maximally entangled state, Alice can choose a random bit d and send that to Bob as well. (We assume Alice can flip a random coin without being effected from a malicious Bob.) Then Alice takes the XOR of d with her measurement outcome, and proceeds as usual. Bob can easily adjust his guess for ${x}_{t}^{\prime }$ using the value of d.

It is easy to see that this preparation of the PO-RACn$\{{\sigma }_{x^{\prime} }\;:x^{\prime} \in \{0,1\}{}^{n}\}$ hides the even parities, in the device-independent model, against any malicious Bob respecting the no-signalling principle, thus proving theorem 6.

It is not the case that the preparation of the PO-RACn$\{{\sigma }_{x^{\prime} }:x^{\prime} \in \{0,1\}{}^{n}\}$ hides the odd parities as well in the device-independent model. For example, suppose Bob controls the measurement such that $a:= {x}_{1}\oplus {x}_{2}\oplus {x}_{3}$. Then

for any choice of x. Thus, if Bob controls the measurements, he can make it so that Alice's first three bits always have parity d, of which he knows the value!

We leave it as an open problem to see if there exists a preparation of an optimal PO-RACn that is still parity-oblivious in the device-independent model against either no-signalling or quantum adversaries.

Conclusion

We have provided the optimal bounds for parity-oblivious quantum ${\rm{RAC}}{\rm{s}}$ and showed that they are asymptotically better than the optimal classical ones. We discussed how these optimal ${\rm{RAC}}{\rm{s}}$ provide a large non-contextuality inequality violation and resolve the main open problem in a work of Spekkens et al [SBK+09]. We also studied optimal bounds for a related version of these ${\rm{RAC}}{\rm{s}}$ which only hide the even parities. We showed their equivalence to a non-local game and explained why even-parity-obliviousness was the correct notion of ${\rm{RAC}}$ in this setting. After constructing a family of optimal parity-oblivious RACs, we discussed how to make the even parities secure in the device-independent model.

We end with an open question. We have seen how even-parity-obliviousness plays a key role in our analysis, especially when making ${\rm{RAC}}{\rm{s}}$ secure in the device-independent setting. This raises the question: what can be said about the optimal bias for odd-parity-oblivious RACs? We have discussed how they can have greater bias than the two variants studied in this work, but perhaps the optimal bias can still be expressed as some nice function of the number of encoded bits. Also, it would be interesting to see if they can be made secure in the device-independent model as well.

Acknowledgments

This research was supported by the French National Research Agency, through CRYQ (ANR-09-JCJC-0067) and by the European Union through the ERC project QCC.

Footnotes

  • We note here that the choice of these observables is not unique and there are applications in the literature that use slightly different observables. However, this particular choice reduces the ${\rm{RAC}}$ dimension by one qubit when n is odd. For example, for n = 3 our ${\rm{RAC}}$ uses $\lfloor n/2\rfloor =1$ qubit (as opposed to 2) just as in the well-known quantum ${\rm{RAC}}$ of three classical bits into one qubit.

Please wait… references are loading.
10.1088/1367-2630/18/4/045003