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

Privacy Against Hypothesis-Testing Adversaries for Quantum Computing

Farhad Farokhi
Abstract

A novel definition for data privacy in quantum computing based on quantum hypothesis testing is presented in this paper. The parameters in this privacy notion possess an operational interpretation based on the success/failure of an omnipotent adversary being able to distinguish the private categories to which the data belongs using arbitrary measurements on quantum states. Important properties of post processing and composition are then proved for the new notion of privacy. The relationship between privacy against hypothesis-testing adversaries, defined in this paper, and quantum differential privacy are then examined. It is shown that these definitions are intertwined in some parameter regimes. This enables us to provide an interpretation for the privacy budget in quantum differential privacy based on its relationship with privacy against hypothesis testing adversaries.

I Introduction

Quantum computing algorithms have garnered huge attention due to their considerable speedups in several classically difficult problems, such as factorising [1]. These breakthroughs and the added attention has paved the way for development of new algorithms for big data processing, such as quantum machine learning [2, 3, 4]. However, data processing can result in unintended information leakage [5]. This is an important issue because, as quantum hardware becomes more commercially available, these algorithms can be implemented on real-world sensitive, private, or proprietary datasets. Therefore, there is a need for development of frameworks to better understand private information leakage in quantum computing algorithms and to construct privacy-preserving algorithms.

In the classical computing literature, differential privacy has become the gold standard of privacy analysis and private algorithm design [6, 7, 8]. This is often attributed to the fact that differential privacy makes minimal assumptions about the data (e.g., range rather than distribution) and meets important properties of post processing and compositions [9]. Although possessing powerful guarantees, differential privacy has been polarizing [10, 11, 12]. Criticisms surroundings conservativeness of differential privacy have motivated a host of studies on privacy in information theory that can handle privacy-utility trade-off better in certain situations [13, 14, 15, 16, 17]. In fact, adoption of hypothesis-testing and estimation-based adversaries have been proposed as less conservative alternatives to differential privacy by social scientists following implementation of differential privacy in the 2020 United States Decennial Census of Population and Housing [10]. Nonetheless, differential privacy has been recently extended to quantum computing [18, 19, 20]. However, very little attention has been paid other forms of privacy in quantum systems. In this paper, we investigate privacy against hypothesis-testing adversaries. This is of particular interest to us due to the need for providing an operational, intuitive notion of privacy with real-world interpretations for privacy analysis and guarantees, which is somewhat absent in the differential privacy literature.

In this paper, we particularly propose a novel definition for data privacy for quantum computing based on quantum hypothesis testing. The design parameters in this notion of privacy possess an operational interpretation (specifically for general lay-users) based on the success/failure of an omnipotent adversary being able to distinguish the private class to which the data belongs (e.g., suffering from a certain disease in health datasets or belonging to the training dataset in membership inference attacks) based on the arbitrary measurement operators. We prove two very important properties for the new notion of privacy: post processing and composition. These properties are highly sought-after in privacy definitions [18] and information leakage metrics [13]. Subsequently, we investigate the relationship between privacy against hypothesis-testing adversaries and quantum differential privacy. This enables us to provide an interpretation for parameters of differential privacy based on its relationship with privacy against hypothesis-testing adversaries in certain parameter regimes. We will finally investigate the effectiveness of differential privacy against hypothesis-testing adversaries.

The remainder of this paper is organized as follows. We provide a review of basic concepts in quantum computing and information in Section II. The definition and results on privacy against hypothesis-testing adversaries is presented in Section III. Section IV presents quantum differential privacy and its relationship with privacy against hypothesis-testing adversaries. Finally, we present some concluding remarks and future directions for research in Section V.

II Quantum states and channels

The definitions and preliminary results in this review section are mostly borrowed from [21]. When the results or definitions are from outside this source, appropriate citations are presented.

A quantum system is modelled by a Hilbert space \mathcal{H}, i.e., a complex vector space, equipped with an inner product, that is complete with respect to the norm defined by the inner product. Throughout this paper, Dirac’s notation is used to denote quantum states. That is, a pure quantum state, which is an element (i.e., vector) of Hilbert space \mathcal{H} with unit norm, is denoted by ‘ket’ |\ket{\cdot}, e.g., |ψ\ket{\psi}\in\mathcal{H}. The inner product of two states |ϕ\ket{\phi} and |ψ\ket{\psi} is denoted by ϕ|ψ\innerproduct{\phi}{\psi}. Here, ‘bra’ ψ|\bra{\psi} is used to refer to conjugate transpose of |ψ\ket{\psi} and ϕ|ψ:=ϕ||ψ\innerproduct{\phi}{\psi}:=\bra{\phi}\,\ket{\psi}\in\mathbb{C}.

The basic element of interest in quantum information theory is a quantum bit, which is often referred to as the qubit. A qubit is a 2-dimensional quantum state. Any qubit can be written in terms of the so-called computational basis |0\ket{0} and |1\ket{1} that form an orthonormal basis for the two dimensional Hilbert space modelling the qubit, that is, any qubit can be written as |ψ=α|0+β|1\ket{\psi}=\alpha\ket{0}+\beta\ket{1} with α,β\alpha,\beta\in\mathbb{C} such that |α|2+|β|2=1|\alpha|^{2}+|\beta|^{2}=1. Combination of two qubits |ϕ\ket{\phi} and |ψ\ket{\psi} is denoted by their tensor product |ϕ|ψ\ket{\phi}\otimes\ket{\psi}, where \otimes is the Kronecker or tensor product. For the sake of brevity, we sometimes refer to |ϕ|ψ\ket{\phi}\otimes\ket{\psi} as |ϕ|ψ\ket{\phi}\ket{\psi} or |ϕ,ψ\ket{\phi,\psi}. When two qubits |ϕ\ket{\phi} and |ψ\ket{\psi} belong to or assigned to two distinct registers AA and BB (e.g., qubits used by two separate parties), and this information is either unclear from the context or must be emphasized, we write |ϕA|ψB\ket{\phi}_{A}\otimes\ket{\psi}_{B} or |ϕA|ψB\ket{\phi}_{A}\ket{\psi}_{B}. A quantum (logic) gate is any unitary operator, e.g., UU such that UU=IU^{\dagger}U=I, that acts on a quantum state. Note that, here, UU^{\dagger} denotes the conjugate transpose of UU.

A mixed quantum state is represented by an ensemble {(p1,|ψ1),,(pk,|ψk)}\{(p_{1},\ket{\psi_{1}}),\dots,(p_{k},\ket{\psi_{k}})\} such that pi0p_{i}\geq 0 for all i[k]:={1,,k}i\in[k]:=\{1,\dots,k\} and i[k]pi=1\sum_{i\in[k]}p_{i}=1. A mixed quantum state implies that the quantum system is in pure state |ψi\ket{\psi_{i}} with probability pip_{i} for all i[k]i\in[k]. A convenient way to model and analyse mixed quantum state is to use density operators. The density operator corresponding to ensemble {(p1,|ψ1),,(pk,|ψk)}\{(p_{1},\ket{\psi_{1}}),\dots,(p_{k},\ket{\psi_{k}})\} is given by ρ:=i[k]pi|ψiψk|\rho:=\sum_{i\in[k]}p_{i}\ket{\psi_{i}}\bra{\psi_{k}}. Evidently, by construction, tr(ρ)=1\tr(\rho)=1. Note that pure quantum states |ϕ\ket{\phi} can also be modelled using rank-one density operator ρ=|ϕϕ|\rho=\ket{\phi}\bra{\phi}. Therefore, there is no loss of generality to work with density operators even when dealing with pure quantum states. Combination of two density operators ρ\rho and σ\sigma is denoted by their tensor product ρσ\rho\otimes\sigma.

A basic operation in quantum systems is measurement, which enables extraction of information about the quantum states of the systems. A measurement is modelled by a set of operators M={Ki}i[m]M=\{K_{i}\}_{i\in[m]} with normalization constraint that i[m]KiKi=I\sum_{i\in[m]}K_{i}^{\dagger}K_{i}=I. By performing measurement MM on a quantum system with state ρ\rho, we observe output i[m]i\in[m] with probability tr(KiρKi)\tr(K_{i}\rho K_{i}^{\dagger}) in which case, after the measurement, the state of the quantum system is KiρKi/tr(KiρKi)K_{i}\rho K_{i}^{\dagger}/\tr(K_{i}\rho K_{i}^{\dagger}). When the post-measurement state of the quantum system is of no interest, we can use the positive operator-valued measure (POVM) framework, which is a set of positive semi-definite Hermitian matrices F={Fi}i[m]F=\{F_{i}\}_{i\in[m]} such that i[m]Fi=I\sum_{i\in[m]}F_{i}=I. In this case, the probability of obtaining output i[m]i\in[m] when taking a measurement on a system with quantum state ρ\rho is given by tr(ρFi)=tr(Fiρ)\tr(\rho F_{i})=\tr(F_{i}\rho).

A quantum channel is the most general quantum operation. A quantum channel is a mapping from the space of density operators to potentially another space of density operators that is both completely positive and trace preserving. Quantum channels model open quantum systems, i.e., quantum systems that interact with environment, and thus can model noisy quantum behaviours. According to Choi-Kraus theorem [21, Theorem 4.4.1], for each quantum channel \mathcal{E}, there exists a family of linear operators {Ej}j[n]\{E_{j}\}_{j\in[n]} for some nn\in\mathbb{N} such that jEjEj=I\sum_{j}E_{j}^{\dagger}E_{j}=I and (ρ)=j[n]EjρEj\mathcal{E}(\rho)=\sum_{j\in[n]}E_{j}\rho E_{j}^{\dagger} for all density operators ρ\rho. This is referred to as the Kraus representation of quantum channels. For instance, a quantum (logic) gate with unitary operator UU can be represented by (ρ)=UρU\mathcal{E}(\rho)=U\rho U^{\dagger}. Similarly, if we discard or delete the outcome of measurement M={Ki}i[m]M=\{K_{i}\}_{i\in[m]}, the quantum state transition can be modelled by quantum channel (ρ)=i[k]KiρKi\mathcal{E}(\rho)=\sum_{i\in[k]}K_{i}\rho K_{i}^{\dagger}. We define the tensor product of quantum channels 1\mathcal{E}_{1} and 2\mathcal{E}_{2} as 12(ρ1ρ2):=1(ρ1)2(ρ2)\mathcal{E}_{1}\otimes\mathcal{E}_{2}(\rho_{1}\otimes\rho_{2}):=\mathcal{E}_{1}(\rho_{1})\otimes\mathcal{E}_{2}(\rho_{2}) for all density operators ρ1\rho_{1} and ρ2\rho_{2}.

The trace norm or Schatten 1-norm of any linear operator MM is defined as M1:=tr(|M|)=tr(MM)\|M\|_{1}:=\tr(|M|)=\tr(\sqrt{M^{\dagger}M}). Based on this, we can define the trace distance between any two density operators ρ\rho and σ\sigma with 𝒯(ρ,σ):=12ρσ1[0,1].\mathcal{T}(\rho,\sigma):=\frac{1}{2}\|\rho-\sigma\|_{1}\in[0,1]. Recall that density operators belong to the set of linear operators (i.e., matrices). The distance is equal to zero when two quantum states are equal. However, the distance attains its maximum value when two quantum states have support on orthogonal subspaces. For υ[0,1]\upsilon\in[0,1], the υ\upsilon-relative entropy between two quantum states ρ\rho and σ\sigma is defined as Dυ(ρσ)=log(min{tr(Qσ)|0QI,tr(Qρ)1υ}).D^{\upsilon}(\rho\|\sigma)\!=\!-\!\log\left(\min\{\tr(Q\sigma)|0\!\preceq\!Q\!\preceq\!I,\tr(Q\rho)\geq 1\!-\!\upsilon\}\right). The υ\upsilon-relative entropy satisfies a few important properties that we will use in this paper. These properties are borrowed from [22]. First, Dυ(ρσ)0D^{\upsilon}(\rho\|\sigma)\geq 0 with equality if ρ=σ\rho=\sigma and υ=0\upsilon=0. Second, υ\upsilon-relative entropy enjoys data processing inequality, i.e., Dυ((ρ)(σ))Dυ(ρσ)D^{\upsilon}(\mathcal{E}(\rho)\|\mathcal{E}(\sigma))\leq D^{\upsilon}(\rho\|\sigma) for all density operators ρ,σ\rho,\sigma and all quantum channels \mathcal{E}. Also, Dυ(ρσ)(S(ρσ)+Hb(υ))/(1υ),D^{\upsilon}(\rho\|\sigma)\leq({S(\rho\|\sigma)+H_{b}(\upsilon)})/({1-\upsilon}), where Hb(υ)=υlog(υ)(1υ)log(1υ)H_{b}(\upsilon)=-\upsilon\log(\upsilon)-(1-\upsilon)\log(1-\upsilon) is the binary entropy function and S(ρσ):=tr(ρ(log(ρ)log(σ)))S(\rho\|\sigma):=\tr(\rho(\log(\rho)-\log(\sigma))) is the usual relative entropy in quantum information theory. The υ\upsilon-relative entropy and the trace distance also satisfy the following relationship ν/(1ν)ρσ1Dυ(ρσ)\nu/(1-\nu)\|\rho-\sigma\|_{1}\leq D^{\upsilon}(\rho\|\sigma) [23]. The smooth max-relative entropy is defined as Dmaxυ(ρσ)=infτυ(ρ)Dmax(τσ)D^{\upsilon}_{\max}(\rho\|\sigma)=\inf_{\tau\in\mathcal{B}^{\upsilon}(\rho)}D_{\max}(\tau\|\sigma), where Dmax(τσ)=inf{λ0|ρexp(λ)σ}D_{\max}(\tau\|\sigma)=\inf\{\lambda\geq 0\,|\,\rho\preceq\exp(\lambda)\sigma\} and υ(ρ):={τ|τ=τ0,ρτ12υ}\mathcal{B}^{\upsilon}(\rho):=\{\tau\,|\,\tau^{\dagger}=\tau\succeq 0,\|\rho-\tau\|_{1}\leq 2\upsilon\}.

Depolarizing channel is an important type of quantum noise that is represented by

Dep(ρ):=pDI+(1p)ρ,\displaystyle\mathcal{E}_{\rm Dep}(\rho):=\frac{p}{D}I+(1-p)\rho, (1)

where DD is the dimension of the Hilbert space to which the system belongs and p[0,1]p\in[0,1] is a probability parameter.

III Quantum Hypothesis Testing and Privacy

Consider a quantum hypothesis testing scenario where a decision maker aims to distinguish between two quantum states ρ\rho (null hypothesis) and σ\sigma (alternative hypothesis). This is done by performing POVM M:={M1,M2}M:=\{M_{1},M_{2}\} with M1+M2=IM_{1}+M_{2}=I and 0MiI0\preceq M_{i}\preceq I for i=1,2i=1,2. If measurement outcome corresponding to the operator M1M_{1} is realized, the decision maker guesses that the state is ρ\rho while, if measurement outcome corresponding to the operator M2M_{2} is realized, the decision maker guesses that the state is σ\sigma. The probability of a type-I error (false positive) is equal

α(M2):=tr(M2ρ).\displaystyle\alpha(M_{2}):=\tr(M_{2}\rho). (2)

The probability of a type-II error (false negative) is given by

β(M1):=tr(M1σ).\displaystyle\beta(M_{1}):=\tr(M_{1}\sigma). (3)

The optimal test, which seeks to minimize the false negative probability subject to a constraint on maintaining the false positive probability below η[0,1]\eta\in[0,1], is given by

βη(ρ,σ):=minM1,M20\displaystyle\beta_{\eta}(\rho,\sigma):=\min_{M_{1},M_{2}\succeq 0} β(M1),\displaystyle\;\beta(M_{1}), (4a)
s.t.\displaystyle\mathrm{s.t.}\quad M1+M2=I,\displaystyle\;M_{1}+M_{2}=I, (4b)
α(M2)η.\displaystyle\;\alpha(M_{2})\leq\eta. (4c)

This is referred to as asymmetric quantum hypothesis testing [24]. The following well-known result (see, e.g., [22]) can be easily derived based on the definition of βη(ρ,σ)\beta_{\eta}(\rho,\sigma) and η\eta-relative entropy Dη(ρσ)D^{\eta}(\rho\|\sigma).

Proposition 1

βη(ρ,σ)=2Dη(ρσ).\beta_{\eta}(\rho,\sigma)=2^{-D^{\eta}(\rho\|\sigma)}.

Proof:

Note that

βη(ρ,σ)=\displaystyle\beta_{\eta}(\rho,\sigma)\!= minM1,M20{tr(M1σ)|M1+M2=I,tr(M2ρ)η}\displaystyle\min_{M_{1},M_{2}\succeq 0}\{\tr(M_{1}\sigma)|M_{1}\!+\!M_{2}\!=\!I,\tr(M_{2}\rho)\!\leq\!\eta\}
=\displaystyle= minIM10{tr(M1σ)|tr((IM1)ρ)η}\displaystyle\min_{I\succeq M_{1}\succeq 0}\{\tr(M_{1}\sigma)|\tr((I-M_{1})\rho)\leq\eta\}
=\displaystyle= minIM10{tr(M1σ)|1ηtr(M1ρ)}\displaystyle\min_{I\succeq M_{1}\succeq 0}\{\tr(M_{1}\sigma)|1-\eta\leq\tr(M_{1}\rho)\}
=\displaystyle= 2Dη(ρσ).\displaystyle 2^{-D^{\eta}(\rho\|\sigma)}.

This concludes the proof. ∎

Alternatively, a combination of false positive and false negative probabilities can be minimized:

perr(ρ,σ):=minM1,M20\displaystyle p_{\rm err}(\rho,\sigma):=\min_{M_{1},M_{2}\succeq 0} pρα(M2)+pσβ(M1),\displaystyle\;p_{\rho}\alpha(M_{2})+p_{\sigma}\beta(M_{1}),
s.t.\displaystyle\mathrm{s.t.}\quad M1+M2=I,\displaystyle\;M_{1}+M_{2}=I,

where pρ[0,1]p_{\rho}\in[0,1] and pσ[0,1]p_{\sigma}\in[0,1], respectively, denote the prior probability that quantum state ρ\rho and the prior probability that quantum state σ\sigma are prepared. Clearly, by construct, pρ+pσ=1p_{\rho}+p_{\sigma}=1. This is referred to as symmetric quantum hypothesis testing [24].

Theorem 1 (Helstrom-Holevo theorem [21, p. 254-255])

perr(ρ,σ)=12(1pρρpσσ1).p_{\rm err}(\rho,\sigma)=\frac{1}{2}\left(1-\|p_{\rho}\rho-p_{\sigma}\sigma\|_{1}\right).

The most indistinguishable quantum states are ρ=σ\rho=\sigma. In this case, a decision maker would not be able to identify the quantum states because their observable are equivalent. Therefore, we can define

pmax:=\displaystyle p_{\rm max}:= perr(ρ,ρ)\displaystyle p_{\rm err}(\rho,\rho)
=\displaystyle= 12(1(pρpσ)ρ1)\displaystyle\frac{1}{2}\left(1-\|(p_{\rho}-p_{\sigma})\rho\|_{1}\right)
=\displaystyle= 12(1|pρpσ|ρ1)\displaystyle\frac{1}{2}\left(1-|p_{\rho}-p_{\sigma}|\|\rho\|_{1}\right)
=\displaystyle= 12(1|pρpσ|).\displaystyle\frac{1}{2}\left(1-|p_{\rho}-p_{\sigma}|\right).

Therefore, for general density operators, we have

perr(ρ,σ)=pmax+12(|pρpσ|pρρpσσ1).\displaystyle p_{\rm err}(\rho,\sigma)=p_{\rm max}+\frac{1}{2}\left(|p_{\rho}-p_{\sigma}|-\|p_{\rho}\rho-p_{\sigma}\sigma\|_{1}\right).

In quantum data privacy, it is desired to protect the quantum state of a system (which is being used for quantum computation) from being accurately estimated. Particularly, given a quantum state ρ\rho, we want to make sure that no decision maker can identify whether the quantum state of the system is ρ\rho or another similar quantum state σ\sigma. Similarity is modelled or captured using the neighbourhood relationship, c.f., differential privacy [20].

Definition 1 (Neighbouring Relationship)

A neighbouring or similarity relationship over the set of density operators is a mathematical relation that is both reflective and symmetric. The notation ρσ\rho\sim\sigma signifies that two quantum states ρ\rho and σ\sigma are neighbouring or similar. Note that, by definition, ρρ\rho\sim\rho (reflectivity) and ρσ\rho\sim\sigma implies σρ\sigma\sim\rho (symmetry).

An example of neighbouring or similarity relationship is the notion defined using trace distance in [18]. In this case, we say ρσ\rho\sim\sigma if and only if 𝒯(ρ,σ)d\mathcal{T}(\rho,\sigma)\leq d for some constant d>0d>0. However, we may select another notion of similarity that ensures that two quantum states are neighbouring if they are constructed based on two private datasets that differ in the data of one individual. Such a definition is well-suited for quantum machine learning with privacy guarantees [25].

Definition 2 ((ε,η)(\varepsilon,\eta)-Privacy Against Hypothesis-Testing Adversary)

For any ε0\varepsilon\geq 0 and η[0,1]\eta\in[0,1], a quantum channel \mathcal{E} is (ε,η)(\varepsilon,\eta)-private (against hypothesis-testing adversary) if Dη((ρ)(σ))εD^{\eta}(\mathcal{E}(\rho)\|\mathcal{E}(\sigma))\leq\varepsilon for all neighbouring states ρσ\rho\sim\sigma.

This definition implies that, if two states are similar ρσ\rho\sim\sigma, a quantum channel \mathcal{E} is private if it makes distinguishing the reported or output states (ρ)\mathcal{E}(\rho) and (σ)\mathcal{E}(\sigma) difficult by any decision maker. In fact, Proposition 1 shows that probability of false negatives β(M1)\beta(M_{1}) for any detection mechanism M={M1,M2}M=\{M_{1},M_{2}\} is lower bounded by 2ϵ2^{-\epsilon} if the probability of false positives bounded by α(M2)η\alpha(M_{2})\leq\eta. Therefore, as ϵ\epsilon tends to zero (privacy guarantee is strengthened/privacy budget is reduced), the probability of false negatives move towards one (i.e., the decision maker would become overwhelmed by false negatives).

Proposition 2

Assume that a quantum channel \mathcal{E} is (ε,η)(\varepsilon,\eta)-private. Then, the quantum channel \mathcal{E} is (ε,η)(\varepsilon^{\prime},\eta^{\prime})-private if ηη\eta^{\prime}\geq\eta and εε\varepsilon\leq\varepsilon^{\prime}.

Proof:

First note that, if ηη\eta^{\prime}\geq\eta, we have

2Dη(ρ,σ)\displaystyle 2^{-D^{\eta}(\rho,\sigma)} =min{tr(Qσ)|0QI,tr(Qρ)1η}\displaystyle=\min\{\tr(Q\sigma)|0\!\preceq\!Q\!\preceq\!I,\tr(Q\rho)\geq 1\!-\!\eta\}
min{tr(Qσ)|0QI,tr(Qρ)1η}\displaystyle\leq\min\{\tr(Q\sigma)|0\!\preceq\!Q\!\preceq\!I,\tr(Q\rho)\geq 1\!-\!\eta^{\prime}\}
=2Dη(ρ,σ),\displaystyle=2^{-D^{\eta^{\prime}}(\rho,\sigma)},

where the inequality follows from that {Q|0QI,tr(Qρ)1η}{Q|0QI,tr(Qρ)1η}\{Q|0\!\preceq\!Q\!\preceq\!I,\tr(Q\rho)\geq 1\!-\!\eta\}\subseteq\{Q|0\!\preceq\!Q\!\preceq\!I,\tr(Q\rho)\geq 1\!-\!\eta^{\prime}\}. Therefore, for all σρ\sigma\sim\rho, we get Dη(ρ,σ)Dη(ρ,σ)ϵϵ.D^{\eta^{\prime}}(\rho,\sigma)\leq D^{\eta}(\rho,\sigma)\leq\epsilon\leq\epsilon^{\prime}.

The following corollary, building on Proposition 2, shows that (ε,0)(\varepsilon,0)-privacy against hypothesis testing adversary is the strongest notion of privacy and thus, (ε,η)(\varepsilon,\eta)-privacy can be thought of as relaxations of (ε,0)(\varepsilon,0)-privacy.

Corollary 1

Assume that a quantum channel \mathcal{E} is (ε,0)(\varepsilon,0)-private. Then, the quantum channel \mathcal{E} is (ε,η)(\varepsilon,\eta)-private for all η[0,1]\eta\in[0,1].

Although privacy is here defined in terms of asymmetric quantum hypothesis testing, we prove the following important bound on the power of symmetric quantum hypothesis testing.

Theorem 2

For any (ε,η)(\varepsilon,\eta)-private quantum channel \mathcal{E},

perr((ρ),(σ))Γpρ,pσ(ε,η),\displaystyle p_{\rm err}(\mathcal{E}(\rho),\mathcal{E}(\sigma))\geq\Gamma_{p_{\rho},p_{\sigma}}(\varepsilon,\eta), (5)

where

Γpρ,pσ(ε,η):=\displaystyle\Gamma_{p_{\rho},p_{\sigma}}(\varepsilon,\eta):= max{pmaxεmin{pρ,pσ}(1η)2η,0}.\displaystyle\max\left\{p_{\rm max}\!-\!\frac{\varepsilon\min\{p_{\rho},p_{\sigma}\}(1\!-\!\eta)}{2\eta},0\right\}.
Proof:

Using [23], we have

η1η(ρ)(σ)1Dη((ρ)(σ)).\displaystyle\frac{\eta}{1-\eta}\|\mathcal{E}(\rho)-\mathcal{E}(\sigma)\|_{1}\leq D^{\eta}(\mathcal{E}(\rho)\|\mathcal{E}(\sigma)).

Therefore, if \mathcal{E} is (ε,η)(\varepsilon,\eta)-private (against hypothesis-testing adversary), we get

(ρ)(σ)11ηηε.\displaystyle\|\mathcal{E}(\rho)-\mathcal{E}(\sigma)\|_{1}\leq\frac{1-\eta}{\eta}\varepsilon.

We have

pρ(ρ)pσ(σ)1=\displaystyle\|p_{\rho}\mathcal{E}(\rho)\!-\!p_{\sigma}\mathcal{E}(\sigma)\|_{1}\!= pσpρpσ(ρ)(σ)1\displaystyle p_{\sigma}\left\|\frac{p_{\rho}}{p_{\sigma}}\mathcal{E}(\rho)\!-\!\mathcal{E}(\sigma)\right\|_{1}
=\displaystyle= pσpρpσpσ(ρ)+(ρ)(σ)1\displaystyle p_{\sigma}\left\|\frac{p_{\rho}\!-\!p_{\sigma}}{p_{\sigma}}\mathcal{E}(\rho)\!+\!\mathcal{E}(\rho)\!-\!\mathcal{E}(\sigma)\right\|_{1}
\displaystyle\leq |pρpσ|(ρ)1+pσ(ρ)(σ)1\displaystyle|p_{\rho}\!-\!p_{\sigma}|\|\mathcal{E}(\rho)\|_{1}\!+\!p_{\sigma}\|\mathcal{E}(\rho)\!-\!\mathcal{E}(\sigma)\|_{1}
\displaystyle\leq |pρpσ|+1ηηεpσ.\displaystyle|p_{\rho}\!-\!p_{\sigma}|\!+\!\frac{1-\eta}{\eta}\varepsilon p_{\sigma}. (6)

Following the same line of reasoning, we can also show that

pρ(ρ)pσ(σ)1|pρpσ|+1ηηεpρ.\displaystyle\|p_{\rho}\mathcal{E}(\rho)-p_{\sigma}\mathcal{E}(\sigma)\|_{1}\leq|p_{\rho}-p_{\sigma}|+\frac{1-\eta}{\eta}\varepsilon p_{\rho}. (7)

Combining (6) and (7), we get

pρ(ρ)pσ(σ)1|pρpσ|+1ηηεmin{pρ,pσ}.\displaystyle\|p_{\rho}\mathcal{E}(\rho)-p_{\sigma}\mathcal{E}(\sigma)\|_{1}\leq|p_{\rho}-p_{\sigma}|+\frac{1-\eta}{\eta}\varepsilon\min\{p_{\rho},p_{\sigma}\}.

Therefore,

perr((ρ),(σ))=\displaystyle p_{\rm err}(\mathcal{E}(\rho),\mathcal{E}(\sigma))= pmax+12(|pρpσ|\displaystyle p_{\rm max}+\frac{1}{2}\Big{(}|p_{\rho}-p_{\sigma}|
pρ(ρ)pσ(σ)1)\displaystyle\hskip 50.58878pt-\|p_{\rho}\mathcal{E}(\rho)-p_{\sigma}\mathcal{E}(\sigma)\|_{1}\Big{)}
\displaystyle\geq pmaxεmin{pρ,pσ}(1η)2η.\displaystyle p_{\rm max}-\frac{\varepsilon\min\{p_{\rho},p_{\sigma}\}(1-\eta)}{2\eta}.

This concludes the proof. ∎

Theorem 2 shows that, by decreasing ε\varepsilon, the combined probabilities of false positive and false negative denoted by perr((ρ),(σ))p_{\rm err}(\mathcal{E}(\rho),\mathcal{E}(\sigma)) increases towards its maximum value pmaxp_{\max}. Figure 1 illustrates the lower bound Γpρ,pσ(ε,η)\Gamma_{p_{\rho},p_{\sigma}}(\varepsilon,\eta) on perr((ρ),(σ))p_{\rm err}(\mathcal{E}(\rho),\mathcal{E}(\sigma)) versus the privacy budget ε\varepsilon for various choices of η\eta for the case that pρ=pσ=12p_{\rho}=p_{\sigma}=\frac{1}{2}. As expected, reducing the privacy budget ε\varepsilon strengthens the privacy guarantees.

Refer to captionΓ12,12(ε,η)\Gamma_{\frac{1}{2},\frac{1}{2}}(\varepsilon,\eta) in Theorem 2 ε\varepsilon
Figure 1: The lower bound Γ12,12(ε,η)\Gamma_{\frac{1}{2},\frac{1}{2}}(\varepsilon,\eta) on perr((ρ),(σ))p_{\rm err}(\mathcal{E}(\rho),\mathcal{E}(\sigma)) in Theorem 2 versus the privacy budget ε\varepsilon for various choices of η\eta. As expected, reducing the privacy budget ε\varepsilon strengthens the privacy guarantees.

It is stipulated that any useful notion of privacy should admit two important properties of post processing and composition [20]. In the remainder of this section, we discuss these properties and their application to privacy against hypothesis testing adversary.

Theorem 3 (Post Processing)

Let \mathcal{E} be any (ε,η)(\varepsilon,\eta)-private and 𝒩\mathcal{N} be an arbitrary quantum channel, then 𝒩\mathcal{N}\circ\mathcal{E} is (ε,η)(\varepsilon,\eta)-private.

Proof:

The proof follows from that, for all ρ\rho and σ\sigma, Dη(𝒩((ρ))𝒩((σ)))Dη(((ρ)(σ))D^{\eta}(\mathcal{N}(\mathcal{E}(\rho))\|\mathcal{N}(\mathcal{E}(\sigma)))\leq D^{\eta}(\mathcal{E}((\rho)\|\mathcal{E}(\sigma)) [22]. ∎

Theorem 3 shows that an adversary cannot weaken the privacy guarantees by processing the received quantum information in any way.

Theorem 4 (Composition)

Let 1\mathcal{E}_{1} be any (ε1,0)(\varepsilon_{1},0)-private and 2\mathcal{E}_{2} be any (ε2,0)(\varepsilon_{2},0)-private. Assume that ρ1ρ2σ1σ2\rho_{1}\otimes\rho_{2}\sim\sigma_{1}\otimes\sigma_{2} if ρ1σ1\rho_{1}\sim\sigma_{1} and ρ2σ2\rho_{2}\sim\sigma_{2}. Then, Then 12\mathcal{E}_{1}\otimes\mathcal{E}_{2} is (ε1+ε2,0)(\varepsilon_{1}+\varepsilon_{2},0)-private.

Proof:

Using the additivity results in [26, Appendix A], we get D0(ρ1ρ2σ1σ2)=D0(ρ1σ1)+D0(ρ2σ2).D^{0}(\rho_{1}\otimes\rho_{2}\|\sigma_{1}\otimes\sigma_{2})=D^{0}(\rho_{1}\|\sigma_{1})+D^{0}(\rho_{2}\|\sigma_{2}). Therefore, if D0(ρiσi)εiD^{0}(\rho_{i}\|\sigma_{i})\leq\varepsilon_{i} for i=1,2i=1,2, then D0(ρ1ρ2σ1σ2)ε1+ε2D^{0}(\rho_{1}\otimes\rho_{2}\|\sigma_{1}\otimes\sigma_{2})\leq\varepsilon_{1}+\varepsilon_{2}. ∎

In practical data processing applications, there is often a need to deal with complicated algorithms in which responses from several queries based on private user data are fused together to extract useful statistical information from the data. For instance, when training machine learning models, iterative gradient descent algorithms can be used and the gradient at each epoch can be modelled as a query on the private data used for training [27]. In this case, it is desirable to establish composition rules for combination of several privacy-preserving quantum operations. Theorem 4 provides such a result for privacy against hypothesis testing adversaries.

IV Quantum Differential Privacy

The gold standard of privacy analysis and enforcement in the computer science literature is differential privacy, which has been recently extended to quantum computing algorithms [18]. In this section, we establish a relationship between differential privacy and privacy against hypothesis testing adversaries.

Definition 3

For any ϵ,δ0\epsilon,\delta\geq 0, a quantum channel \mathcal{E} is (ϵ,δ)(\epsilon,\delta)-differentially private if

tr(M(ρ))exp(ϵ)tr(M(σ))+δ,\displaystyle\tr(M\mathcal{E}(\rho))\leq\exp(\epsilon)\tr(M\mathcal{E}(\sigma))+\delta, (8)

for all measurements 0MI0\preceq M\preceq I and neighbouring density operators ρσ\rho\sim\sigma.

We can prove the following result regarding the relationship between quantum differential privacy and privacy against hypothesis testing adversaries.

Theorem 5

The following two statements hold:

  • If \mathcal{E} be (ε,η)(\varepsilon,\eta)-private, then \mathcal{E} is (ε,2η)(\varepsilon,\sqrt{2\eta})-differentially private.

  • If \mathcal{E} be (ϵ,0)(\epsilon,0)-differentially private, then \mathcal{E} is (ε,η)(\varepsilon,\eta)-private for all η[0,1]\eta\in[0,1].

Proof:

First, Dmax2ν((ρ),(σ))Dν((ρ),(σ))D_{\max}^{\sqrt{2\nu}}(\mathcal{E}(\rho),\mathcal{E}(\sigma))\leq D^{\nu}(\mathcal{E}(\rho),\mathcal{E}(\sigma)) [23, Proposition 4.1]. Therefore, if \mathcal{E} is (ε,η)(\varepsilon,\eta)-private, we get Dmax2η((ρ),(σ))Dη((ρ),(σ))εD_{\max}^{\sqrt{2\eta}}(\mathcal{E}(\rho),\mathcal{E}(\sigma))\leq D^{\eta}(\mathcal{E}(\rho),\mathcal{E}(\sigma))\leq\varepsilon. From Lemma III.2 in [20], a quantum channel \mathcal{E} is (ϵ,δ)(\epsilon,\delta)-differentially private if and only if Dmaxδ((ρ),(σ))ϵD_{\max}^{\delta}(\mathcal{E}(\rho),\mathcal{E}(\sigma))\leq\epsilon. This proves that \mathcal{E} is (ε,2η)(\varepsilon,\sqrt{2\eta})-differentially private.

For the second part, note that Dη((ρ),(σ))Dmax0((ρ),(σ))D^{\eta}(\mathcal{E}(\rho),\mathcal{E}(\sigma))\leq D_{\max}^{0}(\mathcal{E}(\rho),\mathcal{E}(\sigma)) [23, Proposition 4.1]. Therefore, if \mathcal{E} is (ϵ,0)(\epsilon,0)-differentially private, we have Dη((ρ),(σ))Dmax0((ρ),(σ))ϵD^{\eta}(\mathcal{E}(\rho),\mathcal{E}(\sigma))\leq D_{\max}^{0}(\mathcal{E}(\rho),\mathcal{E}(\sigma))\leq\epsilon. This implies that \mathcal{E} is (ϵ,η)(\epsilon,\eta)-private for all η[0,1]\eta\in[0,1]. ∎

Theorem 6 (Lemma IV.2 [20])

Consider neighbourhood notion that ρσ\rho\sim\sigma if 𝒯(ρ,σ)d\mathcal{T}(\rho,\sigma)\leq d. Then, the depolarizing channel Dep(ρ)\mathcal{E}_{\rm Dep}(\rho) is (ϵ,δ)(\epsilon,\delta)-differentially private with δ=max{0,(1exp(ϵ))p/D+(1p)κ}.\delta=\max\{0,(1-\exp(\epsilon))p/D+(1-p)\kappa\}.

Corollary 2

Consider neighbourhood notion that ρσ\rho\sim\sigma if 𝒯(ρ,σ)d\mathcal{T}(\rho,\sigma)\leq d. Then, the depolarizing channel Dep(ρ)\mathcal{E}_{\rm Dep}(\rho) is (ε,η)(\varepsilon,\eta)-private with ε=log(1+(1p)Dκ/p)\varepsilon=\log(1+(1-p)D\kappa/p) and all η[0,1]\eta\in[0,1].

Proof:

First, note that Theorem 6 shows that the depolarizing channel Dep(ρ)\mathcal{E}_{\rm Dep}(\rho) is (ϵ,δ)(\epsilon,\delta)-differentially private with δ=max{0,(1exp(ϵ))p/D+(1p)κ}.\delta=\max\{0,(1-\exp(\epsilon))p/D+(1-p)\kappa\}. If we select ϵ=log(1+(1p)Dκ/p)\epsilon=\log(1+(1-p)D\kappa/p), we get δ=0\delta=0. Using Theorem 5, then \mathcal{E} is (ϵ,η)(\epsilon,\eta)-private for all η[0,1]\eta\in[0,1]. ∎

We finish this section with analysing the performance of hypothesis testing adversaries for differentially-private quantum channels.

Theorem 7

For any (ϵ,δ)(\epsilon,\delta)-differentially private quantum channel \mathcal{E},

βη((ρ),(σ))Ωη(ϵ,δ),\displaystyle\beta_{\eta}(\mathcal{E}(\rho),\mathcal{E}(\sigma))\geq\Omega_{\eta}(\epsilon,\delta), (9)

where Ωη(ϵ,δ):=exp(ϵ)(1ηδ).\Omega_{\eta}(\epsilon,\delta):=\exp(-\epsilon)(1-\eta-\delta).

Proof:

Assume that ρσ\rho\sim\sigma. Because of (ϵ,δ)(\epsilon,\delta)-differential privacy, tr(M(σ))exp(ϵ)(tr(M(ρ))δ)\tr(M\mathcal{E}(\sigma))\geq\exp(-\epsilon)(\tr(M\mathcal{E}(\rho))-\delta) for all measurements 0MI0\preceq M\preceq I. Therefore, βη((ρ),(σ))=minIM0{tr(M(σ))|1ηtr(M(ρ))}exp(ϵ)(1ηδ).\beta_{\eta}(\mathcal{E}(\rho),\mathcal{E}(\sigma))\!=\!\min_{I\succeq M\succeq 0}\{\tr(M\mathcal{E}(\sigma))|1\!-\!\eta\leq\tr(M\mathcal{E}(\rho))\}\geq\!\exp(-\epsilon)(1-\eta-\delta).

Refer to captionΩη(ϵ,δ)\Omega_{\eta}(\epsilon,\delta) in Theorem 7Refer to captionΘ(ϵ,δ)\Theta(\epsilon,\delta) in Theorem 8ϵ\epsilonϵ\epsilon
Figure 2: Lower bound on βη((ρ),(σ))\beta_{\eta}(\mathcal{E}(\rho),\mathcal{E}(\sigma)) in Theorem 7 [top] and lower bound on perr((ρ),(σ))p_{\rm err}(\mathcal{E}(\rho),\mathcal{E}(\sigma)) in Theorem 8 [bottom] versus privacy budget ϵ\epsilon for various choices of δ\delta.

Theorem 7 provides a lower bound for the false negative rate for the best asymmetric hypothesis testing mechanism. The lower bound grows, and thus the decision maker would get overwhelmed by false negatives, with decreasing ϵ\epsilon and δ\delta. Therefore, the privacy guarantees strengthens as the privacy budget reduces in quantum differential privacy. This is illustrated in Figure 2 [top].

Theorem 8

For any (ϵ,δ)(\epsilon,\delta)-differentially private quantum channel \mathcal{E},

perr((ρ),(σ))\displaystyle p_{\rm err}(\mathcal{E}(\rho),\mathcal{E}(\sigma))\!\geq Θ(ϵ,δ),\displaystyle\Theta(\epsilon,\delta), (10)

where Θ(ϵ,δ):=max{pmax+max{pρ,pσ}(1exp(ϵ)δ),0}.\Theta(\epsilon,\delta):=\max\{p_{\rm max}+\max\{p_{\rho},p_{\sigma}\}(1-\exp(\epsilon)-\delta),0\}.

Proof:

First, assume that pρpσp_{\rho}\geq p_{\sigma}. The definition of differential privacy implies that tr(Λ(ρ))exp(ϵ)tr(Λ(σ))+δ\tr(\Lambda\mathcal{E}(\rho))\leq\exp(\epsilon)\tr(\Lambda\mathcal{E}(\sigma))+\delta for all 0ΛI0\preceq\Lambda\preceq I. As a result,

tr(Λ(pρ(ρ)pσ(σ)))\displaystyle\tr(\Lambda(p_{\rho}\mathcal{E}(\rho)-p_{\sigma}\mathcal{E}(\sigma)))\leq (pρexp(ϵ)pσ)tr(Λ(σ))+pρδ\displaystyle(p_{\rho}\exp(\epsilon)\!-\!p_{\sigma})\tr(\Lambda\mathcal{E}(\sigma))\!+\!p_{\rho}\delta
\displaystyle\leq pρexp(ϵ)pσ+pρδ,\displaystyle p_{\rho}\exp(\epsilon)-p_{\sigma}+p_{\rho}\delta,

where the last inequality follows from that pρexp(ϵ)pσpρpσ0p_{\rho}\exp(\epsilon)-p_{\sigma}\geq p_{\rho}-p_{\sigma}\geq 0 and that tr(Λ(σ))1\tr(\Lambda\mathcal{E}(\sigma))\leq 1 because 0ΛI0\preceq\Lambda\preceq I. Therefore, using Lemma 1 in the appendix, we have

12pρ(ρ)pσ(σ)1\displaystyle\frac{1}{2}\|p_{\rho}\mathcal{E}(\rho)-p_{\sigma}\mathcal{E}(\sigma)\|_{1}\leq pρexp(ϵ)pσ+pρδ+pσpρ2\displaystyle p_{\rho}\exp(\epsilon)-p_{\sigma}+p_{\rho}\delta+\frac{p_{\sigma}-p_{\rho}}{2}
\displaystyle\leq pρ(exp(ϵ)+δ)pσ+pρ2.\displaystyle p_{\rho}(\exp(\epsilon)+\delta)-\frac{p_{\sigma}+p_{\rho}}{2}. (11)

Alternatively, assume that pσpρp_{\sigma}\geq p_{\rho}. Following the same line of reasoning, we get

12pσ(σ)pρ(ρ)1\displaystyle\frac{1}{2}\|p_{\sigma}\mathcal{E}(\sigma)-p_{\rho}\mathcal{E}(\rho)\|_{1}\leq pσ(exp(ϵ)+δ)pσ+pρ2.\displaystyle p_{\sigma}(\exp(\epsilon)+\delta)-\frac{p_{\sigma}+p_{\rho}}{2}. (12)

Combining (11) and (12) gives

12pσ(σ)pρ(ρ)1\displaystyle\frac{1}{2}\|p_{\sigma}\mathcal{E}(\sigma)-p_{\rho}\mathcal{E}(\rho)\|_{1}\leq max{pρ,pσ}(exp(ϵ)+δ)\displaystyle\max\{p_{\rho},p_{\sigma}\}(\exp(\epsilon)+\delta)
pσ+pρ2.\displaystyle-\frac{p_{\sigma}+p_{\rho}}{2}.

Therefore,

|pρ\displaystyle|p_{\rho}- pσ|pρ(ρ)pσ(σ)1\displaystyle p_{\sigma}|-\|p_{\rho}\mathcal{E}(\rho)-p_{\sigma}\mathcal{E}(\sigma)\|_{1}
\displaystyle\geq |pρpσ|+(pσ+pρ)2max{pρ,pσ}(exp(ϵ)+δ)\displaystyle|p_{\rho}-p_{\sigma}|+(p_{\sigma}+p_{\rho})-2\max\{p_{\rho},p_{\sigma}\}(\exp(\epsilon)+\delta)
=\displaystyle= 2max{pρ,pσ}(1exp(ϵ)δ).\displaystyle 2\max\{p_{\rho},p_{\sigma}\}(1-\exp(\epsilon)-\delta).

This concludes the proof. ∎

Theorem 8 provides a lower bound for the combined false positive and negative rates of the best symmetric hypothesis testing mechanism. The lower bound grows towards pmaxp_{\max} as ϵ\epsilon and δ\delta become smaller, which demonstrates that the privacy guarantees strengthen as the privacy budget reduces in quantum differential privacy. This is illustrated in Figure 2 [bottom].

V Conclusions and Future Work

We presented a novel definition for privacy in quantum computing based on quantum hypothesis testing. Important properties of post processing and composition were proved for this new notion of privacy. We then examined the relationship between privacy against hypothesis-testing adversaries, defined in this paper, and quantum differential privacy are then examined. In the composition rules for privacy against hypothesis adversaries, we only considered the case of η=0\eta=0. Future work can expand these results for general case of η[0,1]\eta\in[0,1]. Furthermore, we only showed that (ϵ,0)(\epsilon,0)-differential privacy can be translated to privacy against hypothesis testing adversaries (the inverse results are more general in this paper). Therefore, another avenue for future research is to expand these results to general (ϵ,δ)(\epsilon,\delta)-differential privacy. Finally, an important direction for future research is to use the proposed framework in numerical setups based on real-world data.

References

  • [1] P. W. Shor, “Algorithms for quantum computation: discrete logarithms and factoring,” in Proceedings 35th Annual Symposium on Foundations of Computer Science, pp. 124–134, Ieee, 1994.
  • [2] S. Lloyd, M. Mohseni, and P. Rebentrost, “Quantum principal component analysis,” Nature Physics, vol. 10, no. 9, pp. 631–633, 2014.
  • [3] H.-Y. Huang, R. Kueng, and J. Preskill, “Information-theoretic bounds on quantum advantage in machine learning,” Physical Review Letters, vol. 126, no. 19, p. 190505, 2021.
  • [4] J. Biamonte, P. Wittek, N. Pancotti, P. Rebentrost, N. Wiebe, and S. Lloyd, “Quantum machine learning,” Nature, vol. 549, no. 7671, pp. 195–202, 2017.
  • [5] M. Kearns and A. Roth, The ethical algorithm: The science of socially aware algorithm design. Oxford University Press, 2019.
  • [6] C. Dwork, F. McSherry, K. Nissim, and A. Smith, “Calibrating noise to sensitivity in private data analysis,” in Theory of Cryptography: Third Theory of Cryptography Conference, TCC 2006, New York, NY, USA, March 4-7, 2006. Proceedings 3, pp. 265–284, Springer, 2006.
  • [7] C. Dwork, “Differential privacy: A survey of results,” in Theory and Applications of Models of Computation: 5th International Conference, TAMC 2008, Xi’an, China, April 25-29, 2008. Proceedings 5, pp. 1–19, Springer, 2008.
  • [8] M. Abadi, A. Chu, I. Goodfellow, H. B. McMahan, I. Mironov, K. Talwar, and L. Zhang, “Deep learning with differential privacy,” in Proceedings of the 2016 ACM SIGSAC conference on computer and communications security, pp. 308–318, 2016.
  • [9] C. Dwork and A. Roth, “The algorithmic foundations of differential privacy,” Foundations and Trends® in Theoretical Computer Science, vol. 9, no. 3–4, pp. 211–407, 2014.
  • [10] V. J. Hotz, C. R. Bollinger, T. Komarova, C. F. Manski, R. A. Moffitt, D. Nekipelov, A. Sojourner, and B. D. Spencer, “Balancing data privacy and usability in the federal statistical system,” Proceedings of the National Academy of Sciences, vol. 119, no. 31, p. e2104906119, 2022.
  • [11] R. Bhaskar, A. Bhowmick, V. Goyal, S. Laxman, and A. Thakurta, “Noiseless database privacy,” in International Conference on the Theory and Application of Cryptology and Information Security, pp. 215–232, Springer, 2011.
  • [12] S. U. Nabar, B. Marthi, K. Kenthapadi, N. Mishra, and R. Motwani, “Towards robustness in query auditing,” in Proceedings of the 32nd international conference on Very large data bases, pp. 151–162, VLDB Endowment, 2006.
  • [13] I. Issa, A. B. Wagner, and S. Kamath, “An operational approach to information leakage,” IEEE Transactions on Information Theory, vol. 66, no. 3, pp. 1625–1657, 2019.
  • [14] F. Farokhi and N. Ding, “Measuring information leakage in non-stochastic brute-force guessing,” in 2020 IEEE Information Theory Workshop (ITW), pp. 1–5, IEEE, 2021.
  • [15] J. Liao, O. Kosut, L. Sankar, and F. du Pin Calmon, “Tunable measures for information leakage and applications to privacy-utility tradeoffs,” IEEE Transactions on Information Theory, vol. 65, no. 12, pp. 8043–8066, 2019.
  • [16] Z. Li, T. J. Oechtering, and D. Gündüz, “Privacy against a hypothesis testing adversary,” IEEE Transactions on Information Forensics and Security, vol. 14, no. 6, pp. 1567–1581, 2018.
  • [17] F. Farokhi and H. Sandberg, “Fisher information as a measure of privacy: Preserving privacy of households with smart meters using batteries,” IEEE Transactions on Smart Grid, vol. 9, no. 5, pp. 4726–4734, 2017.
  • [18] L. Zhou and M. Ying, “Differential privacy in quantum computation,” in 2017 IEEE 30th Computer Security Foundations Symposium (CSF), pp. 249–262, IEEE, 2017.
  • [19] S. Aaronson and G. N. Rothblum, “Gentle measurement of quantum states and differential privacy,” in Proceedings of the 51st Annual ACM SIGACT Symposium on Theory of Computing, pp. 322–333, 2019.
  • [20] C. Hirche, C. Rouzé, and D. S. França, “Quantum differential privacy: An information theory perspective,” arXiv preprint arXiv:2202.10717, 2022.
  • [21] M. Wilde, Quantum Information Theory. Quantum Information Theory, Cambridge University Press, 2013.
  • [22] L. Wang and R. Renner, “One-shot classical-quantum capacity and hypothesis testing,” Physical Review Letters, vol. 108, no. 20, p. 200501, 2012.
  • [23] F. Dupuis, L. Kraemer, P. Faist, J. M. Renes, and R. Renner, “Generalized entropies,” in XVIIth international congress on mathematical physics, pp. 134–153, World Scientific, 2014.
  • [24] B. Regula, L. Lami, and M. M. Wilde, “Postselected quantum hypothesis testing,” arXiv preprint arXiv:2209.10550, 2022.
  • [25] W. M. Watkins, S. Y.-C. Chen, and S. Yoo, “Quantum machine learning with differential privacy,” arXiv preprint arXiv:2103.06232, 2021.
  • [26] X. Yuan, “Hypothesis testing and entropies of quantum channels,” Physical Review A, vol. 99, no. 3, p. 032317, 2019.
  • [27] N. Wu, F. Farokhi, D. Smith, and M. A. Kaafar, “The value of collaboration in convex machine learning with differential privacy,” in 2020 IEEE Symposium on Security and Privacy (SP), pp. 304–317, IEEE, 2020.
Lemma 1

The following identity holds:

12pρρpσσ1=max0ΛItr(Λ(pρρpσσ))+pσpρ2.\displaystyle\frac{1}{2}\|p_{\rho}\rho-p_{\sigma}\sigma\|_{1}=\max_{0\preceq\Lambda\preceq I}\tr(\Lambda(p_{\rho}\rho-p_{\sigma}\sigma))+\frac{p_{\sigma}-p_{\rho}}{2}.
Proof:

The proof is similar to the standard argument for the trace distance. Note that the difference operator pρρpσσp_{\rho}\rho-p_{\sigma}\sigma is Hermitian. So we can diagonalize it as pρρpσσ=iλi|ii|p_{\rho}\rho-p_{\sigma}\sigma=\sum_{i}\lambda_{i}\ket{i}\bra{i}, where {|i}i\{\ket{i}\}_{i} is an orthonormal basis of eigenvectors and {λi}i\{\lambda_{i}\}_{i} is a set of real eigenvalues. Define matrices P:=i:λi>0λi|ii|0P:=\sum_{i:\lambda_{i}>0}\lambda_{i}\ket{i}\bra{i}\succeq 0 and Q:=i:λi<0(λi)|ii|0Q:=\sum_{i:\lambda_{i}<0}(-\lambda_{i})\ket{i}\bra{i}\succeq 0. Evidently, by construction, pρρpσσ=PQp_{\rho}\rho-p_{\sigma}\sigma=P-Q. Note that,

pρρpσσ1=\displaystyle\|p_{\rho}\rho-p_{\sigma}\sigma\|_{1}= tr(|pρρpσσ|)\displaystyle\tr(|p_{\rho}\rho-p_{\sigma}\sigma|)
=\displaystyle= tr(|PQ|)\displaystyle\tr(|P-Q|)
=\displaystyle= tr(P+Q)\displaystyle\tr(P+Q)
=\displaystyle= 2tr(P)+(pσpρ),\displaystyle 2\tr(P)+(p_{\sigma}-p_{\rho}),

where the last equality follows from

tr(P)tr(Q)=\displaystyle\tr(P)-\tr(Q)= tr(PQ)\displaystyle\tr(P-Q)
=\displaystyle= tr(pρρpσσ)\displaystyle\tr(p_{\rho}\rho-p_{\sigma}\sigma)
=\displaystyle= pρtr(ρ)pσtr(σ)\displaystyle p_{\rho}\tr(\rho)-p_{\sigma}\tr(\sigma)
=\displaystyle= pρpσ.\displaystyle p_{\rho}-p_{\sigma}.

For all 0ΛI0\preceq\Lambda\preceq I, we have

tr(Λ(pρρpσσ))=\displaystyle\tr(\Lambda(p_{\rho}\rho-p_{\sigma}\sigma))= tr(Λ(PQ))\displaystyle\tr(\Lambda(P-Q))
\displaystyle\leq tr(ΛP)\displaystyle\tr(\Lambda P)
\displaystyle\leq tr(P)\displaystyle\tr(P)
=\displaystyle= 12pρρpσσ1+pρpσ2,\displaystyle\frac{1}{2}\|p_{\rho}\rho-p_{\sigma}\sigma\|_{1}+\frac{p_{\rho}-p_{\sigma}}{2},

with equality achieved if P=i:λi>0|ii|P=\sum_{i:\lambda_{i}>0}\ket{i}\bra{i}. This implies that

12pρρpσσ1=max0ΛItr(Λ(pρρpσσ))+pσpρ2.\displaystyle\frac{1}{2}\|p_{\rho}\rho-p_{\sigma}\sigma\|_{1}=\max_{0\preceq\Lambda\preceq I}\tr(\Lambda(p_{\rho}\rho-p_{\sigma}\sigma))+\frac{p_{\sigma}-p_{\rho}}{2}.