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

Quantum Sampling for Optimistic Finite Key Rates in High Dimensional Quantum Cryptography

Keegan Yao Department of Computer Science and Engineering
University of Connecticut
Storrs, CT 06269 USA
Walter O. Krawec111Email: walter.krawec@gmail.com Department of Computer Science and Engineering
University of Connecticut
Storrs, CT 06269 USA
Jiadong Zhu Department of Computer Science and Engineering
University of Connecticut
Storrs, CT 06269 USA
Abstract

It has been shown recently that the framework of quantum sampling, as introduced by Bouman and Fehr, can lead to new entropic uncertainty relations highly applicable to finite-key cryptographic analyses. Here we revisit these so-called sampling-based entropic uncertainty relations, deriving newer, more powerful, relations and applying them to source-independent quantum random number generators and high-dimensional quantum key distribution protocols. Along the way, we prove several interesting results in the asymptotic case for our entropic uncertainty relations. These sampling-based approaches to entropic uncertainty, and their application to quantum cryptography, hold great potential for deriving proofs of security for quantum cryptographic systems, and the approaches we use here may be applicable to an even wider range of scenarios.

1 Introduction

Quantum sampling, as introduced by Bouman and Fehr in [1], is a framework allowing for the analysis of quantum systems through classical statistical sampling methods. Informally, it was shown that when sampling a quantum state (via measuring some subset of it in a particular basis), the remaining, unmeasured, portion of the state behaves like a superposition of states that are “close” (with respect to some target value such as Hamming weight) to the observed sample. How close they are depends, in fact, on the error probability of the classical sampling protocol used (where the classical sampling strategy would observe a portion of a classical word in some alphabet and argue about how the remaining, unobserved, portion of the word looks). At a high level, suppose one measures a random portion of some quantum state |ψ\ket{\psi} in the Z={|0,,|d1}Z=\{\ket{0},\cdots,\ket{d-1}\} basis and always observes |0\ket{0}. Then, one would expect that the remainder of the state (the unmeasured portion) should be a superposition of states that are relatively close to the all |00\ket{0\cdots 0} state. Bouman and Fehr’s framework formalizes this notion, even when the state is entangled with an environment system (e.g., an adversary).

Besides being fascinating on its own, there are now several interesting applications of this work. In their original paper [1], the authors showed some applications to quantum cryptography, namely a security proof of the entanglement-based BB84 QKD protocol for qubits (dimension two systems). Recently in [2, 3], we showed how the quantum sampling framework may be used to derive novel quantum entropic uncertainty relations which are highly applicable to finite-key quantum cryptographic security analyses. Informally, quantum entropic uncertainty relations bound the amount of uncertainty in two different measurement outcomes performed on some quantum system. For instance, the famous Maassen and Uffink relation [4] (which, itself, followed from a conjecture by Kraus in [5] and was an improvement over an uncertainty relation proposed first by Deutsch [6]) states that, given a quantum state ρ\rho acting on a dd-dimensional Hilbert space d\mathcal{H}_{d}, then if two measurements are performed on the system resulting in random variables MM and NN respectively, it holds that H(M)+H(N)γH(M)+H(N)\geq\gamma, where γ\gamma is a function of the two measurements performed (namely their overlap, though we will formally define this later for our applications). In particular, one cannot in general be certain of the outcome of both measurements of the system. By now there are numerous quantum entropic uncertainty relations with various fascinating properties and applications; for a general survey, the reader is referred to [7, 8, 9].

The so-called sampling-based entropic uncertainty relations we introduced in our earlier work [2, 3] turn out to be highly useful in finding optimistic secure bit generation rates for quantum random number generation (QRNG) protocols in the source-independent security model [10]. Our relations bounded the quantum min-entropy Hmin(A|E)H_{\text{min}}(A|E) as a function of the Shannon entropy of a particular measurement outcome and the measurement overlap. Since min entropy is a highly valuable resource in quantum cryptography (in particular, it can be used to determine how many uniform random bits one may extract from a source, independent of any adversary [11]), finding tight bounds on this quantity is highly desirable when analyzing quantum cryptographic protocols. As we’ve shown in our earlier work, our relations often out-perform prior work in cryptographic settings, producing more optimistic bit generation rates for QRNG protocols leading, potentially, to more rapid implementations of such systems (though here, and in our prior work, we focus only on theoretical analyses - practical settings, though interesting, are outside the scope of this current work). Furthermore, our sampling-based relations incorporate all needed finite sampling effects thus making them easy to use “out of the box.”

Here, we revisit sampling-based entropic uncertainty relations. These relations involve a quantum state ρ\rho, possibly entangled with an adversary, whereby a random sample is chosen and a test is performed by measuring a portion of ρ\rho resulting in some outcome qq. In this work, we show a highly general, two-party entropic uncertainty relation (Theorem 3.1) which, informally, states that with high probability (based on the failure probability of a classical sampling strategy):

Hminϵ(A|E)+log2|Jq|nγ,H_{\text{min}}^{\epsilon}(A|E)+\log_{2}|J_{q}|\geq n\gamma, (1)

where JqJ_{q} is the set of all words in some alphabet that are “close” to the observed string qq; nn is the number of qudits that were not measured in the test state; and γ\gamma is a function of the overlap between the two measurements. One of the strong advantages to our new sampling-based relation is that one may design classical sampling strategies suitable to a quantum cryptographic purpose and simply insert it directly into the above; all one needs to do is analyze the classical error probability and bound or evaluate the size of the set JqJ_{q} (which is typically a combinatorial proof). Though this result is more general than our original, it turns out the proof of this is nearly identical to our prior work in [2, 3]. However the novelty is, first, in the generality of the result that it works for any classical sampling strategy (whereas in [3] only a particular sampling strategy was proven); second in its applications, we show that this new bound is powerful enough to analyze a particular source-independent (a form of partial device independence introduced first in [10]) QRNG protocol producing more optimistic bit-generation rates than prior work using alternative entropic uncertainty relations and, furthermore, unlike our previous work, can provide an alternative proof of the previously mentioned Maassen-Uffink relation for dimensions strictly greater than 22 (in [2] we showed this for dimension 22 systems only).

Our second main contribution is to show a novel three-party sampling-based entropic uncertainty relation involving Alice, Bob, and Eve. Here, Alice and Bob perform a test measurement on some portion of their shared quantum state, resulting in outcome qAq_{A} and qBq_{B} respectively (these are words in some dd-character alphabet). Then, informally, our new entropic uncertainty relation (Theorem 4.1) states that, with high probability:

Hminϵ(A|E)+ηdHd[ΔH(qA,qB)+δ]n0γ+n1γ^,H_{\text{min}}^{\epsilon}(A|E)+\eta_{d}H_{d}\left[\Delta_{H}(q_{A},q_{B})+\delta\right]\geq n_{0}\gamma+n_{1}\hat{\gamma}, (2)

where n0+n1=nn_{0}+n_{1}=n, the number of systems not measured initially; ηd\eta_{d} is a constant depending on the dimension (dd) of the individual systems measured; δ\delta takes into account imperfect, finite samples; HdH_{d} is the dd-ary Shannon entropy; and ΔH(x,y)\Delta_{H}(x,y) is the Hamming distance of words xx and yy. Our entropic uncertainty relation can actually incorporate the maximal measurement overlap γ^\hat{\gamma} and the second-maximal overlap γ\gamma, making it useful if the two measurement bases have a similar basis element (e.g., a “vacuum” element, useful in QKD when considering channel loss). This ability shows the great promise in using the Quantum Sampling framework of Bouman and Fehr, augmented with our proof techniques developed here and in our prior work [2, 3] to prove interesting, and useful, entropic uncertainty relations. Indeed, our proof method can even be extended to support additional measurement overlap quantities.

Note that, if qA=qBq_{A}=q_{B}, then our result shows that the min-entropy conditioned on the adversary’s system EE must be high. We use our entropic uncertainty relation to provide a proof of security, in the finite key setting, of the High-Dimensional BB84 protocol [12, 13, 14, 15]. Our security proof is valid against arbitrary attacks by an adversary and applies easily to any dimension dd of the signal states and can even take into account lossy channels. Since high-dimensional QKD protocols exhibit many fascinating and useful properties (such as increased noise tolerance [13, 16]), and are experimentally feasible today [17, 18, 19, 20], our new analysis may provide even further benefits to these systems. We note that in [1], the sampling framework was used to provide a proof of security for the standard (qubit-based) BB84 using alternative methods which were specific to the qubit-BB84 protocol. Our method provides, first, a novel entropic uncertainty relation which may have numerous other applications to quantum cryptographic protocols outside of HD-BB84; and, secondly, provides as an application a simple proof of security for the high-dimensional variant of BB84 for any dimension dd of the system.

This work makes several contributions, not the least of which is showing yet further fascinating, and highly applicable, connections between the quantum sampling framework of Bouman and Fehr [1] and quantum information theory, in particular entropic uncertainty. Furthermore, our relations are immediately applicable to quantum cryptography in the finite key setting, leading to composable security [11] and, as we show, in most typical scenarios also highly optimistic secure bit-generation rates for source-independent QRNG protocols and QKD protocols. In practice, such sampling-based approaches show that quantum communication systems may run at higher bit-generation rates than previously thought. Thus, not only does this work provide interesting theoretical contributions, but also potential practical ones (though, as stated, we are not considering practical experimental imperfections here, leaving this as interesting future work). We suspect that there are even more connections and applications of the quantum sampling framework which may shed further light on problems in general information theory and applied quantum cryptography. This paper attempts to take a step forward in that direction.

1.1 Notation

We start with some notation and definitions that we will use throughout this work. An alphabet 𝒜d\mathcal{A}_{d} is a set of dd characters which we typically label {0,1,,d1}\{0,1,\cdots,d-1\}. Given a word q𝒜dnq\in\mathcal{A}_{d}^{n}, the substring qtq_{t} indexed by t{1,,n}t\subset\{1,\dots,n\} is the string qt=qt1qt2qt|t|q_{t}=q_{t_{1}}q_{t_{2}}\dots q_{t_{|t|}}. The substring qtq_{-t} denotes the substring indexed by the complement of tt.

Much of our work involves arguing about the properties of a given word. In particular, given a string q𝒜dnq\in\mathcal{A}_{d}^{n}, the relative Hamming weight is defined as w(q)=|{j | qj0}|nw(q)=\frac{|\{j\text{ }|\text{ }q_{j}\neq 0\}|}{n} and the relative character count with respect to i𝒜di\in\mathcal{A}_{d} is defined as ci(q)=|{j | qj=i}|nc_{i}(q)=\frac{|\{j\text{ }|\text{ }q_{j}=i\}|}{n}. Note that w(q)=1c0(x)w(q)=1-c_{0}(x). We will use c(q)c(q) to denote the dd-tuple of all relative counts, namely c(q)=(c0(q),,cd1(q))c(q)=(c_{0}(q),\cdots,c_{d-1}(q)). The Hamming distance between two strings x,y𝒜dnx,y\in\mathcal{A}_{d}^{n} is ΔH(x,y)=|{i | xiyi}|n\Delta_{H}(x,y)=\frac{|\{i\text{ }|\text{ }x_{i}\neq y_{i}\}|}{n}.

A density operator ρ\rho is a positive semi-definite Hermitian operator with trace equal to one, acting on some Hilbert space \mathcal{H}. If ρAE\rho_{AE} acts on some Hilbert space AE\mathcal{H}_{A}\otimes\mathcal{H}_{E}, we write ρE\rho_{E} to mean the partial trace of ρAE\rho_{AE} over AA (similarly for other systems).

We use d\mathcal{H}_{d} to denote a dd-dimensional Hilbert space. Given a basis {|v0,,|vd1}\{\ket{v_{0}},\cdots,\ket{v_{d-1}}\} of d\mathcal{H}_{d}, and given a word i𝒜dni\in\mathcal{A}_{d}^{n}, we write |vi\ket{v_{i}} to mean |vi1|vin\ket{v_{i_{1}}}\otimes\cdots\otimes\ket{v_{i_{n}}}. If the basis under consideration is clear, we will sometimes write |i\ket{i} to mean |vi\ket{v_{i}}.

The Shannon entropy of a random variable XX is denoted by H(X)H(X). The dd-ary entropy function HdH_{d} is defined as Hd(x)=dlogd(d1)xlogdx(1x)logd(1x)H_{d}(x)=d\log_{d}(d-1)-x\log_{d}x-(1-x)\log_{d}(1-x). Note that when d=2d=2 this is simply the binary Shannon entropy. Finally, we define the extended dd-ary entropy H¯d(x)\bar{H}_{d}(x) to be Hd(x)H_{d}(x) if 0x11/d0\leq x\leq 1-1/d; otherwise H¯d(x)=0\bar{H}_{d}(x)=0 if x<0x<0 or H¯d(x)=1\bar{H}_{d}(x)=1 if x>11/dx>1-1/d.

Given ρAE\rho_{AE} acting on AE\mathcal{H}_{A}\otimes\mathcal{H}_{E}, then the conditional quantum min entropy [11] is defined to be:

Hmin(A|E)ρ=supσEmax{λ | 2λIAσEρAE0}.H_{\text{min}}(A|E)_{\rho}=\sup_{\sigma_{E}}\max\{\lambda\in\mathbb{R}\text{ }|\text{ }2^{-\lambda}I_{A}\otimes\sigma_{E}-\rho_{AE}\geq 0\}. (3)

When the EE system is trivial, we have Hmin(A|E)ρ=Hmin(A)ρ=logmaxλH_{\text{min}}(A|E)_{\rho}=H_{\text{min}}(A)_{\rho}=-\log\max\lambda, where the maximum is taken over all eigenvalues λ\lambda of ρA\rho_{A}. In particular, if ρA\rho_{A} is a classical system (that is, ρA=apa|aa|\rho_{A}=\sum_{a}p_{a}\ket{a}\bra{a}), then Hmin(A)ρ=logmaxpaH_{\text{min}}(A)_{\rho}=-\log\max p_{a}. Note that, for any quantum-quantum-classical state ρAEC=c=0NpcρAE(c)|cc|\rho_{AEC}=\sum_{c=0}^{N}p_{c}\rho_{AE}^{(c)}\otimes\ket{c}\bra{c}, then it is easy to prove from the definition of min entropy that the following holds:

Hmin(A|EC)ρmincHmin(A|E)ρ(c).H_{\text{min}}(A|EC)_{\rho}\geq\min_{c}H_{\text{min}}(A|E)_{\rho^{(c)}}. (4)

Though we will not need it here, a useful interpretation of Hmin(A|E)H_{\text{min}}(A|E) for classical-quantum states (cq-states) ρAE\rho_{AE} (that is, states of the form ρAE=apa|aa|ρE(a)\rho_{AE}=\sum_{a}p_{a}\ket{a}\bra{a}\otimes\rho_{E}^{(a)}) was given in [21] as:

Hmin(A|E)ρ=logPg(ρAE),H_{\text{min}}(A|E)_{\rho}=-\log P_{g}(\rho_{AE}),

where Pg(ρAE)P_{g}(\rho_{AE}) is the maximal guessing probability that Eve can guess the value of Alice’s register, namely:

Pg(ρAE)=max{Ma}apatr(MaρE(a)),P_{g}(\rho_{AE})=\max_{\{M_{a}\}}\sum_{a}p_{a}tr\left(M_{a}\rho_{E}^{(a)}\right),

where the maximum is over all POVM operators on E\mathcal{H}_{E}.

Finally, the conditional smooth min entropy is defined to be [11]

Hminϵ(A|E)ρ=supσΓϵ(ρ)Hmin(A|E)σ.H_{\text{min}}^{\epsilon}(A|E)_{\rho}=\sup_{\sigma\in\Gamma_{\epsilon}(\rho)}H_{\text{min}}(A|E)_{\sigma}. (5)

where Γϵ(ρ)={σ | ρσϵ}\Gamma_{\epsilon}(\rho)=\{\sigma\text{ }|\text{ }\left|\left|\rho-\sigma\right|\right|\leq\epsilon\} and here X\left|\left|X\right|\right| is the trace distance of operator XX.

For additional notation, given a quantum state ρAE\rho_{AE} and an orthonormal basis ZZ of the AA register, we write Hmin(Z|E)ρH_{\text{min}}(Z|E)_{\rho} to mean the conditional min entropy of ρAE\rho_{AE} after measuring the AA system using the ZZ basis. If the state ρAE\rho_{AE} is pure, namely ρAE=|ψψ|AE\rho_{AE}=\ket{\psi}\bra{\psi}_{AE}, we write Hmin(A|E)ψH_{\text{min}}(A|E)_{\psi}. This notation is similar for smooth min entropy.

The following Lemma relating the min entropies of mixed and pure states will be useful to our work later as it will allow us to bound the min entropy of a superposition of states by, instead, computing the min entropy of a corresponding mixture of states:

Lemma 1.1.

(From [1] based also on a Lemma in [11]) Let Z={|i}Z=\{\ket{i}\} and X={|xi}X=\{\ket{x_{i}}\} be two orthonormal bases of A\mathcal{H}_{A}. Then for any pure state |ψ=iJαi|i|ϕiEAE\ket{\psi}=\sum_{i\in J}\alpha_{i}\ket{i}\otimes\ket{\phi_{i}}_{E}\in\mathcal{H}_{A}\otimes\mathcal{H}_{E} (where |ϕiE\ket{\phi_{i}}_{E} are arbitrary, normalized states in E\mathcal{H}_{E}), if we define the mixed state ρ=iJ|αi|2|ii||ϕiϕi|\rho=\sum_{i\in J}|\alpha_{i}|^{2}\ket{i}\bra{i}\otimes\ket{\phi_{i}}\bra{\phi_{i}}, then

Hmin(X|E)ψHmin(X|E)ρlog2|J|.H_{\text{min}}(X|E)_{\psi}\geq H_{\text{min}}(X|E)_{\rho}-\log_{2}|J|.

Quantum min entropy is of vital importance to quantum cryptography as it allows one to determine how many uniform random bits one may extract from a cqcq-state ρAE\rho_{AE} that are also independent of Eve. In particular, given a cqcq-state (which, itself, is typically the result of running some quantum cryptographic protocol where the AA register may not be uniform random or completely independent of the EE register), one may apply the process of privacy amplification (typically running the AA register through a randomly chosen two-universal hash function) to establish the required uniform and independent random string. If σKE\sigma_{KE} is the result of applying privacy amplification to the initial ρAE\rho_{AE} system, where the KK register is of size \ell bits, it was shown in [11] that:

σKEIK2σE2(Hminϵ(A|E)ρ)+2ϵ.\left|\left|\sigma_{KE}-\frac{I_{K}}{2^{\ell}}\otimes\sigma_{E}\right|\right|\leq\sqrt{2^{(H_{\text{min}}^{\epsilon}(A|E)_{\rho}-\ell)}}+2\epsilon. (6)

Thus, by deriving a lower-bound on the min entropy of the initial state ρAE\rho_{AE} before privacy amplification, one may establish how many uniform and independent bits may be extracted (namely, \ell) from the state to satisfy the above trace distance inequality up to a desired level of security; e.g., so that the difference between the real state σKE\sigma_{KE} and the “ideal” state IK/2σEI_{K}/2^{\ell}\otimes\sigma_{E} (which represents a uniform random string, independent of any other system) is no more than some ϵPA\epsilon_{PA}.

2 Quantum Sampling

In [1], Bouman and Fehr discovered a fascinating connection between classical sampling strategies and quantum sampling. Since our work utilizes this as a foundation to prove our entropic uncertainty relations (later used to prove security of QRNG and QKD protocols), we take the time in this section to provide a review of their main results. Everything in this section, definitions, concepts, and theorems, come from [1] except when explicitly mentioned. Occasionally, we will make some generalizations and simplifications, however wherever we do so, it will be made clear in the narrative.

Let 𝒜d\mathcal{A}_{d} be an alphabet with dd characters and NN\in\mathbb{N} be fixed. A classical sampling strategy is a triple Ψ=(PT,PS,f)\Psi=(P_{T},P_{S},f), where PTP_{T} is a probability distribution over subsets of {1,2,,N}\{1,2,\cdots,N\}, PSP_{S} is a probability distribution over some set {0,1}\{0,1\}^{*} called seed values, and ff is a function:

f:{0,1}×𝒜dk.f:\{0,1\}^{*}\times\mathcal{A}_{d}^{*}\rightarrow\mathbb{R}^{k}. (7)

Given a string q𝒜Nq\in\mathcal{A}^{N}, the strategy consists of, first, sampling a subset tt according to PTP_{T}; sampling a seed value ss according to PSP_{S}, observing the value of qtq_{t} and evaluating f(s,qt)f(s,q_{t}). This evaluation should lead to a “guess” of the value of some target function g:𝒜dkg:\mathcal{A}_{d}^{*}\rightarrow\mathbb{R}^{k} evaluated on the unobserved portion of qq, namely qtq_{-t}. Informally, a good sampling strategy will ensure that, with high probability, maxi|fi(s,qt)gi(qt)|δ\max_{i}|f_{i}(s,q_{t})-g_{i}(q_{-t})|\leq\delta (i.e., the difference in all coordinates of the output function evaluated on the sampled portion of qq, compared to the target function evaluated on the unobserved portion, are no greater than δ\delta). Note that above, we are generalizing the sampling result of [1] to include more general target and guess functions; in [1], k=1k=1 and g(x)=w(x)g(x)=w(x), the Hamming weight of xx. However, the proof of their main result is easily seen to hold in this more general case, so long as suitable classical strategies are analyzed appropriately (as we do later in this section). Finally, note that in our work, we do not make use of this additional random seed value (which is useful when implementing randomized guess functions ff); thus, we disregard writing it from here on out and, instead, our function ff simply maps strings from 𝒜d|t|\mathcal{A}_{d}^{|t|} to values in k\mathbb{R}^{k}.

Now, fix a subset t{1,2,,N}t\subset\{1,2,\cdots,N\} and δ0\delta\geq 0 and consider the set:

𝒢t,δf,g=𝒢t,δ={i𝒜dN | maxj|fj(it)gj(it)|δ}.\mathcal{G}_{t,\delta}^{f,g}=\mathcal{G}_{t,\delta}=\{i\in\mathcal{A}_{d}^{N}\text{ }|\text{ }\max_{j}|f_{j}(i_{t})-g_{j}(i_{-t})|\leq\delta\}. (8)

This set consists of all “good” words in 𝒜dN\mathcal{A}_{d}^{N} where, for the given choice of tt, the estimate produced by ff is δ\delta close to the desired target function on the unobserved portion. Note that, when the context is clear, we will forgo writing the ff and gg superscripts. From this, the error probability of the given classical sampling strategy is defined to be:

ϵδcl(Ψ)=maxq𝒜dNPr(q𝒢T,δ),\epsilon_{\delta}^{cl}(\Psi)=\max_{q\in\mathcal{A}_{d}^{N}}Pr\left(q\not\in\mathcal{G}_{T,\delta}\right), (9)

where the probability is over the choice of subsets tt drawn according to PTP_{T} (the notation 𝒢t,δ\mathcal{G}_{t,\delta} is used to denote the set defined above for a fixed tt whereas 𝒢T,δ\mathcal{G}_{T,\delta} denotes a random variable over the choice of subset tt). Note that the randomness here is only over the choice of subset; if the function ff need also make random choices, this could be incorporated through the use of the additional seed value. Since our strategies we use here do not need this, we forgo considering it.

From the above definition, it is clear that for any q𝒜dNq\in\mathcal{A}_{d}^{N}, the probability that the sampling strategy fails to produce an accurate estimate of the target function is at most ϵδcl\epsilon_{\delta}^{cl}. The “cl” superscript is used to denote that this is the failure probability of the classical sampling strategy.

These notions may be adapted to quantum states. Let d\mathcal{H}_{d} be the dd-dimensional Hilbert space spanned by some orthonormal basis ={|0,,|d1}\mathcal{B}=\{\ket{0},\cdots,\ket{d-1}\}. The choice of basis may be arbitrary, however all following definitions are taken with respect to the chosen basis.

Given a classical sampling strategy (PT,f)(P_{T},f) (again, disregarding the seed PSP_{S} which we do not use) and a quantum input state |ψdNE\ket{\psi}\in\mathcal{H}_{d}^{\otimes N}\otimes\mathcal{H}_{E}, a quantum sampling strategy may be constructed as follows: first, sample tt according to PTP_{T}; second, measure those qudits in dN\mathcal{H}_{d}^{\otimes N} indexed by tt using basis \mathcal{B} to produce measurement result qt𝒜d|t|q_{t}\in\mathcal{A}_{d}^{|t|}; finally, evaluate the function f(qt)f(q_{t}). The main result from [1], informally, is that the remaining unmeasured portion of the input state should behave like a superposition of states that are δ\delta close in the target function g()g(\cdot) to the estimated value f(qt)f(q_{t}).

More formally, consider:

span(𝒢t,δ)=span{|b | b𝒢t,δ},span\left(\mathcal{G}_{t,\delta}\right)=span\left\{\ket{b}\text{ }|\text{ }b\in\mathcal{G}_{t,\delta}\right\},

where, by |b\ket{b}, we mean |b1|bN\ket{b_{1}}\otimes\cdots\otimes\ket{b_{N}} (again, with respect to the given basis). Note that, if |ψAEspan(𝒢t,δ)E\ket{\psi}_{AE}\in span\left(\mathcal{G}_{t,\delta}\right)\otimes\mathcal{H}_{E}, and if subset tt is actually the one chosen by the sampling strategy, then it is guaranteed that, after measuring those qudits indexed by tt in the given basis \mathcal{B} resulting in outcome qtq_{t}, the remaining unmeasured portion will be in a superposition of states of the form:

|ψqAtE=iJqαi|i|Ei,\ket{\psi_{q}}_{A_{-t}E}=\sum_{i\in J_{q}}\alpha_{i}\ket{i}\otimes\ket{E_{i}},

where:

Jq={i𝒜dN|t| | maxj|fj(q)gj(i)|δ}.J_{q}=\{i\in\mathcal{A}_{d}^{N-|t|}\text{ }|\text{ }\max_{j}|f_{j}(q)-g_{j}(i)|\leq\delta\}.

Formally, the main result from [1] is stated below, which argues that the input state will be ϵ\epsilon close in trace distance to an ideal state where this sampling process always yields the correct guess and this collapse always happens. Furthermore, the ϵ\epsilon depends on the error probability of the underlying classical sampling strategy.

Theorem 2.1.

(From [1], though reworded for our application): Let Ψ=(PT,f)\Psi=(P_{T},f) be a classical sampling strategy with classical failure probability ϵδcl\epsilon_{\delta}^{cl} for given δ>0\delta>0. Then, for every state |ψAEAE\ket{\psi}_{AE}\in\mathcal{H}_{A}\otimes\mathcal{H}_{E} with AdN\mathcal{H}_{A}\cong\mathcal{H}_{d}^{\otimes N}, there exists a collection of states {|ϕAEt}t\{\ket{\phi^{t}_{AE}}\}_{t} indexed by subsets tt of {1,,N}\{1,\cdots,N\} with each |ϕAEtspan(𝒢t,δ)E\ket{\phi_{AE}^{t}}\in span\left(\mathcal{G}_{t,\delta}\right)\otimes\mathcal{H}_{E} such that

12tPT(t)|tt||ψψ|tPT(t)|tt||ϕAEtϕAEt|ϵδcl(Ψ),\frac{1}{2}\bigg{|}\bigg{|}\sum_{t}P_{T}(t)\ket{t}\bra{t}\otimes\ket{\psi}\bra{\psi}-\sum_{t}P_{T}(t)\ket{t}\bra{t}\otimes\ket{\phi^{t}_{AE}}\bra{\phi^{t}_{AE}}\bigg{|}\bigg{|}\leq\sqrt{\epsilon_{\delta}^{cl}(\Psi)}, (10)

where tt represents a sampled subset of {1,,N}\{1,\dots,N\}.

Proof.

In Bouman and Fehr’s work [1], it was shown that for a fixed |ψAE\ket{\psi}_{AE} it holds that

min{|ϕAEt}tPT(t)|tt||ψψ|AEtPT(t)|tt||ϕAEtϕAEt|ϵδcl\min_{\{\ket{\phi^{t}_{AE}}\}}\left|\left|\sum_{t}P_{T}(t)\ket{t}\bra{t}\otimes\ket{\psi}\bra{\psi}_{AE}-\sum_{t}P_{T}(t)\ket{t}\bra{t}\otimes\ket{\phi^{t}_{AE}}\bra{\phi^{t}_{AE}}\right|\right|\leq\sqrt{\epsilon_{\delta}^{cl}} (11)

where the minimum is over all {|ϕAEt}span(𝒢t,δ)E\{\ket{\phi^{t}_{AE}}\}\subset span\left(\mathcal{G}_{t,\delta}\right)\otimes\mathcal{H}_{E}, for a sampling strategy where the target function was g(x)=w(x)g(x)=w(x). However, in their proof, the above is shown directly by projecting the input |ψAE\ket{\psi}_{AE} into the space span(𝒢t,δ)Espan\left(\mathcal{G}_{t,\delta}\right)\otimes\mathcal{H}_{E}, thus directly constructing the ideal states. Namely, the ideal states were defined by the decomposition |ψAE=ϕAEt~|ψAE|ϕAEt~+ϕAEt|ψAE|ϕAEt\ket{\psi}_{AE}=\braket{\widetilde{\phi^{t}_{AE}}}{\psi_{AE}}\ket{\widetilde{\phi^{t}_{AE}}}+\braket{{\phi^{t}_{AE}}}{\psi_{AE}}\ket{\phi^{t}_{AE}} where the |ϕAEt~\ket{\widetilde{\phi^{t}_{AE}}} lives in a space orthogonal to the ideal. This minimum is therefore attained by these ideal states. Furthermore, there is no specific reason in this construction to restrict to target functions that are the Hamming weight, nor to target functions that are one-dimensional. Indeed, by considering any definition of 𝒢t,δ\mathcal{G}_{t,\delta}, their construction and the subsequent analysis follows identically assuming the error probability is defined as in Equation 9 based on the set 𝒢t,δ\mathcal{G}_{t,\delta}. The important difference comes in the analysis of the classical sampling strategy in order to compute ϵδcl\epsilon_{\delta}^{cl}. ∎

The fascinating thing about Theorem 2.1 is that, by choosing suitable classical sampling strategies, one may analyze the behavior of ideal states which always behave appropriately for the given strategy. From this, and the fact that the real state is close, in trace distance, to these ideal states (on average over the randomness in the sampling strategy), one may then promote the analysis from the ideal state to the actual input. Already in [2, 3], we used this to prove novel, and useful, quantum entropic uncertainty relations which were then used to analyze particular QRNG protocols. We now generalize these results, analyze a more powerful QRNG protocol, and also show how this can be used to develop three-party entropic uncertainty relations (involving AA, BB, and EE) with applications to high-dimensional QKD protocols. We show that, furthermore, this provides highly optimistic secure bit generation rates for both the QRNG and QKD protocols in a variety of scenarios. However, to analyze these protocols, we first require some important classical sampling strategies.

2.1 Classical Sampling Strategies

As discussed, Theorem 2.1 allows us to consider classical sampling strategies and use these to analyze quantum protocols. Here we discuss four classical sampling strategies which we denote Ψ0,Ψ1,\Psi_{0},\Psi_{1}, Ψ2,\Psi_{2}, and Ψ2+0\Psi_{2+0}. Strategy Ψ0\Psi_{0} was analyzed in [1] and we use this to bound the error of the other strategies. The other strategies involve one party (Ψ1\Psi_{1}) or two parties (Ψ2\Psi_{2} and Ψ2+0\Psi_{2+0}) and will be used later when deriving our entropic uncertainty relations.


One-Party HD-Restricted-Sampling Ψ0\Psi_{0}: In [1], the following natural sampling strategy was analyzed which we denote here as Ψ0\Psi_{0}. We use this result to bound the error in our other sampling strategies to be discussed next. Let q𝒜dn+mq\in\mathcal{A}_{d}^{n+m} be a string and the target function g(x)=w(x)g(x)=w(x). The strategy, first, chooses a subset tt of {1,,n+m}\{1,\cdots,n+m\} of size mm, uniformly at random and observes string qtq_{t}. Next, it outputs f(qt)=w(qt)f(q_{t})=w(q_{t}), an estimate of the Hamming weight of the unobserved portion, namely w(qt)w(q_{-t}). We call this the HD-Restricted-Sampling strategy as it is high-dimensional, however it only looks at the Hamming weight, ignoring the counts of other characters. The following Lemma was proven in [1]:

Lemma 2.1.

(From [1]): Let δ>0\delta>0 and d2d\geq 2. Then the failure probability of the above described sampling strategy Ψ0\Psi_{0} for mnm\leq n is:

ϵδcl(Ψ0)2exp(δ2m(n+m)m+n+2).\epsilon_{\delta}^{cl}(\Psi_{0})\leq 2\exp\left(\frac{-\delta^{2}m(n+m)}{m+n+2}\right).

We comment that there is nothing special in the above sampling strategy, or their proof, about the use of the Hamming weight in the above Lemma; instead one could replace the target function g(x)g(x) with any single cj(x)c_{j}(x) or 1cj(x)1-c_{j}(x) (to count the number of letters equal to, or not equal to, jj respectively) and the same bound will follow (for a single, fixed but arbitrary, jj). See [1].


One-Party HD-Full-Sampling Ψ1\Psi_{1}: In our work, here, we will need three additional sampling strategies. The first sampling strategy, which we denote Ψ1\Psi_{1}, is a one-party strategy involving Alice only and will be used for our QRNG analysis later. The strategy works for strings in 𝒜dN\mathcal{A}_{d}^{N}, where N=n+mN=n+m and the target function is g(x)=(c0(x),,cd1(x))g(x)=(c_{0}(x),\dots,c_{d-1}(x)) where ci(x)c_{i}(x) is the relative number of times symbol ii appears in the word xx (as defined in Section 1.1). First, the strategy Ψ1\Psi_{1} chooses a subset tt of size mm from {1,,N}\{1,\cdots,N\} uniformly at random and observes the string qt𝒜dmq_{t}\in\mathcal{A}_{d}^{m}. Finally, Ψ1\Psi_{1} outputs f(qt)=(c0(qt),,cd1(qt))f(q_{t})=(c_{0}(q_{t}),\dots,c_{d-1}(q_{t})) as an estimate of the relative counts of the unobserved qtq_{-t}. The proceeding Lemma determines an upper bound on the error probability of the sampling strategy Ψ1\Psi_{1}.

Lemma 2.2.

Let δ>0\delta>0 and d2d\geq 2. Then the failure probability of the above described sampling strategy Ψ1\Psi_{1} when mnm\leq n is:

ϵδcl(Ψ1)2dexp(mδ2m+nm+n+2).\epsilon_{\delta}^{cl}(\Psi_{1})\leq 2d\exp\left(-m\delta^{2}\frac{m+n}{m+n+2}\right).
Proof.

Note that, for any jj, (PT,cj)(P_{T},c_{j}) is exactly the strategy Ψ0\Psi_{0} (though, instead of looking at the number of strings with a certain Hamming weight, we are looking at the number of strings with a certain character count). Thus, using the bound provided by Lemma 2.1 we find

ϵδcl\displaystyle\epsilon_{\delta}^{cl} =maxq𝒜dm+nPr(q𝒢T,δ(Ψ1))\displaystyle=\max_{q\in\mathcal{A}_{d}^{m+n}}Pr\left(q\not\in\mathcal{G}_{T,\delta}(\Psi_{1})\right)
jmaxq𝒜dm+nPr(|fj(qt)gj(qt)|>δ)\displaystyle\leq\sum_{j}\max_{q\in\mathcal{A}_{d}^{m+n}}Pr\left(|f_{j}(q_{t})-g_{j}(q_{-t})|>\delta\right)
2dexp(mδ2m+nm+n+2).\displaystyle\leq 2d\exp\left(-m\delta^{2}\frac{m+n}{m+n+2}\right).


Two-Party HD-Sampling Ψ2\Psi_{2}: The second strategy we require will be used for our two-party applications later and we denote by Ψ2\Psi_{2}. Here, we have an input string q=(qA,qB)𝒜dN×𝒜dNq=(q^{A},q^{B})\in\mathcal{A}_{d}^{N}\times\mathcal{A}_{d}^{N}, where N=n+mN=n+m. The strategy will first choose a subset t{1,,N}t\subset\{1,\cdots,N\} of size mm uniformly at random. The strategy will then sample qtAq^{A}_{t} and qtBq^{B}_{t}; that is, it will observe the qAq^{A} portion and qBq^{B} portion individually, using the same subset (this may be written strictly using our earlier definitions, however such strict formality is not enlightening). The target function is g(qtA,qtB)=ΔH(qtA,qtB)g(q^{A}_{-t},q^{B}_{-t})=\Delta_{H}(q^{A}_{-t},q^{B}_{-t}) (where ΔH(x,y)\Delta_{H}(x,y) is the relative Hamming distance of words xx and yy as defined in Section 1.1) and the output will be f(qtA,qtB)=ΔH(qtA,qtB)f(q^{A}_{t},q^{B}_{t})=\Delta_{H}(q^{A}_{t},q^{B}_{t}). Again, we may bound the error probability of this strategy using Lemma 2.1.

Lemma 2.3.

Let Ψ2\Psi_{2} be the strategy defined above; δ>0\delta>0 and mnm\leq n. Then ϵδcl(Ψ2)ϵδcl(Ψ0)\epsilon^{cl}_{\delta}(\Psi_{2})\leq\epsilon^{cl}_{\delta}(\Psi_{0}).

Proof.

Let N=n+mN=n+m and 𝒢t,δ={(i,j)𝒜dN×𝒜dN | |ΔH(it,jt)ΔH(it,jt)|δ}\mathcal{G}_{t,\delta}=\{(i,j)\in\mathcal{A}_{d}^{N}\times\mathcal{A}_{d}^{N}\text{ }|\text{ }|\Delta_{H}(i_{t},j_{t})-\Delta_{H}(i_{-t},j_{-t})|\leq\delta\} and 𝒢t,δ={i𝒜N | |w(it)w(it)|δ}.\mathcal{G}_{t,\delta}^{\prime}=\{i\in\mathcal{A}^{N}\text{ }|\text{ }|w(i_{t})-w(i_{-t})|\leq\delta\}. Pick q=(qA,qB)𝒜dN×𝒜dNq=(q^{A},q^{B})\in\mathcal{A}_{d}^{N}\times\mathcal{A}_{d}^{N} and let x=qAqBx=q^{A}-q^{B}, where the subtraction here is character-wise, modulo dd, in the given alphabet. Clearly w(xt)=ΔH(qtA,qtB)w(x_{t})=\Delta_{H}(q^{A}_{t},q^{B}_{t}), and similarly for xtx_{-t}. Thus, q𝒢t,δq\in\mathcal{G}_{t,\delta} if and only if x𝒢t,δx\in\mathcal{G}_{t,\delta}^{\prime}. Hence, for every q=(qA,qB)q=(q^{A},q^{B}), it holds that:

Pr(qAqB𝒢T,δ)=Pr(qAqB𝒢T,δ)maxx𝒜dNPr(x𝒢T,δ)=ϵδcl(Ψ0).Pr\left(q^{A}q^{B}\not\in\mathcal{G}_{T,\delta}\right)=Pr\left(q^{A}-q^{B}\not\in\mathcal{G}^{\prime}_{T,\delta}\right)\leq\max_{x\in\mathcal{A}_{d}^{N}}Pr\left(x\not\in\mathcal{G}^{\prime}_{T,\delta}\right)=\epsilon^{cl}_{\delta}(\Psi_{0}).

Since this holds for any q=(qA,qB)q=(q^{A},q^{B}), we’re done. ∎

Finally, we define a second two-party sampling strategy which combines Ψ2\Psi_{2} with Ψ0\Psi_{0}; we denote this strategy by Ψ2+0\Psi_{2+0}. For this strategy, the target function is now g(qtA,qtB)=(ΔH(qtA,qtB),cb(qtA))g(q_{-t}^{A},q_{-t}^{B})=(\Delta_{H}(q_{-t}^{A},q_{-t}^{B}),c_{b^{*}}(q_{-t}^{A})) for some given, fixed, distinguished index b𝒜db^{*}\in\mathcal{A}_{d} (we later call this the “count index”). This sampling strategy chooses a subset according to Ψ2\Psi_{2} and outputs a guess f(qtA,qtB)=(ΔH(qtA,qtB),cb(qtA)).f(q^{A}_{t},q^{B}_{t})=(\Delta_{H}(q^{A}_{t},q^{B}_{t}),c_{b^{*}}(q^{A}_{t})). It is not difficult to show from Lemmas 2.1 and 2.3 that the error probability of this strategy is:

ϵδcl(Ψ2+0)ϵδcl(Ψ2)+ϵδcl(Ψ0)4exp(δ2m(n+m)m+n+2).\epsilon_{\delta}^{cl}(\Psi_{2+0})\leq\epsilon_{\delta}^{cl}(\Psi_{2})+\epsilon_{\delta}^{cl}(\Psi_{0})\leq 4\exp\left(\frac{-\delta^{2}m(n+m)}{m+n+2}\right). (12)

3 Quantum Sampling Based Entropic Uncertainty

In [2, 3], we showed how the technique of quantum sampling, introduced in [1] and discussed in the previous section, can be used to prove entropic uncertainty relations bounding the smooth quantum min entropy and the Shannon entropy, as a function of the overlap of two projective measurements. Our first work [2] introduced a novel entropic uncertainty relation applicable to qubits (i.e., d=2d=2) only and with a fixed sampling strategy; in [3], we expanded the result to work for qudits (d2d\geq 2), however only with a partial basis measurement and a particular, fixed, sampling strategy. Here, we discuss and generalize this result to work with more general sampling strategies allowing a “plug-and-play” entropic uncertainty relation for various classical sampling strategies. Indeed, as shown in this section, one may introduce an arbitrary classical sampling strategy (perhaps one that is useful for a particular cryptographic application); one need only compute the error probability of the given classical strategy, along with the size of a set similar to 𝒢\mathcal{G} (generally a classical combinatorial proof) to derive a result applicable to a quantum system. The proof of this follows the same two-step approach we introduced in [2, 3] only with suitable generalizations at certain points.

To describe our sampling based entropic uncertainty relations, we require an experiment which takes as input a quantum state ρ\rho acting on TAE\mathcal{H}_{T}\otimes\mathcal{H}_{A}\otimes\mathcal{H}_{E} where the AA portion is an NN-fold tensor of some smaller dd-dimensional Hilbert space and the TT register is a Hilbert space spanned by orthonormal basis {|t}\{\ket{t}\} where t{1,,N}t\subset\{1,\cdots,N\}. The experiment also requires an orthonormal basis X={|x0,|xd1}X=\{\ket{x_{0}},\cdots\ket{x_{d-1}}\}.

The experiment will first choose a random subset tt by measuring the TT register. It will then measure the AA portion of ρ\rho, indexed by tt, using the given XX basis. This measurement results in outcome q𝒜d|t|q\in\mathcal{A}_{d}^{|t|} and a post-measurement state ρ(q,t)\rho(q,t), acting on the unmeasured portion of A\mathcal{H}_{A} and E\mathcal{H}_{E}. We denote this experiment by (t,q,ρAE(q,t))Exp(ρTAE,X)(t,q,\rho_{A^{\prime}E}(q,t))\leftarrow\textbf{{Exp}}\left(\rho_{TAE},X\right). Note that the experiment also returns the subset chosen. Sampling based entropic uncertainty relations allow one to bound the min entropy in the remaining post-measured state, assuming an alternative measurement were to be made on the AA portion of it. This bound is a function of the measurement overlap and the classical measurement outcome qq.

The main result from [2, 3] was to relate the min entropy in the remaining portion of the system as a function of the measurement overlap and the binary Shannon entropy (or, in the case of [3], the dd-ary Shannon entropy) of the relative Hamming weight of the observed outcome qq after running the experiment. However, the proof technique used there can be applied to a more general setting allowing for arbitrary sampling strategies and, in particular, to bound the min-entropy as a function of the measurement overlap and the size of a particular set JqJ_{q} of classical strings that are δ\delta-close to the observed qq.

Theorem 3.1.

Let 0<β<1/20<\beta<1/2 and Ψ\Psi be a classical sampling strategy with error probability ϵδcl\epsilon_{\delta}^{cl} for given δ>0\delta>0. Let ϵ=ϵδcl\epsilon=\sqrt{\epsilon^{cl}_{\delta}}, and let ρAE\rho_{AE} be an arbitrary quantum state acting on space AE\mathcal{H}_{A}\otimes\mathcal{H}_{E}, where AdN\mathcal{H}_{A}\cong\mathcal{H}_{d}^{\otimes N} for d2d\geq 2. Let Z={|zi}i=0d1Z=\{\ket{z_{i}}\}_{i=0}^{d-1} and X={|xi}i=0d1X=\{\ket{x_{i}}\}_{i=0}^{d-1} be two orthonormal bases of d\mathcal{H}_{d}. Furthermore, let (t,q,ρ(t,q))Exp(tPT(t)|tt|ρAE,X)(t,q,\rho(t,q))\leftarrow\textbf{Exp}(\sum_{t}P_{T}(t)\ket{t}\bra{t}\otimes\rho_{AE},X), where the sum is over all possible subsets of {1,2,,N}\{1,2,\dots,N\} that could be chosen by Ψ\Psi and PT(t)P_{T}(t) is the probability of subset tt being chosen as determined by the given classical sampling strategy. Finally, let γ=log2maxa,b|za|xb|2\gamma=-\log_{2}\max_{a,b}|\braket{z_{a}}{x_{b}}|^{2}. Then, it holds that:

Pr(Hmin4ϵ+2ϵβ(Z|E)ρ(t,q)+log2|Jq(N|t|)|(N|t|)γ)12ϵ12β,Pr\left(H_{\text{min}}^{4\epsilon+2\epsilon^{\beta}}(Z|E)_{\rho(t,q)}+\log_{2}|J_{q}^{(N-|t|)}|\geq(N-|t|)\gamma\right)\geq 1-2\epsilon^{1-2\beta}, (13)

where

Jq(n)={i𝒜dn | maxj|fj(i)gj(q)|δ}.J_{q}^{(n)}=\{i\in\mathcal{A}_{d}^{n}\text{ }|\text{ }\max_{j}|f_{j}(i)-g_{j}(q)|\leq\delta\}. (14)

Above the probability is over the randomness in the experiment (namely the subset chosen and the resulting measurement outcome qq).

Proof.

The proof follows the same two-step argument we developed in [2, 3]. In fact, most of the proof is identical with the exception of a few generalizations; we provide the proof here at a high-level only for completeness, referring the reader to [2, 3] for complete technical details when needed.


First Step - Ideal Analysis: We begin by considering the case when the input state ρAE\rho_{AE} is pure; the mixed case then follows through standard purification techniques.

By applying Theorem 2.1 with respect to the given XX basis and sampling strategy Ψ\Psi, there exist ideal states {|ϕAEt}\{\ket{\phi_{AE}^{t}}\} such that for every tt, the state |ϕAEtspan{|xi | i𝒜dN and maxj|fj(it)gj(it)|δ}E\ket{\phi^{t}_{AE}}\in\text{span}\{\ket{x_{i}}\text{ }|\text{ }i\in\mathcal{A}_{d}^{N}\text{ and }\max_{j}|f_{j}(i_{t})-g_{j}(i_{t})|\leq\delta\}\otimes\mathcal{H}_{E}. Note that the target function g(x)=(g1(x),,gk(x))g(x)=(g_{1}(x),\cdots,g_{k}(x)) also depends on the sampling strategy. Furthermore, from this application of Theorem 2.1, if we define σTAE=tPT(t)|tt||ϕAEtϕAEt|,\sigma_{TAE}=\sum_{t}P_{T}(t)\ket{t}\bra{t}\otimes\ket{\phi_{AE}^{t}}\bra{\phi_{AE}^{t}}, then it holds that:

tPT(t)|tt|ρAEσTAEϵδcl(Ψ)=ϵ.\left|\left|\sum_{t}P_{T}(t)\ket{t}\bra{t}\otimes\rho_{AE}-\sigma_{TAE}\right|\right|\leq\sqrt{\epsilon^{cl}_{\delta}(\Psi)}=\epsilon. (15)

Consider the output of running (t,q,σ(t,q))Exp(σ,X)(t,q,\sigma(t,q))\leftarrow\textbf{{Exp}}\left(\sigma,X\right). Here q𝒜d|t|q\in\mathcal{A}_{d}^{|t|}. It is not difficult to see that the resulting state, after tracing out the measured portion, is of the form:

σ(t,q)=iJq(N|t|)αi|xi|Ei,\sigma(t,q)=\sum_{i\in J_{q}^{(N-|t|)}}\alpha_{i}\ket{x_{i}}\otimes\ket{E_{i}}, (16)

where Jq(n)={i𝒜dn | maxj|fj(i)gj(q)|δ}J_{q}^{(n)}=\{i\in\mathcal{A}_{d}^{n}\text{ }|\text{ }\max_{j}|f_{j}(i)-g_{j}(q)|\leq\delta\} (note that some of the αi\alpha_{i}’s may be zero).

Let n=N|t|n=N-|t|. From Lemma 1.1, we have Hmin(Z|E)σ(t,q)H(Z|E)χlog|Jq(n)|H_{\text{min}}(Z|E)_{\sigma(t,q)}\geq H(Z|E)_{\chi}-\log|J_{q}^{(n)}|, where χ\chi is the mixed state:

χAE=iJq(n)|αi|2|xixi||EiEi|.\chi_{AE}=\sum_{i\in J_{q}^{(n)}}|\alpha_{i}|^{2}\ket{x_{i}}\bra{x_{i}}\otimes\ket{E_{i}}\bra{E_{i}}.

It is straight-forward to show that Hmin(Z|E)χ=nγ=(N|t|)γH_{\text{min}}(Z|E)_{\chi}=n\gamma=(N-|t|)\gamma. This is done by conditioning on an additional classical system, writing out the probability distribution of the ZZ basis measurement given χ\chi and taking advantage of Equation 4 (see [3] for explicit details on how this computation is done given a mixed state of this form). Thus, with certainty, the ideal case, after choosing subset tt and observing qq, will have min entropy no less than (N|t|)γlog|Jq(n)|(N-|t|)\gamma-\log|J_{q}^{(n)}|.


Second Step - Real Case Analysis: The second step involves arguing that the real state cannot behave too differently from the ideal state we just analyzed. We make use of Chebyshev’s inequality while also switching to smooth min entropy to complete the analysis.

Consider the real state ρ=1Tt|tt|ρAE\rho=\frac{1}{T}\sum_{t}\ket{t}\bra{t}\otimes\rho_{AE} where ρAE\rho_{AE} is given as input to the theorem (note that, here, the input state is independent of the subset chosen unlike in the ideal case). The process of choosing a subset tt, measuring, and observing qq (resulting in post-measurement state ρ(t,q)\rho(t,q)) may be described, entirely, by the mixed state:

ρTQR=tPT(t)|tt|q𝒜d|t|p(q|t)|qq|ρ(t,q),\rho_{TQR}=\sum_{t}P_{T}(t)\ket{t}\bra{t}\otimes\sum_{q\in\mathcal{A}_{d}^{|t|}}p(q|t)\ket{q}\bra{q}\otimes\rho(t,q),

where p(q|t)p(q|t) is the probability of observing outcome qq given that the subset tt was sampled; here we use the “R” register to denote the remaining, unmeasured, portion of the state. Likewise, the ideal state, after performing this experiment, may be written as the mixed state: σTQR=tPT(t)|tt|qp~(q|t)|qq|σ(t,q)\sigma_{TQR}=\sum_{t}P_{T}(t)\ket{t}\bra{t}\otimes\sum_{q}\tilde{p}(q|t)\ket{q}\bra{q}\otimes\sigma(t,q). We define Δq,t=12ρ(t,q)σ(t,q)\Delta_{q,t}=\frac{1}{2}||\rho(t,q)-\sigma(t,q)||, which may be treated as a random variable over the choice of tt and observed qq. We want to show that, with high probability, Δq,t\Delta_{q,t} is “small.”

It is not difficult to show that the expected value of Δq,t\Delta_{q,t} is 𝔼(Δq,t)=μ2ϵ{\mathbb{E}}(\Delta_{q,t})=\mu\leq 2\epsilon. Furthermore, the variance V2V^{2} of this random variable has the property that V2μ2ϵV^{2}\leq\mu\leq 2\epsilon (see our proof in [2] for both these computations, though they follow immediately from properties of trace distance and the fact that Δt,q1\Delta_{t,q}\leq 1).

Now, by Chebyshev’s inequality, we have:

Pr(|Δq,tμ|ϵβ)V2ϵ2β2ϵ12β,\Pr(|\Delta_{q,t}-\mu|\geq\epsilon^{\beta})\leq\frac{V^{2}}{\epsilon^{2\beta}}\leq 2\epsilon^{1-2\beta}, (17)

(the last inequality follows since β<12\beta<\frac{1}{2}); note that this probability is over all subsets tt and measurement outcomes qq. Thus, except with probability at most 2ϵ12β2\epsilon^{1-2\beta}, after choosing tt and observing qq, it holds that |Δq,tμ|ϵβ|\Delta_{q,t}-\mu|\leq\epsilon^{\beta} which implies:

12ρ(t,q)σ(t,q)=Δq,tμ+ϵβ2ϵ+ϵβ.\frac{1}{2}||\rho(t,q)-\sigma(t,q)||=\Delta_{q,t}\leq\mu+\epsilon^{\beta}\leq 2\epsilon+\epsilon^{\beta}.

Thus, we may conclude that Hmin4ϵ+2ϵβ(AZ|E)ρHmin(AZ|E)σH_{\text{min}}^{4\epsilon+2\epsilon^{\beta}}(A_{Z}|E)_{\rho}\geq H_{\text{min}}(A_{Z}|E)_{\sigma}, completing the second step of the proof.

Of course, the above analysis assumed the input state ρAE\rho_{AE} was pure. However, if the state is not pure, it may be purified and, incorporating this extra system to EE, the result above follows. ∎

Notice that one may choose sampling strategies suitable to a particular application and, then, need only to analyze the classical strategy to attain a result in the quantum setting. Furthermore, arbitrary sampling strategies may be employed with arbitrary target functions, leading to a potential wide-range of applications. One simply needs to analyze the failure probabilities of the resulting classical sampling strategy (Equation 9). We demonstrate this by analyzing a QRNG protocol in the next section.

3.1 Application to Quantum Random Number Generators

Quantum Random Number Generators (QRNG) are protocols which, by utilizing a physical source of randomness in particular quantum sources, attempt to distill a uniform random string. For a cryptographic QRNG, the string should be uniform random and also independent of any adversary. At the most basic level, a QRNG protocol could consist of a source emitting a photon passing through a beam splitter connected to two photon counters. Such a system will lead to a random measurement on one detector or the other, producing a random stream of 0’s and 11’s. Such a setup assumes fully trusted devices (both the source and measurement apparatus are fully trusted and characterized and outside the control or influence of any adversary).

On the opposite extreme is the fully device independent model [22, 23] whereby the source and measurement apparatus are not trusted (perhaps manufactured by the adversary - though one must still assume, of course, that the actual measurement outcome reported by the untrusted device cannot be sent to the adversary). Fully device independent protocols are obviously highly desirable from a cryptographic standpoint; however in practice, they are slow to implement [24, 25]. This leads to a middle-ground between these two extremes known as the source-independent (SI) model introduced originally in [10] and studied further in several works including [26, 27]. Here, the quantum source is not trusted, however the measurement devices used are trusted and characterized. Such protocols are a step up from the fully trusted scenario (as they can take into account physical imperfections, but also the fact that an adversary may be entangled with the source and, thus, attempt to gain information on the resulting random string). Furthermore, they are highly practical, leading to Gbps implementations [28]. Finally, by not trusting the source, several fascinating possibilities are open, including the use of sunlight as the source [29]. For a general survey of QRNG protocols and their security models, the reader is referred to [30].

In previous work, we showed that sampling-based entropic uncertainty relations provide optimistic results for QRNG protocols. In [2], we analyzed a qubit-based protocol but without an adversary. In [3], we analyzed a SI-QRNG protocol with an adversarial source and qudits (dd-level systems), however where Alice was restricted to performing only a partial basis measurement (our previous relation could not take into account a full basis measurement for the sampling stage of the protocol). Here, we show how our entropic uncertainty relation can be used to provide highly optimistic bit generation rates for the full high-dimensional SI-QRNG protocol introduced in [10] (where a full basis measurement is required for the test stage). The protocol we analyze requires Alice to be able to measure in two bases Z={|0,,|d1}Z=\{\ket{0},\cdots,\ket{d-1}\} and X={|x0,,|xd1}X=\{\ket{x_{0}},\cdots,\ket{x_{d-1}}\}. We assume the measurement devices are fully characterized and so maxi,j|i|xj|\max_{i,j}|\braket{i}{x_{j}}| is known. In the following we will assume that |i|xj|=1/d|\braket{i}{x_{j}}|=1/\sqrt{d} for all i,ji,j however our analysis works identically for other scenarios. The protocol, then, operates as follows:

  1. 1.

    Preparation: An adversary prepares a quantum state |ψ0AE\ket{\psi_{0}}\in\mathcal{H}_{A}\otimes\mathcal{H}_{E}, where the A\mathcal{H}_{A} portion is an (n+m)(n+m)-fold tensor of d\mathcal{H}_{d} (i.e., the AA register consists of n+mn+m qudits of dimension dd for a known d2d\geq 2). The AA portion is sent to Alice while the EE portion remains with the adversary. An ideal source should prepare the state |ψ0=|x0(n+m)|χE\ket{\psi_{0}}=\ket{x_{0}}^{\otimes(n+m)}\otimes\ket{\chi}_{E} - that is, a state independent of Eve and with n+mn+m perfect copies of the qudit state |x0\ket{x_{0}}. As the source is adversarial, we do not assume anything about the structure of |ψ0\ket{\psi_{0}} other than it lives in AE\mathcal{H}_{A}\otimes\mathcal{H}_{E}.

  2. 2.

    Sampling and Measurements: Alice chooses a random subset tt of size mm and measures those qudits indexed by tt in the XX basis, recording the outcome as q𝒜dmq\in\mathcal{A}_{d}^{m}. The character counts of this will be used to determine how much information an adversary has (it should be that c0(q)c_{0}(q) is high). The remaining qudits she measures in the ZZ basis, saving the resulting string as r𝒜dnr\in\mathcal{A}_{d}^{n}. Note we are not considering experimental imperfections on the devices such as dark counts or low-efficiency detectors - we are only interested in the theoretical bound of ideal measurements, leaving these interesting practical measurement concerns as potential future work.

  3. 3.

    Post-Processing: Alice runs a privacy amplification protocol, applying a two-universal hash function ff to the string rr, resulting in her final random string s=f(r)s=f(r). As proven in [31], for a QRNG protocol of this nature, the hash function ff need only be chosen randomly once and then reused, so no additional randomness is needed here.

The sampling portion of this protocol is easily seen to be Ψ1\Psi_{1} introduced in Section 2.1 with target function g(x)=(c0(x),,cd1(x))g(x)=(c_{0}(x),\cdots,c_{d-1}(x)). In this case, the size of the chosen subset tt is always mm leaving nn qudits unmeasured. So we write JqJ_{q} in place of Jq(n)J_{q}^{(n)} from Theorem 3.1 and its definition is:

Jq={i𝒜dn | maxj|cj(i)cj(q)|δ}.J_{q}=\{i\in\mathcal{A}_{d}^{n}\text{ }|\text{ }\max_{j}|c_{j}(i)-c_{j}(q)|\leq\delta\}. (18)

To apply the sampling based entropic uncertainty relation of Theorem 3.1, we first bound the size of this set. Of course JqIq={i𝒜dn | |w(i)w(q)|δ}J_{q}\subset I_{q}=\{i\in\mathcal{A}_{d}^{n}\text{ }|\text{ }|w(i)-w(q)|\leq\delta\} where w(x)w(x) is the relative Hamming weight of xx. Then, using the well-known volume of a Hamming ball, we may bound |Jq||Iq|dnH¯(w(q)+δ)|J_{q}|\leq|I_{q}|\leq d^{n\bar{H}(w(q)+\delta)}. This is the bound we used in our entropic uncertainty relation in [3] (which was based on the set IqI_{q} not the full JqJ_{q} since full measurements were not supported in our earlier work). However, when we have full information on the string qq, we may attempt to derive a tighter bound on JqJ_{q} itself for use in analyzing this QRNG protocol. Theorem 3.2 provides an alternative bound on |Jq||J_{q}| which is tighter in some scenarios as we discuss later.

Theorem 3.2.

Let 1d>δ>0\frac{1}{d}>\delta>0 and q𝒜dmq\in\mathcal{A}^{m}_{d} be given. Define the functions νi\nu_{i} for each i𝒜di\in\mathcal{A}_{d}, dependent on the choice of qq, to be

νi={0,ci(q)δ0ci(q)δ,otherwise.\nu_{i}=\begin{cases}0,&c_{i}(q)-\delta\leq 0\\ c_{i}(q)-\delta,&\text{otherwise}.\end{cases}

then, for Jq=Jq(n)J_{q}=J_{q}^{(n)} defined in Equation 18, we have:

log2|Jq|ni𝒜dνilog2νi+nlog2n(1i𝒜dνi)+(d+1)log2ed2log2(1dδd).\log_{2}|J_{q}|\leq-n\sum_{i\in\mathcal{A}_{d}}\nu_{i}\log_{2}\nu_{i}+n\log_{2}n\left(1-\sum_{i\in\mathcal{A}_{d}}\nu_{i}\right)+(d+1)\log_{2}e-\frac{d}{2}\log_{2}\left(\frac{1-d\delta}{d}\right). (19)
Proof.

To prove this, we count the total number of ways one may construct a string with the required counts. Let 𝒦q={(x0,,xd1)d:|xinci(q)|nδ and xi=n}\mathcal{K}_{q}=\left\{(x_{0},\dots,x_{d-1})\in\mathbb{N}^{d}:|x_{i}-nc_{i}(q)|\leq n\delta\text{ and }\sum x_{i}=n\right\} and observe that

|Jq|\displaystyle|J_{q}| =k𝒦qkik(nj=0i1kjki)\displaystyle=\sum_{k\in\mathcal{K}_{q}}\prod_{k_{i}\in k}{n-\sum_{j=0}^{i-1}k_{j}\choose k_{i}}
=k𝒦qn!k0!(nk0)!(nk0)!k1!(nk0k1)!(nk0k1)!k2!(nk0k1k2)!\displaystyle=\sum_{k\in\mathcal{K}_{q}}\frac{n!}{k_{0}!(n-k_{0})!}\cdot\frac{(n-k_{0})!}{k_{1}!(n-k_{0}-k_{1})!}\cdot\frac{(n-k_{0}-k_{1})!}{k_{2}!(n-k_{0}-k_{1}-k_{2})!}\dots
=k𝒦qn!k0!k1!k2!=n!k𝒦qkik1ki!.\displaystyle=\sum_{k\in\mathcal{K}_{q}}\frac{n!}{k_{0}!k_{1}!k_{2}!\dots}\;\;=\;\;n!\sum_{k\in\mathcal{K}_{q}}\prod_{k_{i}\in k}\frac{1}{k_{i}!}.

Let q={(x0,,xd1)d:|xinci(q)|nδ}\mathcal{M}_{q}=\{(x_{0},\dots,x_{d-1})\in\mathbb{N}^{d}:|x_{i}-nc_{i}(q)|\leq n\delta\}. Of course 𝒦qq\mathcal{K}_{q}\subset\mathcal{M}_{q}. This immediately implies

n!k𝒦qkik1ki!n!xqxix1xi!.n!\sum_{k\in\mathcal{K}_{q}}\prod_{k_{i}\in k}\frac{1}{k_{i}!}\leq n!\sum_{x\in\mathcal{M}_{q}}\prod_{x_{i}\in x}\frac{1}{x_{i}!}.

Now let {xi1,xi2,,ximi}\{x_{i}^{1},x_{i}^{2},\dots,x_{i}^{m_{i}}\}\subset\mathbb{N} be the values in increasing order which satisfy |xijnci(q)|nδ|x_{i}^{j}-nc_{i}(q)|\leq n\delta for all j{1,,mi}j\in\{1,\dots,m_{i}\}. We can enumerate the set q\mathcal{M}_{q} as

q={(x0j0,x1j1,,xd1jd1) | ji{1,,mi}i{0,,d1}}.\mathcal{M}_{q}=\{(x_{0}^{j_{0}},x_{1}^{j_{1}},\dots,x_{d-1}^{j_{d-1}})\text{ }|\text{ }j_{i}\in\{1,\dots,m_{i}\}\;\forall i\in\{0,\dots,d-1\}\}.

Then

xqi=0d11xi!\displaystyle\sum_{x\in\mathcal{M}_{q}}\prod_{i=0}^{d-1}\frac{1}{x_{i}!} =j0,,jd1(1x0j0!1x1j1!1xd1jd1!)\displaystyle=\sum_{j_{0},\dots,j_{d-1}}\left(\frac{1}{x_{0}^{j_{0}}!}\cdot\frac{1}{x_{1}^{j_{1}}!}\cdot\ldots\cdot\frac{1}{x_{d-1}^{j_{d-1}}!}\right)
=i=0d1(1xi1!+1xi2!++1ximi!)=i=0d1ji=1mi1xiji!\displaystyle=\prod_{i=0}^{d-1}\left(\frac{1}{x_{i}^{1}!}+\frac{1}{x_{i}^{2}!}+\ldots+\frac{1}{x_{i}^{m_{i}}!}\right)\;=\;\prod_{i=0}^{d-1}\sum_{j_{i}=1}^{m_{i}}\frac{1}{x_{i}^{j_{i}}!}

The benefit of isolating these partial sums of 1/xiji!1/x_{i}^{j_{i}}! is that we can take advantage of the Taylor series for exe^{x} to bound this partial sum. We can expand on this to get the following:

n!i=0d1(ji=1mi1xiji!)\displaystyle n!\prod_{i=0}^{d-1}\left(\sum_{j_{i}=1}^{m_{i}}\frac{1}{x_{i}^{j_{i}}!}\right)\; =n!i=0d11xi1!(ji=1mi1(xiji!)/(xi1!))n!i=0d11xi1!(ji=1mi1ji!)\displaystyle=\;n!\prod_{i=0}^{d-1}\frac{1}{x_{i}^{1}!}\left(\sum_{j_{i}=1}^{m_{i}}\frac{1}{(x_{i}^{j_{i}}!)/(x_{i}^{1}!)}\right)\;\leq\;n!\prod_{i=0}^{d-1}\frac{1}{x_{i}^{1}!}\left(\sum_{j_{i}=1}^{m_{i}}\frac{1}{j_{i}!}\right)
n!i=0d1exi1!=n!edi=0d11xi1!.\displaystyle\leq\;n!\prod_{i=0}^{d-1}\frac{e}{x_{i}^{1}!}\;=\;n!\cdot e^{d}\cdot\prod_{i=0}^{d-1}\frac{1}{x_{i}^{1}!}.

Since each xi0x_{i}\geq 0, we replace the value of xi1x_{i}^{1} with the value nνin\nu_{i} for each ii, where νi\nu_{i} is defined in the Theorem statement. Furthermore, below, since 0!=10!=1, we only need to multiply by those νi>0\nu_{i}>0. Then,

log2|Jq|\displaystyle\log_{2}|J_{q}| log2(n!edi=0d11xi1!)\displaystyle\leq\log_{2}\left(n!\cdot e^{d}\cdot\prod_{i=0}^{d-1}\frac{1}{x_{i}^{1}!}\right)
=log2(n!edνi01nνi!)\displaystyle=\log_{2}\left(n!\cdot e^{d}\cdot\prod_{\nu_{i}\neq 0}\frac{1}{\lceil n\nu_{i}\rceil!}\right)
=log2(n!)+dlog2eνi0log2(nνi!)\displaystyle=\log_{2}(n!)+d\log_{2}e-\sum_{\nu_{i}\neq 0}\log_{2}(\lceil n\nu_{i}\rceil!)
log2(enn+1/2en)+dlog2eνi0log2(2π(nνi)nνi+1/2enνi)\displaystyle\leq\log_{2}\left(en^{n+1/2}e^{-n}\right)+d\log_{2}e-\sum_{\nu_{i}\neq 0}\log_{2}\left(\sqrt{2\pi}(n\nu_{i})^{n\nu_{i}+1/2}e^{-n\nu_{i}}\right) (20)
nlogn+(d+1n)log2e+12log2nνi0((nνi+1/2)log2nνinνilog2e)\displaystyle\leq n\log n+(d+1-n)\log_{2}e+\frac{1}{2}\log_{2}n-\sum_{\nu_{i}\neq 0}\left((n\nu_{i}+1/2)\log_{2}n\nu_{i}-n\nu_{i}\log_{2}e\right)
nνi0νilog2νi+nlog2n(1νi0νi)+(d+1)log2e\displaystyle\leq-n\sum_{\nu_{i}\neq 0}\nu_{i}\log_{2}\nu_{i}+n\log_{2}n\left(1-\sum_{\nu_{i}\neq 0}\nu_{i}\right)+(d+1)\log_{2}e
+12(log2nνi0log2nνi)\displaystyle\;\;+\frac{1}{2}\left(\log_{2}n-\sum_{\nu_{i}\neq 0}\log_{2}n\nu_{i}\right) (21)
ni𝒜dνilog2νi+nlog2n(1i𝒜dνi)+(d+1)log2e\displaystyle\leq-n\sum_{i\in\mathcal{A}_{d}}\nu_{i}\log_{2}\nu_{i}+n\log_{2}n\left(1-\sum_{i\in\mathcal{A}_{d}}\nu_{i}\right)+(d+1)\log_{2}e
+12(log2ndlog2(n(1dδ)d))\displaystyle\;\;+\frac{1}{2}\left(\log_{2}n-d\log_{2}\left(\frac{n(1-d\delta)}{d}\right)\right) (22)
ni𝒜dνilog2νi+nlog2n(1i𝒜dνi)+(d+1)log2ed2log2(1dδd).\displaystyle\leq-n\sum_{i\in\mathcal{A}_{d}}\nu_{i}\log_{2}\nu_{i}+n\log_{2}n\left(1-\sum_{i\in\mathcal{A}_{d}}\nu_{i}\right)+(d+1)\log_{2}e-\frac{d}{2}\log_{2}\left(\frac{1-d\delta}{d}\right).

Inequality 20 follows from the Stirling upper and lower bounds. Then, the (d+1)(d+1) in inequality 21 follows from n(1iνi)log2e0-n(1-\sum_{i}\nu_{i})\log_{2}e\leq 0. Jensen’s Inequality and concavity of the logarithm imply inequality 22. ∎

Now we use Theorems 3.1 and 3.2 to analyze the protocol described above. Let ϵ>0\epsilon>0 be arbitrarily chosen by the user (this will determine the user’s desired failure probability and security properties). We use

δ=(m+n+2)ln(2d/ϵ2)m(m+n),\delta=\sqrt{\frac{(m+n+2)\ln(2d/\epsilon^{2})}{m(m+n)}}, (23)

which, by Lemma 2.2 implies that the failure probability will be ϵ2\epsilon^{2} (and so the ϵ\epsilon in Theorem 3.1 will match the chosen value of ϵ\epsilon here). Finally, let ϵPA=4ϵβ+9ϵ\epsilon_{PA}=4\epsilon^{\beta}+9\epsilon be the distance from an ideal uniform random string of size \ell independent of EE’s system.

Using Theorem 3.1 along with privacy amplification (Equation 6), we have that, except with probability at most 2ϵ12β2\epsilon^{1-2\beta}, the number of uniform random bits extracted from the protocol leading to an ϵPA\epsilon_{PA} secure string is:

ours=nlog2dlog2|Jq|2log21ϵ,\ell_{\text{ours}}=n\log_{2}d-\log_{2}|J_{q}|-2\log_{2}\frac{1}{\epsilon}, (24)

where

log2|Jq|min{,𝒢},\log_{2}|J_{q}|\leq\min\left\{\mathcal{F},\mathcal{G}\right\}, (25)
\displaystyle\mathcal{F} =ni𝒜dνilog2νi+nlog2n(1i𝒜dνi)+(d+1)log2ed2log2(1dδd),\displaystyle=-n\sum_{i\in\mathcal{A}_{d}}\nu_{i}\log_{2}\nu_{i}+n\log_{2}n\left(1-\sum_{i\in\mathcal{A}_{d}}\nu_{i}\right)+(d+1)\log_{2}e-\frac{d}{2}\log_{2}\left(\frac{1-d\delta}{d}\right), (26)
𝒢\displaystyle\mathcal{G} =nH¯d(1ν0)log2d\displaystyle=n\bar{H}_{d}(1-\nu_{0})\log_{2}d (27)

by Theorem 3.2 and the standard bound on the volume of a Hamming ball as discussed earlier. In our evaluations, we set ϵ=1036\epsilon=10^{-36} and β=1/3\beta=1/3 which balances the failure probability of Theorem 3.1 (namely, the probability of failure is 2ϵ12β2\epsilon^{1-2\beta}) and the smoothing parameter used in the min entropy. With these settings, the failure probability and the value of ϵPA\epsilon_{PA} are on the order of 101210^{-12}.

We compare our new lower bound ours\ell_{\text{ours}} for this protocol against the lower bound provided in [10] using alternative methods and an alternative entropic uncertainty relation. We also compare with another high-dimensional SI-QRNG from [32]. Note that, due to our bound on JqJ_{q} in Equation 25, our new result here will never be worse then the SI-QRNG protocol analyzed in our prior work [3] (which used Equation 27 only) and so we do not compare with that here.

A lower bound for the SI-QRNG protocol of [10], which we denote here as 1\ell_{1}, was given in that reference by:

1n(log2d2log2[Γ(m+d)Γ(m+d+12)i=0d1Γ(ci+32)Γ(ci+1)]),\ell_{1}\geq n\left(\log_{2}d-2\log_{2}\left[\frac{\Gamma(m+d)}{\Gamma(m+d+\frac{1}{2})}\sum_{i=0}^{d-1}\frac{\Gamma(c_{i}+\frac{3}{2})}{\Gamma(c_{i}+1)}\right]\right),

where mm is the test size and cic_{i} represents the number of measurement outcomes that result in outcome |xi\ket{x_{i}}.

The protocol of [32] is slightly different from the one we analyze. Here, an adversarial source prepares an entangled high-dimensional state (if the source were honest, it would prepare n+mn+m copies of the state |ψ0=1di=0d1|i,iA1A2\ket{\psi_{0}}=\frac{1}{\sqrt{d}}\sum_{i=0}^{d-1}\ket{i,i}_{A_{1}A_{2}}) sending the A1A_{1} and A2A_{2} registers to Alice. Alice chooses a random subset and measures the A1A_{1} and A2A_{2} qudit systems each in a dd-dimensional basis XX resulting in classical characters cA1(i)c_{A_{1}}(i) and cA2(i)c_{A_{2}}(i) corresponding to the iith iteration of registers A1A_{1} and A2A_{2}. For the remaining unmeasured systems, she discards the A2A_{2} system and measures only the A1A_{1} system in the ZZ basis resulting in her secret string. If the source were honest, it should be that the XX basis measurement outcomes of the A1A_{1} and A2A_{2} register are fully correlated. She then applies privacy amplification to the result of the ZZ basis measurement. A lower bound for the number of random bits that may be extracted from this protocol, which we denote here as 2\ell_{2}, was computed in [32]:

2=nlog2dlog2γ(d0+δ),\ell_{2}=n\log_{2}d-\log_{2}\gamma(d_{0}+\delta^{\prime}),

where

γ(x)=(x+1+x2)(x1+x21)x\gamma(x)=(x+\sqrt{1+x^{2}})\left(\frac{x}{\sqrt{1+x^{2}}-1}\right)^{x} (28)

and

δ=dN2n2mln(4ϵ).\delta^{\prime}=d\sqrt{\frac{N^{2}}{n^{2}m}\ln\left(\frac{4}{\epsilon}\right)}. (29)

The term d0d_{0} is computed as the average difference between the measurements values of the pairs, cA1(i)c_{A_{1}}(i) and cA2(i)c_{A_{2}}(i) for ii from 1 to mm. That is, d0=1mi=1m|cA1(i)cA2(i)|.d_{0}=\frac{1}{m}\sum_{i=1}^{m}|c_{A_{1}}(i)-c_{A_{2}}(i)|.

For all protocols, we assume a failure probability on the order of 101210^{-12}. The difference from the ideal random string (Equation 6) is also set to be 101210^{-12}. As we are only interested in comparing the relative performance, we do not consider the additional randomness used to choose a random subset of size mm. Since all protocols in our evaluation are using the same process for this and same sampling sizes (in particular we use 7%7\% of all signals for sampling), they will each lose the same amount from their respective \ell values and so the comparison remains unchanged.

Note that for each of the three bounds, no assumption is needed on the noise in the channel - Alice simply uses the direct measurement result from the test case (in the XX basis) and evaluates \ell. To compare, however, we will simulate certain noise scenarios. We first compare these protocols assuming a depolarization channel acting on each qudit state independently and identically. Such a channel will cause the qudit to become the completely mixed state with some probability QQ; otherwise it remains in its original state. In this setting, we see the protocol of [10], but augmented using our new entropic uncertainty relation here outperforms both 1\ell_{1} and 2\ell_{2}. Since 1\ell_{1} is the same protocol we are analyzing with ours\ell_{ours} this shows the great benefit of sampling-based entropic uncertainty relations. This evaluation is shown in Figures 1, 2. Next, we evaluate on asymmetric channels which are more likely to add noise towards one basis vector over another (i.e., it is more likely to change a |0\ket{0} to a |1\ket{1} as opposed to changing a |0\ket{0} to a |3\ket{3}). Depending on the state favored by the channel, our bound generally outperforms prior work as shown in Figures 3 (right), 4 and 5 (right), though there are scenarios where the protocol of [32] can outperform our analysis as shown in Figures 3 (left) and 5 (left). Note, however, that the protocol of [32] is a different protocol; our methods applied to that protocol may provide a boost in performance in this scenario also, a question we leave as future work. Comparing ours\ell_{ours} with 1\ell_{1}, which is the generation rate for the same protocol of [10], shows that our new entropic uncertainty relation always leads to more optimistic bit generation rates in every scenario we simulated. Also, note that in all cases (including in the case highlighted in Figures 3 and 5), if we take the number of signals to be high enough, our bound outperforms.

Refer to caption
Refer to caption
Figure 1: Random bit generation rates. xx-axis: Total number of signals NN from which .07N.07N are used for sampling; yy-axis: Random bit generation rate /N\ell/N. Solid black: ours/N\ell_{\text{ours}}/N; Dotted red: 1/N\ell_{1}/N from [10] (same protocol, different security analysis); Dashed blue: 2/N\ell_{2}/N from [32] (different protocol and different security analysis method). Both graphs plot d=4d=4 with c(q)=(0.8,1/15,1/15,1/15)c(q)=(0.8,1/15,1/15,1/15) (recall that c(q)c(q) denotes the dd-tuple of character counts as discussed in Section 1.1). The left and right graphs plot N108N\leq 10^{8} and N1010N\leq 10^{10} respectively.
Refer to caption
Refer to caption
Figure 2: Random bit generation rates. xx-axis: Total number of signals NN; yy-axis: Random bit generation rate /N\ell/N. Solid black: ours/N\ell_{\text{ours}}/N; Dotted red: 1/N\ell_{1}/N from [10]; Dashed blue: 2/N\ell_{2}/N from [32]. Both graphs plot d=16d=16 with c(q)=(0.7,0.02,0.02,,0.02)c(q)=(0.7,0.02,0.02,\dots,0.02). The left and right graphs plot N108N\leq 10^{8} and N1010N\leq 10^{10} respectively.
Refer to caption
Refer to caption
Figure 3: Random bit generation rates. xx-axis: Total number of signals NN; yy-axis: Random bit generation rate /N\ell/N. Solid black: ours/N\ell_{\text{ours}}/N; Dotted red: 1/N\ell_{1}/N from [10]; Dashed blue: 2/N\ell_{2}/N from [32]. Both graphs plot d=4d=4 with c(q)=(0.8,0.19,0.005,0.005)c(q)=(0.8,0.19,0.005,0.005). The left and right graphs plot N108N\leq 10^{8} and 109N101010^{9}\leq N\leq 10^{10} respectively.
Refer to caption
Refer to caption
Figure 4: Random bit generation rates. xx-axis: Total number of signals NN; yy-axis: Random bit generation rate /N\ell/N. Solid black: ours/N\ell_{\text{ours}}/N; Dotted red: 1/N\ell_{1}/N from [10]; Dashed blue: 2/N\ell_{2}/N from [32]. Both graphs plot d=4d=4 with c(q)=(0.8,0.15,0.025,0.025)c(q)=(0.8,0.15,0.025,0.025). The left and right graphs plot N108N\leq 10^{8} and 3×109N10103\times 10^{9}\leq N\leq 10^{10} respectively.
Refer to caption
Refer to caption
Figure 5: Random bit generation rates. xx-axis: Total number of signals NN; yy-axis: Random bit generation rate /N\ell/N. Solid black: ours/N\ell_{\text{ours}}/N; Dotted red: 1/N\ell_{1}/N from [10]; Dashed blue: 2/N\ell_{2}/N from [32]. Both graphs plot d=16d=16 with c(q)=(0.7,0.16,0.075,0.035,0.0025,0.0025,,0.0025)c(q)=(0.7,0.16,0.075,0.035,0.0025,0.0025,\dots,0.0025). The left and right graphs plot N4×107N\leq 4\times 10^{7} and 1.05×109N10101.05\times 10^{9}\leq N\leq 10^{10} respectively.

In summary, Figures 1 and 2 highlight how ours\ell_{\text{ours}} consistently outperforms 1\ell_{1} [10] and 2\ell_{2} [32] on a depolarization channel for different dimensions dd. Our bound for ours\ell_{\text{ours}} still performs very well on systems far from depolarization, as shown in Figure 4. However, there can exist quantum channels which lead to ours\ell_{\text{ours}} producing a lower random bit generation rate than 2\ell_{2} for certain NN. Even in these cases, Figures 3 and 5 highlight that, assuming sufficient computational power to process larger blocks in the post-processing stage of the protocol, ours\ell_{\text{ours}} can produce a much higher random bit generation rate than 1\ell_{1} and 2\ell_{2} on a large block of signals.

3.2 Asymptotic Behavior and Analysis

In Equation 25, we take the bound for log2|Jq|\log_{2}|J_{q}| to be the minimum of Theorem 3.2 and the size of a Hamming ball. The reason for doing so is that while the bound on the size of a Hamming ball is tighter for some scenarios, the bound from Theorem 3.2 is significantly better in others, especially, as our numerical simulations show, for large numbers of signals. In this section, we analyze and compare the asymptotic behavior of both bounds. We will also use this work to show an alternative proof of the famous Maassen-Uffink relation [4] for high dimensional systems.

First, we prove a technical lemma about the relation between the dd-ary entropy function HdH_{d} and the Shannon entropy, which will be needed to analyze the asymptotic behavior.

Lemma 3.1.

Let XX be a discrete random variable with dd possible outcomes {x0,,xd1}\{x_{0},\dots,x_{d-1}\} such that the probability of observing outcome xjx_{j} is pjp_{j} for each jj. Then for any ii, it holds that

Hd(1pi)logd2H(X)H_{d}(1-p_{i})\geq\log_{d}2\cdot H(X)

where equality holds if and only if pj=pkp_{j}=p_{k} for all j,kij,k\neq i (i.e., if the distribution is uniform on the other outcomes not equal to ii).

Proof.

Fix i{0,,d1}i\in\{0,\dots,d-1\} and let YY be the random variable where the probability of observing xix_{i} is qi=piq_{i}=p_{i} and the probability of observing xjx_{j} for any jij\neq i is qj=1pid1q_{j}=\frac{1-p_{i}}{d-1}. Then

Hd(1pi)\displaystyle H_{d}(1-p_{i}) =(1pi)logd(d1)(1pi)logd(1pi)pilogd(pi)\displaystyle=(1-p_{i})\log_{d}(d-1)-(1-p_{i})\log_{d}(1-p_{i})-p_{i}\log_{d}(p_{i})
=logd2[pilog2(pi)ji1pid1log2(1pid1)]\displaystyle=\log_{d}2\left[-p_{i}\log_{2}(p_{i})-\sum_{j\neq i}\frac{1-p_{i}}{d-1}\log_{2}\left(\frac{1-p_{i}}{d-1}\right)\right]
=logd2[pilog2(pi)jiqjlog2(qj)]=H(Y)logd2.\displaystyle=\log_{d}2\left[-p_{i}\log_{2}(p_{i})-\sum_{j\neq i}q_{j}\log_{2}\left(q_{j}\right)\right]\;=\;H(Y)\cdot\log_{d}2.

Moreover, observe that

H(X)\displaystyle H(X)\; =pilog2(pi)jipjlog2(pj)\displaystyle=\;-p_{i}\log_{2}(p_{i})-\sum_{j\neq i}p_{j}\log_{2}(p_{j})
pilog2(pi)jiqjlog2(qj)=H(Y).\displaystyle\leq\;-p_{i}\log_{2}(p_{i})-\sum_{j\neq i}q_{j}\log_{2}(q_{j})\;=\;H(Y).

Note the inequality is shown by recalling that Shannon entropy is maximal if and only if given a uniform distribution (which also proves equality if the distribution is uniform on outcomes other than xix_{i}). ∎

We now show that our bound for log2|Jq|/n\log_{2}|J_{q}|/n converges to the Shannon entropy of the random variable induced by a measurement on some i.i.d. system. This will then lead us to an alternative proof of the Maassen-Uffink relation from [4].

Lemma 3.2.

Let ρ\rho be a quantum state acting on d\mathcal{H}_{d} and consider the nn-fold tensor state ρ=ρn\rho^{\prime}=\rho^{\otimes n}. Furthermore, let δ=O(1n)\delta=O\left(\frac{1}{\sqrt{n}}\right). Consider measuring all nn qudits of the state ρ\rho^{\prime} in some dd-dimensional orthonormal basis XX resulting in some q𝒜dnq\in\mathcal{A}_{d}^{n} and from this define the set Jq={i𝒜dn:maxj|cj(i)cj(q)|δ}J_{q}=\{i\in\mathcal{A}_{d}^{n}:\max_{j}|c_{j}(i)-c_{j}(q)|\leq\delta\} as before. Then it follows that

limnlog2|Jq|nH(X)ρ.\lim_{n\to\infty}\frac{\log_{2}|J_{q}|}{n}\leq H(X)_{\rho}.
Proof.

Define νi=max{ci(q)δ,0}\nu_{i}=\max\{c_{i}(q)-\delta,0\} and let

(q,n,d,δ)=ni𝒜dνilog2νi+nlog2n(1i𝒜dνi)+(d+1)log2ed2log2(1dδd).\mathcal{F}(q,n,d,\delta)=-n\sum_{i\in\mathcal{A}_{d}}\nu_{i}\log_{2}\nu_{i}+n\log_{2}n\left(1-\sum_{i\in\mathcal{A}_{d}}\nu_{i}\right)+(d+1)\log_{2}e-\frac{d}{2}\log_{2}\left(\frac{1-d\delta}{d}\right).

Observe that (q,n,d,δ)niνilog2(νi)+dδlog2n+O(1)n\frac{\mathcal{F}(q,n,d,\delta)}{n}\leq-\sum_{i}\nu_{i}\log_{2}(\nu_{i})+d\delta\log_{2}n+\frac{O(1)}{n} where we used the fact that iνi1dδ\sum_{i}\nu_{i}\geq 1-d\delta. Since δ=O(1n)\delta=O\left(\frac{1}{\sqrt{n}}\right) we have

(q,n,d,δ)n=i𝒜dνilog2(νi)+O(log2nn)+O(1n).\frac{\mathcal{F}(q,n,d,\delta)}{n}=-\sum_{i\in\mathcal{A}_{d}}\nu_{i}\log_{2}(\nu_{i})+O\left(\frac{\log_{2}n}{\sqrt{n}}\right)+O\left(\frac{1}{n}\right).

Then, νi=max{ci(q)δ,0}pi\nu_{i}=\max\{c_{i}(q)-\delta,0\}\to p_{i} as nn\to\infty by the law of large numbers and the assumption on δ\delta, where we use pip_{i} to denote the probability of observing |xi\ket{x_{i}}, the ii’th basis vector in the measurement basis XX. Hence,

i𝒜dνilog2(νi)i𝒜dpilog2(pi).\sum_{i\in\mathcal{A}_{d}}\nu_{i}\log_{2}(\nu_{i})\to\sum_{i\in\mathcal{A}_{d}}p_{i}\log_{2}(p_{i}).

Finally, by Theorem 3.2, we have log2|Jq|(q,n,d,δ)\log_{2}|J_{q}|\leq\mathcal{F}(q,n,d,\delta), and so we conclude

limnlog2|Jq|nlimn(q,n,d,δ)n\displaystyle\lim_{n\to\infty}\frac{\log_{2}|J_{q}|}{n}\leq\lim_{n\to\infty}\frac{\mathcal{F}(q,n,d,\delta)}{n} limn(νi0νilog2(νi)+O(log2nn)+O(1n))\displaystyle\leq\lim_{n\to\infty}\left(-\sum_{\nu_{i}\neq 0}\nu_{i}\log_{2}(\nu_{i})+O\left(\frac{\log_{2}n}{\sqrt{n}}\right)+O\left(\frac{1}{n}\right)\right)
=i𝒜dpilog2(pi)=H(X)ρ\displaystyle=-\sum_{i\in\mathcal{A}_{d}}p_{i}\log_{2}(p_{i})\;\;=\;\;H(X)_{\rho}

Now we are ready to show that our bound for log2|Jq|\log_{2}|J_{q}| grows at most as quickly as the volume of a Hamming ball (used in our earlier work in [3]). How much slower our bound grows asymptotically depends on the observed relative counts from the sampled qq.

Theorem 3.3.

Let νi=max{ci(q)δ,0}\nu_{i}=\max\{c_{i}(q)-\delta,0\} for any i𝒜di\in\mathcal{A}_{d}. Let 𝒢(q,n,d,δ)=nH¯d(1νa)logd2\mathcal{G}(q,n,d,\delta)=\frac{n\bar{H}_{d}(1-\nu_{a})}{\log_{d}2}, where aa is the element of 𝒜d\mathcal{A}_{d} such that ca(q)=maxi𝒜dci(q)c_{a}(q)=\max_{i\in\mathcal{A}_{d}}c_{i}(q), and

(q,n,d,δ)=ni𝒜dνilog2νi+nlog2n(1i𝒜dνi)+(d+1)log2ed2log2(1dδd).\mathcal{F}(q,n,d,\delta)=-n\sum_{i\in\mathcal{A}_{d}}\nu_{i}\log_{2}\nu_{i}+n\log_{2}n\left(1-\sum_{i\in\mathcal{A}_{d}}\nu_{i}\right)+(d+1)\log_{2}e-\frac{d}{2}\log_{2}\left(\frac{1-d\delta}{d}\right).

Let δ\delta depend on nn and δ=O(1n)\delta=O\left(\frac{1}{\sqrt{n}}\right) asymptotically. Then for arbitrary quantum state ρ\rho acting on d\mathcal{H}_{d}, we have that

limn(q,n,d,δ)𝒢(q,n,d,δ)1\lim_{n\to\infty}\frac{\mathcal{F}(q,n,d,\delta)}{\mathcal{G}(q,n,d,\delta)}\leq 1

where equality holds if and only if pi=1pad1p_{i}=\frac{1-p_{a}}{d-1} for all iai\neq a given pkp_{k} is the probability of observing outcome kk in basis XX (measuring ρ\rho).

Proof.

Notice that by the proof of Lemma 3.2

limn(q,n,d,δ)𝒢(q,n,d,δ)\displaystyle\lim_{n\to\infty}\frac{\mathcal{F}(q,n,d,\delta)}{\mathcal{G}(q,n,d,\delta)} =limnlogd2H¯d(1νa)limn(q,n,d,δ)n\displaystyle=\lim_{n\to\infty}\frac{\log_{d}2}{\bar{H}_{d}(1-\nu_{a})}\cdot\lim_{n\to\infty}\frac{\mathcal{F}(q,n,d,\delta)}{n}
H(X)ρlimnlogd2H¯d(1νa).\displaystyle\leq H(X)_{\rho}\cdot\lim_{n\to\infty}\frac{\log_{d}2}{\bar{H}_{d}(1-\nu_{a})}.

Then, by Lemma 3.1 and the definition of H¯d\bar{H}_{d}, it follows that

H¯d(1νa)Hd(1νa)logd2H(X)ρ\bar{H}_{d}(1-\nu_{a})\geq H_{d}(1-\nu_{a})\geq\log_{d}2\cdot H(X)_{\rho}

and hence

H(X)ρlimnlogd2H¯d(1νa)=limnlogd2H(X)ρH¯d(1νa)1.H(X)_{\rho}\cdot\lim_{n\to\infty}\frac{\log_{d}2}{\bar{H}_{d}(1-\nu_{a})}=\lim_{n\to\infty}\frac{\log_{d}2\cdot H(X)_{\rho}}{\bar{H}_{d}(1-\nu_{a})}\leq 1.

where equality holds if and only if logd2H(X)=limnH¯d(1νa)\log_{d}2\cdot H(X)=\lim_{n\to\infty}\bar{H}_{d}(1-\nu_{a}). This is true if and only if pi=1pad1p_{i}=\frac{1-p_{a}}{d-1} by Lemma 3.1 and the fact that νapa\nu_{a}\to p_{a} as nn\to\infty. ∎

3.2.1 Alternative Proof of Maassen-Uffink Relation

With the above analysis, our Theorems 3.1 and 3.2 can be used to provide an alternative proof of the Maassen and Uffink entropic uncertainty relation for projective basis measurements of dd-dimensional states. Note that in [2] we showed quantum sampling can be used to provide an alternative proof of this relation but only for the qubit case. Furthermore, our earlier work in [3] also cannot lead to an alternative proof of this relation in the high dimensional (d3d\geq 3) case as Theorem 3.3 shows.

Corollary 3.1.

Let Z={|zi}i𝒜dZ=\{\ket{z_{i}}\}_{i\in\mathcal{A}_{d}} and X={|xi}i𝒜dX=\{\ket{x_{i}}\}_{i\in\mathcal{A}_{d}} be two orthonormal bases and let ρ\rho be a density operator acting on d\mathcal{H}_{d}. Then, except with arbitrarily small probability, it holds that

H(Z)ρ+H(X)ργ,H(Z)_{\rho}+H(X)_{\rho}\geq\gamma,

where γ=logmaxi,j|zi|xj|2\gamma=-\log\max_{i,j}|\braket{z_{i}}{x_{j}}|^{2}.

Proof.

Consider the state ρ=ρ2n\rho^{\prime}=\rho^{\otimes 2n}. We apply Theorem 3.1 to ρ\rho^{\prime} using sampling strategy Ψ1\Psi_{1} with m=nm=n. Since ρ\rho^{\prime} is i.i.d., for any subset tt of size nn and any measurement outcome qq on that subset, the post-measurement state is simply ρn\rho^{\otimes n}.

Fix ϵ^>0\widehat{\epsilon}>0 and 0<β<1/20<\beta<1/2. Then, for any nn and ϵϵ^\epsilon\leq\widehat{\epsilon}, setting δ=(n+1)ln(2d/ϵ2)n2\delta=\sqrt{\frac{(n+1)\ln(2d/\epsilon^{2})}{n^{2}}}, Theorem 3.1 implies that, except with probability at most ϵ^12β\widehat{\epsilon}^{1-2\beta}, the inequality

1nHmin4ϵ+2ϵβ(Z|E)ρ(t,q)+1nlog2|Jq(n)|γ\frac{1}{n}H_{\text{min}}^{4\epsilon+2\epsilon^{\beta}}(Z|E)_{\rho(t,q)}+\frac{1}{n}\log_{2}|J_{q}^{(n)}|\geq\gamma

holds, where qq is the observed value after measuring using ZZ. Then, by the asymptotic equipartition property, it follows that

limϵ0limn1nHmin4ϵ+2ϵβ(Z|E)ρn=H(Z|E)ρ.\lim_{\epsilon\to 0}\lim_{n\to\infty}\frac{1}{n}H_{\text{min}}^{4\epsilon+2\epsilon^{\beta}}(Z|E)_{\rho^{\otimes n}}=H(Z|E)_{\rho}.

This, combined with Lemma 3.2 and the fact that H(Z)H(Z|E)H(Z)\geq H(Z|E), completes the proof.

4 A Three-Party Sampling-Based Entropic Uncertainty Relation

We now turn our attention to deriving a new three-party sampling-based entropic uncertainty relation involving Alice, Bob, and Eve. Later we show an application to a finite key analysis of the high-dimensional BB84 [13]. To begin, consider the following experiment, extending an earlier version to this three party case: on an input state of the form ρTABE=tA,tBp(tA,tB)|tA,tBtA,tB|ρABEtA,tB\rho_{TABE}=\sum_{t_{A},t_{B}}p(t_{A},t_{B})\ket{t_{A},t_{B}}\bra{t_{A},t_{B}}\otimes\rho_{ABE}^{t_{A},t_{B}}, choose a random subset t=(tA,tB)t=(t_{A},t_{B}) by measuring the TT register, causing the state to collapse to ρABEtA,tB\rho_{ABE}^{t_{A},t_{B}} (though, as before, this ρ\rho portion may be independent of the chosen subset in which case a random subset is chosen which does not affect the rest of the input state). We assume |tA|=|tB|=m|t_{A}|=|t_{B}|=m. Next, a portion of the AA and BB registers, indexed by the chosen subsets, are measured in basis X={|x0,,|xd1}X=\{\ket{x_{0}},\cdots,\ket{x_{d-1}}\} resulting in outcome qA,qB𝒜dmq_{A},q_{B}\in\mathcal{A}_{d}^{m}. This measurement causes the remaining state to collapse to ρABE(t,qA,qB)\rho_{ABE}(t,q_{A},q_{B}). The experiment outputs (t,qA,qB,ρABE(t,qA,qB))Exp(ρTABE,X)(t,q_{A},q_{B},\rho_{ABE}(t,q_{A},q_{B}))\leftarrow\textbf{{Exp}}\left(\rho_{TABE},X\right).

Note that technically, by considering Alice and Bob as one party for the sampling portion, one could potentially use Theorem 3.1 with a suitable sampling strategy similar to Ψ0\Psi_{0} or Ψ1\Psi_{1}. However, this would bound the resulting min entropy as a function of the set JqA,qB={(i,j)𝒜d2n | |ΔH(qA,qB)ΔH(i,j)|δ}J_{q_{A},q_{B}}=\{(i,j)\in\mathcal{A}_{d}^{2n}\text{ }|\text{ }|\Delta_{H}(q_{A},q_{B})-\Delta_{H}(i,j)|\leq\delta\}. It is not difficult to see that |JqA,qB|dn|J_{q_{A},q_{B}}|\geq d^{n} (since for any fixed qAq_{A} and qBq_{B}, and for every i𝒜dni\in\mathcal{A}_{d}^{n}, one may find a j𝒜dnj\in\mathcal{A}_{d}^{n} satisfying (i,j)JqA,qB(i,j)\in J_{q_{A},q_{B}}). Recalling from our Theorem that the min entropy is higher when the size of this set is smaller. This would always produce the trivial bound of Hmin(A|E)0H_{\text{min}}(A|E)\geq 0 and so Theorem 3.1 cannot be used for the three-party case. We prove that sampling can provide an entropic uncertainty relation in this scenario for high-dimensional states by suitably modifying the first step of our proof method. Furthermore, we show how our proof method can lead to relations incorporating more than one overlap, useful in case the two bases have a shared vector in common (e.g., a “vaccuum” state vector for QKD).

Theorem 4.1.

Let ϵ>0\epsilon>0, 0<β<1/20<\beta<1/2, and ρABE\rho_{ABE} be an arbitrary quantum state acting on ABE\mathcal{H}_{A}\otimes\mathcal{H}_{B}\otimes\mathcal{H}_{E}, where ABd(n+m)\mathcal{H}_{A}\cong\mathcal{H}_{B}\cong\mathcal{H}_{d}^{\otimes(n+m)} with d2d\geq 2 and mnm\leq n. Let ZZ and XX be two orthonormal bases of d\mathcal{H}_{d} and define the maximal overlap γ^\hat{\gamma} as γ^=log2maxa,b|za|xb|2\hat{\gamma}=-\log_{2}\max_{a,b}|\braket{z_{a}}{x_{b}}|^{2}. Let aa^{*} and bb^{*} be a pair that attains this maximum, then we define the second-greatest overlap as:

γ=log2maxaabb|za|xb|2.\gamma=-\log_{2}\max_{\begin{subarray}{c}a\neq a^{*}\\ b\neq b^{*}\end{subarray}}|\braket{z_{a}}{x_{b}}|^{2}.

(It is possible that γ=γ^\gamma=\hat{\gamma} for some bases.) Let

δ=(m+n+2)ln(4/ϵ2)m(m+n).\delta=\sqrt{\frac{(m+n+2)\ln(4/\epsilon^{2})}{m(m+n)}}.

Finally, let ρTABE=1Tt|tt|ρABE\rho_{TABE}=\frac{1}{T}\sum_{t}\ket{t}\bra{t}\otimes\rho_{ABE}, where the sum is over all subsets of the form t=(tA,tB)t=(t_{A},t_{B}) with tA=tBt_{A}=t_{B} (over their respective subspaces) and T=(n+mm)T={n+m\choose m}. Then, except with probability at most 2ϵ12β2\epsilon^{1-2\beta}, after running (t,qA,qB,ρABE(t,qA,qB))Exp(ρTABE,X)(t,q_{A},q_{B},\rho_{ABE}(t,q_{A},q_{B}))\leftarrow\textbf{{Exp}}\left(\rho_{TABE},X\right), it holds that:

Hmin4ϵ+2ϵβ(AZ|E)ρ(t,qA,qB)+nH¯(ΔH(qA,qB)+δ)logd2n(cb(qA)+δ)γ^+n(1cb(qA)δ)γH_{\text{min}}^{4\epsilon+2\epsilon^{\beta}}(A_{Z}|E)_{\rho(t,q_{A},q_{B})}+\frac{n\bar{H}(\Delta_{H}(q_{A},q_{B})+\delta)}{\log_{d}2}\geq n(c_{b^{*}}(q_{A})+\delta)\hat{\gamma}+n(1-c_{b^{*}}(q_{A})-\delta)\gamma

where AZA_{Z} above, denotes the random variable resulting from measuring the remainder of the AA system of ρ(t,qA,qB)\rho(t,q_{A},q_{B}) in the ZZ basis and the probability is over all choices of subsets and measurement outcomes within the experiment. If γ^=γ\hat{\gamma}=\gamma, then the above simplifies to:

Hmin4ϵ+2ϵβ(AZ|E)ρ(t,qA,qB)+nH¯(ΔH(qA,qB)+δ)logd2nγH_{\text{min}}^{4\epsilon+2\epsilon^{\beta}}(A_{Z}|E)_{\rho(t,q_{A},q_{B})}+\frac{n\bar{H}(\Delta_{H}(q_{A},q_{B})+\delta)}{\log_{d}2}\geq n\gamma
Proof.

As with our other proofs of sampling based entropic uncertainty relations, this one follows the same two-step structure where, first, we analyze the ideal case, proving the result there; then, finally, we argue that the real case must follow the ideal except with small probability of failure. For this three party version, only the first step changes from our proof of Theorem 3.1, the second step is identical.

Consider sampling strategy Ψ2+0\Psi_{2+0} defined in Section 2.1 with the count index set to bb^{*} which is the classical strategy we will employ in this scenario. By Theorem 2.1, there exist ideal states {|ϕABEt}\{\ket{\phi^{t}_{ABE}}\}, indexed over all subsets t=(tA,tB)t=(t_{A},t_{B}), such that |ϕABEtspan(𝒢t,δ)E\ket{\phi^{t}_{ABE}}\in\text{span}(\mathcal{G}_{t,\delta})\otimes\mathcal{H}_{E} and, in this case as we are using Ψ2+0\Psi_{2+0}, the set

𝒢t,δ={(i,j)𝒜dN×𝒜dN | |ΔH(itA,jtB)ΔH(itA,jtB)|δ and |cb(itA)cb(itA)|δ}.\mathcal{G}_{t,\delta}=\{(i,j)\in\mathcal{A}_{d}^{N}\times\mathcal{A}_{d}^{N}\text{ }|\text{ }|\Delta_{H}(i_{t_{A}},j_{t_{B}})-\Delta_{H}(i_{-t_{A}},j_{-t_{B}})|\leq\delta\text{ and }|c_{b^{*}}(i_{t_{A}})-c_{b^{*}}(i_{-t_{A}})|\leq\delta\}.

Furthermore, by our choice of δ\delta, and the failure probability of Ψ2+0\Psi_{2+0} (from Equation 12), we have: 12ρTABEσTABEϵ,\frac{1}{2}\left|\left|\rho_{TABE}-\sigma_{TABE}\right|\right|\leq\epsilon, where σTABE\sigma_{TABE} is the ideal state defined over all subsets and individual ideal states above (as in Theorem 2.1). If we consider performing the given experiment on this ideal state, afterwards, we will receive as output the chosen subset tt, the measurement results qA,qBq_{A},q_{B}, and the post-measurement state |ϕt(qA,qB)ABE\ket{\phi^{t}(q_{A},q_{B})}_{ABE} which is guaranteed to be of the form:

|ϕt(qA,qB)=(i,j)JqA,qBαi,j|iA|jB|Ei,j,\ket{\phi^{t}(q_{A},q_{B})}=\sum_{(i,j)\in J_{q_{A},q_{B}}}\alpha_{i,j}\ket{i}_{A}\ket{j}_{B}\ket{E_{i,j}},

where the above |iA\ket{i}_{A} and |jB\ket{j}_{B} are XX basis vectors (i.e., |xi\ket{x_{i}} and |xj\ket{x_{j}}), and:

JqA,qB={(i,j)𝒜d2n | |ΔH(qA,qB)ΔH(i,j)|δ and |cb(qA)cb(i)|δ}.J_{q_{A},q_{B}}=\{(i,j)\in\mathcal{A}_{d}^{2n}\text{ }|\text{ }|\Delta_{H}(q_{A},q_{B})-\Delta_{H}(i,j)|\leq\delta\text{ and }|c_{b^{*}}(q_{A})-c_{b^{*}}(i)|\leq\delta\}.

Rearranging terms and permuting the AA and BB subspaces, we may write the above state as:

|ϕt(qA,qB)|ϕ~t(qA,qB)=jYα~j|jBiJqA,qB(j)βi(j)|iA|E~i,j,\ket{\phi^{t}(q_{A},q_{B})}\cong\ket{\widetilde{\phi}^{t}(q_{A},q_{B})}=\sum_{j\in Y}\widetilde{\alpha}_{j}\ket{j}_{B}\otimes\sum_{i\in J_{q_{A},q_{B}}^{(j)}}\beta_{i}^{(j)}\ket{i}_{A}\ket{\widetilde{E}_{i,j}},

where Y𝒜dnY\subset\mathcal{A}_{d}^{n} and JqA,qB(j){i𝒜dn | |ΔH(i,j)ΔH(qA,qB)|δ and |cb(qA)cb(i)|δ}J_{q_{A},q_{B}}^{(j)}\subset\{i\in\mathcal{A}_{d}^{n}\text{ }|\text{ }|\Delta_{H}(i,j)-\Delta_{H}(q_{A},q_{B})|\leq\delta\text{ and }|c_{b^{*}}(q_{A})-c_{b^{*}}(i)|\leq\delta\}. Note that some of the α~\widetilde{\alpha} and β\beta’s may be zero. Tracing out BB leaves us with:

σAE=jY|α~j|2P[iJqA,qB(j)βi(j)|iA|E~i,j]=jY|α~j|2σAE(j),\sigma_{AE}=\sum_{j\in Y}|\widetilde{\alpha}_{j}|^{2}P\left[\sum_{i\in J_{q_{A},q_{B}}^{(j)}}\beta_{i}^{(j)}\ket{i}_{A}\ket{\widetilde{E}_{i,j}}\right]=\sum_{j\in Y}|\widetilde{\alpha}_{j}|^{2}\sigma_{AE}^{(j)},

where P(z)=zzP(z)=zz^{*}. At this point, AA measures the remaining portion of her register in the ZZ basis, resulting in jσAZ,E(j)\sum_{j}\sigma_{A_{Z},E}^{(j)}. By appending a suitable classical system and conditioning on it, we may use Equation 4, to show that

Hmin(AZ|E)σminjHmin(AZ|E)σ(j).H_{\text{min}}(A_{Z}|E)_{\sigma}\geq\min_{j}H_{\text{min}}(A_{Z}|E)_{\sigma^{(j)}}.

Consider a particular jj and define χAE(j)=iJ(qA,qB)(j)|βi(j)|2|ii||Ei,j~Ei,j~|\chi^{(j)}_{AE}=\sum_{i\in J_{(q_{A},q_{B})}^{(j)}}|\beta_{i}^{(j)}|^{2}\ket{i}\bra{i}\otimes\ket{\widetilde{E_{i,j}}}\bra{\widetilde{E_{i,j}}}. From Lemma 1.1, we have:

Hmin(AZ|E)σ(j)\displaystyle H_{\text{min}}(A_{Z}|E)_{\sigma^{(j)}} Hmin(AZ|E)χ(j)log2|J(qA,qB)(j)|\displaystyle\geq H_{\text{min}}(A_{Z}|E)_{\chi^{(j)}}-\log_{2}|J_{(q_{A},q_{B})}^{(j)}|

We first bound Hmin(AZ|E)χ(j)H_{\text{min}}(A_{Z}|E)_{\chi^{(j)}}. Taking χ(j)\chi^{(j)} and measuring in the ZZ basis yields:

χZE(j)=iJ(qA,qB)(j)|βi(j)|2(z𝒜dnp(z|i)|zz|)|Ei,j~Ei,j~|\chi^{(j)}_{ZE}=\sum_{i\in J_{(q_{A},q_{B})}^{(j)}}|\beta_{i}^{(j)}|^{2}\left(\sum_{z\in\mathcal{A}_{d}^{n}}p(z|i)\ket{z}\bra{z}\right)\otimes\ket{\widetilde{E_{i,j}}}\bra{\widetilde{E_{i,j}}}

where

p(z|i)=|z|xi|2=k=1n|zk|xk|2p(z|i)=|\braket{z}{x_{i}}|^{2}=\prod_{k=1}^{n}|\braket{z_{k}}{x_{k}}|^{2}

We wish to find an upper bound on p(z|i)p(z|i) for any zz and ii (within our constraints on ii) which will be used shortly to bound the min entropy of the system. Recall, we have two particular overlaps we are considering: one for za|xb\braket{z_{a^{*}}}{x_{b^{*}}} and one for the remaining possible pairs. It is not difficult to see that p(z|i)p(z|i) is maximized if, whenever ik=bi_{k}=b^{*} that we have zk=az_{k}=a^{*}. This can happen at most n(cb(qA)+δ)n(c_{b^{*}}(q_{A})+\delta) times due to our constraint on ii and so the remaining counts (namely n(1cb(qA)δ)n(1-c_{b^{*}}(q_{A})-\delta)) will be bounded using γ\gamma. Thus, we conclude:

p(z|i)=k=1n|zk|xk|2(|za|xb|2)n(cb(qA)+δ)×(maxaabb|za|xb|2)n(1cb(qA)δ).p(z|i)=\prod_{k=1}^{n}|\braket{z_{k}}{x_{k}}|^{2}\leq\left(|\braket{z_{a^{*}}}{x_{b^{*}}}|^{2}\right)^{n(c_{b^{*}}(q_{A})+\delta)}\times\left(\max_{\begin{subarray}{c}a\neq a^{*}\\ b\neq b^{*}\end{subarray}}|\braket{z_{a}}{x_{b}}|^{2}\right)^{n(1-c_{b^{*}}(q_{A})-\delta)}. (30)

Finally, we append a classical system spanned by orthonormal basis {|iI}\{\ket{i}_{I}\} for all iJ(qA,qB)(j)i\in J_{(q_{A},q_{B})}^{(j)} producing state:

χZEI(j)=i|βi(j)|2(zp(z|i)|zz|)|Ei,j~Ei,j~||ii|I.\chi^{(j)}_{ZEI}=\sum_{i}|\beta_{i}^{(j)}|^{2}\left(\sum_{z}p(z|i)\ket{z}\bra{z}\right)\otimes\ket{\widetilde{E_{i,j}}}\bra{\widetilde{E_{i,j}}}\otimes\ket{i}\bra{i}_{I}.

Then, using Equation 4 and the definition of min entropy, we conclude:

Hmin(AZ|E)χ(j)\displaystyle H_{\text{min}}(A_{Z}|E)_{\chi^{(j)}} Hmin(AZ|EI)χ(j)\displaystyle\geq H_{\text{min}}(A_{Z}|EI)_{\chi^{(j)}}
mini(logmaxzp(z|i))\displaystyle\geq\min_{i}(-\log\max_{z}p(z|i))
n(cb(qA)+δ)γ^+n(1cb(qA)δ)γ.\displaystyle\geq n(c_{b^{*}}(q_{A})+\delta)\hat{\gamma}+n(1-c_{b^{*}}(q_{A})-\delta)\gamma.

Finally, it is clear that:

|JqA,qB(j)|\displaystyle|J_{q_{A},q_{B}}^{(j)}| |{i𝒜dn | |ΔH(i,j)ΔH(qA,qB)|δ}\displaystyle\leq|\{i\in\mathcal{A}_{d}^{n}\text{ }|\text{ }|\Delta_{H}(i,j)-\Delta_{H}(q_{A},q_{B})|\leq\delta\}
=|{i𝒜dn | |ΔH(i,0)ΔH(qA,qB)|δ}\displaystyle=|\{i\in\mathcal{A}_{d}^{n}\text{ }|\text{ }|\Delta_{H}(i,0)-\Delta_{H}(q_{A},q_{B})|\leq\delta\}
dnH¯(ΔH(qA,qB)+δ),\displaystyle\leq d^{n\bar{H}(\Delta_{H}(q_{A},q_{B})+\delta)},

where the last inequality follows from the well-known bound on the volume of a Hamming sphere. Since the above analysis holds for any jj, we have therefore computed the resulting min entropy of the ideal case, namely for any chosen tt and observed qA,qBq_{A},q_{B}, it holds that:

Hmin(AZ|E)σn((cb(qA)+δ)γ^+(1cb(qA)δ)γH¯(ΔH(qA,qB)+δ)logd2)H_{\text{min}}(A_{Z}|E)_{\sigma}\geq n\left((c_{b^{*}}(q_{A})+\delta)\hat{\gamma}+(1-c_{b^{*}}(q_{A})-\delta)\gamma-\frac{\bar{H}(\Delta_{H}(q_{A},q_{B})+\delta)}{\log_{d}2}\right) (31)

The second step of the proof involves arguing that the smooth min entropy Hmin4ϵ+2ϵβ(AZ|E)ρH_{\text{min}}^{4\epsilon+2\epsilon^{\beta}}(A_{Z}|E)_{\rho}, for the given input state ρABE\rho_{ABE}, is bounded by the same quantity with high probability. This can be done in the same way as the second step in Theorem 3.1. Since the trace distance between the real and ideal states, for our chosen δ\delta, is no greater than ϵ\epsilon, the same error and smoothing bounds apply as in the second step in Theorem 3.1. thus completing the proof.

4.1 Application to QKD Security

Entropic uncertainty relations involving three parties, AA, BB, and EE have numerous applications, especially in quantum cryptography. Here we demonstrate how our bound produces improved finite-key rate bounds for the High-Dimensional BB84 protocol (HD-BB84) introduced in [13]. High-dimensional QKD protocols have been shown to exhibit several advantages over qubit based protocols in some scenarios, including in noise tolerance. For a general survey of QKD protocols, the reader is referred to [33, 34] while for a survey specific to high-dimensional QKD, the reader is referred to [16].

HD-BB84 involves two orthonormal bases, which we denote Z={|0,,|d1}Z=\{\ket{0},\cdots,\ket{d-1}\} and X={|x0,,|xd1}X=\{\ket{x_{0}},\cdots,\ket{x_{d-1}}\}, each of dimension dd; we will assume the bases are mutually unbiased and so |i|xj|=1/d|\braket{i}{x_{j}}|=1/\sqrt{d} for all i,ji,j. If we are considering lossy channels, then we will also add a |vac\ket{vac} vector to both these bases. Alice chooses a random basis and a random state within that basis (though not the |vac\ket{vac} state if it is there), sending it to BB. BB, on receipt of a quantum state will measure it in the ZZ or XX basis, choosing randomly. Afterwards, a classical authenticated communication channel is used allowing AA and BB to inform each other of their basis choices. If they are incompatible, the round is discarded; otherwise, assuming BB did not observe |vac\ket{vac}, they add logd\log d bits to their raw key. Repeating NN times, each AA and BB has a raw key of size nn bits. However, this key is only partially correlated (there may be errors due to natural noise or adversarial interference) and only partially secret. Thus, an Error Correction protocol is run (leaking additional information to the adversary) and, finally, Privacy Amplification (as discussed in Section 1.1), resulting in a secret key of size \ell bits. Maximizing \ell is vital to efficient performance of QKD systems and, from Equation 6, this involves maximizing our estimate of the min entropy Hminϵ(A|E)H_{\text{min}}^{\epsilon}(A|E).

To analyze this protocol, we consider an equivalent entanglement based version, parameterized by ZZ, XX, nn and mm. We also consider an asymmetric version whereby only ZZ basis measurements contribute to the raw key, while XX basis measurements are used only for estimating the error in the channel. The entanglement based HD-BB84 runs as follows:

  1. 1.

    An adversary prepares a quantum state |ψ0ABE\ket{\psi_{0}}\in\mathcal{H}_{A}\otimes\mathcal{H}_{B}\otimes\mathcal{H}_{E}, where ABdn+m\mathcal{H}_{A}\cong\mathcal{H}_{B}\cong\mathcal{H}_{d}^{\otimes n+m}. The AA portion is sent to Alice; the BB portion is sent to Bob; while Eve keeps the EE portion to herself.

  2. 2.

    AA chooses a random subset tt of size mm and sends it to BB; both parties measure their systems indexed by tt in the XX basis resulting in outcomes qAq_{A} and qBq_{B} respectively (these are strings in 𝒜dm\mathcal{A}_{d}^{m}). These values are disclosed to one another using the authenticated channel.

  3. 3.

    AA and BB measure the remaining portion of their systems in the ZZ basis resulting in their raw-keys rAr_{A} and rBr_{B} of size at most nn bits each (if there are |vac\ket{vac} observations, those will not contribute to the raw key and so it may be smaller than nn in a lossy channel).

  4. 4.

    AA and BB run an error correction protocol capable of correcting up to QQ errors in their raw keys, leaking leakEC\texttt{leak}_{\texttt{EC}} bits to Eve.

  5. 5.

    Finally, privacy amplification is run on the error corrected raw key resulting in their secret key.

Note that when d=2d=2 this is exactly the BB84 protocol. Note also that, by increasing the basis dimension to d+1d+1, we can add an additional “vacuum” state |vac\ket{vac} to both the ZZ and XX basis, such that i|vac=xi|vac=0\braket{i}{vac}=\braket{x_{i}}{vac}=0. In this case the maximal overlap function is γ^=log21=0\hat{\gamma}=-\log_{2}1=0 and the second maximal overlap function is γ=log21/d=log2d\gamma=-\log_{2}1/d=\log_{2}d. (Note that this shows the importance of our relation in being able to handle both cases individually.) Without this vacuum basis state, the dimension will be dd, and γ^=γ=log2d\hat{\gamma}=\gamma=\log_{2}d.

Using Equation 6 and results in [11, 35], if AA and BB wish to have an ϵPA\epsilon_{PA}-secure key, we have:

=Hminϵ(A|E)leakEC2log1ϵPA2ϵ.\ell=H_{\text{min}}^{\epsilon^{\prime}}(A|E)-\texttt{leak}_{\texttt{EC}}-2\log\frac{1}{\epsilon_{PA}-2\epsilon^{\prime}}.

Given ϵ>0\epsilon>0 and using our Theorem 4.1, setting ϵPA=4ϵβ+9ϵ\epsilon_{PA}=4\epsilon^{\beta}+9\epsilon, we have:

ourHDBB84=n(1pvacδ)(logdH¯(ΔH(qA,qB)+δ)logd+12)leakEC2log1ϵ\ell_{our-HD-BB84}=n(1-p_{vac}-\delta)\left(\log d-\frac{\bar{H}(\Delta_{H}(q_{A},q_{B})+\delta)}{\log_{d+1}2}\right)-\texttt{leak}_{\texttt{EC}}-2\log\frac{1}{\epsilon} (32)

where pvacp_{vac} is the number of counts in the observed qAq_{A} of the distinguished vacuum basis state (which is shared between both the ZZ and XX basis making γ^=0\hat{\gamma}=0). In particular, if the privacy amplification function is chosen to produce an output of size ours\ell_{ours}, it is guaranteed, except with probability at most 2ϵ12β2\epsilon^{1-2\beta}, that the secret key will be ϵPA\epsilon_{PA} secure according to Equation 6. Note that if we are not considering lossy channels, then the key-rate equation becomes simply:

ourHDBB84noloss=n(logdH¯(ΔH(qA,qB)+δ)logd2)leakEC2log1ϵ\ell_{our-HD-BB84-no-loss}=n\left(\log d-\frac{\bar{H}(\Delta_{H}(q_{A},q_{B})+\delta)}{\log_{d}2}\right)-\texttt{leak}_{\texttt{EC}}-2\log\frac{1}{\epsilon} (33)

To compare our new key-rate bound with prior work, we compare with results in [36] which is, to our knowledge, the current best bound for the HD-BB84 protocol in the finite key setting (with composable security, as is ours). Note that they used an entropic uncertainty relation from [37], resulting in a key-rate bound of:

priorHDBB84=n[log2dh(Q+ν)(Q+ν)log2(d1)],\ell_{prior-HD-BB84}=n[\log_{2}d-h(Q+\nu)-(Q+\nu)\log_{2}(d-1)], (34)

where:

ν=(n+m)(m+1)ln(2/ϵ)m2n.\nu=\sqrt{\frac{(n+m)(m+1)\ln(2/\epsilon)}{m^{2}n}}.

Where, for our evaluations, QQ is the error parameter of a depolarization channel. Note that this prior work could not handle an additional vacuum basis state in each of the ZZ and XX basis (if it were added, the bound from [37] would become the trivial one as the overlap function would be log21=0-\log_{2}1=0). So, when we evaluate, we will compare our bounds both without the vacuum basis then later by considering this basis state and loss in the channel.

In practice, the value of ΔH(qA,qB)\Delta_{H}(q_{A},q_{B}) or QQ is known and observed based on the actual channel used. However, to evaluate and compare our new key-rate bound we will evaluate assuming a depolarization channel with parameter QQ acting on each qudit independently and identically. Such a channel maps a quantum state ρ\rho to:

Q(ρ)=(1dd1Q)ρ+Qd1I.\mathcal{E}_{Q}(\rho)=\left(1-\frac{d}{d-1}\cdot Q\right)\rho+\frac{Q}{d-1}I.

Of course, our security proof does not require this depolarization assumption - instead, it is simply a channel we use to evaluate our bound and compare with prior work. It is also one of the most common noise models considered in theoretical QKD security proofs. For both protocols, we use leakEC=1.2H(A|B)\texttt{leak}_{\texttt{EC}}=1.2H(A|B) which, for this depolarization channel, is easily found to be H(A|B)=Qlog(d1)+h(Q)H(A|B)=Q\log(d-1)+h(Q).

Refer to caption
Refer to caption
Figure 6: Showing the secret key generation rates (/N\ell/N) of the HD-BB84 protocol when dimension d=22d=2^{2} assuming a depolarization channel with parameter Q=10%Q=10\%. Here, the xx-axis is the total number of qudits NN from which we use m=.07Nm=.07N for sampling. Left and Right are different ranges in the number of signals. Dashed black line (top most in both graphs) is the theoretical asymptotic rate (Equation 35); Solid blue line is our key-rate bound using our new entropic uncertainty relation, namely ourHDBB84noloss/N\ell_{our-HD-BB84-no-loss}/N (Equation 33) for pvac=0p_{vac}=0 (no loss); Dashed red line is the previous best known bound for the HD-BB84 key rate using alternative methods to compute EE’s uncertainty, priorHDBB84/N\ell_{prior-HD-BB84}/N (Equation 34) with no loss (loss is not supported in that prior work); Finally, solid-red line (lowest) is our key-rate bound when pvac=20%p_{vac}=20\% (i.e., a 20%20\% loss in the channel) using Equation 32. For our key-rate evaluation, we use β=1/3\beta=1/3 and ϵ=1036\epsilon=10^{-36} giving a failure probability and a value of ϵPA\epsilon_{PA} both on the order of 101210^{-12}. For Equation 34, we use a failure probability of 101210^{-12}. For both finite key results, we use leakEC=1.2H(A|B)\texttt{leak}_{\texttt{EC}}=1.2H(A|B) which, in the case of a depolarization channel, is leakEC=1.2(Qlog(d1)+h(Q))\texttt{leak}_{\texttt{EC}}=1.2(Q\log(d-1)+h(Q)). For the theoretical upper-bound we use the leakEC=H(A|B)\texttt{leak}_{\texttt{EC}}=H(A|B) (without the additional 1.21.2 scaling factor).
Refer to caption
Refer to caption
Figure 7: Similar to Figure 6 but now showing the secret key generation rates (/N\ell/N) of the HD-BB84 protocol when d=210d=2^{10}. Here the depolarization noise is Q=10%Q=10\%. Dashed black line (top most in both graphs) is the theoretical asymptotic rate (Equation 35); Solid blue line is our key-rate bound with no loss (pvac=0p_{vac}=0); Dashed red line is the previous best known bound for the HD-BB84 key rate (with no loss); finally, solid red line (lowest) is our key-rate bound when pvac=50%p_{vac}=50\%.

Finally, we compare to the theoretical, asymptotic upper-bound using the entropic uncertainty relation of [38]. This disregards all finite-key effects (such as failure probabilities and sampling imprecision), and takes the number of signals NN\rightarrow\infty. This bound works out easily to be:

rasym=logd2H(A|B)=logd2(Qlog(d1)+h(Q)),r_{asym}=\log d-2H(A|B)=\log d-2(Q\log(d-1)+h(Q)), (35)

where again we used the easily verified fact that, for a depolarization channel with parameter QQ, H(A|B)=Qlog(d1)+h(Q)H(A|B)=Q\log(d-1)+h(Q) and, furthermore, we assume perfect error correction whereby leakEC=H(A|B)\texttt{leak}_{\texttt{EC}}=H(A|B).

Comparisons of both our new bound and prior work are shown in Figure 6 (for d=22d=2^{2} dimensions) and Figure 7 (for dimension d=210d=2^{10}). We note that when pvac=0p_{vac}=0, our bound is only slightly lower than Equation 34 and this difference decreases as the number of signals increases. Indeed, the difference turns out to be only that our confidence interval, determined by δ\delta is slightly larger for any particular ϵ\epsilon making our results asymptotically the same, though slightly lower than prior work for this case. However, one of the powers of our new relation is its ability to also handle two overlap functions allowing us to incorporate loss in both ZZ and XX bases. Of course, as the loss increases, the key-rate decreases as expected; our new entropic uncertainty relation can, however, easily handle this scenario. Further refinements to the classical sampling strategy used, may further improve our bound (in both the lossy and loss-less case). Indeed our analysis of Lemma 2.3 is not necessarily tight. Alternative sampling strategies or improved analyses, may be easily incorporated through our methods.

5 Closing Remarks

The quantum sampling framework of Bouman and Fehr, introduced in [1], provides a promising new tool to develop results in general quantum information theory and quantum cryptography. In our prior work [2, 3], we used this framework to introduce so-called sampling-based entropic uncertainty relations. In this paper, we showed how quantum sampling can be used to develop very general quantum entropic uncertainty relations allowing one to insert arbitrary classical sampling strategies, perhaps defined for a specific cryptographic task, which may then be “promoted” to analyze results for quantum systems. Furthermore, we developed an entirely new three-party entropic uncertainty relation using the sampling framework as a foundation, which has applications to high-dimensional QKD as we demonstrated here. Our new relation can also handle two different measurement overlaps, allowing one to work with bases that share common vectors (such as a “vacuum” measurement outcome). Since our relation handles all finite sampling precision, they provide an easy and general purpose framework for other researchers to develop finite-key cryptographic security proofs.

Several interesting future problems remain open. So far we only considered projective basis measurements. Generalizing these results to arbitrary POVM’s would be greatly interesting. However, this would require extending the quantum sampling technique to support such measurements. Furthermore, improving the three-party relation with a tighter sampling strategy would produce even more beneficial results. Finding other interesting theoretical and cryptographic applications of quantum sampling and our sampling-based entropic uncertainty relations would also be highly interesting. We feel that the framework of quantum sampling is powerful and can be employed successfully in other areas of quantum information science, and further exploration of quantum sampling in the domain of quantum information theory can yield even more exciting results in quantum cryptography.

References

  • [1] Niek J Bouman and Serge Fehr. Sampling in a quantum population, and applications. In Annual Cryptology Conference, pages 724–741. Springer, 2010.
  • [2] Walter O Krawec. Quantum sampling and entropic uncertainty. Quantum Information Processing, 18(12):368, 2019.
  • [3] Walter O Krawec. A new high-dimensional quantum entropic uncertainty relation with applications. In IEEE International Symposium on Information Theory, ISIT 2020, pages 1978–1983. IEEE, 2020.
  • [4] Hans Maassen and Jos BM Uffink. Generalized entropic uncertainty relations. Physical Review Letters, 60(12):1103, 1988.
  • [5] K. Kraus. Complementary observables and uncertainty relations. Phys. Rev. D, 35:3070–3075, May 1987.
  • [6] David Deutsch. Uncertainty in quantum measurements. Phys. Rev. Lett., 50:631–633, Feb 1983.
  • [7] Iwo Bialynicki-Birula and Łukasz Rudnicki. Entropic uncertainty relations in quantum physics. In Statistical Complexity, pages 1–34. Springer, 2011.
  • [8] Patrick J. Coles, Mario Berta, Marco Tomamichel, and Stephanie Wehner. Entropic uncertainty relations and their applications. Rev. Mod. Phys., 89:015002, Feb 2017.
  • [9] Stephanie Wehner and Andreas Winter. Entropic uncertainty relations—a survey. New Journal of Physics, 12(2):025009, 2010.
  • [10] Giuseppe Vallone, Davide G Marangon, Marco Tomasin, and Paolo Villoresi. Quantum randomness certified by the uncertainty principle. Physical Review A, 90(5):052327, 2014.
  • [11] Renato Renner. Security of quantum key distribution. International Journal of Quantum Information, 6(01):1–127, 2008.
  • [12] Antonio Acin, Nicolas Gisin, and Valerio Scarani. Security bounds in quantum cryptography using d-level systems. arXiv preprint quant-ph/0303009, 2003.
  • [13] Nicolas J Cerf, Mohamed Bourennane, Anders Karlsson, and Nicolas Gisin. Security of quantum key distribution using d-level systems. Physical review letters, 88(12):127902, 2002.
  • [14] Georgios M Nikolopoulos and Gernot Alber. Security bound of two-basis quantum-key-distribution protocols using qudits. Physical Review A, 72(3):032320, 2005.
  • [15] Georgios M Nikolopoulos, Kedar S Ranade, and Gernot Alber. Error tolerance of two-basis quantum-key-distribution protocols using qudits and two-way classical communication. Physical Review A, 73(3):032325, 2006.
  • [16] Daniele Cozzolino, Beatrice Da Lio, Davide Bacco, and Leif Katsuo Oxenløwe. High-dimensional quantum communication: Benefits, progress, and future challenges. Advanced Quantum Technologies, 2(12):1900038, 2019.
  • [17] Jian-Yu Guan, Zhu Cao, Yang Liu, Guo-Liang Shen-Tu, Jason S Pelc, MM Fejer, Cheng-Zhi Peng, Xiongfeng Ma, Qiang Zhang, and Jian-Wei Pan. Experimental passive round-robin differential phase-shift quantum key distribution. Physical review letters, 114(18):180502, 2015.
  • [18] Hiroki Takesue, Toshihiko Sasaki, Kiyoshi Tamaki, and Masato Koashi. Experimental quantum key distribution without monitoring signal disturbance. Nature Photonics, 9(12):827, 2015.
  • [19] Shuang Wang, Zhen-Qiang Yin, HF Chau, Wei Chen, Chao Wang, Guang-Can Guo, and Zheng-Fu Han. Proof-of-principle experimental realization of a qubit-like qudit-based quantum key distribution scheme. Quantum Science and Technology, 3(2):025006, 2018.
  • [20] Nurul T Islam, Clinton Cahall, Andrés Aragoneses, A Lezama, Jungsang Kim, and Daniel J Gauthier. Robust and stable delay interferometers with application to d-dimensional time-frequency quantum key distribution. Physical Review Applied, 7(4):044010, 2017.
  • [21] Robert Konig, Renato Renner, and Christian Schaffner. The operational meaning of min-and max-entropy. IEEE Transactions on Information theory, 55(9):4337–4347, 2009.
  • [22] Roger Colbeck and Adrian Kent. Private randomness expansion with untrusted devices. Journal of Physics A: Mathematical and Theoretical, 44(9):095305, 2011.
  • [23] Stefano Pironio and Serge Massar. Security of practical private randomness generation. Physical Review A, 87(1):012336, 2013.
  • [24] Peter Bierhorst, Emanuel Knill, Scott Glancy, Yanbao Zhang, Alan Mink, Stephen Jordan, Andrea Rommal, Yi-Kai Liu, Bradley Christensen, Sae Woo Nam, et al. Experimentally generated randomness certified by the impossibility of superluminal signals. Nature, 556(7700):223–226, 2018.
  • [25] Yang Liu, Xiao Yuan, Ming-Han Li, Weijun Zhang, Qi Zhao, Jiaqiang Zhong, Yuan Cao, Yu-Huai Li, Luo-Kan Chen, Hao Li, et al. High-speed device-independent quantum random number generation without a detection loophole. Physical review letters, 120(1):010503, 2018.
  • [26] Jing-Yan Haw, SM Assad, AM Lance, NHY Ng, V Sharma, Ping Koy Lam, and Thomas Symul. Maximization of extractable randomness in a quantum random-number generator. Physical Review Applied, 3(5):054004, 2015.
  • [27] Bingjie Xu, Ziyang Chen, Zhengyu Li, Jie Yang, Qi Su, Wei Huang, Yichen Zhang, and Hong Guo. High speed continuous variable source-independent quantum random number generation. Quantum Science and Technology, 4(2):025013, 2019.
  • [28] Marco Avesani, Davide G Marangon, Giuseppe Vallone, and Paolo Villoresi. Secure heterodyne-based quantum random number generator at 17 gbps. arXiv preprint arXiv:1801.04139, 2018.
  • [29] Yu-Huai Li, Xuan Han, Yuan Cao, Xiao Yuan, Zheng-Ping Li, Jian-Yu Guan, Juan Yin, Qiang Zhang, Xiongfeng Ma, Cheng-Zhi Peng, et al. Quantum random number generation with uncharacterized laser and sunlight. npj Quantum Information, 5(1):1–5, 2019.
  • [30] Miguel Herrero-Collantes and Juan Carlos Garcia-Escartin. Quantum random number generators. Reviews of Modern Physics, 89(1):015004, 2017.
  • [31] Daniela Frauchiger, Renato Renner, and Matthias Troyer. True randomness from realistic quantum devices. arXiv preprint arXiv:1311.4547, 2013.
  • [32] Feihu Xu, Jeffrey H Shapiro, and Franco NC Wong. Experimental fast quantum random number generation using high-dimensional entanglement with entropy monitoring. Optica, 3(11):1266–1269, 2016.
  • [33] Stefano Pirandola, Ulrik L Andersen, Leonardo Banchi, Mario Berta, Darius Bunandar, Roger Colbeck, Dirk Englund, Tobias Gehring, Cosmo Lupo, Carlo Ottaviani, et al. Advances in quantum cryptography. arXiv preprint arXiv:1906.01645, 2019.
  • [34] Valerio Scarani, Helle Bechmann-Pasquinucci, Nicolas J. Cerf, Miloslav Dušek, Norbert Lütkenhaus, and Momtchil Peev. The security of practical quantum key distribution. Rev. Mod. Phys., 81:1301–1350, Sep 2009.
  • [35] Marco Tomamichel, Charles Ci Wen Lim, Nicolas Gisin, and Renato Renner. Tight finite-key analysis for quantum cryptography. Nature communications, 3(1):1–6, 2012.
  • [36] Kamil Brádler, Mohammad Mirhosseini, Robert Fickler, Anne Broadbent, and Robert Boyd. Finite-key security analysis for multilevel quantum key distribution. New Journal of Physics, 18(7):073030, 2016.
  • [37] Marco Tomamichel and Renato Renner. Uncertainty relation for smooth entropies. Physical review letters, 106(11):110506, 2011.
  • [38] Mario Berta, Matthias Christandl, Roger Colbeck, Joseph M Renes, and Renato Renner. The uncertainty principle in the presence of quantum memory. Nature Physics, 6(9):659–662, 2010.