• Aucun résultat trouvé

Self-testing quantum states and measurements in the prepare-and-measure scenario

N/A
N/A
Protected

Academic year: 2022

Partager "Self-testing quantum states and measurements in the prepare-and-measure scenario"

Copied!
17
0
0

Texte intégral

(1)

Article

Reference

Self-testing quantum states and measurements in the prepare-and-measure scenario

TAVAKOLI, Armin, et al.

Abstract

The goal of self-testing is to characterize an a priori unknown quantum system based solely on measurement statistics, i.e. using an uncharacterized measurement device. Here we develop self-testing methods for quantum prepare-and-measure experiments, thus not necessarily relying on entanglement and/or violation of a Bell inequality. We present noise-robust techniques for self-testing sets of quantum states and measurements, assuming an upper bound on the Hilbert space dimension. We discuss in detail the case of a 2→1 random access code with qubits, for which we provide analytically optimal self-tests. The simplicity and noise robustness of our methods should make them directly applicable to experiments.

TAVAKOLI, Armin, et al . Self-testing quantum states and measurements in the prepare-and-measure scenario. Physical Review. A , 2018, vol. 98, no. 062307

DOI : 10.1103/PhysRevA.98.062307

Available at:

http://archive-ouverte.unige.ch/unige:127273

Disclaimer: layout of this document may differ from the published version.

1 / 1

(2)

Armin Tavakoli,1J˛edrzej Kaniewski,2Tamás Vértesi,3 Denis Rosset,4, 5and Nicolas Brunner1

1Département de Physique Appliquée, Université de Genève, CH-1211 Genève, Switzerland

2QMATH, Department of Mathematical Sciences, University of Copenhagen, Universitetsparken 5, 2100 Copenhagen, Denmark

3Institute for Nuclear Research, Hungarian Academy of Sciences, P.O. Box 51, 4001 Debrecen, Hungary

4Perimeter Institute for Theoretical Physics, 31 Caroline St. N, Waterloo, Ontario, Canada, N2L 2Y5

5Institute for Quantum Optics and Quantum Information (IQOQI), Boltzmangasse 3, 1090 Vienna, Austria

The goal of self-testing is to characterise an a priori unknown quantum system based solely on measurement statistics, i.e. using an uncharacterised measurement device. Here we develop self-testing methods for quantum prepare-and-measure experiments, thus not necessarily relying on entanglement and/or violation of a Bell in- equality. We present noise-robust techniques for self-testing sets of quantum states and measurements, assuming an upper bound on the Hilbert space dimension. We discuss in detail the case of a2→1random access code with qubits, for which we provide analytically optimal self-tests. The simplicity and noise robustness of our methods should make them directly applicable to experiments.

Introduction.—Predicting the results of measurements per- formed on a given physical system has traditionally been the main concern of physics. However, with the advent of device- independent quantum information processing [1–3], the op- posite question has become relevant. More specifically, given an initially unknown system and an uncharacterised measure- ment device, what can be inferred about the physics of the experiment based solely on the observed measurement statis- tics? Despite the apparent generality of this question, certain cases do allow for a precise characterization of the system.

This is referred to as self-testing [4,5].

The possibility to self-test quantum states and measure- ments usually relies on quantum nonlocality. Consider two distant observers performing local measurements on a shared quantum state. When the resulting statistics leads to viola- tion of a Bell inequality [6], it is necessarily the case that the shared quantum state is entangled, and moreover, that the local quantum measurements are incompatible; see e.g. [7].

Furthermore, for specific Bell inequalities, maximal violation (i.e. the largest possible value in quantum theory) implies that the quantum state and the measurements can be uniquely iden- tified (up to local isometries). For instance, a maximal vio- lation of the Clauser-Horne-Shimony-Holt (CHSH) Bell in- equality [8] implies maximally incompatible measurements (two anti-commuting Pauli observables) and a shared maxi- mally entangled two-qubit state [9–12]. More recently, it has been demonstrated that all bipartite pure entangled states can be self-tested [13], as well as certain multipartite entangled states [14–16]. Another important progress is the develop- ment of self-testing methods robust to noise [17–23]. For in- stance, given a certain level of violation of a Bell inequality (but not necessarily maximal), the fidelity between the ini- tially unknown state and a given target state can be lower- bounded.

Self-testing thus offers promising perspectives for the cer- tification of quantum systems in experiments (see e.g. [24]), as well as for device-independent quantum information pro- tocols [25]. It is therefore natural to ask whether the concept of self-testing can be applied to more general quantum exper- iments, beyond those based on entanglement and nonlocality.

In the present work, we develop self-testing methods tai- lored to the prepare-and-measure scenario. This covers a

broad class of experiments, where quantum communication schemes (e.g. the BB84 quantum key distribution (QKD) pro- tocol) are prominent examples. In this setting, a preparation device initially prepares a quantum system in different possi- ble states. The system is then transmitted to a measurement device, which performs different possible measurements on it. While it is still possible in this case to characterise certain physical properties of the system based only on statistics, this requires in general an assumption on the devices. One pos- sibility, which we will follow here, is to assume that the set of quantum states and measurements admit a full description in a Hilbert space of given dimension [26–28]. Intuitively this means that the amount of information communicated from the preparation device to the measurement device is assumed to be upper bounded. Such a scenario considering quantum systems of fixed dimension, but otherwise uncharacterised, is referred to as semi-device-independent, and opens interesting possibil- ities for quantum information processing [29–33].

Here we demonstrate techniques for robustly self-testing sets of prepared quantum states, as well as sets of quantum measurements. These methods allow one to (i) assess the compatibility of given sets of preparations and measurements with the observed statistics, and (ii) lower-bound the average fidelity between the unknown preparations (measurements) and a set of ideal quantum states (measurements). We discuss in detail a simple prepare-and-measure scenario, namely the 2→1random access code (RAC). This allows us to provide analytically optimal self-tests for a pair of anti-commuting Pauli observables, and for a set of four qubit states corre- sponding to the eigenstates of two anti-commuting Pauli ob- servables. We then generalise these results to other prepare- and-measure scenarios. The simplicity and robustness of our methods should make them directly applicable to experi- ments. We conclude with a number of open questions.

Scenario.— We consider a quantum prepare-and-measure experiment. Upon receiving input x, a preparation device emits a physical system in a quantum stateρx. The system is then transmitted to a measurement device, which, upon receiv- ing an inputy, performs a quantum measurement returning an outcomeb. Formally, the measurement is described by a set of positive operatorsMyb, that equal identity when summed over b. Importantly both the specific statesρxand measurements

arXiv:1801.08520v2 [quant-ph] 11 Sep 2018

(3)

Mybare a priori unknown to the observer. The statistics of the experiment is then given byP(b|x, y) = tr(ρxMyb). In this setting, any possible probability distribution can be obtained, given that the prepared statesρxcan be taken in a sufficiently large Hilbert space. This is however no longer the case when we limit the Hilbert space dimension; specifically we impose thatρx ∈ L(Cd)for some givend <|x|(where|x|denotes the number of possible inputsx). In this case, limits on the set of possible distributions can be captured via inequalities of the form

A=X

x,y,b

αxybP(b|x, y)≤Qd, (1) where αxyb are real coefficients. These “dimension wit- nesses” allow one to place device-independent lower bounds on the dimension of the quantum system [26].

Subsequently, one can ask what the limitations are on the set of distributionsP(b|x, y)given that the preparations ad- mit a classicald-dimensional representation, i.e. there exists ad-dimensional basis such that all states ρxare diagonal in this basis. We denote byCdthe maximal value of the quan- tityA in this case. Interestingly, for well-chosen quantities A, one finds thatCd < Qd. Thus, for a given system dimen- sion d, quantum systems outperform classical ones, in the sense that certain quantum distributions cannot be reproduced classically [26]. This quantum advantage can be viewed as the origin for the possibility of developing self-testing meth- ods for the prepare-and-measure scenario; in analogy to Bell inequality violation being the root for self-testing entangled states.

In the following we present robust self-testing techniques based on specific dimension witnesses A. Based only on the value ofA, which is directly accessible from the exper- iment statistics, we characterise the (initially unknown) pre- pared states and measurements. In particular, when the max- imal value of the witness is obtained, i.e. A = Qd, then a specific set of pure statesρx = |ψxihψx|and a specific set of projective measurementsMyb must have been used (up to a unitary). Moreover, when a non-maximal valueA< Qdis obtained, the compatibility of given sets of preparations and measurements can be assessed. Finally, one can efficiently lower bound the fidelity between the prepared states and mea- surements and the ideal (or target) states and measurement leading toA=Qd.

Note that a recent series of works followed a related though conceptually different approach, based on hypothesis testing [34–36]. This method does however not allow for self-testing.

The2 → 1random access code.—We discuss in detail a simple prepare-and-measure experiment. This involves four possible preparations, denoted byx= (x0, x1)(wherexj ∈ {0,1}), and two possible binary measurements, y ∈ {0,1}

andb∈ {0,1}. The score is given by A2= 1

8 X

x0,x1,y

P(b=xy|x0, x1, y). (2) This means that, upon receiving inputy, the measurement de- vice should return the outputb =xy, i.e. they-th bit of the

input bit-stringxreceived by the preparation device. Hence the name of a2 → 1RAC [37–39]. Note that all inputs are assumed to be chosen uniformly at random. Indeed, this task is nontrivial only whend <4; here we will consider the case d = 2, i.e. qubits. In this case, one finds the tight bounds C2 = 3/4andQ2 = (1 + 1/√

2)/2 ≈0.85[37]. The clas- sical boundC2can be obtained by simply always sending the bitx0. The quantum boundQ2is obtained via the following

“ideal” strategy. The four qubit preparations correspond to the pure states

ρideal00 =11+σx

2 ρideal01 = 11+σz

2 ρideal11 =11−σx

2 ρideal10 = 11−σz

2 . (3)

These are simply the eigenstates of the Pauli observablesσx

andσz. Next, the measurements are projective and given by two anti-commuting Pauli observables

Myideal= (My0)ideal−(My1)ideal= σx+ (−1)yσz

√2 . (4) These qubit preparations and measurements represent the ideal situation, where the maximal value A2 = Q2 is achieved. In the following we will determine what restric- tions apply to the possible preparations and measurements, given that a particular value ofA2is observed. In particular, when the maximal valueA2 =Q2is attained, both the states and the measurements must be the ideal ones as given above (up to a unitary).

Self-testing preparations.—Here we find restrictions on the set of prepared states given an observed value of A2. For convenience, we write the qubit preparations as ρx0x1 = (11+m~x0x1·~σ)/2, where m~x0x1 denotes the Bloch vector (satisfying|m~x0x1| ≤ 1) and~σ = (σx, σy, σz)denotes the vector of Pauli matrices.

The first step consists in re-expressing A2= 1

2+1 8

X

y

tr My0Vy

≤ 1 2+1

8 X

y

q

tr My0Vy2

tr My0

(5) whereVy = P

x0,x1(−1)xyρx0x1. In the second step we used that for a positive semi-definiteO and a Hermitian op- eratorR, it holds that |tr (OR)|2 ≤ tr OR2

tr (O) [23].

Without loss of generality, we can restrict to extremal qubit measurements, which are here projective rank-one operators.

Consequently, we have thattr My0

= 1. Next, we obtain Vy2= 12(β+ (−1)yα)11, where

β= 1 2

X

x0,x1

|m~x0x1|2−m~00·m~11−m~01·m~10 (6) α= (~m00−m~11)·(~m01−m~10). (7) Finally, we find that Eq. (5) reduces to

A2≤1 2 + 1

8√ 2

hpβ+α+p β−αi

. (8)

(4)

This provides a tight self-test of the prepared states (in terms of their Bloch vectors), for any given value ofA2. Let us start with the caseA2 =Q2. Since√

β+α+√

β−α= q

2β+ 2p

β2−α2, we see that Eq. (8) is maximal iffα= 0 andβ is maximal. This turns out to be achievable. In order to maximiseβ, we need (i)∀x0x1 : |~mx0x1| = 1, i.e. that all preparations are pure states, and (ii) that m~00 ·m~11 =

~

m01·m~10 =−1, i.e. the states correspond to (pairwise) an- tipodal Bloch vectors. We define~r0 = m~00 = −~m11 and

~

r1 = m~01 = −~m10. Consequently, we find α = 4~r0·~r1. Therefore, in order to haveα= 0, we must choose~r0·~r1= 0.

This implies that the right-hand-side of Eq. (8) is upper- bounded byQ2. Therefore, we conclude that when observing maximal valueA2=Q2, the set of four prepared states must be equivalent (up to a unitary rotation) to the set of four ideal states; we note that this was also shown in Ref. [40] in the context of QKD.

More generally, for any value A2, one can find a set of preparations (and corresponding measurements) such that the inequality (8) is saturated, see Supplementary Material (SM) part A. For the case of classical preparations (i.e. diagonal in a given basis), the Bloch vectors can simply be replaced by numbersmx0x1∈[−1,1], and we getA2≤C2.

Self-testing measurements.— Let us now consider self- testing of measurements. Using thatMy = My0−My1, we write

A2= 1 2+ 1

16 X

x0,x1

tr ρx0x1[(−1)x0M0+ (−1)x1M1]

≤ 1 2+ 1

16 X

x0,x1

λmax[(−1)x0M0+ (−1)x1M1], (9) whereλmax[X]is the largest eigenvalue of the (Hermitian) op- eratorX. Since the upper bound corresponds to choosing the optimal preparations for a fixed pair of observables, it sim- ply quantifies the optimal performance achievable using these observables. IfM0andM1are qubit observables the upper bound can be evaluated exactly (see SM part A) to give

A2≤ 1 2+ 1

16 q

2µ+ 2ν−η2++ q

2µ−2ν−η2

, (10) whereµ = tr M02+M12

, ν = tr{M0, M1} andη± = tr(M0±M1). The right-hand side reaches the optimal value Q2 iff µ = 4, η± = 0 and ν = 0, which implies anti- commuting projective observables (i.e. projective measure- ment operators). In other words, observingA2=Q2implies that the measurements are unitarily equivalent to the ideal ones. Moreover, note that inequality (10) is tight; for any value ofA2 one can find measurements (and corresponding states) such that inequality is saturated (see SM part A). It fol- lows that any pair of projective, rank-one observables that is incompatible (|ν|<4) can lead toA2> C2.

Robust self-testing of the preparations.—We now discuss the problem of characterizing the fidelity between the realised preparations and the ideal ones. This will allow us to quantify the distance of the prepared states with respect to the ideal

ones. Again, we want to develop self-testing methods which are based only on the value ofA2.

More formally, given an arbitrary set of preparations, we define the average fidelity with the ideal preparations to be S({ρx0x1}) = maxΛP

x0,x1F(ρidealx

0x1,Λ[ρx0x1])/4, where Λ is a quantum channel i.e. a completely positive trace- preserving map. Here the fidelitiesF(ρ, σ) = tr p√

ρσ√ ρ simplify toF(ρidealx0x1,Λ[ρx0x1]) = tr Λ[ρx0x1idealx0x1

, as the ρidealx0x1are pure states. We derive lower bounds on the smallest possible value ofSgiven a value ofA2, i.e.,

F(A2) = min

x0x1}∈R(A2)

S[{ρx0x1}]. (11) Note that this involves a minimisation over all sets of four preparations R(A2) that are compatible with an observed valueA2.

In order to lower bound F, we use an approach in- spired by Ref. [22]. From Eq. (9), we have A2 = 12 + P

x0,x1tr (Wx0x1ρx0x1), whereWx0x1 = 161 P

y(−1)xyMy. We define operators corresponding to some suitably chosen channel acting on the ideal preparations:

Kx0x1(M0, M1) = Λ(M0, M1)[ρidealx

0x1], (12) whereΛ is the channel dual toΛ. We aim to construct oper- ator inequalities of the form

Kx0x1(M0, M1)≥sWx0x1+tx0x1(M0, M1)11, (13) for all inputs(x0, x1), for any given measurements, where sandtx0x1(M0, M1)are real coefficients. Finding such in- equalities, as well as a suitable channelΛallows us to lower boundSas follows:

S ≥1 4

X

x0,x1

tr (Kx0x1ρx0x1)≥ s 4

X

x0,x1

tr (Wx0x1ρx0x1)

+1 4

X

x0,x1

tx0x1 = s

4(A2−1/2) + 1 4

X

x0,x1

tx0x1, (14) Applying a minimisation overM0andM1to the right-hand- side, the above inequality becomes valid for all preparations.

Consequently, F(A2)≥ s

4(A2−1/2) +t≡L(A2), (15) wheret≡1/4 minM0,M1P

x0,x1tx0x1(M0, M1). In SM part B, we construct explicitly the channel and derive an operator inequality leading to a lower bound, given bys= 4 1 +√

2 andt= 2−√

2 /4.

This provides a robust self-testing for the preparations. A maximal valueA2 = Q2 implies F = 1, i.e. the prepara- tions must be the ideal ones (up to a unitary). ForA2 =C2, i.e. a maximal value given a set of classical states, we get that F ≥ 3/4. This bound can be attained via the set of pure statesρx0x1 = (11+ (−1)x0x1σz)/2 (diagonal in the same basis, hence classical), combined with the measure- mentsM0 = M1 = σz. Therefore, we see that our bound

(5)

FIG. 1. Average fidelityF(F0)for prepared states (measurements), as a function of the observed value ofA2. The black line is our analytical lower bound of Eq. (15). The blue region is accessible via single qubit strategies without shared randomness, as confirmed by strong numerical evidence (see SM). When allowing for shared randomness between the devices, the accessible region (obtained by taking the convex hull of the blue region) now also includes the grey area, and our analytic lower bound is tight in general.

F(A2) ≥L(A2)is optimal, as far as linear inequalities are concerned (see Fig. 1). It is then interesting to consider the intermediate regionC2<A2< Q2. First, focusing on strate- gies involving a single set of states and measurements, we observe numerically that the linear boundF(A2)≥ L(A2) cannot be saturated anymore, and conjecture the form of op- timal states and measurements; see red curve in Fig. 1 and SM part C for details. Second, allowing for shared random- ness between the preparation and measurement device (such that convex combinations of qubit strategies are now possi- ble), the linear bound becomes tight, a direct consequence of the linearity ofFandA2in terms of the states and measure- ments.

Robust self-testing of the measurements.— Similarly, we can quantify the average fidelity of the measure- ments with respect to the ideal ones: S0 {Myb}

= maxΛP

y,bF (Myb)ideal,Λ[Myb]

/4, whereΛmust be a uni- tal channel (i.e. mapping the identity to itself), in order to ensure that measurements are mapped to measurements. In analogy with the case of preparations, our goal is to lower bound the following quantity:

F0(A2) = min

{Myb}∈R0(A2)

S0 {Myb}

, (16)

whereR0(A2)represents all sets of measurements compatible with a certain value ofA2.

We first rewrite A2 = P

y,btr(MybZyb), where Zyb =

1 8

P

x0,x1ρx0x1δb,xy. Next, we construct operator inequalities Kyb({ρx0x1})≥sZyb+ty({ρx0x1})11, (17) given the unital channelKyb = Λ[(Myb)ideal]. Similarly to the case of preparations, strong operator inequalities can be derived by choosing carefully the channel; all details are given in SM part D. Finally, this leads to a lower bound on the aver-

age fidelity

F0(A2)≥ min

x0x1}

1 4

X

y,b

tr KybMyb

≥L(A2). (18)

That is, we find thatF0can be lower-bounded by a linear ex- pression in terms ofA2, which turns out to be the same as for the case of preparations.

This provides a robust self-test for the measurements. Ob- servingA2 = Q2implies that F0 = 1, hence the measure- ments are equivalent to the ideal ones (up to a unitary). For A2 =C2, we have thatF ≥3/4. This lower bound can be attained by choosingM0z andM1 = 11, with the states ρ0001 = (11+σz)/2andρ1011= (11−σz)/2. For C2 <A2 < Q2, we find numerically that the inequality (18) cannot be saturated using a single set of measurements and states (see Fig. 1). Details, in particular a conjecture for the form of the optimal measurements, are given in SM part C.

Similarly as for the case of states, when allowing for convex combinations of qubit strategies, our linear bound is tight.

Generalisations.—The above results can be generalised in several directions. Firstly, a generalisation of the2→1RAC enables self-testing of any pair of incompatible Pauli observ- ables (see SM part E). Secondly, we derive compatibility rela- tions for theN →1RAC with qubits (see SM part F). Thirdly, we self-test qutrit preparations and projective measurements in the2→1RAC (see SM part G).

Finally, we present a numerical method for robust self- testing of preparations applicable in scenarios beyond RACs.

The method is based on semi-definite programing and com- bines (i) the swap-method [21] used for self-testing in Bell scenarios with (ii) the hierarchy of finite-dimensional quan- tum correlations [41–43]. The idea is to first construct a swap operator, based on the measurement operators, which maps the state of the preparation onto an ancilla. The average fi- delity between the ancilla and the ideal states can then be ex- pressed in terms of strings of products of measurements oper- ators and the extracted states. The last step is to miminize this average fidelity over all quantum realisations that are compat- ible with a given witness value, using the hierarchy of Refs [41–43]. Although typically returning sub-optimal bounds on F, this method is widely applicable. In SM part H, we de- scribe in detail the methodology and apply to two examples, including the2→1RAC.

Outlook.— We presented methods for self-testing quan- tum states and measurements in the prepare-and-measure sce- nario. These techniques demonstrate strong robustness to noise, and should therefore be directly amenable to exper- iments, providing useful certification techniques in a semi- device-independent setting. Moreover, these ideas should find applications in quantum communications. Our methods apply to the states and measurements used in QKD (e.g. in BB84), as well as in semi-device-independent QKD and randomness generation protocols [29–33].

It would be interesting to develop robust self-testing tech- niques for more general scenarios, e.g. for higher dimensional quantum systems. Another direction would be to consider sce- narios beyond prepare-and-measure, for instance adding be-

(6)

tween the preparation and measurement devices a transforma- tion device [44,45], and self-test the latter.

Finally, while we have focused here on self-testing based on an assumption on the dimension, one could develop methods based on different assumptions, such as a bound on the mean energy [46], the overlap [47], or the entropy [48].

Acknowledgements.— This work was supported by the Swiss national science foundation (Starting grant DIAQ, QSIT, and Early Postdoc Mobility fellowship

P2GEP2_162060), the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie Action ROSETTA (grant no. 749316), the European Research Council (grant no. 337603), the Danish Council for Independent Research (Sapere Aude), VILLUM FONDEN via the QMATH Centre of Excellence (grant no. 10059), and the National Research, Development and Innovation Office NKFIH (Grants No. K111734, and No.

KH125096).

[1] A. Acín, N. Brunner, N. Gisin, S. Massar, S. Pironio, and V.

Scarani. Phys. Rev. Lett.,98, 23 (2007).

[2] R. Colbeck, PhD thesis, University of Cambridge, 2007. ArXiv:

0911.3814.

[3] S. Pironio, A. Acín, S. Massar, A. Boyer De La Giroday, N. D.

Matsukevich, P. Maunz, S. Olmschenk, D. Hayes, L. Luo, T. A.

Manning, and C. Monroe, Nature46410 (2010).

[4] D. Mayers, and A. Yao, Proc. 39th FOCS (Washington, DC:

IEEE Computer Society) p 503 (1998).

[5] D. Mayers, and A. Yao, Quantum Information and Computa- tion,4(4):273-286, July 2004

[6] J. S. Bell,Physics1, 195 (1964).

[7] N. Brunner, D. Cavalcanti, S. Pironio, V. Scarani, and S. Wehner, Rev. Mod. Phys.86, 419 (2014).

[8] J. F. Clauser, M. A. Horne, A. Shimony, and R. A. Holt, Phys.

Rev. Lett.,23(15), 1969.

[9] S. J. Summers, and R. Werner J. Math. Phys.282440 (1987).

[10] S. Popescu and D. Rohrlich, Phys. Lett. A, 169(6), 1992.

[11] B. S. Tsirelson, Hadronic J. Suppl.,8, 329 (1993).

[12] B. W. Reichardt, F. Unger and U. Vazirani, Nature496, 456 (2013).

[13] A. Coladangelo, K. T. Goh, and V. Scarani, Nature Communi- cations8, 15485 (2017).

[14] M. McKague, Proc. TQC’11, LNCS, 6745, 2014.

[15] K. F. Pál, T. Vértesi, and M. Navascués, Phys. Rev. A, 90, 042340 (2014).

[16] X. Wu, Y. Cai, T. H. Yang, H. N. Le, J.-D. Bancal, and V.

Scarani. Phys. Rev. A,90, 042339 (2014).

[17] C.-E. Bardyn, T. C. H. Liew, S. Massar, M. McKague, and V.

Scarani. Phys. Rev. A80, 062327 (2009).

[18] M. McKague, T. H. Yang, and V. Scarani, J. Phys. A: Math.

Theor.,45455304 (2012).

[19] T. H. Yang, and M. Navascués, Phys. Rev. A,87, 050102(R) (2013).

[20] C. Bamps and S. Pironio. Phys. Rev. A,91, 052111 (2015).

[21] T. H. Yang, T. Vértesi, J-D. Bancal, V. Scarani, and M. Navas- cués, Phys. Rev. Lett.113, 040401 (2014).

[22] J. Kaniewski, Phys. Rev. Lett.,117, 070402 (2016).

[23] J. Kaniewski, Phys. Rev. A95, 062323 (2017).

[24] T. R. Tan, Y. Wan, S. Erickson, P. Bierhorst, D. Kienzler, S.

Glancy, E. Knill, D. Leibfried, and D. J. Wineland, Phys. Rev.

Lett.118, 130403 (2017).

[25] I. Supic, R. Augusiak, A. Salavrakos, A. Acin, New J. Phys.18, 035013 (2016).

[26] R. Gallego, N. Brunner, C. Hadley, and A. Acin, Phys.Rev.Lett.

105, 230501, (2010).

[27] M. Hendrych, R. Gallego, M. Micuda, N. Brunner, A. Acin, and

J. P. Torres, Nature Physics8, 588 (2012).

[28] J. Ahrens, P. Badziag, A. Cabello, and M. Bourennane, Nature Physics8, 592 (2012).

[29] M. Pawłowski and N. Brunner, Phys. Rev. A84, 010302(R) (2011).

[30] H-W Li, M. Pawłowski, Z-Q. Yin, G-C. Guo, and Z-F. Han, Phys. Rev. A84, 034301 (2011).

[31] E. Woodhead, and S. Pironio, Phys. Rev. Lett.115, 150501 (2015).

[32] T. Lunghi, J. B. Brask, C. C. W. Lim, Q. Lavigne, J. Bowles, A. Martin, H. Zbinden, and N. Brunner, Phys. Rev. Lett.114, 150501 (2015).

[33] P. Mironowicz, A. Tavakoli, A. Hameedi, B. Marques, M.

Pawłowski, and M. Bourennane, New J. Phys. 18, 065004 (2016).

[34] M. Dall’Arno, S. Brandsen, F. Buscemi, V. Vedral, Phys. Rev.

Lett.118, 250501 (2017).

[35] M. Dall’Arno, S. Brandsen, F. Buscemi, Proc. R. Soc. A473, 20160721 (2017).

[36] M. Dall’Arno, arXiv:1702.00575.

[37] A. Ambainis, A. Nayak, A. Ta-Shama, U. Vazirani, inProceed- ings of 31st ACM Symposium on Theory of Computing, pp. 376- 383, 1999.

[38] A. Nayak, in Proceedings of the 40th IEEE Symposium on Foundations of Computer Science (FOCS’99), pp. 369-376, 1999.

[39] A. Tavakoli, A. Hameedi, B. Marques, and M. Bourennane, Phys. Rev. Lett.114, 170502 (2015).

[40] E. Woodhead, C. Lim, and S. Pironio, Proc. TQC’12, LNCS, 7582, 107.

[41] M. Navascués and Tamás Vértesi, Phys. Rev. Lett.115, 020501 (2015).

[42] M. Navascués, A. Feix, M. Araújo, and A. Vértesi, Phys. Rev.

A92, 042117 (2015).

[43] A. Tavakoli, D. Rosset, and M. O. Renou, arXiv:1808.02412.

[44] P. Trojek, C. Schmid, M. Bourennane, C. Brukner, M.

Zukowski, and H. Weinfurter, Phys. Rev. A˙ 72, 050305(R) (2005).

[45] J. Bowles, N. Brunner, M. Pawłowski, Phys. Rev. A92, 022351 (2015).

[46] T. Van Himbeeck, E. Woodhead, N. J. Cerf, R. García-Patron, S. Pironio, Quantum1, 33 (2017).

[47] J. B. Brask, A. Martin, W. Esposito, R. Houlmann, J. Bowles, H. Zbinden, and N. Brunner, Phys. Rev. Applied 7, 054018 (2017).

[48] R. Chaves, J. Bohr Brask, and N. Brunner, Phys. Rev. Lett.115, 110501 (2015).

(7)

Appendix A: Self-testing relations for preparations and measurements

In this section we provide a simple example of preparations that saturate the compatibility bound forA2given in the main text. Moreover, we derive the upper bound for compatibility of measurements given in the main text.

First, we consider the case of preparations. Consider preparations such thatρ00 andρ11, and ρ01 andρ10 correspond to antipodal Bloch vectors with a relative angleθ, the maximal quantum value ofA2, is obtained from

A2=1 2 +1

8 X

y

λmax[Vy], (A1)

whereVy =P

x0,x1(−1)xyρx0x1. We represent the preparations on the Bloch sphere asρx0x1 = 1/2 (11+m~x0x1·~σ), where

~

m00 = [cos(θ/2),0,sin(θ/2)]andm~01 = [cos(θ/2),0,−sin(θ/2)], withm~11 =−m~00andm~10 =−m~01. This givesV0 = 2 cos(θ/2)σxandV1= 2 sin(θ/2)σz. The respective largest eigenvalues areλmax[V0] = 2 cos(θ/2)andλmax[V1] = 2 sin(θ/2), leading to

A2=1 2 + 1

4√ 2

h√

1 + cosθ+√

1−cosθi

. (A2)

It is straightforward to see that this achieves the upper bound in the main text; indeed the above choice of preparations leads to β= 4andα= 4 cosθ.

In order to derive the upper bound onA2for compatibility of measurements in the main text we evaluate X

x0,x1

λmax

(−1)x0M0+ (−1)x1M1

(A3) for arbitrary qubit observablesM0, M1. We take advantage of the fact that

λmax[T] +λmax[−T] =λmax[T]−λmin[T], (A4) which for a2×2matrix can be evaluated analytically. More specifically, ifT is a 2×2Hermitian matrix with eigenvalues λ0≥λ1, let

χ:= trT =λ01, ζ:= trT22021

and then

λ0−λ1=p

2ζ−χ2. (A5)

Evaluating this expression forT =M0±M1gives the desired upper bound given.

Appendix B: Operator inequalities for robust self-testing of preparations In this section we provide a detailed derivation of the lower bound on the average fidelityF(A2).

For a real constants >0, to be chosen later, consider for each pair(x0, x1)the operatorKx0x1 −sWx0x1, whereWx0x1 =

1 16

P

y(−1)xyMyandKx0x1 = Λidealx

0x1], for some channelΛ. Suppose now thattx0x1 ∈Ris a lower bound on its eigenvalues, or, equivalently, that the operator inequality

Kx0x1 ≥s Wx0x1+tx0x111 (B1) holds. Then, computing the trace of this inequality withρx0x1and averaging over inputs leads to

S≥ 1 4

X

x0,x1

F(ρidealx

0x1,Λ[ρx0x1])≥ s 4

A2−1

2

+t, t≡1 4

X

x0,x1

tx0x1 , (B2) where the first inequality holds becauseS is defined as maximisation over all possible channels, and the Λused there is one possible choice. In turn, if (B1) holds as an operator inequality, it is valid for any set of preparations {ρx0x1}, and thus F(A2)≥ s4 A212

+t. Note that (B1) has a dependence onM0,M1throughWx0x1. If (B1) holds for a particular choice of measurement operatorsM0,M1, then the bound onF(A2)holds for all preparations, for that particular choice ofM0,M1.

(8)

However, if (B1) holds forallpossibleM0,M1, then the bound onF(A2)is valid for all quantum setups and is thus a robust self-testing inequality. To derive the appropriate constantssandtx0x1, we first allowtx0x1andΛto have a dependence onM0

andM1. We then minimise overM0andM1the constantstx0x1, for a suitable choice ofs, such that, at the end, (B1) holds regardless of the choice of measurement operators.

We choose a dephasing channel of the form

Λθ(ρ) = 1 +c(θ)

2 ρ+1−c(θ)

2 Γ(θ)ρΓ(θ), (B3)

where for0≤θ≤π/4we useΓ =σx, while forπ/4< θ≤π/2we useΓ =σz. The functionc(θ)∈[−1,1]will be specified later.

In the interval0≤θ≤π/4, the action of the channel leads to K00= 11+σx

2 K01= 11+c(θ)σz

2 K10= 11−c(θ)σz

2 K11= 11−σx

2 , (B4)

whereas in the intervalπ/4< θ≤π/2, we have K00= 11+c(θ)σx

2 K01=11+σz

2 K10= 11−σz

2 K11=11−c(θ)σx

2 . (B5)

As discussed in the main text, for any given set of preparations, the optimal measurements are projective and rank-one.

Furthermore, any two such measurements can be represented on an equator of the Bloch sphere. Due to the freedom of setting the reference frame, we can without loss of generality represent the two measurements in thexz-plane, i.e.

M0= cosθ σx+ sinθ σz M1= cosθ σx−sinθ σz.

We can therefore writeWx0x1 as W00=1

8cosθσx W01=1

8sinθσz W10=−1

8sinθσz W11=−1

8cosθσx. (B6) We can reduce the number of operator inequalities (B1) by exploiting the apparent symmetries in the expressions forWx0x1 and Kx0x1: we restrict ourselves so thatto≡t01=t10andte≡t00=t11. Thus, we have to consider two operator inequalities in each intervalθ∈[0, π/4]andθ∈(π/4, π/2]. In the first interval, the two operator inequalities are

1 +σx

2 −s

8cosθσx−te11≥0 1 +c(θ)σz

2 −s

8sinθσz−to11≥0. (B7) In the second interval, the two operator inequalities are

1 +c(θ)σx

2 −s

8cosθσx−te11≥0 1 +σz

2 −s

8sinθσz−to11≥0. (B8) We now focus on the former interval. Solving the two inequalities fortoandtewe obtain

te≤1−s

8cosθ, to≤ 1

8(4 + 4c(θ)−ssinθ), (B9) te≤ s

8cosθ, to≤ 1

8(4−4c(θ) +ssinθ). (B10) Any choice oftoandte satisfying these constraints gives rise to valid operator inequalities. In order to obtain the strongest bound, we choose the largest values oftoandteconsistent with their respective constraints, i.e.,

te= minn 1−s

8cosθ,s 8cosθo

to= min 1

8(4 + 4c(θ)−ssinθ),1

8(4−4c(θ) +ssinθ)

. (B11)

A similar procedure for the intervalθ∈(π/4, π/2]leads to te= min

1

8(4 + 4c(θ)−scosθ),1

8(4−4c(θ) +scosθ)

to= minn 1−s

8sinθ,s 8sinθo

. (B12)

It is worth pointing out that the two intervals only differ by exchangingte↔toandsinθ↔cosθ. Hence, for any givenθ, we have constructed operator inequalities of the form (B1).

(9)

As shown in the main text, we obtain our lower bound on the average fidelity from F(A2)≥ s

4(A2−1/2) + min

M0,M1t(M0, M1)≡L(A2), (B13)

wheret(M0, M1) = (te+to)/2. To compute this quantity we fix the value ofsto be s= 4 1 +√

2

(B14) and choose the dephasing function asc(θ) = min{1,4ssinθ}whenever θ ∈ [0, π/4], and c(θ) = min{1,s4cosθ} whenever θ∈(π/4, π/2]. It is easy to see thatc(θ)∈[0,1], which ensures thatΛθis a valid quantum channel, and thatc(θ)is continuous atθ=π/4. A simple calculation shows that in this case

t= 2−√ 2

4 , (B15)

which gives the lower bound

F(A2)≥ 1 +√

2

A2− 3 2√

2 ≡L(A2). (B16)

One can check that choosing distinct values ofswill not lead to improved lower bounds.

Appendix C: Tightness of fidelity bounds

In the main text, we have derived fidelity bounds for both the preparations and the measurements, based on operator inequali- ties. Specifically, we obtain a lower bound on the average fidelityFof the prepared states (with respect to the ideal ones) given by the linear expression

F(A2)≥ 1 +√

2

A2− 3 2√

2 ≡L(A2). (C1)

For measurements, a similar bound is obtained on the average fidelityF0with respect to the ideal ones. In the present appendix, we discuss the tightness of these bounds.

We start with our bound on the fidelity of the states. As discussed in the main text, obtainingA2=Q2impliesF = 1, i.e. the states are the ideal ones (up to a unitary). Let us refer to the optimal strategy (with the ideal states) as strategyS1. Then, forA2= C2, our bound givesF ≥3/4. This bound is tight and can be obtained via the set of pure statesρx0x1 = (11+ (−1)x0x1σz)/2 (diagonal in the same basis, hence classical), combined with the measurementsM0 =M1z. Let us refer to this strategy as S2.

The above shows that our bound (C1) is tight as far as linear inequalities are concerned. More generally, the bound is in fact tight in general, when shared randomness between the preparation and measurement devices is taken into account. In this case, taking a convex combination between strategiesS1andS2allows us to get any point on the line (i.e. pair of valuesFandA2) betweenS1andS2.

It is also interesting to understand what happens when shared randomness between the devices is not taken into account. In this case, the end points (A2 =Q2,F = 1) and (A2 =C2,F = 3/4) can still be obtained. To understand what happens in the intermediate regionC2 <A2 < Q2, we first performed a numerical analysis. Specifically, we chose randomly four qubit states, and compute (i) the maximal value ofA2(optimizing over the measurements), and (ii) the average fidelityF (where the optimization over channels is restricted here to unitaries). The resulting points are shown on Fig. 2 (blue circles). This indicates that forC2<A2 < Q2, the bound (C1) cannot be saturated anymore. Moreover, we conjecture that an optimal class of strategies is given by the pure states

00i=|0i |ψ11i=|1i |ψ01i= cosθ|0i+ sinθ|1i |ψ10i= cosθ|0i −sinθ|1i (C2) and the measurementsMy= cos(ϕ)σz+ (−1)ysin(ϕ)σx. Straightforward calculations show that, takingtanϕ= sin 2θ, leads to

A2= 1 2+1

4 q

1 + tan2(ϕ) and F= 1

4(3 + tanϕ). (C3)

This gives a parametric curve, as a function ofϕ∈[0, π/4], given by the red curve in Fig. 2. This curve is in excellent agreement with the numerical results obtained before. Note that this class of strategies interpolates between the strategiesS1(settingϕ= 0) andS2(settingϕ=π/4).

(10)

FIG. 2. The black line is the analytic lower bound on the average fidelityF(F0)for prepared states (measurements), as a function of the observed value ofA2. To characterise the region accessible via pure qubit strategies (i.e. without shared randomness), we perform numerics generating randomly sets of qubit preparations (blue circles and crosses); here we show the numerical results for the case of states, but similar results are obtained for the case of measurements. In the regionC2 <A2 < Q2, we conjecture that the class of strategies given in the text (corresponding to the red curve) are optimal; both forFandF0. Finally, the green dashed line is our conjectured upper bound on the average fidelity.

Next we discuss the bound on the average fidelity of measurements. As discussed in the main text, the linear boundF0(A2)≥ L(A2)is optimal as far as linear inequalities are concerned. Moreover, when allowing for shared randomness the bound is tight in general forC2 ≤ A2 ≤ Q2. This is obtained by considering convex combinations of strategyS01(defined as the optimal strategyS1, up to a rotation ofπ/8around theyaxis; see below), and the following strategy (referred to asS3): takeM0z

andM1=11, with the statesρ0001= (11+σz)/2andρ1011= (11−σz)/2.

Similarly to the case of states, we now consider the situation where shared randomness between the devices is not allowed.

Performing a numerical analysis similar to the one described above (except that measurements are no generated randomly), we observe that the accessible region (in terms ofF0vsA2) appears to be exactly the same as for the case of states (i.e. the blue region in Fig. 2). We conjecture that the lower bound is given by the following class of optimal strategies: take the measurements:

M0z and M1=ησx+ (1−η)11 (C4) with the states|ψ00i= cosθ|0i+ sinθ|1i,|ψ01i= cosθ|0i −sinθ|1i,|ψ10i = cosθ|1i+ sinθ|0i, and|ψ11i= cosθ|1i − sinθ|0i. Settingη= tan 2θ, we get

A2= cos2(θ) 2 +1

4 +sin2(2θ)

cos(2θ) and F =1

4(3 + tan(2θ)). (C5)

This gives a parametric curve, as a function ofθ∈[0, π/8], given by the red curve in Fig. 2. This curve is in excellent agreement with the numerical results obtained before. Also, this curve turns out to be exactly the same as the curve we obtained above for the case of states. Note that this class of strategies interpolates between the strategiesS10 (settingθ =π/8) andS3(setting θ= 0).

Finally, note that the numerics also suggests that there is a linear upper bound on the average fidelitiesF(F0) as a function of A2(see Fig. 2); specificallyF ≤ Q1−Q2

2−3/4A2+QQ22−3/4

2−3/4and similarly forF0. It would be interesting to provide a proof of these upper bounds.

Appendix D: Operator inequalities for robust self-testing of measurements

In this section, we account for the detailed derivation of the lower bound on the average fidelity of the measurementsF0(A2).

The approach bears significant resemblance to the case of robustly self-testing preparations, as outlined in AppendixB.

We aim to derive operator inequalities of the form

Kyb({ρx0x1})≥sZyb+ty({ρx0x1})11, (D1)

(11)

where Zyb = 18P

x0,x1ρx0x1δb,xy and Kyb({ρx0x1}) = Λ[(Myb)ideal]. For the sake of simplicity, we first apply a uni- tary channel to(Myb)ideal align these operators with the eigenstates of σx and σz. Then, we adopt the same (unital, trace- preserving) channelΛas specified in the main text, with the same coefficients as used to robustly self-test the preparations:

c(θ) = min{1,s4sinθ}whenθ∈[0, π/4]andc(θ) = min{1,4scosθ}whenθ∈(π/4, π/2].

It is straightforward to see that, for any given pair of measurements, the optimal choice of preparations are four pure qubit states, such thatρ00andρ11, andρ01andρ10respectively, correspond to antipodal vectors on the Bloch sphere. Therefore, we can without loss of generality restrict to such preparations since these impose the weakest constraints on the measurements of our interest. We can therefore parametrise the preparationsρx0x1= 1/2 (11+m~x0x1·~σ)by Bloch vectors

~

m00= [cosθ,0,sinθ] m~11=−[cosθ,0,sinθ] m~01= [cosθ,0,−sinθ] m~10= [−cosθ,0,sinθ]. (D2) ExpressingZybin terms of these preparations gives

Z00=1

8(11+ cosθσx) Z01=1

8(11−cosθσx) Z10=1

8(11+ sinθσzz Z11=1

8(11−sinθσz). (D3) Due to symmetries, we restrict ourselves so thatto ≡t01 =t10andte ≡t00 =t11. Thus, we have to consider two operator inequalities in each intervalθ∈[0, π/4]andθ∈(π/4, π/2]. In the first interval, the two operator inequalities are

1 +σx

2 −s

8(11+ cosθσx)−te11≥0 1 +c(θ)σz

2 −s

8(11+ sinθσz)−to11≥0. (D4) In the second interval, the two operator inequalities are

1 +c(θ)σx

2 −s

8(11+ cosθσx)−te11≥0 1 +σz 2 −s

8(11+ sinθσz)−to11≥0. (D5) Just as in AppendixB, we solve these inequalities forteandto, and choose the largest value compatible with the solutions. In the first interval, this gives

te= min 1

8(8−s−scosθ),s

8(cosθ−1)

to= min 1

8(4c(θ)−ssinθ−s+ 4),1

8(−4c(θ) +ssinθ−s+ 4)

. (D6) A similar procedure for the intervalθ∈(π/4, π/2]leads to

te= min 1

8(4c(θ)−scosθ−s+ 4),1

8(−4c(θ) +scosθ−s+ 4)

to= min s

8(sinθ−1),1

8(8−s−ssinθ)

. (D7) For any choice ofθ, we have constructed operator inequalities of the form (D1).

In order to obtain our lower bound onF0, we must minimise the quantityt(θ) = (te+to)/2for a specific choice ofs. In analogy with the procedure in AppendixD, we chooses= 4 1 +√

2

, which returnsminθt(θ) =−3/(2√

2). Hence, we have obtained the lower bound

F0(A2)≥ 1 +√

2

A2− 3 2√

2 =L(A2). (D8)

Appendix E: Self-testing all pairs of incompatible Pauli observables

Consider a generalisation of the2→1RAC, in which we introduce a bias on the score associated to certain inputs. Specifi- cally, whenever the game is successful, i.e.b=xy, the awarded score isq/2ifx0⊕x1= 0, and(1−q)/2ifx0⊕x1= 1, for someq∈[0,1]. The average score reads

Aq2=1 2

X

x0,x1,y

r(x0, x1)P(b=xy|x0, x1, y), (E1) wherer(x0, x1) =q/2ifx0⊕x1= 0andr(x0, x1) = (1−q)/2ifx0⊕x1= 1. Note that forq= 1/2, we recover the standard 2→1RAC. Based on the quantityAq2, we will now see how to derive a self-testing condition for any pair of incompatible Pauli observables, i.e. any pair of non-commuting projective rank-one qubit measurements.

(12)

We start by expressingAq2for a quantum strategy:

Aq2=1 2+1

4 X

x0,x1

r(x0, x1) tr (ρx0x1((−1)x0M0+ (−1)x1M1))≤ 1 2+1

4 X

x0,x1

r(x0, x1max[(−1)x0M0+ (−1)x1M1].

(E2) Denotingµkmin

M0+ (−1)kM1

andνkmax

M0+ (−1)kM1

, fork= 0,1, we obtain Aq2≤ 1

2 +1

8[q(µ0−ν0) + (1−q) (µ1−ν1)]. (E3) Following a derivation analogous to that appearing in AppendixAto obtain, we obtain

Aq2≤1 2 +1

8 h

qp

β+α+ (1−q)p β−αi

, (E4)

whereβ = 2 tr M02+M12

−tr (M0)2−tr (M1)2 andα = 2 tr ({M0, M1})−2 tr (M0) tr (M1). Treatingαandβ as independent variables, we obtain the largest value of the right-hand-side of Eq. (E4) by demanding that the derivative with respect toαequals zero, and checking that the second derivative is negative at this point. We obtain the optimality constraint

α= 2q−1

1−2q+ 2q2β. (E5)

Inserting this value back into Eq. (E4), we find an upper bound onAq2as obtained by independent variablesαandβ. It turns out that this bound can be saturated by the de facto coupled variablesαandβ. From Eq. (E4), it is clear that a necessary condition for optimality is to maximiseβ. This amounts to the observablesM0andM1being traceless and such thatM02=M12=11, leading toβ = 8. This implies that the observables represent projective rank-one measurements. Hence, we can writeMy =~ny·~σ where the Bloch vector satisfies|~ny|= 1. Hence, we haveα= 8~n0·~n1. Thus, Eq.(E5) becomes

~

n0·~n1= 2q−1

1−2q+ 2q2, (E6)

which has a solution for any choice ofq. Note that settingq= 1/2reduces the above to~n0·~n1= 0which we recognise as the optimality constraint for the standard2→1random access code. In conclusion, for any pair of incompatible Pauli observables (characterised by the scalar product~n0·~n1), we have a gameAq2(whereqis chosen in order to satisfy the above equation), such that the maximal score can only be attained by using that specific pair of Pauli observables. We thus obtain a general class of self-tests for any pair of Pauli observables, corresponding to saturating the maximal quantum value ofAq2for a given value ofq:

Aq2≤ 1 2

1 +p

1−2q+ 2q2

. (E7)

Appendix F: Self testing for theN→1random access code

In this appendix, we extend the results presented in the main text to self-test the preparations and measurements in anN →1 RAC. The latter is a straightforward generalisation of the2→1RAC considered in the main text. The input of the preparation device is a randomN-bit stringx≡(x1, . . . , xN), while the input of the measurement device isy∈ {1, . . . , N}. The average score is

AN = 1 N2N

X

x,y

P(b=xy|x, y). (F1)

Considering qubit statesρx, and measurement observablesMy, we get AN = 1

2+ 1 N2N+1

X

x,y

(−1)xytr (ρxMy). (F2)

Références

Documents relatifs

Given this type of facts, theories on the incompatibility of measure phrases and negative/ non neutral adjectives usually claim that the semantics of measures

We develop methods for (i) robustly self-testing extremal qubit POVMs (which feature either three or four outcomes), and (ii) certify that an uncharacterised qubit measurement

We consider a prepare-and-measure scenario with a bound on the Hilbert space dimension and develop methods for (i) robustly self-testing extremal qubit POVMs and (ii) certifying

A comparison was carried out between the CV measurements obtained with imaging (aorta; heart cavities and cardiac wall thicknesses; maximum cardiac diameter and cardiothoracic

To prove Theorem 2.1 we shall follow [13] and we shall study the relation between the flat of α of slope α (c) and the Aubry set at c... 2.1. The Aubry sets and the flats

For every p-pass Shellsort algorithm and every increment sequence, every subset of n!/2 n input permutations of n keys contains an input permutation that uses ⍀ ( pn 1⫹(1⫺⑀)/p

This paper shows experimental results obtained by EIS measurements on 2 types of Lithium-ion batteries, which had different aging processes and may differ at

This article generalizes a previous work in which the author obtained a large lower bound for the lifespan of the solutions to the Primitive Equations, and proved convergence to the