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

Sample efficient tomography via Pauli Measurements

Nengkun Yu Centre for Quantum Software and Information, Faculty of Engineering and Information Technology, University of Technology, Sydney, NSW 2007, Australia
Abstract

Pauli Measurements are the most important measurements in both theoretical and experimental aspects of quantum information science. In this paper, we explore the power of Pauli measurements in the state tomography related problems. Firstly, we show that the quantum state tomography problem of nn-qubit system can be accomplished with 𝒪(10nϵ2){\mathcal{O}}(\frac{10^{n}}{\epsilon^{2}}) copies of the unknown state using Pauli measurements. As a direct application, we studied the quantum overlapping tomography problem introduced by Cotler and Wilczek in Ref. Cotler and Wilczek (2020). We show that the sample complexity is 𝒪(10klog((nk)/δ))ϵ2)\mathcal{O}(\frac{10^{k}\cdot\log({{n}\choose{k}}/\delta))}{\epsilon^{2}}) for quantum overlapping tomography of kk-qubit reduced density matrices among nn is quantum system, where 1δ1-\delta is the confidential level, and ϵ\epsilon is the trace distance error. This can be achieved using Pauli measurements. Moreover, we prove that Ω(log(n/δ)ϵ2)\Omega(\frac{\log(n/\delta)}{\epsilon^{2}}) copies are needed. In other words, for constant kk, joint, highly entangled, measurements are not asymptotically more efficient than Pauli measurements.

pacs:
03.65.Wj, 03.67.-a

I Introduction

It is increasingly important to understand how the cost scales of learning useful information as the experiments can control larger and larger quantum systems. Quantum state tomography refers to a procedure of reconstructing the density matrix of a quantum state from various measurements on multiple copies of the state. This fundamental task is crucial for quantum information experiments and theoretically goes back at least to the work of Helstrom, Holevo, and others from around 1970.

In quantum state tomography, one is given kk copies of an nn-qubit quantum state ρ\rho and is required to output a classical description of density matrix ρ^\hat{\rho} close to ρ\rho by performing quantum measurements on ρk\rho^{\otimes k}. Broadly speaking, there are three categories of measurements: one consists of joint (entangled) measurements on ρk\rho^{\otimes k}, as in Bagan et al. (2004); Keyl (2006); Guţă and Kahn (2008). In Haah et al. (2016); O’Donnell and Wright (2016, 2017), the authors showed that the optimal scaling of the sample number kk as a function of trace distance goal ϵ\epsilon is n4nϵ2n\propto\frac{4^{n}}{\epsilon^{2}}. The second category consists of measurements on each copy of the state ρ\rho, whose results are to be combined to reconstruct the density matrix, as in Flammia et al. (2012a). Ref. Kueng et al. (2017) showed that if one can perform many-outcome measurements, tomography is possible using n8n/δ2n\propto 8^{n}/\delta^{2} copies, and this is optimal if the measurements on every copy are independent as showed in Haah et al. (2016). The third category local measurements consists of measurements on each qubit of each copy of the state ρ\rho, whose results are to be combined to reconstruct the density matrix Guţă et al. (2020). Generally, joint measurements over several copies of the state can usually achieve lower sampling rates Haah et al. (2016); O’Donnell and Wright (2016, 2017), but are much more challenging to implement. On the other hand, Pauli measurements are experimental friendly, therefore, extremely important both theoretically and experimentally. If one is restricted to use Pauli measurements, Flammia, Gross, Liu and Eisert observed that the quantum state tomography can be accomplished using 𝒪(n16nϵ2){\mathcal{O}}(\frac{n\cdot 16^{n}}{\epsilon^{2}}) copies in Flammia et al. (2012b). This was improved the bound to 𝒪(n12nϵ2){\mathcal{O}}(\frac{n\cdot 12^{n}}{\epsilon^{2}}) in Guţă et al. (2020).

In Cotler and Wilczek (2020), Cotler and Wilczek introduced the problem called quantum overlapping tomography problem. The goal is to output the classical description of all kk-qubit reduced density matrix of an nn-qubit system. This problem is also of great importance in quantum information science, because many important physical quantities, for instance, energy and entropy, depend on very small parts of the whole system only, in other words, depends on the set of local reduced density matrices. By using the perfect hash families, they showed that one only needs to use e𝒪(k)log2ne^{\mathcal{O}(k)}\log^{2}n rounds of parallel measurements to achieve this goal. Ref. García-Pérez et al. (2020) proposed a measurement scheme to perform two-qubit tomography of all pairs. Later Ref. Evans et al. (2019) provided an algorithm to estimate the expectation value of mm Pauli operators with weight k\leq k using 𝒪(3klogm)\mathcal{O}(3^{k}\log m) measurements for small kk. All of these bounds can be achieved using Pauli measurements.

I.1 Our results

In this paper, we study the power of Pauli measurements, the most important class of measurements, in two tomography related problems.

In the first part of this paper, we revisit the sample complexity of quantum state tomography problem,

Problem 1.

Given an unknown nn-qubit quantum mixed state ρ\rho, the goal is to output density matrices σ\sigma such that

ρσ1<ϵ\displaystyle||\rho-\sigma||_{1}<\epsilon (1)

for a given ϵ>0\epsilon>0. How many copies of ρ\rho are needed to achieve this goal, with high probability?

The sample complexity of this problem is nearly resolved in the general joint measurement setting and independent measurement setting Haah et al. (2016). However, it is still unclear in the local measurement, in particular, the Pauli measurement setting. We show that

Theorem 1.

The quantum state tomography problem can be solved using 𝒪(10nlog1δϵ2)\mathcal{O}(\frac{10^{n}\log\frac{1}{\delta}}{\epsilon^{2}}) copies of ρ\rho via Pauli measurement on each qubit to success with probability at least 1δ1-\delta.

Secondly, we study the sample complexity of the quantum overlapping tomography problem: “How many copies of states are necessary and sufficient to reconstruct all the kk-reduced density matrices of unknown nn-qubit state ρ\rho, to within additive error ϵ\epsilon, for constant kk?” More precisely, the formula of this problem is

Problem 2.

Given an unknown nn-qubit quantum mixed state ρ1,2,,n\rho_{1,2,\dots,n}, and 1kn1\leq k\leq n, the goal of quantum overlapping tomography is to output density matrices σS\sigma_{S} for all S{1,2,,n}S\subseteq\{1,2,\dots,n\} with |S|k|S|\leq k such that

ρSσS1<ϵ\displaystyle||\rho_{S}-\sigma_{S}||_{1}<\epsilon (2)

for a given ϵ>0\epsilon>0, where ρS\rho_{S} denotes the reduced density matrix of ρ1,2,,n\rho_{1,2,\dots,n} on SS.

If one only cares about MM density matrices σS1,σS2,,σSM\sigma_{S_{1}},\sigma_{S_{2}},\cdots,\sigma_{S_{M}} with |Si|k|S_{i}|\leq k, this problem is called partial quantum overlapping tomography.

How many copies of ρ\rho are needed to achieve this goal, with probability at least 1δ1-\delta for δ>0\delta>0?

We show that,

Theorem 2.

The sample complexity of quantum overlapping tomography of nn-qubit system is Θ(ϵ2log(n/δ))\Theta(\epsilon^{-2}\cdot\log(n/\delta)) for constant kk to succeed with probability at least 1δ1-\delta, even using general joint measurement schemes. General quantum overlapping tomography problem for 1kn1\leq k\leq n can be accomplished by performing Pauli measurements on 𝒪(ϵ210klog((nk)/δ))\mathcal{O}(\epsilon^{-2}\cdot 10^{k}\cdot\log({{n}\choose{k}}/\delta)) copies. Moreover, for partial quantum overlapping tomography with MM outcomes, the sample complexity is 𝒪(ϵ210klog(2M/δ))\mathcal{O}(\epsilon^{-2}\cdot 10^{k}\cdot\log(2M/\delta)) using Pauli measurements.

For the lower bound part, we show that to succeed with probability at least 1δ1-\delta, it is necessary to have Clog(n/δ)ϵ2C_{\ell}\frac{\log(n/\delta)}{\epsilon^{2}} copies even if joint measurement on many copies of ρ\rho are allowed, where C>0C_{\ell}>0 is a constant. For example, if mm copies of the nn-qubit system are prepared and measured. In the joint measure setting, the mnmn qubits may be accessed collectively.

The upper bound can be achieved by the following algorithm,

1Repeat the following measurement 𝒪(g(k)ϵ2log((nk)/δ))\mathcal{O}(g(k)\cdot\epsilon^{-2}\cdot\log({{n}\choose{k}}/\delta)) times;
2 Measuring each qubit using some informationally complete measurement for any copy of ρ\rho;
Algorithm 1 Quantum Overlapping Tomography

Here g(k)>0g(k)>0 is a function depends on the informationally complete measurements and kk only. Therefore, for constant kk and fixed informationally complete measurements, the used number of copies becomes 𝒪(ϵ2log(n/δ))\mathcal{O}(\epsilon^{-2}\cdot\log(n/\delta)).

As an example, we choose informationally complete measurement =16{σI+σX,σIσX,σI+σY,σIσY,σI+σZ,σIσZ}\mathcal{M}=\frac{1}{6}\{\sigma_{I}+\sigma_{X},\sigma_{I}-\sigma_{X},\sigma_{I}+\sigma_{Y},\sigma_{I}-\sigma_{Y},\sigma_{I}+\sigma_{Z},\sigma_{I}-\sigma_{Z}\}, which can be regarded as random Pauli measurement, on each qubit. Then we can obtain the measurement scheme using 𝒪(ϵ210klog(M/δ))\mathcal{O}(\epsilon^{-2}\cdot 10^{k}\cdot\log(M/\delta)) copies of ρ\rho.

This is the first nontrivial example that Pauli measurements are as powerful as general joint measurements asymptotically. For this example, the number of unknown parameters of the output is exponentially larger than the number of copies.

This implies that for all observables OSIS¯O_{S}\otimes I_{\bar{S}} with SS being a set consisting of at most kk-qubit, S¯\bar{S} being the complementary set of SS and IOSI-I\leq O_{S}\leq I, we can output estimation oSo_{S} such that

|tr[ρ(OSIS¯)]oS|ϵ\displaystyle|\operatorname{tr}[\rho(O_{S}\otimes I_{\bar{S}})]-o_{S}|\leq\epsilon

II Preliminaries

A positive-operator valued measure (POVM) on a finite dimensional Hilbert space \mathcal{H} is a set of positive semi-definite matrices ={Mi}\mathcal{M}=\{M_{i}\} such that

Mi=I.\displaystyle\sum M_{i}=I_{\mathcal{H}}.

We need the concept of the informationally complete POVM as follows,

Definition 3.

An informationally complete POVM is a POVM whose outcome probabilities are sufficient to determine any state.

Equaivalently, a POVM ={Mi}\mathcal{M}=\{M_{i}\} on dd-dimensional \mathcal{H} is informationally complete if the linear span of {Mi}\{M_{i}\} equals to the whole d×dd\times d matrix space. In qubit system, it means σI,σX,σY\sigma_{I},\sigma_{X},\sigma_{Y} and σZ\sigma_{Z} all live in the linear span of {Mi}\{M_{i}\}, where σI\sigma_{I}, σX,σY\sigma_{X},\sigma_{Y} and σZ\sigma_{Z} are Pauli matrices,

σI=[1001],σX=[0110],σZ=[1001],σY=[0ii0].\displaystyle\sigma_{I}=\begin{bmatrix}1&0\\ 0&1\end{bmatrix},\sigma_{X}=\begin{bmatrix}0&1\\ 1&0\end{bmatrix},\sigma_{Z}=\begin{bmatrix}1&0\\ 0&-1\end{bmatrix},\sigma_{Y}=\begin{bmatrix}0&i\\ -i&0\end{bmatrix}.

It is direct to observe that =16{σI+σX,σIσX,σI+σY,σIσY,σI+σZ,σIσZ}\mathcal{M}=\frac{1}{6}\{\sigma_{I}+\sigma_{X},\sigma_{I}-\sigma_{X},\sigma_{I}+\sigma_{Y},\sigma_{I}-\sigma_{Y},\sigma_{I}+\sigma_{Z},\sigma_{I}-\sigma_{Z}\} is an informationally complete measurement.

Observation 1.

Given sufficient measurement outcomes of an informationally complete POVM, one can determine the state with high accuracy and confidence.

Observation 2.

For information complete POVMs, 1\mathcal{M}_{1} on 1\mathcal{H}_{1} and 2\mathcal{M}_{2} on 2\mathcal{H}_{2}, 12\mathcal{M}_{1}\otimes\mathcal{M}_{2} is an informationally complete POVM on 12\mathcal{H}_{1}\otimes\mathcal{H}_{2}.

Directly, n==16n{σI+σX,σIσX,σI+σY,σIσY,σI+σZ,σIσZ}n\mathcal{M}^{\otimes n}=\mathcal{M}=\frac{1}{6^{n}}\{\sigma_{I}+\sigma_{X},\sigma_{I}-\sigma_{X},\sigma_{I}+\sigma_{Y},\sigma_{I}-\sigma_{Y},\sigma_{I}+\sigma_{Z},\sigma_{I}-\sigma_{Z}\}^{\otimes n} is an informationally complete POVM of nn-qubit system.

Definition 4.

Let X1,X2,,XnX_{1},X_{2},\cdots,X_{n} be nn samples of a distribution on {1,2,,n}\{1,2,\cdots,n\}. Then the empirical distribution is defined as

p(i)^=numberofoccurrencesofin\displaystyle\hat{p(i)}=\frac{\mathrm{number~{}of~{}occurrences~{}of}~{}i}{n}

The following McDiarmid’s inequality McDiarmid (1989) will be used in this paper.

Lemma 5.

Consider independent random variables X1,X2,Xn{\displaystyle X_{1},X_{2},\dots X_{n}} on probability space (Ω,,P){\displaystyle(\Omega,{\mathcal{F}},{\text{P}})} where Xi𝒳i{\displaystyle X_{i}\in{\mathcal{X}}_{i}} for all i{\displaystyle i} and a mapping f:𝒳1×𝒳2××𝒳n{\displaystyle f:{\mathcal{X}}_{1}\times{\mathcal{X}}_{2}\times\cdots\times{\mathcal{X}}_{n}\rightarrow\mathbb{R}}. Assume there exist constant c1,c2,,cn{\displaystyle c_{1},c_{2},\dots,c_{n}} such that for all i{\displaystyle i},

supx1,,xi1,xi,xi,xi+1,,xn|f(x1,,xi1,xi,xi+1,,xn)f(x1,,xi1,xi,xi+1,,xn)|ci.\displaystyle{\displaystyle{\underset{x_{1},\cdots,x_{i-1},x_{i},x_{i}^{\prime},x_{i+1},\cdots,x_{n}}{\sup}}|f(x_{1},\dots,x_{i-1},x_{i},x_{i+1},\cdots,x_{n})-f(x_{1},\dots,x_{i-1},x_{i}^{\prime},x_{i+1},\cdots,x_{n})|\leq c_{i}.} (3)

(In other words, changing the value of the i{\displaystyle i}-th coordinate xi{\displaystyle x_{i}} changes the value of f{\displaystyle f} by at most ci{\displaystyle c_{i}}.) Then, for any ϵ>0{\displaystyle\epsilon>0},

Pr(f(X1,X2,,Xn)𝔼[f(X1,X2,,Xn)]ϵ)exp(2ϵ2i=1nci2).\displaystyle{\displaystyle{\mathrm{Pr}}(f(X_{1},X_{2},\cdots,X_{n})-\mathbb{E}[f(X_{1},X_{2},\cdots,X_{n})]\geq\epsilon)\leq\exp\left(-{\frac{2\epsilon^{2}}{\sum_{i=1}^{n}c_{i}^{2}}}\right)}. (4)

III Quantum Tomography Using Pauli Measurement

We start from the following observation:

When we measure an element of Pauli group, for instance σXσY\sigma_{X}\sigma_{Y}, on a two-qubit state ρ\rho, the outcome is a sample from a 44-dimensional probability distribution, says (p00,p01,p10,p11)(p_{00},p_{01},p_{10},p_{11}), such that

tr(ρ(σXσY))=p00p01p10+p11.\displaystyle\operatorname{tr}(\rho(\sigma_{X}\otimes\sigma_{Y}))=p_{00}-p_{01}-p_{10}+p_{11}.

One can easily observe that

tr(ρ(σXσI))=p00+p01p10p11,\displaystyle\operatorname{tr}(\rho(\sigma_{X}\otimes\sigma_{I}))=p_{00}+p_{01}-p_{10}-p_{11},
tr(ρ(σIσY))=p00p01+p10p11,\displaystyle\operatorname{tr}(\rho(\sigma_{I}\otimes\sigma_{Y}))=p_{00}-p_{01}+p_{10}-p_{11},
tr(ρ(σIσI))=p00+p01+p10+p11.\displaystyle\operatorname{tr}(\rho(\sigma_{I}\otimes\sigma_{I}))=p_{00}+p_{01}+p_{10}+p_{11}.

In other words, by measuring XYXY, we actually obtained a sample of σXσI\sigma_{X}\sigma_{I}, a sample of σYσI\sigma_{Y}\sigma_{I}, and a sample of σIσI\sigma_{I}\sigma_{I}.

This is also true for general nn-qubit system as the following observation shows,

Observation 3.

For any P=P1P2Pn{σX,σY,σZ}nP=P_{1}\otimes P_{2}\otimes\cdots\otimes P_{n}\in\{\sigma_{X},\sigma_{Y},\sigma_{Z}\}^{\otimes n}, the measurement result of performing measurement PiP_{i} on the ii-th qubit is an nn-bit string ss. One can interpretes the measurement outcome of performing Qi{σI,σX,σY,σZ}Q_{i}\in\{\sigma_{I},\sigma_{X},\sigma_{Y},\sigma_{Z}\} on the ii-th qubit if Qi=PiQ_{i}=P_{i} or Qi=σIQ_{i}=\sigma_{I}. We call these Q=Q1Q2QnQ=Q_{1}\otimes Q_{2}\otimes\cdots\otimes Q_{n} corresponds to PP.

Our measurement scheme is as follows: For any ϵ>0\epsilon>0, fix an integer m=1610nlog1δ3nϵ2m=16\cdot\frac{10^{n}\log\frac{1}{\delta}}{3^{n}\cdot\epsilon^{2}}.

For any P{σX,σY,σZ}nP\in\{\sigma_{X},\sigma_{Y},\sigma_{Z}\}^{\otimes n}, one performs mm times PP on ρ\rho, and records the mm samples of the 2n2^{n} dimensional outcome distribution.

According to the key observation, this measurement scheme provides m3nwm\cdot 3^{n-w} samples of the expectation tr(ρP)\operatorname{tr}(\rho P), says μPm3nw\frac{\mu_{P}}{m\cdot 3^{n-w}} for each Pauli operator P{σI,σX,σY,σZ}nP\in\{\sigma_{I},\sigma_{X},\sigma_{Y},\sigma_{Z}\}^{\otimes n} with weight ww, where m3nwμPm3nw-m\cdot 3^{n-w}\leq\mu_{P}\leq m\cdot 3^{n-w}.

Output

σ=PμPm3nw2nP.\displaystyle\sigma=\sum_{P}\frac{\mu_{P}}{m\cdot 3^{n-w}\cdot 2^{n}}P.

Using this scheme, we obtained m3nm\cdot 3^{n} independent samples,

X1,X2,,Xm3n\displaystyle X_{1},X_{2},\cdots,X_{m\cdot 3^{n}}

where each 0Xi2n10\leq X_{i}\leq 2^{n}-1.

Fruthermore, X1,X2,,XmX_{1},X_{2},\cdots,X_{m} corresponds to the measurement σXn\sigma_{X}^{\otimes n}; Xm+1,Xm+2,,X2mX_{m+1},X_{m+2},\cdots,X_{2m} corresponds to the measurement σXn1σY\sigma_{X}^{\otimes n-1}\otimes\sigma_{Y}; \cdots

It is direct to observe that, for any PP of weight ww, μP=j=0m3nw1Zj\mu_{P}=\sum_{j=0}^{m\cdot 3^{n-w}-1}Z_{j}, where ZjZ_{j} are independent samples from distribution ZZ

Pr(Z=1)=1+tr(ρP)2\displaystyle\mathrm{Pr}(Z=1)=\frac{1+\operatorname{tr}(\rho P)}{2}
Pr(Z=1)=1tr(ρP)2.\displaystyle\mathrm{Pr}(Z=-1)=\frac{1-\operatorname{tr}(\rho P)}{2}.

Furthermore, ZjZ_{j} can be obtained from samples

X1,X2,,Xm3n.\displaystyle X_{1},X_{2},\cdots,X_{m\cdot 3^{n}}.

Therefore,

σ=PμPm3nwP2nP.\displaystyle\sigma=\sum_{P}\frac{\mu_{P}}{m\cdot 3^{n-w_{P}}\cdot 2^{n}}P.

is defined according to samples

X1,X2,,Xm3n.\displaystyle X_{1},X_{2},\cdots,X_{m\cdot 3^{n}}.

It is direct to verify that

𝔼σ=ρ,\displaystyle\mathbb{E}\sigma=\rho,

where the expectation is taken over the probabilistic distribution according to the measurements.

For any

ρ=PαP2nP\displaystyle\rho=\sum_{P}\frac{\alpha_{P}}{2^{n}}P

we can define the function f:X1×X2××Xm3nf:X_{1}\times X_{2}\times\cdots\times X_{m\cdot 3^{n}}\mapsto\mathbb{R}

f=ρσ2\displaystyle f=||\rho-\sigma||_{2}

According to Cauchy inequality, we have

𝔼f\displaystyle\mathbb{E}f
\displaystyle\leq 𝔼f2\displaystyle\sqrt{\mathbb{E}f^{2}}
=\displaystyle= 𝔼(trρ22trρσ+trσ2)\displaystyle\sqrt{\mathbb{E}(\operatorname{tr}\rho^{2}-2\operatorname{tr}\rho\sigma+\operatorname{tr}\sigma^{2})}
=\displaystyle= 𝔼trσ2trρ2\displaystyle\sqrt{\mathbb{E}\operatorname{tr}\sigma^{2}-\operatorname{tr}\rho^{2}}
=\displaystyle= 12nP(𝔼μP2m29nwPαP2)\displaystyle\sqrt{\frac{1}{2^{n}}\sum_{P}(\mathbb{E}\frac{\mu_{P}^{2}}{m^{2}\cdot 9^{n-w_{P}}}-\alpha_{P}^{2})}
=\displaystyle= 1m2nP1αP23nwP\displaystyle\sqrt{\frac{1}{m\cdot 2^{n}}\cdot{\sum_{P}\frac{1-\alpha_{P}^{2}}{3^{n-w_{P}}}}}
<\displaystyle< 1m2nP13nwP\displaystyle\sqrt{\frac{1}{m\cdot 2^{n}}\cdot{\sum_{P}\frac{1}{3^{n-w_{P}}}}}
=\displaystyle= 1m2nwP=0n13nwP(nwP)3wP\displaystyle\sqrt{\frac{1}{m\cdot 2^{n}}\cdot{\sum_{w_{P}=0}^{n}\frac{1}{3^{n-w_{P}}}{{n}\choose{w_{P}}}3^{w_{P}}}}
=\displaystyle= 1m6nwP=0n(1+9)n\displaystyle\sqrt{\frac{1}{m\cdot 6^{n}}\cdot\sum_{w_{P}=0}^{n}(1+9)^{n}}
=\displaystyle= 5nm3n\displaystyle\sqrt{\frac{5^{n}}{m\cdot 3^{n}}}
<\displaystyle< ϵ22n.\displaystyle\frac{\epsilon}{2\cdot\sqrt{2^{n}}}.

For any sample XiX_{i} corresponding to P{σX,σY,σZ}nP\in\{\sigma_{X},\sigma_{Y},\sigma_{Z}\}^{\otimes n}. If only XiX_{i} is changed, it would only change μQ\mu_{Q} for those Q{σI,σX,σY,σZ}nQ\in\{\sigma_{I},\sigma_{X},\sigma_{Y},\sigma_{Z}\}^{\otimes n} where QQ is obtained by replacing some {σX,σY,σZ}\{\sigma_{X},\sigma_{Y},\sigma_{Z}\}s of PP by σI\sigma_{I}. Moreover, those μQ\mu_{Q} would change at most 22. According to triangle inequality, ff would change at most

QΔμQm3nwQ2nQ2\displaystyle||\sum_{Q}\frac{\Delta\mu_{Q}}{m\cdot 3^{n-w_{Q}}\cdot 2^{n}}Q||_{2}
=\displaystyle= QΔμQ2m29nwQ2n\displaystyle\sqrt{\sum_{Q}\frac{\Delta\mu_{Q}^{2}}{m^{2}\cdot 9^{n-w_{Q}}\cdot 2^{n}}}
\displaystyle\leq Q22m29nwQ2n\displaystyle\sqrt{\sum_{Q}\frac{2^{2}}{m^{2}\cdot 9^{n-w_{Q}}\cdot 2^{n}}}
=\displaystyle= wQ=0n22m29nwQ2n(nwQ)\displaystyle\sqrt{\sum_{w_{Q}=0}^{n}\frac{2^{2}}{m^{2}\cdot 9^{n-w_{Q}}\cdot 2^{n}}{{n}\choose{w_{Q}}}}
=\displaystyle= 25nm3n,\displaystyle\frac{2\cdot\sqrt{5}^{n}}{m\cdot 3^{n}},

where QQ ranges over all Paulis which corresponds to PP, and δμQ\delta\mu_{Q} denotes the difference of μQ\mu_{Q} when XiX_{i} is changed.

For ϵ>0\epsilon>0, let ϵ=ϵ2n\epsilon^{\prime}=\frac{\epsilon}{\sqrt{2^{n}}}. We have

Pr(ρσ1>ϵ)\displaystyle\mathrm{Pr}(||\rho-\sigma||_{1}>\epsilon)
\displaystyle\leq Pr(ρσ2>ϵ)\displaystyle\mathrm{Pr}(||\rho-\sigma||_{2}>\epsilon^{\prime})
=\displaystyle= Pr(f>ϵ)\displaystyle\mathrm{Pr}(f>\epsilon^{\prime})
\displaystyle\leq Pr(f𝔼f>ϵ2)\displaystyle\mathrm{Pr}(f-\mathbb{E}f>\frac{\epsilon^{\prime}}{2})
<\displaystyle< exp(ϵ2445nm29nm3n)\displaystyle\exp(-\frac{\epsilon^{\prime 2}}{4\cdot\frac{4\cdot 5^{n}}{m^{2}\cdot 9^{n}}\cdot m\cdot 3^{n}})
=\displaystyle= exp(m3nϵ2165n)\displaystyle\exp(-\frac{m\cdot 3^{n}\cdot\epsilon^{\prime 2}}{16\cdot 5^{n}})
<\displaystyle< δ,\displaystyle\delta,

where the first step is by the relation between the trace norm and 2 norm; the third step is by 𝔼fϵ2\mathbb{E}f\leq\frac{\epsilon^{\prime}}{2}; the fourth step is by Lemma 5 (McDiarmid’s inequality).

The total number of used copies is

m3n=1610nlog1δϵ2.\displaystyle m\cdot 3^{n}=16\cdot\frac{10^{n}\log\frac{1}{\delta}}{\epsilon^{2}}.

IV Quantum Overlapping Tomography

In this section, we study the quantum overlapping tomography.

We first analyze Algorithm 1.

Fixing nn informationally complete POVM of one-qubit system, says 1,2,,n\mathcal{M}_{1},\mathcal{M}_{2},\cdots,\mathcal{M}_{n}. For any given unknown nn-qubit state ρ\rho, we perform 12n\mathcal{M}_{1}\otimes\mathcal{M}_{2}\otimes\cdots\otimes\mathcal{M}_{n} on ρ\rho and obtain a string s1s2sns_{1}s_{2}\cdots s_{n} with sis_{i} denoting the measurement outcome of i\mathcal{M}_{i}. Assume the output probability distribution is p1,2,,np_{1,2,\cdots,n}. we know that s1s2sns_{1}s_{2}\cdots s_{n} obeys the distribution of p1,2,,np_{1,2,\cdots,n}.

The key observation is that measuring each qubit independently preserves the local structure. More precisely,

Observation 4.

For any S={i1,i2,,ir}{1,2,,n}S=\{i_{1},i_{2},\cdots,i_{r}\}\subseteq\{1,2,\cdots,n\}, the distribution of si1si2sirs_{i_{1}}s_{i_{2}}\cdots s_{i_{r}} obeys the outcome distribution of performing i1i2ir\mathcal{M}_{i_{1}}\otimes\mathcal{M}_{i_{2}}\otimes\cdots\otimes\mathcal{M}_{i_{r}} on ρS\rho_{S}. Moreover, si1si2sirs_{i_{1}}s_{i_{2}}\cdots s_{i_{r}} obeys the marginal distribution pSp_{S}.

By Algorithm 1, we repeat the measurement 12n\mathcal{M}_{1}\otimes\mathcal{M}_{2}\otimes\cdots\otimes\mathcal{M}_{n} mm times to obtain mm samples of p1,2,,np_{1,2,\cdots,n}. For any sets S1,S2,,SM{1,2,,n}S_{1},S_{2},\cdots,S_{M}\subseteq\{1,2,\cdots,n\}, Obervation 3 implies that we have mm samples of the marginal distributions pS1p_{S_{1}}, pS2p_{S_{2}}\cdots, and pSMp_{S_{M}}, not independent. Although, mm samples maybe not enough to recover p1,2,,np_{1,2,\cdots,n}, but enough to recover pSip_{S_{i}} for each SiS_{i} with high accuracy and high confidence using Chernoff bound. According to Observation 1 and Observation 2, this implies one can recover ρSi\rho_{S_{i}} with high accuracy, trace distance error ϵ\epsilon, and sucessful probability at least 1δ1-\delta^{\prime} for each SiS_{i} using m=𝒪(g(k)logδϵ2)m=\mathcal{O}(g(k)\cdot\frac{\log\delta^{\prime}}{\epsilon^{2}}).

To successfully recover ρSi\rho_{S_{i}} simultaneously, we only need the recovery of each ρSi\rho_{S_{i}} with high accuracy and probability at least 1δ(nk)1-\frac{\delta}{{{n}\choose{k}}} according to union bound.

Therefore, we only need m=𝒪(g(k)log(nk)/δϵ2)m=\mathcal{O}(g(k)\cdot\frac{\log{{{n}\choose{k}}}/{\delta}}{\epsilon^{2}}) copies.

IV.1 Joint Measurement Lower bound

In this subsection, we show that Ω(log(n/δ)ϵ2)\Omega(\frac{\log(n/\delta)}{\epsilon^{2}}) copies are necessary to quantum overlapping tomography by proving that: Ω(log(n/δ)ϵ2)\Omega(\frac{\log(n/\delta)}{\epsilon^{2}}) copies are necessary for quantum overlapping tomography with k=1k=1, even if general joint measurements are used.

Because trace distance is non-increasing according to partial trace, we can conclude that any measurement scheme can solve the quantum overlapping tomography problem for k1k\geq 1, automatically, it solves the case that k=1k=1. Moreover, to deal with general joint measurement schemes, we focus on classical distributions.

We first consider a simple question: Given a binary random variable XX obeys distribution q0=(1/2ϵ,1/2+ϵ)q_{0}=(1/2-\epsilon,1/2+\epsilon) or q1=(1/2+ϵ,1/2ϵ)q_{1}=(1/2+\epsilon,1/2-\epsilon). The goal is to find out which distribution is the true distribution. It is widely known that: For any fixed mm as the number of tossing this coin, the best strategy is to toss the coin mm times and declare the index (0 or 11) that appears less.

Let the X1,X2,,XmX_{1},X_{2},\dots,X_{m} be the mm samples of q1q_{1} and any 0t2mϵ0\leq t\leq 2m\epsilon, Ref. Mousavi (2016) proved the following:

Pr(i=1mXi>t+m(1/2ϵ))14exp(2t2m(1/2ϵ))\displaystyle\mathrm{Pr}(\sum_{i=1}^{m}X_{i}>t+m(1/2-\epsilon))\geq\frac{1}{4}\cdot\exp(-\frac{2t^{2}}{m(1/2-\epsilon)}) (5)

By choosing t=mϵt=m\epsilon, Pr(i=1mXi>t+m(1/2ϵ))=Pr(i=1mXi>m/2)\mathrm{Pr}(\sum_{i=1}^{m}X_{i}>t+m(1/2-\epsilon))=\mathrm{Pr}(\sum_{i=1}^{m}X_{i}>m/2) is the probability of answering q0q_{0}, a lower bound of the failure probability.

To success with probability at least 1δ1-\delta^{\prime}, we must have

δ14exp(2mϵ21/2ϵ)\displaystyle\delta^{\prime}\geq\frac{1}{4}\cdot\exp(-\frac{2m\epsilon^{2}}{1/2-\epsilon}) (6)

That is, 12ϵ4ϵ2log(14δ)\frac{1-2\epsilon}{4\epsilon^{2}}\log(\frac{1}{4\delta^{\prime}}) samples are needed to distinguish q0q_{0} and q1q_{1} with confidence at least 1δ1-\delta^{\prime}.

Back to our problem of showing Ω(log(n/δ)ϵ2)\Omega(\frac{\log(n/\delta)}{\epsilon^{2}}) samples of nn-qubit state ρ\rho are necessary to solve the quantum overlapping tomography for k1k\geq 1, to within additive error ϵ\epsilon and confidence at least 1δ1-\delta, we consider the classical distributions pi1,i2,,in=qi1qi2qinp_{i_{1},i_{2},\cdots,i_{n}}=q_{i_{1}}\otimes q_{i_{2}}\otimes\cdots\otimes q_{i_{n}} where q0=(1/2ϵ,1/2+ϵ)q_{0}=(1/2-\epsilon,1/2+\epsilon) and q1=(1/2+ϵ,1/2ϵ)q_{1}=(1/2+\epsilon,1/2-\epsilon). In total, there are 2n2^{n} different distributions.

Suppose there is a quantum procedure 𝒜\mathcal{A} which uses mm copies of ρ\rho to accomplish the quantum overlapping tomography with probability at least 1δ1-\delta.

Let Z1,Z2,,ZnZ_{1},Z_{2},\cdots,Z_{n} be random variables which obey uniform binary distribution. Choose each pZ1,Z2,,Znp_{Z_{1},Z_{2},\cdots,Z_{n}} with probability 1/2n1/2^{n}, and apply 𝒜\mathcal{A} on mm copies of pZ1,Z2,,Znp_{Z_{1},Z_{2},\cdots,Z_{n}}. Because the 1\ell_{1} norm is non-increasing under partial trace, we know that according to the output of 𝒜\mathcal{A}, we can sucessfully recover the indices Z1,Z2,,ZnZ_{1},Z_{2},\cdots,Z_{n} with probability at least 1δ1-\delta.

In the following, we first observe that quantum procedure does not help in recovering Z1,Z2,,ZnZ_{1},Z_{2},\cdots,Z_{n} from samples of pZ1,Z2,,Znp_{Z_{1},Z_{2},\cdots,Z_{n}}. We assume the joint measurement (M0,0,,0,,M1,1,,1)(M_{0,0,\cdots,0},\cdots,M_{1,1,\cdots,1}) applied on mm copies (samples) of pp such that the measurement outcome Mi1,i2,,inM_{i_{1},i_{2},\cdots,i_{n}} allows us to answer Z1=i1,Z2=i2,,Zn=inZ_{1}=i_{1},Z_{2}=i_{2},\cdots,Z_{n}=i_{n}. Here Mi1,i2,,inM_{i_{1},i_{2},\cdots,i_{n}}s are 2mn×2mn2^{mn}\times 2^{mn} matrices.

We first observe that pZ1,Z2,,Znp_{Z_{1},Z_{2},\cdots,Z_{n}}s are all diagonal, so are pZ1,Z2,,Znmp_{Z_{1},Z_{2},\cdots,Z_{n}}^{\otimes m}. Then, the off diagonal elements of Mi1,i2,,inM_{i_{1},i_{2},\cdots,i_{n}} has no effect for this task. Therefore, we only need to consider the procedure in the following two steps: At the first step, measure mm copies of pZ1,Z2,,Znp_{Z_{1},Z_{2},\cdots,Z_{n}}s in the diagonal basis; at the second step, output according to some probability distributions.

The first step ensures that we only need to measure each copy of pZ1,Z2,,Znp_{Z_{1},Z_{2},\cdots,Z_{n}}s in the diagonal basis, as there is no difference. By the convexity of the successful probability, we know that deterministic function works best in the second step, that is declare the index (0 or 11) that appears less for each 1jn1\leq j\leq n.

Now, we assume the output random virable is Y1,Y2,,YnY_{1},Y_{2},\cdots,Y_{n}, our goal is

Pr(Y1=Z1,Y2=Z2,,Yn=Zn)1δ.\displaystyle\mathrm{Pr}(Y_{1}=Z_{1},Y_{2}=Z_{2},\cdots,Y_{n}=Z_{n})\geq 1-\delta. (7)

By Bayes’ theorem, we know that

Pr(Y1=Z1,Y2=Z2,,Yn=Zn)\displaystyle\mathrm{Pr}(Y_{1}=Z_{1},Y_{2}=Z_{2},\cdots,Y_{n}=Z_{n})
=\displaystyle= Pr(Y1=Z1)×Pr(Y2=Z2|Y1=Z1)××Pr(Yn=Zn|Y1=Z1,,Yn1=Zn1)\displaystyle\mathrm{Pr}(Y_{1}=Z_{1})\times\mathrm{Pr}(Y_{2}=Z_{2}|Y_{1}=Z_{1})\times\cdots\times\mathrm{Pr}(Y_{n}=Z_{n}|Y_{1}=Z_{1},\cdots,Y_{n-1}=Z_{n-1})
\displaystyle\leq (1δ)(1δ)(1δ)\displaystyle(1-\delta^{\prime})\cdot(1-\delta^{\prime})\cdot\cdots\cdot(1-\delta^{\prime})
=\displaystyle= (1δ)m,\displaystyle(1-\delta^{\prime})^{m},

where we use the fact that pZ1,Z2,,Znp_{Z_{1},Z_{2},\cdots,Z_{n}}s are all in tensor product form, therefore, Pr(Y2=Z2|Y1=Z1)=Pr(Y2=Z2)\mathrm{Pr}(Y_{2}=Z_{2}|Y_{1}=Z_{1})=\mathrm{Pr}(Y_{2}=Z_{2}), and so on. (1δ)(1-\delta^{\prime}) denotes the sucessful probability of discriminate q0q_{0} and q1q_{1} with mm copies.

Therefore, we require that

(1δ)n>1δ.\displaystyle(1-\delta^{\prime})^{n}>1-\delta.

That is δ=Θ(δn)\delta^{\prime}=\Theta(\frac{\delta}{n}), this implies the bound m=Ω(log(n/δ)ϵ2)m=\Omega(\frac{\log(n/\delta)}{\epsilon^{2}}).

IV.2 Pauli Measurement Upper bound

In this subsection, we analyze Algorithm 1 using the measurement scheme {M0,M1,M2,M3,M4,M5}n\{M_{0},M_{1},M_{2},M_{3},M_{4},M_{5}\}^{\otimes n} as a special case, where

M0=σI+σX6,M1=σIσX6,\displaystyle M_{0}=\frac{\sigma_{I}+\sigma_{X}}{6},\ \ \ \ \ M_{1}=\frac{\sigma_{I}-\sigma_{X}}{6},
M2=σI+σY6,M3=σIσY6,\displaystyle M_{2}=\frac{\sigma_{I}+\sigma_{Y}}{6},\ \ \ \ \ M_{3}=\frac{\sigma_{I}-\sigma_{Y}}{6},
M4=σI+σZ6,M5=σIσZ6.\displaystyle M_{4}=\frac{\sigma_{I}+\sigma_{Z}}{6},\ \ \ \ \ M_{5}=\frac{\sigma_{I}-\sigma_{Z}}{6}.

This can be regarded as a random chosen of P{σX,σY,σZ}P\in\{\sigma_{X},\sigma_{Y},\sigma_{Z}\}^{\otimes} and measure ρ\rho in the basis of PP, and repeat it mm times. That is,

1Repeat the following measurement 3210kϵ2log(2(nk)/δ)32\cdot 10^{k}\cdot\epsilon^{-2}\cdot\log(2{{n}\choose{k}}/\delta) times;
2 For i=1i=1 to nn: measure the ii-th qubit in a random chosen basis from{σX,σY,σZ}\{\sigma_{X},\sigma_{Y},\sigma_{Z}\};
Algorithm 2 Quantum Overlapping Tomography by Pauli Measurements

We observe the following,

Observation 5.

For any S={i1,i2,,ik}{1,2,,n}S=\{i_{1},i_{2},\cdots,i_{k}\}\subseteq\{1,2,\cdots,n\}, with probability at least 12mem3k1-\frac{2^{m}}{e^{m}}\cdot 3^{k}, each P{σX,σY,σZ}kP\in\{\sigma_{X},\sigma_{Y},\sigma_{Z}\}^{\otimes k} was measured for ρS\rho_{S} at least m=1610klog(2(nk)/δ)3kϵ2m=16\cdot\frac{10^{k}\cdot\log(2{{n}\choose{k}}/\delta)}{3^{k}\cdot\epsilon^{2}} times.

Proof.

It is direct to observe that for any P{σX,σY,σZ}kP\in\{\sigma_{X},\sigma_{Y},\sigma_{Z}\}^{\otimes k}, it was chosen as a basis to measure is no more than mm times with probability at most

i=0m(2m3ki)(113k)2m3ki13ki\displaystyle\sum_{i=0}^{m}{{2m\cdot 3^{k}}\choose{i}}(1-\frac{1}{3^{k}})^{2m\cdot 3^{k}-i}\frac{1}{3^{ki}}
=\displaystyle= (113k)2m3ki=0m(2m3ki)(113k)i13ki\displaystyle(1-\frac{1}{3^{k}})^{2m\cdot 3^{k}}\cdot\sum_{i=0}^{m}{{2m\cdot 3^{k}}\choose{i}}(1-\frac{1}{3^{k}})^{-i}\frac{1}{3^{ki}}
=\displaystyle= (113k)2m3ki=0m(2m3ki)1(3k1)i\displaystyle(1-\frac{1}{3^{k}})^{2m\cdot 3^{k}}\cdot\sum_{i=0}^{m}{{2m\cdot 3^{k}}\choose{i}}\frac{1}{(3^{k}-1)^{i}}

The ratio of the i+1i+1-th term and the ii-th term is

13k2m3kii+12\displaystyle\frac{1}{3^{k}}\cdot\frac{2m\cdot 3^{k}-i}{i+1}\geq 2

Therefore, the serie grows faster than a geometric progression with common ratio 22. Then,

(113k)2m3ki=0m(2m3ki)1(3k1)i\displaystyle(1-\frac{1}{3^{k}})^{2m\cdot 3^{k}}\cdot\sum_{i=0}^{m}{{2m\cdot 3^{k}}\choose{i}}\frac{1}{(3^{k}-1)^{i}}
\displaystyle\leq 2(113k)2m3k(2m3km)1(3k1)m\displaystyle 2\cdot(1-\frac{1}{3^{k}})^{2m\cdot 3^{k}}{{2m\cdot 3^{k}}\choose{m}}\frac{1}{(3^{k}-1)^{m}}
\displaystyle\leq 323k2πm(23k1)[2(1123k1)23k1]m\displaystyle 3\sqrt{\frac{2\cdot 3^{k}}{2\pi\cdot m\cdot(2\cdot 3^{k}-1)}}[2(1-\frac{1}{2\cdot 3^{k}-1})^{2\cdot 3^{k}-1}]^{m}
\displaystyle\leq 323k2πm(23k1)2mem\displaystyle 3\sqrt{\frac{2\cdot 3^{k}}{2\pi\cdot m\cdot(2\cdot 3^{k}-1)}}\frac{2^{m}}{e^{m}}
\displaystyle\leq 2mem.\displaystyle\frac{2^{m}}{e^{m}}.

where we use the bound that

2πnnnene112n<n!<2πnnnene112n+1\displaystyle\sqrt{2\pi n}\cdot n^{n}\cdot e^{-n}\cdot e^{\frac{1}{12n}}<n!<\sqrt{2\pi n}\cdot n^{n}\cdot e^{-n}\cdot e^{\frac{1}{12n+1}}

and the number series

(11d)d\displaystyle(1-\frac{1}{d})^{d}

is increasing and goes to 1e\frac{1}{e}.

Then, by union bound, we know that for any SS of size kk,with probability at least 12mem3k1-\frac{2^{m}}{e^{m}}\cdot 3^{k}, each P{σX,σY,σZ}kP\in\{\sigma_{X},\sigma_{Y},\sigma_{Z}\}^{\otimes k} was measured for ρS\rho_{S} at least mm times. ∎

According to Theorem 1, the tomography of ρS\rho_{S} with trace distance error ϵ\epsilon was successful with probability at least

1[3k2mem+δ2(nk)]>1δ(nk),\displaystyle 1-[3^{k}\cdot\frac{2^{m}}{e^{m}}+\frac{\delta}{2{{n}\choose{k}}}]>1-\frac{\delta}{{{n}\choose{k}}},

the last inequality follows from

3k2mem\displaystyle 3^{k}\cdot\frac{2^{m}}{e^{m}}
<\displaystyle< 3k(2e)123k(2e)43klog(2(nk)/δ)\displaystyle 3^{k}(\frac{2}{e})^{12\cdot 3^{k}}\cdot(\frac{2}{e})^{4\cdot 3^{k}\cdot\log(2{{n}\choose{k}}/\delta)}
<\displaystyle< 3k(2e)12k(1e)log(2(nk)/δ)\displaystyle 3^{k}(\frac{2}{e})^{12\cdot k}\cdot(\frac{1}{e})^{\log(2{{n}\choose{k}}/\delta)}
<\displaystyle< δ2(nk).\displaystyle\frac{\delta}{2{{n}\choose{k}}}.

By union bound, the quantum overlapping tomography was with trace distance error ϵ\epsilon was successful with probability at least

1(nk)δ(nk)=1δ\displaystyle 1-{{n}\choose{k}}\cdot\frac{\delta}{{{n}\choose{k}}}=1-\delta

V Discussion

For the general quantum state tomography problem, our measurement scheme is much more efficient than previous schemes in the Pauli measurements setting. Our result raises an important question on the performance of local measurements, in particular Pauli measurements, in which each qubit is measured at a time.

The sample complexity of the quantum overlapping tomography problem is nearly resolved here up to a constant factor.

An independent work Huang et al. (2020) analyzes random Clifford measurement and random Pauli measurement and proves that it only requires 𝒪(log(M)sϵ2)\mathcal{O}(\frac{\log(M)\cdot s}{\epsilon^{2}}) copies to achieve ϵ\epsilon accuracy estimation for tr(O1ρ),,tr(OMρ)\operatorname{tr}(O_{1}\rho),\cdots,\operatorname{tr}(O_{M}\rho), a more general question, where ss depends on the structure of OiO_{i}s. They also show that this is tight if only single-copy measurements are allowed. If OiO_{i}s are kk-qubit Paulis, 𝒪(log(M/δ)3kϵ2)\mathcal{O}(\frac{\log(M/\delta)\cdot 3^{k}}{\epsilon^{2}}) which coinsides with the bound provided in Ref. Evans et al. (2019). These upper bound does not cover ours, because a 4k4^{k} coefficient would be added when we move from Pauli estimation to state tomography, thus, these results would lead to a 𝒪(log(M/δ)12kϵ2)\mathcal{O}(\frac{\log(M/\delta)\cdot 12^{k}}{\epsilon^{2}}) measurement scheme. Also, this lower bound does not imply our lower bound, for the specific quantum overlapping tomography problem, because this bound is not for joint measurement and does not consider the successful probability parameter.

VI Acknowledgments

We thank Steve Flammia for the discussion on quantum overlapping tomography. We thank Richard Keung for clasifying their bound in Guţă et al. (2020). This work was supported by DE180100156.

References

  • Cotler and Wilczek (2020) Jordan Cotler and Frank Wilczek, “Quantum overlapping tomography,” Physical Review Letters 124 (2020), 10.1103/physrevlett.124.100401.
  • Bagan et al. (2004) E. Bagan, M. Baig, R. Muñoz Tapia,  and A. Rodriguez, “Collective versus local measurements in a qubit mixed-state estimation,” Phys. Rev. A 69, 010304 (2004).
  • Keyl (2006) M. Keyl, “Quantum state estimation and large deviations,” Reveiws in Mathematical Physics 18, 19–60 (2006).
  • Guţă and Kahn (2008) Mădălin Guţă and Jonas Kahn, “Optimal estimation of qubit states with continuous time measurements,” Communications in Mathematical Physics 277, 127–160 (2008)quant-ph/0608074 .
  • Haah et al. (2016) J. Haah, A. W. Harrow, Z. Ji, X. Wu, ,  and N. Yu, “Sample-optimal tomography of quantum states,” in Proceedings of the Forty-Seventh Annual ACM on Symposium on Theory of Computing, STOC ’16 (2016) pp. 913–925.
  • O’Donnell and Wright (2016) R. O’Donnell and J. Wright, “Efficient quantum tomography,” in Proceedings of the Forty-Seventh Annual ACM on Symposium on Theory of Computing, STOC ’16 (2016) pp. 899–912.
  • O’Donnell and Wright (2017) R. O’Donnell and J. Wright, “Efficient quantum tomography ii,” in Proceedings of the Forty-Seventh Annual ACM on Symposium on Theory of Computing, STOC ’17 (2017) pp. 962–974.
  • Flammia et al. (2012a) S. T. Flammia, D. Gross, Y. Liu,  and J. Eisert, “Quantum tomography via compressed sensing: Error bounds, sample complexity, and efficient estimators,” New J. Phys. 14, 095022 (2012a).
  • Kueng et al. (2017) R. Kueng, H. Rauhut,  and U. Terstiege, “Low rank matrix recovery from rank one measurements,” Applied and Computational Harmonic Analysis 42, 88–116 (2017).
  • Guţă et al. (2020) M Guţă, J Kahn, R Kueng,  and J A Tropp, “Fast state tomography with optimal error bounds,” Journal of Physics A: Mathematical and Theoretical 53, 204001 (2020).
  • Flammia et al. (2012b) Steven T Flammia, David Gross, Yi-Kai Liu,  and Jens Eisert, “Quantum tomography via compressed sensing: error bounds, sample complexity and efficient estimators,” New Journal of Physics 14, 095022 (2012b).
  • García-Pérez et al. (2020) Guillermo García-Pérez, Matteo A. C. Rossi, Boris Sokolov, Elsi-Mari Borrelli,  and Sabrina Maniscalco, “Pairwise tomography networks for many-body quantum systems,” Physical Review Research 2 (2020), 10.1103/physrevresearch.2.023393.
  • Evans et al. (2019) Tim J. Evans, Robin Harper,  and Steven T. Flammia, “Scalable bayesian hamiltonian learning,”  (2019), arXiv:1912.07636 [quant-ph] .
  • McDiarmid (1989) Colin McDiarmid, “On the method of bounded differences,” in Surveys in Combinatorics, 1989: Invited Papers at the Twelfth British Combinatorial Conference, London Mathematical Society Lecture Note Series, edited by J.Editor Siemons (Cambridge University Press, 1989) p. 148–188.
  • Mousavi (2016) Nima Mousavi, “How tight is chernoff bound?”   (2016), https://ece.uwaterloo.ca/ nmousavi/Papers/Chernoff-Tightness.pdf .
  • Huang et al. (2020) Hsin-Yuan Huang, Richard Kueng,  and John Preskill, “Predicting many properties of a quantum system from very few measurements,” Nature Physics  (2020), 10.1038/s41567-020-0932-7.