scispace - formally typeset
Open AccessJournal ArticleDOI

Predicting many properties of a quantum system from very few measurements

Reads0
Chats0
TLDR
In this article, the authors presented a method for constructing an approximate classical description of a quantum state using very few measurements of the state, called a "classical shadow", which can be used to predict many different properties.
Abstract
Predicting the properties of complex, large-scale quantum systems is essential for developing quantum technologies. We present an efficient method for constructing an approximate classical description of a quantum state using very few measurements of the state. This description, called a ‘classical shadow’, can be used to predict many different properties; order $${\mathrm{log}}\,(M)$$ measurements suffice to accurately predict M different functions of the state with high success probability. The number of measurements is independent of the system size and saturates information-theoretic lower bounds. Moreover, target properties to predict can be selected after the measurements are completed. We support our theoretical findings with extensive numerical experiments. We apply classical shadows to predict quantum fidelities, entanglement entropies, two-point correlation functions, expectation values of local observables and the energy variance of many-body local Hamiltonians. The numerical results highlight the advantages of classical shadows relative to previously known methods. An efficient method has been proposed through which the properties of a complex, large-scale quantum system can be predicted without fully characterizing the quantum state.

read more

Content maybe subject to copyright    Report

Predicting Many Properties of a Quantum System from Very Few Measurements
Hsin-Yuan Huang,
1, 2,
Richard Kueng,
1, 2, 3
and John Preskill
1, 2, 4
1
Institute for Quantum Information and Matter, Caltech, Pasadena, CA, USA
2
Department of Computing and Mathematical Sciences, Caltech, Pasadena, CA, USA
3
Institute for Integrated Circuits, Johannes Kepler University Linz, Austria
4
Walter Burke Institute for Theoretical Physics, Caltech, Pasadena, CA, USA
(Dated: April 23, 2020)
Predicting properties of complex, large-scale quantum systems is essential for developing quantum
technologies. We present an efficient method for constructing an approximate classical description
of a quantum state using very few measurements of the state. This description, called a classical
shadow, can be used to predict many different properties: order log M measurements suffice to
accurately predict M different functions of the state with high success probability. The number of
measurements is independent of the system size, and saturates information-theoretic lower bounds.
Moreover, target properties to predict can be selected after the measurements are completed. We
support our theoretical findings with extensive numerical experiments. We apply classical shadows
to predict quantum fidelities, entanglement entropies, two-point correlation functions, expectation
values of local observables, and the energy variance of many-body local Hamiltonians. The numerical
results highlight the advantages of classical shadows relative to previously known methods.
Making predictions based on empirical observations is a central topic in statistical learning theory and is
at the heart of many scientific disciplines, including quantum physics. There, predictive tasks, like estimating
target fidelities, verifying entanglement, and measuring correlations, are essential for building, calibrating and
controlling quantum systems. Recent advances in the size of quantum platforms [59] have pushed traditional
prediction techniques like quantum state tomography to the limit of their capabilities. This is mainly due
to a curse of dimensionality: the number of parameters needed to describe a quantum system scales exponen-
tially with the number of its constituents. Moreover, these parameters cannot be accessed directly, but must
be estimated by measuring the system. An informative quantum mechanical measurement is both destructive
(wave-function collapse) and only yields probabilistic outcomes (Born’s rule). Hence, many identically prepared
samples are required to estimate accurately even a single parameter of the underlying quantum state. Further-
more, all of these measurement outcomes must be processed and stored in memory for subsequent prediction of
relevant features. In summary, reconstructing a full description of a quantum system with n constituents (e.g.
qubits) necessitates a number of measurement repetitions exponential in n, as well as an exponential amount
of classical memory and computing power.
Several approaches have been proposed to overcome this fundamental scaling problem. These include matrix
product state (MPS) tomography [18] and neural network tomography [15, 69]. Both only require a polynomial
number of samples, provided that the underlying state has suitable properties. However, for general quantum
systems, these techniques still require an exponential number of samples. We refer to the related work section
(Supplementary Section 3) for details.
Pioneering a conceptually very different line of research, Aaronson [1] pointed out that demanding full classical
descriptions of quantum systems may be excessive for many concrete tasks. Instead it is often sufficient to
accurately predict certain properties of the quantum system. In quantum mechanics, interesting properties
are often linear functions of the underlying density matrix ρ, such as the expectation values {o
i
} of a set of
observables {O
i
}:
o
i
(ρ) =trace(O
i
ρ) 1 i M. (1)
The fidelity with a pure target state, entanglement witnesses, and the probability distribution governing the
possible outcomes of a measurement are all examples that fit this framework. A nonlinear function of ρ such
as entanglement entropy, may also be of interest. Aaronson coined the term [1, 3] shadow tomography
1
for the
task of predicting properties without necessarily fully characterizing the quantum state, and he showed that a
polynomial number of state copies already suffice to predict an exponential number of target functions. While
very efficient in terms of samples, Aaronson’s procedure is very demanding in terms of quantum hardware
a concrete implementation of the proposed protocol requires exponentially long quantum circuits that act
collectively on all the copies of the unknown state stored in a quantum memory.
In this work, we combine the mindset of shadow tomography [1] (predict target functions, not the full state)
with recent insights from quantum state tomography [35] (rigorous statistical convergence guarantees) and
Electronic address: hsinyuan@caltech.edu
1
According to Ref. [1] it was actually S.T. Flammia who originally suggested the name shadow tomography.
arXiv:2002.08953v2 [quant-ph] 22 Apr 2020

2
Measurements
Few Repetitions
Predicting …
Quantum System
Local Observables
Entanglement
Entropy
2-point Correlations
Hamiltonian
Possible Properties
Data Acquisition Phase Prediction Phase
Quantum Fidelity
Entanglement
Witness
Unitary
Evolution
Random
Unitary
Classical
Representation
Figure 1: An illustration for constructing a classical representation, the classical shadow, of a quantum system from
randomized measurements. In the data acquisition phase, we perform a random unitary evolution and measurements
on independent copies of an n-qubit system to obtain a classical representation of the quantum system the classical
shadow. Such classical shadows facilitate accurate prediction of a large number of different properties using a simple
median-of-means protocol.
the stabilizer formalism [31] (efficient implementation). The result is a highly efficient protocol that learns a
minimal classical sketch S
ρ
the classical shadow of an unknown quantum state ρ that can be used to predict
arbitrary linear function values (1) by a simple median-of-means protocol. A classical shadow is created by
repeatedly performing a simple procedure: Apply a unitary transformation ρ 7→ U ρU
, and then measure all
the qubits in the computational basis. The number of times this procedure is repeated is called the size of
the classical shadow. The transformation U is randomly selected from an ensemble of unitaries, and different
ensembles lead to different versions of the procedure that have characteristic strengths and weaknesses. In
a practical scheme, each ensemble unitary should be realizable as an efficient quantum circuit. We consider
random n-qubit Clifford circuits and tensor products of random single-qubit Clifford circuits as important
special cases. These two procedures turn out to complement each other nicely. We refer to Figure 1 for a
visualization and a list of important properties that can be predicted efficiently.
Our main theoretical contribution equips this procedure with rigorous performance guarantees. Classical
shadows with size of order log(M) suffice to predict M target functions in Eq. (1) simultaneously. Most impor-
tantly, the actual system size (number of qubits) does not enter directly. Instead, the number of measurement
repetitions N is determined by a (squared) norm kO
i
k
2
shadow
. This norm depends on the target functions and
the particular measurement procedure used to produce the classical shadow. For example, random n-qubit
Clifford circuits lead to the Hilbert-Schmidt norm. On the other hand, random single-qubit Clifford circuits
produce a norm that scales exponentially in the locality of target functions, but is independent of system
size. The resulting prediction technique is applicable to current laboratory experiments and facilitates the
efficient prediction of few-body properties, such as two-point correlation functions, entanglement entropy of
small subsystems, and expectation values of local observables.
In some cases, this scaling may seem unfavorable. However, we rigorously prove that this is not a flaw of the
method, but an unavoidable limitation rooted in quantum information theory. By relating the prediction task
to a communication task [25], we establish fundamental lower bounds highlighting that classical shadows are
(asymptotically) optimal.
We support our theoretical findings by conducting numerical simulations for predicting various physically
relevant properties over a wide range of system sizes. These include quantum fidelity, two-point correlation
functions, entanglement entropy, and local observables. We confirm that prediction via classical shadows scales
favorably and improves on powerful existing techniques such as machine learning in a variety of well-
motivated test cases. An open source release for predicting many properties from very few measurements is
available at https://github.com/momohuang/predicting-quantum-properties.

3
Algorithm 1 Median of means prediction based on a classical shadow S(ρ, N).
1 function LinearPredictions(O
1
, . . . , O
M
, S(ρ; N), K)
2 Import S(ρ; N) = [ˆρ
1
, . . . , ˆρ
N
] Load classical shadow
3 Split the shadow into K equally-sized parts and set Construct K estimators of ρ
ˆρ
(k)
=
1
bN/Kc
kbN/Kc
X
i=(k1)bN/Kc+1
ˆρ
i
4 for i = 1 to M do
5 Output ˆo
i
(N, K) = median
tr
O
i
ˆρ
(1)
, . . . , tr
O
i
ˆρ
(K)

. Median of means estimation
PROCEDURE
Throughout this work we restrict attention to n-qubit systems and ρ is a fixed, but unknown, quantum state in
d = 2
n
dimensions. To extract meaningful information, we repeatedly perform a simple measurement procedure:
apply a random unitary to rotate the state (ρ 7→ UρU
) and perform a computational-basis measurement.
The unitary U is selected randomly from a fixed ensemble. Upon receiving the n-bit measurement outcome
|
ˆ
bi {0, 1}
n
, we store an (efficient) classical description of U
|
ˆ
bih
ˆ
b|U in classical memory. It is instructive to
view the average (over both the choice of unitary and the outcome distribution) mapping from ρ to its classical
snapshot U
|
ˆ
bih
ˆ
b|U as a quantum channel:
E
h
U
|
ˆ
bih
ˆ
b|U
i
= M(ρ) = ρ = E
h
M
1
U
|
ˆ
bih
ˆ
b|U
i
. (2)
This quantum channel M depends on the ensemble of (random) unitary transformations. Although the inverted
channel M
1
is not physical (it is not completely positive), we can still apply M
1
to the (classically stored)
measurement outcome U
|
ˆ
bih
ˆ
b|U in a completely classical post-processing step.
2
In doing so, we produce a single
classical snapshot ˆρ = M
1
U
|
ˆ
bih
ˆ
b|U
of the unknown state ρ from a single measurement. By construction,
this snapshot exactly reproduces the underlying state in expectation (over both unitaries and measurement
outcomes): E[ˆρ] = ρ. Repeating this procedure N times results in an array of N independent, classical
snapshots of ρ:
S(ρ; N) =
n
ˆρ
1
= M
1
U
1
|
ˆ
b
1
ih
ˆ
b
1
|U
1
, . . . , ˆρ
N
= M
1
U
N
|
ˆ
b
N
ih
ˆ
b
N
|U
N
o
. (3)
We call this array the classical shadow of ρ. Classical shadows of sufficient size N are expressive enough
to predict many properties of the unknown quantum state efficiently. To avoid outlier corruption, we split
the classical shadow up into equally-sized chunks and construct several, independent sample mean estimators.
Subsequently, we predict linear function values (1) via median of means estimation [41, 55]. This procedure
is summarized in Algorithm 1. For many physically relevant properties O
i
and measurement channels M,
Algorithm 1 can be carried out very efficiently without explicitly constructing the large matrix ˆρ
i
.
Median of means prediction with classical shadows can be defined for any distribution of random unitary
transformations. Two prominent examples are: (i) random n-qubit Clifford circuits; and (ii) tensor products
of random single-qubit Clifford circuits. Example (i) results in a clean and powerful theory, but also practical
drawbacks, because n
2
/ log(n) entangling gates are needed to sample from n-qubit Clifford unitaries. The
corresponding inverted quantum channel is M
1
n
(X) = (2
n
+ 1)X I. Example (ii) is equivalent to measuring
each qubit independently in a random Pauli basis. Such measurements can be routinely carried out in many
experimental platforms. The corresponding inverted quantum channel is M
1
P
=
N
n
i=1
M
1
1
. We refer to
examples (i) / (ii) as random Clifford / Pauli measurements, respectively. In both cases, the resulting classical
shadow can be stored efficiently in a classical memory using the stabilizer formalism.
RIGOROUS PERFORMANCE GUARANTEES
Theorem 1 (informal version). Classical shadows of size N suffice to predict M arbitrary linear target functions
tr(O
1
ρ), . . . , tr(O
M
ρ) up to additive error given that N (order) log(M) max
i
kO
i
k
2
shadow
/
2
. The definition
2
M is invertible if the ensemble of unitary transformations defines a tomographically complete set of measurements. See Supple-
mentary Section 1.

4
of the norm kO
i
k
shadow
depends on the ensemble of unitary transformations used to create the classical shadow.
We refer to Section 1 in the Supplementary Information for background, a detailed statement and proofs.
Theorem 1 is most powerful when the linear functions have a bounded norm that is independent of system size.
In this case, classical shadows allow for predicting a large number of properties from only a logarithmic number
of quantum measurements.
The norm kO
i
k
shadow
in Theorem 1 plays an important role in defining the space of linear functions that can
be predicted efficiently. For random Clifford measurements, kOk
2
shadow
is closely related to the Hilbert-Schmidt
norm tr(O
2
). As a result, a large collection of (global) observables with a bounded Hilbert-Schmidt norm can
be predicted efficiently. For random Pauli measurements, the norm scales exponentially in the locality of the
observable, not the actual number of qubits. For an observable O
i
that acts non-trivially on (at most) k qubits,
kO
i
k
2
shadow
4
k
kO
i
k
2
, where k·k
denotes the operator norm
3
. This guarantees the accurate prediction of
many local observables from only a much smaller number of measurements.
ILLUSTRATIVE EXAMPLE APPLICATIONS
Quantum fidelity estimation. Suppose we wish to certify that an experimental device prepares a desired
n-qubit state. Typically, this target state |ψihψ| is pure and highly structured, e.g. a a GHZ state [32] for
quantum communication protocols, or a toric code ground state [21] for fault-tolerant quantum computation.
Theorem 1 asserts that a classical shadow (Clifford measurements) of dimension-independent size suffices to
accurately predict the fidelity of any state in the lab with any pure target state. This improves on the best
existing result on direct fidelity estimation [27] which requires O(2
n
/
4
) samples in the worst case. Moreover,
a classical shadow of polynomial size allows for estimating an exponential number of (pure) target fidelities all
at once.
Entanglement verification. Fidelities with pure target states can also serve as (bipartite) entanglement
witnesses [36]. For every (bipartite) entangled state ρ, there exists a constant α and an observable O = |ψihψ|
such that tr(Oρ) > α tr(Oρ
s
), for all (bipartite) separable states ρ
s
. Establishing tr(Oρ) > α verifies the
existence of entanglement in the state ρ. Any O = |ψihψ| that satisfies the above condition is known as an
entanglement witness for the state ρ. Classical shadows (Clifford measurements) of logarithmic size allow for
checking a large number of potential entanglement witnesses simultaneously.
Predicting expectation values of local observables. Many near-term applications of quantum devices rely on
repeatedly estimating a large number of local observables. For example, low-energy eigenstates of a many-body
Hamiltonian may be prepared and studied using a variational method, in which the Hamiltonian, a sum of
local terms, is measured many times. Classical shadows constructed from a logarithmic number of random
Pauli measurements can efficiently estimate polynomially many such local observables. Because only single-
qubit Pauli measurements suffice, this measurement procedure is highly efficient. Potential applications include
quantum chemistry [43] and lattice gauge theory [46].
Predicting expectation values of global observables (non-example). Classical shadows are not without limi-
tations. In our examples, the size of classical shadows must either scale with tr(O
2
i
) (Clifford measurements)
or must scale exponentially in the locality of O
i
(Pauli measurements). Both quantities can simultaneously
become exponentially large for nonlocal observables with large Hilbert-Schmidt norm. A concrete example is
the Pauli expectation value of a spin chain: hP
i
1
··· P
i
n
i
ρ
= tr (O
1
ρ), where tr(O
2
1
) = 2
n
and k = n
(non-local observable). In this case, classical shadows of exponential size may be required to accurately predict
a single expectation value. In contrast, a direct spin measurement achieves the same accuracy with only of
order 1/
2
copies of the state ρ.
MATCHING INFORMATION-THEORETIC LOWER BOUNDS
The non-example above raises an important question: does the scaling of the required number of measure-
ments with Hilbert-Schmidt norm or with the locality of observables arise from a fundamental limitation, or
is it merely an artifact of prediction with classical shadows? A rigorous analysis reveals that this scaling is no
mere artifact; rather it stems from information-theoretic reasons.
Theorem 2 (informal version). Any procedure based on single-copy measurements, that can predict any M lin-
ear functions tr(O
i
ρ) up to additive error , requires at least (order) log(M) max
i
kO
i
k
2
shadow
/
2
measurements.
3
This scaling can be further improved to 3
k
if O
i
is a tensor product of k single-qubit observables.

5
Here, kO
i
k
2
shadow
could be taken as the Hilbert-Schmidt norm tr(O
2
i
) or as a function scaling exponentially in
the locality of O
i
. The proof results from embedding the abstract prediction procedure into a communication
protocol. Quantum information theory imposes fundamental restrictions on any quantum communication
protocol and allows us to deduce stringent lower bounds. We refer to Supplementary Section 7 and 8 for details
and proofs.
The two main technical results complement each other nicely. Theorem 1 equips classical shadows with a
constructive performance guarantee: an order of log(M ) max
i
kO
i
k
2
shadow
/
2
single-copy measurements suffice
to accurately predict an arbitrary collection of M target functions. Theorem 2 highlights that this number of
measurements is unavoidable in general.
PREDICTING NONLINEAR FUNCTIONS
The classical shadow S(ρ; N) = {ˆρ
1
, . . . , ˆρ
N
} of the unknown quantum state ρ may also be used to predict
non-linear functions f(ρ). We illustrate this with a quadratic function f (ρ) = tr(Oρ ρ), where O acts on two
copies of the state. Because ˆρ
i
is equal to the quantum state ρ in expectation, one could predict tr( ρ)
using two independent snapshots ˆρ
i
, ˆρ
j
, i 6= j. Because of independence, tr(Oˆρ
i
ˆρ
j
) correctly predicts the
quadratic function in expectation:
E tr(Oˆρ
i
ˆρ
j
) = tr(O E ˆρ
i
E ˆρ
j
) = tr(Oρ ρ). (4)
To reduce the prediction error, we use N independent snapshots and symmetrize over all possible pairs:
1
N(N1)
P
i6=j
tr(Oˆρ
i
ˆρ
j
). We then repeat this procedure several times and form their median to further
reduce the likelihood of outlier corruption (similar to median of means). Rigorous performance guarantees
are given in Supplementary Section 6. This approach readily generalizes to higher order polynomials using
U-statistics [38].
One particularly interesting nonlinear function is the second-order Rényi entanglement entropy:
log(tr(ρ
2
A
)), where A is a subsystem of the n-qubit quantum system. We can rewrite the argument in
the log as tr(ρ
2
A
) = tr (S
A
ρ ρ) where S
A
is the local swap operator of two copies of the subsystem A
and use classical shadows to obtain very accurate predictions. The required number of measurements scales
exponentially in the size of the subsystem A, but is independent of total system size. Probing this entanglement
entropy is a useful task and a highly efficient specialized approach has been proposed in [12]. We compare this
Brydges et al. method to classical shadows in the numerical experiments.
For nonlinear functions, unlike linear ones, we have have not derived an information-theoretic lower bound
on the number of measurements needed, though it may be possible to do so by generalizing our methods.
NUMERICAL EXPERIMENTS
One of the key features of prediction with classical shadows is scalability. The data acquisition phase is
designed to be tractable for state of the art platforms (Pauli measurements) and future quantum computers
(Clifford measurements), respectively. The resulting classical shadow can be stored efficiently in classical
memory. For may important features such as local observables or global features with efficient stabilizer
decompositions scalability moreover extends to the computational cost associated with median of means
prediction.
These design features allowed us to conduct numerical experiments for a wide range of problems and sys-
tem sizes (up to 160 qubits). The computational bottleneck is not feature prediction with classical shadows,
but generating synthetic data, i.e. classically generating target states and simulating quantum measurements.
Needless to say, this classical bottle-neck does not occur in actual experiments. We then use this synthetic data
to learn a classical representation of ρ and use this representation to predict various interesting properties.
Machine learning based approaches [15, 69] are among the most promising alternative methods that have
applications in this regime, where the Hilbert space dimension is roughly comparable to the total number
of silicon atoms on earth (2
160
' 10
48
). For example, a recent version of neural network quantum state
tomography (NNQST) is a generative model that is based on a deep neural network trained on independent
quantum measurement outcomes (local SIC/tetrahedral POVMs [64]). In this section, we consider the task
of learning a classical representation of an unknown quantum state, and using the representation to predict
various properties, addressing the relative merit of classical shadows and alternative methods.

Figures
Citations
More filters
Journal ArticleDOI

Noisy intermediate-scale quantum algorithms

TL;DR: In this article , the authors discuss what is possible in this ''noisy intermediate scale'' quantum (NISQ) era, including simulation of many-body physics and chemistry, combinatorial optimization, and machine learning.
Journal ArticleDOI

Power of data in quantum machine learning

TL;DR: In this article, the authors show that some problems that are classically hard to compute can be easily predicted by classical machines learning from data using rigorous prediction error bounds as a foundation, and develop a methodology for assessing potential quantum advantage in learning tasks.
References
More filters
Journal ArticleDOI

Quantum Computing in the NISQ era and beyond

TL;DR: Noisy Intermediate-Scale Quantum (NISQ) technology will be available in the near future, and the 100-qubit quantum computer will not change the world right away - but it should be regarded as a significant step toward the more powerful quantum technologies of the future.
Book ChapterDOI

A Class of Statistics with Asymptotically Normal Distribution

TL;DR: In this article, the authors considered the problem of estimating a U-statistic of the population characteristic of a regular functional function, where the sum ∑″ is extended over all permutations (α 1, α m ) of different integers, 1 α≤ (αi≤ n, n).
Journal ArticleDOI

Hardware-efficient variational quantum eigensolver for small molecules and quantum magnets

TL;DR: The experimental optimization of Hamiltonian problems with up to six qubits and more than one hundred Pauli terms is demonstrated, determining the ground-state energy for molecules of increasing size, up to BeH2.
Journal ArticleDOI

Many-Body Localization and Thermalization in Quantum Statistical Mechanics

TL;DR: In this paper, the authors provide a brief introduction to quantum thermalization, paying particular attention to the eigenstate thermalization hypothesis (ETH) and the resulting single-eigenstate statistical mechanics.
Journal ArticleDOI

Stabilizer Codes and Quantum Error Correction

TL;DR: In this paper, the authors give an overview of the field of quantum error correction and the formalism of stabilizer codes, discuss a number of known codes, the capacity of a quantum channel, bounds on quantum codes, and fault-tolerant quantum computation.
Related Papers (5)

Quantum supremacy using a programmable superconducting processor

Frank Arute, +85 more
- 24 Oct 2019 - 
Frequently Asked Questions (10)
Q1. What are the contributions in "Predicting many properties of a quantum system from very few measurements" ?

The authors present an efficient method for constructing an approximate classical description of a quantum state using very few measurements of the state. 

(4)To reduce the prediction error, the authors use N independent snapshots and symmetrize over all possible pairs: 1 N(N−1) ∑ i 6=j tr(Oρ̂i ⊗ ρ̂j). 

due to the dependence on the specific set of observables for choosing the measurement bases, the derandomized version can exploit advantageous structures in the set of observables the authors want to measure. 

(S27)One of the most widely used conjectures for building post-quantum cryptography is the hardness of learning with error (LWE) [63]. 

To circumvent the exponential scaling in representing quantum states, Gosset and Smolin [30] have proposed a stabilizer sketching approach that compresses a classical description of quantum states to an accurate sketch of subexponential size. 

To test the performance of feature prediction with classical shadows the authors first have to simulate the (quantum) data acquisition phase. 

The resulting prediction technique is applicable to current laboratory experiments and facilitates the efficient prediction of few-body properties, such as two-point correlation functions, entanglement entropy of small subsystems, and expectation values of local observables. 

This is mainly due to a curse of dimensionality: the number of parameters needed to describe a quantum system scales exponentially with the number of its constituents. 

Hence the authors only need a total number of Ntot = O(log(M)4k/ 2) random Pauli basis measurements to predict M quadratic functions tr(Oiρ⊗ ρ). 

The authors therefore anticipate that classical shadows will be useful in near-term experiments characterizing noise in quantum devices and exploring variational quantum algorithms for optimization, materials science, and chemisty.