Skip to main content
National Science Review logoLink to National Science Review
. 2019 Apr 9;6(4):719–729. doi: 10.1093/nsr/nwz048

Universal bound on sampling bosons in linear optics and its computational implications

Man-Hong Yung 1,2,3,, Xun Gao 4, Joonsuk Huh 5,
PMCID: PMC8291458  PMID: 34691927

ABSTRACT

In linear optics, photons are scattered in a network through passive optical elements including beam splitters and phase shifters, leading to many intriguing applications in physics, such as Mach–Zehnder interferometry, the Hong–Ou–Mandel effect, and tests of fundamental quantum mechanics. Here we present the fundamental limit in the transition amplitudes of bosons, applicable to all physical linear optical networks. Apart from boson sampling, this transition bound results in many other interesting applications, including behaviors of Bose–Einstein condensates (BEC) in optical networks, counterparts of Hong–Ou–Mandel effects for multiple photons, and approximating permanents of matrices. In addition, this general bound implies the existence of a polynomial-time randomized algorithm for estimating the transition amplitudes of bosons, which represents a solution to an open problem raised by Aaronson and Hance (Quantum Inf Comput 2012; 14: 541–59). Consequently, this bound implies that computational decision problems encoded in linear optics, prepared and detected in the Fock basis, can be solved efficiently by classical computers within additive errors. Furthermore, our result also leads to a classical sampling algorithm that can be applied to calculate the many-body wave functions and the S-matrix of bosonic particles.

Keywords: boson sampling, quantum optics, quantum supremacy, linear optics, computational complexity

INTRODUCTION

Apart from being of fundamental interest in physics, linear optics has become a simple but powerful tool for processing quantum information [1–5] and quantum simulation [6–11]. One of the major advantages for encoding information with light is that photons are highly robust against decoherence, which makes it an ideal system to study quantum coherence [12–16]. Furthermore, in linear optical networks, all possible transformations can be achieved with simple operations involving at most a pair of modes; more precisely, every optical circuit can be implemented with beam splitters and phase shifters only [17]. Linear optical networks have been routinely built in photonic chips [18–20] using standard semiconductor fabrication technology. In particular, a reprogrammable linear optical circuit has been integrated into a photonic chip [20], which can perform universal operations on six photonic modes with up to six photons. In addition, photonic chips can also be applied for demonstrating different physics [21,22].

What about the computational power of linear optics? Knill, Laflamme and Milburn (KLM) [23] proved a universality theorem showing that linear optics, together with post-selected measurements, is powerful enough for universal quantum computation. Aaronson adopted this fact to prove that computation of the permanent of a matrix is #P hard. However, can linear optics outperform a classical device in some computational problems, even without adaptive measurements? This is the main question that we address here.

Recently, it was found that boson sampling [24–28], as a novel application of linear optics, can be regarded as evidence for proving the inefficiency of classical devices to perform quantum simulation, which represents a serious challenge to the validity of the extended Church–Turing thesis [1]. Furthermore, a special type of correlation, called forrelation [29,30] has been identified as being able to exhibit the largest possible separation between quantum and classical query complexities. In boson sampling [24], a product of single-photon states is injected into a linear photonic network that encodes an instance of complex matrices. In fact, the famous Hong–Ou–Mandel interference can be viewed as a special instance of boson sampling (see Fig. 1). The ability to approximate the corresponding permanents of matrices with a multiplicative error implies the ability of simulating boson sampling, which is widely believed to be impossible, based on computational complexity assumptions [24]. With this motivation, much progress [20,31–36] has been made in the experimental realization of boson samplers using linear optical devices.

Figure 1.

Figure 1.

(a) The Hong–Ou–Mandel interference can be viewed as a special instance of boson sampling, where the number of modes is two, and the linear network contains a 50:50 beam splitter only. (b) Summary of our main result: an upper bound of the transition amplitudes for linear optics. The initial and final states are products of Fock states. The matrix U presents any realizable unitary transformation in linear optics.

Specifically, the problem of interest in this work is described as follows: let us suppose that we are given a product of Fock states containing a total of n identical photons (or generally bosons) distributed over m different modes, i.e.

graphic file with name M1.gif (1)

where Inline graphic contains tk photons for tk = 0, 1, 2, …, n. Moreover, the state is subject to the constraint of particle conservation:

graphic file with name M3.gif (2)

Here Inline graphic creates a boson in kth mode and satisfies the commutation relations Inline graphic, Inline graphic.

Let us consider any member U in the set of all possible unitary operators (i.e. linear optics) that induces a linear transformation (i.e. non-interacting) for the boson modes, i.e.

graphic file with name M7.gif (3)

The central problem is to give an upper bound for the absolute value of the transition amplitude, |〈s1s2 ⋅ ⋅ ⋅ sm|U|t1t2 ⋅ ⋅ ⋅ tm〉|, for locating the resulting state in another given product state, |s1s2 ⋅ ⋅ ⋅ sm〉, subject to the same particle-conserving constraint, Inline graphic, for sk = 0, 1, 2, …, n. This problem is solved in this work, where an explicit form of the bound is given in Eq. (4), which limits the efficiency of sampling bosons for all possible linear optical networks, including the behaviors of Bose–Einstein condensates (BEC) in linear networks, and the counterparts of Hong–Ou–Mandel effects for multiple photons. Furthermore, this bound is important for our proof on the existence of a polynomial-time randomized algorithm for approximating permanents of matrices, which represents a solution to an open problem posed by Aaronson and Hance [37].

In terms of computational complexity, this algorithm establishes that the computational complexity class of decision problems encoded in an optical network cannot exhibit quantum supremacy [38], as it can be solved by a classical randomized algorithm. Previously, it was known that a quantum computational model with Clifford gates [39], sparse distribution [40], and fermionic (matchgates) [41–43] can be simulated by classical computers. Our result completes the picture by showing that computational decision problems encoded with non-interacting bosons can also be simulated classically (but sampling bosons remain a hard problem under computational-theoretic assumptions).

We remark that our results are built on the existing results of Aaronson and Hance [37], who have already demonstrated an efficient classical algorithm for a special case where the initial state is confined to be those with either 1 or 0, e.g. |11 ⋅ ⋅ ⋅ 10 ⋅ ⋅ ⋅ 0〉, but not for the states with one or more photons, e.g. |22 ⋅ ⋅ ⋅ 20 ⋅ ⋅ ⋅ 0〉 or |n00 ⋅ ⋅ ⋅ 0〉 for n ≥ 2. In addition, Aaronson and Hance posed an open question [37] asking whether it is possible to extend the result to the cases where the initial state can be an arbitrary Fock product state, i.e. |t1t2 ⋅ ⋅ ⋅ tm〉. In this work, we adopted a different mathematical technique to achieve the goal of constructing such a classical algorithm and answering the open question positively. However, this does not imply that linear optical devices cannot perform hard computational problems, but it does have implications for decision problems (which are discussed later in the paper).

In practice, such a classical algorithm can be applied to calculate the many-body wave functions and the S-matrix of bosonic particles; this is made possible by a technique called ‘operator-to-number conversion’ developed in this work. In the online supplementary material, we provide a compact introduction to various concepts related to this work.

RESULTS

Statement of the main result

We shall prove that the upper bound of the boson transition amplitude is given by the following expression:

graphic file with name M9.gif (4)

where Inline graphic is a product of m factors generated from the elements in the list Inline graphic,

graphic file with name M12.gif (5)

and defined similarly for Inline graphic (see Fig. 1). If one of the modes is unoccupied, e.g. sk = 0, then we simply set Inline graphic.

An immediate consequence of our bound is that a necessary condition for a perfect transition from a general Fock state to another Fock state is that Inline graphic.

In the context of boson sampling [24], the initial state is always a product of single-photon states, i.e. |t1t2 ⋅ ⋅ ⋅ tm〉 = |111 ⋅ ⋅ ⋅ 00〉. In this case, we can recover the result obtained previously by Aaronson and Hance [37], which is a special case of our result, i.e. Inline graphic.

Before we go into the details of the proof of the bound, we first discuss the physical and computational implications of the bound.

Absence of exact boson bunching

If we further set s1 = n, and s2 = s3 ⋅ ⋅ ⋅  = sm = 0, the probability of putting all bosons into the same mode from |111 ⋅ ⋅ ⋅ 00〉 is exponentially low, as (previously obtained in Ref. [37])

graphic file with name M17.gif (6)

using the Stirling approximation, Inline graphic. Consequently, for n ≥ 3,

graphic file with name M19.gif (7)

one cannot observe the generalization of the Hong–Ou–Mandel effect with linear optics, i.e. going from the state |11 ⋅ ⋅ ⋅ 1〉 to |n0 ⋅ ⋅ ⋅ 〉 + |0n0 ⋅ ⋅ ⋅ 〉 + |00n0 ⋅ ⋅ ⋅ 〉; the reason that the Hong–Ou–Mandel effect is possible for the case of n = 2 is because the bound is given by

graphic file with name M20.gif (8)

but there are two modes; the total probability can therefore reach unity. This result is complementary to a previous result [44] showing the absence of the Hong–Ou–Mandel dip with the Bell-multiport beam splitter.

Boson bunching limits

Furthermore, we can find the upper limit on the transition probabilities in general scenarios. For example, imagine that there are p bosons in one mode and q bosons in another mode. Suppose that we are interested in the case where all bosons are grouped into a single mode, i.e.

graphic file with name M21.gif (9)

From Eq. (4), the probability of getting p + q in a single mode through linear optics is then bounded by the following:

graphic file with name M22.gif (10)

or its inverse. As far as we are aware, this bound is new in quantum optics. To be specific, let us consider the following cases.

Case 1: creation of a mode with 2n bosons from two separate modes with n bosons each

In this case, we have

graphic file with name M23.gif (11)

In the limit of a Bose–Einstein condensate (BEC), meaning a very large Fock state |n〉 with n ≫ 1, the probability bound

graphic file with name M24.gif (12)

decreases as Inline graphic. An optimal strategy for achieving the bound is to apply the 50:50 beam splitter, i.e. Inline graphic and Inline graphic. Note that the reverse process, i.e. splitting a BEC into halves, is equally inefficient with linear optics.

Case 2: adding one extra boson to a BEC using linear operations

Supposing that p = n and q = 1, the bound is given by

graphic file with name M28.gif (13)

which approaches a constant limit, e−1, when n → ∞. In fact, this bound can be saturated by the following transformation: Inline graphic, where sin 2θn = (n + 1)−1.

Absence of deterministic boson adders

Note that the process of boson bunching as described in Eq. (9) can be viewed as a process of arithmetic addition using bosons. These results impose limitations to the efficiency of performing arithmetic operations using pure linear optics, which is necessarily a probabilistic process as we have seen, unless extra degrees of freedom are allowed (see e.g. Ref. [45]).

Quantum superposition ≠ classically intractable

By showing that classical algorithms can solve the class of decision problems of sampling bosons, we can establish the following fact: although the key feature of quantum computation comes from the ability to create a superposition of an exponential number of states, our results provide explicit evidence that this quantum ability cannot be a sufficient condition for exhibiting quantum advantage in computational tasks over classical devices. Similar conclusions can be achieved for other quantum computing models, such as with Clifford gates [39], sparse distribution [40], and fermionic (matchgates) [41–43]. Our result is complementary to these models in the context of linear optics.

In other words, non-classical decision problems in linear optics, assuming that they are encoded in polynomially many output states, can be solved.

Bounding size of matrix permanents

Another implication of our main result is related to permanents of matrices. The transition amplitude in Eq. (4) is known (see e.g. Ref. [24]) to be related to a permanent of a matrix regarding the unitary operator U:

graphic file with name M30.gif (14)

where Inline graphic is an n × n matrix constructed by the transformation elements ukj (see Eq. (3)) of the unitary operator U in the following way: create sk copies of a row of vectors Inline graphic that contain tj copies of ukj. The special case where all ti ∈ {0, 1} was discussed in Ref. [37].

For example, if Inline graphic and Inline graphic, then the matrix Inline graphic is of the following form:

graphic file with name M36.gif (15)

Note that if all s and t equal unity, then the transition probability is exactly the same as the permanent of the matrix defined in Eq. (3), i.e.

graphic file with name M37.gif (16)

In other words, our bound also implies an upper bound of the permanent of a matrix:

graphic file with name M38.gif (17)

Efficient representation and calculations of many-body wave functions

As a further application, we have developed a new operator-to-number technique, which can be employed to represent bosonic many-body wave functions,

graphic file with name M39.gif (18)

in terms of a summation of complex numbers (see Eq. (21)). Moreover, the many-body wave function Inline graphic can be estimated efficiently by a random sampling of the complex numbers (see Eq. (23)).

Additionally, the S-matrix in quantum field theory can be reduced to a problem of calculating the permanents of certain matrices [46], which can also be solved by our classical randomized algorithm in polynomial time.

Implication of the open problem on computational complexity

Before this work, it was not known if it was possible to create classically hard ‘decision problems’ with linear quantum optics that involve the determination of the transition amplitudes. One of the major questions in the field of linear optics is whether boson sampling can be extended to solving decisions problems, such as factoring, etc., within additive errors.

Now, the existence of the classical randomized algorithm presented in this work implies that (S. Aaronson, private communication) any decision problem involving only one (or a polynomial number) of boson transition amplitudes can be solved efficiently with a classical device, in the context of computational complexity theory. To be more specific, the open problem [37] raised by Aaronson and Hance asks,

Can we estimate any linear-optical amplitude (see Eq. (14)) to ±1/poly(n) additive error (or better) in polynomial time?

With our bound shown in Eq. (4), we confirm that there does exist a polynomial-time randomized algorithm for the general cases. Thus, the open problem is now settled for those decision problems encoded in (a polynomial number of) the transition amplitudes.

Decision problems of boson sampling

Recall that we are dealing with decision problems instead of sampling problems for linear optics. The sampling problems require an estimation of the transition amplitude to a multiplicative error, or equivalently the ability to reproduce the target distribution, i.e. weak sampling. For decision problems, one is required to estimate the probability directly, i.e. strong simulation (see e.g. Refs. [47,48] for a further discussion on the relationship between strong and weak simulation).

To elaborate further, recall the definition of the complex class BQP (bounded-error quantum polynomial time), which represents the class of languages that can be decided with high probability by polynomial-size uniform quantum circuit families. More precisely, a language L is in BQP if and only if there exists a polynomial-time uniform family of quantum circuits Qn(x), which takes n qubits and outputs 1 bit, such that

  1. for all xL, Inline graphic,

  2. for all xL, Inline graphic.

For example, one may start with the all-zero state, |0n〉 ≡ |000 ⋅ ⋅ ⋅ 0〉. The quantum circuit associated with the instance x can be represented by a unitary transformation Ux. The quantum circuit may accept when one of the outcomes, e.g. |y〉, appears. In other words, we may write

graphic file with name M43.gif (19)

In our case, we replace the quantum circuit in BQP with a linear optical network Inline graphic, which is a subset of BQP, for a fixed number of bosons. However, we consider not only a fixed initial state, but any arbitrary Fock state |t1t2 ⋅ ⋅ ⋅ tm〉 of n photons, which means that a decision problem encoded in linear quantum optics can be formulated as follows:

graphic file with name M45.gif (20)

where Inline graphic denotes a linear optical network of m modes.

Note that an alternative definition of decision problems for boson sampling has been defined [49]. There, decision problems of boson sampling are defined as problems to decide if a function f(x, y) satisfies a certain property. Here x is generated by the most probable bin (MPB) of a boson-sampling instance, and y is an optional string. In other words, by grouping the outcomes of boson sampling into bins, one is required to find the bin associated with the largest probability.

In fact, if we can calculate the individual transition amplitude (i.e. Eq. (14)) accurately, then we can also find such an MPB by generating a probability distribution that approximates the original distribution (see Refs. [49–51]).

Given the existence of the polynomial-time classical algorithm in this work, which can estimate boson transition amplitude within additive error, conditions (i) and (ii) can be readily satisfied with a classical computer, which means that they are actually problems inside the complexity class BPP (see Fig. 2). Finally, we note that our classical algorithm is designed to estimate the amplitude up to an additive error. However, to solve the sampling problem of linear optics, i.e. boson sampling where the amplitudes become exponentially small, it will require the estimation of the amplitudes to within multiplicative errors. As a result, our algorithm will need to run for an exponential time to estimate the amplitudes to within multiplicative errors.

Figure 2.

Figure 2.

Relationship between the complexity class of estimating boson amplitude, and classical and quantum computation. Our result establishes that calculating the boson amplitude, with a polynomial additive error, is a problem inside BPP.

Overall, our result imposes a new constraint for obtaining quantum computational supremacy with linear optics over classical computers, in the context of solving decision problems.

DERIVATION OF MAIN RESULTS

Let us now establish a general theorem that is crucial for our result (see Fig. 3 for a summary). Similar to Feynman’s path integrals, this theorem tells us how to perform operator-to-number conversion for bosonic transition amplitudes. However, unlike Feynman’s path integrals, the number of variables depends on the number of modes, instead of infinite-dimensional integrals, and there is no sign problem.

Figure 3.

Figure 3.

Summary of the relationships with the transition amplitude. (a) The transition matrix element can be transformed into a sum of complex variables, through the operator-to-number conversion technique. (b) The absolute value of the transition can be bounded by calculating Inline graphic and Inline graphic.

Generalized transition amplitude

Given any polynomial function, Inline graphic, of multi-mode creation operators Inline graphic, the vacuum-to-vacuum transition amplitude (unnormalized),

graphic file with name M51.gif (21)

can always be expressed as a sum involving a set of weighted complex roots of unity, by mapping the boson operator,

graphic file with name M52.gif (22)

to a complex number zk, and similarly Inline graphic to its complex conjugate Inline graphic:

graphic file with name M55.gif (23)

where

graphic file with name M56.gif (24)

is related to one of the complex roots of unity ω ≡ e−2πi/d, weighted by a factor Inline graphic. Here d is chosen to be an integer larger than the degree of the function and the sum Inline graphic.

Alternatively, we can write Inline graphic in the form of an expectation value:

graphic file with name M60.gif (25)

which allows us to devise a sampling algorithm to estimate its value, as we shall discuss later.

Technique of operator-to-number conversion

Here we introduce a technique of operator-to-number conversion. Since all the terms in the function Inline graphic commute with one another, we can e.g. sort out the first creation operator Inline graphic as if it was just a real number, and write

graphic file with name M63.gif (26)

where Inline graphic is a resulting polynomial function without Inline graphic. Consequently, we have

graphic file with name M66.gif (27)

Note that there is only one non-zero term in the summation, as

graphic file with name M67.gif (28)

Now, since the Kronecker delta function can be expressed (by the representation through discrete Fourier transform) as follows: Inline graphic, we can therefore write the inner product (with Inline graphic),

graphic file with name M70.gif (29)

as a sum over all values of z1, which implies that we can replace the operators with complex numbers, i.e. Inline graphic. Next, we can define a new polynomial function,

graphic file with name M72.gif (30)

and repeat the same procedure for Inline graphic, and so on, which yields the result in Eq. (23) at the end.

Three-step derivation of the universal bound

We are now ready to present the proof for the bound in Eq. (4). For this purpose, we express the transition amplitude explicitly with bonsonic operators, i.e.

graphic file with name M74.gif (31)

where we have defined an operator function:

graphic file with name M75.gif (32)

The proof can be completed with only three steps as follows.

Step 1 (operator-to-number conversion)

With the transformation rule given in Eq. (3), we have

graphic file with name M76.gif (33)

which is exactly a polynomial function of the creation operators. Therefore, the theorem above implies that

graphic file with name M77.gif (34)

where the function Inline graphic is defined as follows:

graphic file with name M79.gif (35)

In order to bound the absolute value of Inline graphic, it is sufficient to bound the function Inline graphic by writing its absolute value in the following form: Inline graphic.

Step 2 (arithmetic–geometric inequality)

Recall that the weighted arithmetic–geometric inequality suggests that

graphic file with name M83.gif (36)

for all non-negative Ak and λk, subject to the constraint Inline graphic. In terms of our t (by setting λk = tk/n), we have Inline graphic. Now, let us denote

graphic file with name M86.gif (37)

Then, we have,

graphic file with name M87.gif (38)

Step 3 (bounding the norms)

Note that the right-hand side is related to the 2-norm of a vector, which is defined by: Inline graphic. To take a step further, we can always define a unitary matrix V such that

graphic file with name M89.gif (39)

which implies that

graphic file with name M90.gif (40)

Since ‖V‖ = 1 for unitary matrices, and Inline graphic, we have

graphic file with name M92.gif (41)

Consequently, as Inline graphic, we have Inline graphic, and hence

graphic file with name M95.gif (42)

which implies part of the advertised result of the bound Inline graphic in Eq. (4). We can repeat essentially the same procedure for the complex conjugate, 〈t1 ⋅ ⋅ ⋅ tm| U|s1 ⋅ ⋅ ⋅ sm〉, of the transition amplitude, in order to obtain the other part, Inline graphic. This completes our proof for the bound in Eq. (4).

Generalized Glynn’s estimators

In Ref. [37], Aaronson and Hance proposed a generalization of Gurvits’s algorithm by defining a generalized Glynn’s estimator, namely

graphic file with name M98.gif (43)

where Inline graphic is defined in Eq. (5). Sampling the generalized Glynn’s estimator over the complex values, the permanent, Perm(V), of a matrix V, which is obtained by repeating si times the ith row of the m × m matrix W = (wi, j), can be estimated in polynomial time with an additive error Inline graphic. With the new bound, we are able to construct a more general estimator (see Eq. (46)) and bound the size of it. As a result, we are able to estimate matrix permanents and the boson transition amplitude through the relation shown in Eq. (14).

Our classical algorithm

Comparing the right-hand sides of Eqs. (14) and (31), we conclude that Inline graphic is equal to the permanent of the matrix Inline graphic, i.e.

graphic file with name M103.gif (44)

In other words, with Eqs. (31), (34), and (35), we can express the transition amplitude as the following summation:

graphic file with name M104.gif (45)

where Inline graphic, and the constant is given by Inline graphic. Note that it is possible to extend our formalism for an arbitrary m × m matrix W = (wi, j) from the transformation in Eq. (3), which implies that we can define an even more general Glynn estimator,

graphic file with name M107.gif (46)

which is reduced to the estimator, Inline graphic, of Aaronson and Hance for the special cases where t1 = t2 =  ⋅ ⋅ ⋅  = tm = 1, and further reduced to the estimator, Inline graphic, of Gurvits, when s1 = s2 =  ⋅ ⋅ ⋅  = sm = 1 in addition. An alternative estimator can be found in Huh [52].

We return to the case of quantum optics, where the transformation is necessarily a unitary matrix U, with ‖U‖ = 1. Now, let us write

graphic file with name M110.gif (47)

where

graphic file with name M111.gif (48)

Here the important point is that we can bound the left-hand side by the following:

graphic file with name M112.gif (49)

which is due to Eq. (41) and Inline graphic.

Next, we shall show that one can approximate the transition amplitude with a high probability, by uniformly sampling the more general Glynn’s estimator in Eq. (46), with Inline graphic, i.e.

graphic file with name M115.gif (50)

The point is to determine the number of terms T for a given error Inline graphic. Our analysis makes use of the standard Chebyshev bound: for a set of identical, independent, and random variables {Xi}, the probability of taking the average, Inline graphic, of T variables to deviate from the expectation value, μ = 〈Xi〉, by an amount Inline graphic is given by

graphic file with name M119.gif (51)

where σ2 = 〈(Xi − μ)2〉 is the variance, which can be bounded by the second moment Inline graphic, as Inline graphic.

To apply the Chebyshev bound, we identify the transition amplitude as the expectation value

graphic file with name M122.gif (52)

and the random variable as follows:

graphic file with name M123.gif (53)

The efficiency of the sampling algorithm depends on the size of the variance; the remaining task is to determine the size of Inline graphic, which can be used to bound the size of the variance. From Eq. (49), we have

graphic file with name M125.gif (54)

If Inline graphic, we can always repeat the argument for the complex conjugate of the transition amplitude, 〈s1 ⋅ ⋅ ⋅ sm|U|t1 ⋅ ⋅ ⋅ tm〉* = 〈t1 ⋅ ⋅ ⋅ tm|U|s1 ⋅ ⋅ ⋅ sm〉, which means that we can always bound the left-hand side by a value smaller than 1, i.e. Inline graphic.

Consequently, from the Chebyshev bound, by taking a total of Inline graphic samples, the error of the approximation in Eq. (50) can be made to be within Inline graphic with a high probability close to 1. Note that the evaluation of each sample requires O(m2) steps, as in Eq. (46), the calculation of the summation takes m steps and there are m factors to multiply. Therefore, the existence of this polynomial-time algorithm, scaling as Inline graphic, represents a solution to an open problem raised in the work of Aaronson and Hance [37].

DISCUSSION

We have presented a general upper bound (Eq. (4)) on the transition amplitudes in sampling bosons for any linear optical network (Eq. (3)). This bound plays the main role in establishing the fact that the computational complexity of decision problems encoded in linear optics cannot be hard problems for classical computers, when the decision problems are encoded in a transition amplitude. However, it may still be possible to encode decision problems in different ways, e.g. problems that may involve many (exponentially) more such amplitudes. For those cases, a direct application of our algorithm may become inefficient.

Nevertheless, this bound yields many implications in quantum physics and computational complexity. The crucial step in proving the bound involves a general theorem (see Eq. (21)) that makes it possible to convert any vacuum-to-vacuum transition amplitude, for some polynomial functions of the boson operators, into a sum of discrete random variables (Eq. (23)). In addition to boson sampling, this theorem is applicable to sampling problems of spin systems [26] and the calculation of elements of the S-matrix in quantum electrodynamics [46].

The connection between the transition amplitudes and the permanents makes it possible to bound the absolute value of the corresponding permanents of matrices (Eq. (17)). Moreover, the classical algorithm proposed by Gurvits [53] can be extended (Eq. (50)) with our bound; the existence of such an algorithm implies that the open problem of Aaronson and Hance in Ref. [37] can now be settled. However, we note that our goal is to estimate the size of the amplitude (probability), a task known as strong simulation, instead of weak sampling where the target distribution is output. The connection between strong and weak simulation has been thoroughly studied in the context of (extended) Clifford circuits [47,48]. Perhaps a similar analysis can also be carried out in the context of linear optics.

Finally, we note that it is straightforward to show that our bound can also be applied to generalize the de-randomizing algorithm for approximating permanents of non-negative matrices, which was discussed by Aaronson and Hance [37].

Supplementary Material

nwz048_Supplemental_File

ACKNOWLEDGEMENTS

We thank Scott Aaronson and Radu Ionicioiu for the valuable comments on the manuscript.

FUNDING

M.-H.Y. acknowledges support from the National Natural Science Foundation of China (11875160), the Guangdong Innovative and Entrepreneurial Research Team Program (2016ZT06D348), the Natural Science Foundation of Guangdong Province (2017B030308003), and the Science, Technology and Innovation Commission of Shenzhen Municipality (ZDSYS20170303165926217, JCYJ20170412152620376, JCYJ20170817105046702). J.H. acknowledges support from the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (NRF-2015R1A6A3A04059773).

AUTHOR CONTRIBUTIONS

All authors researched, collated, and wrote this paper.

Conflict of interest statement . None declared.

REFERENCES

  • 1. Nielsen MA, Chuang IL. Quantum Computation and Quantum Information. Cambridge: Cambridge University Press, 2011. [Google Scholar]
  • 2. Kok P, Munro WJ, Nemoto Ket al.. Linear optical quantum computing with photonic qubits. Rev Mod Phys 2007; 79: 135–74. [Google Scholar]
  • 3. Peruzzo A, McClean J, Shadbolt Pet al.. A variational eigenvalue solver on a photonic quantum processor. Nat Commun 2014; 5: 4213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Chapman RJ, Santandrea M, Huang Zet al.. Experimental perfect state transfer of an entangled photonic qubit. Nat Commun 2016; 7: 11339. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Gao J, Qiao LF, Jiao ZQet al.. Experimental machine learning of quantum states. Phys Rev Lett 2018; 120: 240501. [DOI] [PubMed] [Google Scholar]
  • 6. Lu CY, Gao WB, Gühne Oet al.. Demonstrating anyonic fractional statistics with a six-qubit quantum simulator. Phys Rev Lett 2009; 102: 030502. [DOI] [PubMed] [Google Scholar]
  • 7. Lanyon BP, Whitfield JD, Gillett GGet al.. Towards quantum chemistry on a quantum computer. Nat Chem 2010; 2: 106–11. [DOI] [PubMed] [Google Scholar]
  • 8. Aspuru-Guzik A, Walther P. Photonic quantum simulators. Nat Phys 2012; 8: 285–91. [Google Scholar]
  • 9. Xu JS, Yung MH, Xu XYet al.. Demon-like algorithmic quantum cooling and its realization with quantum optics. Nat Photon 2014; 8: 113–8. [Google Scholar]
  • 10. Huh J, Mostame S, Fujita Tet al.. Linear-algebraic bath transformation for simulating complex open quantum systems. New J Phys 2014; 16: 123008. [Google Scholar]
  • 11. Li Y, Hu J, Zhang XMet al.. Variational quantum simulation for quantum chemistry. Adv Theory Simul 2019; 2: 1800182. [Google Scholar]
  • 12. Scully MO, Zubairy MS. Quantum Optics. Cambridge: Cambridge University Press, 1997. [Google Scholar]
  • 13. Xu JS, Yung MH, Xu XYet al.. Robust bidirectional links for photonic quantum networks. Sci Adv 2016; 2: e1500672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14. Tillmann M, Tan SH, Stoeckl SEet al.. Generalized multiphoton quantum interference. Phys Rev X 2015; 5: 041015. [Google Scholar]
  • 15. Biggerstaff DN, Heilmann R, Zecevik AAet al.. Enhancing coherent transport in a photonic network using controllable decoherence. Nat Commun 2016; 7: 11282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Song XK, Huang Y, Ling Jet al.. Quantifying quantum coherence in experimentally observed neutrino oscillations. Phys Rev A 2018; 98: 050302. [Google Scholar]
  • 17. Reck M, Zeilinger A, Bernstein HJet al.. Experimental realization of any discrete unitary operator. Phys Rev Lett 1994; 73: 58–61. [DOI] [PubMed] [Google Scholar]
  • 18. Crespi A, Ramponi R, Osellame Ret al.. Integrated photonic quantum gates for polarization qubits. Nat Commun 2011; 2: 566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Shadbolt PJ, Verde MR, Peruzzo Aet al.. Generating, manipulating and measuring entanglement and mixture with a reconfigurable photonic circuit. Nat Photon 2011; 6: 45–9. [Google Scholar]
  • 20. Carolan J, Harrold C, Sparrow Cet al.. Universal linear optics. Science 2015; 349: 711–6. [DOI] [PubMed] [Google Scholar]
  • 21. Tang H, Lin XF, Feng Zet al.. Experimental two-dimensional quantum walk on a photonic chip. Sci Adv 2018; 4: eaat3174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Feng Z, Gao ZW, Wu LAet al.. Photonic Newton’s cradle for remote energy transport. arXiv:1901.07574. [Google Scholar]
  • 23. Knill E, Laflamme R, Milburn GJ. A scheme for efficient quantum computation with linear optics. Nature 2001; 409: 46–52. [DOI] [PubMed] [Google Scholar]
  • 24. Aaronson S, Arkhipov A. The computational complexity of linear optics. Theory Comput 2013; 9: 143–252. [Google Scholar]
  • 25. Gard BT, Motes KR, Olson JPet al.. An Introduction to Boson-Sampling. Singapore: World Scientific, 2015, 167–92. [Google Scholar]
  • 26. Peropadre B, Aspuru-Guzik A, Garcia-Ripoll JJ. Spin models and boson sampling. arXiv:1509.02703. [Google Scholar]
  • 27. Walschaers M, Kuipers J, Urbina JDet al.. Statistical benchmark for BosonSampling. New J Phys 2016; 18: 032001. [Google Scholar]
  • 28. Huh J, Yung MH. Vibronic boson sampling: generalized Gaussian boson sampling for molecular vibronic spectra at finite temperature. Sci Rep 2017; 7: 7462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Aaronson S, Ambainis A. Forrelation: a problem that optimally separates quantum from classical computing. SIAM J Comput 2018; 47: 982–1038. [Google Scholar]
  • 30. Li H, Gao X, Xin Tet al.. Experimental study of forrelation in nuclear spins. Sci Bull 2017; 62: 497–502. [DOI] [PubMed] [Google Scholar]
  • 31. Broome MA, Fedrizzi A, Rahimi-Keshari Set al.. Photonic boson sampling in a tunable circuit. Science 2013; 339: 794–8. [DOI] [PubMed] [Google Scholar]
  • 32. Spring JB, Metcalf BJ, Humphreys PCet al.. Boson sampling on a photonic chip. Science 2013; 339: 798–801. [DOI] [PubMed] [Google Scholar]
  • 33. Tillmann M, Dakić B, Heilmann Ret al.. Experimental boson sampling. Nat Photon 2013; 7: 540–4. [Google Scholar]
  • 34. Crespi A, Osellame R, Ramponi Ret al.. Integrated multimode interferometers with arbitrary designs for photonic boson sampling. Nat Photon 2013; 7: 545–9. [Google Scholar]
  • 35. Spagnolo N, Vitelli C, Bentivegna Met al.. Experimental validation of photonic boson sampling. Nat Photon 2014; 8: 615–20. [Google Scholar]
  • 36. Bentivegna M, Spagnolo N, Vitelli Cet al.. Experimental scattershot boson sampling. Sci Adv 2015; 1: e1400255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Aaronson S, Hance T. Generalizing and derandomizing Gurvits’s approximation algorithm for the permanent. Quantum Inf Comput 2012; 14: 541–59. [Google Scholar]
  • 38. Yung MH. Quantum supremacy: some fundamental concepts. Natl Sci Rev 2019; 6: 22–3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Gottesman D. The Heisenberg representation of quantum computers. arXiv:quant-ph/9807006. [Google Scholar]
  • 40. Schwarz M, den Nest MV. Simulating quantum circuits with sparse output distributions. arXiv:1310.6749. [Google Scholar]
  • 41. Terhal BM, DiVincenzo DP. Classical simulation of noninteracting-fermion quantum circuits. Phys Rev A 2002; 65: 032325. [Google Scholar]
  • 42. Knill E. Fermionic linear optics and matchgates. arXiv:quant-ph/0108033. [Google Scholar]
  • 43. Jozsa R, Miyake A. Matchgates and classical simulation of quantum circuits. Proc R Soc Lond A 2008; 464: 3089–106. [Google Scholar]
  • 44. Lim YL, Beige A. Generalized Hong–Ou–Mandel experiments with bosons and fermions. New J Phys 2005; 7: 155. [Google Scholar]
  • 45. Um M, Zhang J, Lv Det al.. Phonon arithmetic in a trapped ion system. Nat Commun 2016; 7: 11410. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Caianiello ER. On quantum field theory — I: explicit solution of Dyson’s equation in electrodynamics without use of Feynman graphs. Nuovo Cimento 1953; 10: 1634–52. [Google Scholar]
  • 47. Jozsa R, Van den Nest M. Classical simulation complexity of extended Clifford circuits. Quantum Inf Comput 2014; 14: 633–48. [Google Scholar]
  • 48. Koh DE. Further extensions of Clifford circuits and their classical simulation complexities. Quantum Inf Comput 2015; 17: 1–19. [Google Scholar]
  • 49. Nikolopoulos GM, Brougham T. Decision and function problems based on boson sampling. Phys Rev A 2016; 94: 012315. [Google Scholar]
  • 50. Bremner MJ, Montanaro A, Shepherd DJ. Achieving quantum supremacy with sparse and noisy commuting quantum computations. Quantum 2017; 1: 8. [Google Scholar]
  • 51. Yung MH, Gao X. Can chaotic quantum circuits maintain quantum supremacy under noise? arXiv:1706.08913. [Google Scholar]
  • 52. Chin S, Huh J. Generalized concurrence in boson sampling. Sci Rep 2018; 8: 6101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53. Gurvits L. On the complexity of mixed discriminants and related problems. In: Jȩdrzejowicz J and Szepietowski A (eds.). Mathematical Foundations of Computer Science 2005. Berlin: Springer, 2005, 447–58. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

nwz048_Supplemental_File

Articles from National Science Review are provided here courtesy of Oxford University Press

RESOURCES