Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: CC BY 4.0
arXiv:2403.02614v1 [quant-ph] 05 Mar 2024

Generation of True Quantum Random Numbers with On-Demand Probability Distributions via Single-Photon Quantum Walks

Chaoying Meng    \authormark1,3,{\dagger} Miao Cai    \authormark2,3,{\dagger} Yufang Yang    \authormark1,3 Haodong Wu    \authormark2,3 Zhixiang Li    \authormark2,3 Yaping Ruan    \authormark2,3 Yong Zhang    \authormark2 Han Zhang    \authormark1,3,* Keyu Xia    \authormark2,3,4* Franco Nori    \authormark5,6 \authormark1School of Physics, Nanjing University, Nanjing 210023, China
\authormark2College of Engineering and Applied Sciences, and National Laboratory of Solid State Microstructures, Nanjing University, Nanjing 210023, China
\authormark3National Laboratory of Solid State Microstructures, Nanjing University, Nanjing 210093, China
\authormark4Shishan Laboratory, Suzhou Campus of Nanjing University, Suzhou 215000, China
\authormark5Quantum Computing Center, Cluster for Pioneering Research, RIKEN, Wakoshi, Saitama 351-0198, Japan
\authormark6Physics Department, The University of Michigan, Ann Arbor, Michigan 48109-1040, USA
\authormark{\dagger}These two authors contributed equally
\authormark*zhanghan@nju.edu.cn; keyu.xia@nju.edu.cn
journal: opticajournalarticletype: Research Article{abstract*}

Random numbers are at the heart of diverse fields, ranging from simulations of stochastic processes to classical and quantum cryptography. The requirement for true randomness in these applications has motivated various proposals for generating random numbers based on the inherent randomness of quantum systems. The generation of true random numbers with arbitrarily defined probability distributions is highly desirable for applications, but it is very challenging. Here we show that single-photon quantum walks can generate multi-bit random numbers with on-demand probability distributions, when the required “coin” parameters are found with the gradient descent (GD) algorithm. Our theoretical and experimental results exhibit high fidelity for various selected distributions. This GD-enhanced single-photon system provides a convenient way for building flexible and reliable quantum random number generators. Multi-bit random numbers are a necessary resource for high-dimensional quantum key distribution.

1 Introduction

Random numbers are important for science research and engineering applications, such as Monte-Carlo simulations [1, 2], cryptography [3, 4] and tests of fundamental physics [5, 6]. For example, quantum key distribution (QKD) technology highly relies on the availability of true random numbers to protect its communication security [7, 8, 9, 10]. Theoretically, pseudo-random number generators, due to their deterministic and predictable nature, cannot satisfy the requirement for building perfectly secure communication systems. Therefore, the inherent randomness of a quantum system makes it a promising platform for generating faithful random numbers. This is known as quantum random number generator (RNG)[11].

Practical quantum RNGs using various sources of randomness have been demonstrated. Discrete generators can use branching paths[12, 13, 14], arrival times[15, 16, 17, 18], photon counting[19, 20, 21, 22], and attenuated pulse[23, 24]; whereas continuous approaches exploit quantum vacuum fluctuations[25, 26, 27], phase noise of lasers[28, 29, 30], amplified spontaneous emission[10, 31], and Raman scattering[32]. Among these schemes, quantum RNG based on quantum walks promise a convenient and fast way to generate true random numbers[33].

The applications of a RNG strongly rely on the probability distribution used. Different distributions are indispensable in various fields. Uniformly distributed random numbers are most desirable and particularly useful in practical applications [11] because these avoid inherent bias. A Gaussian distributed RNG is of most significance in the modulation of coherent states in continuous-variable QKD systems [34, 35, 36], simulations of communication channels, and stochastic processes (e.g. noise) [37].

It is highly valuable to develop a quantum RNG with an on-demand probability distribution. Based on quantum walks, significant efforts have been made for this task [33, 38]. However, it is challenging to find the proper parameter numbers for a complex system to generate true random numbers with a given distribution. In contrast, the gradient descent (GD) algorithm, as a highly adaptive optimization algorithm that has been widely utilized in many fields [39, 40, 41, 42], can provide a more general and efficient way to accomplish this challenging task.

In this work, we propose a GD-enhanced quantum walk for realizing quantum RNG with, in principle, an on-demand probability distribution. Our GD-based scheme can be implemented by using a linear optical system without the need of time-bin encoding and dynamical modulation. We further experimentally demonstrate the generation of true random numbers with various selected probability distributions by using quantum walks of heralded single photons.

2 System and model

In quantum walks, the walker is located in the Hilbert space pctensor-productsubscriptpsubscriptc\mathcal{H}\equiv\mathcal{H}_{\text{p}}\otimes\mathcal{H}_{\text{c}}caligraphic_H ≡ caligraphic_H start_POSTSUBSCRIPT p end_POSTSUBSCRIPT ⊗ caligraphic_H start_POSTSUBSCRIPT c end_POSTSUBSCRIPT, where psubscriptp\mathcal{H}_{\text{p}}caligraphic_H start_POSTSUBSCRIPT p end_POSTSUBSCRIPT is position space and csubscriptc\mathcal{H}_{\text{c}}caligraphic_H start_POSTSUBSCRIPT c end_POSTSUBSCRIPT is the coin space. The coin space contains two basis vectors {|L,|R}ket𝐿ket𝑅\{|L\rangle,|R\rangle\}{ | italic_L ⟩ , | italic_R ⟩ }, which represent the eigenstate of the coin. Therefore, the definite position and classical coins are both replaced by position states and coin operators in a quantum walk system.

In a one-dimensional (1D) discrete-time quantum walk system, the quantum walker’s state can be described by a product state |Ψ=|ψ|cketΨtensor-productket𝜓ket𝑐|\Psi\rangle=|\psi\rangle\otimes|c\rangle| roman_Ψ ⟩ = | italic_ψ ⟩ ⊗ | italic_c ⟩, where |c=αL|L+αR|Rket𝑐subscript𝛼𝐿ket𝐿subscript𝛼𝑅ket𝑅|c\rangle=\alpha_{L}|L\rangle+\alpha_{R}|R\rangle| italic_c ⟩ = italic_α start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT | italic_L ⟩ + italic_α start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT | italic_R ⟩ is the coin state and |ψ=xαx|xket𝜓subscript𝑥subscript𝛼𝑥ket𝑥|\psi\rangle=\sum_{x}{\alpha_{x}|x\rangle}| italic_ψ ⟩ = ∑ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT | italic_x ⟩ is the position state. Each walking step consists of a unitary operator U^=S^C^^𝑈^𝑆^𝐶\hat{U}=\hat{S}\hat{C}over^ start_ARG italic_U end_ARG = over^ start_ARG italic_S end_ARG over^ start_ARG italic_C end_ARG, where S^^𝑆\hat{S}over^ start_ARG italic_S end_ARG is the conditional shift operator and C^^𝐶\hat{C}over^ start_ARG italic_C end_ARG is the coin operator. The coin operator C^^𝐶\hat{C}over^ start_ARG italic_C end_ARG rotates the coin state and its most general form can be expressed as

C^=x|xx|eiβ(eiξcos(θ)eiζsin(θ)eiζsin(θ)eiξcos(θ)),^𝐶subscript𝑥tensor-productket𝑥bra𝑥superscripte𝑖𝛽matrixsuperscripte𝑖𝜉𝜃superscripte𝑖𝜁𝜃superscripte𝑖𝜁𝜃superscripte𝑖𝜉𝜃\hat{C}=\sum_{x}{|x\rangle\langle x|}\otimes\text{e}^{i\beta}\left(\begin{% matrix}\text{e}^{i\xi}\cos(\theta)&\text{e}^{i\zeta}\sin(\theta)\\ -\text{e}^{i\zeta}\sin(\theta)&\text{e}^{-i\xi}\cos(\theta)\end{matrix}\right)\;,over^ start_ARG italic_C end_ARG = ∑ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT | italic_x ⟩ ⟨ italic_x | ⊗ e start_POSTSUPERSCRIPT italic_i italic_β end_POSTSUPERSCRIPT ( start_ARG start_ROW start_CELL e start_POSTSUPERSCRIPT italic_i italic_ξ end_POSTSUPERSCRIPT roman_cos ( italic_θ ) end_CELL start_CELL e start_POSTSUPERSCRIPT italic_i italic_ζ end_POSTSUPERSCRIPT roman_sin ( italic_θ ) end_CELL end_ROW start_ROW start_CELL - e start_POSTSUPERSCRIPT italic_i italic_ζ end_POSTSUPERSCRIPT roman_sin ( italic_θ ) end_CELL start_CELL e start_POSTSUPERSCRIPT - italic_i italic_ξ end_POSTSUPERSCRIPT roman_cos ( italic_θ ) end_CELL end_ROW end_ARG ) , (1)

where ξ,ζ[0,2π]𝜉𝜁02𝜋\xi,\zeta\in[0,2\pi]italic_ξ , italic_ζ ∈ [ 0 , 2 italic_π ] and θ[0,π/2]𝜃0𝜋2\theta\in[0,\pi/2]italic_θ ∈ [ 0 , italic_π / 2 ] are the parameters of the rotation and β𝛽\betaitalic_β fixes the global phase. The conditional shift operator S^^𝑆\hat{S}over^ start_ARG italic_S end_ARG moves the walker either to the left or right depending on the coin state and has the form

S^=x|x1,Lx,R|+|x+1,Rx,L|.^𝑆subscript𝑥ket𝑥1𝐿bra𝑥𝑅ket𝑥1𝑅bra𝑥𝐿\hat{S}=\sum_{x}{|x-1,L\rangle\langle x,R|+|x+1,R\rangle\langle x,L|}\;.over^ start_ARG italic_S end_ARG = ∑ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT | italic_x - 1 , italic_L ⟩ ⟨ italic_x , italic_R | + | italic_x + 1 , italic_R ⟩ ⟨ italic_x , italic_L | . (2)

It leads to the conditional shift operation S^|x,L=|x+1,L^𝑆ket𝑥𝐿ket𝑥1𝐿\hat{S}|x,L\rangle=|x+1,L\rangleover^ start_ARG italic_S end_ARG | italic_x , italic_L ⟩ = | italic_x + 1 , italic_L ⟩ and S^|x,R=|x1,R^𝑆ket𝑥𝑅ket𝑥1𝑅\hat{S}|x,R\rangle=|x-1,R\rangleover^ start_ARG italic_S end_ARG | italic_x , italic_R ⟩ = | italic_x - 1 , italic_R ⟩. In the following, we fix the parameters β=π/2𝛽𝜋2\beta=\pi/2italic_β = italic_π / 2 and ξ=ζ=π/2𝜉𝜁𝜋2\xi=\zeta=-\pi/2italic_ξ = italic_ζ = - italic_π / 2, so that we obtain the coin determined by one parameter θ𝜃\thetaitalic_θ. If θ=π/4𝜃𝜋4\theta=\pi/4italic_θ = italic_π / 4, the coin then becomes the Hadamard coin : H^=[1111]/2^𝐻delimited-[]matrix11112\hat{H}=\left[\begin{matrix}1&1\\ 1&-1\end{matrix}\right]/\sqrt{2}over^ start_ARG italic_H end_ARG = [ start_ARG start_ROW start_CELL 1 end_CELL start_CELL 1 end_CELL end_ROW start_ROW start_CELL 1 end_CELL start_CELL - 1 end_CELL end_ROW end_ARG ] / square-root start_ARG 2 end_ARG.

After n𝑛nitalic_n walking steps, the state of a quantum walk system becomes |ΨnketsubscriptΨ𝑛|\Psi_{n}\rangle| roman_Ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩. The quantum walker remains in a superposition of many positions until the final measurement is performed. The measured probability for the walker being at xksubscript𝑥𝑘x_{k}italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT after n𝑛nitalic_n walking steps can be written as

𝒫(xk)=|R|xk|Ψn|2+|L|xk|Ψn|2.𝒫subscript𝑥𝑘superscriptbra𝑅inner-productsubscript𝑥𝑘subscriptΨ𝑛2superscriptbra𝐿inner-productsubscript𝑥𝑘subscriptΨ𝑛2\mathcal{P}(x_{k})=|\langle R|\langle x_{k}|\Psi_{n}\rangle|^{2}+|\langle L|% \langle x_{k}|\Psi_{n}\rangle|^{2}\;.caligraphic_P ( italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) = | ⟨ italic_R | ⟨ italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | roman_Ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | ⟨ italic_L | ⟨ italic_x start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | roman_Ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (3)

The probability distribution is determined by the choice of coin parameter set in each walking step. It is difficult to adjust the coin parameters to obtain desired probability distributions because the number of coin parameters grows rapidly when increasing the walking steps. In this work, we exploit the gradient descent algorithm to solve this challenging problem.

3 Algorithm

Generally, a GD algorithm consists of three elements [43]: a system function F𝐹Fitalic_F, system parameters {θi}(i=1,2,,k)subscript𝜃𝑖𝑖12𝑘\{\theta_{i}\}\;(i=1,2,...,k){ italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } ( italic_i = 1 , 2 , … , italic_k ), and a loss function \mathcal{L}caligraphic_L. The system function F𝐹Fitalic_F defines the input-output relation of the system and is parameterized by the parameters {θi}(i=1,2,,k)subscript𝜃𝑖𝑖12𝑘\{\theta_{i}\}\;(i=1,2,...,k){ italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } ( italic_i = 1 , 2 , … , italic_k ). The loss function \mathcal{L}caligraphic_L evaluates the system output and compares it to the target. Here, we write the system function F𝐹Fitalic_F as y=F(x;θ1,θ2,,θk)𝑦𝐹𝑥subscript𝜃1subscript𝜃2subscript𝜃𝑘y=F(x\;;\theta_{1},\theta_{2},...,\theta_{k})italic_y = italic_F ( italic_x ; italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ), where x𝑥xitalic_x and y𝑦yitalic_y are the input and output of the system, respectively. Hereafter we use the mean square error function =12(Ty)212superscript𝑇𝑦2\mathcal{L}=\frac{1}{2}(T-y)^{2}caligraphic_L = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_T - italic_y ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT as the loss function, where T𝑇Titalic_T is the target. Therefore, the loss function \mathcal{L}caligraphic_L is also parameterized by {θi}(i=1,2,,k)subscript𝜃𝑖𝑖12𝑘\{\theta_{i}\}\;(i=1,2,...,k){ italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } ( italic_i = 1 , 2 , … , italic_k ) and can be written as ({θi})subscript𝜃𝑖\mathcal{L}(\{\theta_{i}\})caligraphic_L ( { italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } ). Essentially, the gradient descent method minimizes the loss function ({θi})subscript𝜃𝑖\mathcal{L}(\{\theta_{i}\})caligraphic_L ( { italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } ) by updating {θi}subscript𝜃𝑖\{\theta_{i}\}{ italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } in the opposite direction of the gradient of the loss function. In each iteration of the gradient descent method, the parameters θisubscript𝜃𝑖\theta_{i}italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are updated according to [θiηθi(θi)]θidelimited-[]subscript𝜃𝑖𝜂subscriptsubscript𝜃𝑖subscript𝜃𝑖subscript𝜃𝑖[\theta_{i}-\eta\cdot\nabla_{\theta_{i}}\mathcal{L}(\theta_{i})]\rightarrow% \theta_{i}[ italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_η ⋅ ∇ start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT caligraphic_L ( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ] → italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT[43], where η(0,1]𝜂01\eta\in(0,1]italic_η ∈ ( 0 , 1 ] is the learning rate.

Refer to caption
Figure 1: (a) Schematic of a one-dimensional discrete-time quantum walk process. The red arrows represent the walking directions. The vertical and horizontal gray dashed lines denote the position states and the walking steps, respectively. (b) Details of the quantum state transfer in a quantum walk. The symbols next to the red arrows describe the coin state transfer in each walking step. The symbol rmnsuperscriptsubscript𝑟𝑚𝑛r_{m}^{n}italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT represents the coin bias ratio of the n𝑛nitalic_n-th walking step starting from position m𝑚mitalic_m.

A 1D discrete-time quantum walk process is depicted in Fig. 1 (a). The blue circles denote different position states, and the red arrows indicate the directions of the walk starting from different position states in each walking step. Without loss of generality, we assume the coin bias ratio can be adjusted for every coin operation at different position states in different walking steps. This assumption can be experimentally realized in a linear optical system [44].

The specific description of quantum state transfer in a quantum walk system is shown in Fig. 1(b). The letters c𝑐citalic_c and r𝑟ritalic_r represent the complex amplitude and coin bias ratio, respectively. The notation cm,R(n)superscriptsubscript𝑐𝑚𝑅𝑛c_{m,R}^{(n)}\;italic_c start_POSTSUBSCRIPT italic_m , italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT(cm,Lnsuperscriptsubscript𝑐𝑚𝐿𝑛c_{m,L}^{n}italic_c start_POSTSUBSCRIPT italic_m , italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT) represents the complex amplitude of the coin state |Rket𝑅|R\rangle| italic_R ⟩ (|Lket𝐿|L\rangle| italic_L ⟩) at the position m𝑚mitalic_m in the n𝑛nitalic_n-th walking step; while rm(n)superscriptsubscript𝑟𝑚𝑛r_{m}^{(n)}italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT is the coin bias ratio of the n𝑛nitalic_n-th walking step starting from position m𝑚mitalic_m, and Pmsubscript𝑃𝑚P_{m}italic_P start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT is the measured probability at the detector located at position m𝑚mitalic_m. The coin bias ratio r𝑟ritalic_r is defined as r=cos2θ𝑟superscript2𝜃r=\cos^{2}\thetaitalic_r = roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ. Thus, the state transformation with coin bias ratio r𝑟ritalic_r can be modeled as c|Lrc|L+1rc|R𝑐ket𝐿𝑟𝑐ket𝐿1𝑟𝑐ket𝑅c|L\rangle\rightarrow\sqrt{r}\cdot c|L\rangle+\sqrt{1-r}\cdot c|R\rangleitalic_c | italic_L ⟩ → square-root start_ARG italic_r end_ARG ⋅ italic_c | italic_L ⟩ + square-root start_ARG 1 - italic_r end_ARG ⋅ italic_c | italic_R ⟩, and c|R1rc|Lrc|R𝑐ket𝑅1𝑟𝑐ket𝐿𝑟𝑐ket𝑅c|R\rangle\rightarrow\sqrt{1-r}\cdot c|L\rangle-\sqrt{r}\cdot c|R\rangleitalic_c | italic_R ⟩ → square-root start_ARG 1 - italic_r end_ARG ⋅ italic_c | italic_L ⟩ - square-root start_ARG italic_r end_ARG ⋅ italic_c | italic_R ⟩. Then the measured probability Pmsubscript𝑃𝑚P_{m}italic_P start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT becomes

Pm=[1rm1(n+1)am2,R(n)rm1(n+1)am,L(n)]2+[1rm1(n+1)bm2,R(n)rm1(n+1)bm,L(n)]2+[rm+1(n+1)am,R(n)+1rm+1(n+1)am+2,L(n)]2+[rm+1(n+1)bm,R(n)+1rm+1(n+1)bm+2,L(n)]2,subscript𝑃𝑚superscriptdelimited-[]1superscriptsubscript𝑟𝑚1𝑛1superscriptsubscript𝑎𝑚2𝑅𝑛superscriptsubscript𝑟𝑚1𝑛1superscriptsubscript𝑎𝑚𝐿𝑛2superscriptdelimited-[]1superscriptsubscript𝑟𝑚1𝑛1superscriptsubscript𝑏𝑚2𝑅𝑛superscriptsubscript𝑟𝑚1𝑛1superscriptsubscript𝑏𝑚𝐿𝑛2superscriptdelimited-[]superscriptsubscript𝑟𝑚1𝑛1superscriptsubscript𝑎𝑚𝑅𝑛1superscriptsubscript𝑟𝑚1𝑛1superscriptsubscript𝑎𝑚2𝐿𝑛2superscriptdelimited-[]superscriptsubscript𝑟𝑚1𝑛1superscriptsubscript𝑏𝑚𝑅𝑛1superscriptsubscript𝑟𝑚1𝑛1superscriptsubscript𝑏𝑚2𝐿𝑛2\begin{split}P_{m}=&\left[\sqrt{1-r_{m-1}^{(n+1)}}a_{m-2,R}^{(n)}-\sqrt{r_{m-1% }^{(n+1)}}a_{m,L}^{(n)}\right]^{2}+\left[\sqrt{1-r_{m-1}^{(n+1)}}b_{m-2,R}^{(n% )}-\sqrt{r_{m-1}^{(n+1)}}b_{m,L}^{(n)}\right]^{2}\\ &+\left[\sqrt{r_{m+1}^{(n+1)}}a_{m,R}^{(n)}+\sqrt{1-r_{m+1}^{(n+1)}}a_{m+2,L}^% {(n)}\right]^{2}+\left[\sqrt{r_{m+1}^{(n+1)}}b_{m,R}^{(n)}+\sqrt{1-r_{m+1}^{(n% +1)}}b_{m+2,L}^{(n)}\right]^{2}\;,\end{split}start_ROW start_CELL italic_P start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = end_CELL start_CELL [ square-root start_ARG 1 - italic_r start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n + 1 ) end_POSTSUPERSCRIPT end_ARG italic_a start_POSTSUBSCRIPT italic_m - 2 , italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT - square-root start_ARG italic_r start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n + 1 ) end_POSTSUPERSCRIPT end_ARG italic_a start_POSTSUBSCRIPT italic_m , italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + [ square-root start_ARG 1 - italic_r start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n + 1 ) end_POSTSUPERSCRIPT end_ARG italic_b start_POSTSUBSCRIPT italic_m - 2 , italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT - square-root start_ARG italic_r start_POSTSUBSCRIPT italic_m - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n + 1 ) end_POSTSUPERSCRIPT end_ARG italic_b start_POSTSUBSCRIPT italic_m , italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + [ square-root start_ARG italic_r start_POSTSUBSCRIPT italic_m + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n + 1 ) end_POSTSUPERSCRIPT end_ARG italic_a start_POSTSUBSCRIPT italic_m , italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT + square-root start_ARG 1 - italic_r start_POSTSUBSCRIPT italic_m + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n + 1 ) end_POSTSUPERSCRIPT end_ARG italic_a start_POSTSUBSCRIPT italic_m + 2 , italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + [ square-root start_ARG italic_r start_POSTSUBSCRIPT italic_m + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n + 1 ) end_POSTSUPERSCRIPT end_ARG italic_b start_POSTSUBSCRIPT italic_m , italic_R end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT + square-root start_ARG 1 - italic_r start_POSTSUBSCRIPT italic_m + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n + 1 ) end_POSTSUPERSCRIPT end_ARG italic_b start_POSTSUBSCRIPT italic_m + 2 , italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , end_CELL end_ROW (4)

where a𝑎aitalic_a and b𝑏bitalic_b are the real and imaginary components of c𝑐citalic_c, respectively.

According to the GD algorithm, the updated value of rm(n)superscriptsubscript𝑟𝑚𝑛r_{m}^{(n)}italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT with respect to Pjsubscript𝑃𝑗P_{j}italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is

Δrm,Pj(n)=ηrm(n)=ηPjPjrm(n)=η(TjPj)Pjrm(n),Δsuperscriptsubscript𝑟𝑚subscript𝑃𝑗𝑛𝜂superscriptsubscript𝑟𝑚𝑛𝜂subscript𝑃𝑗subscript𝑃𝑗superscriptsubscript𝑟𝑚𝑛𝜂subscript𝑇𝑗subscript𝑃𝑗subscript𝑃𝑗superscriptsubscript𝑟𝑚𝑛\Delta r_{m,P_{j}}^{(n)}=-\eta\frac{\partial\mathcal{L}}{\partial r_{m}^{(n)}}% =-\eta\frac{\partial\mathcal{L}}{\partial P_{j}}\frac{\partial P_{j}}{\partial r% _{m}^{(n)}}=\eta(T_{j}-P_{j})\frac{\partial P_{j}}{\partial r_{m}^{(n)}}\;,roman_Δ italic_r start_POSTSUBSCRIPT italic_m , italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT = - italic_η divide start_ARG ∂ caligraphic_L end_ARG start_ARG ∂ italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT end_ARG = - italic_η divide start_ARG ∂ caligraphic_L end_ARG start_ARG ∂ italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG divide start_ARG ∂ italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT end_ARG = italic_η ( italic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) divide start_ARG ∂ italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT end_ARG , (5)

where here the loss function becomes =12j(TjPj)212subscript𝑗superscriptsubscript𝑇𝑗subscript𝑃𝑗2\mathcal{L}=\frac{1}{2}\sum_{j}(T_{j}-P_{j})^{2}caligraphic_L = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, and Tjsubscript𝑇𝑗T_{j}italic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the target probability at position j𝑗jitalic_j. Then the overall updated value of rm(n)superscriptsubscript𝑟𝑚𝑛r_{m}^{(n)}italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT is obtained by summing Eq. (5),

jΔrm,Pj(n)=jη(TjPj)Pjrm(n).subscript𝑗Δsuperscriptsubscript𝑟𝑚subscript𝑃𝑗𝑛subscript𝑗𝜂subscript𝑇𝑗subscript𝑃𝑗subscript𝑃𝑗superscriptsubscript𝑟𝑚𝑛\sum_{j}\Delta r_{m,P_{j}}^{(n)}=\sum_{j}\eta(T_{j}-P_{j})\frac{\partial P_{j}% }{\partial r_{m}^{(n)}}\;.∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT roman_Δ italic_r start_POSTSUBSCRIPT italic_m , italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_η ( italic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) divide start_ARG ∂ italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT end_ARG . (6)

The details of the derivation are presented in the supplemental document. Therefore, during each iteration of our algorithm, rm(n)superscriptsubscript𝑟𝑚𝑛r_{m}^{(n)}italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT updates according to the following relation

[rm(n)+jη(TjPj)Pjrm(n)]rm(n).delimited-[]superscriptsubscript𝑟𝑚𝑛subscript𝑗𝜂subscript𝑇𝑗subscript𝑃𝑗subscript𝑃𝑗superscriptsubscript𝑟𝑚𝑛superscriptsubscript𝑟𝑚𝑛\left[r_{m}^{(n)}+\sum_{j}\eta(T_{j}-P_{j})\frac{\partial P_{j}}{\partial r_{m% }^{(n)}}\right]\rightarrow r_{m}^{(n)}\;.[ italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_η ( italic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) divide start_ARG ∂ italic_P start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT end_ARG ] → italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT . (7)

The training finishes when the simulated quantum walk probability distribution reaches the target distribution. After the training is completed, the theoretical values of the coin bias ratios for generating the desired probability distribution are obtained.

4 Experimental setup

Quantum walks lay the natural foundation for studying plenty of novel quantum phenomena and can be realized in various systems[45, 46, 47, 48, 49, 50]. Among these, linear-optics-based quantum walks have advantages in convenience of implementation and compatibility. Therefore, we use this platform for our GD-based quantum RNG scheme.

In linear optical implementations of quantum walks, we use single photons as the quantum walker that moves in both directions in every position. The polarization states {|H,|V}ket𝐻ket𝑉\{|H\rangle,|V\rangle\}{ | italic_H ⟩ , | italic_V ⟩ } are introduced to represent two orthogonal coin states {|L,|R}ket𝐿ket𝑅\{|L\rangle,|R\rangle\}{ | italic_L ⟩ , | italic_R ⟩ }, respectively. We use single-photon spatial modes to represent the position of the walker |xket𝑥|x\rangle| italic_x ⟩.

The schematic of our experimental setup is shown in Fig. 2(a). Pairs of single photons are created via type-II spontaneous parametric down-conversion in a periodically poled potassium titanyl phosphate (PPKTP) crystal. This crystal is pumped by a diode laser centered at 397.5397.5397.5397.5 \unit\milli and emits orthogonally polarized photon pairs (i.e., horizontal and vertical polarized, or left- and right-circularly polarized) with a wavelength of 795 \unit\nano and a FWHM bandwidth of 0.3 \unit\nano. The photon pairs are separated by a polarized beam splitter. One photon from each pair served as a trigger while the other photon is launched into the quantum walk system.

Refer to caption
Figure 2: (a) Schematic of experimental setup. PPKTP: periodically poled potassium titanyl phosphate crystal, PBS: polarized beam splitter, SPCM: single photon counting module, HWP: half-wave plate, QWP: quarter-wave plate, AWP: adjustable wave plate, M: mirror, BD: beam displacer. Here AWP is designed as a HWP in the middle of two QWPs in order to compensate for the phase shift caused by the fiber twist, and can convert circular polarized light to horizontal polarized light without loss. (b) Details of the first two quantum walk steps in our experiment. (c) Measured (colorful bars) and theoretical (red dashes) probability distribution for a four-step quantum walk. Different colors represent different polarization states of the initial input single photons. (d) The second-order correlation function gc2(τ)superscriptsubscript𝑔𝑐2𝜏g_{c}^{2}(\tau)italic_g start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_τ ) versus the delay τ𝜏\tauitalic_τ for our single-photon source. The time window length is approximately 3333 \unit\nano and gc2(0)superscriptsubscript𝑔𝑐20g_{c}^{2}(0)italic_g start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 0 ) is 0.0286±0.001plus-or-minus0.02860.0010.0286\pm 0.0010.0286 ± 0.001.

The position states of the quantum walk are represented by spatial modes of the single photons. The shift operator S^^𝑆\hat{S}over^ start_ARG italic_S end_ARG acting on these modes is implemented by a 37.7 \unit\milli long, birefringent calcite beam displacer. The optical axis of each calcite prism is cut so that vertically polarized light was directly transmitted, and horizontal light underwent a 4 \unit\milli lateral displacement into a neighboring spatial mode. Here, we place the half wave plates in front of each beam displacer to adjust the coin bias ratio in the quantum walk. The aperture diameter of our half wave plate is small so that each half wave plate can change the polarization state of one beam of light without affecting adjacent beams. Therefore, we can adjust the coin bias ratio at different positions during each walking step.

The details of the first two quantum walk steps are depicted in Fig. 2(b). The spatial modes after step 1111 are recombined interferometrically in step 2222. Repetition of these steps then forms an interferometric network as in Fig. 2(a). The lattice sites are labeled so that there are odd sites at odd walking steps, and even sites at even steps. After an n𝑛nitalic_n-step quantum walk, the photons output in (n+1𝑛1n+1italic_n + 1) spatial modes are coupled into an optical fiber and subsequently detected by a single-photon photodiode, in coincidence with the trigger photon. We measure the final probability distribution of the quantum walk by manually moving the fiber coupler between individual output spatial modes.

For a four-step quantum walk with an unbiased coin (θ=π/4𝜃𝜋4\theta=\pi/4italic_θ = italic_π / 4), the measured probability distribution at given sites is shown in Fig. 2(c). Here we choose four initial polarization states to verify our experimental system: horizontal polarization, vertical polarization, right-circular polarization, and left-circular polarization. The experimental data (bars with colors) are in excellent agreement with theoretical simulations (red dashes). To characterize the single-photon purity in the experiment, we also measure the second-order correlation function gc2(τ)superscriptsubscript𝑔𝑐2𝜏g_{c}^{2}(\tau)italic_g start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_τ ) for our single-photon source, as depicted in Fig. 2(d).

5 Results

5.1 Uniform distribution

Quantum RNGs with a uniform distribution [20, 51] are of importance for applications without inherent bias, such as quantum secure communication [52, 11]. Therefore, we first evaluate the performance of our algorithm for generating a uniform distribution in a four-step quantum walk system. Here we use the fidelity \mathcal{F}caligraphic_F, defined to evaluate the similarity between the output (simulated or measured output) and the target probability distribution,

=my(m)T(m)mmax(y(m),T(m))2,subscript𝑚𝑦𝑚𝑇𝑚subscript𝑚maxsuperscript𝑦𝑚𝑇𝑚2\mathcal{F}=\frac{\sum_{m}{y(m)\cdot T(m)}}{\sum_{m}\text{max}(y(m),T(m))^{2}}\;,caligraphic_F = divide start_ARG ∑ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_y ( italic_m ) ⋅ italic_T ( italic_m ) end_ARG start_ARG ∑ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT max ( italic_y ( italic_m ) , italic_T ( italic_m ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (8)

where y𝑦yitalic_y is the system output, T𝑇Titalic_T is the target distribution, and m𝑚mitalic_m represents the position.

Refer to caption
Figure 3: Uniform probability distribution generation in a four-step quantum walk system. (a) Fidelity as the iteration increases. (b) Values of the coin bias ratio of each position and walking step, obtained with our GD algorithm. The black arrow points out the direction of the quantum walking process. The number displayed on the cell is the value of the corresponding coin bias ratio r𝑟ritalic_r. (c) Measured probability distribution of the quantum walk for right-circular (green) and left-polarized (orange) single photons. The red dashes represent the values of the target probability distribution.

For generating a uniform probability distribution, the fidelity curve during the training of our GD algorithm is shown in Fig. 3(a). The “iterations" represent the accumulating time step when the training progresses. From Fig. 3(a) we can see that the fidelity increases rapidly as the training process goes on. It exceeds 0.950.950.950.95 after 5555 iterations and finally approaches unity within 20202020 iterations. The learning rate of the training process is set as 0.10.10.10.1. The convergence rate of the training can be further improved by appropriately choosing the learning rate η𝜂\etaitalic_η.

When the training is completed, we obtain the values of the coin bias ratio for generating a uniform probability distribution in the quantum walk. The values are shown in Fig. 3(b). Obviously, these values are unlikely to be found manually, while our algorithm can find proper values to obtain a high fidelity. According to these values, we adjust {r}𝑟\{r\}{ italic_r } in the quantum walk experimental setup by rotating the half-wave plates in front of the BDs. The minimal adjustable angle of our half-wave plate is 0.250.250.250.25 degree, which leads to a slight deviation between the actual coin bias ratio in the experiment and the theoretical values. But this does not affect the performance of our experiment because our system has strong robustness (See supplemental document for details of the experimental system robustness analysis). We perform experiments with right-circular and left-circular polarized single photons, respectively. The measured probability distributions for detecting the photon at given positions are shown in Fig. 3(c). It is clear that the measured probability distributions are in good agreement with the target distribution. The fidelities of the experimental results are 96.5%percent96.596.5\%96.5 % for right-circular polarized photons and 95.8%percent95.895.8\%95.8 % for left-circular polarized photons.

Refer to caption
Figure 4: Gaussian probability distribution generation in a four-step quantum walk system. (a) Fidelity as the iteration increases. (b) Values of the coin bias ratio for each walking step and position, obtained with our GD algorithm. (c) Measured probability distribution of the quantum walk for right-circular (green) and left-polarized (orange) single photons.

5.2 Gaussian distribution

Gaussian RNGs, as another important RNG, also have diverse useful applications, including Monte Carlo simulation of Gaussian noises. Specific to quantum information, this type of RNGs provide Gaussian distributed randomness for coherent states modulation in continuous-variable quantum key distribution systems [34, 35, 36]. In the following, we show that our GD algorithm can find the parameter set for the quantum walk based RNG to generate Gaussian distributed single-photon outputs.

We set the Gaussian distribution as the target probability distribution for the GD algorithm. The fidelity change during the training process is shown in Fig. 4(a). It can be seen that the fidelity rapidly increases to 95%percent9595\%95 % at the 10101010th iteration. The coin bias ratios can be found in Fig. 4(b). Figure 4(c) presents the measured probability distribution of single photons in a quantum walk with GD-optimized coin bias ratios. Right- and left-circularly polarized photons are chosen as input photons to perform the quantum walk experiment. The experimentally measured probability distribution is again in good agreement with the target distribution. The fidelities of the experiment results are 94.1%percent94.194.1\%94.1 % and 95.8%percent95.895.8\%95.8 % for the right- and left-circular polarized input photons, respectively. These results show that our algorithm can be utilized to adjust a quantum walk system to generate single photons with desired distributions. This allows one to build an effective quantum RNG that conforms to arbitrary probability distributions.

6 Conclusion

We have reported a GD-enhanced quantum RNG based on quantum walks of single photons in a linear optical system. Our multi-bit quantum RNG can generate true random numbers with an arbitrarily defined probability distribution with nearly unitary fidelity. The promised faithful randomness of our quantum RNG can determine the random measurement basis in high-dimensional quantum communications [53, 54, 55, 56]. We note that quantum walks with a uniform distribution can be used to generate quantum random numbers [57]. In comparison with this method, our GD-enhanced quantum walk can generate quantum random numbers with flexible probability distribution.

\bmsection

Funding National Key R&D Program of China (Grant No. 2019YFA0308700), the National Natural Science Foundation of China (Grant No.11890704), the Program for Innovative Talents and Teams in Jiangsu (Grant No. JSSCTD202138), Nippon Telegraph and Telephone Corporation (NTT) Research, the Japan Science and Technology Agency (JST) (via the Quantum Leap Flagship Program (Q-LEAP), and the Moonshot R&D Grant Number JPMJMS2061), the Asian Office of Aerospace Research and Development (AOARD) (Grant No. FA2386-20-1-4069), and the Foundational Questions Institute Fund (FQXi) (Grant No. FQXi-IAF19-06).

\bmsection

Acknowledgements The authors thank Lijian Zhang and Ben Wang for helpful discussions. We thank the High Performance Computing Center of Nanjing University for allowing the numerical calculations on its blade cluster system.

\bmsection

Disclosures The authors declare that there are no conflicts of interest related to this article.

\bmsection

Supplemental document See Supplement 1 for supporting content.

References

  • [1] N. Metropolis and S. Ulam, “The Monte Carlo method,” \JournalTitleJ. Am. Statist. Assoc. 44, 335–341 (1949).
  • [2] H. Niederreiter, “Quasi-Monte Carlo methods and pseudo-random numbers,” \JournalTitleBull. Amer. Math. Soc. 84, 957–1041 (1978).
  • [3] C. E. Shannon, “Communication theory of secrecy systems,” \JournalTitleBell Syst. Tech. J. 28, 656–715 (1949).
  • [4] R. Gennaro, “Randomness in cryptography,” \JournalTitleIEEE Secur. Priv. 4, 64–67 (2006).
  • [5] J. S. Bell, “On the Einstein-Podolsky-Rosen paradox,” \JournalTitlePhysica 1, 195 (1964).
  • [6] P. Shadbolt, J. C. F. Mathews, A. Laing, and J. L. O’Brien, “Testing foundations of quantum mechanics with photons,” \JournalTitleNat. Phys. 10, 278–286 (2014).
  • [7] N. Gisin, G. Ribordy, W. Tittel, and H. Zbinden, “Quantum cryptography,” \JournalTitleRev. Mod. Phys. 74, 145–195 (2002).
  • [8] V. Scarani, H. Bechmann-Pasquinucci, N. J. Cerf, et al., “The security of practical quantum key distribution,” \JournalTitleRev. Mod. Phys. 81, 1301–1350 (2009).
  • [9] H.-K. Lo, M. Curty, and K. Tamaki, “Secure quantum key distribution,” \JournalTitleNat. Photonics 8, 595–604 (2014).
  • [10] A. Martin, B. Sanguinetti, C. C. W. Lim, et al., “Quantum random number generation for 1.25-GHz quantum key distribution systems,” \JournalTitleJ. Lightwave Technol. 33, 2855–2859 (2015).
  • [11] M. Herrero-Collantes and J. C. Garcia-Escartin, “Quantum random number generators,” \JournalTitleRev. Mod. Phys. 89, 015004 (2017).
  • [12] J. G. Rarity, P. C. M. Owens, and P. R. Tapster, “Quantum random-number generation and key sharing,” \JournalTitleJ. Mod. Opt. 41, 2435–2444 (1994).
  • [13] T. Jennewein, U. Achleitner, G. Weihs, et al., “A fast and compact quantum random number generator,” \JournalTitleRev. Sci. Instrum. 71, 1675–1680 (2000).
  • [14] A. Stefanov, N. Gisin, O. Guinnard, et al., “Optical quantum random number generator,” \JournalTitleJ. Mod. Opt. 47, 595–598 (2000).
  • [15] H.-Q. Ma, Y. Xie, and L.-A. Wu, “Random number generation based on the time of arrival of single photons,” \JournalTitleAppl. Opt. 44, 7760–7763 (2005).
  • [16] M. Stipčević and B. M. Rogina, “Quantum random number generator based on photonic emission in semiconductors,” \JournalTitleRev. Sci. Instrum. 78, 045104 (2007).
  • [17] J. F. Dynes, Z. L. Yuan, A. W. Sharpe, and A. J. Shields, “A high speed, postprocessing free, quantum random number generator,” \JournalTitleAppl. Phys. Lett. 93, 031109 (2008).
  • [18] M. A. Wayne, E. R. Jeffrey, G. M. Akselrod, and P. G. Kwiat, “Photon arrival time quantum random number generation,” \JournalTitleJ. Mod. Opt. 56, 516–522 (2009).
  • [19] H. Fürst, H. Weier, S. Nauerth, et al., “High speed optical quantum random number generation,” \JournalTitleOpt. Express 18, 13029–13037 (2010).
  • [20] M. Ren, E. Wu, Y. Liang, et al., “Quantum random-number generator based on a photon-number-resolving detector,” \JournalTitlePhys. Rev. A 83, 023820 (2011).
  • [21] Y. Jian, M. Ren, E. Wu, et al., “Two-bit quantum random number generator based on photon-number-resolving detection,” \JournalTitleRev. Sci. Instrum. 82, 073109 (2011).
  • [22] S. Tisa, F. Villa, A. Giudice, et al., “High-speed quantum random number generation using CMOS photon counting detectors,” \JournalTitleIEEE J. Sel. Top. Quantum Electron. 21, 23–29 (2015).
  • [23] W. Wei and H. Guo, “Bias-free true random-number generator,” \JournalTitleOpt. Lett. 34, 1876–1878 (2009).
  • [24] Z. Bisadi, A. Meneghetti, G. Fontana, et al., “Quantum random number generator based on silicon nanocrystals LED,” in Integrated Photonics: Materials, Devices, and Applications III, vol. 9520 J.-M. Fédéli, ed. (2015), p. 952004.
  • [25] Y. Shen, L. Tian, and H. Zou, “Practical quantum random number generator based on measuring the shot noise of vacuum states,” \JournalTitlePhys. Rev. A 81, 063814 (2010).
  • [26] C. Gabriel, C. Wittmann, D. Sych, et al., ‘‘A generator for unique quantum random numbers based on vacuum states,” \JournalTitleNat. Photonics 4, 711–715 (2010).
  • [27] Y. Zhu, G. He, and G. Zeng, “Unbiased quantum random number generation based on squeezed vacuum state,” \JournalTitleInt. J. Quantum Inf. 10, 1250012 (2012).
  • [28] H. Guo, W. Tang, Y. Liu, and W. Wei, “Truly random number generation based on measurement of phase noise of a laser,” \JournalTitlePhys. Rev. E 81, 051137 (2010).
  • [29] B. Qi, Y.-M. Chi, H.-K. Lo, and L. Qian, “High-speed quantum random number generation by measuring phase noise of a single-mode laser,” \JournalTitleOpt. Lett. 35, 312–314 (2010).
  • [30] Y.-Q. Nie, L. Huang, Y. Liu, et al., “The generation of 68 Gbps quantum random number by measuring laser phase fluctuations,” \JournalTitleRev. Sci. Instrum. 86, 063105 (2015).
  • [31] C. R. S. Williams, J. C. Salevan, X. Li, et al., ‘‘Fast physical random number generator using amplified spontaneous emission,” \JournalTitleOpt. Express 18, 23584–23597 (2010).
  • [32] P. J. Bustard, D. Moffatt, R. Lausten, et al., “Quantum random bit generation using stimulated Raman scattering,” \JournalTitleOpt. Express 19, 25173–25180 (2011).
  • [33] A. Sarkar and C. M. Chandrashekar, “Multi-bit quantum random number generation from a single qubit quantum walk,” \JournalTitleSci. Rep. 9, 12323 (2019).
  • [34] Y. Zhang, Z. Li, Z. Chen, et al., “Continuous-variable QKD over 50 km commercial fiber,” \JournalTitleQuantum Sci. Technol. 4, 035006 (2019).
  • [35] Y. Zhang, Z. Chen, S. Pirandola, et al., “Long-distance continuous-variable quantum key distribution over 202.81 km of fiber,” \JournalTitlePhys. Rev. Lett. 125, 010502 (2020).
  • [36] M. Huang, Z. Chen, Y. Zhang, and H. Guo, ‘‘A Gaussian-distributed quantum random number generator using vacuum shot noise,” \JournalTitleEntropy 22 (2020).
  • [37] D. B. Thomas, W. Luk, P. H. Leong, and J. D. Villasenor, “Gaussian random number generators,” \JournalTitleACM Comput. Surv. 39, 11–49 (2007).
  • [38] R. Zhang, R. Yang, J. Guo, et al., “Arbitrary coherent distributions in a programmable quantum walk,” \JournalTitlePhys. Rev. Res. 4, 023042 (2022).
  • [39] J. Biamonte, P. Wittek, N. Pancotti, et al., “Quantum machine learning,” \JournalTitleNature 549, 195–202 (2017).
  • [40] P. Palittapongarnpim, P. Wittek, E. Zahedinejad, et al., “Learning in quantum control: High-dimensional global optimization for noisy quantum dynamics,” \JournalTitleNeurocomputing 268, 116–126 (2017).
  • [41] I. Kerenidis and A. Prakash, ‘‘Quantum gradient descent for linear systems and least squares,” \JournalTitlePhys. Rev. A 101, 022316 (2020).
  • [42] M. Cai, Y. Lu, M. Xiao, and K. Xia, “Optimizing single-photon generation and storage with machine learning,” \JournalTitlePhys. Rev. A 104, 053707 (2021).
  • [43] S. Ruder, “An overview of gradient descent optimization algorithms,” \JournalTitlearXiv: 1609, 04747 (2016).
  • [44] H. Jeong, M. Paternostro, and M. S. Kim, “Simulation of quantum random walks using the interference of a classical field,” \JournalTitlePhys. Rev. A 69, 012310 (2004).
  • [45] H. B. Perets, Y. Lahini, F. Pozzi, et al., “Realization of quantum walks with negligible decoherence in waveguide lattices,” \JournalTitlePhys. Rev. Lett. 100, 170506 (2008).
  • [46] A. Peruzzo, M. Lobino, J. C. F. Matthews, et al., “Quantum walks of correlated photons,” \JournalTitleScience 329, 1500–1503 (2010).
  • [47] H. Tang, C. Di Franco, Z.-Y. Shi, et al., “Experimental quantum fast hitting on hexagonal graphs,” \JournalTitleNat. Photonics 12, 754–758 (2018).
  • [48] Q.-P. Su, Y. Zhang, L. Yu, et al., “Experimental demonstration of quantum walks with initial superposition states,” \JournalTitlenpj Quantum Inf. 5, 40 (2019).
  • [49] Z. Yan, Y.-R. Zhang, M. Gong, et al., “Strongly correlated quantum walks with a 12-qubit superconducting processor,” \JournalTitleScience 364, 753–756 (2019).
  • [50] Q.-P. Su, S.-C. Wang, Y. Chi, et al., “Implementing quantum walks with a single qubit,” \JournalTitlearXiv: 2206, 03642 (2022).
  • [51] M. Eaton, A. Hossameldin, R. J. Birrittella, et al., “Resolution of 100 photons and quantum generation of unbiased random numbers,” \JournalTitleNat. Photonics 17, 106–111 (2022).
  • [52] Z. Tang, Z. Liao, F. Xu, et al., “Experimental demonstration of polarization encoding measurement-device-independent quantum key distribution,” \JournalTitlePhys. Rev. Lett. 112, 190503 (2014).
  • [53] A. Vaziri, J.-W. Pan, T. Jennewein, et al., “Concentration of higher dimensional entanglement: Qutrits of photon orbital angular momentum,” \JournalTitlePhys. Rev. Lett. 91, 227902 (2003).
  • [54] X.-L. Wang, Y.-H. Luo, H.-L. Huang, et al., “18-qubit entanglement with six photons’ three degrees of freedom,” \JournalTitlePhys. Rev. Lett. 120, 260502 (2018).
  • [55] Z.-F. Liu, C. Chen, J.-M. Xu, et al., “Hong-Ou-Mandel interference between two hyperentangled photons enables observation of symmetric and antisymmetric particle exchange phases,” \JournalTitlePhys. Rev. Lett. 129, 263602 (2022).
  • [56] Z.-X. Li, D. Zhu, P.-C. Lin, et al., “High-dimensional entanglement generation based on a Pancharatnam-Berry phase metasurface,” \JournalTitlePhotonics Res. 10, 2702–2707 (2022).
  • [57] M. Grafe, R. Heilmann, A. Perez-Leija, et al., “On-chip generation of high-order single-photon W-states,” \JournalTitleNat. Photonics 8, 791–795 (2014).