Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: arXiv.org perpetual non-exclusive license
arXiv:2306.02660v2 [math.NA] 11 Mar 2024

Automated Importance Sampling via Optimal Control for Stochastic Reaction Networks: A Markovian Projection–based Approach

Chiheb Ben Hammouda Utrecht University, Mathematical Institute, 3584 CD Utrecht, The Netherlands. Nadhir Ben Rached University of Leeds, School of Mathematics, Woodhouse, Leeds LS2 9JT, UK. Raúl Tempone King Abdullah University of Science and Technology (KAUST), Computer, Electrical and Mathematical Sciences & Engineering Division (CEMSE), Thuwal 23955-6900, Saudi Arabia. RWTH Aachen University, Alexander von Humboldt Professor in Mathematics for Uncertainty Quantification, 52062 Aachen, Germany. Sophia Wiechert Corresponding author: wiechert@uq.rwth-aachen.de RWTH Aachen University, Chair of Mathematics for Uncertainty Quantification, Pontdriesch 14-16, 52062 Aachen, Germany.
Abstract

We propose a novel alternative approach to our previous work (Ben Hammouda et al., 2023) to improve the efficiency of Monte Carlo (MC) estimators for rare event probabilities for stochastic reaction networks (SRNs). In the same spirit of (Ben Hammouda et al., 2023), an efficient path-dependent measure change is derived based on a connection between determining optimal importance sampling (IS) parameters within a class of probability measures and a stochastic optimal control formulation, corresponding to solving a variance minimization problem. In this work, we propose a novel approach to address the encountered curse of dimensionality by mapping the problem to a significantly lower-dimensional space via a Markovian projection (MP) idea. The output of this model reduction technique is a low-dimensional SRN (potentially even one dimensional) that preserves the marginal distribution of the original high-dimensional SRN system. The dynamics of the projected process are obtained by solving a related optimization problem via a discrete L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT regression. By solving the resulting projected Hamilton–Jacobi–Bellman (HJB) equations for the reduced-dimensional SRN, we obtain projected IS parameters, which are then mapped back to the original full-dimensional SRN system, resulting in an efficient IS-MC estimator for rare events probabilities of the full-dimensional SRN. Our analysis and numerical experiments reveal that the proposed MP-HJB-IS approach substantially reduces the MC estimator variance, resulting in a lower computational complexity in the rare event regime than standard MC estimators.

Keywords: stochastic reaction networks, tau-leap, importance sampling, stochastic optimal control, Markovian projection, rare event.

1 Introduction

This paper proposes an efficient estimator for rare event probabilities for a particular class of continuous-time Markov processes, stochastic reaction networks (SRNs). We design an automated importance sampling (IS) approach based on the approximate explicit tau-leap (TL) scheme to build a Monte Carlo (MC) estimator for rare event probabilities of SRNs. The used IS change of measure was introduced in [11], wherein the optimal IS controls were determined via a stochastic optimal control (SOC) formulation. In that same work, we also presented a learning-based approach to avoid the curse of dimensionality. Building on that work, we propose an alternative method for high-dimensional SRNs that leverages dimension reduction through Markovian projection (MP) and then recovers the optimal IS controls of the full-dimensional SRNs as a mapping from the solution in lower-dimensional space, potentially one. To the best of our knowledge, we are the first to establish the MP framework for the SRN setting to solve an IS problem.

An SRN (refer to Section 1.1 for a brief introduction and [9] for more details) describes the time evolution of a set of species through reactions and can be found in a wide range of applications, such as biochemical reactions, epidemic processes [15, 5], and transcription and translation in genomics and virus kinetics [48, 32]. For a d𝑑ditalic_d-dimensional SRN, 𝐗:[0,T]d:𝐗0𝑇superscript𝑑\mathbf{X}:[0,T]\rightarrow\mathbb{N}^{d}bold_X : [ 0 , italic_T ] → blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, with the given final time T>0𝑇0T>0italic_T > 0, we aim to determine accurate and computationally efficient MC estimations for the expected value 𝔼[g(𝐗(T))]𝔼delimited-[]𝑔𝐗𝑇\mathbb{E}[g(\mathbf{X}(T))]blackboard_E [ italic_g ( bold_X ( italic_T ) ) ]. The observable g:d:𝑔superscript𝑑g:\mathbb{N}^{d}\to\mathbb{R}italic_g : blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT → blackboard_R is a given scalar function of 𝐗𝐗\mathbf{X}bold_X, where indicator functions g(𝐱)=𝟙{𝐱}𝑔𝐱subscript1𝐱g(\mathbf{x})=\mathbbm{1}_{\{\mathbf{x}\in\mathcal{B}\}}italic_g ( bold_x ) = blackboard_1 start_POSTSUBSCRIPT { bold_x ∈ caligraphic_B } end_POSTSUBSCRIPT are of interest to estimate the rare event probability (𝐗(T))1much-less-than𝐗𝑇1\mathbb{P}(\mathbf{X}(T)\in\mathcal{B})\ll 1blackboard_P ( bold_X ( italic_T ) ∈ caligraphic_B ) ≪ 1, where dsuperscript𝑑\mathcal{B}\subset\mathbb{R}^{d}caligraphic_B ⊂ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT.

The quantity of interest, 𝔼[g(𝐗(T))]𝔼delimited-[]𝑔𝐗𝑇\mathbb{E}[g(\mathbf{X}(T))]blackboard_E [ italic_g ( bold_X ( italic_T ) ) ], is the solution to the corresponding Kolmogorov backward equations [8]. Solving these ordinary differential equations (ODEs) in closed form is infeasible for most SRNs; thus, numerical approximations based on discretized schemes are used to derive solutions. A drawback of these approaches is that, without using dimension reduction techniques, the computational cost scales exponentially with the number of species d𝑑ditalic_d. To avoid the curse of dimensionality, we propose estimating 𝔼[g(𝐗(T))]𝔼delimited-[]𝑔𝐗𝑇\mathbb{E}[g(\mathbf{X}(T))]blackboard_E [ italic_g ( bold_X ( italic_T ) ) ] using MC methods.

Numerous schemes have been developed to simulate the exact sample paths of SRNs. These include the stochastic simulation algorithm introduced by Gillespie in [26] and the modified next reaction method proposed by Anderson in [3]. However, when SRNs involve reaction channels with high reaction rates, simulating exact realizations of the system can be computationally expensive. To address this issue, Gillespie [27] and Aparicio and Solari [6] independently proposed the explicit-TL method (see Section 1.2), which approximates the paths of 𝐗𝐗\mathbf{X}bold_X by evolving the process with fixed time steps while maintaining constant reaction rates within each time step. Additionally, other simulation schemes have been proposed to handle situations with well-separated fast and slow time scales [16, 45, 1, 2, 42, 12].

In order to compute MC estimates of 𝔼[g(𝐗(T))]𝔼delimited-[]𝑔𝐗𝑇\mathbb{E}[g(\mathbf{X}(T))]blackboard_E [ italic_g ( bold_X ( italic_T ) ) ] more efficiently, different variance reduction techniques have been proposed in the context of SRNs. In the spirit of the multilevel MC (MLMC) idea [22, 23], various MLMC-based methods [4, 38, 42, 12, 10] have been introduced to overcome different challenges in this context. Moreover, as the naive MC and MLMC estimators have high computational costs when used for estimating rare event probabilities, different IS approaches [36, 25, 47, 19, 17, 24, 46, 11] have been proposed.

To estimate various statistical quantities efficiently for SRNs (specifically rare event probabilities), we use the path-dependent IS approach originally introduced in [11]. This class of probability measure change is based on modifying the rates of the Poisson random variables used to construct the TL paths. In [11], it is shown how optimal IS controls are obtained by minimizing the second moment of the IS estimator (equivalently, the variance), representing the cost function of the associated SOC problem, and that the corresponding value function solves a dynamic programming relation (see Section 2.1 for revising these results). In this work, we generalize the discrete-time dynamic programming relation by a set of continuous-time ODEs, the Hamilton–Jacobi–Bellman (HJB) equations, allowing the formulation of optimal IS controls in continuous time. Compared to the discrete-time IS control formulation presented in [11], the continuous-time formulation offers the advantage that it provides a curve of IS controls over time instead of a discrete set. This allows its application for any time stepping in the IS-TL paths and thereby eliminates the need for ad-hoc interpolations often needed in the discrete setting.

In the multidimensional setting, the cost of solving the backward HJB equations increases exponentially with respect to the dimension d𝑑ditalic_d (curse of dimensionality). In [11], we proposed a learning-based approach to reduce this effect. In that approach, the value function is approximated using an ansatz function, the parameters of which are learned through a stochastic optimization algorithm (see Figure 1.1 for a schematic illustration of the approach). In this work, we present an alternative method using a dimension reduction approach for SRNs (see Figure 1.2 for a schematic illustration of the approach). The proposed methodology is to adapt the MP idea originally introduced in [29] for the setting of diffusion-type stochastic differential equations (SDEs) to the SRN framework, resulting in a significantly lower-dimensional process, preserving the marginal distribution of the original full-dimensional SRN. The propensities characterizing the lower-dimensional MP process can be approximated using L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT regression. Using the resulting low-dimensional SRN, we derive an approximate value function and, consequently, near-optimal IS controls while reducing the effect of the curse of dimensionality. By mapping the IS controls to the original full-dimensional SRNs, we derive an unbiased IS-MC estimator for the TL scheme. Compared to the learning-based approach presented in [11], this novel MP-IS approach eliminates the need for an ansatz function to model the value function. This approach allows its application to general observables g𝑔gitalic_g that differ from indicator functions for rare event estimation, because no prior knowledge regarding the shape of the value function and suitable ansatz functions is required.

To the best of our knowledge, we are the first to establish the MP idea for SRNs and apply it to derive an efficient pathwise IS for MC methods. Initially, the MP idea was introduced for Itô stochastic processes in [34, 29] and was later generalized to martingales and semimartingales [35, 14]. In addition, MP has been widely applied for dimension reduction in SDEs [37], particularly in financial applications [43, 20]. For instance, in [7], solving HJB equations for an MP process was pursued but in the setting of Itô SDEs with the application of pricing American options. In [31], MP was used for control problems and IS problems for rare events in high-dimensional diffusion processes with multiple time scales. In this work, we introduce the general dimension reduction framework of MP for SRNs such that it can be applied to other problems beyond the selected IS application. (e.g., solving the chemical master equation [40] or the Kolmogorov backward equations [41]).

The remainder of this work is organized as follows. Sections 1.1, 1.2, 1.3, and 1.4 recall the relevant SRN, TL, MC, and IS notations and definitions from [11]. Next, Section 2 reviews the connection between IS and SOC by introducting the IS scheme, the value function, and the corresponding dynamic programming theorem from [11] in Section 2.1. Then, Section 2.2 extends the framework to a continuous-time formulation leading to the continuous-time value function and deriving the corresponding HJB equations. Section 3 presents the MP technique for SRNs and shows how the projected dynamics can be computed using L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT regression. Next, Section 4 addresses the curse of dimensionality of high-dimensional SRNs occurring from the optimal IS scheme in Section 2 by combining the IS scheme with MP (Section 3) to derive near-optimal IS controls. Finally, Section 5 presents the numerical results for the rare event probability estimation to demonstrate the efficiency of the proposed MP-IS approach compared to a standard TL-MC estimator.

Figure 1.1: Schematic diagram of the learning-based approach in [11].
Refer to caption
Figure 1.2: Schematic diagram of the MP-IS approach.
Refer to caption

1.1 Stochastic Reaction Networks (SRNs)

We recall from [11] that an SRN describes the time evolution for a homogeneously mixed chemical reaction system, in which d𝑑ditalic_d distinct species interact through J𝐽Jitalic_J reaction channels. Each reaction channel jsubscript𝑗\mathcal{R}_{j}caligraphic_R start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , j=1,J𝑗1𝐽j=1\dots,Jitalic_j = 1 … , italic_J, is given by the relation

αj,1S1++αj,dSdθjβj,1S1++βj,dSd,subscript𝛼𝑗1subscript𝑆1subscript𝛼𝑗𝑑subscript𝑆𝑑subscript𝜃𝑗subscript𝛽𝑗1subscript𝑆1subscript𝛽𝑗𝑑subscript𝑆𝑑\alpha_{j,1}S_{1}+\dots+\alpha_{j,d}S_{d}\overset{\theta_{j}}{\rightarrow}% \beta_{j,1}S_{1}+\dots+\beta_{j,d}S_{d},italic_α start_POSTSUBSCRIPT italic_j , 1 end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + ⋯ + italic_α start_POSTSUBSCRIPT italic_j , italic_d end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_OVERACCENT italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG italic_β start_POSTSUBSCRIPT italic_j , 1 end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + ⋯ + italic_β start_POSTSUBSCRIPT italic_j , italic_d end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT , (1.1)

where αj,isubscript𝛼𝑗𝑖\alpha_{j,i}italic_α start_POSTSUBSCRIPT italic_j , italic_i end_POSTSUBSCRIPT molecules of species Sisubscript𝑆𝑖S_{i}italic_S start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are consumed and βj,isubscript𝛽𝑗𝑖\beta_{j,i}italic_β start_POSTSUBSCRIPT italic_j , italic_i end_POSTSUBSCRIPT molecules are produced. The positive constants {θj}j=1Jsuperscriptsubscriptsubscript𝜃𝑗𝑗1𝐽\{\theta_{j}\}_{j=1}^{J}{ italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT represent the reaction rates.

This process can be modeled by a Markovian pure jump process, 𝐗:[0,T]×Ωd:𝐗0𝑇Ωsuperscript𝑑\mathbf{X}:[0,T]\times\Omega\to\mathbb{N}^{d}bold_X : [ 0 , italic_T ] × roman_Ω → blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, where (ΩΩ\Omegaroman_Ω, \mathcal{F}caligraphic_F, \mathbb{P}blackboard_P) is a probability space. We are interested in the time evolution of the state vector,

𝐗(t)=(X1(t),,Xd(t))d,𝐗𝑡subscript𝑋1𝑡subscript𝑋𝑑𝑡superscript𝑑\mathbf{X}(t)=\left(X_{1}(t),\ldots,X_{d}(t)\right)\in\mathbb{N}^{d},bold_X ( italic_t ) = ( italic_X start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) , … , italic_X start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_t ) ) ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT , (1.2)

where the i𝑖iitalic_i-th component, Xi(t)subscript𝑋𝑖𝑡X_{i}(t)italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ), describes the abundance of the i𝑖iitalic_ith species present in the system at time t𝑡titalic_t. The process 𝐗𝐗\mathbf{X}bold_X is a continuous-time, discrete-space Markov process characterized by Kurtz’s random time change representation [21]:

𝐗(t)=𝐱0+j=1JYj(0taj(𝐗(s))𝑑s)𝝂j,𝐗𝑡subscript𝐱0superscriptsubscript𝑗1𝐽subscript𝑌𝑗superscriptsubscript0𝑡subscript𝑎𝑗𝐗𝑠differential-d𝑠subscript𝝂𝑗\mathbf{X}(t)=\mathbf{x}_{0}+\sum_{j=1}^{J}Y_{j}\left(\int_{0}^{t}a_{j}(% \mathbf{X}(s))\,ds\right)\boldsymbol{\nu}_{j},bold_X ( italic_t ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_Y start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_s ) ) italic_d italic_s ) bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , (1.3)

where Yj:+×Ω:subscript𝑌𝑗subscriptΩY_{j}:\mathbb{R}_{+}{\times}\Omega\to\mathbb{N}italic_Y start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT : blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT × roman_Ω → blackboard_N are independent unit-rate Poisson processes and the stoichiometric vector is defined as 𝝂j=(βj,1αj,1,,βj,dαj,d)dsubscript𝝂𝑗subscript𝛽𝑗1subscript𝛼𝑗1subscript𝛽𝑗𝑑subscript𝛼𝑗𝑑superscript𝑑\boldsymbol{\nu}_{j}=\left(\beta_{j,1}-\alpha_{j,1},\dots,\beta_{j,d}-\alpha_{% j,d}\right)\in\mathbb{Z}^{d}bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = ( italic_β start_POSTSUBSCRIPT italic_j , 1 end_POSTSUBSCRIPT - italic_α start_POSTSUBSCRIPT italic_j , 1 end_POSTSUBSCRIPT , … , italic_β start_POSTSUBSCRIPT italic_j , italic_d end_POSTSUBSCRIPT - italic_α start_POSTSUBSCRIPT italic_j , italic_d end_POSTSUBSCRIPT ) ∈ blackboard_Z start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT.

For each reaction channel jsubscript𝑗\mathcal{R}_{j}caligraphic_R start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, the propensity function aj:d+:subscript𝑎𝑗superscript𝑑subscripta_{j}:\mathbb{N}^{d}\rightarrow\mathbb{R}_{+}italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT : blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT → blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT obeys the non-negativity assumption (i.e. the system can not attain negative states), which is that aj(𝐗)=0subscript𝑎𝑗𝐗0a_{j}(\mathbf{X})=0italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ) = 0 for 𝐱𝐱\mathbf{x}bold_x such that 𝐗+νjd𝐗subscript𝜈𝑗superscript𝑑\mathbf{X}+\mathbf{\nu}_{j}\notin\mathbb{N}^{d}bold_X + italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ∉ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT. In our numerical simulations, we consider a propensity derived from the stochastic mass-action kinetic principle

aj(𝐱):=θji=1dxi!(xiαj,i)!𝟏{xiαj,i},assignsubscript𝑎𝑗𝐱subscript𝜃𝑗superscriptsubscriptproduct𝑖1𝑑subscript𝑥𝑖subscript𝑥𝑖subscript𝛼𝑗𝑖subscript1subscript𝑥𝑖subscript𝛼𝑗𝑖a_{j}(\mathbf{x}):=\theta_{j}\prod_{i=1}^{d}\frac{x_{i}!}{(x_{i}-\alpha_{j,i})% !}\mathbf{1}_{\{x_{i}\geq\alpha_{j,i}\}},italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) := italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ∏ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT divide start_ARG italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ! end_ARG start_ARG ( italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_α start_POSTSUBSCRIPT italic_j , italic_i end_POSTSUBSCRIPT ) ! end_ARG bold_1 start_POSTSUBSCRIPT { italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≥ italic_α start_POSTSUBSCRIPT italic_j , italic_i end_POSTSUBSCRIPT } end_POSTSUBSCRIPT , (1.4)

where xisubscript𝑥𝑖x_{i}italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the counting number for species Sisubscript𝑆𝑖S_{i}italic_S start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. However, the approach presented in this work is not restricted to the particular structure of the propensity function in (1.4) (see Remark 3.4).

1.2 Explicit Tau-Leap Approximation

The explicit-TL scheme is a pathwise approximate method based on Kurtz’s random time change representation (1.3) [27, 6]. It was originally introduced to overcome the computational drawbacks of exact methods, which become computationally expensive when many reactions fire during a short time interval. For a uniform time mesh {t0=0,t1,,tN=T}formulae-sequencesubscript𝑡00subscript𝑡1subscript𝑡𝑁𝑇\{t_{0}=0,t_{1},...,t_{N}=T\}{ italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 , italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_t start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = italic_T } with step size Δt=TNΔ𝑡𝑇𝑁\Delta t=\frac{T}{N}roman_Δ italic_t = divide start_ARG italic_T end_ARG start_ARG italic_N end_ARG and a given initial value 𝐗(0)=𝐱0𝐗0subscript𝐱0\mathbf{X}(0)=\mathbf{x}_{0}bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , the explicit-TL approximation for 𝐗𝐗\mathbf{X}bold_X is defined by

𝐗^0subscript^𝐗0\displaystyle\widehat{\mathbf{X}}_{0}over^ start_ARG bold_X end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT :=𝐱0assignabsentsubscript𝐱0\displaystyle:=\mathbf{x}_{0}:= bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT
𝐗^kΔtsubscriptsuperscript^𝐗Δ𝑡𝑘\displaystyle\widehat{\mathbf{X}}^{\Delta t}_{k}over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT :=max(𝟎,𝐗^k1Δt+j=1J𝒫k1,j(aj(𝐗^k1Δt)Δt)𝝂j), 1kN,formulae-sequenceassignabsent𝟎subscriptsuperscript^𝐗Δ𝑡𝑘1superscriptsubscript𝑗1𝐽subscript𝒫𝑘1𝑗subscript𝑎𝑗subscriptsuperscript^𝐗Δ𝑡𝑘1Δ𝑡subscript𝝂𝑗1𝑘𝑁\displaystyle:=\max\left(\textbf{0},\widehat{\mathbf{X}}^{\Delta t}_{k-1}+\sum% _{j=1}^{J}\mathcal{P}_{k-1,j}\left(a_{j}(\widehat{\mathbf{X}}^{\Delta t}_{k-1}% )\Delta t\right)\boldsymbol{\nu}_{j}\right),\>1\leq k\leq N,:= roman_max ( 0 , over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT caligraphic_P start_POSTSUBSCRIPT italic_k - 1 , italic_j end_POSTSUBSCRIPT ( italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k - 1 end_POSTSUBSCRIPT ) roman_Δ italic_t ) bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) , 1 ≤ italic_k ≤ italic_N , (1.5)

where {𝒫k,j(rk,j)}{1jJ}subscriptsubscript𝒫𝑘𝑗subscript𝑟𝑘𝑗1𝑗𝐽\{\mathcal{P}_{k,j}(r_{k,j})\}_{\{1\leq j\leq J\}}{ caligraphic_P start_POSTSUBSCRIPT italic_k , italic_j end_POSTSUBSCRIPT ( italic_r start_POSTSUBSCRIPT italic_k , italic_j end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT { 1 ≤ italic_j ≤ italic_J } end_POSTSUBSCRIPT are independent Poisson random variables with respective rates rk,j:=aj(𝐗^kΔt)Δtassignsubscript𝑟𝑘𝑗subscript𝑎𝑗subscriptsuperscript^𝐗Δ𝑡𝑘Δ𝑡r_{k,j}:=a_{j}(\widehat{\mathbf{X}}^{\Delta t}_{k})\Delta titalic_r start_POSTSUBSCRIPT italic_k , italic_j end_POSTSUBSCRIPT := italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) roman_Δ italic_t conditioned on the current state 𝐗^kΔtsubscriptsuperscript^𝐗Δ𝑡𝑘\widehat{\mathbf{X}}^{\Delta t}_{k}over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. The maximum in (1.2) is applied entry-wise. In each TL step, the current state is projected to zero to prevent the process from exiting the lattice (i.e., producing negative values).

1.3 Biased Monte Carlo estimator

We let 𝐗𝐗\mathbf{X}bold_X be an SRN and g:d:𝑔superscript𝑑g:\mathbb{R}^{d}\rightarrow\mathbb{R}italic_g : blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT → blackboard_R be a scalar observable. For a given final time T𝑇Titalic_T, we estimate 𝔼[g(𝐗(T))]𝔼delimited-[]𝑔𝐗𝑇\mathbb{E}\left[g(\mathbf{X}(T))\right]blackboard_E [ italic_g ( bold_X ( italic_T ) ) ] using the standard MC-TL estimator:

μM:=1Mm=1Mg(𝐗^[m]Δt(T)),assignsubscript𝜇𝑀1𝑀superscriptsubscript𝑚1𝑀𝑔subscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑇\mu_{M}:=\frac{1}{M}\sum_{m=1}^{M}g(\widehat{\mathbf{X}}^{\Delta t}_{[m]}(T)),italic_μ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT := divide start_ARG 1 end_ARG start_ARG italic_M end_ARG ∑ start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT italic_g ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] end_POSTSUBSCRIPT ( italic_T ) ) , (1.6)

where {𝐗^[m]Δt(T)}m=1Msuperscriptsubscriptsubscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑇𝑚1𝑀\{\widehat{\mathbf{X}}^{\Delta t}_{[m]}(T)\}_{m=1}^{M}{ over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] end_POSTSUBSCRIPT ( italic_T ) } start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT are independent TL samples.

The global error for the proposed MC estimator has the following error decomposition:

|𝔼[g(𝐗(T))]μM||𝔼[g(𝐗(T))]𝔼[g(𝐗^Δt(T))]|Bias+|𝔼[g(𝐗^Δt(T))]μM|Statistical Error.𝔼delimited-[]𝑔𝐗𝑇subscript𝜇𝑀subscript𝔼delimited-[]𝑔𝐗𝑇𝔼delimited-[]𝑔superscript^𝐗Δ𝑡𝑇Biassubscript𝔼delimited-[]𝑔superscript^𝐗Δ𝑡𝑇subscript𝜇𝑀Statistical Error\displaystyle\left|\mathbb{E}[g(\mathbf{X}(T))]-\mu_{M}\right|\leq\underbrace{% \left|\mathbb{E}[g(\mathbf{X}(T))]-\mathbb{E}[g(\widehat{\mathbf{X}}^{\Delta t% }(T))]\right|}_{\text{Bias}}+\underbrace{\left|\mathbb{E}[g(\widehat{\mathbf{X% }}^{\Delta t}(T))]-\mu_{M}\right|}_{\text{Statistical Error}}.| blackboard_E [ italic_g ( bold_X ( italic_T ) ) ] - italic_μ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT | ≤ under⏟ start_ARG | blackboard_E [ italic_g ( bold_X ( italic_T ) ) ] - blackboard_E [ italic_g ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( italic_T ) ) ] | end_ARG start_POSTSUBSCRIPT Bias end_POSTSUBSCRIPT + under⏟ start_ARG | blackboard_E [ italic_g ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( italic_T ) ) ] - italic_μ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT | end_ARG start_POSTSUBSCRIPT Statistical Error end_POSTSUBSCRIPT . (1.7)

Under some assumptions, the TL scheme has a weak order, 𝒪(Δt)𝒪Δ𝑡{\mathcal{O}}\left(\Delta t\right)caligraphic_O ( roman_Δ italic_t ) [39], that is, for sufficiently small ΔtΔ𝑡\Delta troman_Δ italic_t,

|𝔼[g(𝐗(T))g(𝐗^Δt(T))]|CΔt𝔼delimited-[]𝑔𝐗𝑇𝑔superscript^𝐗Δ𝑡𝑇𝐶Δ𝑡\displaystyle\left|\mathbb{E}\left[g(\mathbf{X}(T))-g(\widehat{\mathbf{X}}^{% \Delta t}(T))\right]\right|\leq C\Delta t| blackboard_E [ italic_g ( bold_X ( italic_T ) ) - italic_g ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( italic_T ) ) ] | ≤ italic_C roman_Δ italic_t (1.8)

where C>0𝐶0C>0italic_C > 0.

The bias and statistical error can be bound equally using TOL2𝑇𝑂𝐿2\frac{TOL}{2}divide start_ARG italic_T italic_O italic_L end_ARG start_ARG 2 end_ARG to achieve the desired accuracy, TOL, with a confidence level of 1α1𝛼1-\alpha1 - italic_α for α(0,1)𝛼01\alpha\in(0,1)italic_α ∈ ( 0 , 1 ), which can be achieved by the step size:

Δt(TOL)=TOL2CΔ𝑡TOLTOL2𝐶\displaystyle\Delta t(\text{TOL})=\frac{\text{TOL}}{2\cdot C}roman_Δ italic_t ( TOL ) = divide start_ARG TOL end_ARG start_ARG 2 ⋅ italic_C end_ARG (1.9)

and

M*(TOL)=Cα24Var[g(𝐗^Δt(T))]TOL2superscript𝑀TOLsuperscriptsubscript𝐶𝛼24Vardelimited-[]𝑔superscript^𝐗Δ𝑡𝑇superscriptTOL2\displaystyle M^{*}(\text{TOL})=C_{\alpha}^{2}\frac{4\cdot\text{Var}[g(% \widehat{\mathbf{X}}^{\Delta t}(T))]}{\text{TOL}^{2}}italic_M start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ( TOL ) = italic_C start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG 4 ⋅ Var [ italic_g ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( italic_T ) ) ] end_ARG start_ARG TOL start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (1.10)

sample paths, where the constant Cαsubscript𝐶𝛼C_{\alpha}italic_C start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is the (1α2)limit-from1𝛼2(1-\frac{\alpha}{2})-( 1 - divide start_ARG italic_α end_ARG start_ARG 2 end_ARG ) -quantile for the standard normal distribution. We select Cα=1.96subscript𝐶𝛼1.96C_{\alpha}=1.96italic_C start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = 1.96 for a 95%percent9595\%95 % confidence level corresponding to α=0.05𝛼0.05\alpha=0.05italic_α = 0.05.

When estimating rare event probabilities, we are interested in the relative error

|𝔼[g(𝐗(T))]μM||𝔼[g(𝐗(T))]|.𝔼delimited-[]𝑔𝐗𝑇subscript𝜇𝑀𝔼delimited-[]𝑔𝐗𝑇\displaystyle\frac{\left|\mathbb{E}[g(\mathbf{X}(T))]-\mu_{M}\right|}{\left|% \mathbb{E}[g(\mathbf{X}(T))]\right|}.divide start_ARG | blackboard_E [ italic_g ( bold_X ( italic_T ) ) ] - italic_μ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT | end_ARG start_ARG | blackboard_E [ italic_g ( bold_X ( italic_T ) ) ] | end_ARG .

In this context, to achieve a prescribed relative tolerance TOLrelsubscriptTOL𝑟𝑒𝑙\text{TOL}_{rel}TOL start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT, we use step size

Δtrel(TOLrel)=TOLrel|𝔼[g(𝐗(T))]|2CΔsubscript𝑡𝑟𝑒𝑙subscriptTOL𝑟𝑒𝑙subscriptTOL𝑟𝑒𝑙𝔼delimited-[]𝑔𝐗𝑇2𝐶\displaystyle\Delta t_{rel}(\text{TOL}_{rel})=\frac{\text{TOL}_{rel}\left|% \mathbb{E}[g(\mathbf{X}(T))]\right|}{2\cdot C}roman_Δ italic_t start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT ( TOL start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT ) = divide start_ARG TOL start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT | blackboard_E [ italic_g ( bold_X ( italic_T ) ) ] | end_ARG start_ARG 2 ⋅ italic_C end_ARG (1.11)

and

Mrel*(TOLrel)=Cα24Var[g(𝐗^Δt(T))]TOLrel2|𝔼[g(𝐗(T))]|2subscriptsuperscript𝑀𝑟𝑒𝑙subscriptTOL𝑟𝑒𝑙superscriptsubscript𝐶𝛼24Vardelimited-[]𝑔superscript^𝐗Δ𝑡𝑇superscriptsubscriptTOL𝑟𝑒𝑙2superscript𝔼delimited-[]𝑔𝐗𝑇2\displaystyle M^{*}_{rel}(\text{TOL}_{rel})=C_{\alpha}^{2}\frac{4\cdot\text{% Var}[g(\widehat{\mathbf{X}}^{\Delta t}(T))]}{\text{TOL}_{rel}^{2}\left|\mathbb% {E}[g(\mathbf{X}(T))]\right|^{2}}italic_M start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT ( TOL start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT ) = italic_C start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG 4 ⋅ Var [ italic_g ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( italic_T ) ) ] end_ARG start_ARG TOL start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | blackboard_E [ italic_g ( bold_X ( italic_T ) ) ] | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (1.12)

sample paths.

Given that the computational cost to simulate a single path is 𝒪(Δt1)𝒪Δsuperscript𝑡1{\mathcal{O}}\left({\Delta t}^{-1}\right)caligraphic_O ( roman_Δ italic_t start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ), the expected total computational complexity is 𝒪(TOL3)𝒪superscriptTOL3{\mathcal{O}}\left(\text{TOL}^{-3}\right)caligraphic_O ( TOL start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ) and 𝒪(TOLrel3)𝒪superscriptsubscriptTOL𝑟𝑒𝑙3{\mathcal{O}}\left(\text{TOL}_{rel}^{-3}\right)caligraphic_O ( TOL start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ) for the absolute and relative errors, respectively.

1.4 Importance Sampling

Using IS techniques [36, 25, 24, 19, 17, 24, 46] can improve the computational costs for the crude MC estimator through variance reduction in (1.10). For a general motivation, we refer to [11] Section 1.4. For illustrating the IS method, let us consider the general problem of estimating 𝔼[g(Y)]𝔼delimited-[]𝑔𝑌\mathbb{E}[g(Y)]blackboard_E [ italic_g ( italic_Y ) ], where g𝑔gitalic_g is a given observable and Y𝑌Yitalic_Y is a random variable taking values in \mathbb{R}blackboard_R with the probability density function ρYsubscript𝜌𝑌\rho_{Y}italic_ρ start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT. We let ρ^Zsubscript^𝜌𝑍\widehat{\rho}_{Z}over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT be the probability density function for an auxiliary real random variable Z𝑍Zitalic_Z. The MC estimator under the IS measure is

μMIS=1Mj=1ML(Z[j])g(Z[j]),superscriptsubscript𝜇𝑀𝐼𝑆1𝑀superscriptsubscript𝑗1𝑀𝐿subscript𝑍delimited-[]𝑗𝑔subscript𝑍delimited-[]𝑗\displaystyle\mu_{M}^{IS}=\frac{1}{M}\sum_{j=1}^{M}L(Z_{[j]})\cdot g(Z_{[j]}),italic_μ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_I italic_S end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_M end_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT italic_L ( italic_Z start_POSTSUBSCRIPT [ italic_j ] end_POSTSUBSCRIPT ) ⋅ italic_g ( italic_Z start_POSTSUBSCRIPT [ italic_j ] end_POSTSUBSCRIPT ) , (1.13)

where Z[j]subscript𝑍delimited-[]𝑗Z_{[j]}italic_Z start_POSTSUBSCRIPT [ italic_j ] end_POSTSUBSCRIPT are independent and identically distributed samples from ρ^Zsubscript^𝜌𝑍\widehat{\rho}_{Z}over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT for j=1,,M𝑗1𝑀j=1,\dots,Mitalic_j = 1 , … , italic_M and the likelihood factor is given by L(Z[j]):=ρY(Z[j])ρ^Z(Z[j])assign𝐿subscript𝑍delimited-[]𝑗subscript𝜌𝑌subscript𝑍delimited-[]𝑗subscript^𝜌𝑍subscript𝑍delimited-[]𝑗L(Z_{[j]}):=\frac{\rho_{Y}(Z_{[j]})}{\widehat{\rho}_{Z}(Z_{[j]})}italic_L ( italic_Z start_POSTSUBSCRIPT [ italic_j ] end_POSTSUBSCRIPT ) := divide start_ARG italic_ρ start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_Z start_POSTSUBSCRIPT [ italic_j ] end_POSTSUBSCRIPT ) end_ARG start_ARG over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT ( italic_Z start_POSTSUBSCRIPT [ italic_j ] end_POSTSUBSCRIPT ) end_ARG. The IS estimator retains the expected value of (1.6) (i.e., 𝔼[L(Z)g(Z)]=𝔼[g(Y)]𝔼delimited-[]𝐿𝑍𝑔𝑍𝔼delimited-[]𝑔𝑌\mathbb{E}[L(Z)g(Z)]=\mathbb{E}[g(Y)]blackboard_E [ italic_L ( italic_Z ) italic_g ( italic_Z ) ] = blackboard_E [ italic_g ( italic_Y ) ]), but the variance can be reduced due to a different second moment 𝔼[(L(Z)g(Z))2]𝔼delimited-[]superscript𝐿𝑍𝑔𝑍2\mathbb{E}\left[\left(L(Z)\cdot g(Z)\right)^{2}\right]blackboard_E [ ( italic_L ( italic_Z ) ⋅ italic_g ( italic_Z ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ].

Determining an auxiliary probability measure that substantially reduces the variance compared with the original measure is challenging and strongly depends on the structure of the considered problem. In addition, the derivation of the new measure must come with a moderate additional computational cost to ensure an efficient IS scheme. This work uses the path-dependent change of probability measure introduced in [11], employing an IS measure derived from changing the Poisson random variable rates in the TL paths. Section 2.1 recalls the SOC formulation for optimal IS parameters from [11] and extends it with a novel HJB formulation. We conclude this consideration in Section 4, combining the IS scheme with a dimension reduction approach to reduce the computational cost.

2 Importance Sampling via Stochastic Optimal Control Formulation

2.1 Dynamic Programming for Importance Sampling Parameters

This section revisits the connection between optimal IS measure determination within a class of probability measures, and the SOC formulated originally in [11]. We let 𝐗𝐗\mathbf{X}bold_X be an SRN as defined in Section 1.1 and let 𝐗^Δtsuperscript^𝐗Δ𝑡\widehat{\mathbf{X}}^{\Delta t}over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT denote its TL approximation as given by (1.2). Then, the goal is to derive a near-optimal IS measure to estimate 𝔼[g(𝐗(T))]𝔼delimited-[]𝑔𝐗𝑇\mathbb{E}\left[g(\mathbf{X}(T))\right]blackboard_E [ italic_g ( bold_X ( italic_T ) ) ]. We limit ourselves to the parameterized class of IS schemes used in [10, 11]:

𝐗¯n+1Δtsuperscriptsubscript¯𝐗𝑛1Δ𝑡\displaystyle\overline{\mathbf{X}}_{n+1}^{\Delta t}over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT =max(𝟎,𝐗¯nΔt+j=1JP¯n,j𝝂j),n=0,,N1,formulae-sequenceabsent𝟎superscriptsubscript¯𝐗𝑛Δ𝑡superscriptsubscript𝑗1𝐽subscript¯𝑃𝑛𝑗subscript𝝂𝑗𝑛0𝑁1\displaystyle=\max\left(\textbf{0},\overline{\mathbf{X}}_{n}^{\Delta t}+\sum_{% j=1}^{J}\overline{P}_{n,j}\boldsymbol{\nu}_{j}\right),~{}~{}~{}n=0,\dots,N-1,= roman_max ( 0 , over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT over¯ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) , italic_n = 0 , … , italic_N - 1 , (2.1)
𝐗¯0Δtsuperscriptsubscript¯𝐗0Δ𝑡\displaystyle\overline{\mathbf{X}}_{0}^{\Delta t}over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT =𝐱0,absentsubscript𝐱0\displaystyle=\mathbf{x}_{0},= bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ,

where the measure change is obtained by modifying the Poisson random variable rates of the TL paths:

P¯n,j=𝒫¯n,j(δn,jΔt(𝐗¯nΔt)Δt),n=0,,N1,j=1,,J.formulae-sequencesubscript¯𝑃𝑛𝑗subscript¯𝒫𝑛𝑗superscriptsubscript𝛿𝑛𝑗Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑛Δ𝑡formulae-sequence𝑛0𝑁1𝑗1𝐽\overline{P}_{n,j}=\overline{\mathcal{P}}_{n,j}\left(\delta_{n,j}^{\Delta t}(% \overline{\mathbf{X}}^{\Delta t}_{n})\Delta t\right),~{}~{}~{}n=0,\dots,N-1,j=% 1,\dots,J.over¯ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT = over¯ start_ARG caligraphic_P end_ARG start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT ( italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) roman_Δ italic_t ) , italic_n = 0 , … , italic_N - 1 , italic_j = 1 , … , italic_J . (2.2)

In (2.2), δn,jΔt(𝐱)𝒜𝐱,jsuperscriptsubscript𝛿𝑛𝑗Δ𝑡𝐱subscript𝒜𝐱𝑗\delta_{n,j}^{\Delta t}(\mathbf{x})\in\mathcal{A}_{\mathbf{x},j}italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( bold_x ) ∈ caligraphic_A start_POSTSUBSCRIPT bold_x , italic_j end_POSTSUBSCRIPT is the control parameter at time step n𝑛nitalic_n, under reaction j𝑗jitalic_j, and in state 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT. In addition, {𝒫¯n,j(rn,j)}{1jJ}subscriptsubscript¯𝒫𝑛𝑗subscript𝑟𝑛𝑗1𝑗𝐽\{\overline{\mathcal{P}}_{n,j}(r_{n,j})\}_{\{1\leq j\leq J\}}{ over¯ start_ARG caligraphic_P end_ARG start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT ( italic_r start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT ) } start_POSTSUBSCRIPT { 1 ≤ italic_j ≤ italic_J } end_POSTSUBSCRIPT are independent Poisson random variables, conditioned on 𝐗¯nΔtsubscriptsuperscript¯𝐗Δ𝑡𝑛\overline{\mathbf{X}}^{\Delta t}_{n}over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, with the respective rates rn,j:=δn,jΔt(𝐗¯nΔt)Δtassignsubscript𝑟𝑛𝑗superscriptsubscript𝛿𝑛𝑗Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑛Δ𝑡r_{n,j}:=\delta_{n,j}^{\Delta t}(\overline{\mathbf{X}}^{\Delta t}_{n})\Delta titalic_r start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT := italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) roman_Δ italic_t. The set of admissible controls is

𝒜𝐱,j={{0},if aj(𝐱)=0{y:y>0},otherwise.\displaystyle\mathcal{A}_{\mathbf{x},j}=\begin{cases}\{0\}&,\text{if }a_{j}(% \mathbf{x})=0\\ \{y\in\mathbb{R}:y>0\}&,\text{otherwise}.\end{cases}caligraphic_A start_POSTSUBSCRIPT bold_x , italic_j end_POSTSUBSCRIPT = { start_ROW start_CELL { 0 } end_CELL start_CELL , if italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) = 0 end_CELL end_ROW start_ROW start_CELL { italic_y ∈ blackboard_R : italic_y > 0 } end_CELL start_CELL , otherwise . end_CELL end_ROW (2.3)

In the following, we use the vector notation (𝜹nΔt(𝐱))j:=δn,jΔt(𝐱)assignsubscriptsuperscriptsubscript𝜹𝑛Δ𝑡𝐱𝑗superscriptsubscript𝛿𝑛𝑗Δ𝑡𝐱\left(\boldsymbol{\delta}_{n}^{\Delta t}(\mathbf{x})\right)_{j}:=\delta_{n,j}^% {\Delta t}(\mathbf{x})( bold_italic_δ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( bold_x ) ) start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT := italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( bold_x ) and (𝐏¯n)j:=P¯n,jassignsubscriptsubscript¯𝐏𝑛𝑗subscript¯𝑃𝑛𝑗\left(\overline{\mathbf{P}}_{n}\right)_{j}:=\overline{P}_{n,j}( over¯ start_ARG bold_P end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT := over¯ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT for j=1,,J𝑗1𝐽j=1,\dots,Jitalic_j = 1 , … , italic_J.

The corresponding likelihood ratio of the path {𝐗¯nΔt:n=0,,N}conditional-setsubscriptsuperscript¯𝐗Δ𝑡𝑛𝑛0𝑁\{\overline{\mathbf{X}}^{\Delta t}_{n}:n=0,\dots,N\}{ over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT : italic_n = 0 , … , italic_N } for the IS parameters 𝜹nΔt(𝐱)×j=1J𝒜𝐱,j\boldsymbol{\delta}_{n}^{\Delta t}(\mathbf{x})\in\times_{j=1}^{J}\mathcal{A}_{% \mathbf{x},j}bold_italic_δ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( bold_x ) ∈ × start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT caligraphic_A start_POSTSUBSCRIPT bold_x , italic_j end_POSTSUBSCRIPT is

L((𝐏¯0,,𝐏¯N1),(𝜹0Δt(𝐗¯0Δt),,𝜹N1Δt(𝐗¯N1Δt)))=n=0N1Ln(𝐏¯n,𝜹nΔt(𝐗¯nΔt)),𝐿subscript¯𝐏0subscript¯𝐏𝑁1superscriptsubscript𝜹0Δ𝑡subscriptsuperscript¯𝐗Δ𝑡0superscriptsubscript𝜹𝑁1Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑁1superscriptsubscriptproduct𝑛0𝑁1subscript𝐿𝑛subscript¯𝐏𝑛superscriptsubscript𝜹𝑛Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑛L\left(\left(\overline{\mathbf{P}}_{0},\dots,\overline{\mathbf{P}}_{N-1}\right% ),\left(\boldsymbol{\delta}_{0}^{\Delta t}(\overline{\mathbf{X}}^{\Delta t}_{0% }),\dots,\boldsymbol{\delta}_{N-1}^{\Delta t}(\overline{\mathbf{X}}^{\Delta t}% _{N-1})\right)\right)=\prod_{n=0}^{N-1}L_{n}(\overline{\mathbf{P}}_{n},% \boldsymbol{\delta}_{n}^{\Delta t}(\overline{\mathbf{X}}^{\Delta t}_{n})),italic_L ( ( over¯ start_ARG bold_P end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , … , over¯ start_ARG bold_P end_ARG start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT ) , ( bold_italic_δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , … , bold_italic_δ start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT ) ) ) = ∏ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT italic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( over¯ start_ARG bold_P end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_italic_δ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ) , (2.4)

where the likelihood ratio update at time step n𝑛nitalic_n is

Ln(𝐏¯n,𝜹nΔt(𝐗¯nΔt))subscript𝐿𝑛subscript¯𝐏𝑛superscriptsubscript𝜹𝑛Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑛\displaystyle L_{n}(\overline{\mathbf{P}}_{n},\boldsymbol{\delta}_{n}^{\Delta t% }(\overline{\mathbf{X}}^{\Delta t}_{n}))italic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( over¯ start_ARG bold_P end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_italic_δ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ) =j=1Jexp((aj(𝐗¯nΔt)δn,jΔt(𝐗¯nΔt))Δt)(aj(𝐗¯nΔt)δn,jΔt(𝐗¯nΔt))P¯n,jabsentsuperscriptsubscriptproduct𝑗1𝐽subscript𝑎𝑗superscriptsubscript¯𝐗𝑛Δ𝑡superscriptsubscript𝛿𝑛𝑗Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑛Δ𝑡superscriptsubscript𝑎𝑗superscriptsubscript¯𝐗𝑛Δ𝑡superscriptsubscript𝛿𝑛𝑗Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑛subscript¯𝑃𝑛𝑗\displaystyle=\prod_{j=1}^{J}\exp\left(-(a_{j}(\overline{\mathbf{X}}_{n}^{% \Delta t})-\delta_{n,j}^{\Delta t}(\overline{\mathbf{X}}^{\Delta t}_{n}))% \Delta t\right)\left(\frac{a_{j}(\overline{\mathbf{X}}_{n}^{\Delta t})}{\delta% _{n,j}^{\Delta t}(\overline{\mathbf{X}}^{\Delta t}_{n})}\right)^{\overline{P}_% {n,j}}= ∏ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT roman_exp ( - ( italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ) - italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ) roman_Δ italic_t ) ( divide start_ARG italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) end_ARG ) start_POSTSUPERSCRIPT over¯ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT
=exp((j=1Jaj(𝐗¯nΔt)δn,jΔt(𝐗¯nΔt))Δt)j=1J(aj(𝐗¯nΔt)δn,jΔt(𝐗¯nΔt))P¯n,j.absentsuperscriptsubscript𝑗1𝐽subscript𝑎𝑗superscriptsubscript¯𝐗𝑛Δ𝑡superscriptsubscript𝛿𝑛𝑗Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑛Δ𝑡superscriptsubscriptproduct𝑗1𝐽superscriptsubscript𝑎𝑗superscriptsubscript¯𝐗𝑛Δ𝑡superscriptsubscript𝛿𝑛𝑗Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑛subscript¯𝑃𝑛𝑗\displaystyle=\exp\left(-\left(\sum_{j=1}^{J}a_{j}(\overline{\mathbf{X}}_{n}^{% \Delta t})-\delta_{n,j}^{\Delta t}(\overline{\mathbf{X}}^{\Delta t}_{n})\right% )\Delta t\right)\cdot\prod_{j=1}^{J}\left(\frac{a_{j}(\overline{\mathbf{X}}_{n% }^{\Delta t})}{\delta_{n,j}^{\Delta t}(\overline{\mathbf{X}}^{\Delta t}_{n})}% \right)^{\overline{P}_{n,j}}.= roman_exp ( - ( ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ) - italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ) roman_Δ italic_t ) ⋅ ∏ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT ( divide start_ARG italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) end_ARG ) start_POSTSUPERSCRIPT over¯ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT . (2.5)

To simplify the notation, we use the convention that aj(𝐗¯nΔt)δn,jΔt(𝐗¯nΔt)=1subscript𝑎𝑗superscriptsubscript¯𝐗𝑛Δ𝑡superscriptsubscript𝛿𝑛𝑗Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑛1\frac{a_{j}(\overline{\mathbf{X}}_{n}^{\Delta t})}{\delta_{n,j}^{\Delta t}(% \overline{\mathbf{X}}^{\Delta t}_{n})}=1divide start_ARG italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) end_ARG = 1, whenever both aj(𝐗¯nΔt)=0subscript𝑎𝑗superscriptsubscript¯𝐗𝑛Δ𝑡0a_{j}(\overline{\mathbf{X}}_{n}^{\Delta t})=0italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ) = 0 and δn,jΔt(𝐗¯nΔt)=0superscriptsubscript𝛿𝑛𝑗Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑛0\delta_{n,j}^{\Delta t}(\overline{\mathbf{X}}^{\Delta t}_{n})=0italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = 0 in (2.1). From (2.3), this results in a factor of 1111 in the likelihood ratio for reactions where aj(𝐗¯nΔt)=0subscript𝑎𝑗superscriptsubscript¯𝐗𝑛Δ𝑡0a_{j}(\overline{\mathbf{X}}_{n}^{\Delta t})=0italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ) = 0.

Using the introduced change of measure (2.2), the quantity of interest can be expressed with respect to the new measure:

𝔼[g(𝐗^NΔt)]=𝔼[L((𝐏¯0,,𝐏¯N1),(𝜹0Δt(𝐗¯0Δt),,𝜹N1Δt(𝐗¯N1Δt)))g(𝐗¯NΔt)],𝔼delimited-[]𝑔subscriptsuperscript^𝐗Δ𝑡𝑁𝔼delimited-[]𝐿subscript¯𝐏0subscript¯𝐏𝑁1superscriptsubscript𝜹0Δ𝑡subscriptsuperscript¯𝐗Δ𝑡0superscriptsubscript𝜹𝑁1Δ𝑡subscriptsuperscript¯𝐗Δ𝑡𝑁1𝑔subscriptsuperscript¯𝐗Δ𝑡𝑁\mathbb{E}[g(\widehat{\mathbf{X}}^{\Delta t}_{N})]=\mathbb{E}\left[L\left(% \left(\overline{\mathbf{P}}_{0},\dots,\overline{\mathbf{P}}_{N-1}\right),\left% (\boldsymbol{\delta}_{0}^{\Delta t}(\overline{\mathbf{X}}^{\Delta t}_{0}),% \dots,\boldsymbol{\delta}_{N-1}^{\Delta t}(\overline{\mathbf{X}}^{\Delta t}_{N% -1})\right)\right)\cdot g(\overline{\mathbf{X}}^{\Delta t}_{N})\right],blackboard_E [ italic_g ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) ] = blackboard_E [ italic_L ( ( over¯ start_ARG bold_P end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , … , over¯ start_ARG bold_P end_ARG start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT ) , ( bold_italic_δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , … , bold_italic_δ start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT ) ) ) ⋅ italic_g ( over¯ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) ] , (2.6)

with the expectation on the right-hand side of (2.6) taken with respect to the dynamics in (2.1).

Next, we recall the connection between the optimal second moment minimizing IS parameters {𝜹nΔt(𝐱)}n=0,,N1;𝐱dsubscriptsuperscriptsubscript𝜹𝑛Δ𝑡𝐱formulae-sequence𝑛0𝑁1𝐱superscript𝑑\{\boldsymbol{\delta}_{n}^{\Delta t}(\mathbf{x})\}_{n=0,\dots,N-1;\mathbf{x}% \in\mathbb{N}^{d}}{ bold_italic_δ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( bold_x ) } start_POSTSUBSCRIPT italic_n = 0 , … , italic_N - 1 ; bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT end_POSTSUBSCRIPT and the corresponding discrete-time dynamic programming relation from [11]. We revisit the definition of the discrete-time value function uΔt(,)subscript𝑢Δ𝑡u_{\Delta t}(\cdot,\cdot)italic_u start_POSTSUBSCRIPT roman_Δ italic_t end_POSTSUBSCRIPT ( ⋅ , ⋅ ) in Definition 2.1, allowing the formulation of the dynamic programming equations in Theorem 2.2. The proof and further details for Theorem 2.2 are provided in [11].

Definition 2.1 (Value function).

For a given Δt>0Δ𝑡0\Delta t>0roman_Δ italic_t > 0, the discrete-time value function uΔt(,)subscript𝑢Δ𝑡u_{\Delta t}(\cdot,\cdot)italic_u start_POSTSUBSCRIPT roman_Δ italic_t end_POSTSUBSCRIPT ( ⋅ , ⋅ ) is defined as the optimal (infimum) second moment for the proposed IS estimator. For time step 0nN0𝑛𝑁0\leq n\leq N0 ≤ italic_n ≤ italic_N and state 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT,

uΔt(n,𝐱)subscript𝑢Δ𝑡𝑛𝐱\displaystyle u_{\Delta t}(n,\mathbf{x})italic_u start_POSTSUBSCRIPT roman_Δ italic_t end_POSTSUBSCRIPT ( italic_n , bold_x ) =inf{𝜹kΔt}k=n,,N1𝒜Nn𝔼[g2(𝐗¯NΔt)k=nN1Lk2(𝐏¯k,𝜹kΔt(𝐗¯kΔt))|𝐗¯nΔt=𝐱],\displaystyle=\inf_{\{\boldsymbol{\delta}^{\Delta t}_{k}\}_{k=n,\dots,N-1}\in% \mathcal{A}^{N-n}}\mathbb{E}\left[g^{2}\left(\overline{\mathbf{X}}_{N}^{\Delta t% }\right)\prod_{k=n}^{N-1}L_{k}^{2}\left(\overline{\mathbf{P}}_{k},\boldsymbol{% \delta}_{k}^{\Delta t}(\overline{\mathbf{X}}_{k}^{\Delta t})\right)\middle|% \overline{\mathbf{X}}_{n}^{\Delta t}=\mathbf{x}\right],= roman_inf start_POSTSUBSCRIPT { bold_italic_δ start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_k = italic_n , … , italic_N - 1 end_POSTSUBSCRIPT ∈ caligraphic_A start_POSTSUPERSCRIPT italic_N - italic_n end_POSTSUPERSCRIPT end_POSTSUBSCRIPT blackboard_E [ italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ) ∏ start_POSTSUBSCRIPT italic_k = italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT italic_L start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( over¯ start_ARG bold_P end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , bold_italic_δ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ) ) | over¯ start_ARG bold_X end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT = bold_x ] , (2.7)

where 𝒜=×𝐱d×j=1J𝒜𝐱,jd×J\mathcal{A}=\bigtimes_{\mathbf{x}\in\mathbb{N}^{d}}\bigtimes_{j=1}^{J}\mathcal% {A}_{\mathbf{x},j}\in\mathbb{R}^{\mathbb{N}^{d}\times J}caligraphic_A = × start_POSTSUBSCRIPT bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT end_POSTSUBSCRIPT × start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT caligraphic_A start_POSTSUBSCRIPT bold_x , italic_j end_POSTSUBSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT × italic_J end_POSTSUPERSCRIPT is the admissible set for the IS parameters, and uΔt(N,𝐱)=g2(𝐱)subscript𝑢Δ𝑡𝑁𝐱superscript𝑔2𝐱u_{\Delta t}(N,\mathbf{x})=g^{2}(\mathbf{x})italic_u start_POSTSUBSCRIPT roman_Δ italic_t end_POSTSUBSCRIPT ( italic_N , bold_x ) = italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_x ), for any 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT.

Theorem 2.2 (Dynamic programming for IS parameters [11]).

For 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT and the given step size Δt>0normal-Δ𝑡0\Delta t>0roman_Δ italic_t > 0, the discrete-time value function uΔt(n,𝐱)subscript𝑢normal-Δ𝑡𝑛𝐱u_{\Delta t}(n,\mathbf{x})italic_u start_POSTSUBSCRIPT roman_Δ italic_t end_POSTSUBSCRIPT ( italic_n , bold_x ) fulfills the dynamic programming relation:

uΔt(N,𝐱)subscript𝑢Δ𝑡𝑁𝐱\displaystyle u_{\Delta t}(N,\mathbf{x})italic_u start_POSTSUBSCRIPT roman_Δ italic_t end_POSTSUBSCRIPT ( italic_N , bold_x ) =g2(𝐱)absentsuperscript𝑔2𝐱\displaystyle=g^{2}(\mathbf{x})= italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_x )
and for nand for 𝑛\displaystyle\text{and for }nand for italic_n =N1,,0,𝑎𝑛𝑑𝒜𝐱:=×j=1J𝒜𝐱,j,\displaystyle=N-1,\dots,0,\>\text{and}\>\mathcal{A}_{\mathbf{x}}:=\bigtimes_{j% =1}^{J}\mathcal{A}_{\mathbf{x},j},= italic_N - 1 , … , 0 , and caligraphic_A start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT := × start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT caligraphic_A start_POSTSUBSCRIPT bold_x , italic_j end_POSTSUBSCRIPT ,
uΔt(n,𝐱)subscript𝑢Δ𝑡𝑛𝐱\displaystyle u_{\Delta t}(n,\mathbf{x})italic_u start_POSTSUBSCRIPT roman_Δ italic_t end_POSTSUBSCRIPT ( italic_n , bold_x ) =inf𝜹nΔt(𝐱)𝒜𝐱exp((2j=1Jaj(𝐱)+j=1Jδn,jΔt(𝐱))Δt)absentsubscriptinfimumsuperscriptsubscript𝜹𝑛Δ𝑡𝐱subscript𝒜𝐱2superscriptsubscript𝑗1𝐽subscript𝑎𝑗𝐱superscriptsubscript𝑗1𝐽superscriptsubscript𝛿𝑛𝑗Δ𝑡𝐱Δ𝑡\displaystyle=\inf_{\boldsymbol{\delta}_{n}^{\Delta t}(\mathbf{x})\in\mathcal{% A}_{\mathbf{x}}}\exp\left(\left(-2\sum_{j=1}^{J}a_{j}(\mathbf{x})+\sum_{j=1}^{% J}\delta_{n,j}^{\Delta t}(\mathbf{x})\right)\Delta t\right)= roman_inf start_POSTSUBSCRIPT bold_italic_δ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( bold_x ) ∈ caligraphic_A start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_exp ( ( - 2 ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( bold_x ) ) roman_Δ italic_t ) (2.8)
×𝐩J(j=1J(Δtδn,jΔt(𝐱))pjpj!(aj(𝐱)δn,jΔt(𝐱))2pj)uΔt(n+1,max(𝟎,𝐱+𝝂𝐩)),\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\times\sum_{\mathbf{p}\in% \mathbb{N}^{J}}\left(\prod_{j=1}^{J}\frac{(\Delta t\cdot\delta_{n,j}^{\Delta t% }(\mathbf{x}))^{p_{j}}}{p_{j}!}(\frac{a_{j}(\mathbf{x})}{\delta_{n,j}^{\Delta t% }(\mathbf{x})})^{2p_{j}}\right)\cdot u_{\Delta t}(n+1,\max(\mathbf{0},\mathbf{% x}+\boldsymbol{\nu}\mathbf{p})),× ∑ start_POSTSUBSCRIPT bold_p ∈ blackboard_N start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( ∏ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT divide start_ARG ( roman_Δ italic_t ⋅ italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( bold_x ) ) start_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ! end_ARG ( divide start_ARG italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) end_ARG start_ARG italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( bold_x ) end_ARG ) start_POSTSUPERSCRIPT 2 italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) ⋅ italic_u start_POSTSUBSCRIPT roman_Δ italic_t end_POSTSUBSCRIPT ( italic_n + 1 , roman_max ( bold_0 , bold_x + bold_italic_ν bold_p ) ) ,

where 𝛎=(𝛎1,,𝛎J)d×J𝛎subscript𝛎1normal-…subscript𝛎𝐽superscript𝑑𝐽\boldsymbol{\nu}=\left(\boldsymbol{\nu}_{1},\dots,\boldsymbol{\nu}_{J}\right)% \in\mathbb{Z}^{d\times J}bold_italic_ν = ( bold_italic_ν start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , bold_italic_ν start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT ) ∈ blackboard_Z start_POSTSUPERSCRIPT italic_d × italic_J end_POSTSUPERSCRIPT.

Analytically solving the minimization problem (2.2) is challenging due to the infinite sum. In [11], the problem is solved by approximating the value function (2.7) using a truncated Taylor expansion of the dynamic programming (2.2). To overcome the curse of dimensionaliy, a learning-based approach for the value function was proposed. Instead, in this work, we utilize a continuous-time SOC formulation, leading to a set of coupled d𝑑ditalic_d-dimensional ODEs, the HJB equations (refer to Section 2.2). We deal with the curse of dimensionality issue by using a dimension reduction technique, namely the MP, as explained in Section 3.

2.2 Derivation of Hamilton–Jacobi–Bellman (HJB) Equations

In Corollary 2.3, the discrete-time dynamic programming relation in Theorem 2.2 is replaced by its analogous continuous-time relation, resulting in a set of ODEs known as the HJB equations. The continuous-time value function u~(,𝐱):[0,T]:~𝑢𝐱0𝑇\tilde{u}(\cdot,\mathbf{x}):[0,T]\rightarrow\mathbb{R}over~ start_ARG italic_u end_ARG ( ⋅ , bold_x ) : [ 0 , italic_T ] → blackboard_R, 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, is the limit of the discrete value function uΔt(,𝐱)subscript𝑢Δ𝑡𝐱u_{\Delta t}(\cdot,\mathbf{x})italic_u start_POSTSUBSCRIPT roman_Δ italic_t end_POSTSUBSCRIPT ( ⋅ , bold_x ) as the step size ΔtΔ𝑡\Delta troman_Δ italic_t approaches zero. In addition, the IS controls 𝜹(,𝐱):[0,T]𝒜𝐱:𝜹𝐱0𝑇subscript𝒜𝐱\boldsymbol{\delta}(\cdot,\mathbf{x}):[0,T]\rightarrow\mathcal{A}_{\mathbf{x}}bold_italic_δ ( ⋅ , bold_x ) : [ 0 , italic_T ] → caligraphic_A start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT become time-continuous curves for 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT.

Corollary 2.3 (HJB equations for IS parameters).

For all 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, the continuous-time value function u~(t,𝐱)normal-~𝑢𝑡𝐱\tilde{u}(t,\mathbf{x})over~ start_ARG italic_u end_ARG ( italic_t , bold_x ) fulfills (2.3) for t[0,T]𝑡0𝑇t\in[0,T]italic_t ∈ [ 0 , italic_T ]:

u~(T,𝐱)~𝑢𝑇𝐱\displaystyle\tilde{u}(T,\mathbf{x})over~ start_ARG italic_u end_ARG ( italic_T , bold_x ) =g2(𝐱)absentsuperscript𝑔2𝐱\displaystyle=g^{2}(\mathbf{x})= italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_x )
du~dt(t,𝐱)𝑑~𝑢𝑑𝑡𝑡𝐱\displaystyle-\frac{d\tilde{u}}{dt}(t,\mathbf{x})- divide start_ARG italic_d over~ start_ARG italic_u end_ARG end_ARG start_ARG italic_d italic_t end_ARG ( italic_t , bold_x ) =inf𝜹(t,𝐱)𝒜𝐱(2j=1Jaj(𝐱)+j=1Jδj(t,𝐱))u~(t,𝐱)+j=1Jaj(𝐱)2δj(t,𝐱)u~(t,max(0,𝐱+νj)),absentsubscriptinfimum𝜹𝑡𝐱subscript𝒜𝐱2superscriptsubscript𝑗1𝐽subscript𝑎𝑗𝐱superscriptsubscript𝑗1𝐽subscript𝛿𝑗𝑡𝐱~𝑢𝑡𝐱superscriptsubscript𝑗1𝐽subscript𝑎𝑗superscript𝐱2subscript𝛿𝑗𝑡𝐱~𝑢𝑡0𝐱subscript𝜈𝑗\displaystyle=\inf_{\boldsymbol{\delta}(t,\mathbf{x})\in\mathcal{A}_{\mathbf{x% }}}\left(-2\sum_{j=1}^{J}a_{j}(\mathbf{x})+\sum_{j=1}^{J}\delta_{j}(t,\mathbf{% x})\right)\tilde{u}(t,\mathbf{x})+\sum_{j=1}^{J}\frac{a_{j}(\mathbf{x})^{2}}{% \delta_{j}(t,\mathbf{x})}\tilde{u}\left(t,\max\left(0,\mathbf{x}+\nu_{j}\right% )\right),= roman_inf start_POSTSUBSCRIPT bold_italic_δ ( italic_t , bold_x ) ∈ caligraphic_A start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( - 2 ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) ) over~ start_ARG italic_u end_ARG ( italic_t , bold_x ) + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT divide start_ARG italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) end_ARG over~ start_ARG italic_u end_ARG ( italic_t , roman_max ( 0 , bold_x + italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) , (2.9)

where δj(t,𝐱):=(𝛅(t,𝐱))jassignsubscript𝛿𝑗𝑡𝐱subscript𝛅𝑡𝐱𝑗\delta_{j}(t,\mathbf{x}):=\left(\boldsymbol{\delta}(t,\mathbf{x})\right)_{j}italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) := ( bold_italic_δ ( italic_t , bold_x ) ) start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT.

Proof.

The proof of the corollary is presented in Appendix A. ∎

If u~(t,𝐱)>0~𝑢𝑡𝐱0\tilde{u}(t,\mathbf{x})>0over~ start_ARG italic_u end_ARG ( italic_t , bold_x ) > 0 for all 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT and t[0,T]𝑡0𝑇t\in[0,T]italic_t ∈ [ 0 , italic_T ], we can solve the minimization problem in (2.3) in closed form, such that the optimal controls are given by

δ~j(t,𝐱)subscript~𝛿𝑗𝑡𝐱\displaystyle\tilde{\delta}_{j}(t,\mathbf{x})over~ start_ARG italic_δ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) =aj(𝐱)u~(t,max(0,𝐱+νj))u~(t,𝐱)absentsubscript𝑎𝑗𝐱~𝑢𝑡0𝐱subscript𝜈𝑗~𝑢𝑡𝐱\displaystyle=a_{j}(\mathbf{x})\sqrt{\frac{\tilde{u}\left(t,\max\left(0,% \mathbf{x}+\nu_{j}\right)\right)}{\tilde{u}\left(t,\mathbf{x}\right)}}= italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) square-root start_ARG divide start_ARG over~ start_ARG italic_u end_ARG ( italic_t , roman_max ( 0 , bold_x + italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) end_ARG start_ARG over~ start_ARG italic_u end_ARG ( italic_t , bold_x ) end_ARG end_ARG (2.10)

and (2.3) simplifies to

du~dt(t,𝐱)𝑑~𝑢𝑑𝑡𝑡𝐱\displaystyle\frac{d\tilde{u}}{dt}(t,\mathbf{x})divide start_ARG italic_d over~ start_ARG italic_u end_ARG end_ARG start_ARG italic_d italic_t end_ARG ( italic_t , bold_x ) =2j=1Jaj(𝐱)(u~(t,𝐱)u~(t,max(0,𝐱+νj))u~(t,𝐱)).absent2superscriptsubscript𝑗1𝐽subscript𝑎𝑗𝐱~𝑢𝑡𝐱~𝑢𝑡0𝐱subscript𝜈𝑗~𝑢𝑡𝐱\displaystyle=-2\sum_{j=1}^{J}a_{j}(\mathbf{x})\left(\sqrt{\tilde{u}(t,\mathbf% {x})\tilde{u}(t,\max(0,\mathbf{x}+\nu_{j}))}-\tilde{u}(t,\mathbf{x})\right).= - 2 ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) ( square-root start_ARG over~ start_ARG italic_u end_ARG ( italic_t , bold_x ) over~ start_ARG italic_u end_ARG ( italic_t , roman_max ( 0 , bold_x + italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) end_ARG - over~ start_ARG italic_u end_ARG ( italic_t , bold_x ) ) . (2.11)

To estimate rare event probabilities with an observable g(𝐱)=𝟙{xi>γ}𝑔𝐱subscript1subscript𝑥𝑖𝛾g(\mathbf{x})=\mathbbm{1}_{\{x_{i}>\gamma\}}italic_g ( bold_x ) = blackboard_1 start_POSTSUBSCRIPT { italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT > italic_γ } end_POSTSUBSCRIPT, we encounter u~(t,𝐱)=0~𝑢𝑡𝐱0\tilde{u}(t,\mathbf{x})=0over~ start_ARG italic_u end_ARG ( italic_t , bold_x ) = 0 for some 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT; therefore, we modify (2.3) by approximating the final condition g(𝐱)𝑔𝐱g(\mathbf{x})italic_g ( bold_x ) using a sigmoid:

g~(𝐱)=11+exp(bβxi)>0~𝑔𝐱11𝑏𝛽subscript𝑥𝑖0\displaystyle\tilde{g}(\mathbf{x})=\frac{1}{1+\exp(-b-\beta x_{i})}>0over~ start_ARG italic_g end_ARG ( bold_x ) = divide start_ARG 1 end_ARG start_ARG 1 + roman_exp ( - italic_b - italic_β italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_ARG > 0 (2.12)

with appropriately chosen parameters b𝑏b\in\mathbb{R}italic_b ∈ blackboard_R and β𝛽\beta\in\mathbb{R}italic_β ∈ blackboard_R. By incorporating the modified final condition, we obtain an approximate value function by solving (2.11) using an ODE solver (e.g. ode23s from MATLAB). When using the numerical solver, we truncate the infinite state space d¯superscript¯𝑑\mathbb{N}^{\overline{d}}blackboard_N start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT using sufficiently large upper bounds. The approximated near-optimal IS controls are then expressed by (2.10). By the truncation of the infinite state space and the approximation of the final condition g𝑔gitalic_g by g~~𝑔\tilde{g}over~ start_ARG italic_g end_ARG, we introduce a bias to the value function. This can impact the amount of variance reduction in the IS-MC forward run, however, the IS-MC estimator is bias-free.

The cost for the ODE solver scales exponentially with the dimension d𝑑ditalic_d of the SRNs, making this approach infeasible for high-dimensional SRNs. Section 3 presents a dimension reduction approach for SRNs employed in Section 4 to derive suboptimal IS controls for a lower-dimensional SRN. We later demonstrate how these controls are mapped to the full-dimensional SRN system.

Remark 2.4 (Continuous-time IS controls).

In Corollary 2.3 and Theorem 2.2, we present two alternative methods to express the value function (2.7) and the IS controls. Utilizing the HJB framework, we can derive continuous controls across time. This allows any time stepping ΔtΔ𝑡\Delta troman_Δ italic_t in the IS-TL forward run and eliminates the need for ad-hoc interpolations.

3 Markovian Projection for Stochastic Reaction Networks

3.1 Formulation

To address the curse of dimensionality problem when deriving near-optimal IS controls, we project the SRN to a lower-dimensional network while preserving the marginal distribution of the original high-dimensional SRN system. We adapt the MP idea originally introduced in [29] for the setting of diffusion type stochastic differential equations to the SRNs framework. For an d𝑑ditalic_d-dimensional SRN state vector, 𝐗(t)𝐗𝑡\mathbf{X}(t)bold_X ( italic_t ), we introduce a projection to a d¯¯𝑑\overline{d}over¯ start_ARG italic_d end_ARG-dimensional space such that 1d¯d1¯𝑑much-less-than𝑑1\leq\overline{d}\ll d1 ≤ over¯ start_ARG italic_d end_ARG ≪ italic_d:

P:dd¯:𝐱𝐏𝐱,:𝑃superscript𝑑superscript¯𝑑:maps-to𝐱𝐏𝐱\displaystyle P:\mathbb{R}^{d}\rightarrow\mathbb{R}^{\overline{d}}:\>\mathbf{x% }\mapsto\mathbf{P}\mathbf{x},italic_P : blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT → blackboard_R start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT : bold_x ↦ bold_Px ,

where 𝐏d¯×d𝐏superscript¯𝑑𝑑\mathbf{P}\in\mathbb{R}^{\overline{d}\times d}bold_P ∈ blackboard_R start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG × italic_d end_POSTSUPERSCRIPT is a given matrix. This section develops a general MP framework for arbitrary projections with d¯1¯𝑑1\overline{d}\geq 1over¯ start_ARG italic_d end_ARG ≥ 1. However, the choice of the projection depends on the quantity of interest. In particular, when considering rare event probabilities with an observable g(𝐱)=𝟙{xi>γ},γformulae-sequence𝑔𝐱subscript1subscript𝑥𝑖𝛾𝛾g(\mathbf{x})=\mathbbm{1}_{\{x_{i}>\gamma\}},\gamma\in\mathbb{R}italic_g ( bold_x ) = blackboard_1 start_POSTSUBSCRIPT { italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT > italic_γ } end_POSTSUBSCRIPT , italic_γ ∈ blackboard_R as we do in Section 4, the projection operator is of the form

P(𝐱)=(0,,0i1,1𝑖,0i+1,,0)𝐱.𝑃𝐱0𝑖10𝑖1𝑖100𝐱\displaystyle P(\mathbf{x})=\left(0,\dots,\underset{\color[rgb]{.5,.5,.5}% \definecolor[named]{pgfstrokecolor}{rgb}{.5,.5,.5}\pgfsys@color@gray@stroke{.5% }\pgfsys@color@gray@fill{.5}{i-1}}{0},\underset{\color[rgb]{.5,.5,.5}% \definecolor[named]{pgfstrokecolor}{rgb}{.5,.5,.5}\pgfsys@color@gray@stroke{.5% }\pgfsys@color@gray@fill{.5}{i}}{1},\>\underset{\color[rgb]{.5,.5,.5}% \definecolor[named]{pgfstrokecolor}{rgb}{.5,.5,.5}\pgfsys@color@gray@stroke{.5% }\pgfsys@color@gray@fill{.5}{i+1}}{0},\dots,0\right)\mathbf{x}.italic_P ( bold_x ) = ( 0 , … , start_UNDERACCENT italic_i - 1 end_UNDERACCENT start_ARG 0 end_ARG , underitalic_i start_ARG 1 end_ARG , start_UNDERACCENT italic_i + 1 end_UNDERACCENT start_ARG 0 end_ARG , … , 0 ) bold_x . (3.1)

In the derivation of the MP, we assume that

𝔼[aj(𝐗(t))P(𝐗(t))=𝐬,𝐗(0)=𝐱0]<𝔼delimited-[]formulae-sequenceconditionalsubscript𝑎𝑗𝐗𝑡𝑃𝐗𝑡𝐬𝐗0subscript𝐱0\mathbb{E}[a_{j}(\mathbf{X}(t))\mid P(\mathbf{X}(t))=\mathbf{s},\mathbf{X}(0)=% \mathbf{x}_{0}]<\inftyblackboard_E [ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_t ) ) ∣ italic_P ( bold_X ( italic_t ) ) = bold_s , bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] < ∞ (3.2)

for 𝐬d¯𝐬superscript¯𝑑\mathbf{s}\in\mathbb{N}^{\overline{d}}bold_s ∈ blackboard_N start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT, t[0,T]𝑡0𝑇t\in[0,T]italic_t ∈ [ 0 , italic_T ] and 1jJ1𝑗𝐽1\leq j\leq J1 ≤ italic_j ≤ italic_J.

Theorem 3.1 shows that a d¯¯𝑑\overline{d}over¯ start_ARG italic_d end_ARG dimensional SRN, 𝑺¯(t)¯𝑺𝑡\overline{\boldsymbol{S}}(t)over¯ start_ARG bold_italic_S end_ARG ( italic_t ) exists that follows the same conditional distribution as 𝑺(t):=P(𝐗(t))assign𝑺𝑡𝑃𝐗𝑡\boldsymbol{S}(t):=P(\mathbf{X}(t))bold_italic_S ( italic_t ) := italic_P ( bold_X ( italic_t ) ) conditioned on the initial state 𝐗(0)=𝐱0𝐗0subscript𝐱0\mathbf{X}(0)=\mathbf{x}_{0}bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for all t[0,T]𝑡0𝑇t\in[0,T]italic_t ∈ [ 0 , italic_T ].

Theorem 3.1 (MP for SRNs).

We let 𝐒¯(t)normal-¯𝐒𝑡\overline{\boldsymbol{S}}(t)over¯ start_ARG bold_italic_S end_ARG ( italic_t ) be a d¯normal-¯𝑑\overline{d}over¯ start_ARG italic_d end_ARG-dimensional stochastic process whose dynamics are given by

𝑺¯(t)=P(𝐱0)+j=1JY¯j(0ta¯j(τ,𝑺¯(τ))𝑑τ)P(𝝂j)=:𝝂¯j,¯𝑺𝑡𝑃subscript𝐱0superscriptsubscript𝑗1𝐽subscript¯𝑌𝑗superscriptsubscript0𝑡subscript¯𝑎𝑗𝜏¯𝑺𝜏differential-d𝜏subscript𝑃subscript𝝂𝑗:absentsubscript¯𝝂𝑗\displaystyle\overline{\boldsymbol{S}}(t)=P(\mathbf{x}_{0})+\sum_{j=1}^{J}% \overline{Y}_{j}\left(\int_{0}^{t}\overline{a}_{j}(\tau,\overline{\boldsymbol{% S}}(\tau))d\tau\right)\underbrace{P(\boldsymbol{\nu}_{j})}_{=:\overline{% \boldsymbol{\nu}}_{j}},over¯ start_ARG bold_italic_S end_ARG ( italic_t ) = italic_P ( bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT over¯ start_ARG italic_Y end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_τ , over¯ start_ARG bold_italic_S end_ARG ( italic_τ ) ) italic_d italic_τ ) under⏟ start_ARG italic_P ( bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_ARG start_POSTSUBSCRIPT = : over¯ start_ARG bold_italic_ν end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUBSCRIPT , (3.3)

for t[0,T]𝑡0𝑇t\in[0,T]italic_t ∈ [ 0 , italic_T ], where Y¯jsubscriptnormal-¯𝑌𝑗\overline{Y}_{j}over¯ start_ARG italic_Y end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT denotes independent unit-rate Poisson processes and a¯jsubscriptnormal-¯𝑎𝑗\overline{a}_{j}over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, j=1,,J𝑗1normal-…𝐽j=1,\dots,Jitalic_j = 1 , … , italic_J, are characterized by

a¯j(t,𝒔):=𝔼[aj(𝐗(t))|P(𝐗(t))=𝒔,𝐗(0)=𝐱0], for 1jJ,𝒔d¯.\displaystyle\overline{a}_{j}(t,\boldsymbol{s}):=\mathbb{E}\left[a_{j}(\mathbf% {X}(t))\middle|P\left(\mathbf{X}(t)\right)=\boldsymbol{s},\mathbf{X}(0)=% \mathbf{x}_{0}\right]\text{, for }1\leq j\leq J,\boldsymbol{s}\in\mathbb{N}^{% \overline{d}}.over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_italic_s ) := blackboard_E [ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_t ) ) | italic_P ( bold_X ( italic_t ) ) = bold_italic_s , bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] , for 1 ≤ italic_j ≤ italic_J , bold_italic_s ∈ blackboard_N start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT . (3.4)

Thus, 𝐒(t){𝐗(0)=𝐱0}=P(𝐗(t)){𝐗(0)=𝐱0}evaluated-at𝐒𝑡𝐗0subscript𝐱0evaluated-at𝑃𝐗𝑡𝐗0subscript𝐱0\boldsymbol{S}(t)\mid_{\{\mathbf{X}(0)=\mathbf{x}_{0}\}}=P(\mathbf{X}(t))\mid_% {\{\mathbf{X}(0)=\mathbf{x}_{0}\}}bold_italic_S ( italic_t ) ∣ start_POSTSUBSCRIPT { bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT = italic_P ( bold_X ( italic_t ) ) ∣ start_POSTSUBSCRIPT { bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT and 𝐒¯(t){𝐗(0)=𝐱0}evaluated-atnormal-¯𝐒𝑡𝐗0subscript𝐱0\overline{\boldsymbol{S}}(t)\mid_{\{\mathbf{X}(0)=\mathbf{x}_{0}\}}over¯ start_ARG bold_italic_S end_ARG ( italic_t ) ∣ start_POSTSUBSCRIPT { bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT have the same distribution for all t[0,T]𝑡0𝑇t\in[0,T]italic_t ∈ [ 0 , italic_T ].

Proof.

The proof for Theorem 3.1 is given in Appendix B. ∎

In Theorem 3.1, we require the assumption in (3.2) to hold in order to ensure that the MP propensity in (3.4) is well-defined. Assumption (3.2) does not hold for all SRNs. However, in Remark 3.2, we present sufficient conditions guaranteeing (3.2).

Remark 3.2 (Sufficient conditions for assumption (3.2)).

If the propensity functions {aj()}j=1Jsuperscriptsubscriptsubscript𝑎𝑗𝑗1𝐽\{a_{j}(\cdot)\}_{j=1}^{J}{ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( ⋅ ) } start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT are bounded, i.e., it exist bounds Kj+subscript𝐾𝑗superscriptK_{j}\in\mathbb{R}^{+}italic_K start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT such that for any 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT aj(𝐱)<Kjsubscript𝑎𝑗𝐱subscript𝐾𝑗a_{j}(\mathbf{x})<K_{j}italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) < italic_K start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT for 1jJ1𝑗𝐽1\leq j\leq J1 ≤ italic_j ≤ italic_J. Then, the expectation in (3.2) is bounded by the same bounds. An alternative condition is that the set

𝒦t,𝐬:={𝐲d:P(𝐲)=𝐬and(𝐗(t)=𝐲)>0}assignsubscript𝒦𝑡𝐬conditional-set𝐲superscript𝑑formulae-sequence𝑃𝐲𝐬and𝐗𝑡𝐲0\displaystyle\mathcal{K}_{t,\mathbf{s}}:=\left\{\mathbf{y}\in\mathbb{N}^{d}:P(% \mathbf{y})=\mathbf{s}\quad\text{and}\quad\mathbb{P}(\mathbf{X}(t)=\mathbf{y})% >0\right\}caligraphic_K start_POSTSUBSCRIPT italic_t , bold_s end_POSTSUBSCRIPT := { bold_y ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT : italic_P ( bold_y ) = bold_s and blackboard_P ( bold_X ( italic_t ) = bold_y ) > 0 }

is finite for all 𝐬d¯𝐬superscript¯𝑑\mathbf{s}\in\mathbb{N}^{\overline{d}}bold_s ∈ blackboard_N start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT and t[0,T]𝑡0𝑇t\in[0,T]italic_t ∈ [ 0 , italic_T ]. Consequently, the expectation in (3.2) corresponds to a finite sum and is finite. An in-depth analysis is required to establish sharp conditions under which (3.2) holds. Our numerical analysis show that the MP propensity is well defined in many examples.

The mimicking process {𝑺¯(t)}t[0,T]subscript¯𝑺𝑡𝑡0𝑇\{\overline{\boldsymbol{S}}(t)\}_{t\in[0,T]}{ over¯ start_ARG bold_italic_S end_ARG ( italic_t ) } start_POSTSUBSCRIPT italic_t ∈ [ 0 , italic_T ] end_POSTSUBSCRIPT is a SRN and consequently Markovian, whereas the projected full-dimensional process {𝑺(t)}t[0,T]subscript𝑺𝑡𝑡0𝑇\{\boldsymbol{S}(t)\}_{t\in[0,T]}{ bold_italic_S ( italic_t ) } start_POSTSUBSCRIPT italic_t ∈ [ 0 , italic_T ] end_POSTSUBSCRIPT is non-Markovian. Moreover, the propensities of the full-dimensional process {aj}j=1Jsuperscriptsubscriptsubscript𝑎𝑗𝑗1𝐽\{a_{j}\}_{j=1}^{J}{ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT are time-homogeneous functions of the state, whereas the resulting propensities, {a¯j}j=1Jsuperscriptsubscriptsubscript¯𝑎𝑗𝑗1𝐽\{\overline{a}_{j}\}_{j=1}^{J}{ over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT of the MP-SRN 𝑺¯¯𝑺\overline{\boldsymbol{S}}over¯ start_ARG bold_italic_S end_ARG are time-dependent (see (3.4)). Reactions with P(𝝂j)=0𝑃subscript𝝂𝑗0P(\boldsymbol{\nu}_{j})=0italic_P ( bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = 0 do not contribute to the MP propensity in (3.4). For reactions with P(𝝂j)0𝑃subscript𝝂𝑗0P(\boldsymbol{\nu}_{j})\neq 0italic_P ( bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ≠ 0, it may occur that their corresponding projected propensity is known analytically. We denote the index set of reactions requiring an estimation of (3.4) (e.g., via a L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT regression as described in Section 3.2) by 𝒥MPsubscript𝒥𝑀𝑃\mathcal{J}_{MP}caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT. This index set is described as follows:

𝒥MP:={1jJ:P(𝝂j)0andaj(𝐱)f(P(𝐱)) for all functions f:d¯(*)},\displaystyle\mathcal{J}_{MP}:=\left\{1\leq j\leq J:\quad P(\boldsymbol{\nu}_{% j})\neq 0\quad\text{and}\quad\underbrace{a_{j}(\mathbf{x})\neq f(P(\mathbf{x})% )\text{ for all functions }f:\mathbb{R}^{\overline{d}}\rightarrow\mathbb{R}}_{% (*)}\right\},caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT := { 1 ≤ italic_j ≤ italic_J : italic_P ( bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ≠ 0 and under⏟ start_ARG italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) ≠ italic_f ( italic_P ( bold_x ) ) for all functions italic_f : blackboard_R start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT → blackboard_R end_ARG start_POSTSUBSCRIPT ( * ) end_POSTSUBSCRIPT } , (3.5)

where condition (*) excludes reaction channels for which the MP propensity is only dependent on s𝑠sitalic_s and given in closed form by a¯j(t,s)=f(s)subscript¯𝑎𝑗𝑡𝑠𝑓𝑠\overline{a}_{j}(t,s)=f(s)over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , italic_s ) = italic_f ( italic_s ) for the function f𝑓fitalic_f.

3.2 Discrete L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT Regression for Approximating Projected Propensities

To approximate the Markovian propensity a¯jsubscript¯𝑎𝑗\overline{a}_{j}over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT for j𝒥MP𝑗subscript𝒥𝑀𝑃j\in\mathcal{J}_{MP}italic_j ∈ caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT, we reformulate (3.4) as a minimization problem and then use discrete L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT regression as described below.

We let V:={f:[0,T]×d¯:0T𝔼[f(t,P(𝐗(t)))2]𝑑t<}assign𝑉conditional-set𝑓:0𝑇superscript¯𝑑superscriptsubscript0𝑇𝔼delimited-[]𝑓superscript𝑡𝑃𝐗𝑡2differential-d𝑡V:=\left\{f:[0,T]\times\mathbb{R}^{\overline{d}}\rightarrow\mathbb{R}:\int_{0}% ^{T}\mathbb{E}[f(t,P(\mathbf{X}(t)))^{2}]dt<\infty\right\}italic_V := { italic_f : [ 0 , italic_T ] × blackboard_R start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT → blackboard_R : ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT blackboard_E [ italic_f ( italic_t , italic_P ( bold_X ( italic_t ) ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_d italic_t < ∞ }. Then, the projected propensities via the MP for j𝒥MP𝑗subscript𝒥𝑀𝑃j\in\mathcal{J}_{MP}italic_j ∈ caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT are approximated by

a¯j(,)subscript¯𝑎𝑗\displaystyle\overline{a}_{j}(\cdot,\cdot)over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( ⋅ , ⋅ ) =argminhV0T𝔼[(aj(𝐗(t))h(t,P(𝐗(t))))2]𝑑tabsentsubscriptargmin𝑉superscriptsubscript0𝑇𝔼delimited-[]superscriptsubscript𝑎𝑗𝐗𝑡𝑡𝑃𝐗𝑡2differential-d𝑡\displaystyle=\text{argmin}_{h\in V}\int_{0}^{T}\mathbb{E}\left[\left(a_{j}(% \mathbf{X}(t))-h(t,P(\mathbf{X}(t)))\right)^{2}\right]dt= argmin start_POSTSUBSCRIPT italic_h ∈ italic_V end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT blackboard_E [ ( italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_t ) ) - italic_h ( italic_t , italic_P ( bold_X ( italic_t ) ) ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_d italic_t
argminhV𝔼[1Nn=0N1(aj(𝐗^nΔt)h(tn,P(𝐗^nΔt)))2]absentsubscriptargmin𝑉𝔼delimited-[]1𝑁superscriptsubscript𝑛0𝑁1superscriptsubscript𝑎𝑗subscriptsuperscript^𝐗Δ𝑡𝑛subscript𝑡𝑛𝑃subscriptsuperscript^𝐗Δ𝑡𝑛2\displaystyle\approx\text{argmin}_{h\in V}\mathbb{E}\left[\frac{1}{{N}}\sum_{{% n=0}}^{N-1}\left(a_{j}(\widehat{\mathbf{X}}^{\Delta t}_{n})-h(t_{n},P(\widehat% {\mathbf{X}}^{\Delta t}_{n}))\right)^{2}\right]≈ argmin start_POSTSUBSCRIPT italic_h ∈ italic_V end_POSTSUBSCRIPT blackboard_E [ divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT ( italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) - italic_h ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_P ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ]
argminhV1Mm=1M1Nn=0N1(aj(𝐗^[m],nΔt)h(tn,P(𝐗^[m],nΔt)))2,absentsubscriptargmin𝑉1𝑀superscriptsubscript𝑚1𝑀1𝑁superscriptsubscript𝑛0𝑁1superscriptsubscript𝑎𝑗subscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑛subscript𝑡𝑛𝑃subscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑛2\displaystyle\approx\text{argmin}_{h\in V}\frac{1}{M}\sum_{m=1}^{M}\frac{1}{{N% }}\sum_{{n=0}}^{N-1}\left(a_{j}(\widehat{\mathbf{X}}^{\Delta t}_{[m],n})-h(t_{% n},P(\widehat{\mathbf{X}}^{\Delta t}_{[m],n}))\right)^{2},≈ argmin start_POSTSUBSCRIPT italic_h ∈ italic_V end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG italic_M end_ARG ∑ start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT ( italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] , italic_n end_POSTSUBSCRIPT ) - italic_h ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_P ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] , italic_n end_POSTSUBSCRIPT ) ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (3.6)

where {𝐗^[m]Δt}m=1Msuperscriptsubscriptsubscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑚1𝑀\left\{\widehat{\mathbf{X}}^{\Delta t}_{[m]}\right\}_{m=1}^{M}{ over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT are M𝑀Mitalic_M independent TL paths with a uniform time grid 0=t0<t1<<tN=T0subscript𝑡0subscript𝑡1subscript𝑡𝑁𝑇0=t_{0}<t_{1}<\dots<t_{N}=T0 = italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < ⋯ < italic_t start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = italic_T with step size ΔtΔ𝑡\Delta troman_Δ italic_t.

To solve (3.2), we use a discrete L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT regression approach. For the case d¯=1¯𝑑1\overline{d}=1over¯ start_ARG italic_d end_ARG = 1, we employ a set of basis functions in V𝑉Vitalic_V, {ϕp(,)}pΛsubscriptsubscriptitalic-ϕ𝑝𝑝Λ\{\phi_{p}(\cdot,\cdot)\}_{p\in\Lambda}{ italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( ⋅ , ⋅ ) } start_POSTSUBSCRIPT italic_p ∈ roman_Λ end_POSTSUBSCRIPT, where Λ2Λsuperscript2\Lambda\subset\mathbb{N}^{2}roman_Λ ⊂ blackboard_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is a finite index set. In Remark 3.3, we provide more details on the choice of the basis. Consequently, the projected propensities via MP are approximated by

a¯j(t,s)pΛcp(j)ϕp(t,s),j𝒥MPformulae-sequencesubscript¯𝑎𝑗𝑡𝑠subscript𝑝Λsuperscriptsubscript𝑐𝑝𝑗subscriptitalic-ϕ𝑝𝑡𝑠𝑗subscript𝒥𝑀𝑃\displaystyle\overline{a}_{j}(t,s)\approx\sum_{p\in\Lambda}c_{p}^{(j)}{\phi}_{% p}(t,s),j\in\mathcal{J}_{MP}over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , italic_s ) ≈ ∑ start_POSTSUBSCRIPT italic_p ∈ roman_Λ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t , italic_s ) , italic_j ∈ caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT (3.7)

where the coefficients cp(j)superscriptsubscript𝑐𝑝𝑗c_{p}^{(j)}italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT must be derived for j𝒥MP𝑗subscript𝒥𝑀𝑃j\in\mathcal{J}_{MP}italic_j ∈ caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT and pΛ𝑝Λp\in\Lambdaitalic_p ∈ roman_Λ.

Next, we derive the linear systems of equations, solved by {cp(j)}pΛsubscriptsuperscriptsubscript𝑐𝑝𝑗𝑝Λ\{c_{p}^{(j)}\}_{p\in\Lambda}{ italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT } start_POSTSUBSCRIPT italic_p ∈ roman_Λ end_POSTSUBSCRIPT from (3.7) for j𝒥MP𝑗subscript𝒥𝑀𝑃j\in\mathcal{J}_{MP}italic_j ∈ caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT. For a given one-dimensional indexing of {1,,M}×{0,,N1}1𝑀0𝑁1\{1,\dots,M\}\times\{0,\dots,N-1\}{ 1 , … , italic_M } × { 0 , … , italic_N - 1 }, the corresponding design matrix 𝑫MN×|Λ|𝑫superscript𝑀𝑁Λ\boldsymbol{D}\in\mathbb{R}^{MN\times|\Lambda|}bold_italic_D ∈ blackboard_R start_POSTSUPERSCRIPT italic_M italic_N × | roman_Λ | end_POSTSUPERSCRIPT is given by

Dk,p=ϕp(tn,P(𝐗^[m],nΔt)), for k=(m,n){1,,M}×{0,,N1},pΛ.formulae-sequenceformulae-sequencesubscript𝐷𝑘𝑝subscriptitalic-ϕ𝑝subscript𝑡𝑛𝑃subscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑛 for 𝑘𝑚𝑛1𝑀0𝑁1𝑝Λ\displaystyle D_{k,p}={\phi}_{p}(t_{n},P(\widehat{\mathbf{X}}^{\Delta t}_{[m],% n})),\text{ for }k=(m,n)\in\{1,\dots,M\}\times\{0,\dots,N-1\},\quad p\in\Lambda.italic_D start_POSTSUBSCRIPT italic_k , italic_p end_POSTSUBSCRIPT = italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_P ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] , italic_n end_POSTSUBSCRIPT ) ) , for italic_k = ( italic_m , italic_n ) ∈ { 1 , … , italic_M } × { 0 , … , italic_N - 1 } , italic_p ∈ roman_Λ .

Further, we set ψk(j)=aj(𝐗^[m],nΔt)superscriptsubscript𝜓𝑘𝑗subscript𝑎𝑗subscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑛\psi_{k}^{(j)}=a_{j}(\widehat{\mathbf{X}}^{\Delta t}_{[m],n})italic_ψ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT = italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] , italic_n end_POSTSUBSCRIPT ) (𝝍(j)MNsuperscript𝝍𝑗superscript𝑀𝑁\boldsymbol{\psi}^{(j)}\in\mathbb{R}^{MN}bold_italic_ψ start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_M italic_N end_POSTSUPERSCRIPT) for k{1,,M}×{0,,N1}𝑘1𝑀0𝑁1k\in\{1,\dots,M\}\times\{0,\dots,N-1\}italic_k ∈ { 1 , … , italic_M } × { 0 , … , italic_N - 1 }, and j𝒥MP𝑗subscript𝒥𝑀𝑃j\in\mathcal{J}_{MP}italic_j ∈ caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT.

Then, the minimization problem in (3.2) becomes

𝐜(j)superscript𝐜𝑗\displaystyle\textbf{c}^{(j)}c start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT =argmin{cp}pΛ1MNm=1Mn=0N1(aj(𝐗^[m],nΔt)pΛcpϕ(tn,P(𝐗^[m],nΔt)))2absentsubscriptargminsubscriptsubscript𝑐𝑝𝑝Λ1𝑀𝑁superscriptsubscript𝑚1𝑀superscriptsubscript𝑛0𝑁1superscriptsubscript𝑎𝑗subscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑛subscript𝑝Λsubscript𝑐𝑝italic-ϕsubscript𝑡𝑛𝑃subscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑛2\displaystyle=\text{argmin}_{\{c_{p}\}_{p\in\Lambda}}\frac{1}{MN}\sum_{m=1}^{M% }\sum_{n=0}^{N-1}\left(a_{j}(\widehat{\mathbf{X}}^{\Delta t}_{[m],n})-\sum_{p% \in\Lambda}c_{p}{\phi}\left(t_{n},P(\widehat{\mathbf{X}}^{\Delta t}_{[m],n})% \right)\right)^{2}= argmin start_POSTSUBSCRIPT { italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_p ∈ roman_Λ end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG italic_M italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT ( italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] , italic_n end_POSTSUBSCRIPT ) - ∑ start_POSTSUBSCRIPT italic_p ∈ roman_Λ end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_ϕ ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_P ( over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] , italic_n end_POSTSUBSCRIPT ) ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
=argmin𝐜#Λ(𝝍(j)𝑫𝐜)(𝝍(j)𝑫𝐜)absentsubscriptargmin𝐜superscript#Λsuperscriptsuperscript𝝍𝑗𝑫𝐜topsuperscript𝝍𝑗𝑫𝐜\displaystyle=\text{argmin}_{\mathbf{c}\in\mathbb{R}^{\#\Lambda}}\left(% \boldsymbol{\psi}^{(j)}-\boldsymbol{D}\textbf{c}\right)^{\top}\left(% \boldsymbol{\psi}^{(j)}-\boldsymbol{D}\textbf{c}\right)= argmin start_POSTSUBSCRIPT bold_c ∈ blackboard_R start_POSTSUPERSCRIPT # roman_Λ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( bold_italic_ψ start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT - bold_italic_D c ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT ( bold_italic_ψ start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT - bold_italic_D c )
=argmin𝐜#Λ𝝍(j)𝝍(j)2𝐜𝑫𝝍(j)+𝐜𝑫𝑫𝐜=:I(𝐜).absentsubscriptargmin𝐜superscript#Λsubscriptsuperscriptsuperscript𝝍𝑗topsuperscript𝝍𝑗2superscript𝐜topsuperscript𝑫topsuperscript𝝍𝑗superscript𝐜topsuperscript𝑫top𝑫𝐜:absent𝐼𝐜\displaystyle=\text{argmin}_{\mathbf{c}\in\mathbb{R}^{\#\Lambda}}\underbrace{{% \boldsymbol{\psi}^{(j)}}^{\top}\boldsymbol{\psi}^{(j)}-2\mathbf{c}^{\top}% \boldsymbol{D}^{\top}\boldsymbol{\psi}^{(j)}+\mathbf{c}^{\top}\boldsymbol{D}^{% \top}\boldsymbol{D}\mathbf{c}}_{=:I(\mathbf{c})}.= argmin start_POSTSUBSCRIPT bold_c ∈ blackboard_R start_POSTSUPERSCRIPT # roman_Λ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT under⏟ start_ARG bold_italic_ψ start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_italic_ψ start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT - 2 bold_c start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_italic_D start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_italic_ψ start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT + bold_c start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_italic_D start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_italic_D bold_c end_ARG start_POSTSUBSCRIPT = : italic_I ( bold_c ) end_POSTSUBSCRIPT .

We minimize I(𝐜)𝐼𝐜I(\mathbf{c})italic_I ( bold_c ) with respect to 𝐜𝐜\mathbf{c}bold_c by solving

I(𝐜)𝐜=2𝑫𝝍(j)+2𝑫𝑫𝐜=0𝐼𝐜𝐜2superscript𝑫topsuperscript𝝍𝑗2superscript𝑫top𝑫𝐜0\displaystyle\frac{\partial I(\mathbf{c})}{\partial\mathbf{c}}=-2{\boldsymbol{% D}}^{\top}\boldsymbol{\psi}^{(j)}+2{\boldsymbol{D}}^{\top}\boldsymbol{% \boldsymbol{D}}\mathbf{c}=0divide start_ARG ∂ italic_I ( bold_c ) end_ARG start_ARG ∂ bold_c end_ARG = - 2 bold_italic_D start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_italic_ψ start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT + 2 bold_italic_D start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_italic_D bold_c = 0

and obtain the normal equation for j𝒥MP𝑗subscript𝒥𝑀𝑃j\in\mathcal{J}_{MP}italic_j ∈ caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT:

(𝑫𝑫)𝐜(j)=𝑫𝝍(j).superscript𝑫top𝑫superscript𝐜𝑗superscript𝑫topsuperscript𝝍𝑗\displaystyle\left(\boldsymbol{D}^{\top}\boldsymbol{D}\right)\mathbf{c}^{(j)}=% \boldsymbol{D}^{\top}\boldsymbol{\psi}^{(j)}.( bold_italic_D start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_italic_D ) bold_c start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT = bold_italic_D start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_italic_ψ start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT . (3.8)
Remark 3.3 (Orthonormal basis approach via empirical inner product).

For the case d¯=1¯𝑑1\overline{d}=1over¯ start_ARG italic_d end_ARG = 1, the normal equation with a set of polynomials {ϕp}pΛsubscriptsubscriptitalic-ϕ𝑝𝑝Λ\{\phi_{p}\}_{p\in\Lambda}{ italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_p ∈ roman_Λ end_POSTSUBSCRIPT in 2superscript2\mathbb{R}^{2}blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT can be used to derive the MP propensity a¯jsubscript¯𝑎𝑗\overline{a}_{j}over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT for j𝒥MP𝑗subscript𝒥𝑀𝑃j\in\mathcal{J}_{MP}italic_j ∈ caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT. We use the standard basis {ϕ(i1,i2)}(i1,i2)Λsubscriptsubscriptitalic-ϕsubscript𝑖1subscript𝑖2subscript𝑖1subscript𝑖2Λ\left\{\phi_{(i_{1},i_{2})}\right\}_{(i_{1},i_{2})\in\Lambda}{ italic_ϕ start_POSTSUBSCRIPT ( italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT } start_POSTSUBSCRIPT ( italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ∈ roman_Λ end_POSTSUBSCRIPT for a two-dimensional index set ΛΛ\Lambdaroman_Λ, where

ϕ(i1,i2):2,(t,x)ti1xi2.:subscriptitalic-ϕsubscript𝑖1subscript𝑖2formulae-sequencesuperscript2maps-to𝑡𝑥superscript𝑡subscript𝑖1superscript𝑥subscript𝑖2\phi_{(i_{1},i_{2})}:\mathbb{R}^{2}\rightarrow\mathbb{R},(t,x)\mapsto t^{i_{1}% }x^{i_{2}}.italic_ϕ start_POSTSUBSCRIPT ( italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT : blackboard_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT → blackboard_R , ( italic_t , italic_x ) ↦ italic_t start_POSTSUPERSCRIPT italic_i start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT italic_i start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT .

For better stability [18], we use the Gram–Schmidt orthogonalization algorithm to determine an orthonormal set of functions for the empirical scalar product:

ϕi,ϕjM=1Nn=0N11Mm=1Mϕi(tn,P𝐗^[m],nΔt)ϕj(tn,P𝐗^[m],nΔt)subscriptsubscriptitalic-ϕ𝑖subscriptitalic-ϕ𝑗𝑀1𝑁superscriptsubscript𝑛0𝑁11𝑀superscriptsubscript𝑚1𝑀subscriptitalic-ϕ𝑖subscript𝑡𝑛𝑃subscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑛subscriptitalic-ϕ𝑗subscript𝑡𝑛𝑃subscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑛\displaystyle\left\langle\phi_{i},\phi_{j}\right\rangle_{{M}}=\frac{1}{N}\sum_% {n=0}^{N-1}\frac{1}{{M}}\sum_{m=1}^{{M}}\phi_{i}\left(t_{n},P\widehat{\mathbf{% X}}^{\Delta t}_{[m],n}\right)\phi_{j}\left(t_{n},P\widehat{\mathbf{X}}^{\Delta t% }_{[m],n}\right)⟨ italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⟩ start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_M end_ARG ∑ start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_M end_POSTSUPERSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_P over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] , italic_n end_POSTSUBSCRIPT ) italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_P over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] , italic_n end_POSTSUBSCRIPT ) (3.9)

to find an orthonormal set of functions. We base the empirical scalar product and the normal equation (3.8) on the same set of TL paths, {𝐗^[m]Δt}m=1,,Msubscriptsubscriptsuperscript^𝐗Δ𝑡delimited-[]𝑚𝑚1𝑀\{\widehat{\mathbf{X}}^{\Delta t}_{[m]}\}_{m=1,\dots,M}{ over^ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT [ italic_m ] end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_m = 1 , … , italic_M end_POSTSUBSCRIPT, such that the matrix condition number becomes cond(𝑫𝑫)=1condsuperscript𝑫top𝑫1\text{cond}(\boldsymbol{D}^{\top}\boldsymbol{D})=1cond ( bold_italic_D start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_italic_D ) = 1 and 𝑫𝑫=diag(TΔtM,,TΔtM)superscript𝑫top𝑫diag𝑇Δ𝑡𝑀𝑇Δ𝑡𝑀\boldsymbol{D}^{\top}\boldsymbol{D}=\text{diag}\left(\frac{T}{\Delta t}M,\dots% ,\frac{T}{\Delta t}M\right)bold_italic_D start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT bold_italic_D = diag ( divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG italic_M , … , divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG italic_M ) [18].

Remark 3.4 (Propensity functions).

The proof of Theorem (3.1) and the derivation of the normal equation (3.8) is general with respect to the structure of the propensity function. The Markovian projection can be applied to SRNs with arbitrary propensity functions, i.e., not restricted to the mass-action-kinetics principle introduced in (1.4) (e.g. Hill-type reaction rate law [49]). The choice of suitable basis functions {ϕp}pΛsubscriptsubscriptitalic-ϕ𝑝𝑝Λ\{\phi_{p}\}_{p\in\Lambda}{ italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_p ∈ roman_Λ end_POSTSUBSCRIPT for the L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT regression depends on the given example and also on the type of propensity.

3.3 Computational Cost of Markovian Projection

The computational work to derive an MP for an SRN with J𝐽Jitalic_J reactions based on a time stepping ΔtΔ𝑡\Delta troman_Δ italic_t based on M𝑀Mitalic_M TL paths and an orthonormal set of polynomials (see Remark 3.3) of size #Λ#Λ\#\Lambda# roman_Λ splits into three types of costs:

WMP(#Λ,Δt,M)MWTL(Δt)+WGS(#Λ,Δt,M)+WL2(#Λ,Δt,M),subscript𝑊𝑀𝑃#ΛΔ𝑡𝑀𝑀subscript𝑊𝑇𝐿Δ𝑡subscript𝑊𝐺𝑆#ΛΔ𝑡𝑀subscript𝑊superscript𝐿2#ΛΔ𝑡𝑀\displaystyle W_{MP}(\#\Lambda,\Delta t,M)\approx M\cdot W_{TL}(\Delta t)+W_{G% -S}(\#\Lambda,\Delta t,M)+W_{L^{2}}(\#\Lambda,\Delta t,M),italic_W start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( # roman_Λ , roman_Δ italic_t , italic_M ) ≈ italic_M ⋅ italic_W start_POSTSUBSCRIPT italic_T italic_L end_POSTSUBSCRIPT ( roman_Δ italic_t ) + italic_W start_POSTSUBSCRIPT italic_G - italic_S end_POSTSUBSCRIPT ( # roman_Λ , roman_Δ italic_t , italic_M ) + italic_W start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( # roman_Λ , roman_Δ italic_t , italic_M ) , (3.10)

where WTLsubscript𝑊𝑇𝐿W_{TL}italic_W start_POSTSUBSCRIPT italic_T italic_L end_POSTSUBSCRIPT, WGSsubscript𝑊𝐺𝑆W_{G-S}italic_W start_POSTSUBSCRIPT italic_G - italic_S end_POSTSUBSCRIPT, and WL2subscript𝑊superscript𝐿2W_{L^{2}}italic_W start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT denote the computational costs to simulate a TL path, derive an orthonormal basis (as described in Remark 3.3), and derive and solve the normal equation in (3.8), respectively. The dominant terms of these costs contribute as follows:

WTL(Δt)subscript𝑊𝑇𝐿Δ𝑡\displaystyle W_{TL}(\Delta t)italic_W start_POSTSUBSCRIPT italic_T italic_L end_POSTSUBSCRIPT ( roman_Δ italic_t ) TΔtJCPoi,absent𝑇Δ𝑡𝐽subscript𝐶𝑃𝑜𝑖\displaystyle\approx\frac{T}{\Delta t}\cdot J\cdot C_{Poi},≈ divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ italic_J ⋅ italic_C start_POSTSUBSCRIPT italic_P italic_o italic_i end_POSTSUBSCRIPT ,
WGS(#Λ,Δt,M)subscript𝑊𝐺𝑆#ΛΔ𝑡𝑀\displaystyle W_{G-S}(\#\Lambda,\Delta t,M)italic_W start_POSTSUBSCRIPT italic_G - italic_S end_POSTSUBSCRIPT ( # roman_Λ , roman_Δ italic_t , italic_M ) MTΔt(#Λ)3,absent𝑀𝑇Δ𝑡superscript#Λ3\displaystyle\approx M\cdot\frac{T}{\Delta t}\cdot\left(\#\Lambda\right)^{3},≈ italic_M ⋅ divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ ( # roman_Λ ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ,
WL2(#Λ,Δt,M)subscript𝑊superscript𝐿2#ΛΔ𝑡𝑀\displaystyle W_{L^{2}}(\#\Lambda,\Delta t,M)italic_W start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( # roman_Λ , roman_Δ italic_t , italic_M ) MTΔt((#Λ)2+#𝒥MP#Λ),absent𝑀𝑇Δ𝑡superscript#Λ2#subscript𝒥𝑀𝑃#Λ\displaystyle\approx M\cdot\frac{T}{\Delta t}\cdot\left(\left(\#\Lambda\right)% ^{2}+\#\mathcal{J}_{MP}\cdot\#\Lambda\right),≈ italic_M ⋅ divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ ( ( # roman_Λ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + # caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ⋅ # roman_Λ ) ,

where CPoisubscript𝐶𝑃𝑜𝑖C_{Poi}italic_C start_POSTSUBSCRIPT italic_P italic_o italic_i end_POSTSUBSCRIPT represents the cost to simulate one realization of a Poisson random variable. The main computational cost results from deriving an orthonormal basis (see Remark 3.3). A more detailed derivation of the cost terms is provided in Appendix C. For many applications, such as the MP-IS approach presented in Section 4, the MP must be computed only once, such that the computational cost WMP(#Λ,Δt,M)subscript𝑊𝑀𝑃#ΛΔ𝑡𝑀W_{MP}(\#\Lambda,\Delta t,M)italic_W start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( # roman_Λ , roman_Δ italic_t , italic_M ) can be regarded as an off-line cost.

Remark 3.5 (Simulating MP paths).

The MP-SRN 𝐒¯(t)¯𝐒𝑡\overline{\mathbf{S}}(t)over¯ start_ARG bold_S end_ARG ( italic_t ) can be simulated as a SRN with inhomogeneous propensity function. The explicit TL scheme in (1.2) can be naturally adapted to this setting. However, we note that the IS approach presented in Section 4 does not require explicitly simulating paths of the MP-SRN.

4 Importance Sampling for Higher-dimensional Stochastic Reaction Networks via Markovian Projection

Next, we employ MP to overcome the curse of dimensionality when deriving IS controls from solving (2.3). Specifically, we solve the HJB equations in (2.11) for a reduced-dimensional MP system (refer to Figure 1.2 for a schematic illustration of the approach). Given a suitable projection P:dd¯:𝑃superscript𝑑superscript¯𝑑P:\mathbb{R}^{d}\rightarrow\mathbb{R}^{\overline{d}}italic_P : blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT → blackboard_R start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT and a corresponding final condition g~:d¯:~𝑔superscript¯𝑑\tilde{g}:\mathbb{N}^{\overline{d}}\rightarrow\mathbb{R}over~ start_ARG italic_g end_ARG : blackboard_N start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT → blackboard_R with g~(P(𝐱))=g(𝐱)~𝑔𝑃𝐱𝑔𝐱\tilde{g}(P(\mathbf{x}))=g(\mathbf{x})over~ start_ARG italic_g end_ARG ( italic_P ( bold_x ) ) = italic_g ( bold_x ), the HJB equations (2.11) for the MP process are

u~d¯(T,𝒔)subscript~𝑢¯𝑑𝑇𝒔\displaystyle\tilde{u}_{\overline{d}}(T,\boldsymbol{s})over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT ( italic_T , bold_italic_s ) =g~2(𝒔),𝒔d¯formulae-sequenceabsentsuperscript~𝑔2𝒔𝒔superscript¯𝑑\displaystyle=\tilde{g}^{2}(\boldsymbol{s}),\quad\boldsymbol{s}\in\mathbb{N}^{% \overline{d}}= over~ start_ARG italic_g end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_italic_s ) , bold_italic_s ∈ blackboard_N start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT
du~d¯dt(t,𝒔)𝑑subscript~𝑢¯𝑑𝑑𝑡𝑡𝒔\displaystyle\frac{d\tilde{u}_{\overline{d}}}{dt}(t,\boldsymbol{s})divide start_ARG italic_d over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG ( italic_t , bold_italic_s ) =2j=1Ja¯j(t,𝒔)(u~d¯(t,𝒔)u~d¯(t,max(0,𝒔+𝝂¯j))u~d¯(t,𝒔)),t[0,T],𝒔d¯.formulae-sequenceabsent2superscriptsubscript𝑗1𝐽subscript¯𝑎𝑗𝑡𝒔subscript~𝑢¯𝑑𝑡𝒔subscript~𝑢¯𝑑𝑡0𝒔subscript¯𝝂𝑗subscript~𝑢¯𝑑𝑡𝒔formulae-sequence𝑡0𝑇𝒔superscript¯𝑑\displaystyle=-2\sum_{j=1}^{J}\overline{a}_{j}(t,\boldsymbol{s})\left(\sqrt{% \tilde{u}_{\overline{d}}(t,\boldsymbol{s})\tilde{u}_{\overline{d}}(t,\max(0,% \boldsymbol{s}+\overline{\boldsymbol{\nu}}_{j}))}-\tilde{u}_{\overline{d}}(t,% \boldsymbol{s})\right),\quad t\in[0,T],\boldsymbol{s}\in\mathbb{N}^{\overline{% d}}.= - 2 ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_italic_s ) ( square-root start_ARG over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT ( italic_t , bold_italic_s ) over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT ( italic_t , roman_max ( 0 , bold_italic_s + over¯ start_ARG bold_italic_ν end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) end_ARG - over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT ( italic_t , bold_italic_s ) ) , italic_t ∈ [ 0 , italic_T ] , bold_italic_s ∈ blackboard_N start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT . (4.1)

For observables of the type g(𝐱)=𝟙{xi>γ}𝑔𝐱subscript1subscript𝑥𝑖𝛾g(\mathbf{x})=\mathbbm{1}_{\{x_{i}>\gamma\}}italic_g ( bold_x ) = blackboard_1 start_POSTSUBSCRIPT { italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT > italic_γ } end_POSTSUBSCRIPT, we use an MP to a (d¯=1¯𝑑1\overline{d}=1over¯ start_ARG italic_d end_ARG = 1)-dimensional process via projection (3.1), and the final condition is approximated by a positive sigmoid (see (2.12)). The solution of (4) is the value function u~d¯subscript~𝑢¯𝑑\tilde{u}_{\overline{d}}over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT of the d¯¯𝑑\overline{d}over¯ start_ARG italic_d end_ARG-dimensional MP process. To obtain continuous-time IS controls for the d𝑑ditalic_d-dimensional SRN, we substitute the value function u~(t,𝐱)~𝑢𝑡𝐱\tilde{u}\left(t,\mathbf{x}\right)over~ start_ARG italic_u end_ARG ( italic_t , bold_x ) of the full-dimensional process in (2.10) with the value function u~d¯(t,P(𝐱))subscript~𝑢¯𝑑𝑡𝑃𝐱\tilde{u}_{\overline{d}}(t,P(\mathbf{x}))over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT ( italic_t , italic_P ( bold_x ) ) of the MP-SRN:

δ¯j(t,𝐱)=aj(𝐱)u~d¯(t,max(0,P(𝐱+𝝂j)))u~d¯(t,P(𝐱)) for 𝐱d,t[0,T].formulae-sequencesubscript¯𝛿𝑗𝑡𝐱subscript𝑎𝑗𝐱subscript~𝑢¯𝑑𝑡0𝑃𝐱subscript𝝂𝑗subscript~𝑢¯𝑑𝑡𝑃𝐱 for 𝐱superscript𝑑𝑡0𝑇\displaystyle\overline{\delta}_{j}(t,\mathbf{x})={a_{j}(\mathbf{x})}\sqrt{% \frac{\tilde{u}_{\overline{d}}\left(t,\max\left(0,P(\mathbf{x}+\boldsymbol{\nu% }_{j})\right)\right)}{\tilde{u}_{\overline{d}}\left(t,P(\mathbf{x})\right)}}% \text{ for }\mathbf{x}\in\mathbb{N}^{d},t\in[0,T].over¯ start_ARG italic_δ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) = italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) square-root start_ARG divide start_ARG over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT ( italic_t , roman_max ( 0 , italic_P ( bold_x + bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) ) end_ARG start_ARG over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT ( italic_t , italic_P ( bold_x ) ) end_ARG end_ARG for bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT , italic_t ∈ [ 0 , italic_T ] . (4.2)
Remark 4.1 (Alternative MP-IS approach).

In the presented approach, we map the value function of the d¯¯𝑑\overline{d}over¯ start_ARG italic_d end_ARG-dimensional MP process to the full-dimensional SRNs. Alternatively, one could also map the optimal controls from the d¯¯𝑑\overline{d}over¯ start_ARG italic_d end_ARG-dimensional MP-SRN to the full-dimensional SRNs, leading to the following controls:

δ~jd¯(t,𝐱)=a¯j(t,P(𝐱))u~d¯(t,max(0,P(𝐱+𝝂j)))u~d¯(t,P(𝐱)), for 𝐱d,t[0,T].formulae-sequencesubscriptsuperscript~𝛿¯𝑑𝑗𝑡𝐱subscript¯𝑎𝑗𝑡𝑃𝐱subscript~𝑢¯𝑑𝑡0𝑃𝐱subscript𝝂𝑗subscript~𝑢¯𝑑𝑡𝑃𝐱formulae-sequence for 𝐱superscript𝑑𝑡0𝑇\displaystyle\tilde{\delta}^{\overline{d}}_{j}(t,\mathbf{x})=\overline{a}_{j}(% t,P(\mathbf{x}))\sqrt{\frac{\tilde{u}_{\overline{d}}\left(t,\max\left(0,P(% \mathbf{x}+\boldsymbol{\nu}_{j})\right)\right)}{\tilde{u}_{\overline{d}}\left(% t,P(\mathbf{x})\right)}},\text{ for }\mathbf{x}\in\mathbb{N}^{d},t\in[0,T].over~ start_ARG italic_δ end_ARG start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) = over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , italic_P ( bold_x ) ) square-root start_ARG divide start_ARG over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT ( italic_t , roman_max ( 0 , italic_P ( bold_x + bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) ) end_ARG start_ARG over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT ( italic_t , italic_P ( bold_x ) ) end_ARG end_ARG , for bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT , italic_t ∈ [ 0 , italic_T ] . (4.3)

The numerical experiments demonstrate that this approach results in a comparable variance reduction to the approach presented in (4.2).

Remark 4.2 (Adaptive MP for d¯>1¯𝑑1\overline{d}>1over¯ start_ARG italic_d end_ARG > 1).

In (4.2), when utilizing u~d¯subscript~𝑢¯𝑑\tilde{u}_{\overline{d}}over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT as the value function for the d𝑑ditalic_d-dimensional control, we introduce a bias to the optimal IS controls by approximating u~(t,𝐱)~𝑢𝑡𝐱\tilde{u}(t,\mathbf{x})over~ start_ARG italic_u end_ARG ( italic_t , bold_x ) by u~d¯(t,P(𝐱))subscript~𝑢¯𝑑𝑡𝑃𝐱\tilde{u}_{\overline{d}}(t,P(\mathbf{x}))over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT ( italic_t , italic_P ( bold_x ) ) for 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT and t[0,T]𝑡0𝑇t\in[0,T]italic_t ∈ [ 0 , italic_T ]. For the case d¯=d¯𝑑𝑑\overline{d}=dover¯ start_ARG italic_d end_ARG = italic_d, we have u~d¯(t,P(𝐱))=u~(t,𝐱)subscript~𝑢¯𝑑𝑡𝑃𝐱~𝑢𝑡𝐱\tilde{u}_{\overline{d}}(t,P(\mathbf{x}))=\tilde{u}(t,\mathbf{x})over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT ( italic_t , italic_P ( bold_x ) ) = over~ start_ARG italic_u end_ARG ( italic_t , bold_x ) and the MP produces the optimal IS control for the full-dimensional SRNs. For d¯<d¯𝑑𝑑\overline{d}<dover¯ start_ARG italic_d end_ARG < italic_d, this equality does not hold, since the interaction (correlation effects) between non-projected species are not taken into account in the MP SRNs, because the MP only ensures that the marginal distributions of P(𝐗(t)){𝐗(0)=𝐱0}evaluated-at𝑃𝐗𝑡𝐗0subscript𝐱0P(\mathbf{X}(t))\mid_{\{\mathbf{X}(0)=\mathbf{x}_{0}\}}italic_P ( bold_X ( italic_t ) ) ∣ start_POSTSUBSCRIPT { bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT and 𝑺¯(t){𝐗(0)=𝐱0}evaluated-at¯𝑺𝑡𝐗0subscript𝐱0\overline{\boldsymbol{S}}(t)\mid_{\{\mathbf{X}(0)=\mathbf{x}_{0}\}}over¯ start_ARG bold_italic_S end_ARG ( italic_t ) ∣ start_POSTSUBSCRIPT { bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT are identical. This can be seen in examples in which reactions occur with P(𝝂j)=0𝑃subscript𝝂𝑗0P(\boldsymbol{\nu}_{j})=0italic_P ( bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = 0. Those reactions, are not present in the MP and; thus, are not included in the IS scheme. For the extreme case, d¯=1¯𝑑1\overline{d}=1over¯ start_ARG italic_d end_ARG = 1, we expect to achieve the least variance reduction which could be already substantial and satisfactory for many examples as we show in our numerical experiments. However, examples could exist where a projection to dimension d¯=1¯𝑑1\overline{d}=1over¯ start_ARG italic_d end_ARG = 1 is insufficient to achieve a desired variance reduction. In this case, we can adaptively choose a better projection with increased dimension d¯=1,2,¯𝑑12\overline{d}=1,2,\dotsover¯ start_ARG italic_d end_ARG = 1 , 2 , … until a sufficient variance reduction is achieved. This will imply an increased computational cost in the MP and in solving the HJB equations (4) for 𝐱d¯𝐱superscript¯𝑑\mathbf{x}\in\mathbb{N}^{\overline{d}}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT. Investigating the effect of d¯¯𝑑\overline{d}over¯ start_ARG italic_d end_ARG on improving the variance reduction of our approach is left for a future work.

To derive an MP-IS-MC estimator for a given uniform time grid 0=t0t1tN=T0subscript𝑡0subscript𝑡1subscript𝑡𝑁𝑇0=t_{0}\leq t_{1}\leq\dots\leq t_{N}=T0 = italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≤ italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ ⋯ ≤ italic_t start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT = italic_T with step size ΔtΔ𝑡\Delta troman_Δ italic_t, we generate IS paths using the scheme in (2.1) with IS control parameters δn,jΔt(𝐱)=δ¯j(tn,𝐱)superscriptsubscript𝛿𝑛𝑗Δ𝑡𝐱subscript¯𝛿𝑗subscript𝑡𝑛𝐱\delta_{n,j}^{\Delta t}(\mathbf{x})=\overline{\delta}_{j}(t_{n},\mathbf{x})italic_δ start_POSTSUBSCRIPT italic_n , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ italic_t end_POSTSUPERSCRIPT ( bold_x ) = over¯ start_ARG italic_δ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , bold_x ), as in (4.2), for j=1,,J,𝐱d,n=0,,N1formulae-sequence𝑗1𝐽formulae-sequence𝐱superscript𝑑𝑛0𝑁1j=1,\dots,J,\mathbf{x}\in\mathbb{R}^{d},n=0,\dots,N-1italic_j = 1 , … , italic_J , bold_x ∈ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT , italic_n = 0 , … , italic_N - 1. Figure 4.1 presents a schematic illustration of the entire derivation of the MP-IS-MC estimator.

This computational work consists of three cost contributions:

WMPISMCsubscript𝑊𝑀𝑃𝐼𝑆𝑀𝐶\displaystyle W_{MP-IS-MC}italic_W start_POSTSUBSCRIPT italic_M italic_P - italic_I italic_S - italic_M italic_C end_POSTSUBSCRIPT (#Λ,Δt,M,Mfw)#ΛΔ𝑡𝑀subscript𝑀𝑓𝑤\displaystyle(\#\Lambda,\Delta t,M,M_{fw})( # roman_Λ , roman_Δ italic_t , italic_M , italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT ) (4.4)
WMP(#Λ,Δt,M)+WHJB(#Λ)+Wforward(Δt,Mfw),absentsubscript𝑊𝑀𝑃#ΛΔ𝑡𝑀subscript𝑊𝐻𝐽𝐵#Λsubscript𝑊𝑓𝑜𝑟𝑤𝑎𝑟𝑑Δ𝑡subscript𝑀𝑓𝑤\displaystyle\approx W_{MP}(\#\Lambda,\Delta t,M)+W_{HJB}(\#\Lambda)+W_{% forward}(\Delta t,M_{fw}),≈ italic_W start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( # roman_Λ , roman_Δ italic_t , italic_M ) + italic_W start_POSTSUBSCRIPT italic_H italic_J italic_B end_POSTSUBSCRIPT ( # roman_Λ ) + italic_W start_POSTSUBSCRIPT italic_f italic_o italic_r italic_w italic_a italic_r italic_d end_POSTSUBSCRIPT ( roman_Δ italic_t , italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT ) ,

where WMP(#Λ,Δt,M)subscript𝑊𝑀𝑃#ΛΔ𝑡𝑀W_{MP}(\#\Lambda,\Delta t,M)italic_W start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ( # roman_Λ , roman_Δ italic_t , italic_M ) denotes the off-line cost to derive the MP (see (3.10)), WHJB(#Λ)subscript𝑊𝐻𝐽𝐵#ΛW_{HJB}(\#\Lambda)italic_W start_POSTSUBSCRIPT italic_H italic_J italic_B end_POSTSUBSCRIPT ( # roman_Λ ) represents the cost to solve the HJB (4) for the d¯¯𝑑\overline{d}over¯ start_ARG italic_d end_ARG-dimensional MP-SRN, and Wforward(Δt,Mfw)subscript𝑊𝑓𝑜𝑟𝑤𝑎𝑟𝑑Δ𝑡subscript𝑀𝑓𝑤W_{forward}(\Delta t,M_{fw})italic_W start_POSTSUBSCRIPT italic_f italic_o italic_r italic_w italic_a italic_r italic_d end_POSTSUBSCRIPT ( roman_Δ italic_t , italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT ) indicates the cost of deriving Mfwsubscript𝑀𝑓𝑤M_{fw}italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT IS paths. The cost to solve the HJB (4) WHJB(#Λ)subscript𝑊𝐻𝐽𝐵#ΛW_{HJB}(\#\Lambda)italic_W start_POSTSUBSCRIPT italic_H italic_J italic_B end_POSTSUBSCRIPT ( # roman_Λ ) depends on the used solver, and the cost for the forward run has the following dominant terms:

Wforward(Δt,Mfw)MfwTΔt(JCPoi+Clik+#𝒥MPCδ),subscript𝑊𝑓𝑜𝑟𝑤𝑎𝑟𝑑Δ𝑡subscript𝑀𝑓𝑤subscript𝑀𝑓𝑤𝑇Δ𝑡𝐽subscript𝐶𝑃𝑜𝑖subscript𝐶𝑙𝑖𝑘#subscript𝒥𝑀𝑃subscript𝐶𝛿\displaystyle W_{forward}(\Delta t,M_{fw})\approx M_{fw}\cdot\frac{T}{\Delta t% }\cdot(J\cdot C_{Poi}+C_{lik}+\#\mathcal{J}_{MP}\cdot C_{\delta}),italic_W start_POSTSUBSCRIPT italic_f italic_o italic_r italic_w italic_a italic_r italic_d end_POSTSUBSCRIPT ( roman_Δ italic_t , italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT ) ≈ italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT ⋅ divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ ( italic_J ⋅ italic_C start_POSTSUBSCRIPT italic_P italic_o italic_i end_POSTSUBSCRIPT + italic_C start_POSTSUBSCRIPT italic_l italic_i italic_k end_POSTSUBSCRIPT + # caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ⋅ italic_C start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT ) , (4.5)

where Cδsubscript𝐶𝛿C_{\delta}italic_C start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT is the cost to evaluate (4.2).

Figure 4.1: Schematic diagram MP-IS-MC. The costs of the operations in the first line (blue boxes) are off-line.
{NoHyper} generate M𝑀Mitalic_M TL paths Gram-Schmidt derive orthogonal polynomials (see Remark 3.3) L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT regression derive MP propensity a¯jsubscript¯𝑎𝑗\overline{a}_{j}over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT for j𝒥MP𝑗subscript𝒥𝑀𝑃j\in\mathcal{J}_{MP}italic_j ∈ caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT (see Section 3.2) solve reduced- dimensional HJB for the approximate value function u~d¯subscript~𝑢¯𝑑\tilde{u}_{\overline{d}}over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUBSCRIPT (see (4)) generate Mfwsubscript𝑀𝑓𝑤M_{fw}italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT IS-TL path using (4.2) MP-IS-MC estimator Markovian Projection
Remark 4.3 (Further applications of MP in the context of SRNs).

In this work, we use the described MP for dimension reduction to derive a sub-optimal change of measure for IS, but the same MP framework can be used for other applications, such as solving the chemical master equation [40] or the Kolmogorov backward equations [41]. We intend to explore these directions in a future work.

5 Numerical Experiments and Results

Through Examples 5.1 and 5.2, we demonstrate the advantages of the proposed MP-IS approach compared with the standard MC approach for rare event estimations. We numerically demonstrate that the proposed approach achieves a substantial variance reduction compared with standard MC estimators when applied to SRNs with various dimensions.

Example 5.1 (Michaelis–Menten enzyme kinetics [44]).

The Michaelis-Menten enzyme kinetics are enzyme-catalyzed reactions describing the interaction of an enzyme E𝐸Eitalic_E with a substrate S𝑆Sitalic_S, resulting in a product P𝑃Pitalic_P:

E+Sθ1C,Cθ2E+S,Cθ3E+P,𝐸𝑆subscript𝜃1𝐶𝐶subscript𝜃2𝐸𝑆𝐶subscript𝜃3𝐸𝑃\displaystyle E+S\overset{\theta_{1}}{\rightarrow}C,~{}~{}C\overset{\theta_{2}% }{\rightarrow}E+S,~{}~{}C\overset{\theta_{3}}{\rightarrow}E+P,italic_E + italic_S start_OVERACCENT italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG italic_C , italic_C start_OVERACCENT italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG italic_E + italic_S , italic_C start_OVERACCENT italic_θ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG italic_E + italic_P ,

where θ=(0.001,0.005,0.01)𝜃superscript0.0010.0050.01top\theta=(0.001,0.005,0.01)^{\top}italic_θ = ( 0.001 , 0.005 , 0.01 ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT. We consider the initial state 𝐗0=(E(0),S(0),C(0),P(0))=(100,100,0,0)subscript𝐗0superscript𝐸0𝑆0𝐶0𝑃0topsuperscript10010000top\mathbf{X}_{0}=(E(0),S(0),C(0),P(0))^{\top}=(100,100,0,0)^{\top}bold_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ( italic_E ( 0 ) , italic_S ( 0 ) , italic_C ( 0 ) , italic_P ( 0 ) ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT = ( 100 , 100 , 0 , 0 ) start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT and the final time T=1𝑇1T=1italic_T = 1. The corresponding propensity and the stoichiometric matrix are given by

a(𝐱)=(θ1ESθ2Cθ3C),𝝂=(111110111001).formulae-sequence𝑎𝐱subscript𝜃1𝐸𝑆subscript𝜃2𝐶subscript𝜃3𝐶𝝂111110111001\displaystyle a(\textbf{x})=\left(\begin{array}[]{c}\theta_{1}ES\\ \theta_{2}C\\ \theta_{3}C\end{array}\right),\quad\boldsymbol{\nu}=\left(\begin{array}[]{ccc}% -1&1&1\\ -1&1&0\\ 1&-1&-1\\ 0&0&1\end{array}\right).italic_a ( x ) = ( start_ARRAY start_ROW start_CELL italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_E italic_S end_CELL end_ROW start_ROW start_CELL italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_C end_CELL end_ROW start_ROW start_CELL italic_θ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_C end_CELL end_ROW end_ARRAY ) , bold_italic_ν = ( start_ARRAY start_ROW start_CELL - 1 end_CELL start_CELL 1 end_CELL start_CELL 1 end_CELL end_ROW start_ROW start_CELL - 1 end_CELL start_CELL 1 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 1 end_CELL start_CELL - 1 end_CELL start_CELL - 1 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 1 end_CELL end_ROW end_ARRAY ) .

The observable of interest is g(𝐱)=𝟏{x3>22}𝑔𝐱subscript1subscript𝑥322g(\mathbf{x})=\mathbf{1}_{\{x_{3}>22\}}italic_g ( bold_x ) = bold_1 start_POSTSUBSCRIPT { italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT > 22 } end_POSTSUBSCRIPT.

Example 5.2 (Goutsias’s model of regulated transcription [28, 33]).

The model describes a transcription regulation through the following six molecules:
Protein monomer (M𝑀Mitalic_M), Transcription factor (D𝐷Ditalic_D), mRNA (RNA𝑅𝑁𝐴RNAitalic_R italic_N italic_A), Unbound DNA (DNA𝐷𝑁𝐴DNAitalic_D italic_N italic_A), DNA bound at one site (DNAD𝐷𝑁𝐴𝐷DNA\cdot Ditalic_D italic_N italic_A ⋅ italic_D), DNA bounded at two sites (DNA2D𝐷𝑁𝐴2𝐷DNA\cdot 2Ditalic_D italic_N italic_A ⋅ 2 italic_D). These species interact through the following 10 reaction channels R N A θ1RNA+M,subscript𝜃1𝑅𝑁𝐴𝑀\overset{\theta_{1}}{\rightarrow}RNA+M,start_OVERACCENT italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG italic_R italic_N italic_A + italic_M , M θ2,subscript𝜃2\overset{\theta_{2}}{\rightarrow}\varnothing,start_OVERACCENT italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG ∅ , D N A \cdot D θ3RNA+DNAD,subscript𝜃3𝑅𝑁𝐴𝐷𝑁𝐴𝐷\overset{\theta_{3}}{\rightarrow}RNA+DNA\cdot D,start_OVERACCENT italic_θ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG italic_R italic_N italic_A + italic_D italic_N italic_A ⋅ italic_D , R N A θ4,subscript𝜃4\overset{\theta_{4}}{\rightarrow}\varnothing,start_OVERACCENT italic_θ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG ∅ , D N A+D θ5DNAD,subscript𝜃5𝐷𝑁𝐴𝐷\overset{\theta_{5}}{\rightarrow}DNA\cdot D,start_OVERACCENT italic_θ start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG italic_D italic_N italic_A ⋅ italic_D , D N A \cdot D θ6DNA+D,subscript𝜃6𝐷𝑁𝐴𝐷\overset{\theta_{6}}{\rightarrow}DNA+D,start_OVERACCENT italic_θ start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG italic_D italic_N italic_A + italic_D , D N A \cdot D+D θ7DNA2D,subscript𝜃7𝐷𝑁𝐴2𝐷\overset{\theta_{7}}{\rightarrow}DNA\cdot 2D,start_OVERACCENT italic_θ start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG italic_D italic_N italic_A ⋅ 2 italic_D , D N A \cdot 2 D θ8DNAD+D,subscript𝜃8𝐷𝑁𝐴𝐷𝐷\overset{\theta_{8}}{\rightarrow}DNA\cdot D+D,start_OVERACCENT italic_θ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG italic_D italic_N italic_A ⋅ italic_D + italic_D , 2M θ9D,subscript𝜃9𝐷\overset{\theta_{9}}{\rightarrow}D,start_OVERACCENT italic_θ start_POSTSUBSCRIPT 9 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG italic_D , D θ102Msubscript𝜃102𝑀\overset{\theta_{10}}{\rightarrow}2Mstart_OVERACCENT italic_θ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT end_OVERACCENT start_ARG → end_ARG 2 italic_M,

where (θ1,,θ10)=(0.043,0.0007,0.0715,0.0039,0.0199,0.479,0.000199,8.77×1012,0.083,0.5)subscript𝜃1subscript𝜃100.0430.00070.07150.00390.01990.4790.0001998.77superscript10120.0830.5(\theta_{1},\dots,\theta_{10})=(0.043,0.0007,0.0715,0.0039,0.0199,0.479,0.0001% 99,8.77\times 10^{-12},0.083,0.5)( italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_θ start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ) = ( 0.043 , 0.0007 , 0.0715 , 0.0039 , 0.0199 , 0.479 , 0.000199 , 8.77 × 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT , 0.083 , 0.5 ). As the initial state, we use 𝐗0=(M(0),D(0),RNA(0),DNA(0),DNAD(0),DNA2D(0))=(2,6,0,0,2,0)subscript𝐗0𝑀0𝐷0𝑅𝑁𝐴0𝐷𝑁𝐴0𝐷𝑁𝐴𝐷0𝐷𝑁𝐴2𝐷0260020\mathbf{X}_{0}=(M(0),D(0),RNA(0),DNA(0),DNA\cdot D(0),DNA\cdot 2D(0))=(2,6,0,0% ,2,0)bold_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ( italic_M ( 0 ) , italic_D ( 0 ) , italic_R italic_N italic_A ( 0 ) , italic_D italic_N italic_A ( 0 ) , italic_D italic_N italic_A ⋅ italic_D ( 0 ) , italic_D italic_N italic_A ⋅ 2 italic_D ( 0 ) ) = ( 2 , 6 , 0 , 0 , 2 , 0 ), and the final time is T=1𝑇1T=1italic_T = 1. We aim to estimate the rare event probability (D(T)>8)𝐷𝑇8\mathbb{P}(D(T)>8)blackboard_P ( italic_D ( italic_T ) > 8 ).

5.1 Markovian Projection Results

Through simulations for Examples 5.1 and 5.2, we numerically demonstrate that the distribution of the MP process 𝑺¯(T){𝐗0=𝐱0}evaluated-at¯𝑺𝑇subscript𝐗0subscript𝐱0\overline{\boldsymbol{S}}(T)\mid_{\{\mathbf{X}_{0}=\mathbf{x}_{0}\}}over¯ start_ARG bold_italic_S end_ARG ( italic_T ) ∣ start_POSTSUBSCRIPT { bold_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT matches the conditional distribution of the projected process 𝑺(t){𝐗0=𝐱0}=P(𝐗(t)){𝐗0=𝐱0}evaluated-at𝑺𝑡subscript𝐗0subscript𝐱0evaluated-at𝑃𝐗𝑡subscript𝐗0subscript𝐱0\boldsymbol{S}(t)\mid_{\{\mathbf{X}_{0}=\mathbf{x}_{0}\}}=P(\mathbf{X}(t))\mid% _{\{\mathbf{X}_{0}=\mathbf{x}_{0}\}}bold_italic_S ( italic_t ) ∣ start_POSTSUBSCRIPT { bold_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT = italic_P ( bold_X ( italic_t ) ) ∣ start_POSTSUBSCRIPT { bold_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT, as shown in Theorem 3.1. For both examples, we use an MP projection with d¯=1¯𝑑1\overline{d}=1over¯ start_ARG italic_d end_ARG = 1 using the projection given in (3.1), where the projected species is indexed as i=3𝑖3i=3italic_i = 3 in Example 5.1 and as i=2𝑖2i=2italic_i = 2 in Example 5.2.

The MP is based on M=104𝑀superscript104M=10^{4}italic_M = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT TL sample paths with a step size of Δt=28Δ𝑡superscript28\Delta t=2^{-8}roman_Δ italic_t = 2 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT and uses the orthonormal basis of polynomials described in Remark 3.3 with Λ={0,1,2}×{0,1,2}Λ012012\Lambda=\{0,1,2\}\times\{0,1,2\}roman_Λ = { 0 , 1 , 2 } × { 0 , 1 , 2 } for the L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT regression. For the numerical comparison of the full-dimensional SRN with the MP-SRN, we use explicit TL paths with step size Δt=28Δ𝑡superscript28\Delta t=2^{-8}roman_Δ italic_t = 2 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT (see Remark 3.5). In Figure 5.1, we show sample paths of the original SRN and the mimicking MP-SRN. Figure 5.2 shows the relative occurrences of states at final time T𝑇Titalic_T with Mfw=104subscript𝑀𝑓𝑤superscript104M_{fw}=10^{4}italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT sample paths, comparing the TL estimate of P(𝐗(t)){𝐗0=𝐱0}evaluated-at𝑃𝐗𝑡subscript𝐗0subscript𝐱0P(\mathbf{X}(t))\mid_{\{\mathbf{X}_{0}=\mathbf{x}_{0}\}}italic_P ( bold_X ( italic_t ) ) ∣ start_POSTSUBSCRIPT { bold_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT and the TL-MP estimate of 𝑺¯(T){𝐗0=𝐱0}evaluated-at¯𝑺𝑇subscript𝐗0subscript𝐱0\overline{\boldsymbol{S}}(T)\mid_{\{\mathbf{X}_{0}=\mathbf{x}_{0}\}}over¯ start_ARG bold_italic_S end_ARG ( italic_T ) ∣ start_POSTSUBSCRIPT { bold_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT. In both examples, the one-dimensional MP process mimics the distribution of the state of interest Xi(T)subscript𝑋𝑖𝑇X_{i}(T)italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_T ) of the original SRNs. Further quantification and analyses of the MP error are left for future work. In this work, a detailed analysis of the MP error is less relevant because the MP is used as a tool to derive IS controls for the full-dimensional process, and the IS is bias-free with respect to the TL scheme.

Figure 5.1: Sample paths of TL and MP-TL for a step size of Δt=28Δ𝑡superscript28\Delta t=2^{-8}roman_Δ italic_t = 2 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT.
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 5.2: Relative occurrences of states at final time T𝑇Titalic_T in Mtest=104subscript𝑀𝑡𝑒𝑠𝑡superscript104M_{test}=10^{4}italic_M start_POSTSUBSCRIPT italic_t italic_e italic_s italic_t end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT sample paths, comparing the TL estimate of P(𝐗(t)){𝐗0=𝐱0}evaluated-at𝑃𝐗𝑡subscript𝐗0subscript𝐱0P(\mathbf{X}(t))\mid_{\{\mathbf{X}_{0}=\mathbf{x}_{0}\}}italic_P ( bold_X ( italic_t ) ) ∣ start_POSTSUBSCRIPT { bold_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT and the MP estimate of 𝑺¯(T){𝐗0=𝐱0}evaluated-at¯𝑺𝑇subscript𝐗0subscript𝐱0\overline{\boldsymbol{S}}(T)\mid_{\{\mathbf{X}_{0}=\mathbf{x}_{0}\}}over¯ start_ARG bold_italic_S end_ARG ( italic_T ) ∣ start_POSTSUBSCRIPT { bold_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT } end_POSTSUBSCRIPT. We set the step size to Δt=28Δ𝑡superscript28\Delta t=2^{-8}roman_Δ italic_t = 2 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT for the sample paths. The MP is based on M=104𝑀superscript104M=10^{4}italic_M = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT TL paths with a step size of Δt=28Δ𝑡superscript28\Delta t=2^{-8}roman_Δ italic_t = 2 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT and uses the orthogonal basis of polynomials described in Remark 3.3, where Λ={0,1,2}×{0,1,2}Λ012012\Lambda=\{0,1,2\}\times\{0,1,2\}roman_Λ = { 0 , 1 , 2 } × { 0 , 1 , 2 }.
Refer to caption
(a)
Refer to caption
(b)

5.2 Makovian Projection-Importance Sampling Results

For the numerical experiments, we use a four-dimensional and a six-dimensional SRNs with the observable g(𝐱)=𝟙{xi>γ}𝑔𝐱subscript1subscript𝑥𝑖𝛾g(\mathbf{x})=\mathbbm{1}_{\{x_{i}>\gamma\}}italic_g ( bold_x ) = blackboard_1 start_POSTSUBSCRIPT { italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT > italic_γ } end_POSTSUBSCRIPT, where i𝑖iitalic_i and γ𝛾\gammaitalic_γ are specified in Examples 5.1 and 5.2. Figure 5.2 indicates that this observable leads to a rare event probability estimation for which an MC estimate is insufficient. We use the workflow in Figure 4.1 with separate simulations for various ΔtΔ𝑡\Delta troman_Δ italic_t values for the MP-IS simulations. The MP is based on M=104𝑀superscript104M=10^{4}italic_M = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT TL sample paths each. The MP-IS-MC estimator, the sample variance, and the kurtosis estimate are based on Mfw=106subscript𝑀𝑓𝑤superscript106M_{fw}=10^{6}italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT IS sample paths.

The relative error is more relevant than the absolute error for rare event probabilities. Therefore, we display the squared coefficient of variation [11, 13] in the simulations results, which is given by the following for a random variable X𝑋Xitalic_X:

Varrel[X]=Var[X]𝔼[X]2.𝑉𝑎subscript𝑟𝑟𝑒𝑙delimited-[]𝑋𝑉𝑎𝑟delimited-[]𝑋𝔼superscriptdelimited-[]𝑋2\displaystyle Var_{rel}[X]=\frac{Var[X]}{\mathbb{E}[X]^{2}}.italic_V italic_a italic_r start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT [ italic_X ] = divide start_ARG italic_V italic_a italic_r [ italic_X ] end_ARG start_ARG blackboard_E [ italic_X ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (5.1)

The kurtosis is a good indicator of the robustness of the variance estimator (see [10, 11] for the connection between the sample variance and kurtosis).

Figure 5.3 shows the simulation results for the four-dimensional Example 5.1 for different step sizes ΔtΔ𝑡\Delta troman_Δ italic_t. The quantity of interest is a rare event probability with a magnitude of 105superscript10510^{-5}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT. For a step size of Δt=210Δ𝑡superscript210\Delta t=2^{-10}roman_Δ italic_t = 2 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT, the proposed MP-IS approach reduces the squared coefficient of variation by a factor of 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT compared to the standard MC-TL approach. The third plot in Figure 5.3 indicates that the kurtosis of the proposed MP-IS approach is below the kurtosis for standard TL for all observed step sizes ΔtΔ𝑡\Delta troman_Δ italic_t, confirming that the proposed approach results in a robust variance estimator. In Figure 5.3 (d), we show the required number of sample paths to reach a prescribed relative tolerance (see (1.12) and (1.11)). We compare the required number of sample paths for standard TL with the sample paths in the forward run of the IS-TL estimator, Mfwsubscript𝑀𝑓𝑤M_{fw}italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT, and the total number of paths M+Mfw𝑀subscript𝑀𝑓𝑤M+M_{fw}italic_M + italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT including M=104𝑀superscript104M=10^{4}italic_M = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT TL paths to derive the MP (see Section 3.3). We observe that for small tolerances, the additional paths to derive the MP become negligible compared to the cost of the forward IS-TL run. Further, we reduce the number of paths significantly compared to the standard TL approach. For small tolerances, the reduction of sample paths is up to a factor of approximately 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT compared to the standard TL approach.

The second application of the proposed IS approach is the six-dimensional Example 5.2. Figure 5.4 shows that this rare event probability has a magnitude of 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. We observe that, for Δt23Δ𝑡superscript23\Delta t\leq 2^{-3}roman_Δ italic_t ≤ 2 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, the squared coefficient of variation of the proposed MP-IS approach is reduced compared to the standard TL-MC approach. For a step size of Δt=210Δ𝑡superscript210\Delta t=2^{-10}roman_Δ italic_t = 2 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT, this is a variance reduction of a factor of approximately 500500500500. Note that, this example achieves less variance reduction than Example 5.1 due to a less rare quantity of interest. For most step sizes ΔtΔ𝑡\Delta troman_Δ italic_t, the kurtosis of the proposed IS approaches is moderately increased compared to the standard TL estimator, with decreasing kurtosis for smaller ΔtΔ𝑡\Delta troman_Δ italic_t. This outcome indicates a potentially insatiable variance estimator for coarse time steps of Δt>27Δ𝑡superscript27\Delta t>2^{-7}roman_Δ italic_t > 2 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT. For finer time steps, we expect a robust variance estimator. For small tolerances the total number of paths is reduced by a factor of approximately 500500500500.

Figure 5.3: Example 5.1 with separate simulations for different step sizes ΔtΔ𝑡\Delta troman_Δ italic_t for the proposed IS methods and the standard TL-MC estimator: (a) sample mean, (b) squared coefficient of variation, (c) kurtosis, and (d) number of required sample paths to reach a prescribed relative tolerance TOLrelsubscriptTOL𝑟𝑒𝑙\text{TOL}_{rel}TOL start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT. The used MP is based on M=104𝑀superscript104M=10^{4}italic_M = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT TL sample paths each, and the estimators in (a), (b) and (c) are derived based on Mfw=106subscript𝑀𝑓𝑤superscript106M_{fw}=10^{6}italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT sample paths. For (d), we use (1.11) and (1.12) and the dashed line represents the total number of sample paths including the derivation of the MP (i.e. M+Mrel*(TOLrel)𝑀subscriptsuperscript𝑀𝑟𝑒𝑙subscriptTOL𝑟𝑒𝑙M+M^{*}_{rel}(\text{TOL}_{rel})italic_M + italic_M start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT ( TOL start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT )).
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 5.4: Example 5.2 with separate simulations for different step sizes ΔtΔ𝑡\Delta troman_Δ italic_t for the proposed IS methods and the standard TL-MC estimator: (a) sample mean, (b) squared coefficient of variation, (c) kurtosis, and (d) number of required sample paths to reach a prescribed relative tolerance TOLrelsubscriptTOL𝑟𝑒𝑙\text{TOL}_{rel}TOL start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT. The used MP is based on M=104𝑀superscript104M=10^{4}italic_M = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT TL sample paths each, and the estimators in (a), (b) and (c) are derived based on Mfw=106subscript𝑀𝑓𝑤superscript106M_{fw}=10^{6}italic_M start_POSTSUBSCRIPT italic_f italic_w end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT sample paths.
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)

6 Conclusion

In conclusion, this work presented an efficient IS scheme for estimating rare event probabilities for SRNs. We utilized a class of parameterized IS measure changes originally introduced in [11], for which near-optimal IS controls can be derived through a SOC formulation. We showed that the value function associated with this formulation can be expressed as a solution of a set of coupled ODEs, the HJB equations. One challenge encountered in solving the HJB equations is the curse of dimensionality, arising from the high-dimensional SRN. To address this issue, we introduced a dimension reduction approach for the setting of SRNs, namely MP. Then, we used a discrete L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT regression to approximate the propensity and the stoichiometric vector of the MP-SRN. We demonstrated how the MP-SRN can be used for solving a significantly lower-dimensional HJB system, and how the resulting parameters are then mapped back to the full-dimensional SRNs to derive near-optimal IS controls. Our numerical simulations showed substantial variance reduction for the MP-IS-MC estimator compared to the standard MC-TL estimator for rare event probability estimations.

Acknowledgments This publication is based upon work supported by the King Abdullah University of Science and Technology (KAUST) Office of Sponsored Research (OSR) under Award No. OSR-2019-CRG8-4033. This work was performed as part of the Helmholtz School for Data Science in Life, Earth and Energy (HDS-LEE) and received funding from the Helmholtz Association of German Research Centres and the Alexander von Humboldt Foundation. For the purpose of open access, the authors have applied a Creative Commons Attribution (CC BY) licence to any Author Accepted Manuscript version arising from this submission.

References Cited

  • [1] Assyr Abdulle, Yucheng Hu, and Tiejun Li. Chebyshev methods with discrete noise: the τ𝜏\tauitalic_τ-ROCK methods. Journal of Computational Mathematics, pages 195–217, 2010.
  • [2] Tae-Hyuk Ahn, Adrian Sandu, and Xiaoying Han. Implicit simulation methods for stochastic chemical kinetics. arXiv preprint arXiv:1303.3614, 2013.
  • [3] David F Anderson. A modified next reaction method for simulating chemical systems with time dependent propensities and delays. The Journal of chemical physics, 127(21):214107, 2007.
  • [4] David F Anderson and Desmond J Higham. Multilevel Monte Carlo for continuous time Markov chains, with applications in biochemical kinetics. Multiscale Modeling & Simulation, 10(1):146–179, 2012.
  • [5] David F Anderson and Thomas G Kurtz. Stochastic analysis of biochemical systems, volume 1. Springer, 2015.
  • [6] Juan P Aparicio and Hernán G Solari. Population dynamics: Poisson approximation and its relation to the Langevin process. Physical Review Letters, 86(18):4183, 2001.
  • [7] Christian Bayer, Juho Häppölä, and Raúl Tempone. Implied stopping rules for american basket options from Markovian projection. Quantitative Finance, 19(3):371–390, 2019.
  • [8] Christian Bayer, Alvaro Moraes, Raúl Tempone, and Pedro Vilanova. An efficient forward–reverse expectation-maximization algorithm for statistical inference in stochastic reaction networks. Stochastic Analysis and Applications, 34(2):193–231, 2016.
  • [9] Chiheb Ben Hammouda. Hierarchical approximation methods for option pricing and stochastic reaction networks. PhD thesis, 2020.
  • [10] Chiheb Ben Hammouda, Nadhir Ben Rached, and Raúl Tempone. Importance sampling for a robust and efficient multilevel monte carlo estimator for stochastic reaction networks. Statistics and Computing, 30:1665–1689, 2020.
  • [11] Chiheb Ben Hammouda, Nadhir Ben Rached, Raúl Tempone, and Sophia Wiechert. Learning-based importance sampling via stochastic optimal control for stochastic reaction networks. Statistics and Computing, 33(3):58, 2023.
  • [12] Chiheb Ben Hammouda, Alvaro Moraes, and Raúl Tempone. Multilevel hybrid split-step implicit tau-leap. Numerical Algorithms, 74(2):527–560, 2017.
  • [13] Nadhir Ben Rached, Abdul-Lateef Haji-Ali, Gerardo Rubino, and Raúl Tempone. Efficient importance sampling for large sums of independent and identically distributed random variables. Statistics and Computing, 31(6):1–13, 2021.
  • [14] Amel Bentata and Rama Cont. Mimicking the marginal distributions of a semimartingale. arXiv preprint arXiv:0910.3992, 2009.
  • [15] Fred Brauer and Carlos Castillo-Chavez. Mathematical models in population biology and epidemiology, volume 40. Springer.
  • [16] Yang Cao and Linda Petzold. Trapezoidal tau-leaping formula for the stochastic simulation of biochemical systems. Proceedings of Foundations of Systems Biology in Engineering (FOSBE 2005), pages 149–152, 2005.
  • [17] Youfang Cao and Jie Liang. Adaptively biased sequential importance sampling for rare events in reaction networks with comparison to exact solutions from finite buffer dCME method. The Journal of chemical physics, 139(2):07B605_1, 2013.
  • [18] Albert Cohen, Mark A Davenport, and Dany Leviatan. On the stability and accuracy of least squares approximations. Foundations of computational mathematics, 13:819–834, 2013.
  • [19] Bernie J Daigle Jr, Min K Roh, Dan T Gillespie, and Linda R Petzold. Automated estimation of rare event probabilities in biochemical systems. The Journal of chemical physics, 134(4):01B628, 2011.
  • [20] Boualem Djehiche and Björn Löfdahl. Risk aggregation and stochastic claims reserving in disability insurance. Insurance: Mathematics and Economics, 59:100–108, 2014.
  • [21] Stewart N Ethier and Thomas G Kurtz. Markov processes : characterization and convergence. Wiley series in probability and mathematical statistics. J. Wiley & Sons, New York, Chichester, 1986.
  • [22] Michael B Giles. Multilevel Monte Carlo path simulation. Operations Research, 56(3):607–617, 2008.
  • [23] Michael B Giles. Multilevel Monte Carlo methods. Acta Numerica, 24:259–328, 2015.
  • [24] Colin S Gillespie and Andrew Golightly. Guided proposals for efficient weighted stochastic simulation. The Journal of chemical physics, 150(22):224103, 2019.
  • [25] Dan T Gillespie, Min Roh, and Linda R Petzold. Refining the weighted stochastic simulation algorithm. The Journal of chemical physics, 130(17):174103, 2009.
  • [26] Daniel T Gillespie. A general method for numerically simulating the stochastic time evolution of coupled chemical reactions. Journal of computational physics, 22(4):403–434, 1976.
  • [27] Daniel T Gillespie. Approximate accelerated stochastic simulation of chemically reacting systems. The Journal of Chemical Physics, 115(4):1716–1733, 2001.
  • [28] John Goutsias. Quasiequilibrium approximation of fast reaction kinetics in stochastic biochemical systems. The Journal of chemical physics, 122(18):184102, 2005.
  • [29] István Gyöngy. Mimicking the one-dimensional marginal distributions of processes having an Itô differential. Probability theory and related fields, 71(4):501–516, 1986.
  • [30] Floyd B. Hanson. Applied Stochastic Processes and Control for Jump-Diffusions: Modeling, Analysis and Computation. SIAM, Philadelphia, PA, 2007.
  • [31] Carsten Hartmann, Christof Schütte, and Wei Zhang. Model reduction algorithms for optimal control and importance sampling of diffusions. Nonlinearity, 29(8):2298, 2016.
  • [32] Sebastian C Hensel, James B Rawlings, and John Yin. Stochastic kinetic modeling of vesicular stomatitis virus intracellular growth. Bulletin of mathematical biology, 71(7):1671–1692, 2009.
  • [33] Hye-Won Kang and Thomas G Kurtz. Separation of time-scales and model reduction for stochastic reaction networks. 2013.
  • [34] Nikolai Vladimirovich Krylov. Nonlinear elliptic and parabolic equations of the second order. Springer, 1987.
  • [35] Thomas Kurtz and Richard Stockbridge. Stationary solutions and forward equations for controlled and singular martingale problems. 2001.
  • [36] Hiroyuki Kuwahara and Ivan Mura. An efficient and exact stochastic simulation method to analyze rare events in biochemical systems. The Journal of chemical physics, 129(16):10B619, 2008.
  • [37] Frédéric Legoll and Tony Lelievre. Effective dynamics using conditional expectations. Nonlinearity, 23(9):2131, 2010.
  • [38] Christopher Lester, Christian Adam Yates, Michael B Giles, and Ruth E Baker. An adaptive multi-level simulation algorithm for stochastic biological systems. The Journal of chemical physics, 142(2):01B612_1, 2015.
  • [39] Tiejun Li. Analysis of explicit tau-leaping schemes for simulating chemically reacting systems. Multiscale Modeling & Simulation, 6(2):417–436, 2007.
  • [40] Linar Mikeev and Werner Sandmann. Approximate numerical integration of the chemical master equation for stochastic reaction networks. arXiv preprint arXiv:1907.10245, 2019.
  • [41] Alvaro Moraes. Simulation and statistical inference of stochastic reaction networks with applications to epidemic models. PhD thesis, 2015.
  • [42] Alvaro Moraes, Raúl Tempone, and Pedro Vilanova. A multilevel adaptive reaction-splitting simulation method for stochastic reaction networks. SIAM Journal on Scientific Computing, 38(4):A2091–A2117, 2016.
  • [43] Vladimir Piterbarg. Markovian projection method for volatility calibration. Available at SSRN 906473, 2006.
  • [44] Christopher V Rao and Adam P Arkin. Stochastic chemical kinetics and the quasi-steady-state assumption: Application to the Gillespie algorithm. The Journal of chemical physics, 118(11):4999–5010, 2003.
  • [45] Muruhan Rathinam and Hana El Samad. Reversible-equivalent-monomolecular tau: A leaping method for “small number and stiff” stochastic chemical systems. Journal of Computational Physics, 224(2):897–923, 2007.
  • [46] Min K Roh. Data-driven method for efficient characterization of rare event probabilities in biochemical systems. Bulletin of mathematical biology, 81(8):3097–3120, 2019.
  • [47] Min K Roh, Dan T Gillespie, and Linda R Petzold. State-dependent biasing method for importance sampling in the weighted stochastic simulation algorithm. The Journal of chemical physics, 133(17):174106, 2010.
  • [48] Ranjan Srivastava, Lingchong You, Jesse Summers, and John Yin. Stochastic vs. deterministic modeling of intracellular viral kinetics. Journal of theoretical biology, 218(3):309–321, 2002.
  • [49] James N Weiss. The hill equation revisited: uses and misuses. The FASEB Journal, 11(11):835–841, 1997.

Appendix A Proof for Corollary 2.3

Proof.

For 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, we define u~(,𝐱;Δt)~𝑢𝐱Δ𝑡\tilde{u}(\cdot,\mathbf{x};\Delta t)over~ start_ARG italic_u end_ARG ( ⋅ , bold_x ; roman_Δ italic_t ) as the continuous smooth extension of uΔt(,𝐱)subscript𝑢Δ𝑡𝐱u_{\Delta t}(\cdot,\mathbf{x})italic_u start_POSTSUBSCRIPT roman_Δ italic_t end_POSTSUBSCRIPT ( ⋅ , bold_x ) (defined in (2.2)) on [0,T]0𝑇[0,T][ 0 , italic_T ]. Consequently, we denote the continuous-time IS controls by 𝜹(,𝐱):[0,T]𝒜𝐱:𝜹𝐱0𝑇subscript𝒜𝐱\boldsymbol{\delta}(\cdot,\mathbf{x}):[0,T]\rightarrow\mathcal{A}_{\mathbf{x}}bold_italic_δ ( ⋅ , bold_x ) : [ 0 , italic_T ] → caligraphic_A start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT for 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT. Then, the Taylor expansion of u~(t+Δt,𝐱;Δt)~𝑢𝑡Δ𝑡𝐱Δ𝑡\tilde{u}(t+\Delta t,\mathbf{x};\Delta t)over~ start_ARG italic_u end_ARG ( italic_t + roman_Δ italic_t , bold_x ; roman_Δ italic_t ) in t𝑡titalic_t results in the following:

u~(t+Δt,𝐱;Δt)=u~(t,𝐱;Δt)+Δttu~(t,𝐱;Δt)+𝒪(Δt2),𝐱d.formulae-sequence~𝑢𝑡Δ𝑡𝐱Δ𝑡~𝑢𝑡𝐱Δ𝑡Δ𝑡subscript𝑡~𝑢𝑡𝐱Δ𝑡𝒪Δsuperscript𝑡2𝐱superscript𝑑\displaystyle\tilde{u}(t+\Delta t,\mathbf{x};\Delta t)=\tilde{u}(t,\mathbf{x};% \Delta t)+\Delta t\partial_{t}\tilde{u}(t,\mathbf{x};\Delta t)+\mathcal{O}% \left(\Delta t^{2}\right),\mathbf{x}\in\mathbb{N}^{d}.over~ start_ARG italic_u end_ARG ( italic_t + roman_Δ italic_t , bold_x ; roman_Δ italic_t ) = over~ start_ARG italic_u end_ARG ( italic_t , bold_x ; roman_Δ italic_t ) + roman_Δ italic_t ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT over~ start_ARG italic_u end_ARG ( italic_t , bold_x ; roman_Δ italic_t ) + caligraphic_O ( roman_Δ italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT . (A.1)

By the definition of the value function 2.1, the final condition is given by

u~(T,𝐱;Δt)=g2(𝐱),𝐱d.formulae-sequence~𝑢𝑇𝐱Δ𝑡superscript𝑔2𝐱𝐱superscript𝑑\displaystyle\tilde{u}(T,\mathbf{x};\Delta t)=g^{2}(\mathbf{x}),\mathbf{x}\in% \mathbb{N}^{d}.over~ start_ARG italic_u end_ARG ( italic_T , bold_x ; roman_Δ italic_t ) = italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_x ) , bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT .

For t=TΔt,,0𝑡𝑇Δ𝑡0t=T-\Delta t,\ldots,0italic_t = italic_T - roman_Δ italic_t , … , 0, we apply to (2.2) from Theorem 2.2 a Taylor expansion around Δt=0Δ𝑡0\Delta t=0roman_Δ italic_t = 0 to the exponential term and (A.1) to u~(t+Δt,𝐱;Δt)~𝑢𝑡Δ𝑡𝐱Δ𝑡\tilde{u}(t+\Delta t,\mathbf{x};\Delta t)over~ start_ARG italic_u end_ARG ( italic_t + roman_Δ italic_t , bold_x ; roman_Δ italic_t ):

u~(t,𝐱;Δt)~𝑢𝑡𝐱Δ𝑡\displaystyle\tilde{u}(t,\mathbf{x};\Delta t)over~ start_ARG italic_u end_ARG ( italic_t , bold_x ; roman_Δ italic_t ) =inf𝜹(t,𝐱)𝒜𝐱exp((2j=1Jaj(𝐱)+j=1Jδj(t,𝐱))Δt)absentsubscriptinfimum𝜹𝑡𝐱subscript𝒜𝐱2superscriptsubscript𝑗1𝐽subscript𝑎𝑗𝐱superscriptsubscript𝑗1𝐽subscript𝛿𝑗𝑡𝐱Δ𝑡\displaystyle=\inf_{\boldsymbol{\delta}(t,\mathbf{x})\in\mathcal{A}_{\mathbf{x% }}}\exp\left(\left(-2\sum_{j=1}^{J}a_{j}(\mathbf{x})+\sum_{j=1}^{J}\delta_{j}(% t,\mathbf{x})\right)\Delta t\right)= roman_inf start_POSTSUBSCRIPT bold_italic_δ ( italic_t , bold_x ) ∈ caligraphic_A start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_exp ( ( - 2 ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) ) roman_Δ italic_t )
×𝐩J(j=1J(Δtδj(t,𝐱))pjpj!(aj(𝐱)δj(t,𝐱))2pj)u~(t+Δt,max(𝟎,𝐱+𝝂𝐩);Δt)\displaystyle\quad\quad\quad\quad\times\sum_{\mathbf{p}\in\mathbb{N}^{J}}\left% (\prod_{j=1}^{J}\frac{(\Delta t\cdot\delta_{j}(t,\mathbf{x}))^{p_{j}}}{p_{j}!}% \left(\frac{a_{j}(\mathbf{x})}{\delta_{j}(t,\mathbf{x})}\right)^{2p_{j}}\right% )\cdot\tilde{u}(t+\Delta t,\max(\mathbf{0},\mathbf{x}+\boldsymbol{\nu}\mathbf{% p});\Delta t)× ∑ start_POSTSUBSCRIPT bold_p ∈ blackboard_N start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( ∏ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT divide start_ARG ( roman_Δ italic_t ⋅ italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) ) start_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ! end_ARG ( divide start_ARG italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) end_ARG start_ARG italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) end_ARG ) start_POSTSUPERSCRIPT 2 italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) ⋅ over~ start_ARG italic_u end_ARG ( italic_t + roman_Δ italic_t , roman_max ( bold_0 , bold_x + bold_italic_ν bold_p ) ; roman_Δ italic_t )
=inf𝜹(t,𝐱)𝒜𝐱(1+(2j=1Jaj(𝐱)+j=1Jδj(t,𝐱))Δt+𝒪(Δt2))absentsubscriptinfimum𝜹𝑡𝐱subscript𝒜𝐱12superscriptsubscript𝑗1𝐽subscript𝑎𝑗𝐱superscriptsubscript𝑗1𝐽subscript𝛿𝑗𝑡𝐱Δ𝑡𝒪Δsuperscript𝑡2\displaystyle=\inf_{\boldsymbol{\delta}(t,\mathbf{x})\in\mathcal{A}_{\mathbf{x% }}}\left(1+\left(-2\sum_{j=1}^{J}a_{j}(\mathbf{x})+\sum_{j=1}^{J}\delta_{j}(t,% \mathbf{x})\right)\Delta t+\mathcal{O}(\Delta t^{2})\right)= roman_inf start_POSTSUBSCRIPT bold_italic_δ ( italic_t , bold_x ) ∈ caligraphic_A start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( 1 + ( - 2 ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) ) roman_Δ italic_t + caligraphic_O ( roman_Δ italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) )
×[𝐩J(j=1J(Δtδj(t,𝐱))pjpj!(aj(𝐱)δj(t,𝐱))2pj)\displaystyle\quad\quad\quad\quad\times\left[\sum_{\mathbf{p}\in\mathbb{N}^{J}% }\left(\prod_{j=1}^{J}\frac{(\Delta t\cdot\delta_{j}(t,\mathbf{x}))^{p_{j}}}{p% _{j}!}\left(\frac{a_{j}(\mathbf{x})}{\delta_{j}(t,\mathbf{x})}\right)^{2p_{j}}% \right)\right.× [ ∑ start_POSTSUBSCRIPT bold_p ∈ blackboard_N start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( ∏ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT divide start_ARG ( roman_Δ italic_t ⋅ italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) ) start_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ! end_ARG ( divide start_ARG italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) end_ARG start_ARG italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) end_ARG ) start_POSTSUPERSCRIPT 2 italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT )
(u~(t,max(𝟎,𝐱+𝝂𝐩);Δt)+Δttu~(t,max(𝟎,𝐱+𝝂𝐩);Δt)+𝒪(Δt2))]\displaystyle\quad\quad\quad\quad\left.\vphantom{\prod_{j=1}^{J}}\cdot\left(% \tilde{u}\left(t,\max(\mathbf{0},\mathbf{x}+\boldsymbol{\nu}\mathbf{p});\Delta t% \right)+\Delta t\partial_{t}\tilde{u}\left(t,\max(\mathbf{0},\mathbf{x}+% \boldsymbol{\nu}\mathbf{p});\Delta t\right)+\mathcal{O}(\Delta t^{2})\right)\right]⋅ ( over~ start_ARG italic_u end_ARG ( italic_t , roman_max ( bold_0 , bold_x + bold_italic_ν bold_p ) ; roman_Δ italic_t ) + roman_Δ italic_t ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT over~ start_ARG italic_u end_ARG ( italic_t , roman_max ( bold_0 , bold_x + bold_italic_ν bold_p ) ; roman_Δ italic_t ) + caligraphic_O ( roman_Δ italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ) ]
(*)tsubscript𝑡\displaystyle\overset{(*)}{\Longrightarrow}-\partial_{t}start_OVERACCENT ( * ) end_OVERACCENT start_ARG ⟹ end_ARG - ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT u~(t,𝐱;Δt)=inf𝜹(t,𝐱)𝒜𝐱(2j=1Jaj(𝐱)+j=1Jδj(t,𝐱))u~(t,𝐱;Δt)+𝒪(Δt)~𝑢𝑡𝐱Δ𝑡subscriptinfimum𝜹𝑡𝐱subscript𝒜𝐱2superscriptsubscript𝑗1𝐽subscript𝑎𝑗𝐱superscriptsubscript𝑗1𝐽subscript𝛿𝑗𝑡𝐱~𝑢𝑡𝐱Δ𝑡𝒪Δ𝑡\displaystyle\tilde{u}(t,\mathbf{x};\Delta t)=\inf_{\boldsymbol{\delta}(t,% \mathbf{x})\in\mathcal{A}_{\mathbf{x}}}\left(-2\sum_{j=1}^{J}a_{j}(\mathbf{x})% +\sum_{j=1}^{J}\delta_{j}(t,\mathbf{x})\right)\tilde{u}(t,\mathbf{x};\Delta t)% +\mathcal{O}(\Delta t)over~ start_ARG italic_u end_ARG ( italic_t , bold_x ; roman_Δ italic_t ) = roman_inf start_POSTSUBSCRIPT bold_italic_δ ( italic_t , bold_x ) ∈ caligraphic_A start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( - 2 ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) + ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) ) over~ start_ARG italic_u end_ARG ( italic_t , bold_x ; roman_Δ italic_t ) + caligraphic_O ( roman_Δ italic_t )
+(1+𝒪(Δt))[𝐩𝟎Δt|𝐩|1(j=1Jaj(𝐱)2pjpj!(δj(t,𝐱))pj)\displaystyle\quad\quad\quad\quad\quad\quad+(1+\mathcal{O}(\Delta t))\left[% \sum_{\mathbf{p}\neq\mathbf{0}}\Delta t^{|\mathbf{p}|-1}\left(\prod_{j=1}^{J}% \frac{a_{j}(\mathbf{x})^{2p_{j}}}{p_{j}!\cdot\left(\delta_{j}(t,\mathbf{x})% \right)^{p_{j}}}\right)\right.+ ( 1 + caligraphic_O ( roman_Δ italic_t ) ) [ ∑ start_POSTSUBSCRIPT bold_p ≠ bold_0 end_POSTSUBSCRIPT roman_Δ italic_t start_POSTSUPERSCRIPT | bold_p | - 1 end_POSTSUPERSCRIPT ( ∏ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT divide start_ARG italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_x ) start_POSTSUPERSCRIPT 2 italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ! ⋅ ( italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) ) start_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG )
(u~(t,max(𝟎,𝐱+𝝂𝐩);Δt)+𝒪(Δt))],\displaystyle\quad\quad\quad\quad\quad\quad\left.\vphantom{\prod_{j=1}^{J}}% \cdot\left(\tilde{u}\left(t,\max(\mathbf{0},\mathbf{x}+\boldsymbol{\nu}\mathbf% {p});\Delta t\right)+\mathcal{O}(\Delta t)\right)\right],⋅ ( over~ start_ARG italic_u end_ARG ( italic_t , roman_max ( bold_0 , bold_x + bold_italic_ν bold_p ) ; roman_Δ italic_t ) + caligraphic_O ( roman_Δ italic_t ) ) ] , (A.2)

where |𝐩|:=i=1Jpjassign𝐩superscriptsubscript𝑖1𝐽subscript𝑝𝑗|\mathbf{p}|:=\sum_{i=1}^{J}p_{j}| bold_p | := ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, and δj(t,𝐱):=(𝜹(t,𝐱))jassignsubscript𝛿𝑗𝑡𝐱subscript𝜹𝑡𝐱𝑗\delta_{j}(t,\mathbf{x}):=\left(\boldsymbol{\delta}(t,\mathbf{x})\right)_{j}italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , bold_x ) := ( bold_italic_δ ( italic_t , bold_x ) ) start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. In (*), we split the sum, rearrange the terms, divide by ΔtΔ𝑡\Delta troman_Δ italic_t, and collect the terms of 𝒪(Δt)𝒪Δ𝑡\mathcal{O}(\Delta t)caligraphic_O ( roman_Δ italic_t ). The limit for Δt0Δ𝑡0\Delta t\rightarrow 0roman_Δ italic_t → 0 in (A) is denoted by u~(t,𝐱)~𝑢𝑡𝐱\tilde{u}(t,\mathbf{x})over~ start_ARG italic_u end_ARG ( italic_t , bold_x ), leading to (2.3) for 0<t<T0𝑡𝑇0<t<T0 < italic_t < italic_T and 𝐱d𝐱superscript𝑑\mathbf{x}\in\mathbb{N}^{d}bold_x ∈ blackboard_N start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT. ∎

Appendix B Proof for Theorem 3.1

Proof.

We let f:d¯:𝑓superscript¯𝑑f:\mathbb{R}^{\overline{d}}\rightarrow\mathbb{R}italic_f : blackboard_R start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT → blackboard_R be an arbitrary bounded continuous function and 𝑺¯¯𝑺\overline{\boldsymbol{S}}over¯ start_ARG bold_italic_S end_ARG be defined in (3.3). We consider the following weak approximation error:

εT:=𝔼[f(P(𝐗(T)))|𝐗(0)=𝐱0]𝔼[f(𝑺¯(T))|𝑺¯(0)=P(𝐱0)].\displaystyle\varepsilon_{T}:=\mathbb{E}\left[f(P(\mathbf{X}(T)))\middle|% \mathbf{X}(0)=\mathbf{x}_{0}\right]-\mathbb{E}\left[f(\overline{\boldsymbol{S}% }(T))\middle|\overline{\boldsymbol{S}}(0)=P(\mathbf{x}_{0})\right].italic_ε start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT := blackboard_E [ italic_f ( italic_P ( bold_X ( italic_T ) ) ) | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] - blackboard_E [ italic_f ( over¯ start_ARG bold_italic_S end_ARG ( italic_T ) ) | over¯ start_ARG bold_italic_S end_ARG ( 0 ) = italic_P ( bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ] . (B.1)

For t[0,T]𝑡0𝑇t\in[0,T]italic_t ∈ [ 0 , italic_T ], we define the cost to go function as

v¯(t,𝒔):=𝔼[f(𝑺¯(T))|𝑺¯(t)=𝒔].\displaystyle\overline{v}(t,\boldsymbol{s}):=\mathbb{E}\left[f(\overline{% \boldsymbol{S}}(T))\middle|\overline{\boldsymbol{S}}(t)=\boldsymbol{s}\right].over¯ start_ARG italic_v end_ARG ( italic_t , bold_italic_s ) := blackboard_E [ italic_f ( over¯ start_ARG bold_italic_S end_ARG ( italic_T ) ) | over¯ start_ARG bold_italic_S end_ARG ( italic_t ) = bold_italic_s ] .

Then, we can represent the weak error in (B.1) as follows:

εT=𝔼[v¯(T,P(𝐗(T)))|𝐗(0)=𝐱0]v¯(0,P(𝐱0)).\displaystyle\varepsilon_{T}=\mathbb{E}\left[\overline{v}(T,P(\mathbf{X}(T)))% \middle|\mathbf{X}(0)=\mathbf{x}_{0}\right]-\overline{v}(0,P(\mathbf{x}_{0})).italic_ε start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = blackboard_E [ over¯ start_ARG italic_v end_ARG ( italic_T , italic_P ( bold_X ( italic_T ) ) ) | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] - over¯ start_ARG italic_v end_ARG ( 0 , italic_P ( bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ) . (B.2)

Using Dynkin’s formula [30] and (1.3), we can express the first term in (B.1) as follows:

𝔼𝔼\displaystyle\mathbb{E}blackboard_E [v¯(T,P(𝐗(T)))|𝐗(0)=𝐱0]\displaystyle\left[\overline{v}(T,P(\mathbf{X}(T)))\middle|\mathbf{X}(0)=% \mathbf{x}_{0}\right][ over¯ start_ARG italic_v end_ARG ( italic_T , italic_P ( bold_X ( italic_T ) ) ) | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ]
=v¯(0,P(𝐱0))+0T𝔼[tv¯(τ,P(𝐗(τ)))\displaystyle=\overline{v}(0,P(\mathbf{x}_{0}))+\int_{0}^{T}\mathbb{E}\left[% \vphantom{\sum_{j=1}^{J}}\partial_{t}\overline{v}(\tau,P(\mathbf{X}(\tau)))\right.= over¯ start_ARG italic_v end_ARG ( 0 , italic_P ( bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ) + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT blackboard_E [ ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) )
+j=1Jaj(𝐗(τ))(v¯(τ,P(𝐗(τ)+𝝂j))v¯(τ,P(𝐗(τ))))|𝐗(0)=𝐱0]dτ.\displaystyle\left.+\sum_{j=1}^{J}a_{j}(\mathbf{X}(\tau))\left(\overline{v}(% \tau,P(\mathbf{X}(\tau)+\boldsymbol{\nu}_{j}))-\overline{v}(\tau,P(\mathbf{X}(% \tau)))\right)\middle|\mathbf{X}(0)=\mathbf{x}_{0}\right]d\tau.+ ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_τ ) ) ( over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) + bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) - over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) ) | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] italic_d italic_τ .

The Kolmogorov backward equations [30] of 𝑺¯¯𝑺\overline{\boldsymbol{S}}over¯ start_ARG bold_italic_S end_ARG are given as

τv¯(τ,𝒔)=j=1Ja¯j(τ,𝒔)(v¯(τ,𝒔+𝝂¯j)v¯(τ,𝒔)),𝐬d¯,formulae-sequencesubscript𝜏¯𝑣𝜏𝒔superscriptsubscript𝑗1𝐽subscript¯𝑎𝑗𝜏𝒔¯𝑣𝜏𝒔subscript¯𝝂𝑗¯𝑣𝜏𝒔𝐬superscript¯𝑑\displaystyle\partial_{\tau}\overline{v}(\tau,\boldsymbol{s})=-\sum_{j=1}^{J}% \overline{a}_{j}(\tau,\boldsymbol{s})\left(\overline{v}(\tau,\boldsymbol{s}+% \overline{\boldsymbol{\nu}}_{j})-\overline{v}(\tau,\boldsymbol{s})\right),% \mathbf{s}\in\mathbb{N}^{\overline{d}},∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT over¯ start_ARG italic_v end_ARG ( italic_τ , bold_italic_s ) = - ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_τ , bold_italic_s ) ( over¯ start_ARG italic_v end_ARG ( italic_τ , bold_italic_s + over¯ start_ARG bold_italic_ν end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) - over¯ start_ARG italic_v end_ARG ( italic_τ , bold_italic_s ) ) , bold_s ∈ blackboard_N start_POSTSUPERSCRIPT over¯ start_ARG italic_d end_ARG end_POSTSUPERSCRIPT ,

implying that the weak error simplifies to

εTsubscript𝜀𝑇\displaystyle\varepsilon_{T}italic_ε start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT =j=1J0T𝔼[aj(𝐗(τ))v¯(τ,P(𝐗(τ)+𝝂j))a¯j(τ,P(𝐗(τ)))v¯(τ,P(𝐗(τ))+𝝂¯j)|𝐗(0)=𝐱0]\displaystyle=\sum_{j=1}^{J}\int_{0}^{T}\mathbb{E}\left[a_{j}(\mathbf{X}(\tau)% )\overline{v}(\tau,P(\mathbf{X}(\tau)+\boldsymbol{\nu}_{j}))-\overline{a}_{j}(% \tau,P(\mathbf{X}(\tau)))\overline{v}(\tau,P(\mathbf{X}(\tau))+\overline{% \boldsymbol{\nu}}_{j})\middle|\mathbf{X}(0)=\mathbf{x}_{0}\right]= ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT blackboard_E [ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_τ ) ) over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) + bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) - over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) + over¯ start_ARG bold_italic_ν end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ]
𝔼[(aj(𝐗(τ))a¯j(τ,P(𝐗(τ))))v¯(τ,P(𝐗(τ)))|𝐗(0)=𝐱0]dτ.\displaystyle-\mathbb{E}\left[\left(a_{j}(\mathbf{X}(\tau))-\overline{a}_{j}(% \tau,P(\mathbf{X}(\tau)))\right)\overline{v}(\tau,P(\mathbf{X}(\tau)))\middle|% \mathbf{X}(0)=\mathbf{x}_{0}\right]d\tau.- blackboard_E [ ( italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_τ ) ) - over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) ) over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] italic_d italic_τ . (B.3)

Next, we choose a¯jsubscript¯𝑎𝑗\overline{a}_{j}over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and 𝝂¯jsubscript¯𝝂𝑗\overline{\boldsymbol{\nu}}_{j}over¯ start_ARG bold_italic_ν end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT for j=1,,J𝑗1𝐽j=1,\dots,Jitalic_j = 1 , … , italic_J such that εT=0subscript𝜀𝑇0\varepsilon_{T}=0italic_ε start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 0 for any function f𝑓fitalic_f. We consider the second term in (B) and use the tower property to obtain

𝔼[(aj(𝐗(τ))a¯j(τ,P(𝐗(τ))))v¯(τ,P(𝐗(τ)))|𝐗(0)=𝐱0]\displaystyle\mathbb{E}\left[\left(a_{j}(\mathbf{X}(\tau))-\overline{a}_{j}(% \tau,P(\mathbf{X}(\tau)))\right)\overline{v}(\tau,P(\mathbf{X}(\tau)))\middle|% \mathbf{X}(0)=\mathbf{x}_{0}\right]blackboard_E [ ( italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_τ ) ) - over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) ) over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ]
=𝔼[𝔼[(aj(𝐗(τ))a¯j(τ,P(𝐗(τ))))v¯(τ,P(𝐗(τ)))|P(𝐗(τ)),𝐗(0)=𝐱0]|𝐗(0)=𝐱0]\displaystyle=\mathbb{E}\left[\mathbb{E}\left[\left(a_{j}(\mathbf{X}(\tau))-% \overline{a}_{j}(\tau,P(\mathbf{X}(\tau)))\right)\overline{v}(\tau,P(\mathbf{X% }(\tau)))\middle|P(\mathbf{X}(\tau)),\mathbf{X}(0)=\mathbf{x}_{0}\right]% \middle|\mathbf{X}(0)=\mathbf{x}_{0}\right]= blackboard_E [ blackboard_E [ ( italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_τ ) ) - over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) ) over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) | italic_P ( bold_X ( italic_τ ) ) , bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ]
=𝔼[(𝔼[aj(𝐗(τ))|P(𝐗(τ)),𝐗(0)=𝐱0]a¯j(τ,P(𝐗(τ))))v¯(τ,P(𝐗(τ)))|𝐗(0)=𝐱0].\displaystyle=\mathbb{E}\left[\left(\mathbb{E}\left[{a}_{j}(\mathbf{X}(\tau))% \middle|P(\mathbf{X}(\tau)),\mathbf{X}(0)=\mathbf{x}_{0}\right]-\overline{a}_{% j}(\tau,P(\mathbf{X}(\tau)))\right)\overline{v}(\tau,P(\mathbf{X}(\tau)))% \middle|\mathbf{X}(0)=\mathbf{x}_{0}\right].= blackboard_E [ ( blackboard_E [ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_τ ) ) | italic_P ( bold_X ( italic_τ ) ) , bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] - over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) ) over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] . (B.4)

To ensure that (B)=0absent0=0= 0 for any function f𝑓fitalic_f, we obtain the following:

a¯j(τ,P(𝐗(τ)))=𝔼[aj(𝐗(τ))|P(𝐗(τ)),𝐗(0)=𝐱0],j=1,,J.\displaystyle\overline{a}_{j}(\tau,P(\mathbf{X}(\tau)))=\mathbb{E}\left[{a}_{j% }(\mathbf{X}(\tau))\middle|P(\mathbf{X}(\tau)),\mathbf{X}(0)=\mathbf{x}_{0}% \right],j=1,\dots,J.over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) = blackboard_E [ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_τ ) ) | italic_P ( bold_X ( italic_τ ) ) , bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] , italic_j = 1 , … , italic_J . (B.5)

Applying (B.5) and the tower property for the first term, we derive

𝔼[aj(𝐗(τ))v¯(τ,P(𝐗(τ)+𝝂j))a¯j(τ,P(𝐗(τ)))v¯(τ,P(𝐗(τ))+𝝂¯j)|𝐗(0)=𝐱0]\displaystyle\mathbb{E}\left[a_{j}(\mathbf{X}(\tau))\overline{v}(\tau,P(% \mathbf{X}(\tau)+\boldsymbol{\nu}_{j}))-\overline{a}_{j}(\tau,P(\mathbf{X}(% \tau)))\overline{v}(\tau,P(\mathbf{X}(\tau))+\overline{\boldsymbol{\nu}}_{j})% \middle|\mathbf{X}(0)=\mathbf{x}_{0}\right]blackboard_E [ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_τ ) ) over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) + bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) - over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) + over¯ start_ARG bold_italic_ν end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ]
=𝔼[𝔼[aj(𝐗(τ))v¯(τ,P(𝐗(τ)+𝝂j))\displaystyle=\mathbb{E}\left[\mathbb{E}\left[a_{j}(\mathbf{X}(\tau))\overline% {v}(\tau,P(\mathbf{X}(\tau)+\boldsymbol{\nu}_{j}))\right.\right.= blackboard_E [ blackboard_E [ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_τ ) ) over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) + bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) )
a¯j(τ,P(𝐗(τ)))v¯(τ,P(𝐗(τ))+𝝂¯j)|P(𝐗(τ)),𝐗(0)=𝐱0]|𝐗(0)=𝐱0]\displaystyle\left.\left.-\overline{a}_{j}(\tau,P(\mathbf{X}(\tau)))\overline{% v}(\tau,P(\mathbf{X}(\tau))+\overline{\boldsymbol{\nu}}_{j})\middle|P(\mathbf{% X}(\tau)),\mathbf{X}(0)=\mathbf{x}_{0}\right]\middle|\mathbf{X}(0)=\mathbf{x}_% {0}\right]- over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) + over¯ start_ARG bold_italic_ν end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) | italic_P ( bold_X ( italic_τ ) ) , bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ]
=𝔼[𝔼[aj(𝐗(τ))|P(𝐗(τ)),𝐗(0)=𝐱0]v¯(τ,P(𝐗(τ))+P(𝝂j)))\displaystyle=\mathbb{E}\left[\mathbb{E}\left[a_{j}(\mathbf{X}(\tau))\middle|P% (\mathbf{X}(\tau)),\mathbf{X}(0)=\mathbf{x}_{0}\right]\overline{v}(\tau,P(% \mathbf{X}(\tau))+P(\boldsymbol{\nu}_{j})))\right.= blackboard_E [ blackboard_E [ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_τ ) ) | italic_P ( bold_X ( italic_τ ) ) , bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) + italic_P ( bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) )
a¯j(τ,P(𝐗(τ)))v¯(τ,P(𝐗(τ))+𝝂¯j)|𝐗(0)=𝐱0]\displaystyle\left.-\overline{a}_{j}(\tau,P(\mathbf{X}(\tau)))\overline{v}(% \tau,P(\mathbf{X}(\tau))+\overline{\boldsymbol{\nu}}_{j})\middle|\mathbf{X}(0)% =\mathbf{x}_{0}\right]- over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_τ , italic_P ( bold_X ( italic_τ ) ) ) over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) + over¯ start_ARG bold_italic_ν end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ]
=𝔼[𝔼[aj(𝐗(τ))|P(𝐗(τ)),𝐗(0)=𝐱0]\displaystyle=\mathbb{E}\left[\vphantom{\underbrace{\overline{v}}_{\overline{% \boldsymbol{\nu}}_{j}}}\mathbb{E}\left[a_{j}(\mathbf{X}(\tau))\middle|P(% \mathbf{X}(\tau)),\mathbf{X}(0)=\mathbf{x}_{0}\right]\right.= blackboard_E [ blackboard_E [ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_X ( italic_τ ) ) | italic_P ( bold_X ( italic_τ ) ) , bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ]
(v¯(τ,P(𝐗(τ))+P(𝝂j))v¯(τ,P(𝐗(τ))+𝝂¯j))|𝐗(0)=𝐱0].\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\cdot\left.\left(\overline{v}(\tau,% P(\mathbf{X}(\tau))+P(\boldsymbol{\nu}_{j}))-\overline{v}(\tau,P(\mathbf{X}(% \tau))+\overline{\boldsymbol{\nu}}_{j})\right)\middle|\mathbf{X}(0)=\mathbf{x}% _{0}\right].⋅ ( over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) + italic_P ( bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) - over¯ start_ARG italic_v end_ARG ( italic_τ , italic_P ( bold_X ( italic_τ ) ) + over¯ start_ARG bold_italic_ν end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ) | bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] . (B.6)

Moreover, Equation (B) becomes zero for any function f𝑓fitalic_f using

𝝂¯j=P(𝝂j),j=1,,J.formulae-sequencesubscript¯𝝂𝑗𝑃subscript𝝂𝑗𝑗1𝐽\displaystyle\overline{\boldsymbol{\nu}}_{j}=P(\boldsymbol{\nu}_{j}),j=1,\dots% ,J.over¯ start_ARG bold_italic_ν end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_P ( bold_italic_ν start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) , italic_j = 1 , … , italic_J . (B.7)

With this choice for a¯jsubscript¯𝑎𝑗\overline{a}_{j}over¯ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and 𝝂¯jsubscript¯𝝂𝑗\overline{\boldsymbol{\nu}}_{j}over¯ start_ARG bold_italic_ν end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, we derive εT=0subscript𝜀𝑇0\varepsilon_{T}=0italic_ε start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 0. The derivation holds for arbitrary bounded and smooth functions f𝑓fitalic_f, for all fixed times T𝑇Titalic_T; thus, the process 𝑺(t)=P(𝐗(t))𝑺𝑡𝑃𝐗𝑡\boldsymbol{S}(t)=P(\mathbf{X}(t))bold_italic_S ( italic_t ) = italic_P ( bold_X ( italic_t ) ) has the same conditional distribution as 𝑺¯(t)¯𝑺𝑡\overline{\boldsymbol{S}}(t)over¯ start_ARG bold_italic_S end_ARG ( italic_t ) conditioned on the initial value 𝐗(0)=𝐱0𝐗0subscript𝐱0\mathbf{X}(0)=\mathbf{x}_{0}bold_X ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

Appendix C Markovian Projection Cost Derivation

We present details on the computational cost of MP, as provided in (3.10):

  • The number of operations to generate one TL paths is given by

    WTL(Δt)=TΔt(Cprop+JCPoi+d(J+2)),subscript𝑊𝑇𝐿Δ𝑡𝑇Δ𝑡subscript𝐶𝑝𝑟𝑜𝑝𝐽subscript𝐶𝑃𝑜𝑖𝑑𝐽2\displaystyle W_{TL}(\Delta t)=\frac{T}{\Delta t}\cdot(C_{prop}+J\cdot C_{Poi}% +d(J+2)),italic_W start_POSTSUBSCRIPT italic_T italic_L end_POSTSUBSCRIPT ( roman_Δ italic_t ) = divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ ( italic_C start_POSTSUBSCRIPT italic_p italic_r italic_o italic_p end_POSTSUBSCRIPT + italic_J ⋅ italic_C start_POSTSUBSCRIPT italic_P italic_o italic_i end_POSTSUBSCRIPT + italic_d ( italic_J + 2 ) ) , (C.1)

    where Cpropsubscript𝐶𝑝𝑟𝑜𝑝C_{prop}italic_C start_POSTSUBSCRIPT italic_p italic_r italic_o italic_p end_POSTSUBSCRIPT is the cost of one evaluation of the propensity function (1.4). The dominant cost in (C.1) is CPoisubscript𝐶𝑃𝑜𝑖C_{Poi}italic_C start_POSTSUBSCRIPT italic_P italic_o italic_i end_POSTSUBSCRIPT (the cost of generating a Poisson random variable).

  • The number of operations for the Gram–Schmidt algorithm, as described in Remark 3.3, is given by

    WGS(#Λ,Δt,M)=#Λ(Cinner+#Λ+1)+(#Λ1)#Λ2(2#Λ+Cinner),subscript𝑊𝐺𝑆#ΛΔ𝑡𝑀#Λsubscript𝐶𝑖𝑛𝑛𝑒𝑟#Λ1#Λ1#Λ22#Λsubscript𝐶𝑖𝑛𝑛𝑒𝑟\displaystyle W_{G-S}(\#\Lambda,\Delta t,M)=\#\Lambda\cdot(C_{inner}+\#\Lambda% +1)+\frac{(\#\Lambda-1)\#\Lambda}{2}(2\#\Lambda+C_{inner}),italic_W start_POSTSUBSCRIPT italic_G - italic_S end_POSTSUBSCRIPT ( # roman_Λ , roman_Δ italic_t , italic_M ) = # roman_Λ ⋅ ( italic_C start_POSTSUBSCRIPT italic_i italic_n italic_n italic_e italic_r end_POSTSUBSCRIPT + # roman_Λ + 1 ) + divide start_ARG ( # roman_Λ - 1 ) # roman_Λ end_ARG start_ARG 2 end_ARG ( 2 # roman_Λ + italic_C start_POSTSUBSCRIPT italic_i italic_n italic_n italic_e italic_r end_POSTSUBSCRIPT ) , (C.2)

    where Cinnersubscript𝐶𝑖𝑛𝑛𝑒𝑟C_{inner}italic_C start_POSTSUBSCRIPT italic_i italic_n italic_n italic_e italic_r end_POSTSUBSCRIPT is the cost of the evaluation of the empirical inner product (3.9) given by

    Cinner=TΔtM(2+2Cpol)+3=𝒪(TΔtM#Λ).subscript𝐶𝑖𝑛𝑛𝑒𝑟𝑇Δ𝑡𝑀22subscript𝐶𝑝𝑜𝑙3𝒪𝑇Δ𝑡𝑀#Λ\displaystyle C_{inner}=\frac{T}{\Delta t}\cdot M(2+2C_{pol})+3=\mathcal{O}(% \frac{T}{\Delta t}\cdot M\cdot\#\Lambda).italic_C start_POSTSUBSCRIPT italic_i italic_n italic_n italic_e italic_r end_POSTSUBSCRIPT = divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ italic_M ( 2 + 2 italic_C start_POSTSUBSCRIPT italic_p italic_o italic_l end_POSTSUBSCRIPT ) + 3 = caligraphic_O ( divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ italic_M ⋅ # roman_Λ ) . (C.3)

    The cost Cpolsubscript𝐶𝑝𝑜𝑙C_{pol}italic_C start_POSTSUBSCRIPT italic_p italic_o italic_l end_POSTSUBSCRIPT in (C.3) is the computational cost for one evaluation of a polynomial in the space <ϕp>pΛsubscriptexpectationsubscriptitalic-ϕ𝑝𝑝Λ<\phi_{p}>_{p\in\Lambda}< italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT > start_POSTSUBSCRIPT italic_p ∈ roman_Λ end_POSTSUBSCRIPT, which is 𝒪(#Λ)𝒪#Λ\mathcal{O}(\#\Lambda)caligraphic_O ( # roman_Λ ).

    In the simulations, we apply the setting #ΛTΔtMmuch-less-than#Λ𝑇Δ𝑡𝑀\#\Lambda\ll\frac{T}{\Delta t}\cdot M# roman_Λ ≪ divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ italic_M (see Section 5.2, using the parameter #Λ=9#Λ9\#\Lambda=9# roman_Λ = 9, M=104𝑀superscript104M=10^{4}italic_M = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT, TΔt=24𝑇Δ𝑡superscript24\frac{T}{\Delta t}=2^{4}divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG = 2 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT). Therefore, the dominant cost in (C.2) is 𝒪(MTΔt(#Λ)3)𝒪𝑀𝑇Δ𝑡superscript#Λ3\mathcal{O}(M\cdot\frac{T}{\Delta t}\cdot\left(\#\Lambda\right)^{3})caligraphic_O ( italic_M ⋅ divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ ( # roman_Λ ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ).

  • The cost WL2(#Λ,Δt,M)subscript𝑊superscript𝐿2#ΛΔ𝑡𝑀W_{L^{2}}(\#\Lambda,\Delta t,M)italic_W start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( # roman_Λ , roman_Δ italic_t , italic_M ) is split into two: the cost to (i) derive and (ii) solve the normal equation (3.8). The number of operations to derive the design matrix D𝐷Ditalic_D is MTΔt#ΛCpol𝑀𝑇Δ𝑡#Λsubscript𝐶𝑝𝑜𝑙M\cdot\frac{T}{\Delta t}\cdot\#\Lambda\cdot C_{pol}italic_M ⋅ divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ # roman_Λ ⋅ italic_C start_POSTSUBSCRIPT italic_p italic_o italic_l end_POSTSUBSCRIPT, and the cost to derive one right-hand side (Ψ(j))j𝒥MPsubscriptsuperscriptΨ𝑗𝑗subscript𝒥𝑀𝑃(\Psi^{(j)})_{j\in\mathcal{J}_{MP}}( roman_Ψ start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_j ∈ caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT end_POSTSUBSCRIPT is MTΔtCprop𝑀𝑇Δ𝑡subscript𝐶𝑝𝑟𝑜𝑝M\cdot\frac{T}{\Delta t}\cdot C_{prop}italic_M ⋅ divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ italic_C start_POSTSUBSCRIPT italic_p italic_r italic_o italic_p end_POSTSUBSCRIPT. In (3.8), the cost for the matrix product DDsuperscript𝐷top𝐷D^{\top}Ditalic_D start_POSTSUPERSCRIPT ⊤ end_POSTSUPERSCRIPT italic_D is 𝒪(#Λ2MTΔt)𝒪#superscriptΛ2𝑀𝑇Δ𝑡\mathcal{O}(\#\Lambda^{2}\cdot M\cdot\frac{T}{\Delta t})caligraphic_O ( # roman_Λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⋅ italic_M ⋅ divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ), and the cost for #𝒥MP#subscript𝒥𝑀𝑃\#\mathcal{J}_{MP}# caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT matrix-vector products is 𝒪(#𝒥MP#ΛMTΔt)𝒪#subscript𝒥𝑀𝑃#Λ𝑀𝑇Δ𝑡\mathcal{O}(\#\mathcal{J}_{MP}\cdot\#\Lambda\cdot M\cdot\frac{T}{\Delta t})caligraphic_O ( # caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ⋅ # roman_Λ ⋅ italic_M ⋅ divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ). Finally, solving (3.8) costs 𝒪(#𝒥MP#Λ3)𝒪#subscript𝒥𝑀𝑃#superscriptΛ3\mathcal{O}(\#\mathcal{J}_{MP}\cdot\#\Lambda^{3})caligraphic_O ( # caligraphic_J start_POSTSUBSCRIPT italic_M italic_P end_POSTSUBSCRIPT ⋅ # roman_Λ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ), which is a nondominant term under the given setting, #ΛTΔtMmuch-less-than#Λ𝑇Δ𝑡𝑀\#\Lambda\ll\frac{T}{\Delta t}\cdot M# roman_Λ ≪ divide start_ARG italic_T end_ARG start_ARG roman_Δ italic_t end_ARG ⋅ italic_M.