Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
\addbibresource

sample.bib

Efficient gPC-based quantification of probabilistic robustness
for systems in neuroscience

Uros Sutulovic, Daniele Proverbio, Rami Katz, and Giulia Giordano This work was funded by the European Union through the ERC INSPIRE grant (project number 101076926). Views and opinions expressed are however those of the authors only and do not necessarily reflect those of the European Union or the European Research Council. Neither the European Union nor the granting authority can be held responsible for them.The authors are with the Department of Industrial Engineering, University of Trento, Italy. uros.sutulovic@unitn.it; daniele.proverbio@unitn.it; ramkatsee@gmail.com; giulia.giordano@unitn.it
Abstract

We introduce and analyze generalised polynomial chaos (gPC), considering both intrusive and non-intrusive approaches, as an uncertainty quantification method in studies of probabilistic robustness. The considered gPC methods are complementary to Monte Carlo (MC) methods and are shown to be fast and scalable, allowing for comprehensive and efficient exploration of parameter spaces. These properties enable robustness analysis of a wider set of models, compared to computationally expensive MC methods, while retaining desired levels of accuracy. We discuss the application of gPC methods to systems in biology and neuroscience, notably subject to multiple parametric uncertainties, and we examine a well-known model of neural dynamics as a case study.

I Introduction and Motivation

The flourishing field of systems biology [alon2019introduction] employs methods from mathematics, physics and engineering to quantitatively understand, predict and control biological systems at all scales. Of great relevance is the study of robustness [Kitano2007towards]. Natural systems showcase an impressive complexity; yet, they manage to thrive despite uncertainties, variability, or external perturbations. Examples include cell homeostasis [Aoki2019], multicellular coordination [proverbio2020assessing], and robust neural modulation [goldman2001global]. Systematically guaranteeing that a property of interest is preserved regardless of parameter values, or for all parameter values within a certain set, is the scope of structural analysis [blanchini2021structural] and robustness analysis [Barmish1994]. However, sometimes a property of interest does not hold structurally, nor robustly, but just with high probability.

Probabilistic robustness has been fruitfully introduced and investigated in engineering [tempo2013randomized], to characterise the likelihood of a property to hold, given the probability distribution of model parameters. The most commonly used method for probabilistic analysis is Monte Carlo (MC), running many simulations with sampled random variables or random processes. However, MC methods suffer from poor scalability and require extensive computational power and time to address even relatively simple models. Developing flexible and efficient alternative methods to quantify probabilistic robustness would scale up the investigation of parametric uncertainties in complex models, and would be particularly relevant for biological systems, characterised by many uncertain and hardly controllable parameters.

To address these challenges, we present and discuss the use of generalised polynomial chaos (gPC), which is a spectral method to approximate the solution to stochastic differential equations with random parametric uncertainties. First developed in the realm of uncertainty quantification [pepper2019multiscale], gPC also found applications in other fields, such as model predictive control for stochastic systems [kim2013generalised], to replace or accelerate MC methods. Analysing uncertain systems by means of gPC is computationally efficient [fisher2011optimal], since no sampling is required and the investigation of large parameter spaces can be drastically accelerated.

Here, we systematically analyse the performance of gPC methods in providing efficient surrogate models for polynomial systems, with a specific focus on models from systems biology and neuroscience. We first discuss the theoretical background and variants of gPC methods, to provide a comprehensive set of guidelines towards systematic applications. We then assess the advantages and performance of gPC methods with respect to Monte Carlo approaches in terms of computing efficiency and accuracy, using the Hindmarsh-Rose (HR) model as a case study from neuroscience. Finally, we employ gPC to assess the persistence of signalling regimes in the HR model, subject to parametric uncertainty.

II Problem Formulation and Background

We consider autonomous ODE systems of the form

𝐱˙=f⁒(𝐱,Z),˙𝐱𝑓𝐱𝑍\dot{\mathbf{x}}=f(\mathbf{x},Z),overΛ™ start_ARG bold_x end_ARG = italic_f ( bold_x , italic_Z ) , (1)

where π±βˆˆβ„n𝐱superscriptℝ𝑛\mathbf{x}\in\mathbb{R}^{n}bold_x ∈ blackboard_R start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT is the state of the system and Zβˆˆβ„d𝑍superscriptℝ𝑑Z\in\mathbb{R}^{d}italic_Z ∈ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT is a parametric uncertainty. Specifically, we assume that Z:Ξ©β†’IZβŠ‚β„d:𝑍→Ωsubscript𝐼𝑍superscriptℝ𝑑Z\colon\Omega\to I_{Z}\subset\mathbb{R}^{d}italic_Z : roman_Ξ© β†’ italic_I start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT βŠ‚ blackboard_R start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT is a vector of mutually-independent random variables defined on some probability space (Ξ©,β„±,β„™)Ξ©β„±β„™(\Omega,\mathcal{F},\mathbb{P})( roman_Ξ© , caligraphic_F , blackboard_P ), where ΩΩ\Omegaroman_Ξ© denotes a sample space containing samples Ο‰βˆˆΞ©πœ”Ξ©\omega\in\Omegaitalic_Ο‰ ∈ roman_Ξ©, β„±βŠ‚2Ξ©β„±superscript2Ξ©\mathcal{F}\subset 2^{\Omega}caligraphic_F βŠ‚ 2 start_POSTSUPERSCRIPT roman_Ξ© end_POSTSUPERSCRIPT is a ΟƒπœŽ\sigmaitalic_Οƒ-algebra and β„™β„™\mathbb{P}blackboard_P is a probability measure. We assume that Z𝑍Zitalic_Z is absolutely continuous, and denote its probability density function (PDF) by ρZ⁒(z)subscriptπœŒπ‘π‘§\rho_{Z}(z)italic_ρ start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT ( italic_z ).

Remark 1

If the random variables Z𝑍Zitalic_Z are not mutually independent, the Rosenblatt transformation [rosenblatt1952remarks] can be applied to transform them into new, mutually independent random variables Z~~𝑍\tilde{Z}over~ start_ARG italic_Z end_ARG. Hence, from a theoretical perspective, assuming mutual independence poses no loss of generality. β‹„β‹„\hfill\diamondβ‹„

Due to the parametric uncertainty, the state in (1) is a stochastic process 𝐱⁒(t;Z)𝐱𝑑𝑍\mathbf{x}(t;Z)bold_x ( italic_t ; italic_Z ). Characterizing the first moments (summary statistics) of 𝐱⁒(t;Z)𝐱𝑑𝑍\mathbf{x}(t;Z)bold_x ( italic_t ; italic_Z ), for example in order to approximate its characteristic function [florescu2013handbook], is of interest.

The estimation of summary statistics is often performed by a combination of explicit formulas and randomised algorithms. Monte Carlo methods are widely used in computational studies, to perform random sampling for numerical simulations and derive statistics post-hoc. Yet, they may be impractical when simulating large models with uncertain parameters: their sample complexity scales poorly with model dimensionality, and is sensitive to the desired accuracy. A lower bound for the number NM⁒Csubscript𝑁𝑀𝐢N_{MC}italic_N start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT of Monte Carlo samples that guarantee a certain level of accuracy Ξ΅πœ€\varepsilonitalic_Ξ΅ and confidence δ𝛿\deltaitalic_Ξ΄ can be estimated with the Chernoff bound [tempo2013randomized]:

NM⁒Cβ‰₯12⁒Ρ2⁒ln⁑2Ξ΄.subscript𝑁𝑀𝐢12superscriptπœ€22𝛿N_{MC}\geq\frac{1}{2\varepsilon^{2}}\ln\frac{2}{\delta}\,.italic_N start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT β‰₯ divide start_ARG 1 end_ARG start_ARG 2 italic_Ξ΅ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_ln divide start_ARG 2 end_ARG start_ARG italic_Ξ΄ end_ARG . (2)

For instance, NM⁒C=2.65β‹…104subscript𝑁𝑀𝐢⋅2.65superscript104N_{MC}=2.65\cdot 10^{4}italic_N start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT = 2.65 β‹… 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT for performance verification with accuracy Ξ΅=10βˆ’2πœ€superscript102\varepsilon=10^{-2}italic_Ξ΅ = 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT and confidence Ξ΄=10βˆ’2𝛿superscript102\delta=10^{-2}italic_Ξ΄ = 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT.

To overcome such complexity limitations, alternative surrogate models need to be identified that are computationally more affordable and allow to retrieve summary statistics a priori. For practical applications, it is also necessary to determine how well such surrogate models perform in uncertainty quantification and robustness analysis tasks. Here we propose the use of gPC methods, in different variants, to achieve summary statistics estimation.

II-A Generalised polynomial chaos

Considering the probability density function of Z𝑍Zitalic_Z in (1), we restrict our attention to the set of functions ψ:IZ→ℝ:πœ“β†’subscript𝐼𝑍ℝ\psi\colon I_{Z}\rightarrow\mathbb{R}italic_ψ : italic_I start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT β†’ blackboard_R that belong to the weighted L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT space

LρZ2⁒(IZ)={ψ:𝔼⁒[ψ2]=∫IZψ2⁒(z)⁒ρZ⁒(z)⁒𝑑z<∞}.subscriptsuperscript𝐿2subscriptπœŒπ‘subscript𝐼𝑍conditional-setπœ“π”Όdelimited-[]superscriptπœ“2subscriptsubscript𝐼𝑍superscriptπœ“2𝑧subscriptπœŒπ‘π‘§differential-d𝑧L^{2}_{\rho_{Z}}(I_{Z})=\left\{\psi\,\colon\,\mathbb{E}[\psi^{2}]=\int_{I_{Z}}% \psi^{2}(z)\rho_{Z}(z)dz<\infty\right\}.italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_I start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT ) = { italic_ψ : blackboard_E [ italic_ψ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] = ∫ start_POSTSUBSCRIPT italic_I start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ψ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_z ) italic_ρ start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT ( italic_z ) italic_d italic_z < ∞ } . (3)

A natural basis for the space LρZ2⁒(IZ)subscriptsuperscript𝐿2subscriptπœŒπ‘subscript𝐼𝑍L^{2}_{\rho_{Z}}(I_{Z})italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_I start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT ) is a set of multivariate polynomials {Φα⁒(z)}Ξ±subscriptsubscriptΦ𝛼𝑧𝛼\{\Phi_{\mathbf{\alpha}}(z)\}_{\alpha}{ roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_z ) } start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT satisfying the orthogonality relation

𝔼⁒[Φα⁒Φβ]=∫IZΦα⁒(z)⁒Φβ⁒(z)⁒ρZ⁒(z)⁒𝑑z=γα⁒δα⁒β,𝔼delimited-[]subscriptΦ𝛼subscriptΦ𝛽subscriptsubscript𝐼𝑍subscriptΦ𝛼𝑧subscriptΦ𝛽𝑧subscriptπœŒπ‘π‘§differential-d𝑧subscript𝛾𝛼subscript𝛿𝛼𝛽\mathbb{E}[{\Phi_{\mathbf{\alpha}}\Phi_{\mathbf{\beta}}}]=\int_{I_{Z}}\Phi_{% \mathbf{\alpha}}(z)\Phi_{\mathbf{\beta}}(z)\rho_{Z}(z)dz=\gamma_{\mathbf{% \alpha}}\delta_{\mathbf{\alpha}\mathbf{\beta}}\,,blackboard_E [ roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ² end_POSTSUBSCRIPT ] = ∫ start_POSTSUBSCRIPT italic_I start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_z ) roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ² end_POSTSUBSCRIPT ( italic_z ) italic_ρ start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT ( italic_z ) italic_d italic_z = italic_Ξ³ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT italic_Ξ΄ start_POSTSUBSCRIPT italic_Ξ± italic_Ξ² end_POSTSUBSCRIPT , (4)

where Ξ³Ξ±=𝔼⁒[Φα2]>0subscript𝛾𝛼𝔼delimited-[]superscriptsubscriptΦ𝛼20\gamma_{\mathbf{\alpha}}=\mathbb{E}[\Phi_{\mathbf{\alpha}}^{2}]>0italic_Ξ³ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT = blackboard_E [ roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] > 0 is a normalizing factor, and δα⁒βsubscript𝛿𝛼𝛽\delta_{\mathbf{\alpha}\mathbf{\beta}}italic_Ξ΄ start_POSTSUBSCRIPT italic_Ξ± italic_Ξ² end_POSTSUBSCRIPT is the d𝑑ditalic_d-variate Kronecker delta. Here, Ξ±=(Ξ±1,…,Ξ±d)βˆˆβ„•0d𝛼subscript𝛼1…subscript𝛼𝑑superscriptsubscriptβ„•0𝑑\mathbf{\alpha}=(\alpha_{1},\ldots,\alpha_{d})\in\mathbb{N}_{0}^{d}italic_Ξ± = ( italic_Ξ± start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_Ξ± start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT is a multi-index. The basis {Φα⁒(z)}Ξ±subscriptsubscriptΦ𝛼𝑧𝛼\{\Phi_{\mathbf{\alpha}}(z)\}_{\alpha}{ roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_z ) } start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT can be obtained from a Gram-Schmidt orthogonalization process applied to the set of monomials {∏i=1dziΞ±i}Ξ±βˆˆβ„•0dsubscriptsuperscriptsubscriptproduct𝑖1𝑑superscriptsubscript𝑧𝑖subscript𝛼𝑖𝛼superscriptsubscriptβ„•0𝑑\{\prod_{i=1}^{d}z_{i}^{\alpha_{i}}\}_{\alpha\in\mathbb{N}_{0}^{d}}{ ∏ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_Ξ± start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT } start_POSTSUBSCRIPT italic_Ξ± ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT end_POSTSUBSCRIPT. Henceforth, we assume that Ξ³Ξ±=1subscript𝛾𝛼1\gamma_{\alpha}=1italic_Ξ³ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT = 1 for all Ξ±βˆˆβ„•0d𝛼superscriptsubscriptβ„•0𝑑\alpha\in\mathbb{N}_{0}^{d}italic_Ξ± ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, meaning that the polynomials {Φα⁒(z)}Ξ±subscriptsubscriptΦ𝛼𝑧𝛼\left\{\Phi_{\alpha}(z)\right\}_{\alpha}{ roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_z ) } start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT are orthonormal.

Wiener [wiener1938homogeneous] proved a PC expansion for a normally distributed Z∼N⁒(𝟎,I)similar-to𝑍𝑁0𝐼Z\sim N(\mathbf{0},I)italic_Z ∼ italic_N ( bold_0 , italic_I ):

ψ⁒(Z)=βˆ‘|Ξ±|=0∞ψ^α⁒Φα⁒(Z),ψ^Ξ±=𝔼⁒[ψ⁒Φα],formulae-sequenceπœ“π‘superscriptsubscript𝛼0subscript^πœ“π›ΌsubscriptΦ𝛼𝑍subscript^πœ“π›Όπ”Όdelimited-[]πœ“subscriptΦ𝛼\psi(Z)=\sum_{|\mathbf{\alpha}|=0}^{\infty}\hat{\psi}_{\mathbf{\alpha}}\Phi_{% \mathbf{\alpha}}(Z)\,,\quad\hat{\psi}_{\mathbf{\alpha}}=\mathbb{E}[\psi\Phi_{% \mathbf{\alpha}}],italic_ψ ( italic_Z ) = βˆ‘ start_POSTSUBSCRIPT | italic_Ξ± | = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_Z ) , over^ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT = blackboard_E [ italic_ψ roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ] , (5)

where {Φα⁒(z)}Ξ±subscriptsubscriptΦ𝛼𝑧𝛼\{\Phi_{\mathbf{\alpha}}(z)\}_{\alpha}{ roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_z ) } start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT is a base of Hermite polynomials and |Ξ±|=βˆ‘i=1dΞ±i𝛼superscriptsubscript𝑖1𝑑subscript𝛼𝑖|\mathbf{\alpha}|=\sum_{i=1}^{d}\alpha_{i}| italic_Ξ± | = βˆ‘ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_Ξ± start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Later, Cameron and Martin [cameron1947orthogonal] generalized this expansion to random variables Z𝑍Zitalic_Z with an arbitrary distribution. Xiu and Karniadakis [xiu2002wiener] proposed a framework that links standard random variables to the orthogonal polynomials of the Wiener-Askey table (Table I). Constructing the basis {Φα⁒(z)}Ξ±subscriptsubscriptΦ𝛼𝑧𝛼\{\Phi_{\mathbf{\alpha}}(z)\}_{\alpha}{ roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_z ) } start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT in case of arbitrarily distributed random variables is the subject of dedicated studies; for instance, we refer to the procedures described in [wan2006multi].

TABLE I: Wiener-Askey scheme
PDF of Z𝑍Zitalic_Z Basis {Φα}Ξ±subscriptsubscriptΦ𝛼𝛼\{\Phi_{\mathbf{\alpha}}\}_{\alpha}{ roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT
Gaussian Hermite polynomials
Uniform Legendre polynomials
Beta Jacobi polynomials
Gamma Laguerre polynomials

For the stochastic process 𝐱=𝐱⁒(t;Z)𝐱𝐱𝑑𝑍\mathbf{x}=\mathbf{x}(t;Z)bold_x = bold_x ( italic_t ; italic_Z ), subject to the dynamics in (1), the gPC expansion reads

𝐱⁒(t;Z)=βˆ‘Ξ±βˆˆβ„•0d𝐱^α⁒(t)⁒Φα⁒(Z),𝐱^α⁒(t)=[𝐱^Ξ±,1⁒(t)…𝐱^Ξ±,n⁒(t)]βŠ€βˆˆβ„n,𝐱^Ξ±,j⁒(t):=𝔼⁒[𝐱j⁒(t,β‹…)⁒Φα],j=1,…,n.missing-subexpression𝐱𝑑𝑍subscript𝛼superscriptsubscriptβ„•0𝑑subscript^𝐱𝛼𝑑subscriptΦ𝛼𝑍missing-subexpressionmissing-subexpressionsubscript^𝐱𝛼𝑑superscriptmatrixsubscript^𝐱𝛼1𝑑…subscript^𝐱𝛼𝑛𝑑topsuperscriptℝ𝑛missing-subexpressionmissing-subexpressionformulae-sequenceassignsubscript^𝐱𝛼𝑗𝑑𝔼delimited-[]subscript𝐱𝑗𝑑⋅subscriptΦ𝛼𝑗1…𝑛missing-subexpression\begin{array}[]{lll}&\mathbf{x}(t;Z)=\sum_{\alpha\in\mathbb{N}_{0}^{d}}\hat{% \mathbf{x}}_{\mathbf{\alpha}}(t)\Phi_{\mathbf{\alpha}}(Z)\,,\\ &\hat{\mathbf{x}}_{\mathbf{\alpha}}(t)=\begin{bmatrix}\hat{\mathbf{x}}_{\alpha% ,1}(t)&\dots&\hat{\mathbf{x}}_{\alpha,n}(t)\end{bmatrix}^{\top}\in\mathbb{R}^{% n},\\ &\hat{\mathbf{x}}_{\alpha,j}(t):=\mathbb{E}[\mathbf{x}_{j}(t,\cdot)\Phi_{% \mathbf{\alpha}}],\ j=1,\dots,n.\end{array}start_ARRAY start_ROW start_CELL end_CELL start_CELL bold_x ( italic_t ; italic_Z ) = βˆ‘ start_POSTSUBSCRIPT italic_Ξ± ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_t ) roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_Z ) , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_t ) = [ start_ARG start_ROW start_CELL over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± , 1 end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL … end_CELL start_CELL over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± , italic_n end_POSTSUBSCRIPT ( italic_t ) end_CELL end_ROW end_ARG ] start_POSTSUPERSCRIPT ⊀ end_POSTSUPERSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± , italic_j end_POSTSUBSCRIPT ( italic_t ) := blackboard_E [ bold_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t , β‹… ) roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ] , italic_j = 1 , … , italic_n . end_CELL start_CELL end_CELL end_ROW end_ARRAY (6)

In a nutshell, gPC methods consists of representing the stochastic process 𝐱⁒(t;Z)𝐱𝑑𝑍\mathbf{x}(t;Z)bold_x ( italic_t ; italic_Z ) as a series expansion with respect to an appropriate basis of orthogonal polynomials depending on the uncertain parameters Z𝑍Zitalic_Z, with coefficients depending on time t𝑑titalic_t, as in equation (6). The spectral coefficients {𝐱^α⁒(t)}Ξ±subscriptsubscript^𝐱𝛼𝑑𝛼\left\{\hat{\mathbf{x}}_{\mathbf{\alpha}}(t)\right\}_{\alpha}{ over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_t ) } start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT contain all the temporal information, while the stochasticity is confined in the orthogonal polynomials {Φα⁒(Z)}Ξ±subscriptsubscriptΦ𝛼𝑍𝛼\left\{\Phi_{\mathbf{\alpha}}(Z)\right\}_{\alpha}{ roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_Z ) } start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT. The relation (4) and the linearity of the spectral representation (6) can thus be used to compute the summary statistics of 𝐱⁒(t;Z)𝐱𝑑𝑍\mathbf{x}(t;Z)bold_x ( italic_t ; italic_Z ). For arbitrary statistical moments, we refer the reader to [lefebvre2020moment]. In this work we only consider mean and variance, extracted as

{μ𝐱⁒(t)=𝐱^𝟎⁒(t)σ𝐱2⁒(t)=βˆ‘|Ξ±|β‰ 0𝐱^Ξ±2⁒(t).casessubscriptπœ‡π±π‘‘subscript^𝐱0𝑑otherwisesuperscriptsubscript𝜎𝐱2𝑑subscript𝛼0superscriptsubscript^𝐱𝛼2𝑑otherwise\begin{cases}\mu_{\mathbf{x}}(t)=\hat{\mathbf{x}}_{\mathbf{0}}(t)\\ \sigma_{\mathbf{x}}^{2}(t)=\sum_{|\mathbf{\alpha}|\neq 0}\hat{\mathbf{x}}_{% \mathbf{\alpha}}^{2}(t)\,.\end{cases}{ start_ROW start_CELL italic_ΞΌ start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT ( italic_t ) = over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_Οƒ start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_t ) = βˆ‘ start_POSTSUBSCRIPT | italic_Ξ± | β‰  0 end_POSTSUBSCRIPT over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_t ) . end_CELL start_CELL end_CELL end_ROW (7)

The expansion (6) holds theoretically; in practice, for numerical implementation, one should truncate it, obtaining an approximation that enables a straightforward computation of the statistical moments through equation (7). Approximations can be computed either by intrusive approaches (e.g. Galerkin), which modify the governing equations of the original system by truncating the gPC expansion, or by non-intrusive approaches (e.g. collocation) that treat the model as a black box and use sampling to estimate the spectral coefficients indirectly, e.g. via least-squares regression.

As a first approach to generate an approximation of 𝐱⁒(t;Z)𝐱𝑑𝑍\mathbf{x}(t;Z)bold_x ( italic_t ; italic_Z ), assume that (1) is a polynomial system, to have

f⁒(𝐱,Z)=βˆ‘|k|≀Lak⁒(Z)⁒x1k1⁒(t,Z)⁒…⁒xnkn⁒(t,Z),ak⁒(Z)=βˆ‘Ξ±βˆˆβ„•0da^k,α⁒Φα⁒(Z)βˆˆβ„n,missing-subexpression𝑓𝐱𝑍subscriptπ‘˜πΏsubscriptπ‘Žπ‘˜π‘superscriptsubscriptπ‘₯1subscriptπ‘˜1𝑑𝑍…superscriptsubscriptπ‘₯𝑛subscriptπ‘˜π‘›π‘‘π‘missing-subexpressionmissing-subexpressionsubscriptπ‘Žπ‘˜π‘subscript𝛼superscriptsubscriptβ„•0𝑑subscript^π‘Žπ‘˜π›ΌsubscriptΦ𝛼𝑍superscriptℝ𝑛missing-subexpression\begin{array}[]{lll}&f(\mathbf{x},Z)=\sum_{|k|\leq L}a_{k}(Z)x_{1}^{k_{1}}(t,Z% )\ldots x_{n}^{k_{n}}(t,Z),\\ &a_{k}(Z)=\sum_{\alpha\in\mathbb{N}_{0}^{d}}\hat{a}_{k,\alpha}\Phi_{\alpha}(Z)% \in\mathbb{R}^{n},\end{array}start_ARRAY start_ROW start_CELL end_CELL start_CELL italic_f ( bold_x , italic_Z ) = βˆ‘ start_POSTSUBSCRIPT | italic_k | ≀ italic_L end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_Z ) italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( italic_t , italic_Z ) … italic_x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( italic_t , italic_Z ) , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL italic_a start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_Z ) = βˆ‘ start_POSTSUBSCRIPT italic_Ξ± ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_k , italic_Ξ± end_POSTSUBSCRIPT roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_Z ) ∈ blackboard_R start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , end_CELL start_CELL end_CELL end_ROW end_ARRAY (8)

where k=(k1,…,kn)βˆˆβ„•0nπ‘˜subscriptπ‘˜1…subscriptπ‘˜π‘›superscriptsubscriptβ„•0𝑛k=(k_{1},\dots,k_{n})\in\mathbb{N}_{0}^{n}italic_k = ( italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_k start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT and Lβˆˆβ„•0𝐿subscriptβ„•0L\in\mathbb{N}_{0}italic_L ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Hence, by substituting (6) and (8) into (1), multiplying both sides by ΦβsubscriptΦ𝛽\Phi_{\beta}roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ² end_POSTSUBSCRIPT and taking the expectation, we have

dd⁒t⁒𝐱^β⁒(t)=𝔼⁒[f⁒(βˆ‘Ξ±βˆˆβ„•0d𝐱^α⁒(t)⁒Φα,Z)⁒Φβ]=βˆ‘|k|≀L𝔼⁒[Ο‡k⁒Φβ],Ξ²βˆˆβ„•0d,Ο‡k:=(βˆ‘Ξ±βˆˆβ„•0da^k,α⁒Φα)⁒∏j=1n(βˆ‘Ξ±βˆˆβ„•0d𝐱^Ξ±,j⁒(t)⁒Φα)kj.missing-subexpression𝑑𝑑𝑑subscript^𝐱𝛽𝑑𝔼delimited-[]𝑓subscript𝛼superscriptsubscriptβ„•0𝑑subscript^𝐱𝛼𝑑subscriptΦ𝛼𝑍subscriptΦ𝛽missing-subexpressionmissing-subexpressionformulae-sequenceabsentsubscriptπ‘˜πΏπ”Όdelimited-[]subscriptπœ’π‘˜subscriptΦ𝛽𝛽superscriptsubscriptβ„•0𝑑missing-subexpressionmissing-subexpressionassignsubscriptπœ’π‘˜subscript𝛼superscriptsubscriptβ„•0𝑑subscript^π‘Žπ‘˜π›ΌsubscriptΦ𝛼superscriptsubscriptproduct𝑗1𝑛superscriptsubscript𝛼superscriptsubscriptβ„•0𝑑subscript^𝐱𝛼𝑗𝑑subscriptΦ𝛼subscriptπ‘˜π‘—missing-subexpression\begin{array}[]{lll}&\frac{d}{dt}\hat{\mathbf{x}}_{\beta}(t)=\mathbb{E}[f(\sum% _{\alpha\in\mathbb{N}_{0}^{d}}\hat{\mathbf{x}}_{\mathbf{\alpha}}(t)\Phi_{% \mathbf{\alpha}},Z)\Phi_{\beta}]\\ &\hskip 34.1433pt=\sum_{|k|\leq L}\mathbb{E}\left[\chi_{k}\Phi_{\beta}\right],% \ \beta\in\mathbb{N}_{0}^{d},\\ &\chi_{k}:=\left(\sum_{\alpha\in\mathbb{N}_{0}^{d}}\hat{a}_{k,\alpha}\Phi_{% \alpha}\right)\prod_{j=1}^{n}\left(\sum_{\alpha\in\mathbb{N}_{0}^{d}}\hat{% \mathbf{x}}_{\alpha,j}(t)\Phi_{\alpha}\right)^{k_{j}}.\end{array}start_ARRAY start_ROW start_CELL end_CELL start_CELL divide start_ARG italic_d end_ARG start_ARG italic_d italic_t end_ARG over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ² end_POSTSUBSCRIPT ( italic_t ) = blackboard_E [ italic_f ( βˆ‘ start_POSTSUBSCRIPT italic_Ξ± ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_t ) roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT , italic_Z ) roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ² end_POSTSUBSCRIPT ] end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = βˆ‘ start_POSTSUBSCRIPT | italic_k | ≀ italic_L end_POSTSUBSCRIPT blackboard_E [ italic_Ο‡ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ² end_POSTSUBSCRIPT ] , italic_Ξ² ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL italic_Ο‡ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT := ( βˆ‘ start_POSTSUBSCRIPT italic_Ξ± ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_k , italic_Ξ± end_POSTSUBSCRIPT roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ) ∏ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ( βˆ‘ start_POSTSUBSCRIPT italic_Ξ± ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over^ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± , italic_j end_POSTSUBSCRIPT ( italic_t ) roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT . end_CELL start_CELL end_CELL end_ROW end_ARRAY (9)

The approximation to 𝐱⁒(t;Z)𝐱𝑑𝑍\mathbf{x}(t;Z)bold_x ( italic_t ; italic_Z ) is then of the form

𝐱N⁒(t;Z)=βˆ‘|Ξ±|≀N𝐱~α⁒(t)⁒Φα⁒(Z),𝐱~α⁒(t)=[𝐱~Ξ±,1⁒(t)…𝐱~Ξ±,n⁒(t)]βŠ€βˆˆβ„n,missing-subexpressionsubscript𝐱𝑁𝑑𝑍subscript𝛼𝑁subscript~𝐱𝛼𝑑subscriptΦ𝛼𝑍missing-subexpressionmissing-subexpressionsubscript~𝐱𝛼𝑑superscriptmatrixsubscript~𝐱𝛼1𝑑…subscript~𝐱𝛼𝑛𝑑topsuperscriptℝ𝑛missing-subexpression\begin{array}[]{lll}&\mathbf{x}_{N}(t;Z)=\sum_{|\alpha|\leq N}\tilde{\mathbf{x% }}_{\alpha}(t)\Phi_{\alpha}(Z),\\ &\tilde{\mathbf{x}}_{\alpha}(t)=\begin{bmatrix}\tilde{\mathbf{x}}_{\alpha,1}(t% )&\dots&\tilde{\mathbf{x}}_{\alpha,n}(t)\end{bmatrix}^{\top}\in\mathbb{R}^{n},% \end{array}start_ARRAY start_ROW start_CELL end_CELL start_CELL bold_x start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_t ; italic_Z ) = βˆ‘ start_POSTSUBSCRIPT | italic_Ξ± | ≀ italic_N end_POSTSUBSCRIPT over~ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_t ) roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_Z ) , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL over~ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_t ) = [ start_ARG start_ROW start_CELL over~ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± , 1 end_POSTSUBSCRIPT ( italic_t ) end_CELL start_CELL … end_CELL start_CELL over~ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± , italic_n end_POSTSUBSCRIPT ( italic_t ) end_CELL end_ROW end_ARG ] start_POSTSUPERSCRIPT ⊀ end_POSTSUPERSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , end_CELL start_CELL end_CELL end_ROW end_ARRAY (10)

where the coefficients {𝐱~i⁒(t)}|Ξ±|≀Nsubscriptsubscript~𝐱𝑖𝑑𝛼𝑁\left\{\tilde{\mathbf{x}}_{i}(t)\right\}_{|\alpha|\leq N}{ over~ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) } start_POSTSUBSCRIPT | italic_Ξ± | ≀ italic_N end_POSTSUBSCRIPT in (10) are obtained as solutions to the truncated system:

dd⁒t⁒𝐱~β⁒(t)=βˆ‘|k|≀L𝔼⁒[Ο‡~k⁒Φβ],|Ξ²|≀N,Ο‡~k:=(βˆ‘|Ξ±|≀Na^k,α⁒Φα)⁒∏j=1n(βˆ‘|Ξ±|≀N𝐱~Ξ±,j⁒(t)⁒Φα)kj.missing-subexpressionformulae-sequence𝑑𝑑𝑑subscript~𝐱𝛽𝑑subscriptπ‘˜πΏπ”Όdelimited-[]subscript~πœ’π‘˜subscriptΦ𝛽𝛽𝑁missing-subexpressionmissing-subexpressionassignsubscript~πœ’π‘˜subscript𝛼𝑁subscript^π‘Žπ‘˜π›ΌsubscriptΦ𝛼superscriptsubscriptproduct𝑗1𝑛superscriptsubscript𝛼𝑁subscript~𝐱𝛼𝑗𝑑subscriptΦ𝛼subscriptπ‘˜π‘—missing-subexpression\begin{array}[]{lll}&\frac{d}{dt}\tilde{\mathbf{x}}_{\beta}(t)=\sum_{|k|\leq L% }\mathbb{E}\left[\tilde{\chi}_{k}\Phi_{\beta}\right],\ |\beta|\leq N,\\ &\tilde{\chi}_{k}:=\left(\sum_{|\alpha|\leq N}\hat{a}_{k,\alpha}\Phi_{\alpha}% \right)\prod_{j=1}^{n}\left(\sum_{|\alpha|\leq N}\tilde{\mathbf{x}}_{\alpha,j}% (t)\Phi_{\alpha}\right)^{k_{j}}.\end{array}start_ARRAY start_ROW start_CELL end_CELL start_CELL divide start_ARG italic_d end_ARG start_ARG italic_d italic_t end_ARG over~ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ² end_POSTSUBSCRIPT ( italic_t ) = βˆ‘ start_POSTSUBSCRIPT | italic_k | ≀ italic_L end_POSTSUBSCRIPT blackboard_E [ over~ start_ARG italic_Ο‡ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ² end_POSTSUBSCRIPT ] , | italic_Ξ² | ≀ italic_N , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL over~ start_ARG italic_Ο‡ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT := ( βˆ‘ start_POSTSUBSCRIPT | italic_Ξ± | ≀ italic_N end_POSTSUBSCRIPT over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_k , italic_Ξ± end_POSTSUBSCRIPT roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ) ∏ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ( βˆ‘ start_POSTSUBSCRIPT | italic_Ξ± | ≀ italic_N end_POSTSUBSCRIPT over~ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± , italic_j end_POSTSUBSCRIPT ( italic_t ) roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT . end_CELL start_CELL end_CELL end_ROW end_ARRAY (11)

System (11) is obtained from system (9) by retaining only polynomials whose degree is less or equal to N𝑁Nitalic_N. This approach is the so-called stochastic Galerkin projection.

A competing approach to generate surrogates of 𝐱⁒(t;Z)𝐱𝑑𝑍\mathbf{x}(t;Z)bold_x ( italic_t ; italic_Z ) is the collocation method, based on sampling the values of the random variable Z𝑍Zitalic_Z. Let {Z(j)}j=1SCβŠ‚IZ,SCβ‰₯(N+dd)formulae-sequencesuperscriptsubscriptsuperscript𝑍𝑗𝑗1subscript𝑆𝐢subscript𝐼𝑍subscript𝑆𝐢binomial𝑁𝑑𝑑\left\{Z^{(j)}\right\}_{j=1}^{S_{C}}\subset I_{Z},\ S_{C}\geq\binom{N+d}{d}{ italic_Z start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT } start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_S start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT end_POSTSUPERSCRIPT βŠ‚ italic_I start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT , italic_S start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT β‰₯ ( FRACOP start_ARG italic_N + italic_d end_ARG start_ARG italic_d end_ARG ), be a grid of nodes sampled in the range of Z𝑍Zitalic_Z. There are multiple ways to generate such a grid – as pseudo-random Monte Carlo sampling according to ρZsubscriptπœŒπ‘\rho_{Z}italic_ρ start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT, as nodes of a pre-selected quadrature rule, the Smolyak sparse grid [smolyak1963quadrature] or a Clenshaw-Curtis grid [engels1980numerical]. For each 1≀j≀SC1𝑗subscript𝑆𝐢1\leq j\leq S_{C}1 ≀ italic_j ≀ italic_S start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, let 𝐱¯(j)⁒(t)superscript¯𝐱𝑗𝑑\bar{\mathbf{x}}^{(j)}(t)overΒ― start_ARG bold_x end_ARG start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_t ) be a solution to the deterministic system (1) with Z𝑍Zitalic_Z replaced by the value Z(j)superscript𝑍𝑗Z^{(j)}italic_Z start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT. The desired approximation to 𝐱⁒(t;Z)𝐱𝑑𝑍\mathbf{x}(t;Z)bold_x ( italic_t ; italic_Z ) is considered again in the form (10), subject to the conditions

𝐱N⁒(t;Z(j))=βˆ‘|Ξ±|≀N𝐱~α⁒(t)⁒Φα⁒(Z(j))=𝐱¯(j)⁒(t)subscript𝐱𝑁𝑑superscript𝑍𝑗subscript𝛼𝑁subscript~𝐱𝛼𝑑subscriptΦ𝛼superscript𝑍𝑗superscript¯𝐱𝑗𝑑missing-subexpressionmissing-subexpression\begin{array}[]{lll}\mathbf{x}_{N}(t;Z^{(j)})=\sum_{|\alpha|\leq N}\tilde{% \mathbf{x}}_{\alpha}(t)\Phi_{\alpha}(Z^{(j)})=\bar{\mathbf{x}}^{(j)}(t)\end{array}start_ARRAY start_ROW start_CELL bold_x start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_t ; italic_Z start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ) = βˆ‘ start_POSTSUBSCRIPT | italic_Ξ± | ≀ italic_N end_POSTSUBSCRIPT over~ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_t ) roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_Z start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ) = overΒ― start_ARG bold_x end_ARG start_POSTSUPERSCRIPT ( italic_j ) end_POSTSUPERSCRIPT ( italic_t ) end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW end_ARRAY (12)

where 1≀j≀SC1𝑗subscript𝑆𝐢1\leq j\leq S_{C}1 ≀ italic_j ≀ italic_S start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT. These conditions impose an over-determined (due to the condition on SCsubscript𝑆𝐢S_{C}italic_S start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT) linear system on the coefficients {𝐱~α⁒(t)}|Ξ±|≀Nsubscriptsubscript~𝐱𝛼𝑑𝛼𝑁\left\{\tilde{\mathbf{x}}_{\alpha}(t)\right\}_{|\alpha|\leq N}{ over~ start_ARG bold_x end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_t ) } start_POSTSUBSCRIPT | italic_Ξ± | ≀ italic_N end_POSTSUBSCRIPT, which can be solved either explicitly or via a least-squares projection.

Remark 2

The collocation approach takes a general form regardless of the structure of the dynamics. Conversely, the Galerkin approach can be made more precise for polynomial systems; when expectations of higher order products of basis elements are explicitly available [Petzke2020_PoCET], explicit formulas can be embedded in (11), without any additional approximation. This can be an advantage also for the uncertainty quantification of input-output maps of nonlinear systems that can be rewritten as polynomial systems in a higher-dimensional space, e.g. via the immersion technique [ohtsuka2005model]. β‹„β‹„\hfill\diamondβ‹„

The rate of convergence of the approximation depends on the smoothness of the function f𝑓fitalic_f and the type of orthogonal polynomial basis functions {Φα}Ξ±subscriptsubscriptΦ𝛼𝛼\{\Phi_{\alpha}\}_{\alpha}{ roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT } start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT [xiu2010numerical]. For a function g∈LρZ2⁒(IZ)𝑔subscriptsuperscript𝐿2subscriptπœŒπ‘subscript𝐼𝑍g\in L^{2}_{\rho_{Z}}(I_{Z})italic_g ∈ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_I start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT ) with expansion g⁒(Z)=βˆ‘Ξ±βˆˆβ„•0dg^α⁒Φα⁒(Z)𝑔𝑍subscript𝛼superscriptsubscriptβ„•0𝑑subscript^𝑔𝛼subscriptΦ𝛼𝑍g(Z)=\sum_{\alpha\in\mathbb{N}_{0}^{d}}\hat{g}_{\alpha}\Phi_{\alpha}(Z)italic_g ( italic_Z ) = βˆ‘ start_POSTSUBSCRIPT italic_Ξ± ∈ blackboard_N start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_g end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT ( italic_Z ), the approximation error βˆ₯gβˆ’βˆ‘|Ξ±|≀Ng^α⁒Φαβˆ₯LρZ⁒(IZ)2subscriptdelimited-βˆ₯βˆ₯𝑔subscript𝛼𝑁subscript^𝑔𝛼subscriptΦ𝛼subscriptsuperscript𝐿2subscriptπœŒπ‘subscript𝐼𝑍\lVert g-\sum_{|\alpha|\leq N}\hat{g}_{\alpha}\Phi_{\alpha}\rVert_{L^{2}_{\rho% _{Z}(I_{Z})}}βˆ₯ italic_g - βˆ‘ start_POSTSUBSCRIPT | italic_Ξ± | ≀ italic_N end_POSTSUBSCRIPT over^ start_ARG italic_g end_ARG start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT roman_Ξ¦ start_POSTSUBSCRIPT italic_Ξ± end_POSTSUBSCRIPT βˆ₯ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT ( italic_I start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT is π’ͺ⁒(Nβˆ’β„“)π’ͺsuperscript𝑁ℓ\mathcal{O}(N^{-\ell})caligraphic_O ( italic_N start_POSTSUPERSCRIPT - roman_β„“ end_POSTSUPERSCRIPT ), where β„“β„“\ellroman_β„“ denotes the differentiability order of g𝑔gitalic_g. For analytic g𝑔gitalic_g, the convergence rate is exponential, i.e. π’ͺ⁒(eβˆ’Οƒβ’N)π’ͺsuperscriptπ‘’πœŽπ‘\mathcal{O}(e^{-\sigma N})caligraphic_O ( italic_e start_POSTSUPERSCRIPT - italic_Οƒ italic_N end_POSTSUPERSCRIPT ) for some constant Οƒ>0𝜎0\sigma>0italic_Οƒ > 0. For solutions to ODE and PDE system, few rigorous results that quantify the convergence rates exist. Some analytic results and numerous numerical studies found in the literature indicate that similar convergence rates hold for ‖𝐱⁒(t;β‹…)βˆ’π±N⁒(t;β‹…)β€–Lρz2⁒(IZ)subscriptnorm𝐱𝑑⋅subscript𝐱𝑁𝑑⋅subscriptsuperscript𝐿2subscriptπœŒπ‘§subscript𝐼𝑍\left\|\mathbf{x}(t;\cdot)-\mathbf{x}_{N}(t;\cdot)\right\|_{L^{2}_{\rho_{z}}(I% _{Z})}βˆ₯ bold_x ( italic_t ; β‹… ) - bold_x start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_t ; β‹… ) βˆ₯ start_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_I start_POSTSUBSCRIPT italic_Z end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT, at least on compact time intervals; see [kim2013generalised, xiu2002wiener] for further details.

Numerical investigation can be employed to gain insight as to how gPC compares with classical MC methods. Such numerical comparisons have been explored rarely, especially for systems emanating from systems biology [SonDu2020], where Monte Carlo has been the main workhorse for decades. Below, as a benchmark to perform a detailed comparison of gPC variants and Monte Carlo methods, we consider the Hindmarsh-Rose model, as a representative for the class of polynomial systems arising in neuroscience.

II-B The Hindmarsh-Rose model as case study

The Hindmarsh-Rose (HR) model [hindmarsh1984model], well-known in neuroscience, reproduces the dynamics of action-potential within single neurons capable of bursting activity [innocenti2007dynamical]. It can display numerous patterns of behaviours, ranging from cyclic spiking to bursting or chaos, depending on the considered parameter sets [montanari2022functional], which are often uncertain due to experimental design or neural activity. The HR model is represented by the (a-dimensionalised) set of ODEs:

{xΛ™=yβˆ’a⁒x3+b⁒x2βˆ’z+I,yΛ™=cβˆ’d⁒x2βˆ’y,zΛ™=r⁒[s⁒(xβˆ’xR)βˆ’z],casesΛ™π‘₯π‘¦π‘Žsuperscriptπ‘₯3𝑏superscriptπ‘₯2𝑧𝐼otherwise˙𝑦𝑐𝑑superscriptπ‘₯2𝑦otherwiseΛ™π‘§π‘Ÿdelimited-[]𝑠π‘₯subscriptπ‘₯𝑅𝑧otherwise\begin{cases}\dot{x}=y-ax^{3}+bx^{2}-z+I,\\ \dot{y}=c-dx^{2}-y,\\ \dot{z}=r[s(x-x_{R})-z]\,,\end{cases}{ start_ROW start_CELL overΛ™ start_ARG italic_x end_ARG = italic_y - italic_a italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + italic_b italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_z + italic_I , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL overΛ™ start_ARG italic_y end_ARG = italic_c - italic_d italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_y , end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL overΛ™ start_ARG italic_z end_ARG = italic_r [ italic_s ( italic_x - italic_x start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ) - italic_z ] , end_CELL start_CELL end_CELL end_ROW (13)

where xπ‘₯xitalic_x represents the membrane potential, y𝑦yitalic_y is a fast recovery current, and z𝑧zitalic_z is a slow adaptation (inward) current; I𝐼Iitalic_I is an externally applied current (either experimental current injection or in-vivo synaptic current), and the other terms are model parameters. Their values can be inferred from fitting on experimental data [gu2013biological], which nonetheless may depend on the considered organism and come with significant uncertainties. Common default values are a=1,b=3,c=1,d=5,r=0.01,s=4,xR=βˆ’8/5formulae-sequenceπ‘Ž1formulae-sequence𝑏3formulae-sequence𝑐1formulae-sequence𝑑5formulae-sequenceπ‘Ÿ0.01formulae-sequence𝑠4subscriptπ‘₯𝑅85a=1,b=3,c=1,d=5,r=0.01,s=4,x_{R}=-8/5italic_a = 1 , italic_b = 3 , italic_c = 1 , italic_d = 5 , italic_r = 0.01 , italic_s = 4 , italic_x start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT = - 8 / 5. If parameter values are changed deterministically, bifurcation studies [gonzalez2007complex, storace2008hindmarsh] allow to identify parameter combinations corresponding to various patterns.

Fixing aπ‘Žaitalic_a, c𝑐citalic_c, d𝑑ditalic_d, rπ‘Ÿritalic_r, s𝑠sitalic_s and xRsubscriptπ‘₯𝑅x_{R}italic_x start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT in (13) to default values and letting I𝐼Iitalic_I and b𝑏bitalic_b vary within given intervals, five dynamical regimes can be recognised in the HR dynamics [de2008predicting, gonzalez2007complex, storace2008hindmarsh]: quiescence (A𝐴Aitalic_A), tonic spiking (B𝐡Bitalic_B), square-wave bursting (C𝐢Citalic_C), plateau bursting (D𝐷Ditalic_D) and chaotic bursting (E𝐸Eitalic_E). Examples are shown in Fig. 1.

Refer to caption
Figure 1: Examples of time series x⁒(t)π‘₯𝑑x(t)italic_x ( italic_t ) for the regimes (A𝐴Aitalic_A)–(E𝐸Eitalic_E) of model (13), described in the main text. The lines in (C𝐢Citalic_C) and (D𝐷Ditalic_D) identify a burst of spikes.

First, we study the advantage of gPC methods in reconstructing stochastic simulations for each individual regime. To do so, we further fix I𝐼Iitalic_I and consider uniform distributions of b∈[bmin,j;bmax,j]𝑏subscript𝑏min𝑗subscript𝑏max𝑗b\in[b_{\text{min},j};b_{\text{max},j}]italic_b ∈ [ italic_b start_POSTSUBSCRIPT min , italic_j end_POSTSUBSCRIPT ; italic_b start_POSTSUBSCRIPT max , italic_j end_POSTSUBSCRIPT ], where j𝑗jitalic_j identifies the regime in {A,B,C}𝐴𝐡𝐢\{A,B,C\}{ italic_A , italic_B , italic_C }. For regime D𝐷Ditalic_D, we fix b𝑏bitalic_b and vary I∈[Imin,D;Imax,D]𝐼subscript𝐼min𝐷subscript𝐼max𝐷I\in[I_{\text{min},D};I_{\text{max},D}]italic_I ∈ [ italic_I start_POSTSUBSCRIPT min , italic_D end_POSTSUBSCRIPT ; italic_I start_POSTSUBSCRIPT max , italic_D end_POSTSUBSCRIPT ] uniformly. Without loss of generality, minimum and maximum values may differ among regimes, due to their persistence set (see e.g. [barrio2011parameter, Figure 3]). Chaos (regime E𝐸Eitalic_E) adds another layer of complexity and is currently not investigated.

For the Galerkin approach, we assess the accuracy of the obtained surrogate model (10) with respect to the expansion order N=NG𝑁subscript𝑁𝐺N=N_{G}italic_N = italic_N start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT. For the collocation approach, we assess the accuracy of the obtained surrogate model (10) with respect to both the expansion order N=NC𝑁subscript𝑁𝐢N=N_{C}italic_N = italic_N start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and the collocation samples SCsubscript𝑆𝐢S_{C}italic_S start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT. As a proxy for the computational complexity of each gPC method mπ‘šmitalic_m, we estimate the running time Ο„msubscriptπœπ‘š\tau_{m}italic_Ο„ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT, from the first call of the polynomial projection to the estimation of the spectral coefficients of interest. This is a worst-case time in the case of fresh simulations: it includes the compilation time for the expansion polynomials, which can be stored in the case of repeated runs, without concurring to additional processing. The running time Ο„msubscriptπœπ‘š\tau_{m}italic_Ο„ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT is compared with the running time Ο„M⁒Csubscriptπœπ‘€πΆ\tau_{MC}italic_Ο„ start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT of a Monte Carlo chain with NM⁒C=5000subscript𝑁𝑀𝐢5000N_{MC}=5000italic_N start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT = 5000, bounding Ξ΄=0.01𝛿0.01\delta=0.01italic_Ξ΄ = 0.01 and Ξ΅<0.08πœ€0.08\varepsilon<0.08italic_Ξ΅ < 0.08 according to (2).

To assess the accuracy of gPC methods, we first construct a benchmark "ground truth" state for each regime, employing a Monte Carlo scheme with NM⁒C=105subscript𝑁𝑀𝐢superscript105N_{MC}=10^{5}italic_N start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT, guaranteeing accuracy Ξ΅<0.01πœ€0.01\varepsilon<0.01italic_Ξ΅ < 0.01 for the same confidence δ𝛿\deltaitalic_Ξ΄. Given a final simulation time T𝑇Titalic_T and a time-step d⁒t𝑑𝑑dtitalic_d italic_t, we then compute the error vector between the "ground truth" and the specific surrogate model results eβˆˆβ„Ne𝑒superscriptℝsubscript𝑁𝑒e\in\mathbb{R}^{N_{e}}italic_e ∈ blackboard_R start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, Ne=Td⁒tsubscript𝑁𝑒𝑇𝑑𝑑N_{e}=\frac{T}{dt}italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = divide start_ARG italic_T end_ARG start_ARG italic_d italic_t end_ARG. Deviations for both mean and variance are estimated using the element-wise root mean square error (RMSE)

eRMS=1Neβ’βˆ‘n=1Neen2.subscript𝑒RMS1subscript𝑁𝑒superscriptsubscript𝑛1subscript𝑁𝑒superscriptsubscript𝑒𝑛2e_{\text{RMS}}=\sqrt{\frac{1}{N_{e}}\sum_{n=1}^{N_{e}}e_{n}^{2}}.italic_e start_POSTSUBSCRIPT RMS end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG βˆ‘ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG .

After assessing the computational advantage for the gPC methods, we select sweet spot settings and show their use in assessing the probabilistic robust behaviour for a selected regime: plateau bursting (D𝐷Ditalic_D). For a given I=4.2𝐼4.2I=4.2italic_I = 4.2, we thus investigate the probability that D𝐷Ditalic_D persists, depending on sampled values b𝑏bitalic_b within a uniform distribution b∈Ib=[2.4;2.8]𝑏subscript𝐼𝑏2.42.8b\in I_{b}=[2.4;2.8]italic_b ∈ italic_I start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = [ 2.4 ; 2.8 ], where 2.4 is associated with typical leftmost values considered in phase planes from literature [barrio2011parameter], and Δ⁒b=0.4Δ𝑏0.4\Delta b=0.4roman_Ξ” italic_b = 0.4 is informed by experimental uncertainties obtained from fitting experimental data [de2008predicting]. Plateau bursting is known to be characterised by a Hopf bifurcation [gonzalez2007complex] marking the end of the burst of spikes at x=0π‘₯0x=0italic_x = 0. We use this signature on the gPC mean to identify if the simulations maintain regime D𝐷Ditalic_D or not. We further characterise the proportion Pβˆ—=size⁒(Ab)size⁒(Ib)superscript𝑃sizesubscript𝐴𝑏sizesubscript𝐼𝑏P^{*}=\frac{\mbox{size}(A_{b})}{\mbox{size}(I_{b})}italic_P start_POSTSUPERSCRIPT βˆ— end_POSTSUPERSCRIPT = divide start_ARG size ( italic_A start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) end_ARG start_ARG size ( italic_I start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) end_ARG, where AbβŠ†Ibsubscript𝐴𝑏subscript𝐼𝑏A_{b}\subseteq I_{b}italic_A start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT βŠ† italic_I start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is the interval for which D𝐷Ditalic_D is preserved, as a measure of robustness probability.

Simulations of system (13) are performed using Matlab ode45 function, with d⁒t=0.01𝑑𝑑0.01dt=0.01italic_d italic_t = 0.01 up to T=1200𝑇1200T=1200italic_T = 1200. Initial conditions are set uniformly at [0,0,0]⊀superscript000top[0,0,0]^{\top}[ 0 , 0 , 0 ] start_POSTSUPERSCRIPT ⊀ end_POSTSUPERSCRIPT for each regime. The initial 50% of each simulation is discarded to cut-off the burn-in period. gPC approaches are implemented via the PoCET Matlab toolbox [Petzke2020_PoCET], already optimised for uncertainty quantification. To maintain consistency, Monte Carlo simulations are also performed using PoCET built-in functions. All simulations were performed on a Dell Inspiron 16 laptop, with 16 GB RAM and 1.90 GHz Intel i5-1340P core, running Windows 11. Simulations were performed with the laptop on charge, to avoid power saving modes.

As a first assessment of MC running time, we notice that completing the construction of the benchmark "ground truth" took more than 7h per regime.

III Results

III-A gPC methods outrun MC and maintain accuracy

We first compare gPC methods with a MC chain with NM⁒C=5000subscript𝑁𝑀𝐢5000N_{MC}=5000italic_N start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT = 5000. On top of the theoretical bound (2) to derive its computational complexity, we estimate the MC empirical running time and RMSE, to perform direct comparison with gPC statistics. The corresponding results are listed in Table II. The running time is in the order of 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPTs, the mean and variance of RMSE are around 0.010.010.010.01.

TABLE II: Performance metrics for MC with NM⁒C=5000subscript𝑁𝑀𝐢5000N_{MC}=5000italic_N start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT = 5000.
Regime Ο„M⁒Csubscriptπœπ‘€πΆ\tau_{MC}italic_Ο„ start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT [s] Mean RMSE Var RMSE
A𝐴Aitalic_A 1238 0.0070 0.013
B𝐡Bitalic_B 1248 0.0062 0.012
C𝐢Citalic_C 1269 0.0089 0.017
D𝐷Ditalic_D 1257 0.015 0.013
E𝐸Eitalic_E 1205 0.010 0.012

For the Galerkin approximation, the computation time in all regimes, Ο„Gsubscript𝜏𝐺\tau_{G}italic_Ο„ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT, scales sub-exponentially for small expansion orders and exponentially for medium-large NGsubscript𝑁𝐺N_{G}italic_N start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT, following a best fit relation

y=a⁒eb⁒NG+c⁒ed⁒NGπ‘¦π‘Žsuperscript𝑒𝑏subscript𝑁𝐺𝑐superscript𝑒𝑑subscript𝑁𝐺y=a\ e^{bN_{G}}+c\ e^{dN_{G}}\,italic_y = italic_a italic_e start_POSTSUPERSCRIPT italic_b italic_N start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + italic_c italic_e start_POSTSUPERSCRIPT italic_d italic_N start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT end_POSTSUPERSCRIPT (14)

with y=Ο„G𝑦subscript𝜏𝐺y=\tau_{G}italic_y = italic_Ο„ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT. We refer to Fig. 2a for the relationship between Ο„Gsubscript𝜏𝐺\tau_{G}italic_Ο„ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT and NGsubscript𝑁𝐺N_{G}italic_N start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT. For any considered expansion order, Ο„Gsubscript𝜏𝐺\tau_{G}italic_Ο„ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT is at least one order of magnitude lower than Ο„M⁒Csubscriptπœπ‘€πΆ\tau_{MC}italic_Ο„ start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT, marking a significant acceleration in computation.

Despite fluctuations related to stochasticity, the mean RMSE decays rapidly (following the same trend as in (14), with different parameters) as the expansion order increases, as shown in Fig. 2b. For "simple" regimes (A𝐴Aitalic_A, B𝐡Bitalic_B), the mean discrepancy with high NGsubscript𝑁𝐺N_{G}italic_N start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT is close to that of the MC chain; for "difficult" ones, like C𝐢Citalic_C and D𝐷Ditalic_D with bursting, it increases without exceeding the order of magnitude. Overall, the variance of RMSE remains rather constant for increasing NGsubscript𝑁𝐺N_{G}italic_N start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT, around values of 0.3, which is about three times that of MC; see Fig. 2c.

These results suggest that the Galerkin approximation drastically speeds up the computation time with respect to MC methods, but at the price of decreasing accuracy.

Refer to caption
Figure 2: Results for the Galerkin gPC approximation, for the considered regimes, depending on the expansion order NGsubscript𝑁𝐺N_{G}italic_N start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT. (a) Running time Ο„Gsubscript𝜏𝐺\tau_{G}italic_Ο„ start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT. The fit example follows (14) for the regime C𝐢Citalic_C; the trend is conserved for all regimes. (b) Mean of RMSE. For MC, it is around 0.01 (Table II). Dashed lines represent empirical fitting following (14). (c) Variance of RMSE. For MC, it is around 0.013 (Table II).
Refer to caption
Figure 3: Results for the collocation gPC approximation. We show only the plots referring to regime C𝐢Citalic_C, which represent the worst-case in absolute values for mean and variance of RMSE. The metrics are estimated in relation to expansion order and number of collocation samples. (Left) Running time Ο„Csubscript𝜏𝐢\tau_{C}italic_Ο„ start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT. (Centre) Mean of RMSE. (Right) Variance of RMSE.

Collocation relies both on expansion order and on collocation samples. In Fig. 3, we display an example referring to the regime C𝐢Citalic_C, which represents the worst-case; all other regimes have better statistics. The running time Ο„Csubscript𝜏𝐢\tau_{C}italic_Ο„ start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and mean RMSE are more significantly influenced by the number of collocation samples, whereas the expansion order has little impact, at least in the considered interval (Fig. 3a). On the other hand, a higher expansion order significantly improves the variance of RMSE. Notably, even for high NCsubscript𝑁𝐢N_{C}italic_N start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and high SCsubscript𝑆𝐢S_{C}italic_S start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, the running time remains around 2 minutes, while MC takes about 20 minutes, highlighting a 10Γ—10\times10 Γ— acceleration.

We have not considered expansion orders above NC>15subscript𝑁𝐢15N_{C}>15italic_N start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT > 15, which would likely yield comparable scaling as to those of the Galerking approach (Fig. 2a). However, this choice is justified by the fact that the mean of RMSE is already comparable to that of MC, even with a relatively low expansion order NCsubscript𝑁𝐢N_{C}italic_N start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT (Fig. 3b, c), at the price of losing one order of magnitude in the variance of RMSE. We recall that Fig. 3 reports a worst case; other regimes have RMSE statistics that are very close to those reported in Table II.

These findings support the use of collocation methods as a good compromise between computing time and accuracy of the results. In fact, Galerkin simulations with similar Ο„πœ\tauitalic_Ο„ reach lower accuracy on all regimes. Moreover, the further acceleration in computation obtained with low NGsubscript𝑁𝐺N_{G}italic_N start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT yields large quantitative discrepancies, which may hinder the investigation of models like the Hindmarsh-Rose (13), which showcases several dynamical patterns. Low-N𝑁Nitalic_N Galerkin approximation may nonetheless be useful to accelerate the analysis of "simpler" models with more uniform dynamics.

In the analysis of pattern robustness for the HR model, we therefore employ collocation gPC with 400 samples and NC=15subscript𝑁𝐢15N_{C}=15italic_N start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = 15. The gPC mean in each interval AbβŠ†Ibsubscript𝐴𝑏subscript𝐼𝑏A_{b}\subseteq I_{b}italic_A start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT βŠ† italic_I start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is immediately computed following equation (7) in PoCET. Fig. 4a shows an example of ΞΌx⁒(t)subscriptπœ‡π‘₯𝑑\mu_{x}(t)italic_ΞΌ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_t ) for b∈[2.4;2.48]𝑏2.42.48b\in[2.4;2.48]italic_b ∈ [ 2.4 ; 2.48 ], for which the burst of spikes ends at ΞΌx⁒(t)=0subscriptπœ‡π‘₯𝑑0\mu_{x}(t)=0italic_ΞΌ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_t ) = 0 (red dot, when the neuron re-polarizes). Fig. 4b shows an example for b∈[2.64;2.72]𝑏2.642.72b\in[2.64;2.72]italic_b ∈ [ 2.64 ; 2.72 ], whose re-polarization occurring at ΞΌx⁒(t)<0subscriptπœ‡π‘₯𝑑0\mu_{x}(t)<0italic_ΞΌ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_t ) < 0 clearly indicates a signature of regime C𝐢Citalic_C (cf. Fig. 1), and not D𝐷Ditalic_D. Other intervals Absubscript𝐴𝑏A_{b}italic_A start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT are systematically checked using the same criterion, to study the persistence of D𝐷Ditalic_D.

This study identifies in Abβˆ—=[2.4;2.56]superscriptsubscript𝐴𝑏2.42.56A_{b}^{*}=[2.4;2.56]italic_A start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT βˆ— end_POSTSUPERSCRIPT = [ 2.4 ; 2.56 ] the robust interval for regime D𝐷Ditalic_D, corresponding to Pβˆ—=superscript𝑃absentP^{*}=italic_P start_POSTSUPERSCRIPT βˆ— end_POSTSUPERSCRIPT =40%, which is consistent with biological studies [de2008predicting] suggesting that, in experiments, square-wave bursts are more common than plateau bursting. The result for Abβˆ—superscriptsubscript𝐴𝑏A_{b}^{*}italic_A start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT βˆ— end_POSTSUPERSCRIPT is in line with studies in the literature using alternative methods [storace2008hindmarsh, barrio2011parameter, gonzalez2007complex]. The whole process took about 8.2 minutes to complete, with clear advantages over MC (recall that a single MC run on an arbitrary distribution of b𝑏bitalic_b takes around 20 minutes). It thus allows for rapid coarse analysis of the probabilistic robustness sets, while further refinements may look deeper into the neighborhood of the transition point.

Refer to caption
Figure 4: Expectation 𝔼⁒[x]=ΞΌx⁒(t)𝔼delimited-[]π‘₯subscriptπœ‡π‘₯𝑑\mathbb{E}[x]=\mu_{x}(t)blackboard_E [ italic_x ] = italic_ΞΌ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_t ) for (a) the interval b∈[2.4;2.48]𝑏2.42.48b\in[2.4;2.48]italic_b ∈ [ 2.4 ; 2.48 ], fully within regime D𝐷Ditalic_D, and (b) b∈[2.64;2.72]𝑏2.642.72b\in[2.64;2.72]italic_b ∈ [ 2.64 ; 2.72 ], fully outside regime D𝐷Ditalic_D. ΞΌxsubscriptπœ‡π‘₯\mu_{x}italic_ΞΌ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT was computed using the collocation method with NC=15subscript𝑁𝐢15N_{C}=15italic_N start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = 15 and SC=400subscript𝑆𝐢400S_{C}=400italic_S start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = 400. The red circles mark the end of the bursts of spikes at ΞΌx=0subscriptπœ‡π‘₯0\mu_{x}=0italic_ΞΌ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = 0, for regime D𝐷Ditalic_D (not happening for regime C𝐢Citalic_C).

IV Concluding Discussion

gPC methods are suitable candidates to accelerate probabilistic robustness studies. In fact, we have observed an acceleration of orders of magnitude with respect to Monte Carlo approaches. This comes with a significant drop in accuracy for the Galerkin approximation (which is however faster than collocation), while it is well-coped by the collocation method employing large numbers of samples (at the price of slower computation). Overall, our analysis provides a set of guidelines to choose the more suitable method and its hyper-parameters, depending on the desired trade-off between computing time and accuracy.

For the case study of interest, the HR model, which is characterised by multiple complex regimes, the tuned collocation method could rapidly identify the robustness region for the plateau bursting regime. The study highlighted a transition point to alternative regimes that is consistent with literature values, supporting the suitability of gPC surrogates to efficiently investigate complex models of biological interest. Future studies may expand the robustness analysis of all HR regimes, extend it to more complex models with unknown robustness properties, and generate additional insights based on the gPC properties.

\printbibliography