
Test 1 (a): SIR model in diffusive regime. First row: expectation (left) and standard deviation (right) obtained at
We study the Vlasov-Poisson-Fokker-Planck (VPFP) system with uncertainty and multiple scales. Here the uncertainty, modeled by multi-dimensional random variables, enters the system through the initial data, while the multiple scales lead the system to its high-field or parabolic regimes. We obtain a sharp decay rate of the solution to the global Maxwellian, which reveals that the VPFP system is decreasingly sensitive to the initial perturbation as the Knudsen number goes to zero. The sharp regularity estimates in terms of the Knudsen number lead to the stability of the generalized Polynomial Chaos stochastic Galerkin (gPC-SG) method. Based on the smoothness of the solution in the random space and the stability of the numerical method, we conclude the gPC-SG method has spectral accuracy uniform in the Knudsen number.
Citation: Yuhua Zhu. A local sensitivity and regularity analysis for the Vlasov-Poisson-Fokker-Planck system with multi-dimensional uncertainty and the spectral convergence of the stochastic Galerkin method[J]. Networks and Heterogeneous Media, 2019, 14(4): 677-707. doi: 10.3934/nhm.2019027
[1] | Robert Carlson . Myopic models of population dynamics on infinite networks. Networks and Heterogeneous Media, 2014, 9(3): 477-499. doi: 10.3934/nhm.2014.9.477 |
[2] | Shi Jin, Min Tang, Houde Han . A uniformly second order numerical method for the one-dimensional discrete-ordinate transport equation and its diffusion limit with interface. Networks and Heterogeneous Media, 2009, 4(1): 35-65. doi: 10.3934/nhm.2009.4.35 |
[3] | Sabrina Bonandin, Mattia Zanella . Effects of heterogeneous opinion interactions in many-agent systems for epidemic dynamics. Networks and Heterogeneous Media, 2024, 19(1): 235-261. doi: 10.3934/nhm.2024011 |
[4] | Nastassia Pouradier Duteil . Mean-field limit of collective dynamics with time-varying weights. Networks and Heterogeneous Media, 2022, 17(2): 129-161. doi: 10.3934/nhm.2022001 |
[5] | Xu Yang, François Golse, Zhongyi Huang, Shi Jin . Numerical study of a domain decomposition method for a two-scale linear transport equation. Networks and Heterogeneous Media, 2006, 1(1): 143-166. doi: 10.3934/nhm.2006.1.143 |
[6] | Christian Budde, Marjeta Kramar Fijavž . Bi-Continuous semigroups for flows on infinite networks. Networks and Heterogeneous Media, 2021, 16(4): 553-567. doi: 10.3934/nhm.2021017 |
[7] | Rong Huang, Zhifeng Weng . A numerical method based on barycentric interpolation collocation for nonlinear convection-diffusion optimal control problems. Networks and Heterogeneous Media, 2023, 18(2): 562-580. doi: 10.3934/nhm.2023024 |
[8] | Raul Borsche, Axel Klar, T. N. Ha Pham . Nonlinear flux-limited models for chemotaxis on networks. Networks and Heterogeneous Media, 2017, 12(3): 381-401. doi: 10.3934/nhm.2017017 |
[9] | Paulo Amorim, Alessandro Margheri, Carlota Rebelo . Modeling disease awareness and variable susceptibility with a structured epidemic model. Networks and Heterogeneous Media, 2024, 19(1): 262-290. doi: 10.3934/nhm.2024012 |
[10] | Elisabeth Logak, Isabelle Passat . An epidemic model with nonlocal diffusion on networks. Networks and Heterogeneous Media, 2016, 11(4): 693-719. doi: 10.3934/nhm.2016014 |
We study the Vlasov-Poisson-Fokker-Planck (VPFP) system with uncertainty and multiple scales. Here the uncertainty, modeled by multi-dimensional random variables, enters the system through the initial data, while the multiple scales lead the system to its high-field or parabolic regimes. We obtain a sharp decay rate of the solution to the global Maxwellian, which reveals that the VPFP system is decreasingly sensitive to the initial perturbation as the Knudsen number goes to zero. The sharp regularity estimates in terms of the Knudsen number lead to the stability of the generalized Polynomial Chaos stochastic Galerkin (gPC-SG) method. Based on the smoothness of the solution in the random space and the stability of the numerical method, we conclude the gPC-SG method has spectral accuracy uniform in the Knudsen number.
Mathematical modeling in epidemiology has certainly experienced an impressive increase in recent times driven by the overwhelming effects of the COVID-19 pandemic. A large part of the research has been directed towards the construction of models capable of describing specific characteristics associated with the pandemic, in particular the presence of asymptomatic individuals, which were largely underestimated, especially in the early stages of the spread of the disease [14,23,38,36,41]. Another line of research was aimed at the construction of models capable of describing the spatial characteristics of the epidemic, in order to be able to properly assess the impact of containment measures, in particular with regard to the mobility on the territory and restrictions in areas with higher risk of infection [13,43,42,16,1,39]. Other approaches took into account the network structure of connections [10], social heterogeneity aspects [21], and the multiscale nature of the pandemic [6]. See also [37,32] and the survey [2] for some recent developments in epidemiological modelling based on the use of kinetic equations.
Regardless of the characteristics of the model, a common aspect concerns the uncertainty of the data, which, especially in the first phase of the pandemic, largely underestimated the number of infected individuals. More precisely, the difficulty in correctly identifying all infected individuals in the early stages of the pandemic, due to structural limitations in performing large-scale screening and the inability to track the number of contacts, required the introduction of stochastic parameters into the models and the construction of related techniques for quantifying uncertainty [3,4,9,7].
Among the various techniques of uncertainty quantification, the approaches based on stochastic strategies that do not necessarily require a-priori knowledge of the probability distribution of the uncertain parameters, as needed in the case of methods based on generalized polynomial chaos [46], are particularly interesting in view of a comparison with experimental data. On the other hand, the low convergence rate of Monte Carlo type sampling techniques poses serious limitations to their practical use.
In this context, multi-fidelity methods have shown to be able to efficiently alleviate such limitations through control variate techniques based on an appropriate use of low-fidelity surrogate models able to accelerate the convergence of stochastic sampling [48,35,30,18,19]. Specifically, general multi-fidelity approaches for kinetic equations have been developed in [18,19], while bi-fidelity techniques with greedy sample selection in [30,31]. We refer also to the recent survey in [17].
The present paper is devoted to extend the bi-fidelity method for transport equations in the diffusive limit developed in [30] to the case of compartmental systems of multiscale equations designed to model mobility dynamics in an epidemic setting with uncertainty [13,7]. For this purpose, the corresponding hyperbolic system recently introduced in [10,9] will be used as a reduced low-fidelity model. The two models allow to correctly describe the hyperbolic dynamics of the movement of individuals over long distances together with the small-scale, high-density, diffusive nature typical of urban areas [13]. In addition, both models share the same diffusive limit in which it is possible to recover classical models of diffusive type [26]. From a numerical viewpoint, a space-time asymptotic-preserving discretization, which work uniformly in all regimes, has been adopted in combination with the bi-fidelity approach. This permits to obtain efficient stochastic asymptotic-preserving methods.
The rest of the manuscript is organized as follows. In Section 2 we introduce the epidemic transport model with uncertainty in the simplified SIR compartmental setting, together with its diffusive limit. The corresponding reduced order two-velocity model used as low-fidelity surrogate is also discussed. Next, we extend the previous modeling to more realistic compartmental settings based on the introduction of the exposed and asymptomatic compartments in Section 3. The details of the bi-fidelity method and the asymptotic-preserving IMEX Finite Volume scheme are then given in Section 4. Section 5 contains several numerical experiments that illustrate the performance of the bi-fidelity approach. Some conclusions are reported at the end of the manuscript.
For simplicity, we first illustrate the modeling in the case of a classic SIR compartmental dynamic and subsequently we will extend our arguments to a more realistic SEIAR model, designed to take into account specific features of the COVID-19 pandemic, in Section 3.
Let us consider a random vector
The kinetic distribution is then given by
$ f(x,v,t,{\bf z}) = f_S(x,v,t,{\bf z})+f_I(x,v,t,{\bf z})+f_R(x,v,t,{\bf z}), $ |
and we recover their total density by integration over the velocity space
$ \rho(x,t,{\bf z}) = \int_{-1}^{1} f(x,v_*,t,{\bf z})\,dv_*. $ |
As a consequence,
$ S(x,t,z)=∫1−1fS(x,v,t,z)dvI(x,t,z)=∫1−1fI(x,v,t,z)dvR(x,t,z)=∫1−1fR(x,v,t,z)dv, $
|
(1) |
with
$ ∂fS∂t+vS∂fS∂x=−F(fS,I)+1τS(S2−fS)∂fI∂t+vI∂fI∂x=F(fS,I)−γfI+1τI(I2−fI)∂fR∂t+vR∂fR∂x=γfI+1τR(R2−fR), $
|
(2) |
where
$ F(g,I)=βgIp1+κI, $
|
(3) |
with the classic bi-linear case corresponding to
The standard threshold of epidemic models is the well-known reproduction number
$ \frac{\partial}{\partial t} \int_{\Omega} I(x,t,{\bf z})\,dx = \int_{\Omega} F(S,I)\,dx-\int_{\Omega} \gamma(x,{\bf z}) I(x,t,{\bf z})\,dx \geq 0 $ |
when
$ R0(t,z)=∫ΩF(S,I)dx∫Ωγ(x,z)I(x,t,z)dx≥1. $
|
(4) |
The above quantity, therefore, defines the stochastic reproduction number for system (2) describing the space averaged instantaneous variation of the number of infectious individuals at time
Let us introduce the flux functions
$ JS=λS∫1−1vfS(x,v,t,z)dv,JI=λI∫1−1vfI(x,v,t,z)dv,JR=λR∫1−1vfR(x,v,t,z)dv. $
|
(5) |
Then, integrating the system (2) against
$ ∂S∂t+∂JS∂x=−F(S,I)∂I∂t+∂JI∂x=F(S,I)−γI∂R∂t+∂JR∂x=γI $
|
(6) |
whereas the flux functions satisfy
$ ∂JS∂t+λS2∫1−1v2∂fS∂xdv=−F(JS,I)−JSτS∂JI∂t+λI2∫1−1v2∂fI∂xdv=−λIλSF(JS,I)−γJI−JIτI∂JR∂t+λR2∫1−1v2∂fR∂xdv=−λRλIγJI−JRτR. $
|
(7) |
By introducing the space dependent diffusion coefficients
$ DS=13λ2SτS,DI=13λ2IτI,DR=13λ2RτR, $
|
(8) |
which characterize the diffusive transport mechanism of susceptible, infectious and removed, respectively, and keeping the above quantities fixed while letting the relaxation times
$ fS=S/2,fI=I/2,fR=R/2, $
|
and consequently, from (7), we obtain a proportionality relation between the fluxes and the spatial derivatives (Fick's law):
$ JS=−DS∂S∂x,JI=−DI∂I∂x,JR=−DR∂R∂x. $
|
(9) |
Thus, substituting (9) into (6) we get the diffusion system [34,40,45,26]
$ ∂S∂t=−F(S,I)+∂∂x(DS∂S∂x)∂I∂t=F(S,I)−γI+∂∂x(DI∂I∂x)∂R∂t=γI+∂∂x(DR∂R∂x). $
|
(10) |
We remark that the model's capability to account for different regimes, hyperbolic or parabolic, accordingly to the space dependent values
The low-fidelity model, is based on considering individuals moving in two opposite directions (indicated by signs "+" and "-"), with velocities
$ ∂S±∂t+λS∂S±∂x=−F(S±,I)∓12τS(S+−S−)∂I±∂t+λI∂I±∂x=F(S±,I)−γI±∓12τI(I+−I−)∂R±∂t+λR∂R±∂x=γI±∓12τR(R+−R−). $
|
(11) |
In the above system, individuals
$ S=S++S−,I=I++I−,R=R++R−. $
|
The transmission of the infection is governed by the same incidence function as in the high-fidelity model, defined in (3). Also the definition of the basic reproduction number
If we now introduce the fluxes, defined by
$ JS=λS(S+−S−),JI=λI(I+−I−),JR=λR(R+−R−), $
|
(12) |
we obtain a hyperbolic model equivalent to (11), but presenting a macroscopic description of the propagation of the epidemic at finite speeds, for which the densities follow system (6) and equations of fluxes read
$ ∂JS∂t+λ2S∂S∂x=−F(JS,I)−JSτS∂JI∂t+λ2I∂I∂x=λIλSF(JS,I)−γJI−JIτI∂JR∂t+λ2R∂R∂x=λRλIγJI−JRτR. $
|
(13) |
Let us now consider the behavior of the low-fidelity model in diffusive regimes. To this aim, we introduce the diffusion coefficients
$ DS=λ2SτS,DI=λ2IτI,DR=λ2RτR. $
|
(14) |
As for the previous model, the diffusion limit of the system is formally recovered letting the relaxation times
Remark 1. An important aspect in the bi-fidelity approach here proposed, is that the high-fidelity model and the low-fidelity one exactly coincide in the diffusive limit. The only difference lays in the definition of the diffusion coefficients (see eqs. (8) and eqs. (14)). As a consequence, in such a regime the bi-fidelity method achieves the maximum accuracy.
To account for more realistic models to analyze the evolution of the ongoing COVID-19 pandemic, we consider extending the simple SIR compartmentalization by taking into account two additional population compartments,
Let us now consider a population at position
$ f(x,v,t,{\bf z}) = f_S(x,v,t,{\bf z})+f_E(x,v,t,{\bf z})+f_I(x,v,t,{\bf z})+f_A(x,v,t,{\bf z})+f_R(x,v,t,{\bf z}). $ |
In addition to (1), we have that
$ E(x,t,z)=∫1−1fE(x,v,t,z)dv,A(x,t,z)=∫1−1fA(x,v,t,z)dv. $
|
In this setting, the kinetic densities satisfy the following transport equations [7]
$ ∂fS∂t+vS∂fS∂x=−F(fS,I)−FA(fS,A)+1τS(S2−fS)∂fE∂t+vE∂fE∂x=F(fS,I)+FA(fS,A)−afE+1τE(E2−fE)∂fI∂t+vI∂fI∂x=aσfE−γIfI+1τI(I2−fI)∂fA∂t+vA∂fA∂x=a(1−σ)fE−γAfA+1τA(A2−fA)∂fR∂t+vR∂fR∂x=γIfI+γAfA+1τR(R2−fR) $
|
(15) |
The quantities
$ FA(g,A)=βAgAp1+κAA, $
|
(16) |
where a different contact rate,
$ R0(t,z)=∫ΩFI(S,I)dx∫ΩγI(x,z)I(x,t,z)dx⋅∫Ωa(x,z)σ(x,z)E(x,t,z)dx∫Ωa(x,z)E(x,t,z)dx+∫ΩFA(S,A)dx∫ΩγA(x,z)A(x,t,z)dx⋅∫Ωa(x,z)(1−σ(x,z))E(x,t,z)dx∫Ωa(x,z)E(x,t,z)dx, $
|
(17) |
the reader is invited to refer to [9,7].
When introducing the same definition (5) of flux for the additional compartments,
$ ∂S∂t+∂JS∂x=−F(S,I)−FA(S,A)∂E∂t+∂JE∂x=F(S,I)+FA(S,A)−aE∂I∂t+∂JI∂x=aσE−γII∂A∂t+∂JA∂x=a(1−σ)E−γAA∂R∂t+∂JR∂x=γII+γAA $
|
(18) |
and for the fluxes
$ ∂JS∂t+λS2∫1−1v2∂fS∂xdv=−F(JS,I)−FA(JS,A)−JSτS∂JE∂t+λE2∫1−1v2∂fE∂xdv=λEλS(F(JS,I)+FA(JS,A))−aJE−JEτE∂JI∂t+λI2∫1−1v2∂fI∂xdv=λIλEaσJE−γIJI−JIτI∂JA∂t+λA2∫1−1v2∂fA∂xdv=λAλEa(1−σ)JE−γAJA−JAτA∂JR∂t+λR2∫1−1v2∂fR∂xdv=λRλIγIJI+λRλAγAJA−JRτR. $
|
(19) |
Moreover, defining also
$ ∂S∂t=−F(S,I)−FA(S,A)+∂∂x(DS∂S∂x)∂E∂t=F(S,I)+FA(S,A)−aE+∂∂x(DE∂E∂x)∂I∂t=aσE−γII+∂∂x(DI∂I∂x)∂A∂t=a(1−σ)E−γAA+∂∂x(DA∂A∂x)∂R∂t=γII+γAA+∂∂x(DR∂R∂x). $
|
(20) |
The low-fidelity model is obtained, also with the more complex SEIAR compartmentalization, considering the discrete-velocity case with only two opposite velocities
$ ∂S±∂t±λS∂S±∂x=−F(S±,I)−FA(S±,A)+12τS(S∓−S±)∂E±∂t±λE∂E±∂x=F(S±,I)+FA(S±,A)−aE±+12τE(E∓−E±)∂I±∂t±λI∂I±∂x=aσE±−γII±+12τI(I∓−I±)∂A±∂t±λA∂A±∂x=a(1−σ)E±−γAA±+12τA(A∓−A±)∂R±∂t±λR∂R±∂x=γII±+γAA±+12τR(R∓−R±). $
|
(21) |
When defining fluxes
$ ∂JS∂t+λ2S∂S∂x=−F(JS,I)−FA(JS,A)−JSτS∂JE∂t+λ2E∂E∂x=λEλS(F(JS,I)+FA(JS,A))−aJE−JEτE∂JI∂t+λ2I∂I∂x=λIλEaσJE−γIJI−JIτI∂JA∂t+λ2A∂A∂x=λAλEa(1−σ)JE−γAJA−JAτA∂JR∂t+λ2R∂R∂x=λRλIγIJI+λRλAγAJA−JRτR. $
|
(22) |
As for the previous cases, the diffusion limit of the system is formally recovered letting the relaxation times
In this Section, we present the details of the numerical method used to solve the stochastic problem following the bi-fidelity asymptotic-preserving scheme proposed in [35,48]. For the high-fidelity model, the numerical scheme is structured in agreement with a discrete ordinate method in velocity with the even and odd parity formulation [20,28]. Both the high-fidelity and low-fidelity solvers use Finite Volume Method (FVM) in space and achieve asymptotic preservation in time using suitable IMEX Runge-Kutta schemes [11,12]. This permits to obtain a numerical scheme able to deal with the diffusion limit of the mathematical models without loosing consistency, and for which the time step size of the temporal discretization is not subject to excessive restrictions related to the smallness of the scaling parameters
In this Section, we present the AP-IMEX Finite Volume scheme adopted to solve the SIR model at each stochastic collocation point selected for the bi-fidelity approximation.
The asymptotic-preserving IMEX method and the corresponding even and odd parities formulation was introduced in [13] for an SIR kinetic transport model in 2D domains. According to [28,24], for
$ rS(v)=12(fS(v)+fS(−v)),jS(v)=λS2(fS(v)−fS(−v)),rI(v)=12(fI(v)+fI(−v)),jI(v)=λI2(fI(v)−fI(−v)),rR(v)=12(fR(v)+fR(−v)),jR(v)=λR2(fR(v)−fR(−v)). $
|
An equivalent formulation of (2) can be written as
$ ∂rS∂t+v∂jS∂x=−F(rS,I)+1τS(12S−rS)∂rI∂t+v∂jI∂x=F(rS,I)−γrI+1τI(12I−rI)∂rR∂t+v∂jR∂x=γrI+1τR(12R−rR)∂jS∂t+λ2Sv∂rS∂x=−F(jS,I)−1τSjS∂jI∂t+λ2Iv∂rI∂x=F(jS,I)−γjI−1τIjI∂jR∂t+λ2Rv∂rR∂x=γjI−1τRjR, $
|
(23) |
where
$ S=2∫10rSdv,I=2∫10rIdv,R=2∫10rRdv. $
|
(24) |
The above densities can be approximated by a Gauss-Legendre quadrature rule. This leads to a discrete velocity setting, usually referred to as the discrete ordinate method, where we approximate
$ S≈SM=NG∑i=1wirS(ζi)I≈IM=NG∑i=1wirI(ζi)R≈RM=NG∑i=1wirR(ζi) $
|
where
Assuming for simplicity of notation that
$ ∂r∂t+v∂j∂x=E(r)−1τ(r−R2)∂j∂t+Λ2v∂r∂x=E(j)−1τj, $
|
(25) |
with
$ \boldsymbol{r} = \left( r_S,r_I,r_R \right)^T, \quad \boldsymbol{j} = \left( j_S,j_I,j_R \right)^T, \quad \boldsymbol{R} = \left( S,I,R \right)^T,\quad \boldsymbol{\Lambda} = \mathrm{diag}\{\lambda_S,\lambda_I,\lambda_R\}, $ |
$ \boldsymbol{E}(\boldsymbol{r}) = \left( - F(r_S, I), F(r_S, I)- \gamma r_I, \gamma r_I \right)^T ,\quad \boldsymbol{E}(\boldsymbol{j})\left( - F(j_S, I), F(j_S, I)- \gamma j_I, \gamma j_I \right)^T. $ |
Following [12], the Implicit-Explicit (IMEX) Runge-Kutta discretization that we consider for system (25) consists in computing the internal stages
$ r(k)=rn−Δtk∑j=1akj(v∂j(j)∂x+1τ(r(j)−R(j)2))+Δtk−1∑j=1˜akjE(r(j))j(k)=jn−Δtk−1∑j=1˜akj(Λ2v∂r(j)∂x−E(j(j)))−Δtk∑j=1akj1τj(j), $
|
(26) |
and evaluating the final numerical solution
$ rn+1=rn−Δts∑k=1bk(v∂j(k)∂x+1τ(r(k)−R(k)2))+Δts∑k=1˜bkE(r(k))jn+1=jn−Δts∑k=1˜bk(Λ2v∂r(k)∂x−E(j(k)))−Δts∑k=1bk1τj(k). $
|
(27) |
To properly compute the implicit terms
$ akj=bj,j=1,…,s,˜akj=˜bj,j=1,…,s−1, $
|
the method is said to be globally stiffly accurate (GSA). It is worth to notice that this definition states also that the numerical solution of a GSA IMEX Runge-Kutta scheme coincides exactly with the last internal stage of the scheme. Since the GSA property is fundamental to preserve the correct diffusion limit and to achieve asymptotic-preservation stability in stiff regimes [13,10], in the sequel, the GSA BPR(4, 4, 2) scheme presented in [12] is chosen, characterized by
At each internal stage of the IMEX scheme (26), we apply a Total-Variation-Diminishing (TVD) Finite Volume discretization to evaluate the numerical fluxes [10,8]. To achieve second order accuracy also in space, while avoiding the occurrence of spurious oscillations, a classical minmod slope limiter has been adopted.
The same AP-IMEX Finite Volume scheme is adopted also to solve the low-fidelity SIR model (6)-(13), as fully presented in [10]. The reader is invited to refer to [10,9,7,13] for further details on the properties of the chosen numerical scheme applied to epidemic models.
In this Section, we briefly review the bi-fidelity method developed in [35,48]. Bi-fidelity methods make use of low-fidelity models to effectively inform the selection of representative points in the parameter space and then employ this information to construct accurate approximations to high-fidelity solutions. To facilitate future discussion, we denote the expensive high-fidelity solution
The basic idea of the bi-fidelity approximation is to construct an inexpensive surrogate
$ uB(z)=n∑k=1ck(z)uH(zk), $
|
(28) |
where
There are two major key questions for the performance of the above bi-fidelity algorithms: (a) how to select the representative points
Subset Selection. Existing predefined or structured nodes (e.g., sparse grids, cubature rules, etc.) often grows fast in high dimensions. Therefore, these options cannot easily accommodate the current situation where we would like the size
We shall identify important points iteratively by a greedy approach [35,48]. Specifically, we denote a candidate sample set
$ zik+1=argmaxz∈ΓNdL(uL(z),UL(γk)),γk+1=γk∪zik+1, $
|
(29) |
where
Bi-fidelity approximation. For any given
$ uL(z)≈PUL(γn)uL(z)=n∑k=1cLk(z)uL(zk),zk∈γn, $
|
(30) |
where the low-fidelity projection coefficients can be computed as follows:
$ GLcL=f,fL=(fLk)1≤k≤n,fLk=⟨uL(z),uL(zk)⟩, $
|
(31) |
and
$ (GL)ij=⟨uL(zi),uL(zj)⟩L,1≤i,j≤n, $
|
(32) |
where
Under certain conditions [35,48],
$ uB(z)=n∑k=1cLk(z)uH(zk). $
|
(33) |
Algorithm 1:A bi-fidelity approximation for a high-fidelity solution at given |
1Given a candidate sample set |
2Select |
3Run high-fidelity simulation only at the point in the selected sample set |
4For any given |
$ |
5Construct the bi-fidelity approximation by applying the same approximation rule as in low-fidelity model with (33): |
$ |
To put things together, we outline the major steps for the bi-fidelity approximation of the high-fidelity sample for a given
Remark 2. As we mentioned above, the bi-fidelity method relies on the assumption that the low-fidelity coefficients are similar to the high-fidelity coefficients in the parameter space under certain conditions, as stated in [35,48]. This may not hold for some problems. In this case, a correction mapping between low-fidelity and high-fidelity coefficient can be constructed by leveraging approximation power of neural network to further improve the accuracy, if the additional high-fidelity data are available. We refer readers to [33] for details on this approach.
To construct the bi-fidelity approximation of the high-fidelity mean, with the bi-fidelity surrogate
● Compute the low-fidelity sample mean (via the Monte Carlo or quadrature rules):
$ μL=M∑i=1wiuL(zi), $
|
(34) |
where the high-fidelity mean
● Compute its best approximation on the low-fidelity approximation space
$ μL≈PUL(γn)μL=n∑i=1cLiuL(zi), $
|
(35) |
where the expansion coefficients
$ GLcL=gL,gL=⟨μL,uL(zj)⟩L,1≤j≤n,zj∈γn. $
|
(36) |
● Using this coefficient
$ μB=n∑k=1cLkuH(zk),zk∈γn. $
|
(37) |
Note that in this way, only one bi-fidelity surrogate evaluation is required. The standard deviation can be computed similarly. We refer readers to [47] for additional details.
To examine the performance of the proposed methodology, two benchmark tests are considered: the first concerning the kinetic transport model with the SIR compartmentalization discussed in Section 2, and the second one regarding the extension to the SEIAR modeling presented in Section 3. It is worth to highlight that, in these tests, we collect all the quantities of interests together in a single vector to choose the
We first consider the initial distributions of the high-fidelity kinetic SIR model (2) as follows:
$ fi(x,v,0)=ci(x,0)e−v22,i∈{S,I,R} $
|
(38) |
where
$ S(x,0) = 1 - I(x,0), \qquad I(x,0) = 0.01 e^{-(x-10)^2}, \qquad R(x,0) = 0, $ |
in the physical domain
We analyze the behavior of the proposed methodology concerning spatially heterogeneous environments, considering a spatially variable contact rate [10,44]
$ β(x,z)=β0(z)(1+0.05sin(13πx20)), $
|
where
$ \beta^0({\bf{z}}) = 11(1 + 0.6 z_1), $ |
and a recovery rate
$ \gamma({\bf{z}}) = 10(1 + 0.4 z_2). $ |
In the incidence function, we set
Test 1 (a): In this case, a parabolic configuration of speeds and relaxation parameters is considered, setting
In the first row of Figure 1, the expectation and standard deviation of the solution of compartment
Test 1 (a): SIR model in diffusive regime. First row: expectation (left) and standard deviation (right) obtained at
Test 1 (b): In this second case, we consider the hyperbolic regime by letting
In Figure 2, the results of Test 1(b) are reported for the infectious compartment
Test 1 (b): SIR model in hyperbolic regime. First row: expectation (left) and standard deviation (right) obtained at
Next, we analyze the effectiveness of the proposed methodology also with the extended SEIAR compartmentalization examining a more realistic epidemic scenario. Let us now consider a 2-dimensional random vector
$ E(x,0,{\bf z}) = \alpha_1({\bf z})\,e^{-(x-x_1)^2} + \alpha_2({\bf z})\,e^{-(x-x_2)^2} + \alpha_3({\bf z})\,e^{-(x-x_3)^2}, $ |
where
$ S(x,0,{\bf z}) = 1 - E(x,0,{\bf z}), \qquad I(x,0) = 0, \qquad A(x,0) = 0, \qquad R(x,0) = 0, $ |
and
Concerning epidemic parameters, to simulate the more challenging scenario in which the incidence function presents sinusoidal oscillations in space as well as being greater in the most populated areas, we consider the following distribution of the contact rate related to asymptomatic individuals:
$ βA(x,z)=β0A(z)(1+12e−(x−x1)2+14e−(x−x2)2+12e−(x−x3)2)+0.05sin(2πx), $
|
where
$ \beta_A^0({\bf{z}}) = 0.5(1 + 0.5 z_2). $ |
Assuming that highly infectious subjects are mostly detected in the most optimistic scenario, being subsequently quarantined or hospitalized, we set
Test 2 (a): In the first case, an intermediate regime between parabolic and hyperbolic is considered, setting
The baseline space-time evolution of the compartments is shown in Figure 3, where the fast propagation of the epidemic can be observed, especially starting from the first city on the left, as expected. In Figure 4, the expectation and standard deviation of infected individuals, divided by each compartment, are presented for each methodology adopted, highlighting the validity of the bi-fidelity approach. In Figure 5,
Test 2 (b): In the second case, we consider a fully hyperbolic regime choosing
The baseline space-time evolution of the compartments of major interest is presented in Figure 6. In this figure, the spatial heterogeneity of the epidemic spread related to the prescribed contact rate can be clearly appreciated. Mean and standard deviation obtained with the 3 approaches are shown for compartments
In this work we introduced a bi-fidelity method for the quantification of uncertainty in epidemiological transport models based on an asymptotic-preserving space-time discretization. In detail, after presenting the high-fidelity epidemiological model, we considered the corresponding bi-fidelity model. Both models share the same diffusive limit and permits to recover classical epidemic models based on diffusion equations. The numerical scheme used allows to have an efficient quantification of the uncertainty in different regimes, using few simulations of the high-fidelity model and several runs of the low-fidelity model for points selection in the random space. Results for one-dimensional transport problems based on realistic compartmentalization in relation to the recent COVID-19 pandemic, which also include asymptomatic individuals, show the validity of the presented approach. Further research will be directed toward extending the present approach to realistic contexts such as those studied in [7,9].
G.B. and L.P. acknowledge the support of MIUR PRIN 2017, Project No. 2017KKJP4X Innovative numerical methods for evolutionary partial differential equations and applications. G.B. holds a Research Fellowship from the Italian National Institute of High Mathematics, INdAM (GNCS). L.L. holds the Direct Grant for Research supported by Chinese University of Hong Kong and Early Career Scheme 2021/22, No. 24301021, from Research Grants Council of Hong Kong. X.Z. was supported by the Simons Foundation (504054).
1. | Giulia Bertaglia, Chuan Lu, Lorenzo Pareschi, Xueyu Zhu, Asymptotic-Preserving Neural Networks for multiscale hyperbolic models of epidemic spread, 2022, 32, 0218-2025, 1949, 10.1142/S0218202522500452 | |
2. | Rinaldo M. Colombo, Mauro Garavello, Infectious Disease Spreading Fought by Multiple Vaccines Having a Prescribed Time Effect, 2023, 71, 0001-5342, 10.1007/s10441-022-09452-4 | |
3. | Giulia Bertaglia, Andrea Bondesan, Diletta Burini, Raluca Eftimie, Lorenzo Pareschi, Giuseppe Toscani, New trends on the systems approach to modeling SARS-CoV-2 pandemics in a globally connected planet, 2024, 34, 0218-2025, 1995, 10.1142/S0218202524500301 | |
4. | Giulia Bertaglia, 2023, Chapter 2, 978-3-031-29874-5, 23, 10.1007/978-3-031-29875-2_2 | |
5. | Giulia Bertaglia, Lorenzo Pareschi, Multiscale Constitutive Framework of One-Dimensional Blood Flow Modeling: Asymptotic Limits and Numerical Methods, 2023, 21, 1540-3459, 1237, 10.1137/23M1554230 | |
6. | Xiaojing Xu, Wenjun Sun, Qi Li, The asymptotic preserving unified gas kinetic scheme for the multi-scale kinetic SIR epidemic model, 2024, 174, 08981221, 298, 10.1016/j.camwa.2024.09.021 | |
7. | Liu Liu, 2023, Chapter 7, 978-3-031-29874-5, 139, 10.1007/978-3-031-29875-2_7 | |
8. | Giulia Bertaglia, Lorenzo Pareschi, Giuseppe Toscani, Modelling contagious viral dynamics: a kinetic approach based on mutual utility, 2024, 21, 1551-0018, 4241, 10.3934/mbe.2024187 | |
9. | Giulia Bertaglia, Lorenzo Pareschi, Russel E. Caflisch, Gradient-Based Monte Carlo Methods for Relaxation Approximations of Hyperbolic Conservation Laws, 2024, 100, 0885-7474, 10.1007/s10915-024-02614-1 |