Home Application of fluid dynamics in modeling the spatial spread of infectious diseases with low mortality rate: A study using MUSCL scheme
Article Open Access

Application of fluid dynamics in modeling the spatial spread of infectious diseases with low mortality rate: A study using MUSCL scheme

  • Daniel Ugochukwu Nnaji EMAIL logo , Phineas Roy Kiogora , Ifeanyi Sunday Onah , Joseph Mung’atu and Nnaemeka Stanley Aguegboh
Published/Copyright: December 31, 2024

Abstract

This study presents a comprehensive mathematical framework that applies fluid dynamics to model the spatial spread of infectious diseases with low mortality rates. By treating susceptible, infected, and treated population densities as fluids governed by a system of partial differential equations, the study simulates the epidemic’s spatial dynamics. The Monotone Upwind Scheme for Conservation Laws is employed to enhance the accuracy of numerical solutions, providing a high-resolution approach for capturing disease transmission patterns. The model’s analogy between fluid flow and epidemic propagation reveals critical insights into how diseases disperse geographically, influenced by factors like human mobility and environmental conditions. Numerical simulations show that the model can predict the evolution of infection and treatment population densities over time, offering practical applications for public health strategies. Sensitivity analysis of the reproduction number highlights the influence of key epidemiological parameters, guiding the development of more efficient disease control measures. This work contributes a novel perspective to spatial epidemiology by integrating principles of fluid dynamics, aiding in the design of targeted interventions for controlling disease outbreaks.

MSC 2010: 92D30; 35Q35; 76S05

1 Introduction

A mathematical framework used to comprehend and forecast the spread of diseases over various geographic areas is referred to as a spatial epidemiological model [49]. The incorporation of spatial data into epidemiological models has become increasingly important due to the need to understand how geographical factors influence the spread and impact of diseases. Spatial epidemiological models provide critical insights into disease dynamics, allowing public health professionals to design better interventions and allocate resources more efficiently. Traditional epidemiological models frequently assume a well-mixed population; in contrast, spatial models take into account the movement and spatial dispersion of individuals or communities. This allows for better control and prevention of disease outbreaks by incorporating spatial data with disease transmission mechanisms to plan and implement public health interventions more effectively.

The early development of Bayesian hierarchical models and related computation techniques was centered around spatial models. Standard analysis techniques were ineffective in a variety of sectors, including disease mapping, agriculture, ecology, and image analysis [24]. The interplay of geography and infectious disease has been given substantial attention in what is termed spatial epidemiology [17,41]. Although early interest in the spatial aspects of infectious disease distribution was primarily descriptive in nature, current research in spatial epidemiology has emphasized more on applied objectives [55].

Infectious diseases are always challenged by many environmental and social factors in their spread [47]. Many elements that can affect the dynamics of infectious disease transmission are taken into consideration by these models, including social interactions, movement patterns, population density, and geographic layout. Spatial analysis can provide an intuitive understanding of the dynamics of epidemics by converting raw data into actionable information. Spatial epidemiological models can help plan and implement public health interventions more effectively by combining spatial data with disease transmission mechanisms [48].

The goal of spatial epidemiology is to condense the geographical aspects of epidemiological processes into a realistic and actionable framework for understanding infectious diseases [6,11,38,46,61]. Such an approach is of interest since many relevant factors underlying disease transmission are largely and inherently spatially defined. Moreover, in some cases, understanding the spatial characteristics is crucial for implementing effective interventions to mitigate disease spread.

A particular focus of this framework is to predict how interactions between human society, the causative pathogen, and its transmission vector dynamics (epidemic flow) work within a one-dimensional space [13,30,53]. In infectious disease epidemiological modeling, taking spatially heterogeneous mixing and interventions into account is crucial. A major factor in the spread of infections to humans is human mobility and how it affects regional contact patterns, particularly when it comes to airborne infectious pathogens like influenza and corona-viruses [10,18,30,36].

One method to simulate human mobility and its influence on the transmission of airborne pathogens in a heterogeneously mixed population is to divide the population into smaller subgroups based on different levels of activity. The effects of human mobility on the spread of airborne pathogens within this context are analyzed using a coupled system of partial differential equations (PDEs) at each residence, employing a susceptible-infected-removed model [15,61]. Today, the spatial spread of infectious diseases is affected by many factors, including the mobility of infected individuals and the flow of susceptible individuals into areas with high disease prevalence.

The aim of this study is to introduce a novel mathematical modeling framework that integrates the principles of fluid dynamics into the analysis of infectious disease spread and transmission. We refer to this concept as “epidemic flow,” which models the transmission of an epidemic analogous to fluid motion. This approach is inspired by previous research in traffic flow that applied fluid dynamics to simulate vehicle movement as a driving force [29,34,43,59]. In this model, the population of infected or infectious hosts is treated as a fluid, with each sick individual representing a fluid element. Our research primarily focuses on a macroscopic perspective, disregarding the specifics of individual components in favor of viewing them as part of a continuum that captures the epidemic’s overall behavior. Specifically, the understanding of principle of displacement of fluid in motion provides insights into the spatial distribution of the disease, while the velocity of fluid motion reflects the rate of epidemic spread. This epidemic flow modeling allows us to predict critical characteristics of the epidemic’s geographic dynamics, offering valuable guidance for the development of effective disease management strategies.

The concept of integrating fluid dynamics into epidemiology, and analogizing an epidemic model to inviscid flood flow, was first introduced by Cheng and Wang [13,14]. In this context, the term “density” was used by the author to describe the compartmentalized human population at a specific study location. This analogy is intended to illustrate the dynamics of disease transmission and human interactions, drawing from traffic flow models, without implying that humans are literally considered as fluid. The methodology involves the development of a single-phase epidemic flow model to examine the spatial spread of infectious diseases. This approach, while grounded in classical mathematical epidemiology using compartmental, population-level models, innovatively incorporates principles of fluid dynamics. Ultimately, spatial epidemiological models offer a robust framework for comprehending and managing the spread of infectious diseases. By accounting for factors such as human mobility, social interactions, and environmental conditions, these models facilitate more effective public health interventions and improved disease control strategies. As research in this field progresses, such models will become increasingly critical in the global effort to combat infectious diseases.

This study explores a one-dimensional extension of the model introduced in [14], by considering three inviscid reactive fluids, which not only increases the model’s complexity but also enhances its realism. Unlike the previous study that used high-order weighted essentially non-oscillatory methods to solve the resulting hyperbolic PDE, this work improves the accuracy of our numerical solution by applying the Monotone Upwind Scheme for Conservation Laws (MUSCL). The remainder of the study is structured as follows: Section 2 focuses on model formulation, integrating mass and momentum conservation laws to develop the spatial epidemic flow model. Section 3 addresses the spatial analysis of the proposed model and the linearized form of the PDE. Section 4 covers the numerical and graphical simulations for the proposed model. Finally, Section 5 discusses the epidemiological implications of the findings presented in this study.

2 Model description and formulation

We consider an infectious disease, which affects a certain human population at a location of interest. Let S = S ( t , x ) , I = I ( t , x ) , T = T ( t , x ) , and R = R ( t , x ) denote the population densities at time t and location x of the susceptible, infected, treated (infected individuals receiving treatment), and recovered class, respectively, where the total population density denoted by N = N ( t , x ) is written as

(2.1) N = S + I + T + R .

However, the individual populations S ( t , x ) , I ( t , x ) , T ( t , x ) , and R ( t , x ) are regarded as the susceptible, infected, treated (infected individuals who receives treatment), and recovered density population at a specific area of interest, otherwise called mass densities [14]. On the other hand, since individual characteristics are not considered here, rather aggregate properties, we treat each host as having the same (averaged) mass and disregard the variety of individual attributes because the macroscopic behavior of the epidemic is the focus of our study. Let us assume that the epidemic lasts for a short period, during which vital dynamics (births and natural deaths) are not included, and the disease mortality rate is low (e.g., Chickenpox, Influenza, Gastroenteritis, Pertussis, etc). Having that the disease-induced death rate is low, the overall population density N ( t , x ) can be roughly treated as a constant since we have presumed a relatively short time period for the epidemic. Consequently, the recovered population density R ( t , x ) can be disregarded in the spatial analysis due to its negligible impact, allowing us to focus exclusively on the variables S ( t , x ) , I ( t , x ) , and T ( t , x ) . We now describe the susceptible, infected, and treated population densities as three inviscid fluids that move, mix, and interact with one another [13,14]. Through this fluid-like contact, the density of the susceptible fluid falls, while the density of the infected (infectious) and treated fluid increases. Thus, the fluid-like motion as assumed in this work, illustrates how the disease spreads spatially. Let the velocity fields of the susceptible, infected, and treated population densities be denoted by the variables u ( t , x ) , v ( t , x ) , and w ( t , x ) , respectively. It is important to reiterate that these velocities are determined by the collective behavior of the density populations, and are defined at a macroscopic scale.

2.1 Model assumptions

Here, we outline the assumptions made for this model.

  1. The model is primarily concerned with the macroscopic behavior of the epidemic, meaning it looks at the large-scale dynamics rather than the detailed characteristics and heterogeneities of individual hosts.

  2. The velocity fields of the fluids are macroscopic and reflect the collective movement of the population densities rather than the motion of individual person.

  3. The total population density is assumed to be constant over short time periods due to the low death rate from the infection. This allows the model to ignore changes in total population size in the spatial domain.

  4. These fluids interact with each other, where susceptible individuals “convert” into infected individuals through contact ( β 1 , β 2 ), modeled as fluid interaction, continuously changing the densities of the populations over time and space.

  5. The susceptible population density initially occupies most of the space, while the infected population density is localized to small areas where the outbreak begins.

  6. It is assumed that the disease spreads from high-prevalence areas to low-prevalence areas, analogous to fluid flow moving from regions of high density to low density.

Let us assume we have a control volume where the quantity represented by the density ρ can be generated or consumed as well as flow in and out of the volume with velocity field q ¯ . If there is an additional source/sink term σ (representing generation or consumption per unit volume), the total rate of change of ρ within this volume can be described by the conservation principle [2,4,7]:

(2.2) t V ρ d V + ρ q ¯ n d = V σ d V ,

where V is the control volume, is the surface enclosing the control volume V , and n is the outward-pointing unit normal vector on the surface . Applying the divergence theorem, equation (2.2) becomes:

(2.3) ρ t + ( ρ q ¯ ) = σ .

Equation (2.3) is usually the general form for continuity equation [3], with ρ q ¯ representing the mass flux. Now, applying the one-dimensional version of equation (2.3) to the susceptible, infected, and treated population densities, respectively, together with their velocities, we have:

(2.4) S t + x ( S u ) = β 1 S I β 2 S T , I t + x ( I v ) = β 1 S I + β 2 S T ( τ + μ ) I , T t + x ( T w ) = τ I ( γ + μ ) T ,

where u , v , w denote the spatial velocities of the susceptible, infected, and treated fluid motion, respectively. These velocities does not depict the contact rate, rather, take into account how fast the epidemic spreads out spatially. With the following initial conditions: S ( 0 , x ) > 0 , I ( 0 , x ) 0 , T ( 0 , x ) 0 , and the Neumann boundary conditions: S ( t , 0 ) = 0 , I ( t , 0 ) = 0 , T ( t , 0 ) = 0 to S x ( t , L ) = 0 , I x ( t , L ) = 0 , T x ( t , L ) = 0 having that t [ 0 , ) and L Ω . This assumes that no individuals enter or leave the boundaries of the domain, meaning that the flux of the susceptible, infected, and treated individuals is zero at the boundaries.

On the other hand, the parameter β 1 denotes the contact rate when infected individuals comes in contact with susceptible individuals, β 2 denotes the contact rate when the infected individuals who are undergoing treatments come in contact with the susceptible persons (note that β 2 < β 1 ), τ represents the ratio of infected individuals who receives treatment, γ denotes the ratio of treated individuals who obtains recovered, while μ represents the disease mortality rate (which is assumed low).

We drew ideas from the work of Cheng and Wang [13,14], which expressed the mathematical form of the ideal gas equation [33,62],

(2.5) P = ℛT d

and incorporated it into the Euler’s equation of motion (having that pressure force is the only body force), and characterized the pressure of the ideal gas in terms of the state variables (S,I,T) [1,3,44,45,54]. Note that is the ideal gas constant, T is the temperature (which is assumed constant), and d is the density. Thus, we can regard the susceptible, infected, and treated individuals as independent “particles” whose movements are unaffected by other particles. On a macroscopic level, the movement of each population density can be likened to the flow of an ideal gas. Additionally, the distribution of fluid density influences the motion of each fluid, as demonstrated by Cheng and Wang [14]. Consequently, the spread of disease will tend to move from areas of high prevalence to those with lower frequency. At a constant temperature and for a given specific gas constant, the one-dimensional velocity field that describes the epidemic flow can then be expressed as [9,16,20,42,64]. The motion of the susceptible, infected, and treated population densities is dictated by the Euler equations from fluid dynamics, where their movement is driven by density gradients, similar to how fluid pressure operates. The velocity field is therefore expressed as

(2.6) S u t + S u u x = k s S x , I v t + I v v x = k i I x , T w t + T w w x = k t T x .

Equation (2.6) represents Euler’s equations of motion for an inviscid (non-viscous) fluid, where u , v , and w are the velocity components corresponding to S , I , and T , respectively. The constants k s , k i , and k t suggest that temperature remains constant throughout the epidemic cycle. It is important to note that equation (2.4) works in tandem with (2.6); the former describes the dynamics of the disease model, while the latter projects the velocity field associated with the compartments S ( t , x ) , I ( t , x ) , and T ( t , x ) , exhibiting fluid-like behavior. Readers interested in how equation (2.6) was derived can refer to [13].

3 Model analysis

3.1 Spatial treatment

In this section, we will explore the one-dimensional spatial analysis of the model. Let us focus on equations (2.4) and (2.6). By applying the velocity vectors to both the left-hand and right-hand sides of equation (2.4) and then combining the outcomes with equation (2.6), we arrive at the following result:

(3.1) t ( S u ) + x ( S u 2 ) = u β 1 S I u β 2 S T k s S x , t ( I v ) + x ( I v 2 ) = v β 1 S I + v β 2 S T v ( τ + μ ) I k i I x , t ( T w ) + x ( T w 2 ) = w τ I w ( γ + μ ) T k t T x .

Then, expressing the terms of equations (2.4) and (3.1) spatially, we have

(3.2) S t + x ( S u ) = β 1 S I β 2 S T , t ( S u ) + x ( S u 2 ) = u β 1 S I u β 2 S T k s S x , I t + x ( I v ) = β 1 S I + β 2 S T ( τ + μ ) I , t ( I v ) + x ( I v 2 ) = v β 1 S I + v β 2 S T v ( τ + μ ) I k i I x , T t + x ( T w ) = τ I ( γ + μ ) T t ( T w ) + x ( T w 2 ) = w τ I w ( γ + μ ) T k t T x .

In its conservative form, equation (3.2) becomes

(3.3) U t + H ( U ) x = Q ( U ) ,

where

(3.4) U = S S u I I v T T w , H ( U ) = S u S u 2 + k s S I v I v 2 + k i I T w T w 2 + k t T , Q ( U ) = β 1 S I β 2 S T u β 1 S I u β 2 S T β 1 S I + β 2 S T ( τ + μ ) I v β 1 S I + v β 2 S T v ( τ + μ ) I τ I ( γ + μ ) T w τ I w ( γ + μ ) T .

For simplicity, we define m 1 = S , m 2 = S u , m 3 = I , m 4 = I v , m 5 = T , and m 6 = T w . We then rewrite

(3.5) U = m 1 m 2 m 3 m 4 m 5 m 6 , H ( U ) = m 2 m 2 2 m 1 + k s m 1 u 4 m 4 2 m 3 + k i m 3 m 6 m 6 2 m 5 + k t m 5 , Q ( U ) = β 1 m 1 m 3 β 2 m 1 m 5 β 1 m 2 m 3 β 2 m 2 m 5 β 1 m 1 m 3 + β 2 m 1 m 5 ( τ + μ ) m 3 β 1 m 1 m 4 + β 2 m 1 m 4 m 5 m 3 ( τ + μ ) u 4 τ m 3 ( γ + μ ) m 5 τ m 3 m 6 m 5 ( γ + μ ) m 6 .

Computing the Jacobian ( J ) matrix of the flux function H ( U ) , we have

(3.6) J H = 0 1 0 0 0 0 m 2 2 m 1 2 + k s 2 m 2 m 1 0 0 0 0 0 0 0 1 0 0 0 0 v m 4 2 m 3 2 + k i 2 m 4 m 3 0 0 0 0 0 0 0 1 0 0 0 0 m 6 2 m 5 2 + k t 2 m 6 m 5 .

Then, re-substituting m i for i = 1 , , 6 in equation (3.6), we have

(3.7) J H = 0 1 0 0 0 0 u 2 + k s 2 u 0 0 0 0 0 0 0 1 0 0 0 0 v 2 + k i 2 v 0 0 0 0 0 0 0 1 0 0 0 0 w 2 + k t 2 w .

It can be easily verified that the eigenvalues of matrix (3.7) are

λ 1 , 2 = w ± k t , λ 3 , 4 = u ± k s , λ 5 , 6 = v ± k i .

Since the eigenvalues are real and distinct, together with the set of eigenvectors, system (3.7) is hyperbolic. Hence, the hyperbolicity indicates the existence of solution and well-posedness of the model equations (2.4)–(2.6) [14].

3.2 Linear analysis

It is often challenging or even impossible to solve nonlinear PDEs analytically. In order to make complex nonlinear PDE problems more manageable for analysis and numerical computation, a usual mathematical strategy is to linearize the PDE. Take for example, if U 0 is an equilibrium state, and U = U 0 + δ (where δ represents small perturbations), then a nonlinear function F ( U ) can be approximated by F ( U ) F ( U 0 ) + F U U 0 δ , where F U U 0 is the Jacobian matrix of F evaluated at U 0 . A typical example can be found in [32,58].

Given that the total population density N ( t , x ) is re-scaled to “1” [67], then the equilibrium state otherwise known as the steady state, for equation (3.3) becomes:

(3.8) U 0 = [ 1 , 0 , 0 , 0 , 0 , 0 ] T .

As such, it can be ascertained from equation (3.8) that at equilibrium state, S = S 0 = 1 , I = I 0 = 0 , T = T 0 = 0 , and u = v = w = 0 for all x Ω and t [ 0 , ) . This is commonly known as the disease-free equilibrium (DFE) state [14]. To linearize equation (3.3), we begin by evaluating the Jacobian of H ( U ) at the equilibrium state. Thus, we have:

(3.9) J H ( U 0 ) = 0 1 0 0 0 0 k s 0 0 0 0 0 0 0 0 1 0 0 0 0 k i 0 0 0 0 0 0 0 0 1 0 0 0 0 k t 0 .

The source term can be approximated as

(3.10) Q ( U ) Q ( U 0 ) + Q U U 0 ð ,

where ð = U U 0 denotes a minute perturbation and Q ( U ) Q ( U 0 ) + Q U U 0 = J Q ( U ) ( U 0 ) . Since Q ( U 0 ) = 0 , the Jacobian of Q ( U ) computed from equation (3.5) becomes

(3.11) J Q = β 1 m 3 β 2 m 5 0 β 1 m 1 0 β 2 m 1 0 0 β 1 m 3 β 2 m 5 β 1 m 2 0 β 2 m 2 0 β 1 m 3 + β 2 m 5 0 β 1 m 1 ( τ + μ ) 0 β 2 m 1 0 β 1 m 4 + β 2 m 4 m 5 m 3 0 β 2 m 1 m 4 m 5 m 3 2 β 1 m 1 + β 2 m 1 m 5 m 3 ( τ + μ ) β 2 m 1 m 4 m 3 0 0 0 τ 0 ( γ + μ ) 0 0 0 τ m 6 m 5 0 τ m 3 m 6 m 5 2 τ m 3 m 5 ( γ + μ ) .

Evaluating equation (3.11) at U 0 , we have

(3.12) J Q ( U 0 ) = 0 0 β 1 0 β 2 0 0 0 0 0 0 0 0 0 β 1 τ μ 0 β 2 0 0 0 0 β 1 τ μ 0 0 0 0 τ 0 ( γ + μ ) 0 0 0 0 0 0 ( γ + μ ) .

Therefore, the linearization of equation (3.3) becomes

(3.13) U t + H U ( U 0 ) U x = Q U ( U 0 ) U ,

where J H = H U , J Q = Q U , and for convenience, ð = U U 0 U . To further proceed, we make an artificial guess of the solution (otherwise called ansatz)

(3.14) U ( t , x ) = U ¯ e ( η t + j c x ) ,

where c is the wave speed in the spatial direction, j is an imaginary unit with the property j 2 = 1 . Then, differentiating the initial guess (3.14) and substituting the results in equation (3.13), the following result was obtained:

(3.15) η Q U ( U 0 ) j c H U ( U 0 ) U ¯ e η t + j c x = 0 .

But e η t + j c x 0 , and in matrix form, η = η (where represents the identity matrix). Therefore, (3.11) becomes

(3.16) η Q U ( U 0 ) j c H U ( U 0 ) U ¯ = 0 .

For a nontrivial solution of U ¯ , η is expected to be an eigenvalue of the resulting matrix Q U ( U 0 ) H U ( U 0 ) represented as

(3.17) 0 j c β 1 0 β 2 0 j c k s 0 0 0 0 0 0 0 β 1 τ μ j c β 2 0 0 0 j c k i β 1 τ μ 0 0 0 0 τ 0 ( γ + μ ) j c 0 0 0 0 0 ( γ + μ )

with eigenvalues

η 1 , 2 = ± j c k s , η 3 , 4 = β 1 ( τ + μ ) ± j c k i , and η 5 , 6 = ( γ + μ ) (twice) .

The eigenvalues η 1 and η 2 represent the wave-front of the susceptible fluid, with a propagation speed c in the direction defined by the unit vector ± ψ , where ψ = k s k s . Similarly, the eigenvalues η 3 and η 4 describe the wave-front of the infected fluid, also moving at speed c in the same direction ± ψ . Assuming that β 1 ( τ + μ ) > 0 , the eigenvalues η 3 and η 4 will have positive real parts, while η 5 and η 6 will have negative real parts. This indicates that the DFE is a saddle point and thus unstable [14]. Trajectories may diverge along directions corresponding to eigenvalues with positive real parts, while they may converge along those associated with negative eigenvalues. In the case of disease spread, this might translate to a scenario where:

  1. A minor perturbation of the illness in the direction of the positive eigenvalue, such as the introduction of a few new infections, may generate an epidemic or outbreak as the perturbation increases.

  2. In contrast, perturbations in the directions corresponding to negative eigenvalues will eventually fade out, resulting in the disease’s final elimination.

In practical terms, a saddle point situation suggests a delicate balance in the dynamics of disease spread. For instance, models with social distancing, immunization, and quarantine policies are examples of public health interventions that might “push” the system in the direction of stability (where the disease dies out). On the other hand, inaction or complacency could cause the system to become unstable (epidemic outbreak).

To reinforce this perspective, we examine the reproduction number of the corresponding ordinary differential equation (ODE) version of the model [19,63]:

(3.18) R 0 = τ β 2 + ( γ + μ ) β 1 ( τ + μ ) ( γ + μ ) ,

R 0 > 1 means the disease persists in the population, and vice versa. Note that with regard to equation (3.8), S 0 = 1 at DFE.

4 Numerical simulation

We now examine the nonlinear system (3.3), derived from applying fluid dynamics principles to epidemiology. This necessitates the use of a computational fluid dynamics method to discretize the system. To accurately capture the spatial dynamics, the MUSCL was selected [2527]. MUSCL is a high-resolution scheme employed to solve hyperbolic PDEs, improving the numerical solution’s accuracy by extending Godunov’s method [21,31,57] to higher orders. This is achieved through piecewise linear reconstruction of the solution, with gradient calculations in each cell and the use of limiters to avoid nonphysical oscillations near discontinuities [12,22,60]. For time integration, we applied the third-order Runge-Kutta scheme, chosen for its balance between stability and accuracy. Let i = 1 , 2 , , Z x represent the uniform mesh divisions in the spatial domain, and let U denote the numerical approximation at grid node i , represented as U i . We begin by introducing the Runge-Kutta time integration scheme, which advances the numerical solution from step n to step n + 1 .

4.1 Time updating

The third-order Runge-Kutta scheme [8,40,66] for time integration

(4.1) U i ( 1 ) = U i n + Δ t L ( U i n ) , U i ( 2 ) = 3 4 U i n + 1 4 ( U i ( 1 ) + Δ t L ( U i ( 1 ) ) ) , U i n + 1 = 1 3 U i n + 2 3 ( U i ( 2 ) + Δ t L ( U i ( 2 ) ) ) ,

where 1 i Z x , n = 0 , 1 , 2 , , and L ( U ) denotes the spatial discretization operator.

4.2 Numerical scheme

To spatially resolve the resulting nonlinear PDE (3.3) with MUSCL scheme, we begin by discretizing the domain: we divided the spatial domain into grid points with cell center x i and cell interface x i + 1 2 , where i denotes the indexes of the grid point, thereby letting Δ x denotes the spatial grid spacing, and Δ t denote the time step. At each grid point x i , we let U i 0 = U ( x i , t = 0 ) represent our initialized solution.

Accuracy of the numerical method for spatial discretization of our PDE is the utmost priority. To achieve this, we reconstruct the variable U and compute the left U i + 1 2 L and right U i + 1 2 R states at the interfaces of the cell x i + 1 2 . To ensure the monotonicity of our solution, we used the van Leer slope limiter [56,65], where the slope ratio r i for each cell i and the van Leer limiter ϕ ( r ) are given by

(4.2) r i = U i U i 1 U i + 1 U i

and

(4.3) ϕ ( r i ) = r i + r i 1 + r i .

Then, considering a given cell i , reconstructing the left and right state at the boundary i + 1 2 , we have

(4.4) U i + 1 2 L = U i + 1 2 ϕ ( r i ) ( Δ U i ) , U i + 1 2 R = U i + 1 2 ϕ ( r i + 1 ) ( Δ U i + 1 ) ,

where Δ U i = U i + 1 U i . Next based on the reconstructed states, we calculate the flux F i + 1 2 at each cell interface using the Harten-Lax-van Leer with contact (HLLC) Riemann solver [5,51]

(4.5) F i + 1 2 = HLLC U i + 1 2 L , U i + 1 2 R

where H i + 1 2 L = H U i + 1 2 L and H i + 1 2 R = H U i + 1 2 R represent the HLLC of the left and right flux functions. The intermediate state flux H * of the HLLC (4.5) is then calculated with the wave speeds S L , S R , and S M , defined by left speed, right speed, and middle (intermediate) speed, respectively, i.e.,

(4.6) S L = ( u L , v L , w L )

and

(4.7) S R = ( u R , v R , w R ) ,

where the left and right velocities were chosen based on the maximum and minimum velocity signals in the system. For simplicity, we chose a unit difference in each of the velocities u R , v R , w R , and u L , v L , w L . For the middle wave speed, we use Rankine-Hugoniot condition [23,28,35]

(4.8) S M = H ( U i + 1 2 R ) H ( U i + 1 2 L ) ( U i + 1 2 R U i + 1 2 L ) U i + 1 2 R U i + 1 2 L .

Let “ a ” denote the sound speed of the system. Therefore, we define a S , a I , and a T to be the sound speed in each compartment, respectively. For the purpose of the system, we approximate each sound speed to be: a S = k s S x , a I = k i I x , and a T = k t T x . Different cases were considered, where ( a S , a I , a T ) = ( 1 , 0 , 0 ) at DFE, and ( a S , a I , a T ) = ( 1 , ε , ε ) at any other equilibrium point, where 0 < ε < 1 . Implementing ( a S , a I , a T ) into equations (4.6) and (4.7), the new left and right speed becomes

(4.9) S L = min { u L a S , v L a I , w L a T }

and

(4.10) S R = min { u R + a S , v R + a I , w R + a T } .

Process (4.3)–(4.8) and (4.1) are repeated until discretization is exhausted.

Then, for the time step Δ t , solution is updated at cell i by:

(4.11) U i n + 1 = U i n Δ t Δ x F i + 1 2 F i 1 2 .

Therefore, the steps in applying the numerical method involves discretizing the spatial domain and initializing the solution; use the van Leer limiter to rebuild the variables at cell interfaces; use the HLLC Riemann solver to compute the numerical fluxes at cell interfaces; and use the third-order Runge-Kutta scheme to integrate the solution in time. The given PDE can be successfully solved using the MUSCL scheme with van Leer limiter, HLLC Riemann solver, and third-order Runge-Kutta time integration by following these steps.

In Figure 1, the two plots were analyzed separately at τ = 0.1 and τ = 0.315 . In both instances, it was assumed that the entire population was susceptible (blue curve) prior to the onset of disease transmission. The graphical plots (Figure 1) illustrate the numerical solutions for the population densities of Susceptible (S), Infected (I), and Treated (T) groups under different treatment response scenarios. The figure compares two cases, where the treatment parameters are set at τ = 0.1 and τ = 0.315 . These plots show how the populations change spatially across the domain. In both treatment scenarios, the densities of S, I, and T vary along the spatial domain x , with noticeable differences in infection levels and treatment effectiveness. Notably, as τ increases from 0.1 to 0.315, the infected population density decreases, suggesting that higher treatment rates are more successful in reducing infections. In contrast, the density of the treated population rises, indicating a direct relationship between the treatment rate and the number of individuals receiving treatment.

Figure 1 
                  Population densities over space at different treatment responses (
                        
                           
                           
                              τ
                              =
                              0.1
                           
                           \tau =0.1
                        
                      and 
                        
                           
                           
                              τ
                              =
                              0.315
                           
                           \tau =0.315
                        
                     ), respectively.
Figure 1

Population densities over space at different treatment responses ( τ = 0.1 and τ = 0.315 ), respectively.

The surface plot (Figure 2) reveals that with a high treatment rate, the infection (infected population density) fades away over time. Conversely, the treated population density plot shows a positive correlation between the treatment rate and the decline in the infected population density. This surface plot offers a visual depiction of how the susceptible, infected, and treated population densities change over time and space. It mathematically demonstrates how these population densities are affected by spatial movement (advection), interaction terms (such as infection and treatment), and natural processes (recovery and death). This visualization makes it easier to assess the outbreak’s spread, the impact of treatments in controlling the disease, and whether the system reaches a steady state or remains dynamic.

Figure 2 
                  Surface plot of the population density over time and space at 
                        
                           
                           
                              τ
                              =
                              0.315
                           
                           \tau =0.315
                        
                     .
Figure 2

Surface plot of the population density over time and space at τ = 0.315 .

The graphical simulation (Figure 3) of the epidemic disease (2.4)–(2.6) with the corresponding initial conditions, is being analyzed. To this effect, we assigned the parameter values k s = k i = k t = 1 and utilized the same parameter data contained in [49] to simulate. Consequently, we now move forward with the spatial simulation.

Figure 3 
                  (a)–(c) The spatio-temporal dynamics of susceptible, infected, and treated population density of model (2.4) with the parameters: 
                        
                           
                           
                              
                                 
                                    β
                                 
                                 
                                    1
                                 
                              
                              =
                              0.5
                           
                           {\beta }_{1}=0.5
                        
                     , 
                        
                           
                           
                              
                                 
                                    β
                                 
                                 
                                    2
                                 
                              
                              =
                              0.23
                           
                           {\beta }_{2}=0.23
                        
                     , 
                        
                           
                           
                              τ
                              =
                              0.12
                           
                           \tau =0.12
                        
                     , 
                        
                           
                           
                              γ
                              =
                              0.42
                           
                           \gamma =0.42
                        
                     , and 
                        
                           
                           
                              μ
                              =
                              0.0005
                           
                           \mu =0.0005
                        
                     . (d)–(f) The velocity wave-front of 
                        
                           
                           
                              S
                           
                           S
                        
                     , 
                        
                           
                           
                              I
                           
                           I
                        
                     , and 
                        
                           
                           
                              T
                           
                           T
                        
                     , while showing the homogeneous distribution of the population level in each compartment. The plots above centers on the endemic equilibrium of the model 
                        
                           
                           
                              
                                 
                                    E
                                 
                                 
                                    *
                                 
                              
                              =
                              
                                 (
                                 
                                    
                                       
                                          S
                                       
                                       
                                          *
                                       
                                    
                                    ,
                                    
                                       
                                          I
                                       
                                       
                                          *
                                       
                                    
                                    ,
                                    T
                                 
                                 )
                              
                           
                           {E}^{* }=\left({S}^{* },{I}^{* },T)
                        
                      with the basic reproduction number 
                        
                           
                           
                              
                                 
                                    R
                                 
                                 
                                    0
                                 
                              
                              =
                              4.6940759123943545
                           
                           {R}_{0}=4.6940759123943545
                        
                     .
Figure 3

(a)–(c) The spatio-temporal dynamics of susceptible, infected, and treated population density of model (2.4) with the parameters: β 1 = 0.5 , β 2 = 0.23 , τ = 0.12 , γ = 0.42 , and μ = 0.0005 . (d)–(f) The velocity wave-front of S , I , and T , while showing the homogeneous distribution of the population level in each compartment. The plots above centers on the endemic equilibrium of the model E * = ( S * , I * , T ) with the basic reproduction number R 0 = 4.6940759123943545 .

The graphical analysis in Figure 3(a)–(c) presents the surface plot for the model equations (2.4)–(2.6), where the system’s transmission rate ( β 1 ) is notably high. In contrast, plots Figure 3(d)–(f) illustrate the flow velocity for each population density, depicting the spatial spread and progression of the disease across different locations ( x ) for compartments S , I , and T . Notably, the treated population density reaches its peak at a specific spatial point in Figure 3(a)–(c). Within a confined space with a constant density population, the spatial dynamics exhibit sharp variations in the endemic state, particularly in the infected and treated density populations Figure 3(d)–(f). In summary, the analysis shows that higher infection rates lead to sustained and widespread infection levels within communities.

The graphical representation in Figure 4(a)--(c) illustrates the relative impact of a low contact rate within the system. This leads to the establishment of a DFE, where the infection is completely eradicated, and the population remains immune to further disease transmission. Lower transmission rates effectively curb the spread of the disease by reducing the likelihood of individuals transitioning from uninfected to infected. As a result, the disease fails to propagate and eventually ceases to affect the population density. In contrast, plots Figure 4(d)–(f) depict the velocity profile and spatial behavior of each compartment across location x . Our findings suggest that successful disease management and eventual eradication depend on maintaining low infection rates among the susceptible density population and keeping key epidemiological factors at consistently low levels. Consequently, the DFE state ( S 0 , 0 , 0 ) is asymptotically stable, with a basic reproduction number R 0 = 0.539419 , which is less than 1.

Figure 4 
                  (a)–(c) The spatio-temporal dynamics of susceptible, infected and treated population densities of model (2.4) with the parameters: 
                        
                           
                           
                              
                                 
                                    β
                                 
                                 
                                    1
                                 
                              
                              =
                              0.05
                           
                           {\beta }_{1}=0.05
                        
                     , 
                        
                           
                           
                              
                                 
                                    β
                                 
                                 
                                    2
                                 
                              
                              =
                              0.02
                           
                           {\beta }_{2}=0.02
                        
                     , 
                        
                           
                           
                              τ
                              =
                              0.1
                           
                           \tau =0.1
                        
                     , 
                        
                           
                           
                              γ
                              =
                              0.5
                           
                           \gamma =0.5
                        
                     , and 
                        
                           
                           
                              μ
                              =
                              0.0005
                           
                           \mu =0.0005
                        
                     . (d)–(f) The velocity wave-front of 
                        
                           
                           
                              S
                           
                           S
                        
                     , 
                        
                           
                           
                              I
                           
                           I
                        
                     , and 
                        
                           
                           
                              T
                           
                           T
                        
                     , while showing the homogeneous distribution of the population level in each compartment. The above plots centers on the DFE of the model 
                        
                           
                           
                              
                                 (
                                 
                                    
                                       
                                          S
                                       
                                       
                                          0
                                       
                                    
                                    ,
                                    
                                       
                                          I
                                       
                                       
                                          0
                                       
                                    
                                    ,
                                    
                                       
                                          T
                                       
                                       
                                          0
                                       
                                    
                                 
                                 )
                              
                              =
                              
                                 (
                                 
                                    
                                       
                                          S
                                       
                                       
                                          0
                                       
                                    
                                    ,
                                    0
                                    ,
                                    0
                                 
                                 )
                              
                           
                           \left({S}_{0},{I}_{0},{T}_{0})=\left({S}_{0},0,0)
                        
                      with the basic reproduction number 
                        
                           
                           
                              
                                 
                                    R
                                 
                                 
                                    0
                                 
                              
                              =
                              0.4901960784313726
                           
                           {R}_{0}=0.4901960784313726
                        
                     .
Figure 4

(a)–(c) The spatio-temporal dynamics of susceptible, infected and treated population densities of model (2.4) with the parameters: β 1 = 0.05 , β 2 = 0.02 , τ = 0.1 , γ = 0.5 , and μ = 0.0005 . (d)–(f) The velocity wave-front of S , I , and T , while showing the homogeneous distribution of the population level in each compartment. The above plots centers on the DFE of the model ( S 0 , I 0 , T 0 ) = ( S 0 , 0 , 0 ) with the basic reproduction number R 0 = 0.4901960784313726 .

It can be seen for the STIR model, the transmission rate ( β ) plays a substantial role in determining the endemic and disease-free dynamics of the system. However, subsidizing the transmission rate and enhancing the treatment parameter is paramount in order to eradicate diseases [39,52].

Figure 5 illustrates the impact of varying treatment parameters τ on the treatment population density. Given the assumption of a short epidemic duration, the figure indicates that as treatment efficiency improves and its objective is met at a specific point in time, it becomes prudent to discontinue the treatment strategy within the system.

Figure 5 
                  Spatial fragmentation of 
                        
                           
                           
                              T
                              
                                 (
                                 
                                    t
                                    ,
                                    x
                                 
                                 )
                              
                           
                           T\left(t,x)
                        
                      at different treatment rates 
                        
                           
                           
                              τ
                           
                           \tau 
                        
                     .
Figure 5

Spatial fragmentation of T ( t , x ) at different treatment rates τ .

4.3 Sensitivity analysis

To perform the sensitivity analysis on the basic reproduction number R 0 of the epidemiological model, we need to calculate the partial derivatives of R 0 with respect to each parameter ( β 1 , β 2 , τ , γ , μ ) and evaluate how changes in each of these parameters influence R 0 . Sensitivity analysis quantifies the responsiveness of R 0 to changes in these parameters [37] (Figure 6).

Figure 6 
                  partial rank correlation coefficient plot depicting the sensitivity analysis of the system’s infection cycle.
Figure 6

partial rank correlation coefficient plot depicting the sensitivity analysis of the system’s infection cycle.

Definition 1

The normalized forward sensitivity index of R 0 , that depends differentiably on a parameter p , is defined by [50]

(4.12) ϒ p R 0 = R 0 p × p R 0 .

Therefore, we have

ϒ β 1 R 0 = β 1 ( γ + μ ) τ β 2 + ( γ + μ ) β 1 , ϒ β 2 R 0 = τ β 2 τ β 2 + ( γ + μ ) β 1 , ϒ τ R 0 = τ [ ( γ + μ ) ( β 1 β 2 ) τ β 2 ] ( τ + μ ) [ τ β 2 + ( γ + μ ) β 1 ] , ϒ γ R 0 = τ β 2 γ ( γ + μ ) [ τ β 2 + ( γ = μ ) β 1 ] , and ϒ μ R 0 = μ [ ( γ + μ ) β 2 + τ ( β 1 β 2 ) γ β 1 ] ( τ + μ ) ( γ + μ ) [ τ β 2 + ( γ + μ ) β 1 ] .

In summary, the spatio-temporal patterns of the velocities reveal that the velocity component u is influenced by the spatial gradient of S ( t , x ) , v is driven by the spatial gradient of I ( t , x ) , while w is modulated by the gradient of T ( t , x ) . In Figure 3(a), due to the steep slope of S ( t , x ) (indicating a rapid spatial variation), the term k s S x becomes large, resulting in significant changes in u over space and time. Conversely, in regions where S ( t , x ) is nearly constant ( S x 0 ), the velocity u remains relatively stable, leading to a more uniform velocity field.

In Figure 3(b), the large gradient driven by a high infection rate β causes rapid variations in I ( t , x ) , leading to substantial shifts in v . Areas with sharp increases or decreases in infection levels will show notable velocity changes, whereas regions with a more uniform infection spread will exhibit a smoother velocity pattern.

For Figure 3(c), it is important to consider that T ( t , x ) is influenced by its transition dynamics and the treatment rate of I ( t , x ) . Regions with higher treatment rates ( τ ) will significantly affect the gradient of T ( t , x ) , thereby impacting the velocity w . Specifically, w will experience large fluctuations in areas where T undergoes rapid transitions, such as at the boundaries of an outbreak. In contrast, w will be more stable in regions where T changes gradually.

4.4 Contribution to knowledge

This work fundamentally connects epidemiology with fluid dynamics, providing a framework that adeptly models the complex spatio-temporal dynamics of epidemics. Advances in population flow modeling, together with high-resolution numerical methods, create a predictive capability that exceeds traditional epidemic models in capturing the real-world intricacies of disease spread. This approach represents a substantial leap forward in epidemic modeling, supporting more adaptable and spatially targeted public health responses.

The system of PDEs used in this context provides an improved mathematical framework for understanding the dynamics of spatial epidemic propagation. By incorporating specific epidemiological characteristics and the effects of spatial mobility, the equations are designed to model the interactions between susceptible (S), infected (I), and treated (T) population densities. Numerical simulations of these complex interactions can be accurately and stably performed by using third-order Runge-Kutta time integration, in combination with advanced numerical schemes, such as the MUSCL scheme with van Leer slope limiter and the HLLC Riemann solver. This work has contributed to knowledge in the following ways:

  • MUSCL scheme with van Leer slope limiter: The scheme MUSCL minimizes the numerical diffusion while preserving stability by improving the spatial resolution of the numerical solution. By stopping oscillations close to steep gradients, the van Leer slope limiter improves this even further and guarantees physical correctness.

  • HLLC Riemann solver: For accurate computation of fluxes across cell interfaces, the HLLC Riemann solver is utilized. The solution’s shocks and discontinuities are captured, which is important for simulating abrupt changes in population density during the course of an epidemic.

  • Third-order Runge-Kutta time integration: The accuracy and computing efficiency of this temporal integration technique are balanced. Even when the PDEs contain nonlinear elements, it permits steady integration over time.

This study has demonstrated that incorporating spatial factors into the model allows for the simulation of an epidemic’s geographical spread. Gaining insights into how diseases traverse geographic boundaries is crucial for shaping public health policies, such as travel restrictions and quarantine measures. By modeling susceptible, infected, and treated population densities separately, our approach captures the complex dynamics between different stages of the disease. This detailed representation enhances the ability to predict the outcomes of various preventive and therapeutic strategies.

5 Conclusion

A one-dimensional epidemiological flow model, designed to examine the spatial spread of infectious diseases, has been proposed. This model, constructed using a system of PDEs, establishes a comprehensive framework for analyzing the spatial dynamics of epidemics when coupled with advanced numerical methods. By integrating Euler’s equations from fluid dynamics, the model captures crucial aspects of disease transmission, treatment, and recovery on a large scale. This approach is particularly valuable for epidemiologists and public health experts, as it aids in designing effective intervention strategies and deepens the understanding of disease transmission across populations and geographical regions.

By simulating various scenarios, the model enhances preparedness and the ability to mitigate future outbreaks. It also introduces innovative techniques for improving outbreak prediction and control, especially in complex and heterogeneous environments. The study addresses previously overlooked factors such as spatial heterogeneity, targeted interventions, environmental influences, and interactions between multiple population limitations often found in earlier ODE models.

In conclusion, integrating spatial factors into infectious disease modeling yields crucial insights into the geographical spread of epidemics, playing a vital role in shaping public health strategies like travel restrictions and quarantines. The research advances the field by refining prediction and management techniques for infectious diseases in spatially complex contexts.

  1. Funding information: This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

  2. Author contributions: Daniel Ugochukwu Nnaji: Conceptualization of the study, model development, and manuscript writing. Phineas Roy Kiogora: Data analysis, simulation setup, and implementation of the MUSCL scheme. Ifeanyi Sunday Onah: Literature review, sensitivity analysis, and manuscript editing. Joseph Mung’atu: Numerical validation, statistical analysis, and critical review of mathematical formulations. Nnaemeka Stanley Aguegboh: Visualization, interpretation of results, and final review of the manuscript.

  3. Conflict of interest: The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this article.

  4. Ethical approval and consent: The study did not involve human participants or animals, and therefore, did not require ethical approval. All mathematical models and simulations were conducted following the standard scientific protocols without direct interaction with biological subjects.

  5. Informed consent: Not applicable, as the research did not involve human participants or collect any personal data requiring consent.

References

[1] Anderson, D., Tannehill, J. C., Pletcher, R. H., Munipalli, R., & Shankar, V. (2020). Computational fluid mechanics and heat transfer. Boca Raton: CRC Press. DOI: https://doi.org/10.1201/9781351124027-7.10.1201/9781351124027Search in Google Scholar

[2] Arbab, A. I., & Widatallah, H. M. (2010). On the generalized continuity equation. Chinese Physics Letters, 27(8), 084703. DOI: https://doi.org/10.1088/0256-307X/27/8/084703.10.1088/0256-307X/27/8/084703Search in Google Scholar

[3] Attard, P. (2012). Non-equilibrium thermodynamics and statistical mechanics: Foundations and applications. OUP, Oxford. DOI: https://doi.org/10.1093/acprof:oso/9780199662760.003.10.1093/acprof:oso/9780199662760.001.0001Search in Google Scholar

[4] Bagchi, B., Quesne, C., & Znojil, M. (2001). Generalized continuity equation and modified normalization in pt-symmetric quantum mechanics. Modern Physics Letters A, 16(31), 2047–2057. DOI: https://doi.org/10.1142/s0217732301005333.10.1142/S0217732301005333Search in Google Scholar

[5] Balsara, D. S., Dumbser, M., & Abgrall, R. (2014). Multidimensional HLLC Riemann solver for unstructured meshes-with application to Euler and MHD flows. Journal of Computational Physics, 261, 172–208. DOI: https://doi.org/10.1016/j.jcp.2013.12.029.10.1016/j.jcp.2013.12.029Search in Google Scholar

[6] Bertaglia, G., Boscheri, W., Dimarco, G., & Pareschi, L. (2021). Spatial spread of COVID-19 outbreak in Italy using multiscale kinetic transport equations with uncertainty. Mathematical Biosciences and Engineering: MBE, 18(5), 7028–7059. DOI: https://doi.org/10.3934/mbe.2021350.10.3934/mbe.2021350Search in Google Scholar PubMed

[7] Boris, J. P. (1976). Flux-corrected transport modules for solving generalized continuity equations. National Technical Information Service. DOI: https://doi.org/10.1016/b978-0-12-460816-0.50008-7.10.1016/B978-0-12-460816-0.50008-7Search in Google Scholar

[8] Boscarino, S. (2009). On an accurate third-order implicit-explicit Runge-Kutta method for stiff problems. Applied Numerical Mathematics, 59(7), 1515–1528. DOI: https://doi.org/10.1016/j.apnum.2008.10.003.10.1016/j.apnum.2008.10.003Search in Google Scholar

[9] Bourguignon, J.-P., & Brezis, H. (1974). Remarks on the Euler equation. Journal of Functional Analysis, 15(4), 341–363. DOI: https://doi.org/10.1016/0022-1236(74)90027-5.10.1016/0022-1236(74)90027-5Search in Google Scholar

[10] Brauer, F., Castillo-Chavez, C., & Feng, Z. (2019). Mathematical Models in Epidemiology, vol. 69. New York: Springer. DOI: https://doi.org/10.1007/978-1-4939-9828-9_17.10.1007/978-1-4939-9828-9_17Search in Google Scholar

[11] Buebos-Esteve, D. E., & Dagamac, N. H. A. (2024). Spatiotemporal models of dengue epidemiology in the Philippines: Integrating remote sensing and interpretable machine learning. Acta Tropica, 255, 107225. DOI: https://doi.org/10.1016/j.actatropica.2024.107225.10.1016/j.actatropica.2024.107225Search in Google Scholar PubMed

[12] Chen, L., & Schaefer, L. (2018). Godunov-type upwind flux schemes of the two-dimensional finite volume discrete Boltzmann method. Computers and Mathematics with Applications, 75, 3105–3126. DOI: https://doi.org/10.1016/j.camwa.2018.01.034.10.1016/j.camwa.2018.01.034Search in Google Scholar

[13] Cheng, Z., & Wang, J. (2022). Modeling epidemic flow with fluid dynamics. Mathematical Biosciences and Engineering, 19, 8334–8360. DOI: https://doi.org/10.3934/mbe.2022388.10.3934/mbe.2022388Search in Google Scholar PubMed

[14] Cheng, Z., & Wang, J. (2023). A two-phase fluid model for epidemic flow. Infectious Disease Modelling, 8, 920–938. DOI: https://doi.org/10.1016/j.idm.2023.07.001.10.1016/j.idm.2023.07.001Search in Google Scholar PubMed PubMed Central

[15] David, J. F., & Iyaniwura, S. A. (2022). Effect of human mobility on the spatial spread of airborne diseases: An epidemic model with indirect transmission. Bulletin of Mathematical Biology, 84(6), 63. DOI: https://doi.org/10.21203/rs.3.rs-965234/v1.10.1007/s11538-022-01020-8Search in Google Scholar PubMed PubMed Central

[16] Duran, I., & Moreau, S. (2013). Solution of the quasi-one-dimensional linearized Euler equations using flow invariants and the Magnus expansion. Journal of Fluid Mechanics, 723, 190–231. DOI: https://doi.org/10.1017/jfm.2013.118.10.1017/jfm.2013.118Search in Google Scholar

[17] Elliott, P., & Wartenberg, D. (2004). Spatial epidemiology: current approaches and future challenges. Environmental health perspectives, 112(9), 998–1006. DOI: https://doi.org/10.1289/ehp.6735.10.1289/ehp.6735Search in Google Scholar PubMed PubMed Central

[18] Ferraccioli, F., Stilianakis, N. I., & Veliov, V. M. (2024). A spatial epidemic model with contact and mobility restrictions. Mathematical and Computer Modelling of Dynamical Systems, 30, 284–302. DOI: https://doi.org/10.1080/13873954.2024.2341693.10.1080/13873954.2024.2341693Search in Google Scholar

[19] Fosu, G. O., Akweittey, E., & Adu-Sackey, A. (2020). Next-generation matrices and basic reproductive numbers for all phases of the coronavirus disease. Gabriel Obed Fosu, Emmanuel Akweittey, and Albert Adu-Sackey. Next-generation matrices and basic reproductive numbers for all phases of the coronavirus disease. Open Journal of Mathematical Sciences, 4(1), 261–272. DOI: https://doi.org/10.2139/ssrn.3595958.10.30538/oms2020.0117Search in Google Scholar

[20] Giles, M. B., & Pierce, N. A. (2001). Analytic adjoint solutions for the quasi-one-dimensional Euler equations. Journal of Fluid Mechanics, 426, 327–345. DOI: https://doi.org/10.1017/s0022112000002366.10.1017/S0022112000002366Search in Google Scholar

[21] Godunov, S. K., & Bohachevsky, I. (1959). Finite difference method for numerical computation of discontinuous solutions of the equations of fluid dynamics. Matematičeskij sbornik, 47(89)(3), 271–306. DOI: https://hal.science/hal-01620642/file/SG1959.pdf.Search in Google Scholar

[22] Gosse, L. (2014). A two-dimensional version of the Godunov scheme for scalar balance laws. SIAM Journal on Numerical Analysis, 52(2), 626–652. DOI: https://doi.org/10.1137/130925906.10.1137/130925906Search in Google Scholar

[23] Gouin, H., & Gavrilyuk, S. (1999). Hamilton’s principle and Rankine-Hugoniot conditions for general motions of mixtures. Meccanica, 34, 39–47. DOI: https://doi.org/10.1023/A:1004370127958.10.1023/A:1004370127958Search in Google Scholar

[24] Green, P. J., Hjort, N. L., & Richardson, S. (2003). Highly Structured Stochastic Systems, vol. 27. Oxford: Oxford University Press. DOI: https://doi.org/10.1093/oso/9780198510550.003.0001.10.1093/oso/9780198510550.001.0001Search in Google Scholar

[25] Hou, J., Liang, Q., Zhang, H., & Hinkelmann, R. (2014). Multislope MUSCL method applied to solve shallow water equations. Computers and Mathematics with Applications, 68, 2012–2027. DOI: https://doi.org/10.1016/j.camwa.2014.09.018.10.1016/j.camwa.2014.09.018Search in Google Scholar

[26] Hou, J., Liang, Q., Zhang, H., & Hinkelmann, R. (2015). An efficient unstructured MUSCL scheme for solving the 2D shallow water equations. Environmental Modelling and Software, 66, 131–152. DOI: https://doi.org/10.1016/j.envsoft.2014.12.007.10.1016/j.envsoft.2014.12.007Search in Google Scholar

[27] Hou, J., Simons, F., Mahgoub, M., & Hinkelmann, R. (2013). A robust well-balanced model on unstructured grids for shallow water flows with wetting and drying over complex topography. Computer Methods in Applied Mechanics and Engineering, 257, 126–149. DOI: https://www.sciencedirect.com/science/article/pii/S0045782513000339.10.1016/j.cma.2013.01.015Search in Google Scholar

[28] Jaisankar, S., & Rao, S. R. (2009). A central Rankine-Hugoniot solver for hyperbolic conservation laws. Journal of Computational Physics, 228(3), 770–798. DOI: https://doi.org/10.1016/j.jcp.2008.10.002.10.1016/j.jcp.2008.10.002Search in Google Scholar

[29] Kessels, F. V. W., van Lint, H., Vuik, K., & Hoogendoorn, S. (2015). Genealogy of traffic flow models. EURO Journal on Transportation and Logistics, 4, 445–473. DOI: https://doi.org/10.1007/s13676-014-0045-5.10.1007/s13676-014-0045-5Search in Google Scholar

[30] Koçak, H., & Pinar, Z. (2022). Wave propagations for dispersive variants of spatial models in epidemiology and ecology. Communications in Nonlinear Science and Numerical Simulation, 109, 106316. DOI: https://doi.org/10.1016/j.cnsns.2022.106316.10.1016/j.cnsns.2022.106316Search in Google Scholar

[31] Sweby, K. (1999). Godunov Methods. . In Toro, E. F., editor. Godunov Methods. New York: Springer. DOI: https://doi.org/10.1007/978-1-4615-0663-8_85.10.1007/978-1-4615-0663-8_85Search in Google Scholar

[32] Langtangen, H. P. (2016). Solving nonlinear ODE and PDE problems. Center for Biomedical Computing, Simula Research Laboratory and Department of Informatics, University of Oslo, Oslo, Norway. http://hplgit.github.io/INF5620/doc/pub/H14/nonlin/pdf/main_nonlin-4print-A4-2up.pdf. Search in Google Scholar

[33] Laugier, A., & Garai, J. (2007). Derivation of the ideal gas law. Journal of Chemical Education, 84(11), 1832. DOI: https://doi.org/10.1021/ed084p1832.10.1021/ed084p1832Search in Google Scholar

[34] Li, J., Xiang, T., & He, L. (2021). Modeling epidemic spread in transportation networks: A review. Journal of Traffic and Transportation Engineering, 8, 139–152. DOI: https://doi.org/10.1016/j.jtte.2020.10.003.10.1016/j.jtte.2020.10.003Search in Google Scholar

[35] Lin, X.-B. (2000). Generalized Rankine-Hugoniot condition and shock solutions for quasilinear hyperbolic systems. Journal of Differential Equations, 168(2), 321–354. DOI: https://doi.org/10.1006/jdeq.2000.3889.10.1006/jdeq.2000.3889Search in Google Scholar

[36] Lloyd, A. L., & May, R. M. (1996). Spatial heterogeneity in epidemic models. Journal of Theoretical Biology, 179(1), 1–11. DOI: https://doi.org/10.1006/jtbi.1996.0042.10.1006/jtbi.1996.0042Search in Google Scholar PubMed

[37] Lu, X., & Borgonovo, E. (2023). Global sensitivity analysis in epidemiological modeling. European Journal of Operational Research, 304(1), 9–24. DOI: https://doi.org/10.1016/j.ejor.2021.11.018.10.1016/j.ejor.2021.11.018Search in Google Scholar PubMed PubMed Central

[38] Mahsin, M., Deardon, R., & Brown, P. (2022). Geographically dependent individual-level models for infectious diseases transmission. Biostatistics, 23(1), 1–17. DOI: https://doi.org/10.1093/biostatistics/kxaa009.10.1093/biostatistics/kxaa009Search in Google Scholar PubMed

[39] Mvogo, A., Tiomela, S. A., Macías-Díaz, J. E., & Bertrand, B. (2023). Dynamics of a cross-superdiffusive SIRS model with delay effects in transmission and treatment. Nonlinear Dynamics, 111(14), 13619–13639. DOI: https://doi.org/10.1007/s11071-023-08530-7.10.1007/s11071-023-08530-7Search in Google Scholar PubMed PubMed Central

[40] Nikitin, N. (2006). Third-order-accurate semi-implicit Runge-Kutta scheme for incompressible Navier-stokes equations. International Journal for Numerical Methods in Fluids, 51(2), 221–233. DOI: https://doi.org/10.1002/fld.1122.10.1002/fld.1122Search in Google Scholar

[41] Ostfeld, R. S., Glass, G. E., & Keesing, F. (2005). Spatial epidemiology: an emerging (or re-emerging) discipline. Trends in Ecology & Evolution, 20(6), 328–336. DOI: https://doi.org/10.1016/j.tree.2005.03.009.10.1016/j.tree.2005.03.009Search in Google Scholar PubMed

[42] Pan, R., & Zhu, Y. (2016). Singularity formation for one dimensional full Euler equations. Journal of Differential Equations, 261(12), 7132–7144. DOI: https://doi.org/10.1016/j.jde.2016.09.015.10.1016/j.jde.2016.09.015Search in Google Scholar

[43] Pereira, M., Kulcsár, B., Lipták, G., Kovacs, M., & Szederkényi, G. (2024). The traffic reaction model: a kinetic compartmental approach to road traffic modeling. Transportation Research Part C: Emerging Technologies, 158, e104435–e104435. DOI: https://doi.org/10.1016/j.trc.2023.104435.10.1016/j.trc.2023.104435Search in Google Scholar

[44] Perrot, P. (1998). A to Z of Thermodynamics. OUP, Oxford. DOI: https://doi.org/10.1093/oso/9780198565567.001.0001.10.1093/oso/9780198565567.001.0001Search in Google Scholar

[45] Pletcher, R. H., Tannehill, J. C., & Anderson, D. (2012). Computational Fluid Mechanics and Heat Transfer. Boca Raton: CRC Press. DOI: https://doi.org/10.1016/0142-727x(86)90038-x.10.1016/0142-727X(86)90038-XSearch in Google Scholar

[46] Pujante-Otalora, L., Canovas-Segura, B., Campos, M., & Juarez, J. M. (2023). The use of networks in spatial and temporal computational models for outbreak spread in epidemiology: A systematic review. Journal of Biomedical Informatics, 143, 104422. DOI: https://doi.org/10.1016/j.jbi.2023.104422.10.1016/j.jbi.2023.104422Search in Google Scholar PubMed

[47] Richardson, S. (2003). Spatial models in epidemiological applications. Oxford Statistical Science Series, pp. 237–259. Oxford: Oxford University Press. DOI: https://doi.org/10.1093/oso/9780198510550.003.0023. 10.1093/oso/9780198510550.003.0023Search in Google Scholar

[48] Ruan, S. (2007). Spatial-temporal dynamics in nonlocal epidemiological models. In: Mathematics for life science and medicine, pp. 97–122. Berlin, Heidelberg: Springer Berlin Heidelberg. DOI: https://doi.org/10.1007/978-3-540-34426-1_5.10.1007/978-3-540-34426-1_5Search in Google Scholar

[49] Salman, A. M., Mohd, M. H., & Muhammad, A. (2023). A novel approach to investigate the stability analysis and the dynamics of reaction-diffusion SVIR epidemic model. Communications in Nonlinear Science and Numerical Simulation, 126, 107517. DOI: https://doi.org/10.1016/j.cnsns.2023.107517.10.1016/j.cnsns.2023.107517Search in Google Scholar

[50] Samsuzzoha, M., Singh, M., & Lucy, D. (2013). Uncertainty and sensitivity analysis of the basic reproduction number of a vaccinated epidemic model of influenza. Applied Mathematical Modelling, 37(3), 903–915. DOI: https://doi.org/10.1016/j.apm.2012.03.029.10.1016/j.apm.2012.03.029Search in Google Scholar

[51] Simon, S., & Mandal, J. (2019). A simple cure for numerical shock instability in the HLLC Riemann solver. Journal of Computational Physics, 378, 477–496. DOI: https://doi.org/10.1016/j.jcp.2018.11.022.10.1016/j.jcp.2018.11.022Search in Google Scholar

[52] Tiomela, S. A., Macías-Díaz, J. E., & Mvogo, A. (2021). Computer simulation of the dynamics of a spatial susceptible-infected-recovered epidemic model with time delays in transmission and treatment. Computer Methods and Programs in Biomedicine, 212, 106469. DOI: https://doi.org/10.1016/j.cmpb.2021.106469.10.1016/j.cmpb.2021.106469Search in Google Scholar PubMed

[53] Tsatsaris, A., Kalogeropoulos, K., & Stathopoulos, N. (2023). Chapter 1 - geoinformatics, spatial epidemiology, and public health. In Stathopoulos, N., Tsatsaris, A., & Kalogeropoulos, K., editors, Geoinformatics for Geosciences, Earth Observation, pp. 3–29. Amsterdam, The Netherlands: Elsevier. DOI: https://doi.org/10.1016/b978-0-323-98983-1.00002-8.10.1016/B978-0-323-98983-1.00002-8Search in Google Scholar

[54] Tschoegl, N. W. (2000). Fundamentals of equilibrium and steady-state thermodynamics. Amsterdam, The Netherlands: Elsevier. DOI: https://doi.org/10.1016/b978-0-444-50426-5.50046-x.10.1016/B978-0-444-50426-5.50046-XSearch in Google Scholar

[55] Tsori, Y., & Granek, R. (2021). Epidemiological model for the inhomogeneous spatial spreading of covid-19 and other diseases. PloS one, 16(2), e0246056. DOI: https://doi.org/10.1371/journal.pone.0246056.10.1371/journal.pone.0246056Search in Google Scholar PubMed PubMed Central

[56] van Leer, B. (2011). A historical oversight: Vladimir P. Kolgan and his high-resolution scheme. Journal of Computational Physics, 230(7), 2378–2383. DOI: https://doi.org/10.1016/j.jcp.2010.12.032.10.1016/j.jcp.2010.12.032Search in Google Scholar

[57] van Wageningen-Kessels, F., Daamen, W., & Hoogendoorn, S. P. (2018). Two-dimensional approximate Godunov scheme and what it means for continuum Pedestrian flow models. Transportation Science, 52(3), 547–563. DOI: https://doi.org/10.1287/trsc.2017.0793.10.1287/trsc.2017.0793Search in Google Scholar

[58] Vazquez, R., & Krstic, M. (2008). Control of 1D parabolic PDEs with Volterra nonlinearities, part i: Design. Automatica, 44(11), 2778–2790. DOI: https://doi.org/10.1016/j.automatica.2008.04.013.10.1016/j.automatica.2008.04.013Search in Google Scholar

[59] Wang, Y., Yu, X., Guo, J., Papamichail, I., Papageorgiou, M., Zhang, L., Hu, S., Li, Y., & Sun, J. (2022). Macroscopic traffic flow modelling of large-scale freeway networks with field data verification: State-of-the-art review, benchmarking framework, and case studies using metanet. Transportation Research Part C: Emerging Technologies, 145, 103904. DOI: https://doi.org/10.1016/j.trc.2022.103904.10.1016/j.trc.2022.103904Search in Google Scholar

[60] Wendroff, B. (1999). A two-dimensional HLLE Riemann solver and associated Godunov-type difference scheme for gas dynamics. Computers & Mathematics with Applications, 38, 175–185. DOI: https://doi.org/10.1016/s0898-1221(99)00296-5.10.1016/S0898-1221(99)00296-5Search in Google Scholar

[61] Wood, S. M., Alston, L., Beks, H., Mc Namara K., Coffee N., Clark R. A., Wong Shee A., & Versace V. L. (2023). Quality appraisal of spatial epidemiology and health geography research: A scoping review of systematic reviews. Health & Place, 83, 103108. DOI: https://doi.org/10.1016/j.healthplace.2023.103108.10.1016/j.healthplace.2023.103108Search in Google Scholar PubMed

[62] Woody, A. I. (2013). How is the ideal gas law explanatory? Science & Education, 22, 1563–1580. DOI: https://doi.org/10.1007/s11191-011-9424-6.10.1007/s11191-011-9424-6Search in Google Scholar

[63] Yang, H. M. (2014). The basic reproduction number obtained from Jacobian and next generation matrices-a case study of dengue transmission modelling. Biosystems, 126, 52–75. DOI: https://doi.org/10.1016/j.biosystems.2014.10.002.10.1016/j.biosystems.2014.10.002Search in Google Scholar PubMed

[64] Yeddula, S. R., Guzmán-Iñigo, J., & Morgans, A. S. (2022). A solution for the quasi-one-dimensional linearised Euler equations with heat transfer. Journal of Fluid Mechanics, 936, R3. DOI: https://doi.org/10.1017/jfm.2022.101.10.1017/jfm.2022.101Search in Google Scholar

[65] Zeng, X. (2016). A general approach to enhance slope limiters in MUSCL schemes on nonuniform rectilinear grids. SIAM Journal on Scientific Computing, 38(2), A789–A813. DOI: https://doi.org/10.1137/140970185.10.1137/140970185Search in Google Scholar

[66] Zhang, Q., & Shu, C.-W. (2010). Stability analysis and a priori error estimates of the third-order explicit Runge-Kutta discontinuous Galerkin method for scalar conservation laws. SIAM Journal on Numerical Analysis, 48(3), 1038–1063. DOI: https://doi.org/10.1137/090771363.10.1137/090771363Search in Google Scholar

[67] Zhuang, Q., & Wang, J. (2021). A spatial epidemic model with a moving boundary. Infectious Disease Modelling, 6, 1046–1060. DOI: https://doi.org/10.1016/j.idm.2021.08.005.10.1016/j.idm.2021.08.005Search in Google Scholar PubMed PubMed Central

Received: 2024-06-23
Revised: 2024-10-26
Accepted: 2024-10-26
Published Online: 2024-12-31

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Special Issue: Recent Trends in Mathematical Biology – Theory, Methods, and Applications
  2. Editorial for the Special Issue: Recent trends in mathematical Biology – Theory, methods, and applications
  3. Behavior of solutions of a discrete population model with mutualistic interaction
  4. Influence of media campaigns efforts to control spread of COVID-19 pandemic with vaccination: A modeling study
  5. Optimal control and bifurcation analysis of SEIHR model for COVID-19 with vaccination strategies and mask efficiency
  6. A mathematical study of the adrenocorticotropic hormone as a regulator of human gene expression in adrenal glands
  7. On building machine learning models for medical dataset with correlated features
  8. Analysis and numerical simulation of fractional-order blood alcohol model with singular and non-singular kernels
  9. Stability and bifurcation analysis of a nested multi-scale model for COVID-19 viral infection
  10. Augmenting heart disease prediction with explainable AI: A study of classification models
  11. Plankton interaction model: Effect of prey refuge and harvesting
  12. Modelling the leadership role of police in controlling COVID-19
  13. Robust H filter-based functional observer design for descriptor systems: An application to cardiovascular system monitoring
  14. Regular Articles
  15. Mathematical modelling of COVID-19 dynamics using SVEAIQHR model
  16. Optimal control of susceptible mature pest concerning disease-induced pest-natural enemy system with cost-effectiveness
  17. Correlated dynamics of immune network and sl(3, R) symmetry algebra
  18. Variational multiscale stabilized FEM for cardiovascular flows in complex arterial vessels under magnetic forces
  19. Assessing the impact of information-induced self-protection on Zika transmission: A mathematical modeling approach
  20. An analysis of hybrid impulsive prey-predator-mutualist system on nonuniform time domains
  21. Modelling the adverse impacts of urbanization on human health
  22. Markov modeling on dynamic state space for genetic disorders and infectious diseases with mutations: Probabilistic framework, parameter estimation, and applications
  23. In silico analysis: Fulleropyrrolidine derivatives against HIV-PR mutants and SARS-CoV-2 Mpro
  24. Tangleoids with quantum field theories in biosystems
  25. Analytic solution of a fractional-order hepatitis model using Laplace Adomian decomposition method and optimal control analysis
  26. Effect of awareness and saturated treatment on the transmission of infectious diseases
  27. Development of Aβ and anti-Aβ dynamics models for Alzheimer’s disease
  28. Compartmental modeling approach for prediction of unreported cases of COVID-19 with awareness through effective testing program
  29. COVID-19 transmission dynamics in close-contacts facilities: Optimizing control strategies
  30. Modeling and analysis of ensemble average solvation energy and solute–solvent interfacial fluctuations
  31. Application of fluid dynamics in modeling the spatial spread of infectious diseases with low mortality rate: A study using MUSCL scheme
Downloaded on 23.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/cmb-2024-0016/html
Scroll to top button