iconOpen Access

ARTICLE

A Time-Continuous Model for an Untreated HIV-Infection and a Novel Non-Standard Finite-Difference-Method for Its Discretization

Benjamin Wacker1, Jan Christian Schlüter2,3,*

1 Department of Engineering and Natural Sciences, University of Applied Sciences Merseburg, Eberhard-Leibnitz-Str. 2, Merseburg, 06217, Germany
2 Faculty of Management, Social Work and Construction, HAWK, Haarmannplatz 3, Holzminden, 37603, Germany
3 Computational Epidemiology and Public Health Research Group, Institute for Medical Epidemiology, Biometrics and Informatics, Interdisciplinary Center for Health Sciences, Martin Luther University Halle-Wittenberg, Magdeburger Str. 8, Halle, 06112, Germany

* Corresponding Author: Jan Christian Schlüter. Email: email

(This article belongs to the Special Issue: Advances in Mathematical Modeling: Numerical Approaches and Simulation for Computational Biology)

Computer Modeling in Engineering & Sciences 2025, 144(2), 2191-2229. https://doi.org/10.32604/cmes.2025.067397

Abstract

In this work, we re-investigate a classical mathematical model of untreated HIV infection suggested by Kirschner and introduce a novel non-standard finite-difference method for its numerical solution. As our first contribution, we establish non-negativity, boundedness of some solution components, existence globally in time, and uniqueness on a time interval for an arbitrary for the time-continuous problem which extends known results of Kirschner’s model in the literature. As our second analytical result, we establish different equilibrium states and examine their stability properties in the time-continuous setting or discuss some numerical tools to evaluate this question. Our third contribution is the formulation of a non-standard finite-difference method which preserves non-negativity, boundedness of some time-discrete solution components, equilibria, and their stabilities. As our final theoretical result, we prove linear convergence of our non-standard finite-difference-formulation towards the time-continuous solution. Conclusively, we present numerical examples to illustrate our theoretical findings.

Keywords

Convergence; dynamical systems; equilibrium state; existence; finite-difference-method; non-local approximations; numerical analysis; uniqueness

1  Introduction

For more than four decades, infections with human immunodeficiency virus (HIV) have become a global epidemic, with approximately 40 million people currently living with this infection [13]. However, nearly 5.5 million people do not know about their HIV infection [3]. In the absence of medical treatment of such infections, cellular mechanisms and their time development within human individuals have been well studied and understood [47]. Over time, many scientists have developed mathematical models for the evolution of an infection with HIV which are summarized in different reviews [810].

1.1 Motivation on Mathematical Modeling

Mathematical models and, in particular, differential equations play an important role in many branches of sciences. A growing field is a mathematical or theoretical biology where, for example, models for brain cancer [11] or models for competing species [12] are considered. Ethanol metabolism in the human body can be described by differential equations as well [13,14]. Temporal development of calcium-ion concentrations in liver cells can be modeled in a similar way as well [15]. In recent years, epidemiology has seen a large growth due to modeling approaches [1618]. Further fields are chemistry [1921] or computer science [22,23]. Mathematical modeling has also been an invaluable tool for understanding HIV-1 dynamics within human individuals. Let us first consider one basic mathematical model of untreated HIV dynamics within human hosts described by Stafford and co-workers [24], Perelson [25] or presented in Alizon’s and Magnus’s review article [8]. Here, we assume human blood as our compartment for our modeling idea as a non-linear system of ordinary differential equations. Since CD4+ T-cells are mainly affected by an HIV infection [4], we denote the density of susceptible, uninfected CD4+ T-cells by T(t), the density of infected CD4+ T-cells by I(t) and the density of virus particles by V(t). Uninfected target CD4+ T-cells are produced by a constant rate λ in the bone marrow, transfer to infected CD4+ T-cells by a constant rate β and die at a constant natural death rate dT. These infected CD4+ T-cells are also destroyed or die by a constant rate dI. Those infected cells can burst and expose new virus particles by a constant rate p. Finally, virus particles are cleared or die naturally from the human system by a constant rate dV. Hence, this fundamental model for the description of untreated HIV dynamics within human individuals reads

{T(t)=λβT(t)V(t)dTT(t),I(t)=βT(t)V(t)dII(t),V(t)=pI(t)dVV(t),T(0)=T0>0,I(0)=I00,V(0)=V0>0,(1)

where T0, I0 and V0 denote non-negative initial conditions. Here, we assume T0 and V0 to be positive.

From a mathematical modeling point of view, different adaptations of (1) exist in the literature [810]. Let us chronologically summarize some important studies in this regard. In 1991, Nowak and May mainly investigated numerically a simple mathematical model for antigenic variation with multiple virus strains in HIV infections [26]. Based on (1), Perelson, Kirschner and De Boer focused on stability and simulation of an extension of (1) by latently and actively infected CD4+ T-cells and a logistic depletion of susceptible, uninfected CD4+ T-cells [27] two years later. In 1994, Bonhoeffer and Nowak concentrated on the simulation and stability of intra-host time evolution of two HIV strains [28].

Before we continue our chronological summary of different contributions to the field of mathematical modeling of primary HIV infection, we want to introduce our model under consideration which is excellently described by Kirschner [29]. Again, we denote the density of susceptible, uninfected CD4+ T-cells by T(t), the density of infected CD4+ T-cells by I(t) and the density of virus particles by V(t). In contrast to (1), Kirschner modified (1) by a term which represents stimulation of CD4+ T-cells due to presence of virus particle. This adjusted model reads

{T(t)=k1(t)k2T(t)+k3T(t)V(t)k4+V(t)k5T(t)V(t),I(t)=k5T(t)V(t)k6I(t)k3I(t)V(t)k4+V(t),V(t)=k7k3I(t)V(t)k4+V(t)k8T(t)V(t)+k9V(t)k10+V(t),T(0)=T0>0,I(0)=I00,V(0)=V0>0,(2)

where T0>0, I00 and V0>0 stand for initial conditions. All meanings of time-dependent functions are presented in Table 1.

images

In the first equation of (2), k1(t) denotes a time-dependent production rate of susceptible CD4+ T-cells while k2 is a constant natural death rate of these cells. In particular, k1(t) should be a decreasing function if the virus load becomes larger which means that k1(t) should be bounded below by zero, bounded above by a positive constant K1 for all t0 and it converges to a non-negative k1~ as t. In addition to that, k3 symbolizes a constant stimulation rate of the susceptible cells by the presence of virus particles. Here, k2>k3 is assumed. k5 is a constant transfer rate of susceptible cells to infected CD4+ T-cells by infiltration of susceptible CD4+ T-cells from virus particles. k4 is an adjustable parameter. In the second equation of (2), k6 symbolizes the clearance and death rate of infected CD4+ T-cells, while the last term k3I(t)V(t)k4+V(t) represents a bursting term of infected CD4+ T-cells after their infiltration by virus particles. In the last equation of (2), k7 stands for the replication rate of virus particles after division of infected CD4+ T-cells while k8 represents a loss rate of virus particles by immune reactions. At last, k9 is the growth rate of virus particles, whereas k10 stands for a saturation rate of this proliferation process. All meanings of constant, positive problem parameters are summarized in Table 2.

images

Now, we can complete our brief overview on mathematical modeling of primary HIV infections. In 1996, Perelson and co-authors suggested an adjustment of (1) by an infectious pool and a non-infectious pool of virions after introducing a pharmacological therapy [30]. One year later, Perelson and co-workers proposed a changed dynamical model of HIV infection under combination therapy [31]. In 1999, Perelson and Nelson reviewed some fundamental mathematical models, especially regarding the stability of systems of ordinary differential equations concerning primary HIV-infections [32]. In 2000, Stafford and co-authors specifically investigated parameter estimation of model (1) concerning data from different patients [24]. Two years later, Perelson reviewed some basic models regarding primary HIV infection with and without medical treatment [25].

A delay-based system of integro-differential equations of cell-to-cell spread of primary HIV infections was introduced by Culshaw et al. in 2003 [33]. Based on (1), Stilianakis and Schenzle recommended a time-dependent infection rate [34]. In 2009, Kamp examined different strains under HIV coreceptor switch from a dynamical perspective, where investigation of stability was mainly focused [35]. One year later, Elaiw proposed an alternation of (1) by latently infected and actively infected CD4+ T-cells. The main focus of this research work was the proof of globally asymptotic stability [36]. Lythgoe and Fraser came up with a model of three host-cell types, namely susceptible CD4+ T-cells, latently infected memory CD4+ T-cells and macrophages [37]. In the same year of 2012, Alizon and Magnus reviewed the target-cell limitation model (1) and general schemes of the HIV within-host model [8]. Perelson and Ribeiro reconsidered (1) and modifications of within-host HIV models [9]. In 2014, Pourbashash and co-authors examined a global analysis of within-host virus models with cell-to-cell viral transmission [38]. Shen et al. analyzed the global stability of an infection-age structured HIV-1 model linking within-host and between-host dynamics in 2015 [39]. In the same year, Wang and co-authors developed a mathematical model for slow CD4+ T-cell decline in HIV-infected individuals [40]. Based on model (1), Pankavich and Parkinson analyzed a modification with Laplacian operators to include spatial heterogeneity in within-host viral dynamics models [41]. In 2017, Wang and co-authors presented an age-structured within-host HIV model with T-cell competition [42]. One year later, Almocera and co-authors established one multiscale within-host and between-host model for viral infectious diseases [43]. In 2023, D’Orso and Forst reviewed different models for HIV infections [10]. In 2024, Clarke and Pankavich proposed a three-stage model for HIV infections and its implications for antiretroviral therapy [44]. Wacker further investigated a simple mathematical model for primary HIV infection concerning classical mathematical questions [45] in 2024. In 2023, Li and co-authors investigated the impact of standard and non-standard finite-difference methods for between-host dynamics of HIV infections within a population [46], which motivates a quick overview of numerical discretizations of different mathematical models of HIV infection. In 2024, Meetei and co-authors investigated one HIV model against data from real cases [47]. In the same year, Ayele and co-authors examined a population-based model for the co-infection of HIV with tuberculosis [48]. In 2025, Hao and co-workers proposed a mathematical model in order to evaluate an international target strategy against HIV [49]. Recently, Teklu and co-workers have suggested a population-based and age-structured HIV model with optimal control [50].

We also like to briefly highlight stochastic approaches for the solution of within-host HIV models, such as those presented by Mbogo [51] and co-authors in 2013 or by Umar and co-workers [52] in 2022.

1.2 Motivation on Numerical Discretizations

While many works on mathematical models of HIV infections exist, numerical solution of these models, such as time-discrete variants of time-continuous models, is an important research topic as well. We want to summarize some studies with regard to this matter.

Many classical time integration methods such as Runge-Kutta schemes can be applied to such systems of ordinary differential equations [53,54]. As an alternative approach, Mickens popularized the concept of non-standard finite-difference methods for solving differential equations [55,56]. Let

y(t)=f(t,y(t))

be a system of ordinary differential equations. Often, we want to find preferably a finite-difference-scheme of the form

Dh(yn)=Fh(f,yn)

where Dh(yn)y(tn), Fh(f,yn)f(tn,yn) and yny(tn) are appropriate approximations at time grid points tn for numerical simulations. A discretization scheme is called a non-standard finite-difference-method if at least one of the conditions

1.   Dh(yn)=yn+1ynφ(hn) is an approximation of the first derivatives of our dynamical system where φ(h)=h+𝒪(h2) holds and hn=tn+1tn is a possibly non-equidistant time-step size;

2.   Fh(f,yn)=g(yn+1,yn,,h) is a non-local approximation of the right-hand-side vectorial function hold.

In 2013, Obaid et al. established a non-standard finite-difference method of development of HIV infections between hosts in a population [57]. In 2016, Yang and co-authors presented one non-standard finite-difference method for a diffusive within-host dynamics model with both virus-to-cell and cell-to-cell transmissions [58]. These authors mainly focused on analyzing stability properties for their suggested discretization method. Elaiw and Alshaikh examined the global stability of different time-discrete HIV dynamical models with three categories of infected CD4+ T-cells [59]. Salman introduced a non-standard finite-difference method and an optimal control problem for one HIV model with Beddington-DeAngelis incidence and cure rates [60]. Attaullah and Sohaib suggested one continuous Petrov-Galerkin method and one Legendre Wavelet Collocation method for the discretization of the within-host HIV model [61]. In 2023, Elaiw and co-authors developed a non-standard finite-difference-method for a co-infection model of HIV and another illness where the vectorial right-hand-side function of the dynamical system is a non-linear polynomial [62]. In that work, the authors established the global asymptotic stability of their four different equilibria by providing different Lyapunov functions. In 2025, Namjoo and co-workers proposed a non-standard finite-difference-method for a bounded, time-continuous HIV model of CD+ T-cells [63].

1.3 Contributions

All in all, many studies focused mainly on stability properties of (2) and only a few non-standard finite-difference methods exist in the literature. Henceforth, our contributions can be stated as follows:

1.   First, we extend some analytical results of (2) by proofs of non-negativity, boundedness of T(t) and I(t), at most linear growth of V(t), global existence in time, global uniqueness in time on a time interval [0,T] for an arbitrary T>0 and discuss further analytical and numerical investigations on stability properties of model (2). Here, we especially mention all results on the boundedness of solutions, global existence in time, and global uniqueness in time, which were not proved in Kirschner’s original work [29]. In particular, we would like to highlight that we, in contrast to Kirschner, establish local stability of equilibria analytically or numerically;

2.   As our main contribution, we suggest a novel non-standard finite-difference-method

{Tj+1Tjhj=k1,j+1k2Tj+1+k3TjVjk4+Vjk5Tj+1Vj,Ij+1Ijhj=k5Tj+1Vjk6Ij+1k3Ij+1Vjk4+Vj,Vj+1Vjhj=k7k3Ij+1Vjk4+Vjk8Tj+1Vj+1+k9Vjk10+Vj,T1=T0,I1=I0,V1=V0,(3)

with non-equidistant time-step sizes hj=tj+1tj and, for example, Tj denotes time-discrete solutions at grid points tj on a time mesh and k1,j+1 is an appropriate time-discretization of k1(t). Additionally, we show that many properties of the time-continuous case such as non-negativity, boundedness or at most linear growth transfer to the time-discrete setting and we demonstrate linear convergence of the time-discrete solution components towards the time-continuous ones;

3.   Finally, we illustrate our theoretical findings by numerical examples.

Our work is organized as follows. After stating our main goals in Section 1, we provide analytical results of the time-continuous model (2) in Section 2. In Section 3, we examine our proposed non-standard finite-difference-method (3) thoroughly and show evidence of our theoretical results by numerical experiments in Section 4. Conclusively, we summarize our findings and possible future research directions in Section 5.

2  Analysis of the Time-Continuous Problem Formulation

Since the vectorial right-hand-side function of (2) is at least continuous, we can conclude that all solution components are at least continuously differentiable.

2.1 Preliminaries

Before providing our analytical results, we want to summarize some important results that we need for our investigation. Let us consider the general initial-value problem

{z(t)=G(t,z(t)),z(0)=z0,(4)

where all state variables are included in z(t)=(z1(t),,zn(t)) and all initial values are given by the initial vector z0Rn. We define a function

G:[0,)×(𝒞([0,)×Rn;R))n(𝒞([0,)×Rn;R))n

where X:=(𝒞([0,)×Rn;R))n denotes the metric space of continuous functions on [0,) equipped with the classical supremum norm . Now, we want to summarize important definitions such as local and global Lipschitz-continuity which can be found in [[64], Subsection 3.2.1]. Let (X,X) and (Y,Y) be normed spaces. A function f:XY is called globally Lipschitz-continuous on X if there exists a constant K0 such that the inequality

f(x1)f(x2)YKx1x2X

holds for all x1,x2X. A function f:XY is called locally Lipschitz-continuous if, for every xX, there exists a neighborhood U of x such that f restricted to U globally Lipschitz-continuous on U. Let us recapitulate one important theorem from [[64], Theorem 4.7.1].

Theorem 1. If G:[0,)×RnRn is locally Lipschitz-continuous in both time and state variables and if there exist non-negative real functions D:[0,)[0,) and K:[0,)[0,) such that

G(t,z(t))RnK(t)z(t)Rn+D(t)(5)

holds for all z(t)Rn where denotes an appropriate norm on the corresponding space. Then it follows a solution of the initial-value problem (4) exists for all t0 and moreover, for all finite T0, it holds

z(t)Rnz0Rnexp(Kmaxt)+DmaxKmax(exp(Kmaxt)1)(6)

for all t[0,T] where

Dmax:=max0sT|D(s)|andKmax:=max0sT|K(s)|.

2.2 Non-Negativity, Boundedness of T(t) and I(t) and Linear Growth of V(t) in Time

First, we want to show that possible solutions of (2) remain non-negative for all t0.

Theorem 2. Solutions of (2) stay non-negative for all t0.

Proof. Since all solution components of (2) are continuous, there exists a time point t>0 such that

t=supt>0{t>0:T(t)0I(t)0V(t)0}

holds. Let us assume that t>0 is finite. We can conclude that

T(t)=k1(t)k2T(t)+k3T(t)V(t)k4+V(t)k5T(t)V(t)k2T(t)k5T(t)V(t)=T(t)(k2+k5V(t))

is valid for all t[0,t] and by the comparison principle, we obtain

T(t)T0exp(0t(k2k5V(τ))dτ).

Additionally, we notice that

I(t)k6I(t)k3k4I(t)=(k6+k3k4)I(t)

holds for all t[0,t] and this implies

I(t)I0exp(0t(k6+k3k4)dτ)=I0exp((k6+k3k4)t)

for all t[0,t]. Furthermore, we see that

V(t)k8T(t)V(t)

holds for all t[0,t] and the comparison principle yields

V(t)V0exp(0tk8T(τ)dτ)

for all t[0,t]. All these inequalities contradict the maximality of t and consequently, all solution components of (2) remain non-negative for all t0. □

Next, we want to prove the boundedness of T(t) globally in time.

Lemma 1. The solution components T(t) and I(t) are bounded for all t0.

Proof. We separate this proof into two parts.

1.   It has already been proven that the inequality T(t)0 is valid for all t0. For that reason, we only need to demonstrate that there exists a constant A>0 such that T(t)A for all t0. Since k1(t)K1 for all t0 and k2>k3 hold, we obtain

T(t)=k1(t)k2T(t)+k3T(t)V(t)k4+V(t)k5T(t)V(t)K1k2T(t)+k3T(t)K1(k2k3)T(t)

for all t0 due to the non-negativity of all solution components. By the comparison principle, we get

T(t)K1k2k3+(T0K1k2k3)exp((k2k3)t)max{T0;K1k2k3}=:Tmax

which proves our first assertion.

2.   It has also been shown that the inequality I(t)0 holds for all t0. Consequently, we solely need to find a constant B>0 such that I(t)B for all t0. First, by adding T(t) and I(t) and applying k1(t)K1 and k2>k3 for all t0, we conclude that

N(t):=T(t)+I(t)={k1(t)k2T(t)+k3T(t)V(t)k4+V(t)k5T(t)V(t)}+{k5T(t)V(t)k6I(t)k3I(t)V(t)k4+V(t)}K1(k2k3)T(t)k6I(t)K1min{k2k3;k6}(T(t)+I(t))

holds because of the non-negativity of all solution components. Again, applying the comparison principle, we obtain

T(t)+I(t)K1min{k2k3;k6}+(T0+I0K1min{k2k3;k6})exp(min{k2k3;k6}t)max{T0+I0;K1min{k2k3;k6}}=:Nmax

for all t0 which shows our second assertion due to the non-negativity of both solution components T(t) and I(t).

This finishes our proof. □

As our last result of this subsection, we prove that V(t) grows at most linearly in time.

Lemma 2. The solution component V(t) grows at most linearly in time.

Proof. The differential equation for V(t) yields the estimate

V(t)=k7k3I(t)V(t)k4+V(t)k8T(t)V(t)+k9V(t)k10+V(t)k3k7Nmax+k9

due to the non-negativity of all solution components. Consequently, it holds

V(t)V0+(k3k7Nmax+k9)t

by the comparison principle for all t0 which proves our claim. □

2.3 Existence Globally in Time

Let us define

F(t,(T(t),I(t),V(t))):=(F1(t,(T(t),I(t),V(t)))F2(t,(T(t),I(t),V(t)))F3(t,(T(t),I(t),V(t))))=(k1(t)k2T(t)+k3T(t)V(t)k4+V(t)k5T(t)V(t)k5T(t)V(t)k6I(t)k3I(t)V(t)k4+V(t)k7k3I(t)V(t)k4+V(t)k8T(t)V(t)+k9V(t)k10+V(t))

as the right-hand-side function of (2). We notice that this function is locally Lipschitz-continuous on [0,)×(𝒞([0,);R))3 because it is continuously differentiable on this set concerning t, T, I and V by [[64], Proposition 3.2.2]. Consequently, we can state the following result.

Theorem 3. Solutions of (2) exist for all t0.

Proof. We need to estimate all components of F(t,(T(t),I(t),V(t))) separately.

1. We see that

F1(t,(T(t),I(t),V(t)))=k1(t)k2T(t)+k3T(t)V(t)k4+V(t)k5T(t)V(t)K1+k2Tmax+k3Tmax+k5TmaxV(t)(K1+(k2+k3)Tmax)+k5Tmax(T(t),I(t),V(t))

holds.

2. Additionally, we notice that

F2(t,(T(t),I(t),V(t)))=k5T(t)V(t)k6I(t)k3I(t)V(t)k4+V(t)k5TmaxV(t)+k6Nmax+k3Nmax(k3+k6)Nmax+k5Tmax(T(t),I(t),V(t))

is valid.

3. Furthermore, we obtain

F3(t,(T(t),I(t),V(t)))=k7k3I(t)V(t)k4+V(t)k8T(t)V(t)+k9V(t)k10+V(t)k3k7Nmax+k8TmaxV(t)+k9(k3k7Nmax+k9)+k8Tmax(T(t),I(t),V(t)).

4. Conclusively, these three inequalities yield

F(t,(T(t),I(t),V(t)))(K1+k9+(k2+2k3+k6+k3k7)Nmax)+(2k5+k8)Tmax(T(t),I(t),V(t)).

Finally, this proves the existence of solutions of (2) globally in time since all assumptions of Theorem 1 are fulfilled. □

2.4 Uniqueness in Time

By applying [[65], Corollary 2, Subchapter 2.4] and [[65], Theorem 4, Subchapter 2.4], we obtain a uniqueness result globally in time on [0,T] for an arbitrary T>0.

Theorem 4. Solutions of (2) exist uniquely for all t[0,T].

Proof. Let us construct the open set E:=(12min{k4;k10},)3 for the solution variables (T,I,V). Our initial conditions (T0,I0,V0)[0,)3E are assumed to be non-negative. Furthermore, we have already observed that our right-hand-side function is continuously differentiable on E. By Lemmas 1 and 2, we see that

(T(t),I(t),V(t))K:=[0,Nmax]×[0,Nmax]×[0,V0+(k3k7Nmax+k9)T]E

holds for all t[0,T] for an arbitrary T>0. If we assume two different solution vectors starting at the same initial condition, we conclude from [[65], Theorem 4, Subchapter 2.4] that all solution components exist uniquely for all t[0,T] for an arbitrary T>0. As a consequence, uniqueness on [0,T] is established. □

2.5 Equilibrium States

For searching equilibrium states of system (2), we need to solve the system of non-linear equations

0=k1~k2T+k3TVk4+Vk5TV,0=k5TVk6Ik3IVk4+V,0=k3k7IVk4+Vk8TV+k9Vk10+V,(7)

where (T,I,V) denotes a possible equilibrium state. Here, we assume that k1(t)=k1~ is a constant for all t0 for simplicity.

2.5.1 Virus-Free Equilibrium State

It can be seen from (7) that the virus-free equilibrium state reads

(T1,I1,V1)=(k1~k2,0,0).(8)

Let us define the basic reproduction number R0 by

R0:=k2k9k1~k8k10(9)

which is going to be motivated in the proof of our next statement.

Theorem 5. If R0<1 of (9) holds, our virus-free equilibrium state (8) of (2) is locally asymptotically stable.

Proof. We divide our proof into different parts.

1. We first calculate the Jacobian matrix, denoted by DF(T,I,V), of the right-hand-side function of our dynamical system at possible equilibria. We notice that

DF(T,I,V)=(AF,1,1AF,1,2AF,1,3AF,2,1AF,2,2AF,2,3AF,3,1AF,3,2AF,3,3)(10)

with entries

AF,1,1:=k2+k3Vk4+Vk5V,AF,1,2:=0,AF,1,3:=k3Tk4+Vk3TV(k4+V)2k5T,AF,2,1:=k5V,AF,2,2:=k6k3Vk4+V,AF,2,3:=k5Tk3Vk4+V+k3IV(k4+V)2,AF,3,1:=k8V,AF,3,2:=k3k7Vk4+V,AF,3,3:=k3k7Ik4+Vk3k7IV(k4+V)2k8T+k91k10+Vk9V(k10+V)2,

holds.

2. If we plug our virus-free equilibrium state (8), we obtain

DF(T1,0,0)=(k20k3k4T1k5T10k6k5T100k8T1+k9k10).

Calculating the characteristic polynomial χ(λ) of this Jacobian at the virus-free equilibrium state, we see that

χ(λ)=det(DF(T1,0,0)λ(100010001))=(k2λ)(k6λ)(k8T1+k9k10λ)=0

is valid. Both eigenvalues λ1=k2<0 and λ2=k6<0 are negative. The last eigenvalue λ3=k9k10k8T=k9k10k1~k8k2 is negative if R0:=k2k9k1~k8k10<1 holds.

This finishes our argument. □

2.5.2 Virus-Endemic Equilibrium State

Here, we assume that

R0:=k2k9k1~k8k10>1

holds. From (7), we obtain

T2=k1~k2+k5V2k3V2k4+V20,I2=k5T2V2k3V2k4+V2+k60,V2(k3k7I2k4+V2k8T2+k9k10+V2)=0.(11)

Since V2>0 should hold in a virus-endemic equilibrium state (T2,I2,V2), it must hold

k3k7I2k4+V2k8T2+k9k10+V2=0.

Plugging all results of (11) into our previous equation, we get

0=k1~k3k5k7V2(k4+V2)(k3V2k4+V2+k6)(k2+k5V2k3V2k4+V2)k1~k8k2+k5V2k3V2k4+V2+k9k10+V2(12)

and we can re-establish zeros of (12) by finding zeros of

0=k1~k3k5k7V2(k3+k6)V2+k4k6k1~k8+k9(k5V2k3V2V2+k4+k2)V2+k10=:G(V2).(13)

Let us define the coefficients

a3:=k1~k3k5k7k1~k3k8k1~k6k8+k3k5k9+k5k6k9,a2:=k1~k3k4k5k7k1~k3k4k82k1~k4k6k8k32k9+k2k3k9+k3k4k5k9+k2k6k9k3k6k9+2k4k5k6k9+k1~k3k5k7k10k1~k3k8k10k1~k6k8k10,a1:=k1~k42k6k8+k2k3k4k9+2k2k4k6k9k3k4k6k9+k42k5k6k9+k1~k3k4k5k7k10k1~k3k4k8k102k1~k4k6k8k10,a0:=k2k42k6k9k1~k42k6k8k10=k1~k4k6k8k10(R01).(14)

Hence, finding zeros of (13) is equivalent to solving

a3(V2)3+a2(V2)2+a1V2+a0=0.(15)

Normalizing this cubic polynomial yields

(V2)3+(a2a3)=:A2(V2)2+(a1a3)=:A1V2+(a0a3)=:A0=0.(16)

This equation can be solved analytically by applying Tschirnhaus-transformation and setting

A:=12A133A12A2254A0A1A2+81A02+12A0A23,B:=36A1A2108A08A23+12A3,p:=(16)(B212A1+4A222A2B6B),q:=A1A2B9A0B+BA+A1B23B2+2A1A2218A0A2+2A2A+12A123B2.(17)

Conclusively, all solutions of (16) read

V2,1=B212A1+4A222A2B6B,V2,2=p2+i(p2)2+q,V2,3=p2i(p2)2+q.(18)

Since there is a complex interplay between initial values, problem parameters and stability of our possibly three computed equilibrium states, we want to provide a numerical algorithm for solving this problem. Our procedure is given in Algorithm 1.

images

We want to conclude our analytical results here with this remark.

Remark 1. As already mentioned above, we have a complex situation in case of R0>1. Later, we are going to demonstrate in Section 4.2 that, depending on initial values and constant problem parameters, we eventually observe different situations such as different equilibrium states or even divergence of V(t) for a large initial amount of virus particles. Additionally, due to rounding errors, there might be a possibility that we obtain complex numbers for the amount of virus particles in equilibrium states. Since our cubic polynomial in (16) has only real coefficients, two complex solutions need to be complex conjugates to each other. If this is not fulfilled, we can neglect the (small) imaginary parts and can solely regard the real parts of those computation solutions. Hence, we notice at this point that further analytical investigations on the stability of these equilibrium states might be of interest for future work on this system of non-linear differential equations. This is a comparable situation to other works [15] on calcium-ion concentrations in liver cells.

2.6 Concluding Remark

We would like to point out that all qualitative properties of our time-continuous model are important features that should transfer to the time-discrete setting. For that reason, we note that it is always important to examine properties of the time-continuous case rigorously. Consequently, it becomes easier to build time-discrete approximations which fulfill all these requirements.

3  Analysis of the Time-Discrete Problem Formulation Based on a Non-Standard Finite-Difference- Method

Here, we want to consider a non-equidistant partition

0=t1<t2<<tN<tN+1=:T

of our simulation time interval [0,T] where T>0 represents our final simulation time. Our non-equidistant time-step size is given by

hj=tj+1tj

for all j{1,2,,N1,N}. For brevity, we denote zj:=z(tj) for time-continuous functions

z:[0,T]R

at a time point tj and zh,j:=zh(tj) for time-discrete functions

zh:{tj|j{1,2,,N,N+1}}R

at a time point tj for all j{1,2,,N,N+1}.

3.1 Time-Discrete Problem Formulation and Unique Solvability

Discretizing (2) by (3), we propose our non-standard finite-difference-method based solely on non-local approximations by

{Th,j+1Th,jhj=k1,j+1k2Th,j+1+k3Th,jVh,jk4+Vh,jk5Th,j+1Vh,j,Ih,j+1Ih,jhj=k5Th,j+1Vh,jk6Ih,j+1k3Ih,j+1Vh,jk4+Vh,j,Vh,j+1Vh,jhj=k3k7Ih,j+1Vh,jk4+Vh,jk8Th,j+1Vh,j+1+k9Vh,jk10+Vh,j,Th,1=T0,Ih,1=I0,Vh,1=V0,(19)

for all time indices j{1,2,,N1,N}. Let us first observe that all these difference equations in (19) can be recast in an explicit manner. We see that

Th,j+1=Th,j+hj(k1,j+1+k3Th,jVh,jk4+Vh,j)(1+hj(k2+k5Vh,j))(20)

holds from the first equation of (19). The second equation of (19) implies

Ih,j+1=Ih,j+hjk5Th,j+1Vh,j(1+hj(k3Vh,jk4+Vh,j+k6)).(21)

From the last equation of (19), we obtain

Vh,j+1=Vh,j+hj(k3k7Ih,j+1Vh,jk4+Vh,j+k9Vh,jk10+Vh,j)(1+hjk8Th,j+1).(22)

Now, we summarize our solution algorithm for (19) in Algorithm 2.

images

Here, we can show the unique solvability of our Algorithm 2.

Theorem 6. Our numerical solution approach in Algorithm 2 is uniquely solvable.

Proof. Since our initial conditions Th,1,Ih,1,Vh,1 are set, we notice from (20)(22) that Th,2,Ih,2,Vh,2 are uniquely determined as well. Let us assume that Th,j,Ih,j,Vh,j are known for an arbitrary j{2,,N1,N}. Finally, we obtain that Th,j+1,Ih,j+1,Vh,j+1 are uniquely solvable by (20)(22) as well. By using induction, we conclude that our claim is proven. □

3.2 Non-Negativity, Boundedness of Th and Ih and at Most Linear Growth of Every Component of Vh

Let us first show that all solution components of Algorithm 2 stay non-negative.

Theorem 7. All components of our solution vectors Th,Ih,VhRN+1, calculated by Algorithm 2, remain non-negative.

Proof. Our initial conditions Th,1:=T00, Ih,1:=I00 and Vh,1:=V00 are non-negative by our assumptions. From (20)(22), we can conclude that Th,20, Ih,20 and Vh,20 hold. Let us assume that Th,j, Ih,j and Vh,j remain non-negative for an arbitrary j{2,3,,N1,N}. We see from (20)(22) that Th,j+1, Ih,j+1 and Vh,j+1 stay non-negative as well. Our claim follows from induction. □

Now, we want to prove that every component of ThRN+1 remains bounded.

Theorem 8. It holds Th,jmax{T0;K1k2k3}=:Tmax for all time-discrete indices j{1,2,,N,N+1}.

Proof. We separate this proof into two steps. Keep in mind that k1,j+1K1 holds for all j{0,1,,N1,N} by assumption, k2>k3 holds by assumption and that all solution components are non-negative by Theorem 7.

Step 1: Let us first consider

Th,j+1Th,jhj=k1,j+1k2Th,j+1+k3Th,jVh,jk4+Vh,jk5Th,j+1Vh,j

for all j{1,2,,N1,N}. We get the estimate

Th,j+1Th,jhjK1k2(Th,j+1Th,j)k2Th,j+k3Th,jVh,jk4+Vh,jK1k2(Th,j+1Th,j)k2Th,j+k3Th,j=K1k2(Th,j+1Th,j)(k2k3)Th,j

and this implies

Th,j+1Th,j+hjK1hj(k2k3)Th,j1+hjk2.(23)

Step 2: We want to prove our claim by induction. It obviously holds Th,1:=T0Tmax. Let us assume that Th,jTmax holds for an arbitrary j{1,2,,N1,N}. Using (23), we obtain

Th,j+1Th,j+hjK1hj(k2k3)Th,j1+hjk2Tmax+hjK1hj(k2k3)Tmax1+hjk2=Tmax+hjK1hj(k2k3)max{T0;K1k2k3}1+hjk2Tmax.

Conclusively, our upper bound Th,jTmax is valid for all j{1,2,,N,N+1} and it finishes our proof. □

Next, we want to show that every component of IhRN+1 stays bounded. Let us define Nh,j:=Th,j+Ih,j for all time-discrete indices j{1,2,,N,N+1} and furthermore, we set Nmax:=max{T0+I0;K1min{k2k3;k6}}.

Theorem 9. It holds Nh,jNmax for all time-discrete indices j{1,2,,N,N+1}. Consequently, every component of IhRN+1 remains bounded.

Proof. We separate this proof into two steps. Keep in mind that k1,j+1K1 holds for all j{0,1,,N1,N} by assumption, k2>k3 holds by assumption and that all solution components are non-negative by Theorem 7.

Step 1: It holds

Nh,j+1Nh,jhj=Th,j+1Th,jhj+Ih,j+1Ih,jhj=k1,j+1k2Th,j+1+k3Th,jVh,jk4+Vh,jk6Ih,j+1k3Ih,j+1Vh,jk4+Vh,jK1k2(Th,j+1Th,j)k2Th,j+k3Th,jk6(Ih,j+1Ih,j)k6Ih,jK1min{k2;k6}(Th,j+1Th,j+Ih,j+1Ih,j)min{k2k3;k6}(Th,j+Ih,j)=K1min{k2;k6}(Nh,j+1Nh,j)min{k2k3;k6}Nh,j

for all time-discrete indices j{1,2,,N1,N}. This implies

Nh,j+1Nh,j+hj(K1min{k2k3;k6}Nh,j)1+hjmin{k2;k6}(24)

for all j{1,2,,N1,N}.

Step 2: We want to prove our claim by induction. It obviously holds Nh,1:=T0+I0Nmax. Let us assume that Nh,jNmax holds for an arbitrary j{1,2,,N1,N}. Using (24), we obtain

Nh,j+1Nh,j+hj(K1min{k2k3;k6}Nh,j)1+hjmin{k2;k6}Nmax+hj(K1min{k2k3;k6}Nmax)1+hjmin{k2;k6}=Nmax+hj(K1min{k2k3;k6}max{T0+I0;K1min{k2k3;k6}})1+hjmin{k2;k6}Nmax.

Conclusively, our upper bound Nh,jNmax is valid for all j{1,2,,N,N+1} and it finishes our proof because of Nh,j:=Th,j+Ih,j for all j{1,2,,N,N+1} and non-negativity of all solution components. □

Finally, we show that every component of VhRN+1 grows at most linearly in time.

Theorem 10. It holds Vh,jV0+(k3k7Nmax+k9)tj for all time-discrete indices j{1,2,,N,N+1}.

Proof. We observe that

Vh,j+1Vh,jhj=k3k7Ih,j+1Vh,jk4+Vh,jk8Th,j+1Vh,j+1+k9Vh,jk10+Vh,jk3k7Nmax+k9

holds for an arbitrary j{1,2,,N1,N} which implies

Vh,j+1Vh,j+hj(k3k7Nmax+k9)

for an arbitrary time-discrete index j{1,2,,N1,N} as well.

Now, we want to show our claim by induction. For j=1, it obviously holds

Vh,1=V0V0+(k3k7Nmax+k9)t1.

Let us now assume that

Vh,jV0+(k3k7Nmax+k9)tj

holds for an arbitrary j{1,2,,N1,N}. It follows

Vh,j+1Vh,j+hj(k3k7Nmax+k9)V0+(k3k7Nmax+k9)tj+hj(k3k7Nmax+k9)=V0+(k3k7Nmax+k9)(tj+hj)=V0+(k3k7Nmax+k9)tj+1

which finishes our claim’s proof. □

3.3 Equilibrium States

For simplicity of presentation, let us assume that k1(t)=k1~ holds for all t0. At first, we want to re-formulate (20)(22). We first observe that

Th,j+1=Th,j+hj(k1~+k3Th,jVh,jk4+Vh,j)1+hjk2+hjk5Vh,j=Th,j+hj(k1~k2Th,j+k3Th,jVh,jk4+Vh,jk5Th,jVh,j)1+hjk2+hjk5Vh,j(25)

holds. Furthermore, we notice that

Ih,j+1=Ih,j+hjk5Th,j+1Vh,j1+hjk3Vh,jk4+Vh,j+hjk6=Ih,j+hj(k5Th,jVh,jk3Ih,jVh,jk4+Vh,jk6Ih,j)1+hjk3Vh,jk4+Vh,j+hjk6+hj2k5Vh,j(k1~k2Th,j+k3Th,jVh,jk4+Vh,jk5Th,jVh,j)(1+hjk2+hjk5Vh,j)(1+hjk3Vh,jk4+Vh,j+hjk6)(26)

is valid by using (25). By applying (25) and (26), we obtain

Vh,j+1=Vh,j+hj(k3k7Ih,j+1Vh,jk4+Vh,j+k9Vh,jk10+Vh,j)1+hjk8Th,j+1=Vh,j+hj(k3k7Ih,jVh,jk4+Vh,jk8Th,jVh,j+k9Vh,jk10+Vh,j)1+hjk8Th,j+1+hj2k3k7(k5Th,jVh,jk3Ih,jVh,jk4+Vh,jk6Ih,j)Vh,jk4+Vh,j(1+hjk3Vh,jk4+Vh,j+hjk6)(1+hjk8Th,j+1)+hj3k3k5k7(k1~k2Th,j+k3Th,jVh,jk4+Vh,jk5Th,jVh,j)Vh,j(1+hjk2+hjk5Vh,j)(1+hjk3Vh,jk4+Vh,j+hjk6)(1+hjk8Th,j+1)hj2k8(k1~k2Th,j+k3Th,jVh,jk4+Vh,jk5Th,jVh,j)Vh,j(1+hjk2+hjk5Vh,j)(1+hjk8Th,j+1).(27)

Hence, comparing (7) and (25)(27), we conclude that our proposed non-standard finite-difference-method possesses the same equilibrium states as the time-continuous model (2).

3.4 Convergence towards the Time-Continuous Problem Solution

To finish our theoretical observations of Algorithm 2, we show that it converges linearly towards the time-continuous solution of model (2). To begin with, we summarize all our assumptions for linear convergence of our time-discrete solution components of Algorithm 2 towards the time-continuous solution components of model (2).

(A1)   We consider the time interval [0,T~] with partition

0=t1<t2<t3<<tN<tN+1=T~;

(A2)   Initial conditions of the time-continuous and the time-discrete problems coincide;

(A3)   The time-continuous rate function k1:[0,T~][0,K1] should be continuously differentiable;

(A4)   Time-continuous solution components T,I,V:[0,T~][0,) should be twice continuously differentiable;

(A5)   Set Δmin:=minp{1,,N+1}hp and Δmax:=maxp{1,,N+1}hp.

Theorem 11. Let all aforementioned assumptions (A1)–(A5) be fulfilled. Then the time-discrete solution components converge linearly towards the time-continuous solution components if Δmax0.

Proof. We briefly outline our proof plan because it is relatively technical. First, let us assume that all time-continuous solution components coincide with all time-discrete solution components at a certain time point tp[0,T~] for an arbitrary p{1,2,,N1,N}. Our starting step is to examine local error propagation to a time point tp+1[0,T~]. Afterward, we investigate error propagation in time and conclusively, and we inspect the accumulation of these errors. We remind ourselves that time-continuous solutions are, for example, denoted by T(tp) and time-discrete solution, for example, by Th,p at a time point tp[0,T~].

1. For local error propagation, we assume that

(tp,Th,p)=(tp,T(tp));(tp,Ih,p)=(tp,I(tp))and(tp,Vh,p)=(tp,V(tp))

hold for an arbitrary p{1,2,,N1,N} on the time interval [tp,tp+1]. Here, we want to apply our relations (25)(27) where k1~ is replaced by our time-continuous rate function k1,p+1 evaluated at the time point tp+1. Since we only consider one-time step, we denote the corresponding time-discrete solution components by Th,p+1~, Ih,p+1~ and Vh,p+1~.

1.1. We see that

Th,p+1~=Th,p+hp(k1,p+1k2Th,p+k3Th,pVh,pk4+Vh,pk5Th,pVh,p)1+hpk2+hpk5Vh,p=T(tp)+hp(k1(tp+1)k2T(tp)+k3T(tp)V(tp)k4+V(tp)k5T(tp)V(tp))(1+hpk2+hpk5V(tp))

is valid. Additionally, we observe that

|T(tp+1)Th,p+1~||T(tp+1)T(tp)hpT(tp)|=:IT,1+|hpT(tp)hp(k1(tp+1)k2T(tp)+k3T(tp)V(tp)k4+V(tp)k5T(tp)V(tp))(1+hpk2+hpk5V(tp))|=:IIT,2=:IT,1+IIT,2

holds. Application of the mean value theorem yields

IT,1:=|T(tp+1)T(tp)hpT(tp)|=|tptp+1T(τ)T(tp)dτ|=|tptp+1T(τ)T(tp)τtp(τtp)dτ|12hp2T(t).

Another application of the mean value theorem implies

IIT,2:=|hpT(tp)hp(k1(tp+1)k2T(tp)+k3T(tp)V(tp)k4+V(tp)k5T(tp)V(tp))(1+hpk2+hpk5V(tp))|=|hp((k1(tp)k1(tp+1))+(hpk2+hpk5V(tp))T(tp))(1+hpk2+hpk5V(tp))|hp2k1(t)+hp2(k2+k5V(t))T(t).

After defining our constant

Cloc,T:=12T(t)+k1(t)+(k2+k5V(t))T(t),

we obtain

|T(tp+1)Th,p+1~|Cloc,Thp2.(28)

1.2. We see that

Ih,p+1~=I(tp)+hp(k5T(tp)V(tp)k3I(tp)V(tp)k4+V(tp)k6I(tp))1+hpk3V(tp)k4+V(tp)+hpk6+hp2k5V(tp)(k1(tp+1)k2T(tp)+k3T(tp)V(tp)k4+V(tp)k5T(tp)V(tp))(1+hpk2+hpk5V(tp))(1+hpk3V(tp)k4+V(tp)+hpk6)

is valid. This implies

|I(tp+1)Ih,p+1~||I(tp+1)I(tp)hpI(tp)|=:II,1+|hpI(tp)hp(k5T(tp)V(tp)k3I(tp)V(tp)k4+V(tp)k6I(tp))1+hpk3V(tp)k4+V(tp)+hpk6|=:III,2+|hp2k5V(tp)(k1(tp+1)k2T(tp)+k3T(tp)V(tp)k4+V(tp)k5T(tp)V(tp))(1+hpk2+hpk5V(tp))(1+hpk3V(tp)k4+V(tp)+hpk6)|=:IIII,3.

By similar argument as before, we obtain

II,112hp2I(t),

III,2=|hpI(tp)hp(k5T(tp)V(tp)k3I(tp)V(tp)k4+V(tp)k6I(tp))1+hpk3V(tp)k4+V(tp)+hpk6|=hp|hp(k3V(tp)k4+V(tp)+k6)I(tp)1+hpk3V(tp)k4+V(tp)+hpk6|hp2(k3+k6)I(t)

and

IIII,3=|hp2k5V(tp)(k1(tp+1)k2T(tp)+k3T(tp)V(tp)k4+V(tp)k5T(tp)V(tp))(1+hpk2+hpk5V(tp))(1+hpk3V(tp)k4+V(tp)+hpk6)|hp2k5V(t)(k1(t)+(k2+k3+k5)T(t)).

By defining the constant

Cloc,I:=12I(t)+(k3+k6)I(t)+k5V(t)(k1(t)+(k2+k3+k5)T(t)),

we conclude

|I(tp+1)Ih,p+1~|Cloc,Ihp2.(29)

1.3. It holds

Vh,p+1~=V(tp)+hp(k3k7I(tp)V(tp)k4+V(tp)k8T(tp)V(tp)+k9V(tp)k10+V(tp))1+hpk8Th,p+1~+hp2k3k7(k5T(tp)V(tp)k3I(tp)V(tp)k4+V(tp)k6I(tp))V(tp)k4+V(tp)(1+hpk3V(tp)k4+V(tp)+hpk6)(1+hpk8Th,p+1~)+hp3k3k5k7(k1(tp+1)k2T(tp)+k3T(tp)V(tp)k4+V(tp)k5T(tp)V(tp))V(tp)(1+hpk2+hpk5V(tp))(1+hpk3V(tp)k4+V(tp)+hpk6)(1+hpk8Th,p+1~)hp2k8(k1(tp+1)k2T(tp)+k3T(tp)V(tp)k4+V(tp)k5T(tp)V(tp))V(tp)(1+hpk2+hpk5V(tp))(1+hpk8Th,p+1~).

Furthermore, we obtain

|V(tp+1)Vh,p+1~||V(tp+1)V(tp)hpV(tp)|=:IV,1+|hpV(tp)hp(k3k7I(tp)V(tp)k4+V(tp)k8T(tp)V(tp)+k9V(tp)k10+V(tp))1+hpk8Th,p+1~|=:IIV,2+|hp2k3k7(k5T(tp)V(tp)k3I(tp)V(tp)k4+V(tp)k6I(tp))V(tp)k4+V(tp)(1+hpk3V(tp)k4+V(tp)+hpk6)(1+hpk8Th,p+1~)|=:IIIV,3+|hp3k3k5k7(k1(tp+1)k2T(tp)+k3T(tp)V(tp)k4+V(tp)k5T(tp)V(tp))V(tp)(1+hpk2+hpk5V(tp))(1+hpk3V(tp)k4+V(tp)+hpk6)(1+hpk8Th,p+1~)|=:IIIV,4+|hp2k8(k1(tp+1)k2T(tp)+k3T(tp)V(tp)k4+V(tp)k5T(tp)V(tp))V(tp)(1+hpk2+hpk5V(tp))(1+hpk8Th,p+1~)|=:IIIV,5.

By similar arguments as before, we get the estimates

  IV,112hp2I(t),IIV,2|hp2k8V(tp)Th,p+1~1+hpk8Th,p+1~|hp2k8V(t)Tmax,IIIV,3hp2k3k7I(t),IVV,4hp2k3k5k7(hp2k1(t)+hpT(t))V(t)

and

VV,5hp2k8(hpk1(t)+T(t))V(t).

Defining the constant

Cloc,V:=12I(t)+k8V(t)Tmax+k3k7I(t)+k3k5k7(hp2k1(t)+hpT(t))V(t)+k8(hpk1(t)+T(t))V(t),

we obtain

|V(tp+1)Vh,p+1~|Cloc,Vhp2.(30)

2. In general, the points (tp,Th,p), (tp,Ih,p) and (tp,Vh,p) do not exactly lie on the graph of the time-continuous solution components. For that reason, we have to examine how procedural errors such as |Th,pT(tp)|, |Ih,pI(tp)| and |Vh,pV(Tp)| propagate to the (p+1)-th time step. Furthermore, we define

Tmax,T,c:=T(t);Imax,I,c:=I(t)andVmax,V,c:=V(t)

by boundedness of all time-continuous solution components on our simulation interval [0,T~] and

Tmax,T,d:=maxpN|Th,p|;Imax,I,d:=maxpN|Ih,p|andVmax,V,d:=maxpN|Vh,p|

by boundedness of all time-discrete solution components for all p{1,2,,N+1} on our simulation interval [0,T~]. In addition to that, we define the norms

Ep+1:=max{|Th,p+1Th,p+1~|;|Ih,p+1Ih,p+1~|;|Vh,p+1Vh,p+1~|}

for all p{0,1,,N},

Fp:=max{|Th,pT(tp)|;|Ih,pI(tp)|;|Vh,pV(tp)|}

and

Gp:=max{|T(tp)Th,p~|;|I(tp)Ih,p~|;|V(tp)Vh,p~|}

for all p{1,2,,N+1}.

2.1. We see that

|Th,p+1Th,p+1~|=|Th,p+hp(k1,p+1k2Th,p+k3Th,pVh,pk4+Vh,pk5Th,pVh,p)(1+hpk2+hpk5Vh,p)T(tp)hp(k1,p+1k2T(tp)+k3T(tp)V(tp)k4+V(tp)k5T(tp)V(tp))(1+hpk2+hpk5V(tp))||Th,pT(tp)|+hp|k1,p+1(1+hpk2+hpk5V(tp))k1,p+1(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5V(tp))|=:IT,A+hp|k2Th,p(1+hpk2+hpk5V(tp))k2T(tp)(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5V(tp))|=:IIT,A+hp|k3Th,pVh,pk4+Vh,p(1+hpk2+hpk5V(tp))(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5V(tp))k3T(tp)V(tp)k4+V(tp)(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5V(tp))|+hp|k5Th,pVh,p(1+hpk2+hpk5V(tp))k5T(tp)V(tp)(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5V(tp))|=:IVT,A

is valid and IIIT,A is defined as the remaining summand which is multiplied with hp. Consequently, we obtain the estimates

IT,Ahpk5k1(t)|Vh,pV(tp)|hpk5k1(t)Fp;IIT,Ahpk2(1+hpk2)|Th,pT(tp)|+hpk2k5|Th,pV(tp)T(tp)Vh,p|hpk2(1+hpk2)Fp+hpk2k5|Th,pT(tp)||V(tp)|+hpk2k5|Vh,pV(tp)||T(tp)|hpk2(1+hpk2)Fp+hpk2k5Vmax,V,cFp+hpk2k5Tmax,T,cFp;IIIT,A:=|k3Th,pVh,pk4+Vh,p(1+hpk2+hpk5V(tp))(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5V(tp))k3T(tp)V(tp)k4+V(tp)(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5Vh,p)(1+hpk2+hpk5V(tp))|k3k4|Th,pVh,pT(tp)V(tp)|+k3k42|Th,pVh,pV(tp)T(tp)V(tp)Vh,p|+hpk2k3k4|Th,pVh,pT(tp)V(tp)|+hpk2k3k42|Th,pVh,pV(tp)T(tp)V(tp)Vh,p|+hpk3k5k4|Th,pVh,pV(tp)T(tp)Vh,pV(tp)|+hpk3k5k42|Th,pVh,pV(tp)V(tp)T(tp)V(tp)Vh,pVh,p|k3k4(Tmax,T,d+Vmax,V,c)Fp+k3k42Vmax,V,cVmax,V,dFp+hpk2k3k4(Tmax,T,d+Vmax,V,c)Fp+hpk2k3k42Vmax,V,cVmax,V,dFp+hpk3k5k4Vmax,V,cVmax,V,dFp+hpk3k5k42Vmax,V,cVmax,V,d|Th,pV(tp)T(tp)Vh,p|k3k4(Tmax,T,d+Vmax,V,c)Fp+k3k42Vmax,V,cVmax,V,dFp+hpk2k3k4(Tmax,T,d+Vmax,V,c)Fp+hpk2k3k42Vmax,V,cVmax,V,dFp+hpk3k5k4Vmax,V,cVmax,V,dFp+hpk3k5k42Vmax,V,cVmax,V,d(Tmax,T,c+Vmax,V,c)Fp

and

IVT,Ak5(1+hpk2)|Th,pVh,pT(tp)V(tp)|+hpk52|Th,pVh,pV(tp)T(tp)V(tp)Vh,p|k5(1+hpk2)Tmax,T,d|Vh,pV(tp)|+k5(1+hpk2)Vmax,V,c|Th,pT(tp)|+hpk52Vmax,V,cVmax,V,d|Th,pT(tp)|k5(1+hpk2)Tmax,T,dFp+k5(1+hpk2)Vmax,V,cFp+hpk52Vmax,V,cVmax,V,dFp.

By defining the constant

Cprop,T~:=hpk5k1(t)+hpk2{1+hpk2+k5Vmax,V,c+k5Tmax,T,c}+k3k4(Tmax,T,d+Vmax,V,c)+k3k42Vmax,V,cVmax,V,d+hpk2k3k4(Tmax,T,d+Vmax,V,c)+hpk2k3k42Vmax,V,cVmax,V,d+hpk3k5k4Vmax,V,cVmax,V,d+hpk3k5k42Vmax,V,cVmax,V,d(Tmax,T,c+Vmax,V,c)+k5(1+hpk2)Tmax,T,d+k5(1+hpk2)Vmax,V,c+hpk52Vmax,V,cVmax,V,d,

we obtain the estimate

|Th,p+1Th,p+1~|(1+hpCprop,T~)Fp.

2.2. With similar, but lengthier arguments as before, there exist non-negative constants Cprop,I~ and Cprop,V~ such that

|Ih,p+1Ih,p+1~|(1+hpCprop,I~)Fp

and

|Vh,p+1Vh,p+1~|(1+hpCprop,V~)Fp

hold. Conclusively, this implies

Ep+1(1+hpmax{Cprop,T~;Cprop,I~;Cprop,V~})Fp.(31)

3. Let us define

Cprop:=max{Cprop,T~;Cprop,I~;Cprop,V~}andCloc:=max{Cloc,T;Cloc,I;Cloc,V}.

It holds

Fp+1Ep+1+Gp+1(1+hpCprop)Fp+Clochp2

by the results of our first two steps and this can be iterated inductively. Finally, we see that

Fp+1exp(CpropT)F1+exp(CpropT)ClocTΔmax=exp(CpropT)ClocTΔmax(32)

is valid because F1=0 since the initial conditions of the time-continuous and the time-discrete problems coincide.

This finishes our proof of convergence. □

We want to conclude our investigation on our proposed non-standard finite-difference method (19) with one final remark.

Remark 2. Here, we managed to show that our proposed non-standard finite-difference-method (19) is unconditionally non-negativity-preserving. Additionally, we were able to demonstrate that this time-discrete version of (2) also preserves upper bounds for T(t) and I(t). Furthermore, our time discretization is also able to capture at most linear growth of V(t). In addition to that, we were able to prove that our numerical solution scheme converges linearly towards the time-continuous solution and we obtain the same equilibrium points as in the time-continuous setting.

However, we would also like to discuss some potential limitations of our time-discrete analysis. In contrast to [62], we were not able to prove the global asymptotic stability of our equilibrium states. While their dynamical system consists of a polynomial right-hand-side vectorial function, we have rational functions. Additionally, we showed that even our time-continuous model possesses at most linear growth of V(t) which is in accordance with numerical results of Kirschner [29]. Conclusively, we can not expect global asymptotic stability of our virus-endemic equilibrium in our time-continuous and our time-discrete settings.

4  Numerical Examples

Here, we want to illustrate our theoretical findings by different numerical examples. In our examples, we choose an equidistant mesh, i.e., that our chosen time step size h>0 is equal for all time steps. All loads of cells have unit cellsmm3, compare [29].

4.1 Non-Negativity

First, we want to show that our non-standard finite-difference-method (NSFDM), in contrast to other traditional explicit time discretizations such as the explicit Eulerian method (EE) or a second-order Runge-Kutta scheme (RK2), preserves non-negativity independent of our chosen time step sizes h>0. All problem parameters for these simulations are given by k1=10, k2=0.02, k3=0.01, k4=100, k5=0.000024, k6=0.265, k7=1000, k8=0.00074, k9=18.2, k10=1, T0=2000, I0=0 and V0=0.001 which are inspired by Kirschner’s work [29]. Our simulation results for h=10 are seen in Fig. 1.

images

Figure 1: Numerical results of a load of target cells for h=10 for all three numerical methods, namely non-standard finite-difference method (NSFDM), explicit Eulerian scheme (EE) and the second-order Runge-Kutta method (RK2)

We observe that only our proposed non-standard finite-difference method from Algorithm 2 can capture non-negativity over a long time interval for a larger equidistant time step size h=10. Both other explicit Runge-Kutta schemes fail to conserve non-negativity after some time steps. However, if we apply a sufficiently small time step size such as h=0.1, we can see that all of these explicit time-stepping methods can capture non-negativity over a large time interval, compare Fig. 2. Conclusively, we want to note that this parameter setting describes a typical untreated HIV infection over the course of time, compared to the simulations given in Kirschner’s work [29].

images

Figure 2: Numerical results of a load of target cells for h=0.1 for all three numerical methods, namely non-standard finite-difference method (NSFDM), explicit Eulerian scheme (EE) and the second-order Runge-Kutta method (RK2)

4.2 Equilibrium States

In this example, we want to distinguish three different situations of equilibrium states.

4.2.1 Situation 1: Virus-Free Equilibrium State

Let us take all problem parameters from Section 4.1. The exceptions are k9=0.25, T0=2000 and I0=100. Let h=10 and our final simulation time T=1000. From our definition of the basic reproduction number in (9), we obtain

R00.6757

for our example. Our numerical results can be seen in Fig. 3. Here, solid lines represent our solution components of a load of target cells, a load of infected target cells, and a load of virus cells. Dashed lines give our coordinates of the virus-free equilibrium state which are computed by

T1=k1k2500,I1=0,V1=0.

images

Figure 3: Numerical results of our setting in Section 4.2.1 where solid lines represent the solution components while dashed lines stand for the coordinates of the virus-free equilibrium state

We observe that our suggested non-standard finite-difference method converges to the computed coordinates of the virus-free equilibrium state.

4.2.2 Situation 2: Virus-Endemic Equilibrium State

Let us take all problem parameters from Section 4.2.1. The exception is k9=17. From our definition of the basic reproduction number in in (9), we obtain

R045.946

in this example. Our numerical results can be seen in Fig. 4. Here, solid lines represent our solution components of a load of target cells, a load of infected target cells, and load of virus cells. Dashed lines give our coordinates of the virus-free equilibrium state which are computed by

T1574.5532,I14.9362,V196.6225.

images

Figure 4: Numerical results of our setting in Section 4.2.2 where solid lines represent the solution components while dashed lines stand for the coordinates of the virus-endemic equilibrium state

We observe that our suggested non-standard finite-difference method converges to the computed coordinates of the virus-endemic equilibrium state.

4.2.3 Situation 3: Linear Growth of V(t)

Let us take all problem parameters from Section 4.2.1. The exception is k9=20. Our numerical results can be seen in Fig. 5. It can be seen that the load of target cells T(t) and the load of infected target cells I(t) remain bounded while we additionally observe that the load of viral cells V(t) grows linearly at most after a certain amount of time. In addition to that, we notice that the number of infected target cells seems to become stationary after a certain period.

images

Figure 5: Numerical results of our setting in Section 4.2.3 where solid lines represent the solution components

4.2.4 Situation 4: Real-World Application with Some Estimated Parameters from Clinical Trials and Some Calibrated Parameters from Numerical Simulations

In this numerical experiment, we take some estimated parameter values from Kirschner’s work [29] and from [27] which have been evaluated from different clinical trials. Here, we want to show that, in particular, the load of susceptible CD4+ T-cells in the acute phase (first 100 days) and the latent phase (1500–3000 days after the ending of the acute phase) is accurately described by this model, compare [29] and [44]. Our parameter choice reads k1=10, k2=0.02, k3=0.01, k4=100, k5=0.000024, k6=0.265, k7=1000, k8=0.00074, k9=17.98 and k10=1.0. Furthermore, we consider an equidistant time grid with h=1 and final simulation time T=4000.

In Fig. 6, we see numerical approximations of our presented setting. As mentioned in [44], Kirschner’s model can capture important features of the acute and the latent phase. We observe a steep decrease of susceptible CD4+ T-cells approximately 8 years after infection with HIV. Hence, Kirschner’s model can capture this feature of this steep decrease long after an untreated infection with HIV.

images

Figure 6: Numerical results of our setting in Section 4.2.4 where solid lines represent the solution components

What can be seen as a drawback of Kirschner’s model, as we have already noticed from our theoretical results in Section 2, is the linear growth of virus cells which can be explained by the last term k9V(t)k10+V(t). This summand can be interpreted as the growth of load of virus particles by infiltration of other immune cells such as macrophages.

4.3 Convergence

Conclusively, we take all problem parameters from Section 4.2.2. Here, we want to demonstrate that our proposed non-standard finite-difference-method really is a first-order method. For that purpose, we use a fine-scale solution of a fourth-order Runge-Kutta method where all solution components read TRK,j, IRK,j and VRK,j for time grid points tj and equidistant time step size hRK=0.001. Furthermore, we define the error norm

error:=maxj{|TNSFDM,jTRK,j|;|INSFDM,jIRK,j|;|VNSFDM,jVRK,j|}.

The calculated errors of our non-standard finite-difference method for different equidistant time step sizes are summarized in Table 3.

images

For small time step sizes, we notice that rate tends to 2 which means that the convergence order is one.

4.4 Some Concluding Remarks

Here, we want to summarize some of our theoretical results and of our numerical observations concerning our suggested non-standard finite-difference-method (19).

As we have already proven, our numerical solution algorithm is unconditionally stable and, in particular, it preserves non-negativity for all positive time-step sizes which is one preferable property of the time-continuous model. Hence, this is one main difference from classical Runge-Kutta time-stepping methods which are only conditionally stable and which conserve non-negativity only for a certain range of positive time-step sizes.

In terms of computational cost, since our proposed numerical solution algorithm is explicit and is of convergence order one, it is comparable to an explicit Eulerian time-stepping method concerning computational time. However, in contrast to the explicit Eulerian time-stepping procedure, our algorithm is unconditionally stable and unconditionally non-negativity-preserving. These are features that our algorithm shares with the implicit Eulerian time-stepping method. In contrast to implicit time-discretizations, our numerical solution procedure is explicit and henceforth, it avoids solutions of non-linear systems of equations in each time step.

Conclusively, our proposed non-standard finite-difference method has some interesting numerical properties which transfer from the time-continuous model to our time-discrete setting.

5  Conclusion and Outlook

In this work, we re-examined a classical dynamical system for untreated HIV infections proposed by Kirschner. In particular, we proved the non-negativity of all solution components, boundedness of two solution components, at most linear growth of one solution component, existence and uniqueness of solutions, and investigated equilibrium states and local stability for the virus-free equilibrium state. With regard to the time-discrete setting, we suggested a new non-standard finite-difference-method, solely based on non-local approximations. We demonstrated that many properties of the time-continuous case transferred to the time-discrete setting. In particular, we showed that our non-standard finite-difference method is of first order. Conclusively, we strengthened our theoretical findings with different numerical examples.

Let us discuss one important limitation of Kirschner’s time-continuous model [29]. Since our proof showed that the viral load V(t) could grow linearly, it might be important to reconsider other potential models. For simplicity, we assumed k1(t) to be constant for investigation of the equilibrium state. However, as Kirschner mentioned in [29], k1(t)=a+bc+V(t) could be chosen in dependence of V(t) with positive constants a, b and c. This would definitely not disturb many of our analytical results and could be a possible consideration for future research besides examinations of different proposed HIV infection models and their time-discretizations.

Finishing our article, we want to sketch some potential future research directions. At first, it could be of interest to construct higher-order non-standard finite-difference-methods, compare, for example, [66]. In that work, the authors applied Richardson’s extrapolation to construct higher-order non-standard finite-difference methods for a model of malware propagation. While they achieve higher rates of convergence, preservation of non-negativity for all time-step sizes was only observed numerically. Henceforth, we stayed with a first-order approach which easily guarantees non-negativity without any restrictions on possible time-step sizes. Furthermore, in contrast to many other works, we applied a non-equidistant time grid.

From a modeling perspective, it seems interesting to add spatial inhomogeneity to Kirschner’s model by the introduction of diffusion or reaction terms. This implies that the investigated model transfers to a partial differential equation [67]. Additionally, it might be worth noticing that it is possible to implement further compartments into Kirschner’s presented model. Possible extensions might be immune responses from different cells [68] or the total amount of protein levels within a human being [69]. Depending on the structure of these higher-order dynamical systems, our ideas, presented in this article, can be extended to such models as well to create non-standard finite-difference methods which preserve important properties of the time-continuous setting. The implementation of drug therapy into Kirschner’s model has already been mentioned by Kirschner herself [29].

Acknowledgement: It is not applicable.

Funding Statement: The authors received no funding for this study.

Author Contributions: Benjamin Wacker: conceptualization, formal analysis, investigation, methodology, software, validation, visualization, writing—original draft; Jan Christian Schlüter: conceptualization, writing—review & editing. All authors reviewed the results and approved the final version of the manuscript.

Availability of Data and Materials: Our GNU Octave codes can be found under the following hyperlink https://github.com/bewa87/2025-Modified-HIV-NSFDM (accessed on 07 July 2025).

Ethics Approval: It is not applicable.

Conflicts of Interest: The authors declare no conflicts of interest to report regarding the present study.

References

1. Deeks SG, Overbaugh J, Phillips A, Buchbinder S. HIV infection. Nat Rev Dis Primers. 2015;1(1):15035. doi:10.1038/nrdp.2015.35. [Google Scholar] [PubMed] [CrossRef]

2. Mody A, Sohn AH, Iwuji C, Tan RKJ, Venter F, Geng EH. HIV epidemiology, prevention, treatment, and implementation strategies for public health. Lancet. 2024;402:471–92. doi:10.1016/S0140-6736(23)01381-8. [Google Scholar] [PubMed] [CrossRef]

3. UNAids. Fact Sheet 2024-Global HIV Statistics. 2024 [Internet]. [cited 2024 Nov 5]. Available from: https://www.unaids.org/sites/default/files/media_asset/UNAIDS_FactSheet_en.pdf. [Google Scholar]

4. Moir S, Chun TW, Fauci AS. Pathogenic mechanisms of HIV disease. Annu Rev Pathol-Mech Dis. 2011;6(1):223–48. doi:10.1146/annurev-pathol-011110-130254. [Google Scholar] [PubMed] [CrossRef]

5. Mellors JW, Rinaldo CR Jr, Gupta P, White RM, Todd JA, Kingsley LA. Prognosis in HIV-1 infection predicted by the quantity of virus in plasma. Science. 1996;272(5265):1167–70. doi:10.1126/science.272.5265.1167. [Google Scholar] [PubMed] [CrossRef]

6. McCune JM. The dynamics of CD4+ T-cell depletion in HIV disease. Nature. 2001;410(6831):974–9. doi:10.1038/35073648. [Google Scholar] [PubMed] [CrossRef]

7. Deeks SG, Walker BD. Human immunodeficiency virus controllers: mechanisms of durable virus control in the absence of antiretroviral therapy. Immunity. 2007;27(3):406–16. doi:10.1016/j.immuni.2007.08.010. [Google Scholar] [PubMed] [CrossRef]

8. Alizon S, Magnus C. Modelling the course of an HIV infection: insights from ecology and evolution. Viruses. 2012;4(10):1984–2013. doi:10.3390/v4101984. [Google Scholar] [PubMed] [CrossRef]

9. Perelson AS, Ribeiro RM. Modeling the within-host dynamics of HIV infection. BMC Biol. 2013;11(1):96. doi:10.1186/1741-7007-11-96. [Google Scholar] [PubMed] [CrossRef]

10. D’Orso I, Forst CV. Mathematical models of HIV-1 dynamics, transcription, and latency. Viruses. 2023;15(10):2119. doi:10.3390/v15102119. [Google Scholar] [PubMed] [CrossRef]

11. Ashi HA, Alahmadi DM. A mathematical model of brain tumor. Math Meth Appl Sci. 2023;46(9):10137–50. doi:10.1002/mma.9107. [Google Scholar] [CrossRef]

12. Hoang MT, Ehrhardt M. A second-order nonstandard finite difference method for a general Rosenzweig-MacArthur predator-prey model. J Comput Appl Math. 2024;444(12):115752. doi:10.1016/j.cam.2024.115752. [Google Scholar] [CrossRef]

13. Wacker B. Analysis of a finite-difference method based on nonlocal approximations for a nonlinear, extended three-compartmental model of ethanol metabolism in the human body. Math Meth Appl Sci. 2025;48(9):9975–92. doi:10.1002/mma.10858. [Google Scholar] [CrossRef]

14. Levitt MD, Levitt DG. Use of a two-compartmental model to assess pharmacokinetics of human ethanol metabolism. Alcohol Clin Exp Res. 1998;22(8):1680–8. doi:10.1111/j.1530-0277.1998.tb03966.x. [Google Scholar] [CrossRef]

15. Wacker B, Schlüter JC. Qualitative analysis of two systems of nonlinear first-order ordinary differential equations for biological systems. Math Meth Appl Sci. 2022;45(8):4597–624. doi:10.1002/mma.8056. [Google Scholar] [CrossRef]

16. Hariharan S, Shangerganesh L, Debbouche A, Antonov V. Dynamic behaviors for fractional epidemiological model featuring vaccination and quarantine compartments. J Appl Math Comput. 2025;71(1):489–509. doi:10.1007/s12190-024-02249-3. [Google Scholar] [CrossRef]

17. Wacker B, Schlüter JC. An age- and sex-structured SIR model: theory and an explicit-implicit numerical solution algorithm. Math Biosci Eng. 2020;17(5):5752–801. doi:10.3934/mbe.2020309. [Google Scholar] [PubMed] [CrossRef]

18. Wacker B, Schlüter JC. A non-standard finite-difference-method for a non-autonomous epidemiological model: analysis, parameter identification and applications. Math Biosci Eng. 2023;20(7):12923–54. doi:10.3934/mbe.2023577. [Google Scholar] [PubMed] [CrossRef]

19. Feinberg M. Foundations of chemical reaction network theory. 1st ed. Cham, Switzerland: Springer-Verlag; 2019. doi:10.1007/978-3-030-03858-8. [Google Scholar] [CrossRef]

20. Formaggia L, Scotti A. Positivity and conservation properties of some integration schemes for mass action kinetics. SIAM J Numer Anal. 2011;49(3):1267–88. doi:10.1137/100789592. [Google Scholar] [CrossRef]

21. Mincheva M, Siegel D. Nonnegativity and positiveness of solutions to mass action reaction-diffusion systems. J Math Chem. 2007;42(4):1135–45. doi:10.1007/s10910-007-9292-0. [Google Scholar] [CrossRef]

22. Hoang MT, Quy Ngo TK, Tran DH. Dynamically consistent nonstandard numerical schemes for solving some computer virus and malware propagation models. Math Found Comput. 2023;6(4):704–27. doi:10.3934/mfc.2022042. [Google Scholar] [CrossRef]

23. Wacker B. Qualitative study of a dynamical system for computer virus propagation—a nonstandard finite-difference-methodological view. Math Meth Appl Sci. 2025;48(8):9272–91. doi:10.1002/mma.10798. [Google Scholar] [CrossRef]

24. Stafford MA, Corey L, Cao Y, Daar ES, Ho DD, Perelson AS. Modeling plasma virus concentration during primary HIV infection. J Theor Biol. 2000;203(3):285–301. doi:10.1006/jtbi.2000.1076. [Google Scholar] [PubMed] [CrossRef]

25. Perelson AS. Modelling viral and immune system dynamics. Nat Rev-Immunol. 2002;2(1):28–36. doi:10.1038/nri700. [Google Scholar] [PubMed] [CrossRef]

26. Nowak MA, May RM. Mathematical biology of HIV infections: antigenic variation and diversity threshold. Math Biosci. 1991;106(1):1–21. doi:10.1016/0025-5564(91)90037-j. [Google Scholar] [PubMed] [CrossRef]

27. Perelson AS, Kirschner DE, De Boer R. Dynamics of HIV infection of CD4+ T cells. Math Biosci. 1993;114(1):81–125. doi:10.1016/0025-5564(93)90043-a. [Google Scholar] [PubMed] [CrossRef]

28. Bonhoeffer S, Nowak MA. Intra-host versus inter-host selection: viral strategies of immune function impairment. Proc Natl Acad Sci U S A. 1994;91(17):8062–6. doi:10.1073/pnas.91.17.8062. [Google Scholar] [PubMed] [CrossRef]

29. Kirschner DE. Using mathematics to understand HIV immune dynamics. Not Amer Math Soc. 1996;43(2):191–202. [Google Scholar]

30. Perelson AS, Neumann AU, Markowitz M, Leonard JM, Ho DD. HIV-1 dynamics in vivo: virion clearance rate, infected cell life-span, and viral generation time. Science. 1996;271(5255):1582–6. doi:10.1126/science.271.5255.1582. [Google Scholar] [PubMed] [CrossRef]

31. Perelson AS, Essunger P, Cao Y, Vesanen M, Hurley A, Saksela K, et al. Decay characteristics of HIV-1 infected compartments during combination therapy. Nature. 1997;387(6629):188–91. doi:10.1038/387188a0. [Google Scholar] [PubMed] [CrossRef]

32. Perelson AS, Nelson PW. Mathematical analysis of HIV-1 dynamics in vivo. SIAM Rev. 1999;41(1):3–44. doi:10.1137/S0036144598335107. [Google Scholar] [CrossRef]

33. Culshaw RV, Ruan S, Webb G. A mathematical model of cell-to-cell spread of HIV-1 that includes a time delay. J Math Biol. 2003;46(5):425–44. doi:10.1007/s00285-002-0191-5. [Google Scholar] [PubMed] [CrossRef]

34. Stilianakis NI, Schenzle D. On the intra-host dynamics of HIV-1 infections. Math Biosci. 2006;199(1):1–25. doi:10.1016/j.mbs.2005.09.003. [Google Scholar] [PubMed] [CrossRef]

35. Kamp C. Understanding the HIV coreceptor switch from a dynamical perspective. BMC Ecol Evol. 2009;9(1):274. doi:10.1186/1471-2148-9-274. [Google Scholar] [PubMed] [CrossRef]

36. Elaiw AM. Global properties of a class of HIV models. Nonlinear Anal Real World Appl. 2010;11(4):2253–63. doi:10.1016/j.nonrwa.2009.07.001. [Google Scholar] [CrossRef]

37. Lythgoe KA, Fraser C. New insights into the evolutionary rate of HIV-1 at the within-host and epidemiological levels. Proc R Soc B. 2012;279(1741):3367–75. doi:10.1098/rspb.2012.0595. [Google Scholar] [PubMed] [CrossRef]

38. Pourbashash H, Pilyugin SS, De Leenheer P, McCluskey C. Global analysis of within host virus models with cell-to-cell viral transmission. Discrete Contin Dyn Syst-B. 2014;19(10):3341–57. doi:10.3934/dcdsb.2014.19.3341. [Google Scholar] [CrossRef]

39. Shen M, Xiao Y, Rong L. Global stability of an infection-age structured HIV-1 model linking within-host and between-host dynamics. Math Biosci. 2015;263:37–50. doi:10.1016/j.mbs.2015.02.003. [Google Scholar] [PubMed] [CrossRef]

40. Wang S, Hottz P, Schechter M, Rong L. Modeling the slow CD4+ cell decline in HIV-infected individuals. PLoS Comput Biol. 2015;11(12):e1004665. doi:10.1371/journal.pcbi.1004665. [Google Scholar] [PubMed] [CrossRef]

41. Pankavich S, Parkinson C. Mathematical analysis of an in-host model of viral dynamics with spatial heterogeneity. Discrete Contin Dyn Syst-B. 2016;21(4):1237–57. doi:10.3934/dcdsb.2016.21.1237. [Google Scholar] [CrossRef]

42. Wang Y, Liu K, Lou Y. An age-structured within-host HIV model with T-cell competition. Nonlinear Anal Real World Appl. 2017;38(1):1–20. doi:10.1016/j.nonrwa.2017.04.002. [Google Scholar] [CrossRef]

43. Almocera AES, Nguyen VK, Hernandez-Vargas EA. Multiscale model within-host and between-host for viral infectious diseases. J Math Biol. 2018;77(4):1035–57. doi:10.1007/s00285-018-1241-y. [Google Scholar] [PubMed] [CrossRef]

44. Clarke C, Pankavich S. Three-stage modeling of HIV infection and implications for antiretroviral therapy. J Math Biol. 2024;88(3):34. doi:10.1007/s00285-024-02056-1. [Google Scholar] [PubMed] [CrossRef]

45. Wacker B. Revisiting the classical target cell limited dynamical within-host HIV model—basic mathematical properties and stability analysis. Math Biosci Eng. 2024;21(12):7805–29. doi:10.3934/mbe.2024343. [Google Scholar] [PubMed] [CrossRef]

46. Li S, Bukhsh I, Khan IU, Asjad MI, Eldin SM, El-Rahman MA, et al. The impact of standard and nonstandard finite difference schemes on HIV nonlinear dynamical model. Chaos Solit Fract. 2023;173(6):113755. doi:10.1016/j.chaos.2023.113755. [Google Scholar] [CrossRef]

47. Meetei MZ, DarAssi MH, Altaf Khan M, Koam ANA, Alzahrani E, Ahmadini AAH. Analysis and simulation study of the HIV/AIDS model using the real cases. PLoS One. 2024;19(6):e0304735. doi:10.1371/journal.pone.0304735. [Google Scholar] [PubMed] [CrossRef]

48. Ayele TK, Doungmo Goufo EF, Mugisha S, Asamoah JKK. Co-infection mathematical model for HIV/AIDS and tuberculosis with optimal control in Ethiopia. PLoS One. 2024;19(12):e0312539. doi:10.1371/journal.pone.0312539. [Google Scholar] [PubMed] [CrossRef]

49. Hao W, Zhang J, Peng Z, Jin Z. Evaluating the current status and strategies of achieving the 95-95-95 target for HIV/AIDS: a mathematical model. Discrete Contin Dyn Syst-B. 2025;30(10):3829–57. doi:10.3934/dcdsb.2025053. [Google Scholar] [CrossRef]

50. Teklu SW, Guya TT, Kotola BS, Lachamo TS. Analyses of an age structure HIV/AIDS compartmental model with optimal control theory. Sci Rep. 2025;15(1):5491. doi:10.1038/s41598-024-82467-8. [Google Scholar] [PubMed] [CrossRef]

51. Mbogo WR, Luboobi LS, Odhiambo JW. Stochastic model for in-host HIV dynamics with therapeutic intervention. Int Sch Res Notices. 2013;2013(1):103708. doi:10.1155/2013/103708. [Google Scholar] [CrossRef]

52. Umar M, Amin F, Al-Mdallah Q, Ali MR. A stochastic computation procedure to solve the dynamics of prevention in HIV system. Biomed Signal Process Control. 2022;78(6):103888. doi:10.1016/j.bspc.2022.103888. [Google Scholar] [CrossRef]

53. Hairer E, Wanner G, Nørsett. SP. Solving ordinary differential equations I-nonstiff problems. 2nd ed. Berlin: Springer-Verlag; 1993. doi:10.1007/978-3-540-78862-1. [Google Scholar] [CrossRef]

54. Hairer E, Wanner G. Solving ordinary differential equations II-stiff and differential-algebraic problems. 2nd ed. Berlin, Germany: Springer-Verlag; 1996. doi:10.1007/978-3-642-05221-7. [Google Scholar] [CrossRef]

55. Mickens RE. Nonstandard finite difference models of differential equations. 1st ed. Singapore: World Scientific; 1994. doi:10.1142/2081. [Google Scholar] [CrossRef]

56. Mickens RE. Advances in the applications of nonstandard finite difference schemes. Singapore: World Scientific; 2005. doi:10.1142/5884. [Google Scholar] [CrossRef]

57. Obaid HA, Ouifki R, Patidar KC. An unconditionally stable nonstandard finite difference method applied to a mathematical model of HIV infection. Int J Appl Math Comput Sci. 2013;23(2):357–72. doi:10.2478/amcs-2013-0027. [Google Scholar] [CrossRef]

58. Yang Y, Zhou J, Ma X, Zhang T. Nonstandard finite difference scheme for a diffusive within-host virus dynamics model with both virus-to-cell and cell-to-cell transmissions. Comput Math Appl. 2016;72(4):1013–20. doi:10.1016/j.camwa.2016.06.015. [Google Scholar] [CrossRef]

59. Elaiw AM, Alshaikh MA. Stability of discrete-time HIV dynamics models with three categories of infected CD4+ T-cells. Adv Differ Equ. 2019;2019(1):407. doi:10.1186/s13662-019-2338-3. [Google Scholar] [CrossRef]

60. Salman SM. A nonstandard finite difference scheme and optimal control for an HIV model with Beddington-DeAngelis incidence and cure rate. The Eur Phys J Plus. 2020;135(10):808. doi:10.1140/epjp/s13360-020-00839-1. [Google Scholar] [CrossRef]

61. Attaullah, Sohaib M. Mathematical modeling and numerical simulation of HIV infection model. Results Appl Math. 2020;7(41):100118. doi:10.1016/j.rinam.2020.100118. [Google Scholar] [CrossRef]

62. Elaiw AM, Aljahdali AK, Hobiny AD. Discretization and analysis of HIV-1 and HTLV-I coinfection model with latent reservoirs. Computation. 2023;11(3):54. doi:10.3390/computation11030054. [Google Scholar] [CrossRef]

63. Namjoo M, Razaghian S, Karami M, Aminian M. Nonstandard finite difference schemes for the HIV infection model of CD4+ T-cells. J Differ Equ Appl. 2025;31(6):864–91. doi:10.1080/10236198.2025.2470179. [Google Scholar] [CrossRef]

64. Schaeffer DG, Cain JW. Ordinary differential equations: basics and beyond. New York, NY, USA: Springer-Verlag; 2016. doi:10.1007/978-1-4939-6389-8. [Google Scholar] [CrossRef]

65. Perko L. Differential equations and dynamical systems. New York, NY, USA: Springer-Verlag; 1991. [Google Scholar]

66. Martín-Vaquero J, Martín del Rey A, Encinas AH, Hernández Guillén JD, Queiruga-Dios A, Rodríguez Sánchez G. Higher-order nonstandard finite difference schemes for a MSEIR model for a malware propagation. J Comput Appl Math. 2017;317(772):146–56. doi:10.1016/j.cam.2016.11.044. [Google Scholar] [CrossRef]

67. Ren X, Tian Y, Liu L, Liu X. A reaction-diffusion within-host HIV model with cell-to-cell transmission. J Math Biol. 2018;76(7):1831–72. doi:10.1007/s00285-017-1202-x. [Google Scholar] [PubMed] [CrossRef]

68. Liyanage YR, Kohan LM, Martcheva M, Tuncer N. Identifiability and parameter estimation of within-host model of HIV with immune response. Mathematics. 2024;12(18):2837. doi:10.3390/math12182837. [Google Scholar] [CrossRef]

69. Timsina AN, Liyanage YR, Martcheva M, Tuncer N. A novel within-host model of HIV and nutrition. Math Biosci Eng. 2024;21(4):5577–603. doi:10.3934/mbe.2024246. [Google Scholar] [PubMed] [CrossRef]


Cite This Article

APA Style
Wacker, B., Schlüter, J.C. (2025). A Time-Continuous Model for an Untreated HIV-Infection and a Novel Non-Standard Finite-Difference-Method for Its Discretization. Computer Modeling in Engineering & Sciences, 144(2), 2191–2229. https://doi.org/10.32604/cmes.2025.067397
Vancouver Style
Wacker B, Schlüter JC. A Time-Continuous Model for an Untreated HIV-Infection and a Novel Non-Standard Finite-Difference-Method for Its Discretization. Comput Model Eng Sci. 2025;144(2):2191–2229. https://doi.org/10.32604/cmes.2025.067397
IEEE Style
B. Wacker and J. C. Schlüter, “A Time-Continuous Model for an Untreated HIV-Infection and a Novel Non-Standard Finite-Difference-Method for Its Discretization,” Comput. Model. Eng. Sci., vol. 144, no. 2, pp. 2191–2229, 2025. https://doi.org/10.32604/cmes.2025.067397


cc Copyright © 2025 The Author(s). Published by Tech Science Press.
This work is licensed under a Creative Commons Attribution 4.0 International License , which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
  • 1881

    View

  • 1494

    Download

  • 0

    Like

Share Link