Skip to main content
Log in

A Fractional Order Recovery SIR Model from a Stochastic Process

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

Over the past several decades, there has been a proliferation of epidemiological models with ordinary derivatives replaced by fractional derivatives in an ad hoc manner. These models may be mathematically interesting, but their relevance is uncertain. Here we develop an SIR model for an epidemic, including vital dynamics, from an underlying stochastic process. We show how fractional differential operators arise naturally in these models whenever the recovery time from the disease is power-law distributed. This can provide a model for a chronic disease process where individuals who are infected for a long time are unlikely to recover. The fractional order recovery model is shown to be consistent with the Kermack–McKendrick age-structured SIR model, and it reduces to the Hethcote–Tudor integral equation SIR model. The derivation from a stochastic process is extended to discrete time, providing a stable numerical method for solving the model equations. We have carried out simulations of the fractional order recovery model showing convergence to equilibrium states. The number of infecteds in the endemic equilibrium state increases as the fractional order of the derivative tends to zero.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

References

  • Andersson H, Britton T (2000) Stochastic epidemic models and their statistical analysis, vol 151. Springer, New York

    MATH  Google Scholar 

  • Angstmann CN, Donnelly IC, Henry BI (2013a) CTRW with reactions, forcing and trapping. Math Model Nat Phenom 8(02):17–27

    Article  MathSciNet  MATH  Google Scholar 

  • Angstmann CN, Donnelly IC, Henry BI (2013b) Pattern formation on networks with reactions: a continuous time random walk approach. Phys Rev E 87(3):032804

    Article  Google Scholar 

  • Angstmann CN, Donnelly IC, Henry BI, Nichols JA (2015) A discrete time random walk model for anomalous diffusion. J Comput Phys 293:53–69

    Article  MathSciNet  Google Scholar 

  • Ansari SP, Agrawal SK, Das S (2015) Stability analysis of fractional-order generalized chaotic susceptible-infected-recovered epidemic model and its synchronization using active control method. Pramana 84(1):23–32

  • Arafa A, Rida S, Khalil M (2012) Solutions of fractional order model of childhood diseases with constant vaccination strategy. Math Sci Lett 1(1):17–23

    Article  Google Scholar 

  • Arqub OA, El-Ajou A (2013) Solution of the fractional epidemic model by homotopy analysis method. J King Saud Univ Sci 25(1):73–81

    Article  Google Scholar 

  • Barkai E, Metzler R, Klafter J (2000) From continuous time random walks to the fractional Fokker–Planck equation. Phys Rev E 61(1):132

    Article  MathSciNet  Google Scholar 

  • Bateman H, Erdelyi A (1953) Higher transcendental functions (Volumes I-III). McGraw-Hill Book Company, New York

    Google Scholar 

  • Demirci E, Unal A, Özalp N (2011) A fractional order SEIR model with density dependent death rate. Hacet J Math Stat 40(2):287–295

    MathSciNet  MATH  Google Scholar 

  • Diethelm K (2013) A fractional calculus based model for the simulation of an outbreak of dengue fever. Nonlinear Dyn 71(4):613–619

    Article  MathSciNet  Google Scholar 

  • Diethelm K, Ford N, Freed A (2002) A predictor-corrector approach for the numerical solution of fractional differential equations. Nonlinear Dyn 29:3–22

    Article  MathSciNet  MATH  Google Scholar 

  • Ding Y, Ye H (2009) A fractional-order differential equation model of HIV infection of cd\(4^{+}\) t-cells. Math Comput Model 50(3–4):386–392

    Article  MathSciNet  MATH  Google Scholar 

  • Fedotov S (2010) Non-markovian random walks and nonlinear reactions: subdiffusion and propagating fronts. Phys Rev E 81:011117

    Article  Google Scholar 

  • Fedotov S (2011) Subdiffusion, chemotaxis, and anomalous aggregation. Phys Rev E 83:021110

    Article  Google Scholar 

  • González-Parra G, Arenas AJ, Chen-Charpentier BM (2014) A fractional order epidemic model for the simulation of outbreaks of influenza A (H1N1). Math Methods Appl Sci 37(15):2218–2226

    Article  MathSciNet  MATH  Google Scholar 

  • Goufo EFD, Maritz R, Munganga J (2014) Some properties of the Kermack–McKendrick epidemic model with fractional derivative and nonlinear incidence. Adv Differ Equ NY 2014(1):1–9

    Article  Google Scholar 

  • Hanert E, Schumacher E, Deleersnijder E (2011) Front dynamics in fractional-order epidemic models. J Theor Biol 279(1):9–16

    Article  MathSciNet  Google Scholar 

  • Henry BI, Langlands TAM, Wearne SL (2006) Anomalous diffusion with linear reaction dynamics: from continuous time random walks to fractional reaction–diffusion equations. Phys Rev E 74(3):031116

    Article  MathSciNet  Google Scholar 

  • Henry BI, Langlands TAM, Straka P (2010) Fractional Fokker–Planck equations for subdiffusion with space- and time-dependent forces. Phys Rev Lett 105:170602

    Article  Google Scholar 

  • Hethcote HW (2000) The mathematics of infectious diseases. SIAM Rev 42(4):599–653

    Article  MathSciNet  MATH  Google Scholar 

  • Hethcote H, Tudor D (1980) Integral equation models for endemic infectious diseases. J Math Biol 9:37–47

    Article  MathSciNet  MATH  Google Scholar 

  • Hilfer R, Anton L (1995) Fractional master equations and fractal time random walks. Phys Rev E 51(2):R848–R851

    Article  Google Scholar 

  • Kang M, Lagakos SW (2007) Statistical methods for panel data from a semi-Markov process, with application to HPV. Biostatistics 8(2):252–264

    Article  MATH  Google Scholar 

  • Kermack WO, McKendrick AG (1927) A contribution to the mathematical theory of epidemics. Proc R Soc Lond A Math 115(772):700–721

    Article  MATH  Google Scholar 

  • Kermack WO, McKendrick AG (1932) Contributions to the mathematical theory of epidemics. II. The problem of endemicity. Proc R Soc Lond A Math 138(834):55–83

    Article  MATH  Google Scholar 

  • Kermack WO, McKendrick AG (1933) Contributions to the mathematical theory of epidemics. III. Further studies of the problem of endemicity. Proc R Soc Lond A Math 141(843):94–122

    Article  MATH  Google Scholar 

  • Langlands TA, Henry BI (2010) Fractional chemotaxis diffusion equations. Phys Rev E 81(5):051102

    Article  MathSciNet  Google Scholar 

  • Metzler R, Klafter J (2000) The random walk’s guide to anomalous diffusion: a fractional dynamics approach. Phys Rep 339:1–77

    Article  MathSciNet  MATH  Google Scholar 

  • Miller R (1968) On the linearization of Volterra integral equations. J Math Anal Appl 23:198–208

    Article  MathSciNet  MATH  Google Scholar 

  • Mitchell C, Hudgens M, King C, Cu-Uvin S, Lo Y, Rompalo A, Sobel J, Smith J (2011) Discrete-time semi-Markov modeling of human papillomavirus persistence. Stat Med 30(17):2160–2170

    Article  MathSciNet  Google Scholar 

  • Montroll E, Weiss G (1965) Random walks on lattices ii. J Math Phys 6:167

    Article  MathSciNet  Google Scholar 

  • Oldham K, Spanier J (1974) The fractional calculus. Academic Press, London

    MATH  Google Scholar 

  • Özalp N, Demirci E (2011) A fractional order SEIR model with vertical transmission. Math Comput Model 54(1–2):1–6

    Article  MathSciNet  MATH  Google Scholar 

  • Podlubny I (1999) Fractional differential equations, mathematics in science and engineering, vol 198. Academic Press, London

    MATH  Google Scholar 

  • Pooseh S, Rodrigues HS, Torres DF (2011) Fractional derivatives in dengue epidemics. AIP Conf Proc 1389(1):739–742

    Article  Google Scholar 

  • Rositch AF, Koshiol J, Hudgens MG, Razzaghi H, Backes DM, Pimenta JM, Franco EL, Poole C, Smith JS (2013) Patterns of persistent genital human papillomavirus infection among women worldwide: a literature review and meta-analysis. Int J Cancer 133(6):1271–1285

    Article  Google Scholar 

  • Scher H, Lax M (1973) Stochastic transport in a disordered solid. I. Theory Phys Rev B 7:4491–4502

    Article  MathSciNet  Google Scholar 

  • Sibuya M (1979) Generalized hypergeometric digamma and trigamma distributions. Ann Inst Stat Math 31:373–390

  • Sokolov IM, Klafter J (2006) Field-induced dispersion in subdiffusion. Phys Rev Lett 97:140602

    Article  Google Scholar 

  • Sokolov IM, Schmidt MGW, Sagués F (2006) Reaction-subdiffusion equations. Phys Rev E 73(3):031102

    Article  Google Scholar 

  • Zeb A, Zaman G, Momani S, Ertürk VS (2013) Solution of an SEIR epidemic model in fractional order. VFAST Trans Math 1(1):7–15

    Google Scholar 

Download references

Acknowledgments

This work was supported by the Australian Research Council (DP130100595, DP140101193).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to B. I. Henry.

Appendices

Appendix 1: Derivation of the Kermack–McKendrick Model for Infecteds in a Age-Structured System

Let i(at) denote the number of individuals who are infected at time t and who have been infected for time a. The total number of infected individuals at time t is given by

$$\begin{aligned} \int _0^t i(a,t)\, \hbox {d}a=\int _0^t \varPhi (t,t')i(0,t')\, \hbox {d}t' \end{aligned}$$
(127)

where

$$\begin{aligned} \varPhi (t,t')=\phi (t-t')\theta (t,t') \end{aligned}$$
(128)

is the product of the probability of surviving death, \(\theta (t,t')\), and of ‘surviving’ the transition out of the next stage, \(\phi (t-t')\). If we differentiate Eq. (127) with respect to t using Leibniz rule, we have

$$\begin{aligned} \int _0^t\frac{\partial i}{\partial t}\, \hbox {d}a+i(t,t)=\int _0^t\frac{\partial \varPhi (t,t')}{\partial t}i(0,t')\, \hbox {d}t' +\varPhi (t,t)i(0,t) \end{aligned}$$
(129)

We also have

$$\begin{aligned} \int _0^t\frac{\partial i}{\partial a}\, \hbox {d}a=i(t,t)-i(0,t). \end{aligned}$$
(130)

But

$$\begin{aligned} \varPhi (t,t)=1 \end{aligned}$$
(131)

so that if we add Eqs. (129) and (130) then we obtain

$$\begin{aligned} \int _0^t\left( \frac{\partial i}{\partial t}+\, \frac{\partial i}{\partial a}\right) \, \hbox {d}a=\int _0^t\frac{\partial \varPhi (t,t')}{\partial t}i(0,t')\, \hbox {d}t' \end{aligned}$$
(132)

We can expand the partial derivative of \(\varPhi (t,t')\) using the product rule in Eq. (128) with

$$\begin{aligned} \frac{\partial \theta (t,t')}{\partial t}=-\gamma (t)\theta (t,t') \end{aligned}$$
(133)

and

$$\begin{aligned} \frac{\partial \phi (t-t')}{\partial t}=-\psi (t-t'), \end{aligned}$$
(134)

to obtain

$$\begin{aligned}&\int _0^t\left( \frac{\partial i}{\partial t}+\, \frac{\partial i}{\partial a}\right) \, \hbox {d}a=-\int _0^t\psi (t-t')\theta (t,t')i(0,t')\, \hbox {d}t'\nonumber \\&\quad -\int _0^t\gamma (t)\theta (t,t')\phi (t-t')i(0,t')\, \hbox {d}t'. \end{aligned}$$
(135)

Using Eq. (128) and then Eq. (127) in the second term on the RHS of Eq. (135), we now have

$$\begin{aligned} \int _0^t\left( \frac{\partial i}{\partial t}+\, \frac{\partial i}{\partial a}\right) \, \hbox {d}a=-\int _0^t\psi (t-t')\theta (t,t')i(0,t')\, \hbox {d}t' -\gamma (t)\int _0^t i(a,t)\, \hbox {d}a. \end{aligned}$$
(136)

We now define

$$\begin{aligned} \beta (t-t')=\frac{\psi (t-t')}{\phi (t-t')} \end{aligned}$$
(137)

and use Eq. (128) in the first term on the RHS of Eq. (136) to obtain

$$\begin{aligned} \int _0^t\left( \frac{\partial i}{\partial t}+\, \frac{\partial i}{\partial a}\right) \, \hbox {d}a=-\int _0^t\beta (t-t')\varPhi (t,t')i(0,t')\, \hbox {d}t' -\gamma (t)\int _0^t i(a,t)\, \hbox {d}a. \end{aligned}$$
(138)

In the first integral on the RHS of Eq. (138), it is convenient to make a change in variables \(t'=t-a\) with \(\hbox {d}t'=-\hbox {d}a\); then, we have

$$\begin{aligned} \int _0^t\left( \frac{\partial i}{\partial t}+\, \frac{\partial i}{\partial a}\right) \, \hbox {d}a=-\int _0^t\beta (a)\varPhi (t,t-a)i(0,t-a)\, \hbox {d}a -\gamma (t)\int _0^t i(a,t)\, \hbox {d}a. \end{aligned}$$
(139)

Finally, we note that

$$\begin{aligned} i(a,t)=\varPhi (t,t-a)i(0,t-a) \end{aligned}$$
(140)

so that Eq. (139) can be written as

$$\begin{aligned} \int _0^t\left( \frac{\partial i}{\partial t}+\, \frac{\partial i}{\partial a}\right) \, \hbox {d}a=-\int _0^t\beta (a)i(a,t)\, \hbox {d}a -\gamma (t)\int _0^t i(a,t)\, \hbox {d}a. \end{aligned}$$
(141)

This is consistent with Eq. (27) for the change in infectives in the age-structured Kermack McKendrick model.

Appendix 2: Limits to Continuous Time

The discrete time fractional recovery SIR model can be shown to limit to the frSIR model by identifying \(t=n\Delta t\) and taking the limit \(\Delta t \rightarrow 0\) with \(r/\Delta t^{\alpha }\) finite. The continuous time equations can be obtained from the discrete time equations using Z star transform methods. The Z star transform of Y(n) is given by

$$\begin{aligned} Z^*[Y(n)|s,\Delta t]=\sum _{n=0}^\infty Y(n) e^{-ns\Delta t} \end{aligned}$$
(142)

It follows that

$$\begin{aligned} \Delta t Z^*[Y(n)|s,\Delta t]= & {} \sum _{n=0}^\infty Y(n) e^{-ns\Delta t}\Delta t \end{aligned}$$
(143)
$$\begin{aligned}= & {} \sum _{n=0}^\infty \tilde{Y}(n\Delta t) e^{-ns\Delta t}\Delta t \end{aligned}$$
(144)

where we have introduced \(\tilde{Y}(t)\) as a function defined over a continuous variable t. We can now take the inverse Laplace transform from s to t

$$\begin{aligned} \mathcal {L}^{-1}_s\left[ \Delta t Z^*[Y(n)|s,\Delta t]\Bigg \vert t\right] =\sum _{n=0}^\infty \tilde{Y}(n\Delta t) \delta (t-n\Delta t)\Delta t \end{aligned}$$
(145)

where \(\delta (t)\) is the Dirac delta function. Here, and in the following, we use the notation \( \mathcal {L}^{-1}_s\left[ Y(s)\Bigg \vert t\right] \) to denote the inverse Laplace transform from s to t and we use the notation \( \mathcal {L}_t\left[ Y(t)\Bigg \vert s\right] \) to denote the Laplace transform from t to s.

It is useful to define the function

$$\begin{aligned} \tilde{Y}(t|\Delta t)=\sum _{n=0}^\infty \tilde{Y}(n\Delta t) \delta (t-n\Delta t)\Delta t. \end{aligned}$$
(146)

In a similar fashion, we have

$$\begin{aligned} \mathcal {L}^{-1}_s\left[ \Delta t Z^*[Y(n-1)|s,\Delta t] \Bigg \vert t \right]= & {} \sum _{n=0}^\infty \tilde{Y}(n\Delta t) \delta (t-(n+1)\Delta t)\Delta t \end{aligned}$$
(147)
$$\begin{aligned}= & {} \sum _{n=0}^\infty \tilde{Y}((n-1)\Delta t)\delta (t-n\Delta t)\Delta t \end{aligned}$$
(148)
$$\begin{aligned}= & {} \tilde{Y}(t-\Delta t|\Delta t). \end{aligned}$$
(149)

Note that, with \(t'=n\Delta t\), in Eq. (146), we have

$$\begin{aligned} \lim _{\Delta t\rightarrow 0}\tilde{Y}(t|\Delta t)= & {} \lim _{\Delta t\rightarrow 0}\sum _{n=0}^\infty \tilde{Y}(n\Delta t) \delta (t-n\Delta t)\Delta t \end{aligned}$$
(150)
$$\begin{aligned}= & {} \int _0^\infty \tilde{Y}(t')\delta (t-t')\, \hbox {d}t' \end{aligned}$$
(151)
$$\begin{aligned}= & {} \tilde{Y}(t). \end{aligned}$$
(152)

This formally identifies

$$\begin{aligned} \tilde{Y}(t)=\lim _{\Delta t\rightarrow 0} \mathcal {L}^{-1}_s\left[ \Delta t Z^*[Y(n)|s,\Delta t] \Bigg \vert t\right] , \end{aligned}$$
(153)

provided that the limit exists.

We further note the product rule

$$\begin{aligned}&\lim _{\Delta t\rightarrow 0}\sum _{n=0}^\infty \tilde{X}(n\Delta t)\tilde{Y}(n\Delta t)\delta (t-n\Delta t)\Delta t\nonumber \\&=\left( \lim _{\Delta t\rightarrow 0}\sum _{n=0}^\infty \tilde{X}(n\Delta t)\delta (t-n\Delta t)\Delta t\right) \left( \lim _{\Delta t\rightarrow 0}\sum _{n=0}^\infty \tilde{Y}(n\Delta t)\delta (t-n\Delta t)\Delta t\right) ,\nonumber \\ \end{aligned}$$
(154)

which equates to \(\tilde{X}(t)\tilde{Y}(t)\) in each case, with \(t'=n\Delta t\), provided that both \(\tilde{X}(t)\) and \(\tilde{Y}(t)\) exist.

We now take the inverse Laplace transform of the Z star transform of Eq. (105) and multiply by \(\frac{\Delta t}{\Delta t}\) to write

$$\begin{aligned} \begin{aligned} \sum _{n=0}^\infty&\frac{\tilde{I}(n\Delta t)-\tilde{I}((n-1)\Delta t)}{\Delta t}\delta (t-n\Delta t)\Delta t\\&=\sum _{n=0}^\infty \frac{\tilde{\omega }(n\Delta t)}{\Delta t} \tilde{S}((n-1)\Delta t)\tilde{I}((n-1)\Delta t)\delta (t-n\Delta t)\Delta t\\&-\frac{r}{\Delta t}\sum _{n=0}^\infty \tilde{\theta }(n\Delta t,0)\left( \sum _{k=0}^{n-1}\tilde{\kappa }((n-k)\Delta t)\frac{\tilde{I}(k\Delta t)-\tilde{I}_0(k\delta t)}{\tilde{\theta }(k\Delta t,0)}\right) \delta (t-n\Delta t)\Delta t\\&-\sum _{n=0}^\infty \frac{\tilde{\gamma }(n\Delta t)}{\Delta t} \left( \tilde{I}((n-1)\Delta t)-\tilde{I}_0((n-1)\delta t)\right) \delta (t-n\Delta t)\Delta t\\&+\sum _{n=0}^\infty \frac{\tilde{I}_0(n\Delta t)-\tilde{I}_0((n-1)\Delta t)}{\Delta t}\delta (t-n\Delta t)\Delta t \end{aligned} \end{aligned}$$
(155)

We now take the continuous time limit of Eq. (155) using \(t'=n\Delta t\) and the product rule in Eq. (154) to obtain

$$\begin{aligned} \begin{aligned} \int _0^\infty&\left( \lim _{\Delta t\rightarrow 0}\frac{\tilde{I}(t')-\tilde{I}(t'-\Delta t)}{\Delta t}\right) \delta (t-t')\, \hbox {d}t'\\&\quad =\int _0^\infty \hat{\omega }(t') \tilde{S}(t')\tilde{I}(t')\delta (t-t')\, \hbox {d}t'\\&\qquad -\left( \int _0^\infty \tilde{\theta }(t',0)\delta (t-t')\, \hbox {d}t'\right) \\&\qquad \times \left( \lim _{\Delta t\rightarrow 0}\frac{r}{\Delta t}\sum _{n=0}^\infty \left( \sum _{k=0}^{n-1}\tilde{\kappa }((n-k)\Delta t)\frac{\tilde{I}(k\Delta t)-\tilde{I}_0(k\Delta t)}{\tilde{\theta }(k\Delta t,0)}\right) \delta (t-n\Delta t)\Delta t\right) \\&\qquad -\int _0^\infty \hat{\gamma }(t')\left( \tilde{I}(t')-\tilde{I}_0(t')\right) \delta (t-t')\, \hbox {d}t'\\&\qquad +\int _0^\infty \left( \lim _{\Delta t\rightarrow 0}\frac{\tilde{I}_0(t')-\tilde{I}_0(t'-\Delta t)}{\Delta t}\right) \delta (t-t')\, \hbox {d}t' \end{aligned} \end{aligned}$$
(156)

where we have defined continuous time rate parameters

$$\begin{aligned} \hat{\omega }(t')=\lim _{\Delta t\rightarrow 0}\frac{\tilde{\omega }(n\Delta t)}{\Delta t} \end{aligned}$$
(157)

and

$$\begin{aligned} \hat{\gamma }(t')=\lim _{\Delta t\rightarrow 0}\frac{\tilde{\gamma }(n\Delta t)}{\Delta t}. \end{aligned}$$
(158)

Equation (156) simplifies further to

$$\begin{aligned} \begin{aligned}&\frac{\hbox {d}\tilde{I}(t)}{\hbox {d}t}= \hat{\omega }(t)\tilde{S}(t)\tilde{I}(t)-\hat{\gamma }(t)\left( \tilde{I}(t)-\tilde{I}_0(t)\right) +\frac{\hbox {d}\tilde{I}_0(t)}{\hbox {d}t}\\&\quad -\tilde{\theta }(t,0)\left( \lim _{\Delta t\rightarrow 0} \frac{r}{\Delta t}\sum _{n=0}^\infty \left( \sum _{k=0}^{n-1}\tilde{\kappa }((n-k)\Delta t)\frac{\tilde{I}(k\Delta t)-\tilde{I}_0(k\Delta t)}{\tilde{\theta }(k\Delta t,0)}\right) \delta (t-n\Delta t)\Delta t\right) \end{aligned} \end{aligned}$$
(159)

The further reduction in this equation depends on the specific form of the memory kernel \(\kappa (n)\). In the case of a jump at each time step, the memory kernel is

$$\begin{aligned} \kappa (n)=\delta _{n,1}. \end{aligned}$$
(160)

In this case, we can perform the sum over k explicitly in Eq. (159) to arrive at

$$\begin{aligned} \frac{\hbox {d}\tilde{I}(t)}{\hbox {d}t}&= \hat{\omega }(t)\tilde{S}(t)\tilde{I}(t)-\hat{\gamma }(t)\left( \tilde{I}(t)-\tilde{I}_0(t)\right) +\frac{\hbox {d}\tilde{I}_0(t)}{\hbox {d}t}\nonumber \\&\quad -\,\tilde{\theta }(t,0)\left( \lim _{\Delta t\rightarrow 0} \frac{r}{\Delta t}\sum _{n=0}^\infty \frac{\tilde{I}((n-1)\Delta t)-\tilde{I}_0((n-1)\Delta t)}{\tilde{\theta }((n-1)\Delta t,0)}\delta (t-n\Delta t)\Delta t\right) \nonumber \\ \end{aligned}$$
(161)

In order for the continuous time limit of the above equation to exist, we define

$$\begin{aligned} \mu =\lim _{\Delta t\rightarrow 0}\frac{r}{\Delta t}. \end{aligned}$$
(162)

Note that r is a free parameter in the range [0, 1], and hence, \(\mu \) is only well defined in this limit if we take r to be a function of \(\Delta t\). With this definition of \(\mu \), we can now perform the limit \(\delta t\rightarrow 0\) to obtain the continuous time equation

$$\begin{aligned} \frac{\hbox {d}\tilde{I}(t)}{\hbox {d}t} =\hat{\omega }(t) \tilde{S}(t)\tilde{I}(t) -\mu (\tilde{I}(t)-\tilde{I}_0(t))-\hat{\gamma }(t)(\tilde{I}(t)-\tilde{I}_0(t))+\frac{\hbox {d}\tilde{I}_0(t)}{\hbox {d}t}. \end{aligned}$$
(163)

This further simplifies to

$$\begin{aligned} \frac{\hbox {d}\tilde{I}}{\hbox {d}t} =\hat{\omega }(t) \tilde{S}(t)\tilde{I}(t) -\mu \tilde{I}(t)-\hat{\gamma }(t)\tilde{I}(t) \end{aligned}$$
(164)

Equation (164) recovers the corresponding equation in the classic SIR model.

We now consider the continuous time limit of Eq. (159) with the Sibuya memory kernel, given by Eq. (104). First, we simplify the double sum in Eq. (159) using Laplace transforms and Z star transforms as follows:

$$\begin{aligned}&\lim _{\Delta t\rightarrow 0}\frac{r}{\Delta t}\sum _{n=0}^\infty \left( \sum _{k=0}^{n-1}\tilde{\kappa }((n-k)\Delta t)\frac{\tilde{I}(k\Delta t)-\tilde{I}_0(k\Delta t)}{\tilde{\theta }(k\Delta t,0)}\right) \delta (t-n\Delta t)\Delta t\nonumber \\&= \mathcal {L}_s^{-1}\left[ \mathcal {L}_t\left[ \lim _{\Delta t\rightarrow 0}\frac{r}{\Delta t} \sum _{n=0}^\infty \left( \sum _{k=0}^{n-1}\tilde{\kappa }((n-k)\Delta t)\frac{\tilde{I}(k\Delta t)-\tilde{I}_0(k\Delta t)}{\tilde{\theta }(k\Delta t,0)}\right) \delta (t-n\Delta t)\Delta t\Bigg \vert s\right] \Bigg \vert t\right] \nonumber \\&=\mathcal {L}_s^{-1}\left[ \lim _{\Delta t\rightarrow 0}\frac{r}{\Delta t} \sum _{n=0}^\infty \left( \sum _{k=0}^{n-1}\tilde{\kappa }((n-k)\Delta t)\frac{\tilde{I}(k\Delta t)-\tilde{I}_0(k\Delta t)}{\tilde{\theta }(k\Delta t,0)}\right) e^{-sn\Delta t}\Delta t\Bigg \vert t\right] \nonumber \\&=\mathcal {L}_s^{-1}\left[ \lim _{\Delta t\rightarrow 0}\frac{r}{\Delta t} Z^*\left[ \sum _{k=0}^{n-1}\kappa (n-k)\frac{I(k)-I_0(k)}{\theta (k,0)}| s,\Delta t \right] \Delta t\Bigg \vert t\right] \nonumber \\&= \mathcal {L}_s^{-1}\left[ \lim _{\Delta t\rightarrow 0}\frac{r}{\Delta t}Z^*[\kappa (n)| s,\Delta t] Z^{*}\left[ \frac{I(n)-I_0(k)}{\theta (n,0)}| s,\Delta t \right] \Delta t\Bigg \vert t\right] . \end{aligned}$$
(165)

The last line in the above follows from the convolution theorem for Z star transforms.

To proceed further, we use the Z transform of the Sibuya memory kernel in Eq. (114) to write

$$\begin{aligned} Z^*[\kappa (n)| s,\Delta t]= & {} \left[ ((1-e^{-s\Delta t})^{1-\alpha }-(1-e^{-s\Delta t}))\right] \end{aligned}$$
(166)
$$\begin{aligned}\approx & {} (s\Delta t)^{1-\alpha }+o(s\Delta t). \end{aligned}$$
(167)

The result in Eq. (165) can now be written as

$$\begin{aligned}&\lim _{\Delta t\rightarrow 0}\frac{r}{\Delta t}\sum _{n=0}^\infty \left( \sum _{k=0}^{n-1}\tilde{\kappa }((n-k)\Delta t)\frac{\tilde{I}(k\Delta t)-\tilde{I}_0(k\Delta t)}{\tilde{\theta }(k\Delta t,0)}\right) \delta (t-n\Delta t)\Delta t\nonumber \\&\quad = \mathcal {L}_s^{-1}\left[ \lim _{\Delta t\rightarrow 0}\frac{r}{\Delta t^\alpha } s^{1-\alpha } \sum _{n=0}^\infty \frac{\tilde{I}(n\Delta t)-\tilde{I}_0(n\Delta t)}{\tilde{\theta }(n\Delta t,0)}e^{-sn\Delta t}\Delta t\Bigg \vert t\right] \nonumber \\&\quad = \mu \mathcal {L}_s^{-1}\left[ s^{1-\alpha }\int _0^\infty \frac{\tilde{I}(t)-\tilde{I}_0(t)}{\tilde{\theta }(t,0)}e^{-st}\, \hbox {d}t\Bigg \vert t\right] \nonumber \\&\quad = \mu \mathcal {L}_s^{-1}\left[ s^{1-\alpha } \mathcal {L}_t \left[ \frac{\tilde{I}(t)-\tilde{I}_0(t)}{\tilde{\theta }(t,0)}\Bigg \vert s\right] \Bigg \vert t\right] , \end{aligned}$$
(168)

where

$$\begin{aligned} \mu =\lim _{\Delta t\rightarrow 0}\frac{r}{\Delta t^\alpha }. \end{aligned}$$
(169)

Finally, we substitute the result of Eq. (168) into Eq. (159) and use the known result (Podlubny 1999)

$$\begin{aligned} \mathcal {L}_t\left[ \, _0D_t^{1-\alpha } Y(t)\Bigg \vert s\right] = s^{1-\alpha } \mathcal {L}_t\left[ Y(t)\Bigg \vert s\right] \end{aligned}$$
(170)

to invert the Laplace transform and obtain

$$\begin{aligned} \frac{\hbox {d}\tilde{I}(t)}{\hbox {d}t} =\hat{\tilde{\omega }}(t) \tilde{S}(t)\tilde{I}(t) - \mu \hat{\tilde{\theta }}(t,0) \, _0D_t^{1-\alpha }\left( \frac{\tilde{I}(t)-\tilde{I}_0(t)}{\hat{\tilde{\theta }}(t,0)}\right) - \hat{\tilde{\gamma }}(t)\left( \tilde{I}(t)-\tilde{I}_0(t)\right) +\frac{\hbox {d}\tilde{I}_0(t)}{\hbox {d}t} \end{aligned}$$
(171)

Equation (171) recovers the continuous time frSIR model equation.

Note that in order for the continuous time limit of the frSIR model equation to exist, we defined

$$\begin{aligned} \mu =\lim _{\Delta t\rightarrow 0}\frac{r}{\Delta t^\alpha } \end{aligned}$$
(172)

which requires \(r\in [0,1]\) to be a function of \(\Delta t\). This is important for numerical simulations based on this DTRW method where we take \(r=\mu \Delta t^\alpha \) and then the requirement that \(r\in [0,1]\) places restrictions on \(\Delta t\) for given \(\alpha \) and \(\mu \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Angstmann, C.N., Henry, B.I. & McGann, A.V. A Fractional Order Recovery SIR Model from a Stochastic Process. Bull Math Biol 78, 468–499 (2016). https://doi.org/10.1007/s11538-016-0151-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-016-0151-7

Keywords

Mathematics Subject Classification

Navigation