Recurrent epidemic cycles driven by intervention in a population of two susceptibility types

Similar documents
Chapter 4 An Introduction to Networks in Epidemic Modeling

Thursday. Threshold and Sensitivity Analysis

Introduction to SEIR Models

Transmission of Hand, Foot and Mouth Disease and Its Potential Driving

Qualitative Analysis of a Discrete SIR Epidemic Model

Introduction to Stochastic SIR Model

Bifurcations in an SEIQR Model for Childhood Diseases

A BINOMIAL MOMENT APPROXIMATION SCHEME FOR EPIDEMIC SPREADING IN NETWORKS

STABILITY ANALYSIS OF A GENERAL SIR EPIDEMIC MODEL

The Existence and Stability Analysis of the Equilibria in Dengue Disease Infection Model

Electronic appendices are refereed with the text. However, no attempt has been made to impose a uniform editorial style on the electronic appendices.

arxiv: v3 [q-bio.pe] 2 Jun 2009

GLOBAL DYNAMICS OF A MATHEMATICAL MODEL OF TUBERCULOSIS

Stability of SEIR Model of Infectious Diseases with Human Immunity

Simulating stochastic epidemics

Forecast and Control of Epidemics in a Globalized World

PARAMETER ESTIMATION IN EPIDEMIC MODELS: SIMPLIFIED FORMULAS

Transmission in finite populations

A Generic Multivariate Distribution for Counting Data

HETEROGENEOUS MIXING IN EPIDEMIC MODELS

A Note on the Spread of Infectious Diseases. in a Large Susceptible Population

Approximation of epidemic models by diffusion processes and their statistical inferencedes

Coherence of Noisy Oscillators with Delayed Feedback Inducing Multistability

Stochastic models in biology and their deterministic analogues

Dynamical Analysis of Plant Disease Model with Roguing, Replanting and Preventive Treatment

A Time Since Recovery Model with Varying Rates of Loss of Immunity

arxiv: v2 [q-bio.pe] 7 Nov 2015

A prudent adaptive behaviour accelerates disease transmission on networks Supplementary Information

Preventive behavioural responses and information dissemination in network epidemic models

Global Analysis of an Epidemic Model with Nonmonotone Incidence Rate

GLOBAL STABILITY OF SIR MODELS WITH NONLINEAR INCIDENCE AND DISCONTINUOUS TREATMENT

Quarantine generated phase transition in epidemic spreading. Abstract

Epidemic reemergence in adaptive complex networks

Three Disguises of 1 x = e λx

arxiv: v1 [physics.soc-ph] 9 Nov 2012

Phase Transitions of an Epidemic Spreading Model in Small-World Networks

A review on Lamb's atmospheric oscillations using initial value problem approach

The Spatial Perspective

Role of GIS in Tracking and Controlling Spread of Disease

Analysis and Monte Carlo simulations of a model for the spread of infectious diseases in heterogeneous metapopulations

Estimating the Exponential Growth Rate and R 0

Final Project Descriptions Introduction to Mathematical Biology Professor: Paul J. Atzberger. Project I: Predator-Prey Equations

Applied Mathematics Letters

Spatial Heterogeneity in Epidemic Models

Models of Infectious Disease Formal Demography Stanford Summer Short Course James Holland Jones, Instructor. August 15, 2005

Non-Linear Models Cont d: Infectious Diseases. Non-Linear Models Cont d: Infectious Diseases

Seasonally forced epidemics Biennial outbreaks of Measles in England and Wales

Lecture 10. Under Attack!

The effect of emigration and immigration on the dynamics of a discrete-generation population

arxiv: v1 [physics.soc-ph] 7 Jul 2015

The Fractional-order SIR and SIRS Epidemic Models with Variable Population Size

1 Disease Spread Model

A comparison of delayed SIR and SEIR epidemic models

An intrinsic connection between Richards model and SIR model

Impact of Travel Between Patches for Spatial Spread of Disease

Mathematical Analysis of Epidemiological Models III

Downloaded from:

Lightlike solitons with spin

Project 1 Modeling of Epidemics

Analysis of Numerical and Exact solutions of certain SIR and SIS Epidemic models

Emergence of diversity in a biological evolution model

Time Evolution of Disease Spread on Networks with Degree Heterogeneity

The death of an epidemic

Bounded Rationality Alters the Dynamics of Paediatric Immunization Acceptance

On modeling two immune effectors two strain antigen interaction

Global Stability of a Computer Virus Model with Cure and Vertical Transmission

Spotlight on Modeling: The Possum Plague

Dynamics of Disease Spread. in a Predator-Prey System

Dynamical models of HIV-AIDS e ect on population growth

Slow and fast dynamics model of a Malaria with Sickle-Cell genetic disease with multi-stage infections of the mosquitoes population

Numerical solution of stochastic epidemiological models

ECS 289 / MAE 298, Lecture 7 April 22, Percolation and Epidemiology on Networks, Part 2 Searching on networks

Exact solution of site and bond percolation. on small-world networks. Abstract

Final source eccentricity measured by HBT interferometry with the event shape selection

3 rd Workshop and Conference on Modeling Infectious Diseases

Research on State-of-Charge (SOC) estimation using current integration based on temperature compensation

KINETICS OF SOCIAL CONTAGION. János Kertész Central European University. SNU, June

Modeling the Existence of Basic Offspring Number on Basic Reproductive Ratio of Dengue without Vertical Transmission

Threshold Parameter for a Random Graph Epidemic Model

Modelling spatio-temporal patterns of disease

Australian Journal of Basic and Applied Sciences. Effect of Personal Hygiene Campaign on the Transmission Model of Hepatitis A

GLOBAL STABILITY OF THE ENDEMIC EQUILIBRIUM OF A TUBERCULOSIS MODEL WITH IMMIGRATION AND TREATMENT

New results on the existences of solutions of the Dirichlet problem with respect to the Schrödinger-prey operator and their applications

Determining the Critical Point of a Sigmoidal Curve via its Fourier Transform

Bifurcation Analysis in Simple SIS Epidemic Model Involving Immigrations with Treatment

THE STOCHASTIC DYNAMICS OF EPIDEMIC MODELS

Design of Oscillator Networks for Generating Signal with Prescribed Statistical Property

Analytic approximation for the modified Bessel function I 2/3

Statistical Inference for Stochastic Epidemic Models

Mathematical Analysis of Epidemiological Models: Introduction

Stochastic models, patterns formation and diffusion

Mathematical modelling and controlling the dynamics of infectious diseases

Dynamics Analysis of Anti-predator Model on Intermediate Predator With Ratio Dependent Functional Responses

CHALMERS, GÖTEBORGS UNIVERSITET. EXAM for COMPUTATIONAL BIOLOGY A. COURSE CODES: FFR 110, FIM740GU, PhD

Epidemiology. Schedule: Friday, 25 May. 9:30-11:00 Session 1 Basic models and in epidemiology. 11:30-13:00 Session 2 Tuberculosis, pertussis, malaria

Age-dependent branching processes with incubation

Transmission Dynamics of an Influenza Model with Vaccination and Antiviral Treatment

Applications in Biology

Seasonal forcing drives spatio-temporal pattern formation in rabies epidemics

Transcription:

Journal of Physics: Conference Series OPEN ACCESS Recurrent epidemic cycles driven by intervention in a population of two susceptibility types To cite this article: Drandreb Earl O Juanico 214 J. Phys.: Conf. Ser. 49 12188 View the article online for updates and enhancements. Related content - Dynamics of infectious diseases Kat Rock, Sam Brand, Jo Moir et al. - Mathematical Model Of Tuberculosis Transmission With Reccurent Infection And Vaccination J Nainggolan, Sudradjat Supian, A K Supriatna et al. - Which is more effective for suppressing an infectious disease: imperfect vaccination or defense against contagion? Kazuki Kuga and Jun Tanimoto This content was downloaded from IP address 148.251.232.83 on 12/11/218 at 6:51

2nd International Conference on Mathematical Modeling in Physical Sciences 213 Journal of Physics: Conference Series 49 (214) 12188 doi:1.188/1742-6596/49/1/12188 Recurrent epidemic cycles driven by intervention in a population of two susceptibility types Drandreb Earl O. Juanico Department of Mathematics, Ateneo de Manila University, Loyola Heights Quezon City, Philippines 118 E-mail: djuanico@ateneo.edu Abstract. Epidemics have been known to persist in the form of recurrence cycles. Despite intervention efforts through vaccination and targeted social distancing, infectious diseases like influenza continue to appear intermittently over time. I have undertaken an analysis of a stochastic epidemic model to explore the hypothesis that intervention efforts actually drive epidemic cycles. Time series from simulations of the model reveal oscillations exhibiting a similar temporal signature as influenza epidemics. The power-spectral density indicates a resonant frequency, which approximately corresponds to the apparent annual seasonality of influenza in temperate zones. Asymptotic solution to the backward Kolmogorov equation of the dynamics corresponds to an exponentially-decaying mean-exit time as a function of the intervention rate. Intervention must be implemented at a sufficiently high rate to extinguish the infection. The results demonstrate that intervention efforts can induce epidemic cycles, and that the temporal signature of cycles can provide early warning of imminent outbreaks. 1. Introduction Sentinel surveillance of influenza-like illnesses (ILI) done in several countries have gathered and presented evidence for recurrent epidemics [1, 2, 3, 4, 5]. Prevailing models based on the Susceptible-Infected-Recovered or SIR framework, however, do not inherently predict the cyclelike variation of R with time unless exogenous seasonal forcing is assumed [6]. The problem with that assumption is the weak correspondence of seasonal forcing with recurrent epidemic activity [7]. On the contrary, the apparent seasonality may emerge endogenously from the hostpathogen feedback interaction. The host s natural immunity along with intervention efforts induce a feedback loop with the pathogen s mutations via antigenic drift. Demographic noise is a likely feature of that interaction. In models of resonant amplification, endogenous forcing due to demographic stochasticity has been demonstrated [8]. However, those models do not clearly address how epidemic cycles can be extinguished through intervention initiatives. Here, I propose and analyze a mathematical model that generates epidemic cycles and could account for the observed time variation in R. The model illustrates how intervention promotes epidemic cycles, and how a time-varying R could serve as a lead indicator of an imminent outbreak. 2. Mathematical model Consider an infected population of size N = N 1 + N 2 with two susceptibility types of subpopulation sizes N 1 and N 2. The population dynamics of both types are similar in all respects but one: type 1 gets infected at a faster rate than type 2. The population dynamics Content from this work may be used under the terms of the Creative Commons Attribution 3. licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI. Published under licence by Ltd 1

2nd International Conference on Mathematical Modeling in Physical Sciences 213 Journal of Physics: Conference Series 49 (214) 12188 doi:1.188/1742-6596/49/1/12188 can be further described in terms of reactions. Recruitment processes are: 1 and 2 λ 22 or 1 λ respectively. Density-dependent removal are: k1 μ 11 or 2 μ 21; 12, wherein μ and λ are per-capita rates of infection of type 1 and 2, δ k or k2 δ k where k =1, 2. Interventioninduced removal is expressed as: k νφ, occurring at the rate νφ, which may be related to vaccine efficacy [9]. The function Φ = k {1,2} N k (N k 1) /[N(N 1)] encapsulates the susceptibility structure of the community relevant to targeted vaccination. The population size of both types are random variables drawn from a joint probability density function P(N 1,N 2 ; t) that evolves over time. The exact time evolution of P is given by the master equation constructed from the recruitment and removal processes. Assuming a community size Ω 1, a diffusion approximation of the master equation is the Fokker-Planck equation: P(n 1,n 2 ; t) t = { 2 j=1 [b j (n 1,n 2 )P]+ 1 n j 2Ω 2 n 2 j [a j (n 1,n 2 )P] }, (1) where n 1 = N 1 /Ωandn 2 = N 2 /Ω are dimensionless concentrations. The drift terms are: b 1 =[μ n 1 νφ] (n 1 + n 2 )andb 2 =[λ n 2 νφ] (n 1 + n 2 ); and the diffusion terms are: a 1 =[μ + n 1 + νφ] (n 1 + n 2 )anda 2 =[λ + n 2 + νφ] (n 1 + n 2 ). A further simplification of the model is made by setting μ + λ =1,withμ ( 1 2, 1) and λ (, 1 2 ). Linking the variables in the model with measurement requires a rescaling of dimensions. The time unit t c in the model is set at the clinical-onset serial interval, which is the period between the onset of symptoms in the index case and the onset of symptoms in any secondary cases. The serial interval is 3.6 days from careful estimates supplemented by laboratory testing [1]. The carrying capacity is set at Ω 1 6, which is the order of the size of a typical community. But the timescale of the system is set using a small dimensionless parameter ε.1, such that the effective capacity is ε Ω 1 4. A transcritical bifurcation of the model occurs when μ is a function of ν, as follows: μ = μ b (ν) = 1 2 + 1 (1 ν) 3. (2) 3 3ν On this manifold, it can be shown that the equilibrium concentration vector ñ =(ñ 1, ñ 2 )is asymptotically stable, where ñ 1 = ( ) ( ) 3 1+ 3ν and ñ 2 = ( ) ( ) 3 1 3ν. 3. Epidemic cycle and resonant frequency The model is examined in a region of the parameter space denoted by the set of ordered pairs, (ν, μ) K =[, 1] ( 1 2, 1). Let B denote the drift Jacobian, which is evaluated at the stable equilibrium ñ. The power spectral density S(ω) of the epidemic time series is inversely proportional to the expression ( ω 2 det B ) 2 +(Tr B) 2 ω 2, the real zero of which corresponds to the square of the resonant frequency [8]: 1 ν ω (μ, ν) 2 =2(2μ 1) ν 2 3ν 3 (1 ν)2. (3) The cycle exists if ω 2 >, which is satisfied in J = { (ν, μ) μ ( 1 2, 1) ; ν>ν b = μ 1 b [μ(ν)] > } K, andwhereν b is the inverse function from Eq. (2). Figure 1 illustrates the cycles generated by the model (using a stochastic simulation algorithm [11]) compared with a real time-series of ILI activity in the U.S. from 1997 to 21. ILI activity reported in other settings exhibit similar time series (not shown): Hong Kong, 1998-26 [1]; Melbourne, 1999-28 [2]; Réunion Island, 29-211 [3]; and several countries in continental Europe [4, 5]. Resonance is indicated by the peak in the spectral density. 2

2nd International Conference on Mathematical Modeling in Physical Sciences 213 Journal of Physics: Conference Series 49 (214) 12188 doi:1.188/1742-6596/49/1/12188 concentration Infected.1.8.6.4.2 (a) Type 1 Type 2 Both types Time average 1 2 3 4 5 6.1.8.6.4.2 (b) USA Weekly ILI 1997-213 Time average 1 2 3 4 5 6 time, t (days) (c) Power spectral density, log 1 S(ω) -1-2 -3-4 -5-6 -7-8 -9 ω.27 day -1 Simulation: ν=.67, μ=.58 USA Weekly ILI, 1997-21 Ensemble average.1.1.1 1 Frequency, ω (day -1 ) Figure 1. Epidemic time series and power-spectral density. (a) Simulated model with (ν, μ) = (.67,.58). The dashed line is a time average; (b) U.S. Weekly ILI data from Center for Disease Control website; (c) Comparison of power spectral density. Higher data resolution accounts for the longer plot for simulated model. The predicted resonant frequency ω 2.7 1 3 day 1 (or about 1 yr 1 ) is determined using Eq. (3). 4. Mean extinction time and stochastic amplification Extinguishing the epidemic is the primary objective of intervention. Extinction refers to the state wherein n 1 = n 2 =. The mean time of such occurrence is represented by T,whichis a solution to the 2D backward Kolmogorov equation (BKE) [12]. The asymptotic solution for Ω 1 is the mean extinction time given by: T = π Ωϕ (ñ 2 ) e 2Ωϕ() A(ñ 2 )ϕ (). (4) At the bifurcation manifold, where μ = μ b (ν) as in Eq. (2), the mean extinction time in Eq. (4) only depends explicitly on ν [see Fig. 2(a)], which simplifies the analysis. Figure 2(b), inset, illustrates how extinction time is determined for a time series from the simulations. Mean extinction time, log 1 T 2-2 -4-6 Time, T(ν) μ=.51 μ=.53 μ=.55 μ=.575.6.65.7.75.8.85.9.95 1 Intervention rate, ν (a).8.7.6.5.4.3.2.1 Resonant frequency, ω Mean Extinction Time, T (days) 14 12 1 8 6 4 2 ν.649 Infected concentration, n = n 1 +n 2 (b).18.16 n =.14 Extinction time, T.12.1.8.6.4.2 4 8 12 time, t (days) Ω = 1 5 Ω = 1 3 Ω = 1 2.6.65.7.75.8.85.9.95 1 Intervention rate, ν Figure 2. Mean extinction time of epidemic. (a) The mean extinction time T (ν) calculated from Eq. (4) superimposed with the resonant frequency ω (ν, μ) calculated from Eq. (3) for several values of μ. T (ν) diverges for ν ν, where ν.649. (b) The mean extinction time T is calculated as the average of several runs for a community size Ω and intervention rate ν. The inset illustrates how the extinction time is determined for a time series. The mean extinction time is measured at the bifurcation manifold, where only ν is the free parameter, in the region J of resonance. The observed extinction time for different Ω is presented in figure 2b. The trend is consistent with the existence of a singularity of T (ν) 3

2nd International Conference on Mathematical Modeling in Physical Sciences 213 Journal of Physics: Conference Series 49 (214) 12188 doi:1.188/1742-6596/49/1/12188 at ν = ν = 1 4 [2 ( 1+ 2 ) 1/3 ( ) 1/3 ] + 1+ 2.649. Cycles persist longer at resonant frequencies for which ν ν. The value of T isconfirmedtodivergeasν ν +. This is indicated by the higher T for higher Ω. It is expected that for ν< ν the extinction time is very large as the epidemic persists in time. On the other hand, for ν> ν the epidemic is extinguished in finite time which is shorter for larger ν. 5. Reproduction number and outbreak prediction The reproduction number denoted as R is commonly employed as a measure of the capability of a disease to spread as an epidemic. Based on the definition of R, the model finds a dynamic reproduction number given by 1 R (t) = n(t)+2νφ(t). (5) Figure 3 shows R (t) superimposed with the n(t) time series. The range of values is.6 < R < 2.2 which is consistent with typical ranges known empirically via parameter estimation methods [13]. Generally, the time evolution of R (t) inrelationton(t) is consistent with those found in empirical studies [14]. Reproduction Number, R 2 1.8 1.6 1.4 1.2 1.8 1 12 14 16 18 2 22 24 Time, t (days) R (t) n 1 (t) n 2 (t) n(t).1.8.6.4.2 Infected Concentration Figure 3. Reproduction number R varies in time. Equation 5 is calculated for each time t and superimposed the concentrations n 1 and n 2.Remarkably,R (t) peaks along with n 2 (t) and thus may serve as a suitable leading indicator for an imminent outbreak when the epidemic has spread to the more susceptible subpopulation of the community. 6. Conclusion The proposed model predicts an extinction time that decreases with increasing ν above a threshold value ν. Epidemic cycles emerge from the confluence of decreasing extinction time and presence of a resonant frequency. Thus, the path to extinction is in the form of cycles. Indeed, epidemic cycles are counterintuitively driven by intervention efforts, which are meant to extinguish the infection. Future work may focus on the analysis, using the proposed model, on epidemics of other communicable diseases requiring person-to-person transmission. One may also address the demographic effects of migration which in the present model is neglected. Moreover, the inclusion of susceptibility structure should enhance the design of existing community surveillance protocols. Consequently, early detection for outbreaks is made possible, which is especially relevant in the tropics where influenza does not display any apparent seasonality. References [1] L. Yang et al., PLoS ONE 3, e1399 (28). [2] M. Coory, K. Grant, and H. Kelly, Euro Surveill. 14, 19387 (29). [3] L. Filleul et al., Euro Surveill. 17, 2212 (212). 4

2nd International Conference on Mathematical Modeling in Physical Sciences 213 Journal of Physics: Conference Series 49 (214) 12188 doi:1.188/1742-6596/49/1/12188 [4] S.P.VanNoort,R.Águas, S. Ballesteros, and M. G. M. Gomes, J. Theor. Biol. 298, 131 (212). [5] E. Bonabeau, L. Toubiana, and A. Flahault, Proc. R. Soc. Lond. B 265, 2421 (1998). [6] L. Stone, R. Olinky, and A. Huppert, Nature 446, 533 (27); R. Olinky, A. Huppert, and L. Stone, J. Math. Biol. 56, 827 (28). [7] L. Willem et al., PLoS ONE 7, e48695 (212). [8] A. J. McKane and T. J. Newman, Phys. Rev. Lett. 94, 21812 (25); D. Alonso, A. J. McKane, and M. Pascual, J. R. Soc. Interface 4, 575 (27) [9] C. A. Diaz-Granados, M. Denis, and S. Plotkin, Vaccine 31, 49 (212). [1] B. J. Cowling et al., Epidemiol. 2, 344 (29). [11] D. T. Gillespie, J. Comput. Phys. 22, 43 (1976). T T [12] The 2D BKE is a boundary-value problem b 1 n 1 b 2 n 2 + a 1 2 T 2Ω n 2 1 + a 2 2 T 2Ω n 2 2 = 1, where T =atthe boundaries. [13] C. Fraser et al. (WHO Rapid Pandemic Assessment Collaboration), Science 324, 1557 (29). [14] N. M. Ferguson, D. A. T. Cummings, S. Cauchemez, C. Fraser, S. Riley, A. Meeyai, S. Iamsirithaworn, and D. S. Burke, Nature 437, 29 (25). 5