Skip to main content
Log in

Social distancing game and insurance investment in a pandemic

  • Original Research
  • Published:
Annals of Operations Research Aims and scope Submit manuscript

Abstract

We consider a heterogeneous Susceptible–Infected–Recovered epidemic model, calibrated to the COVID-19 pandemic characteristics. We study the equilibrium of a voluntary social distancing game on a network of individuals subject to epidemic risk. We quantify the absolute and relative utility gaps across age cohorts. We further introduce life insurance in the model, which serves to mitigate the loss in the severe individual outcomes. We find that in most cases, insurance decreases the risk of contagion in the network because more individuals can be incentivized to self-isolate. On the other hand, in the case when the insurer does not have sufficient information on the self-isolation strategy of the individual, insurance can introduce moral hazard. We find that when premiums cannot be sufficiently differentiated, individuals may choose not to socially distance. This illustrates the importance of disclosing self-isolation strategies to the insurer, or inferring these strategies for the various types. Partial monitoring of social distancing strategy can mitigate the two problems of moral hazard and adverse selection.

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
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. See e.g., https://www.abi.org.uk/products-and-issues/topics-and-issues/coronavirus-hub/life-insurance/ for information from the Association of British Insurers (ABI).

  2. Note that agents have private information. In a multi-period version of the game, it would be thus interesting to explore a Perfect Bayesian Nash Equilbrium concept.

References

  • Acemoglu, D., Chernozhukov, V., Werning, I., & Whinston, M. D. (2020). A multi-risk SIR model with optimally targeted lockdown. Working Paper 27102, National Bureau of Economic Research.

  • Akian, M., Ganassali, L., Gaubert, S., & Massoulié, L. (2020). Probabilistic and mean-field model of COVID-19 epidemics with user mobility and contact tracing. arXiv:2009.05304

  • Aldy, J. E., & Viscusi, W. K. (2007). Age differences in the value of statistical life: Revealed preference evidence. Review of Environmental Economics and Policy, 1(2), 241–260.

    Article  Google Scholar 

  • Aldy, J. E., & Viscusi, W. K. (2008). Adjusting the value of a statistical life for age and cohort effects. The Review of Economics and Statistics, 90(3), 573–581.

    Article  Google Scholar 

  • Alvarez, F. E., Argente, D., & Lippi, F. (2020). A simple planning problem for COVID-19 lockdown. National Bureau of Economic Research: Technical report.

    Book  Google Scholar 

  • Amini, H., & Minca, A. (2022). Epidemic spreading and equilibrium social distancing in heterogeneous networks. Dynamic Games and Applications, 12(1), 258–287.

    Article  Google Scholar 

  • Aron-Dine, A., Einav, L., Finkelstein, A., & Cullen, M. (2015). Moral hazard in health insurance: Do dynamic incentives matter? Review of Economics and Statistics, 97(4), 725–741.

    Article  Google Scholar 

  • Ball, F. & Sirl, D. (2016). Evaluation of vaccination strategies for sir epidemics on random networks incorporating household structure. Journal of Mathematical Biology, 76.

  • Ball, F., Sirl, D., & Trapman, P. (2009). Threshold behaviour and final outcome of an epidemic on a random network with household structure. Advances in Applied Probability, 41(3), 765–796.

    Article  Google Scholar 

  • Ball, F. G., Sirl, D. J., Trapman, P., et al. (2014). Epidemics on random intersection graphs. The Annals of Applied Probability, 24(3), 1081–1128.

    Article  Google Scholar 

  • Britton, T., Janson, S., & Martin-Löf, A. (2007). Graphs with specified degree distributions, simple epidemics, and local vaccination strategies. Advances in Applied Probability, 39(4), 922–948.

    Article  Google Scholar 

  • Cairns, A. J., Blake, D. P., Kessler, A., & Kessler, M. (2020). The impact of COVID-19 on future higher-age mortality. Available at SSRN 3606988.

  • Cevik, M., Tate, M., Lloyd, O., Maraolo, A. E., Schafers, J., & Ho, A. (2020). Sars-cov-2, sars-cov, and mers-cov viral load dynamics, duration of viral shedding, and infectiousness: a systematic review and meta-analysis. The Lancet Microbe.

  • Chen, R., Wong, K. A., & Lee, H. C. (2001). Age, period, and cohort effects on life insurance purchases in the us. Journal of Risk and Insurance, 303–327.

  • Cont, R., Kotlicki, A., & Xu, R. (2020). Modelling COVID-19 contagion: Risk assessment and targeted mitigation policies. Available at SSRN 3681507.

  • Cummins, J. D., & Tennyson, S. (1996). Moral hazard in insurance claiming: Evidence from automobile insurance. Journal of Risk and Uncertainty, 12(1), 29–50.

    Article  Google Scholar 

  • de Blaeij, A., Florax, R. J., Rietveld, P., & Verhoef, E. (2003). The value of statistical life in road safety: A meta-analysis. Accident Analysis & Prevention, 35(6), 973–986.

    Article  Google Scholar 

  • Draief, M., & Massouli, L. (2010). Epidemics and rumours in complex networks. Cambridge University Press.

    Google Scholar 

  • Einav, L., Finkelstein, A., Ryan, S. P., Schrimpf, P., & Cullen, M. R. (2013). Selection on moral hazard in health insurance. American Economic Review, 103(1), 178–219.

    Article  Google Scholar 

  • Fan, V. Y., Jamison, D. T., & Summers, L. H. (2018). Pandemic risk: How large are the expected losses? Bulletin of the World Health Organization, 96(2), 129.

    Article  Google Scholar 

  • Fang, H., Wang, L., & Yang, Y. (2020). Human mobility restrictions and the spread of the novel coronavirus (2019-ncov) in China. Journal of Public Economics, 191, 104272.

    Article  Google Scholar 

  • Farboodi, M., Jarosch, G., & Shimer, R. (2021). Internal and external effects of social distancing in a pandemic. Journal of Economic Theory, 196, 105293.

    Article  Google Scholar 

  • Ferguson, N. M., Cummings, D. A., Fraser, C., Cajka, J. C., Cooley, P. C., & Burke, D. S. (2006). Strategies for mitigating an influenza pandemic. Nature, 442(7101), 448–452.

    Article  Google Scholar 

  • Ferguson, N. M., Laydon, D., Nedjati-Gilani, G., Imai, N., Ainslie, K., Baguelin, M., Bhatia, S., Boonyasiri, A., & Cucunubá, Z., Cuomo-Dannenburg, G., et al. (2020). Impact of non-pharmaceutical interventions (NPIs) to reduce COVID-19 mortality and healthcare demand.

  • Flaxman, S., Mishra, S., Gandy, A., Unwin, H. J. T., Mellan, T. A., Coupland, H., Whittaker, C., Zhu, H., Berah, T., & Eaton, J. W., et al. (2020). Estimating the effects of non-pharmaceutical interventions on COVID-19 in Europe. Nature, 1–8.

  • Goldstein, P., Yeyati, E. L., & Sartorio, L. (2021). Lockdown fatigue: The diminishing effects of quarantines on the spread of COVID-19. Covid Economics, 67, 1–23.

    Google Scholar 

  • Greenstone, M. & Nigam, V. (2020). Does social distancing matter? University of Chicago, Becker Friedman Institute for Economics Working Paper, (2020–2026).

  • Haberman, S., & Renshaw, A. (2009). On age-period-cohort parametric mortality rate projections. Insurance: Mathematics and Economics, 45(2), 255–270.

    Google Scholar 

  • Janson, S., Luczak, M. J., & Windridge, P. (2014). Law of large numbers for the SIR epidemic on a random graph with given degrees. Random Structures & Algorithms, 45(4), 726–763.

    Article  Google Scholar 

  • Jones, C., Philippon, T., & Venkateswaran, V. (2021). Optimal mitigation policies in a pandemic: Social distancing and working from home. The Review of Financial Studies, 34(11), 5188–5223.

    Article  Google Scholar 

  • Kiss, I. Z., Miller, J. C., Simon, P. L., et al. (2017). Mathematics of epidemics on networks (p. 598). Springer.

  • Kniesner, T. J. & Viscusi, W. K. (2019). The value of a statistical life. Oxford Research Encyclopedia of Economics and Finance (pp. 19–15).

  • Linka, K., Peirlinck, M., & Kuhl, E. (2020). The reproduction number of COVID-19 and its correlation with public health interventions. Computational Mechanics, 66(4), 1035–1050.

    Article  Google Scholar 

  • Manski, C. F., & Molinari, F. (2021). Estimating the COVID-19 infection rate: Anatomy of an inference problem. Journal of Econometrics, 220(1), 181–192.

    Article  Google Scholar 

  • Miller, E., Hoschler, K., Hardelid, P., Stanford, E., Andrews, N., & Zambon, M. (2010). Incidence of 2009 pandemic influenza A H1N1 infection in England: A cross-sectional serological study. The Lancet, 375(9720), 1100–1108.

    Article  Google Scholar 

  • Mossong, J., Hens, N., Jit, M., Beutels, P., Auranen, K., Mikolajczyk, R., Massari, M., Salmaso, S., Tomba, G. S., Wallinga, J., et al. (2008). Social contacts and mixing patterns relevant to the spread of infectious diseases. PLoS Medicine, 5(3), e74.

    Article  Google Scholar 

  • Murphy, K. M., & Topel, R. H. (2006). The value of health and longevity. Journal of Political Economy, 114(5), 871–904.

    Article  Google Scholar 

  • Pastor-Satorras, R., Castellano, C., Van Mieghem, P., & Vespignani, A. (2015). Epidemic processes in complex networks. Reviews of Modern Physics, 87, 925–979.

    Article  Google Scholar 

  • Prem, K., Cook, A. R., & Jit, M. (2017). Projecting social contact matrices in 152 countries using contact surveys and demographic data. PLoS Computational Biology, 13(9), e1005697.

    Article  Google Scholar 

  • Prem, K., Liu, Y., Russell, T. W., Kucharski, A. J., Eggo, R. M., Davies, N., Flasche, S., Clifford, S., Pearson, C. A., Munday, J. D., et al. (2020). The effect of control strategies to reduce social mixing on outcomes of the COVID-19 epidemic in Wuhan, China: A modelling study. The Lancet Public Health, 5(5), e261–e270.

    Article  Google Scholar 

  • Rivieccio, B. A., Luconi, E., Boracchi, P., Pariani, E., Romanò, L., Salini, S., Castaldi, S., Biganzoli, E., & Galli, M. (2020). Heterogeneity of COVID-19 outbreak in Italy. Acta Bio Medica: Atenei Parmensis, 91(2), 31.

    Google Scholar 

  • Stegehuis, C., Hofstad, R., & Leeuwaarden, J. (2016). Epidemic spreading on complex networks with community structures. Scientific Reports, 6, 29748.

    Article  Google Scholar 

  • Toxvaerd, F. M. (2020). Equilibrium social distancing.

  • Verity, R., Okell, L. C., Dorigatti, I., Winskill, P., Whittaker, C., Imai, N., Cuomo-Dannenburg, G., Thompson, H., Walker, P. G., Fu, H., et al. (2020). Estimates of the severity of coronavirus disease 2019: A model-based analysis. The Lancet Infectious Diseases, 20(6), 669–677.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Andreea Minca.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

The authors acknowledge financial support from an AXA Research Fund Award on “Mitigating Risk in the wake of the COVID-19 Pandemic”.

A Final outcome of heterogeneous SIR epidemics

A Final outcome of heterogeneous SIR epidemics

In this appendix we state some results on the final outcome of heterogeneous SIR epidemic model, for a given individuals social distancing strategies profile across individuals, when contact takes place on general networks. We also provide an iterative algorithm for our baseline model to compute both the asymptotic Nash equilibrium, as well as the social optimum on random networks with given vertex degrees. We assume that the degrees are bounded from above by \(\Delta \). We are interested in \({\mathcal {S}}^{({\textbf{s}})}_f\) and \({\mathcal {R}}^{({\textbf{s}})}_f\) the final set of susceptible and removed individuals, respectively, when the individuals follow the social distancing strategy \({\textbf{s}}\).

1.1 A.1 Cohort targeting to bound epidemics in general networks

We first state some general conditions on the adjacency matrix of the interaction graph and epidemics parameters for the size of the epidemics to be small compared to the size of the network. This can be used in the case where the entire or part of the realized network is known, in order to understand the potential for a large epidemic. In particular, it is essential to choose a social distancing strategy, imposed by policy makers or by setting various incentives, so that the epidemic dies out. For example, if one population type is prone to a large epidemic, then the incentive to socially distance may come from higher costs of insurance for that population type. In particular, the suggestion is that in the context of a pandemic, one needs to take into account not only the individual risk but the risk at the level of an entire population that engages in a certain (type-dependent) social distancing strategy.

Let A denote the adjacency matrix of the social contact graph. The probability that an infected individual makes infectious contact with any susceptible neighbour with type t and social activity \(s=1\) is given by \(1-e^{-\beta _t}\). Given the social distancing profile \({{\textbf{s}}}\), we define the infection matrix \(B^{({{\textbf{s}}})}\) as

$$\begin{aligned} B^{({{\textbf{s}}})}_{ij}:= \left( 1-e^{-\beta _{t_j}s_j}\right) A_{ij}, \end{aligned}$$
(31)

for all \(i,j \in [n]\). Note that the infection rates are not necessarily symmetric and, in general, the matrix \(B^{({{\textbf{s}}})}\) might not be symmetric even (if) the adjacency matrix A is symmetric.

We first give a condition on the maximum row sum of the matrix B, which gives us an upper bound for the expected amplification of infected individuals \((|{\mathcal {R}}^{({\textbf{s}})}_f| - |{\mathcal {R}}(0)| ) /|{\mathcal {I}}(0)|\).

The set of removed individuals at the end of the epidemic \({\mathcal {R}}^{({\textbf{s}})}_f \setminus {\mathcal {R}}(0) \) is the set of recovered or dead during the epidemic and is same set of all individuals who have ever been infected starting from the initial seed.

Proposition A.1

Let \(B^{({{\textbf{s}}})}_i=\sum _{j=1}^n B^{({{\textbf{s}}})}_{ij}\) and \(B^{({{\textbf{s}}})}_{\max }=\max _i(B^{({{\textbf{s}}})}_i)\). If \(B^{({{\textbf{s}}})}_{\max }<1\), then

$$\begin{aligned} {\mathbb {E}}[|{\mathcal {R}}^{({{\textbf{s}}})}_f|]\le |{\mathcal {R}}(0)|+\frac{1}{1-B^{({{\textbf{s}}})}_{\max }} |{\mathcal {I}}(0)|, \end{aligned}$$
(32)

which in particular implies that for all \(k>0\),

$$\begin{aligned} {\mathbb {P}}\left( |{\mathcal {R}}^{({{\textbf{s}}})}_f|-|{\mathcal {R}}(0)|\ge \frac{k}{1-B^{({{\textbf{s}}})}_{\max }} |{\mathcal {I}}(0)| \right) \le \frac{1}{k}. \end{aligned}$$

Proof

Let \(p_i = {\mathbb {P}}(i \in {\mathcal {R}}_f/{\mathcal {R}}(0))\). Hence, \(p_i = 1\) if \(i \in {\mathcal {I}}(0)\) and otherwise \(p_i \le \sum _{j=1}^n B_{ji}^{({\textbf{s}})} p_j\), which writes for all \(i=1, 2, \dots , n\) as

$$\begin{aligned} p_i \le \mathrm{1\hspace{-0.90ex}1}\left( i \in {\mathcal {I}}(0)\right) + \sum _{j=1}^n B^{({\textbf{s}})}_{ji} p_j. \end{aligned}$$
(33)

We thus obtain

$$\begin{aligned} \sum _{i=1}^n p_i\le & {} \sum _{i=1}^n \mathrm{1\hspace{-0.90ex}1}(i \in {\mathcal {I}}(0)) + \sum _{i=1}^n \sum _{j=1}^n B^{({\textbf{s}})}_{ji} p_j \\= & {} |{\mathcal {I}}(0)| + \sum _{j=1}^n \left( \sum _{i=1}^n B^{({\textbf{s}})}_{ji}\right) p_j \le |{\mathcal {I}}(0)| + B^{({\textbf{s}})}_{\max } \sum _{j=1}^n p_j. \end{aligned}$$

We conclude \({\mathbb {E}}\left[ |{\mathcal {R}}_f|\right] - |{\mathcal {R}}(0)|= \sum _{i=1}^n p_i \le \frac{1}{1-B^{({\textbf{s}})}_{\max }} |{\mathcal {I}}(0)|.\) The second statement follows using the Markov inequality. Namely, for any \(k>0\) we have

$$\begin{aligned} {\mathbb {P}}\left( |{\mathcal {R}}^{({{\textbf{s}}})}_f|-|{\mathcal {R}}(0)|\ge \frac{k}{1-B^{({{\textbf{s}}})}_{\max }} |{\mathcal {I}}(0)| \right) \le \frac{1}{k}. \end{aligned}$$

by the Markov inequality. \(\square \)

We now consider the \(L_2\) norm of the matrix \(B^{({{\textbf{s}}})}\). Let \(\lambda _{\max }(B)=||B||_2\) be the largest singular value of B, which is the square root of the largest eigenvalue of the positive-semidefinite matrix \(B^TB\). The following proposition shows that the expected amplification is \(O(\sqrt{n})\) whenever the largest singular value is smaller than 1.

Proposition A.2

If \(\lambda _{\max }(B^{({{\textbf{s}}})})<1\), then

$$\begin{aligned} {\mathbb {E}}[|{\mathcal {R}}^{({{\textbf{s}}})}_f|]\le |{\mathcal {R}}(0)|+\frac{1}{1-\lambda _{\max }(B^{({{\textbf{s}}})})} \sqrt{n|{\mathcal {I}}(0)|}, \end{aligned}$$
(34)

which in particular implies that for all \(k>0\),

$$\begin{aligned} {\mathbb {P}}\left( |{\mathcal {R}}^{({{\textbf{s}}})}_f|-|{\mathcal {R}}(0)|\ge \frac{k}{1-\lambda _{\max }(B^{({{\textbf{s}}})})} \sqrt{n|{\mathcal {I}}(0)|} \right) \le \frac{1}{k}. \end{aligned}$$

Proof

Recall that from (33) we have

$$\begin{aligned} p_i \le \mathrm{1\hspace{-0.90ex}1}(i \in {\mathcal {I}}(0)) + \sum _{j=1}^n B^{({\textbf{s}})}_{ji} p_j. \end{aligned}$$

Let \({\textbf{p}} = [p_1, p_2, \dots , p_n]\) denote the vector with components \(p_i, {{\textbf{1}}}\) be the vector with all components equal to 1 and \({{\textbf{1}}}_{{\mathcal {I}}(0)}\) be the vector with component 1 for \(i\in {\mathcal {I}}(0)\) and 0 for \(i \notin {\mathcal {I}}(0)\). By Equation (33), we have

$$\begin{aligned} {{\textbf{p}}} \le {{\textbf{1}}}_{{\mathcal {I}}(0)} + B^{({\textbf{s}})}{{\textbf{p}}}. \end{aligned}$$

Denoting by \(|| \cdot ||\) the Euclidean norm, we have

$$\begin{aligned} ||{{\textbf{p}}}|| \le ||{{\textbf{1}}}_{{\mathcal {I}}(0)} + B{{\textbf{p}}} || \le ||{{\textbf{1}}}_{{\mathcal {I}}(0)}|| + ||B {{\textbf{p}}}|| \le \sqrt{|{\mathcal {I}}(0)|} + \lambda _{\max }(B^{({\textbf{s}})}) ||{{\textbf{p}}}||. \end{aligned}$$

We thus have for \(\lambda _{\max }(B^{({\textbf{s}})}) < 1\) that \(||{{\textbf{p}}}|| \le \frac{\sqrt{|{\mathcal {I}}(0)|}}{1-\lambda _{\max }(B^{({\textbf{s}})})}\). Furthermore by the Cauchy-Schwarz inequality,

$$\begin{aligned} {\mathbb {E}}\left[ |{\mathcal {R}}^{({\textbf{s}})}_f|\right] -|{\mathcal {R}}(0)| = \sum _{i \in [n]} p_i = ||{{\textbf{1}}}^T {{\textbf{p}}}|| \le ||{{\textbf{1}}}^T|| \ ||{{\textbf{p}}}|| = \sqrt{n} ||{{\textbf{p}}}||. \end{aligned}$$

We conclude (if \(\lambda _{\max }(B^{({\textbf{s}})}) < 1\))

$$\begin{aligned} {\mathbb {E}}\left[ |{\mathcal {R}}^{({\textbf{s}})}_f|\right] -|{\mathcal {R}}(0)| \le \frac{1}{1-\lambda _{\max }(B)} \sqrt{n|{\mathcal {I}}(0)|}, \end{aligned}$$

and the second statement follows using the Markov inequality.

1.2 A.2 Algorithms to find the Asymptotic Nash equilibrium

In this section, we provide an iterative algorithm for our baseline model to compute both the asymptotic Nash equilibrium, as well as the social optimum. We assume that the degrees are bounded from above by \(\Delta \). Remark that for all \(\epsilon >0\), by Condition \((C_2)\), one can always choose \(\Delta =\Delta _\epsilon \) such that \(\sum _{t\in {\mathcal {T}}}\sum _{d=\Delta _\epsilon }^{\infty }d\mu _{t,d}<\epsilon \).

Under the assumptions of the following algorithm is guaranteed to converge to the unique fixed point of the mapping

$$\begin{aligned} \gamma _{t,d} = 1-G_t\left( \frac{\pi _{t,d}}{\kappa _t(x ^*_\gamma )\left( 1- \left( x^*_\gamma +(1-x^*_\gamma )e^{-\beta _t}\right) ^d \right) }\right) , \end{aligned}$$

where \(x^*_\gamma \) is the unique fixed point in [0, 1] of equation

$$\begin{aligned} f_{\gamma }(x):= \frac{\lambda _R}{\lambda } + \alpha _S \sum _{t\in {\mathcal {T}}}\sum \limits _{d=0}^{\infty } \frac{d}{\lambda } \mu _{t,d} \left[ \gamma _{t,d}+(1-\gamma _{t,d}) \left( x+(1-x)e^{-\beta _t}\right) ^{d-1}\right] . \end{aligned}$$
\({\textbf {Step 0:}}\):

Fix the error tolerance \(\epsilon >0\). Set \(k=0\) and the initial fraction of individuals with degree \(d=0, \dots , \Delta \) and type \(t\in {\mathcal {T}}\) who self-isolate as \(\gamma _{t,d}^{(0)}=0\). Let \(x_0\) be the smallest solution \(x\in [0,1]\) of (fixed-point iteration algorithm with error tolerance \(\epsilon \))

$$\begin{aligned} f_{0}(x):= \frac{\lambda _R}{\lambda } + \alpha _S \sum _{t\in {\mathcal {T}}}\sum \limits _{d=0}^{\Delta } \frac{d}{\lambda } \mu _{t,d} \left[ \left( x+(1-x)e^{-\beta _t}\right) ^{d-1}\right] . \end{aligned}$$

This is the equilibrium under no self-isolation.

\({\textbf {Step 1:}}\):

Set \(k \leftarrow k+1\). For all \(d=0, \dots , \Delta \) and type \(t\in {\mathcal {T}}\), set the fraction of individuals with degree d and type t who self-isolate at step k as

$$\begin{aligned} \gamma _{t,d}^{(k)}:= 1-G_t\left( \frac{\pi _{t,d}}{\kappa _t(x_{k-1})\left( 1- \left( x_{k-1}+(1-x_{k-1})e^{-\beta _t}\right) ^d \right) }\right) , \end{aligned}$$

and let \(x_k\) be the unique fixed-point \(x\in [0,1]\) of

$$\begin{aligned} f_{k}(x) := \frac{\lambda _R}{\lambda } + \alpha _S \sum _{t\in {\mathcal {T}}}\sum \limits _{d=0}^{\Delta } \frac{d}{\lambda } \mu _{t,d} \left[ \gamma ^{(k)}_{t,d}+(1-\gamma ^{(k)}_{t,d}) \left( x+(1-x)e^{-\beta _t}\right) ^{d-1}\right] . \end{aligned}$$
\({\textbf {Step 2:}}\):

If \(x_{k}-x_{k-1}<\epsilon \), terminate the algorithm. Otherwise, return to Step 1.

Note that since F is a strictly increasing cumulative distribution function, the fixed point solutions sequence \(x_k\) will be increasing, converging to \(x^*_\gamma \) which is unique as proven in Amini and Minca (2022), Theorem 3.1.

Further, for individuals with type t and degree d, \(\pi _{t,d}(1-\gamma _{t,d})\) is the total payoff from social activity and \(\kappa _t(x^*_\gamma ) \int _\gamma ^1 G_t^{-1}(1-u)du\) is the average loss faced by the \((1-\gamma _{t,d})\)-fraction of individuals who are not following isolation and therefore are subject to epidemic risk.

It follows that the social utility averaged over the population converges to

$$\begin{aligned} \frac{1}{n} \sum _{i=1}^n u_i ({\varvec{\ell }}, \textbf{s}) {\mathop {\longrightarrow }\limits ^{p}} \bar{u}_\textrm{social}(\gamma ):=\sum _{t\in {\mathcal {T}}}\sum \limits _{d=0}^{\infty } \mu _{t,d}\bar{u}_{t,d}(\gamma ), \end{aligned}$$

with

$$\begin{aligned} \bar{u}_{t,d}(\gamma ):= \pi _{t,d}(1-\gamma _{t,d}) -\kappa _t(x^*_\gamma )\left( 1- \left( x^*_\gamma +(1-x^*_\gamma )e^{-\beta _t}\right) ^d \right) \int _{\gamma _{t,d}}^1 G_t^{-1}(1-u)du. \end{aligned}$$

The following algorithm is used to compute the social optimal network immunity fixed-point solution, as well as for social optimal isolation policy over different individuals:

\({\textbf {Step 0:}}\):

Set again \(k=0\), \(\gamma _{t,d}^{(0)}=0\) and let \(x_0\) be the smallest solution \(x\in [0,1]\) of

$$\begin{aligned} f_{0}(x):= \frac{\lambda _R}{\lambda } + \alpha _S \sum _{t\in {\mathcal {T}}}\sum \limits _{d=0}^{\Delta } \frac{d}{\lambda } \mu _{t,d} \left[ \left( x+(1-x)e^{-\beta _t}\right) ^{d-1}\right] . \end{aligned}$$
\({\textbf {Step 1}}:\):

Set \(k \leftarrow k+1\). For all \(d=0, \dots , \Delta \) and type \(t\in {\mathcal {T}}\), set the fraction of individuals with degree d and type t who self-isolate at step k as

$$\begin{aligned} \gamma _{t,d}^{(k)}:= & {} \underset{\gamma \in [0,1]}{\arg \max } \pi _{t,d}(1-\gamma ) -\kappa _t(x_{k-1})\left( 1- \left( x_{k-1}+(1-x_{k-1})e^{-\beta _t}\right) ^d \right) \\ {}{} & {} \times \int _{\gamma }^1 G_t^{-1}(1-u)du, \end{aligned}$$

and let \(x_k\) be the unique fixed-point \(x\in [0,1]\) of

$$\begin{aligned} f_{k}(x) := \frac{\lambda _R}{\lambda } + \alpha _S \sum _{t\in {\mathcal {T}}}\sum \limits _{d=0}^{\Delta } \frac{d}{\lambda } \mu _{t,d} \left[ \gamma ^{(k)}_{t,d}+(1-\gamma ^{(k)}_{t,d}) \left( x+(1-x)e^{-\beta _t}\right) ^{d-1}\right] . \end{aligned}$$
\({\textbf {Step 2:}}\):

If \(x_{k}-x_{k-1}<\epsilon \), terminate the algorithm. Otherwise, return to Step 1.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Amini, H., Minca, A. Social distancing game and insurance investment in a pandemic. Ann Oper Res 336, 2009–2036 (2024). https://doi.org/10.1007/s10479-023-05485-1

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-023-05485-1

Keywords

Navigation