A Review of Agent-Based Model Simulation for Covid 19 Spread

  • Conference paper
  • First Online: 13 December 2022
  • Cite this conference paper

scientific method mystery virus case study simulation

  • Samar Ibrahim 13  

Part of the book series: Lecture Notes in Networks and Systems ((LNNS,volume 573))

Included in the following conference series:

  • International Conference on Emerging Technologies and Intelligent Systems

646 Accesses

1 Citations

1 Altmetric

Containing the implications of the Covid-19 pandemic has and continues to be a priority around the globe. Pursuant to this, scientists have focused on understanding the virus spread’s behavior and patterns to develop mitigation plans. Artificial Intelligence agent-based simulations (ABS) have been used by scientists to simulate virus spread and control. ABS can model different variables and decisions in specific environments and contexts to find ways to reduce the transmission and diminish the severity of this pandemic. This paper presents a review of literature on ABS modeling to contain the virus through pharmaceutical and non-pharmaceutical interventions and the impact of the virus spread on economies.

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

Access this chapter

  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
  • Available as EPUB and PDF
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Castro, B.M., de Abreu de Melo, Y., Fernanda dos Santos, N., Luiz da Costa Barcellos, A., Choren, R., Salles, R.M.: Multi-agent simulation model for the evaluation of COVID-19 transmission. Comput. Biol. Med. 136 (2021)

Google Scholar  

Villavicencio, C., Macrohon, J.J., Inbaraj, X.A., Jeng, J., Hsieh, J.: Twitter sentiment analysis towards COVID-19 vaccines in the Philippines using Naïve Bayes. Informatiom 12 (2021)

Lorig, F., Johansson, E., Davidsson, P.: Agent-based social simulation of the covid-19 pandemic: a systematic review. Jasss 24 (3) (2021)

Cardoso, R.C., Ferrando, A.: A review of agent-based programming for multi-agent systems. Computers 10 (2), 1–15 (2021)

Article   Google Scholar  

Sebastien, R., Olivier, M., Andrei, D.: Use of fuzzy sets, aggregation operators and multi agent systems to simulate COVID-19 transmission in a context of absence of barrier gestures and social distancing: Application to an island region. Proceedings - 2020 IEEE International Conference Bioinformation Biomedical BIBM 2020, pp. 2298–2305 (2020)

Wei, Y., Wang, J., Song, W., Xiu, C., Ma, L., Pei, T.: Spread of COVID-19 in China: analysis from a city-based epidemic and mobility model. Cities 110 (2021)

Philip, J.R.P., Gressman, T.: Simulating Covid -19 in a university environment. Math. Biosci. 388 , 539–547 (2020)

Albahri, O.S., et al.: Systematic review of artificial intelligence techniques in the detection and classification of COVID-19 medical images in terms of evaluation and benchmarking: Taxonomy analysis, challenges, future solutions, and methodological aspects. J. Infect. Public Health 13 (10), 1381–1396 (2020)

Alamoodi, A.H. et al.: Sentiment analysis and its applications in fighting COVID-19 and infectious diseases: a systematic review. Expert Syst. Appl. 167 (2021)

Shamil Salman, I., Farhanaz, F.: An agent-based modeling of COVID-19 validation, analysis, and recommendations. Cognit. Comput. (2021)

Mukherjee, U.K. et al.: Evaluation of reopening strategies for educational institutions during COVID-19 through agent based simulation. Scientific Reports 11 (1) (2021)

Catching, A., Capponi, S., Te Yeh, M., Bianco, S., Andino, R.: Examining the interplay between face mask usage, asymptomatic transmission, and social distancing on the spread of COVID-19. Scientific Reports 11 (1) (2021)

Li, K.K.F., Jarvis, S.A., Minhas, F.: Elementary effects analysis of factors controlling COVID-19 infections in computational simulation reveals the importance of social distancing and mask usage. Comput. Biol. Med. 134 (2021)

Almagor, J., Picascia, S.: Exploring the effectiveness of a COVID-19 contact tracing app using an agent-based model. Scientific Reports 10 (1) (2020)

D’Orazio, M., Bernardini, G., Quagliarini, E.: Sustainable and resilient strategies for touristic cities against COVID-19: an agent-based approach. Safety Sci. 142 (2021)

Asgary, A., Cojocaru, M.G., Najafabadi, M.M., Wu, J.: Simulating preventative testing of SARS-CoV-2 in schools: policy implications. BMC Public Health 21 (1) (2021)

Silva, P.C.L., Batista, P.V.C., Lima, H.S., Alves, M.A., Guimarães, F.G., Silva, R.C.P.: COVID-ABS: an agent-based model of COVID-19 epidemic to simulate health and economic effects of social distancing interventions. Chaos Solitons Fractals 139 (2020)

Truszkowska, A. et al.: Designing the safe reopening of US towns through high-resolution agent-based modeling. Adv. Theory Simul. 4 (9) (2021)

Sulis, E., Terna, P.: An agent-based decision support for a vaccination campaign. J. Med. Syst. 45 (11) (2021)

Download references

Author information

Authors and affiliations.

British University in Dubai, Dubai, UAE

Samar Ibrahim

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Samar Ibrahim .

Editor information

Editors and affiliations.

Department of Business Analytics, Sunway University, Subang Jaya, Selangor, Malaysia

Mohammed A. Al-Sharafi

The British University in Dubai, Dubai, United Arab Emirates

Mostafa Al-Emran

Al Buraimi University College, Al Buraimi, Oman

Mohammed Naji Al-Kabi

Khaled Shaalan

Rights and permissions

Reprints and permissions

Copyright information

© 2023 The Author(s), under exclusive license to Springer Nature Switzerland AG

About this paper

Cite this paper.

Ibrahim, S. (2023). A Review of Agent-Based Model Simulation for Covid 19 Spread. In: Al-Sharafi, M.A., Al-Emran, M., Al-Kabi, M.N., Shaalan, K. (eds) Proceedings of the 2nd International Conference on Emerging Technologies and Intelligent Systems . ICETIS 2022. Lecture Notes in Networks and Systems, vol 573. Springer, Cham. https://doi.org/10.1007/978-3-031-20429-6_53

Download citation

DOI : https://doi.org/10.1007/978-3-031-20429-6_53

Published : 13 December 2022

Publisher Name : Springer, Cham

Print ISBN : 978-3-031-20428-9

Online ISBN : 978-3-031-20429-6

eBook Packages : Intelligent Technologies and Robotics Intelligent Technologies and Robotics (R0)

Share this paper

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Publish with us

Policies and ethics

  • Find a journal
  • Track your research

STEM Printables - Foldables  Graphic Organizers for Interactive Science Activities

Mystery Virus Epidemiology | Case Study Simulation

by Paul | Digital Classroom

Mystery Virus Epidemiology

This highly engaging online simulation asks students to take on the role of a Public Health Director to determine the cause of a recent mystery killer virus epidemic outbreak at a local High School and Senior Center.

In their role as Public Health Director, students are guided by an expert epidemiologist through the entire case study simulation which consists of:

  • Studying medical reports and determining which ones to follow-up on
  • Collecting information and identifying patterns of infection from patients
  • Using the Scientific Method, develop & test a hypothesis to explain the outbreak.
  • Evaluate epidemiological testing to determine what pathogen is causing the disease.
  • Use epidemiology methods to determine the source and how the pathogen spreads.
  • Make a recommendation on which patients should be isolated and quarantined.

Distance Learning: This simulation case study is appropriate for independent student work and should require minimal or no additional instructions from the teacher, The online simulation takes the student thru the entire process, step-by-step, with clear and concise directions from a virtual “expert epidemiologist”.

Topics Covered : Scientific Method, Epidemiologist, Epidemic, Obtaining Medical records to confirm illness symptoms, Mapping locations of confirmed cases, Establish timelines and travel histories, Developing Hypothesis for what is causing illnesses and how it is spreading, Viral or Bacterial Culture, Blood Cultures, Rapid PCR Test, Blood Test for Antibodies, Case-Definition, Case-Control Study, Longitudinal Study, Analytic Epidemiology, Odds-Ratio

Check out this product on my TpT Store now!

virus outbreak simulation

Library homepage

  • school Campus Bookshelves
  • menu_book Bookshelves
  • perm_media Learning Objects
  • login Login
  • how_to_reg Request Instructor Account
  • hub Instructor Commons

Margin Size

  • Download Page (PDF)
  • Download Full Book (PDF)
  • Periodic Table
  • Physics Constants
  • Scientific Calculator
  • Reference & Cite
  • Tools expand_more
  • Readability

selected template will load here

This action is not available.

Biology LibreTexts

14.7: Activity – Laboratory Simulation of Viral Infection

  • Last updated
  • Save as PDF
  • Page ID 24381

\( \newcommand{\vecs}[1]{\overset { \scriptstyle \rightharpoonup} {\mathbf{#1}} } \)

\( \newcommand{\vecd}[1]{\overset{-\!-\!\rightharpoonup}{\vphantom{a}\smash {#1}}} \)

\( \newcommand{\id}{\mathrm{id}}\) \( \newcommand{\Span}{\mathrm{span}}\)

( \newcommand{\kernel}{\mathrm{null}\,}\) \( \newcommand{\range}{\mathrm{range}\,}\)

\( \newcommand{\RealPart}{\mathrm{Re}}\) \( \newcommand{\ImaginaryPart}{\mathrm{Im}}\)

\( \newcommand{\Argument}{\mathrm{Arg}}\) \( \newcommand{\norm}[1]{\| #1 \|}\)

\( \newcommand{\inner}[2]{\langle #1, #2 \rangle}\)

\( \newcommand{\Span}{\mathrm{span}}\)

\( \newcommand{\id}{\mathrm{id}}\)

\( \newcommand{\kernel}{\mathrm{null}\,}\)

\( \newcommand{\range}{\mathrm{range}\,}\)

\( \newcommand{\RealPart}{\mathrm{Re}}\)

\( \newcommand{\ImaginaryPart}{\mathrm{Im}}\)

\( \newcommand{\Argument}{\mathrm{Arg}}\)

\( \newcommand{\norm}[1]{\| #1 \|}\)

\( \newcommand{\Span}{\mathrm{span}}\) \( \newcommand{\AA}{\unicode[.8,0]{x212B}}\)

\( \newcommand{\vectorA}[1]{\vec{#1}}      % arrow\)

\( \newcommand{\vectorAt}[1]{\vec{\text{#1}}}      % arrow\)

\( \newcommand{\vectorB}[1]{\overset { \scriptstyle \rightharpoonup} {\mathbf{#1}} } \)

\( \newcommand{\vectorC}[1]{\textbf{#1}} \)

\( \newcommand{\vectorD}[1]{\overrightarrow{#1}} \)

\( \newcommand{\vectorDt}[1]{\overrightarrow{\text{#1}}} \)

\( \newcommand{\vectE}[1]{\overset{-\!-\!\rightharpoonup}{\vphantom{a}\smash{\mathbf {#1}}}} \)

Simulation of Viral Infection

The test performed in the lab is a simulation. No virus, living materials, or biohazardous reagents are used. The series of test tubes are each filled with a fluid that represents body fluid one might exchange with another individual.

One tube of fluid is “positive”. You cannot identify this cup by visual inspection. You will perform an epidemiological study to determine the source of the positive fluid.

  • Your instructor will give you a cup of fluid. Record the ID (letter) of the cup.
  • During your interaction, each of you will exchange fluid by pouring some of your fluid into your partners cup, and having your partner pour some of their fluid into your cup.
  • You will be handed a behavior cards. a) Step 1: Monogamous trade with person to their right. b) Step 2: Promiscuous people stand up go to one side of the room and everyone trade with each other. c) Step 3: Cheaters stand up, go next to promiscuous, and trade with either a cheater or promiscuous. d) Step 4: One night stands stand up and trade with one other one night stand or promiscuous then sit down. e) Step 5: Promiscuous trade with each other. f) Step 6: Cheaters trade with same person from before. g) Step 7: Everyone sit down. h) Step 8: Monogamous trade once more with partner.
  • The instructor will perform add pH indicator to your tube, tubes that turn pink indicate infection.
  • Trace back the path the transmission of the disease.

ASU for You, learning resources for everyone

  • News/Events
  • Arts and Sciences
  • Design and the Arts
  • Engineering
  • Global Futures
  • Health Solutions
  • Nursing and Health Innovation
  • Public Service and Community Solutions
  • University College
  • Thunderbird School of Global Management
  • Polytechnic
  • Downtown Phoenix
  • Online and Extended
  • Lake Havasu
  • Research Park
  • Washington D.C.
  • Biology Bits
  • Bird Finder
  • Coloring Pages
  • Experiments and Activities
  • Games and Simulations
  • Quizzes in Other Languages
  • Virtual Reality (VR)
  • World of Biology
  • Meet Our Biologists
  • Listen and Watch
  • PLOSable Biology
  • All About Autism
  • Xs and Ys: How Our Sex Is Decided
  • When Blood Types Shouldn’t Mix: Rh and Pregnancy
  • What Is the Menstrual Cycle?
  • Understanding Intersex
  • The Mysterious Case of the Missing Periods
  • Summarizing Sex Traits
  • Shedding Light on Endometriosis
  • Periods: What Should You Expect?
  • Menstruation Matters
  • Investigating In Vitro Fertilization
  • Introducing the IUD
  • How Fast Do Embryos Grow?
  • Helpful Sex Hormones
  • Getting to Know the Germ Layers
  • Gender versus Biological Sex: What’s the Difference?
  • Gender Identities and Expression
  • Focusing on Female Infertility
  • Fetal Alcohol Syndrome and Pregnancy
  • Ectopic Pregnancy: An Unexpected Path
  • Creating Chimeras
  • Confronting Human Chimerism
  • Cells, Frozen in Time
  • EvMed Edits
  • Stories in Other Languages
  • Virtual Reality
  • Zoom Gallery
  • Ugly Bug Galleries
  • Ask a Question
  • Top Questions
  • Question Guidelines
  • Permissions
  • Information Collected
  • Author and Artist Notes
  • Share Ask A Biologist
  • Articles & News
  • Our Volunteers
  • Teacher Toolbox

Question icon

Practice the Scientific Method

You may not know it, but science is everywhere. You practice the scientific method every day as you move through the world, solving problems and learning things. Want to see for yourself how often you might use the scientific method? Play The Case of the Mystery Images . Then, learn more at Using the Scientific Method to Solve Mysteries .

View Citation

Bibliographic details:.

  • Article: Using the Scientific Method: The Case of the Mystery Images
  • Author(s): Dr. Biology
  • Publisher: Arizona State University School of Life Sciences Ask A Biologist
  • Site name: ASU - Ask A Biologist
  • Date published: July 23, 2019
  • Date accessed: May 15, 2024
  • Link: https://askabiologist.asu.edu/games-and-simulations/mystery-images

Dr. Biology. (2019, July 23). Using the Scientific Method: The Case of the Mystery Images. ASU - Ask A Biologist. Retrieved May 15, 2024 from https://askabiologist.asu.edu/games-and-simulations/mystery-images

Chicago Manual of Style

Dr. Biology. "Using the Scientific Method: The Case of the Mystery Images". ASU - Ask A Biologist. 23 July, 2019. https://askabiologist.asu.edu/games-and-simulations/mystery-images

MLA 2017 Style

Dr. Biology. "Using the Scientific Method: The Case of the Mystery Images". ASU - Ask A Biologist. 23 Jul 2019. ASU - Ask A Biologist, Web. 15 May 2024. https://askabiologist.asu.edu/games-and-simulations/mystery-images

New to the scientific method? You might want to try our Training Room Escape before solving The Case of the Mystery Images.

Be Part of Ask A Biologist

By volunteering, or simply sending us feedback on the site. Scientists, teachers, writers, illustrators, and translators are all important to the program. If you are interested in helping with the website we have a Volunteers page to get the process started.

Share to Google Classroom

  • Open access
  • Published: 07 June 2021

A comparison of five epidemiological models for transmission of SARS-CoV-2 in India

  • Soumik Purkayastha 1 ,
  • Rupam Bhattacharyya 1 ,
  • Ritwik Bhaduri 2 ,
  • Ritoban Kundu 2 ,
  • Xuelin Gu 1 , 3 ,
  • Maxwell Salvatore 1 , 3 , 4 ,
  • Debashree Ray 5 , 6 ,
  • Swapnil Mishra 7 &
  • Bhramar Mukherjee   ORCID: orcid.org/0000-0003-0118-4561 1 , 3 , 4  

BMC Infectious Diseases volume  21 , Article number:  533 ( 2021 ) Cite this article

20k Accesses

25 Citations

15 Altmetric

Metrics details

Many popular disease transmission models have helped nations respond to the COVID-19 pandemic by informing decisions about pandemic planning, resource allocation, implementation of social distancing measures, lockdowns, and other non-pharmaceutical interventions. We study how five epidemiological models forecast and assess the course of the pandemic in India: a baseline curve-fitting model, an extended SIR (eSIR) model, two extended SEIR (SAPHIRE and SEIR-fansy) models, and a semi-mechanistic Bayesian hierarchical model (ICM).

Using COVID-19 case-recovery-death count data reported in India from March 15 to October 15 to train the models, we generate predictions from each of the five models from October 16 to December 31. To compare prediction accuracy with respect to reported cumulative and active case counts and reported cumulative death counts, we compute the symmetric mean absolute prediction error (SMAPE) for each of the five models. For reported cumulative cases and deaths, we compute Pearson’s and Lin’s correlation coefficients to investigate how well the projected and observed reported counts agree. We also present underreporting factors when available, and comment on uncertainty of projections from each model.

For active case counts, SMAPE values are 35.14% (SEIR-fansy) and 37.96% (eSIR). For cumulative case counts, SMAPE values are 6.89% (baseline), 6.59% (eSIR), 2.25% (SAPHIRE) and 2.29% (SEIR-fansy). For cumulative death counts, the SMAPE values are 4.74% (SEIR-fansy), 8.94% (eSIR) and 0.77% (ICM). Three models (SAPHIRE, SEIR-fansy and ICM) return total (sum of reported and unreported) cumulative case counts as well. We compute underreporting factors as of October 31 and note that for cumulative cases, the SEIR-fansy model yields an underreporting factor of 7.25 and ICM model yields 4.54 for the same quantity. For total (sum of reported and unreported) cumulative deaths the SEIR-fansy model reports an underreporting factor of 2.97. On October 31, we observe 8.18 million cumulative reported cases, while the projections (in millions) from the baseline model are 8.71 (95% credible interval: 8.63–8.80), while eSIR yields 8.35 (7.19–9.60), SAPHIRE returns 8.17 (7.90–8.52) and SEIR-fansy projects 8.51 (8.18–8.85) million cases. Cumulative case projections from the eSIR model have the highest uncertainty in terms of width of 95% credible intervals, followed by those from SAPHIRE, the baseline model and finally SEIR-fansy.

Conclusions

In this comparative paper, we describe five different models used to study the transmission dynamics of the SARS-Cov-2 virus in India. While simulation studies are the only gold standard way to compare the accuracy of the models, here we were uniquely poised to compare the projected case-counts against observed data on a test period. The largest variability across models is observed in predicting the “total” number of infections including reported and unreported cases (on which we have no validation data). The degree of under-reporting has been a major concern in India and is characterized in this report. Overall, the SEIR-fansy model appeared to be a good choice with publicly available R-package and desired flexibility plus accuracy.

Peer Review reports

Coronavirus disease 2019 (COVID-19) is an infectious disease caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) [ 1 ]. At the time of revising this paper (March 24, 2021), roughly 124 million cases have been reported worldwide. The disease was first identified in Wuhan, Hubei Province, China in December 2019 [ 2 ]. Since then, more than 2.74 million lives have been lost as a direct consequence of the disease. Notable outbreaks were recorded in the United States of America, Brazil and India -- which remains a crucial battleground against the outbreak. The Indian government imposed very strict lockdown measures early in the course of the pandemic in order to reduce the spread of the virus. Said measures have not been as effective as was intended [ 3 ], with India now reporting the largest number of confirmed cases in Asia, and the third highest number of confirmed cases in the world after the United States and Brazil [ 4 ], with the number of confirmed cases crossing the 10 million mark on December 18, 2020. On March 24, 2020, the Government of India ordered a 21-day nationwide lockdown, later extending it until May 3. This was followed by two-week extensions starting May 3 and 17 with substantial relaxations. From June 1, the government started ‘unlocking’ most regions of the country in five unlock phases. In order to formulate and implement policy geared toward containment and mitigation, it is important to recognize the presence of highly variable contagion patterns across different Indian states [ 5 ]. India saw a decay in the virus curve in September, 2020 with daily number of cases going below 10,000. At the time of revising the paper, the daily incidence curve is sharply rising again, as India faces its second wave. There is a rising interest in studying potential trajectories that the infection can take in India to improve policy decisions.

A spectrum of models for projecting infectious disease spread have become widely popular in wake of the pandemic. Some popular models include the ones developed at the Institute of Health Metrics (IHME) [ 6 ] (University of Washington, Seattle) and at the Imperial College London [ 7 ]. The IHME COVID-19 project initially relied on an extendable nonlinear mixed effects model for fitting parametrized curves to COVID-19 data, before moving to a compartmental model to analyze the pandemic and generate projections. The Imperial College model (henceforth referred to as ICM) works backwards from observed death counts to estimate transmission that occurred several weeks ago, allowing for the time lag between infection and death. A Bayesian mechanistic model is introduced - linking the infection cycle to observed deaths, inferring the total population infected (attack rates) as well as the time-varying reproduction number R ( t ). With the onset of the pandemic, there has been renewed interest in multi-compartment models, which have played a central role in modeling infectious disease dynamics since the twentieth century [ 8 ]. The simplest of compartmental models include the standard SIR [ 9 ] model, which has been extended [ 10 ] to incorporate various types of time-varying quarantine protocols, including government-level macro isolation policies and community-level micro inspection measures. Further extensions include one which adds a spatial component to this temporal model by making use of a cellular automata structure [ 11 ]. Larger compartmental models include those which incorporate different states of transition between susceptible, exposed, infected and removed (SEIR) compartments, which have been used in the early days of the pandemic in the Wuhan province of China [ 12 ]. The SEIR compartmental model has been further extended to the SAPHIRE model [ 13 ], which accounts for the infectiousness of asymptomatic [ 14 ] and pre-symptomatic [ 15 ] individuals in the population (both of which are crucial transmission features of COVID-19), time varying ascertainment rates, transmission rates and population movement.

Researchers and policymakers are relying on these models to plan and implement public health policies at the national and local levels. New models are emerging rapidly. Models often have conflicting messages, and it is hard to distinguish a good model from an unreliable one. Different models operate under different assumptions and provide different deliverables. In light of this, it is important to investigate and compare the findings of various models on a given test dataset. While some work has been done in terms of trying to reconcile results from different models of disease transmission that can be fit to emerging data [ 16 ], more comparisons need to be done to investigate how differences between competing models might lead to differing projections on the same dataset. In the context of India, such head-to-head comparison across models are largely unavailable.

We consider five different models of different genre, starting from the simplest baseline model. The baseline model we investigate relies on curve-fitting methods, with cumulative number of infected cases modeled as an exponential process [ 17 ]. Next, we consider the extended SIR (eSIR) model [ 10 ], which uses a Bayesian hierarchical model to generate projections of proportions of infected and removed people at future time points. The SAPHIRE [ 13 ] model has been demonstrated to reconstruct the full-spectrum dynamics of COVID-19 in Wuhan between January and March 2020 across five periods defined by events and interventions. Using this, we study the evolution of the pandemic in India over nine well-defined lockdown and unlock periods, each with distinct transmission and ascertainment features. Another model, SEIR-fansy [ 18 ] modifies the SEIR model to account for high false negative rate and symptom-based administration of COVID-19 tests. Finally, we study the ICM model, which utilizes a semi-mechanistic Bayesian hierarchical model based on renewal equations that model infections as a latent process and links deaths to infections with the help of survival analysis . Each of the models mentioned above have had appreciable success in being able to satisfactorily analyze and project the trajectory of the pandemic in different countries [ 19 , 20 , 21 ].

In order to fairly compare and contrast the models mentioned above, we study their respective treatment of the different lockdown and unlock periods declared by the Government of India. Additionally, we compare their projections based on reported data, with special emphasis on how the models deal with (if they do, at all) under-reporting and under-detection of COVID-cases, which has been a major point of discussion in the scientific community, particularly for India [ 22 ]. We also compare the uncertainty associated with the projections across the models which is often overlooked in the literature.

The rest of the paper is organized as follows. In Section 2 we provide an overview of the various models considered in our analysis. The supplement has detailed discussion on the formulation, assumptions and estimation methods utilized by each of the models. We present the numerical findings of our comparative investigation of the models in Section 3 by comparing projected COVID-counts (i.e., case and death counts associated with COVID-19) and (wherever possible) parameter estimates which help understand transmission dynamics of the pandemic. Next, in Section 4 we discuss sensitivity analyses and note applications of the models studied in the context of data from countries other than India. Finally, we discuss the implications of our findings in Section 5 .

Overview of models

In this section, we discuss the assumptions and formulation of each of the five classes of models described above. Table 1 provides an overview of the models compared in this article.

Baseline model

The baseline model we investigate aims to predict the evolution of the COVID-19 pandemic by means of a regression-based predictive model [ 17 ]. More specifically, the model relies on a regression analysis of the daily cumulative count of infected cases based on the least-squares fitting. In particular, the growth rate of the infection is modeled as an exponentially decaying process. Figure  1 provides a schematic overview of this model.

figure 1

Schematic overview of the baseline model

Formulation

The baseline model assumes that the following simple differential equation governs the evolution of a disease in a fixed population:

where I ( t ) is defined as the number of infected people at time t and λ is the growth rate of infection. Unlike the other models described in subsequent sections, the baseline model analyses and projects only the cumulative number of infections, and not counts/proportions associated with other compartments like deaths and recoveries. The model uses reported field data of the infections in India over a specific time period. The growth rate can be numerically approximated from Eq. ( 1 ) above as

Having estimated the growth rate, the model uses a least-squares method to fit an exponential time-varying curve to \( \hat{\uplambda_t} \) , obtained from Eq. ( 2 ) above. Since all the other methods involve Bayesian estimation methods and use posterior distributions to obtain estimates and associated credible intervals, we place a non-informative prior on the random error in the above curve fitting method [ 27 ] to ensure comparable results. Specifically, we consider a uniform prior for the log of error variance. Using projected values of \( \hat{\lambda_t}, \) we extrapolate the number of infections which will occur in future. The baseline model described above has been implemented in R [ 28 ] using standard packages for exponential curve fitting.

Extended SIR (eSIR) model

We use an extension of the standard susceptible-infected-removed (SIR) compartmental model known as the extended SIR (eSIR) model [ 10 ]. To implement the eSIR model, a Bayesian hierarchical framework is used to model time series data on the proportion of individuals in the infected and removed compartments. Markov chain Monte Carlo (MCMC) methods are used to implement this model, which provides not only posterior estimation of parameters and prevalence values associated with all three compartments of the SIR model, but also predicted proportions of the infected and the removed people at future time points. Figure  2 is a diagrammatic representation of the eSIR model.

figure 2

The eSIR model with a latent SIR model on the unobserved proportions. Reproduced from Wang et al., 2020 [ 10 ]

The eSIR model assumes the true underlying probabilities of the three compartments follow a latent Markov transition process and require observed daily proportions of infected and removed cases as input.

The observed proportions of infected and removed cases on day t are denoted by \( {Y}_t^I \) and \( {Y}_t^R \) , respectively. Further, we denote the true underlying probabilities of the S, I, and R compartments on day t by \( {\theta}_t^S \) , \( {\theta}_t^I \) , and \( {\theta}_t^R \) , respectively, and assume that for any t , \( {\theta}_t^S+{\theta}_t^I+{\theta}_t^R=1 \) . Assuming a usual SIR model on the true proportions we have the following set of differential equations

where β  > 0 denotes the disease transmission rate, and γ  > 0 denotes the removal rate. The basic reproduction number R 0   ≔   β / γ indicates the expected number of cases generated by one infected case in the absence of any intervention and assuming that the whole population is susceptible. We assume a Beta-Dirichlet state space model for the observed infected and removed proportions, which are conditionally independently distributed as

Further, the Markov process associated with the latent proportions is built as:

where θ t denotes the vector of the underlying population probabilities of the three compartments, whose mean is modeled as an unknown function of the probability vector from the previous time point, along with the transition parameters. \( \boldsymbol{\tau} =\left(\beta, \gamma, {\boldsymbol{\theta}}_{\mathbf{0}}^T,\boldsymbol{\lambda}, \kappa \right) \) denotes the whole set of parameters where λ I , λ R and κ are parameters controlling variability of the observation and latent process, respectively. The function f (·) is then solved as the mean transition probability determined by the SIR dynamic system, using a fourth order Runge-Kutta approximation [ 29 ].

Priors and MCMC algorithm

The prior on the initial vector of latent probabilities is set as \( {\boldsymbol{\theta}}_{\mathbf{0}}\sim \mathrm{Dirichlet}\left(1-{Y}_1^I-{Y}_1^R,{Y}_1^I,{Y}_1^R\right) \) , \( {\theta}_0^S=1-{\theta}_0^I-{\theta}_0^R \) . The prior distribution of the basic reproduction number is lognormal such that E ( R 0 ) = 3.28 [ 30 ] (this value was also confirmed by calculating the average time-varying R(t) by from January 30 till March 24, 2020, using the package developed by [ 31 ]). The prior distribution of the removal rate is also lognormal such that E ( γ ) = 0.5436. We use the proportion of death within the removed compartment as 0.0184 so that the initial infection fatality ratio is 0.01 [ 32 ]. For the variability parameters, the default choice is to set large variances in both observed and latent processes, which may be adjusted over the course of epidemic with more data becoming available: \( \kappa, {\lambda}^I,{\lambda}^R\ \overset{iid}{\sim }\ \mathrm{Gamma}\left(2,{10}^{-4}\right). \)

Denoting t 0 as the last date of data availability, and assuming that the forecast spans over the period [ t 0  + 1,  T ], the eSIR algorithm is as follows.

Step 0. Take M draws from the posterior \( \left[{\boldsymbol{\theta}}_{\mathbf{1}:{\boldsymbol{t}}_{\mathbf{0}}},\boldsymbol{\tau} |{\boldsymbol{Y}}_{\mathbf{1}:{\boldsymbol{t}}_{\mathbf{0}}}\right] \) .

Step 1. For each solution path m   ∈  {1, …,  M }, iterate between the following two steps via MCMC.

Draw \( {\boldsymbol{\theta}}_{\boldsymbol{t}}^{\left(\boldsymbol{m}\right)} \) from \( \left[\left.{\boldsymbol{\theta}}_{\boldsymbol{t}}\right|{\boldsymbol{\theta}}_{t-1}^{\left(m-1\right)},{\boldsymbol{\tau}}^{(m)}\right],t\in \left\{{t}_0+1,\dots, T\right\} \) .

Draw \( {\boldsymbol{Y}}_{\boldsymbol{t}}^{\left(\boldsymbol{m}\right)} \) from \( \left[\left.{\boldsymbol{Y}}_{\boldsymbol{t}}\right|{\boldsymbol{\theta}}_t^{(m)},{\boldsymbol{\tau}}^{(m)}\right],t\in \left\{{t}_0+1,\dots, T\right\} \) .

Implementation

We implement the proposed algorithm in R package rjags [ 33 ] and the differential equations were solved via the fourth-order Runge–Kutta approximation. To ensure the quality of the MCMC procedure, we fix the adaptation number (which denotes the number of MCMC samples discarded by JAGS in order to tune parameters which in turn improves speed or de-correlation of sampling) at 10 4 , thin the chain by keeping one draw from every 10 random draws to further reduce autocorrelation, set a burn-in period of 10 5 draws under 2 × 10 5 iterations for four parallel chains. This implementation provides not only posterior estimation of parameters and prevalence of all the three compartments in the SIR model, but also predicts proportions of the infected and the removed people at future time point(s). The R package for implementing this general model for understanding disease dynamics is publicly available at https://github.com/lilywang1988/eSIR .

SAPHIRE model

This model [ 13 ] extends the classic SEIR model to estimate COVID-related transmission parameters, in addition to projecting COVID-19 case counts, while accounting for pre-symptomatic infectiousness, time-varying ascertainment rates (i.e. reporting rates), transmission rates and population movements. Figure  3 provides a schematic diagram of the compartments and transitions conceptualized in this model. The model includes seven compartments: susceptible (S), exposed (E), pre-symptomatic infectious (P), reported infectious (I), unreported infectious (A), isolation in hospital (H) and removed (R). Compared with the classic SEIR model, SAPHIRE explicitly models population movement and introduce two additional compartments (A and H) to account for the fact that only reported cases would seek medical care and thus be quarantined by hospitalization. The model described and implemented here relies on the same methodology and arguments as presented by [ 13 ]. The only difference is that while the original model analyzed data from China over a time period of December 2019 to March 2020 (which constituted the initial days of the pandemic in China), we analyze data from India. Additionally, the original manuscript adjusted the model to account for population movement. Data on population movement not being available consistently over time and regions in India, we make no such modifications. We further note that the SAPHIRE model returns reported and unreported cumulative COVID-case counts, in addition to cumulative counts of the removed compartment. As such, for the purpose of comparisons, the SAPHIRE model is used only to study cumulative COVID-case counts (reported and unreported). The R package for implementing this general model for understanding disease dynamics is publicly available at https://github.com/chaolongwang/SAPHIRE .

figure 3

The SAPHIRE model includes seven compartments: susceptible (S), exposed (E), pre-symptomatic infectious (P), reported infectious (I), unreported infectious (A), isolation in hospital (H) and removed (R)

The dynamics of the 7 compartments described above at time t are described by the set of ordinary differential equations

in which b is the transmission rate for reported cases (defined as the number of individuals that an reported case can infect per day), α is the ratio of the transmission rate of unreported cases to that of reported cases, r is the ascertainment rate, D e is the latent period, D p is the pre-symptomatic infectious period, D i is the symptomatic infectiousness period, D q is the duration from illness onset to isolation and D h is the isolation period in the hospital. Further, we set N  = 1.34 × 10 9 as the population size for India and set n  = 0 to indicate no incoming or outgoing travelers.

Under this setup, the reproductive number R (as presented in the original manuscript) may be expressed as

in which the three terms represent infections contributed by pre-symptomatic individuals, unreported cases and reported cases, respectively. The model adjusts the infectious periods of each type of case by taking isolation of patients who test positive \( \left(\mathrm{by}\ \mathrm{means}\ \mathrm{of}\ {D}_q^{-1}\right) \) into account.

Initial states and parameter settings

We set α = 0.55, assuming lower transmissibility for unreported cases [ 34 ]. Compartment P contains both reported and unreported cases in the pre-symptomatic phase. We set the transmissibility of P to be the same as unreported cases, because it has previously been reported that the majority of cases are unreported [ 34 ]. We assume an incubation period of 5.2 days and a pre-symptomatic infectious period D p  = 2.3 days [ 35 , 36 ]. The latent period was D e  = 2.9 days. Since pre-symptomatic infectiousness was estimated to account for 44% of the total infections from reported cases [ 35 ], we set the mean of total infectious period as ( D p  +  D i ) =  D p /0.44 = 5.2 days, assuming constant infectiousness across the pre-symptomatic and symptomatic phases of reported cases [ 37 ] – thus the mean symptomatic infectious period was D i  = 2.9 days. We set a long isolation period of D h  = 17 days, based on a study investigating hospitalisation of COVID-19 patients in the state of Karnataka [ 38 ]. The duration from the onset of symptoms to isolation was estimated to be D q  = 7 [ 23 , 39 ] as the median time length from onset to confirmed diagnosis. On the basis of the parameter settings above, the initial state of the model is specified on March 15. The initial number of reported symptomatic cases I (0) is specified as the number of reported cases who experienced symptom onset during 12–14 March. The initial ascertainment rate is assumed to be r 0  = 0.10 [ 40 ], and thus the initial number of unreported cases is \( A(0)={r}_0^{-1}\left(1-{r}_0\right)I(0) \) . P 1 (0) and E 1 (0) denote the numbers of reported cases in which individuals experienced symptom onset during 15–16 March and 17–19 March, respectively. Then, the initial numbers of exposed and pre-symptomatic individuals are set as \( E(0)={r}_0^{-1}{E}_1(0) \) and \( P(0)={r}_0^{-1}{P}_1(0) \) , respectively. The initial number of the hospitalized cases H (0) is set as half of the cumulative reported cases on 8 March since D q  = 7 and there would be more severe cases among the reported cases in the early phase of the epidemic.

Likelihood and MCMC algorithm

Considering the time-varying strength of control measures implemented in India over the trajectory of the pandemic, we chose to break the training period into ten sequential blocks: pre-lockdown (March 15–24), lockdown phases 1, 2, 3, and 4 (March 25 – April 14, April 15 – May 3, May 4–17, and May 18–31 respectively) followed by unlock phases 1, 2, 3, 4 and 5 (June 1–30, July 1–31, August 1–31, September 1–30 and October 1–15 respectively). In other words, the model assumes that the value of b  (and  r ) corresponding to the i th lockdown period to vary as b i (and  r i ) for i  = 1, 2, 3, …, 10. The observed number of reported cases in which individuals experience symptom onset on day t – denoted by x t – is assumed to follow a Poisson distribution with rate \( {\uplambda}_t=r{P}_{t-1}{D}_p^{-1} \) , with P t denoting the expected number of pre-symptomatic individuals on day t . The following likelihood equation is used to fit the model using observed data from March 15 ( T 0 ) to October 15 ( T 1 ).

and the model is used to predict COVID-counts from October 16 to December 31. A non-informative prior of U (0, 2) is used for b 1 , b 2 , …, b 10 . For r 1 , an informative prior of Beta(10, 90) is used based on the findings of [ 40 ]. We reparameterise r 2 , …, r 10 as

where logit( t ) = log( t /(1 −  t )) is the standard logit function. In the MCMC, δ i   ∼   N (0, 1) for i  = 2, 3, …, 10. A burn-in period of 100,000 iterations is fixed, with a total of 200,000 iterations being run.

SEIR-fansy model

One of the problems with applying a standard SIR model in the context of the COVID-19 pandemic is the presence of a long incubation period. As a result, extensions of SIR model like the SEIR model are more applicable. In the previous subsection, we have seen an extension which includes the ‘pre-symptomatic infectious’ compartment (people who are infected at time t and contributing to the spread of the virus, but do not show any symptom yet). In the SEIR-fansy model, we use an alternate formulation by defining an ‘untested infectious’ compartment for infected people who are spreading infection but are not tested after the incubation period. This compartment is necessary because there is a large proportion of infected people who are not being tested (a part of them are asymptomatic or mildly symptomatic but for a country like India there are other reasons like access to care and stigma that can prevent someone from getting tested/diagnosed). We have assumed that after the ‘exposed’ compartment, a person enters either the ‘untested infectious’ compartment or the ‘tested infectious’ compartment. To incorporate the possible effect of misclassifications due to imperfect testing, we include a compartment for false negatives (infected people who are tested but reported as negative). As a result, after being tested, an infected person enters either into the ‘false negative’ compartment or the ‘tested positive’ compartment (infected people who are tested and reported to be positive). We keep separate compartments for the recovered and deceased persons coming from the untested and false negatives compartments which are ‘recovered unreported’ and ‘deceased unreported’ respectively. For the ‘tested positive’ compartment, the recovered and the death compartments are denoted by ‘recovered reported’ and ‘deceased reported’ respectively. Thus, we divide the entire population into ten main compartments: S (Susceptible), E (Exposed), T (Tested), U (Untested), P (Tested positive), F (Tested False Negative), RR (Reported Recovered), RU (Unreported Recovered), DR (Reported Deaths) and DU (Unreported Deaths). This model is implemented using the R package SEIRfansy [ 26 ].

Like most compartmental models, this model assumes exponential times for the duration of an individual staying in a compartment. For simplicity, we approximate this continuous-time process by a discrete-time modeling process. The main parameters of this model are β (rate of transmission of infection by false negative individuals), α p (scaling factor that measures the rate of spread of infection by patients who test positive for COVID-19 relative to infected patients who return false negative test results), α u  (scaling factor for the rate of spread of infection by untested individuals), D e (incubation period in days), D r (mean days till recovery for positive individuals), D t (mean number of days for the test result to come after a person is being tested), μ c (death rate due to COVID-19 which is the inverse of the average number of days for death due to COVID-19 starting from the onset of disease multiplied by the probability death of an infected individual due to COVID), λ and μ (natural birth and death rates respectively, assumed to be equal for the sake of simplicity), r (probability of being tested for infectious individuals), f (false negative probability of RT-PCR test), \( {\beta}_1\ and\ {\beta}_2^{-1} \) (scaling factors for rate of recovery for undetected and false negative individuals respectively), \( {\delta}_1\mathrm{and}\ {\delta}_2^{-1} \) (scaling factors for death rate for undetected and false negative individuals respectively). The number of individuals at the time point t in each compartment is governed by the system of differential equations given by Eqs. ( 8a ) – ( 8i ). To simplify this model, we assume that testing is instantaneous. In other words, we assume there is no time difference from the onset of the disease after the incubation period to getting test results. This is a reasonable assumption to make as the time for testing is about 1–2 days which is much less than the mean duration of stay for the other compartments. Further, once a person shows symptoms for COVID-19 like diseases, they are sent to get tested almost immediately. Figure  4 provides a schematic overview of the model.

figure 4

Schematic diagram for the SEIR-fansy model with imperfect testing and misclassification. The model has ten compartments: S (Susceptible), E (Exposed), T (Tested), U (Untested), P (Tested positive), F (Tested False Negative), RR (Reported Recovered), RU (Unreported Recovered), DR (Reported Deaths) and DU (Unreported Deaths). Reproduced from Bhaduri, Kundu et al., 2020 [ 18 ]

The following differential equations summarize the transmission dynamics being modeled.

Using the Next Generation Matrix Method [ 41 ], we calculate the basic reproduction number

where S 0  = λ/μ = 1 since we assume that natural birth and death rates are equal within this short period of time. Supplementary Table S 1 describes the parameters in greater detail.

Likelihood assumptions and estimation

Parameters are estimated using Bayesian estimation techniques and MCMC methods (namely, Metropolis-Hastings method [ 42 ] with Gaussian proposal distribution). First, we approximated the above set of differential equations by a discrete time approximation using daily differences. After we start with an initial value for each of the compartments on the day 1, using the discrete time recurrence relations we obtain the counts for each of the compartments at the next days. To proceed with the MCMC-based estimation, we specify the likelihood explicitly. We assume (conditional on the parameters) the number of new confirmed cases on day t depend only on the number of exposed individuals on the previous day. Specifically, we use multinomial modeling to incorporate the data on recovered and deceased cases as well. The joint conditional distribution is

A multinomial distribution-like structure is then defined

Note: the expected values of E ( t  − 1) and P ( t  − 1) are obtained by solving the discrete time differential equations specified by Eqs. ( 8a ) – ( 8i ).

Prior assumptions and MCMC

For the parameter r , we assume a U (0, 1) prior, while for β , we assume an improper non-informative flat prior with the set of positive real numbers as support. After specifying the likelihood and the prior distributions of the parameters, we draw samples from the posterior distribution of the parameters using the Metropolis-Hastings algorithm with a Gaussian proposal distribution. We run the algorithm for 200,000 iterations with a burn-in period of 100,000. Finally, the mean of the parameters in each of the iterations are obtained as the final estimates of β and r for the different time periods. As in the case of the SAPHIRE model, we again break the training period into ten sequential blocks: pre-lockdown (March 15–24), lockdown phases 1, 2, 3, and 4 (March 25 – April 14, April 15 – May 3, May 4–17, and May 18–31 respectively) followed by unlock phases 1, 2, 3, 4 and 5 (June 1–30, July 1–31, August 1–31, September 1–30 and October 1–15 respectively).

Imperial College London model (ICM)

We examine a Bayesian semi-mechanistic model for estimating the transmission intensity of SARS-CoV-2 [ 7 ]. The model defines a renewal equation using the time-varying reproduction number R t to generate new infections. As a lot of cases in SARS-CoV-2 are asymptomatic and reported case data is unreliable especially in early part of the epidemic in India, the model relies on observed deaths data and calculates backwards to infer the true number of infections. The latent daily infections are modeled as the product of R t with a discrete convolution of the previous infections, weighted using an infection-to-transmission distribution specific to SARS-CoV-2. We implement this Bayesian semi-mechanistic model in the context of COVID-19 data arising from India in order to estimate the reproduction number over time, along with plausible upper and lower bounds (95% Bayesian credible intervals (CrI)) of the daily infections and the daily number of infectious people. We parametrize R t with a fixed effect and a random effect for each week over the course of the epidemic for each state. The fixed effect accounts for the variations in R t across India as a whole whereas the random effect allows for variations among different states. The weekly effects are encoded as a random walk, where at each successive step the random effect has an equal chance of moving upwards or downwards from its current value. The model is implemented using epidemia [ 43 ], a general purpose R package for semi-mechanistic Bayesian modelling of epidemics. Figure  5 represents a schematic overview of the model.

figure 5

Schematic overview of ICM

The true number of infected individuals, i , is modelled using a discrete renewal process. We specify a generation distribution [ 44 ] g with density g (τ) as g   ∼  Gamma(6.5,0.62). Given the generation distribution, the number of infections i t , m on a given day  t , and state m is given by the discrete. Convolution function:

where the generation distribution is discretized by \( {g}_s={\int}_{s-0.5}^{s+0.5}g\left(\uptau \right)d \) for s  = 2, 3, …,and \( {g}_1={\int}_0^{1.5}g\left(\uptau \right)d\uptau \) . The population of state m is denoted by N m . We include the adjustment factor S t , m to account for the number of susceptible individuals left in the population.

We define daily deaths, D t , m , for days t   ∈  {1, …,  n } and states m   ∈  {1, …,  M }. These daily deaths are modelled using a positive real-valued function d t , m  =  E [ D t , m ] that represents the expected number of deaths attributed to COVID-19. The daily deaths D t , m  are assumed to follow a negative binomial distribution with mean d t , m and variance \( {d}_{t,m}+{d}_{t,m}^2/{\uppsi}_1 \) , where ψ 1  follows a positive half normal distribution, i.e.,

We link our observed deaths mechanistically to transmission [ 7 ]. We use a previously estimated COVID-19 infection fatality ratio (IFR, probability of death given infection) of 0.1% [ 45 , 46 ] together with a distribution of times from infection to death π. To incorporate the uncertainty inherent in this estimate we modify the ifr for every state to have additional noise around the mean, denoted by \( \mathrm{if}{\mathrm{r}}_{\mathrm{m}}^{\ast } \) . Specifically, we assume.

where \( \mathrm{if}{\mathrm{r}}_{\mathrm{m}}^{\ast } \) represents the noise-added analog of ifr. Using estimated epidemiological information from previous studies, we assume the distribution of times from infection to death π (infection-to-death) to be the convolution of an infection-to-onset distribution (π ′ ) [ 47 ] and an onset-to-death distribution [ 32 ].

The expected number of deaths d t , m , on a given day t , for state m is given by the following discrete sum

where i τ , m is the number of new infections on day τ in state m and where, similar to the generation distribution, π is discretized via \( {\uppi}_s={\int}_{s-0.5}^{s+0.5}\uppi \left(\uptau \right)d\uptau \) for s  = 2, 3, …, and \( {\uppi}_1={\int}_0^{1.5}\uppi \left(\uptau \right)\mathrm{d}\uptau \) , where π(τ) is the density of π.

We parametrize R t , m with a random effect for each week of the epidemic as follows

where f ( x ) = 2  exp  ( x )/(1 +   exp  ( x )) is twice the inverse logit function, and ϵ w ( t ) and \( {\epsilon}_{m,w\left(t,m\right)}^{state} \) follow a weekly random walk process, that captures variation between R t , m in each subsequent week. ϵ w ( t ) is a fixed effect estimated across all the states and \( {\epsilon}_{m,w\left(t,m\right)}^{state} \) is the random effect specific to each state in India. The prior distribution for R 0 [ 30 ] was chosen to be

We assume that seeding of new infections begins 30 days before the day after a state has cumulatively observed 10 deaths. From this date, we seed our model with 6 sequential days of an equal number of infections: i 1  = … =  i 6   ∼  Exponential(τ −1 ), where τ  ∼  Exponential(0.03). These seed infections are inferred in our Bayesian posterior distribution. Fitting was done with the R package epidemia [ 43 ] which uses STAN [ 48 ], a probabilistic programming language, using an adaptive Hamiltonian Monte Carlo (HMC) sampler.

Comparing models and evaluating performance

Having established differences in the formulation of the different models, we compare their respective projections and inferences. In order to do so, we use the same data sources [ 49 , 50 ] for all five models. Well-defined time points are used to denote training (March 15 to October 15) and test (October 16 to December 31) periods.

Using the parameter values specified above along with data from the training period as inputs, we compare the projections of the five models with observed data from the test period. In order to do so, we use the symmetric mean absolute prediction error (SMAPE) and mean squared relative prediction error (MSRPE) metrics as measures of accuracy. Given observed time-varying data \( {\left\{{O}_t\right\}}_{t=1}^T \) and an analogous time-series dataset of projections \( {\left\{{P}_t\right\}}_{t=1}^T \) , the SMAPE metric is defined as

where | x | denotes the absolute value of x . The metric MSRPE is defined as

It can be seen that 0 ≤  SMAPE  ≤ 100, with smaller values of both MSRPE and SMAPE indicating a more accurate fit. For active reported cases (cases that are active on a given day which is the difference of cumulative reported cases and cumulative reported counts of recoveries and deaths), we compute and compare the metrics defined above for projections from eSIR and SEIR-fansy models as no other model returns relevant projections. For cumulative reported cases we obtain projections from all models apart from ICM (which yields total, i.e., sum of reported and unreported, cumulative cases). For cumulative reported deaths we compare projections from eSIR, SEIR-fansy and ICM, since the baseline and SAPHIRE models do not yield relevant projections. Supplementary Table S 2 gives an overview of output from each of the models we consider and Table  2 reports the values of accuracy metrics described above.

Further, we compare (when possible) the estimated time-varying reproduction number R ( t ) over the different lockdown and unlock stages in India. Specifically, for each lockdown stage, we report the median R ( t ) value along with the associated 95% credible interval (CrI). The values are presented in Table 2 .

Since we are interested in comparing relative performances of the models (specifically, their projections), we define another metric – the relative mean squared prediction error (Rel-MSPE). Given time series data on observed cumulative cases (or deaths) \( {\left\{{O}_t\right\}}_{t=1}^T \) , projections from a model A \( {\left\{{P}_t^A\right\}}_{t=1}^T \) , and projections from some other model B, \( {\left\{{P}_t^B\right\}}_{t=1}^T \) , the Rel-MSPE of model B with respect to model A is defined as

Higher values of Rel-MSPE(B:A) indicate better performance of model B over model A. Since the baseline model yields projections of cumulative reported cases, we compute Rel-MSPE for the other models with respect to the baseline model for reported cumulative cases. Projections from ICM represent total (i.e., sum of reported and unreported) cumulative cases and are left out of this comparison of reported counts. For cumulative reported deaths, we compute Rel-MSPE of the SEIR-fansy and ICM models relative to the eSIR model. In addition to comparing the accuracy of fits that arise from the different models, we also investigate if projections from the different models are correlated with observed data. We use the standard Pearson’s correlation coefficient and Lin’s concordance correlation coefficient [ 51 ] as summary measures to study said correlation. Higher values of these correlation metrics indicate better concordance of model projections and the observed data from the test period. Rel-MSPE and correlation metrics are presented in Table  3 . Since we have projections for total (sum of reported and unreported cases) for active cases from SEIR-fansy, for cumulative cases from SAPHIRE, SEIR-fansy and ICM, and for cumulative deaths from SEIR-fansy, we present the projected totals along with 95% credible intervals and associated underreporting factors on three specific dates – October 31, November 30 and December 31 in Table  4 . The table also includes projected cumulative reported counts (which are available from all models under investigation apart from ICM) with 95% credible intervals for the three dates mentioned above.

Data source

The data on confirmed cases, recovered cases and deaths for India and the 20 states of interest are taken from COVID-19 India [ 49 ] and the JHU CSSE COVID-19 GitHub repository [ 50 ]. In addition to this and other similar articles concerning the spread of this disease in India, we have created an interactive dashboard [ 52 ] summarizing COVID-19 data and forecasts for India and its states (generated with the eSIR model discussed in this paper). While the models are trained using data from March 15 to October 15, 2020, their performances are compared by examining their respective projections from October 16 to December 31, 2020.

Estimation of the reproduction number

From Table 2 , we compare the mean of the time-varying effective reproduction number R ( t ) over the four phases of lockdown and subsequent unlock phased in India. The eSIR model returns a mean value of 2.08 (95% credible interval: 1.41–2.12) over the entire training period. Factoring in different levels of government interventions which modified transmission dynamics during lockdown, we get period specific estimates ranging from 2.12 (1.44–2.16) in lockdown phase 1, which drops to 1.48 (1.00–1.51) in lockdown phase 2 and then reports a steady decline over the subsequent lockdown and unlock phases. The mean values returned by the SAPHIRE model varied from 2.54 (2.41–2.74) during phase 1 of the lockdown, 1.60 (1.36–2.17) for phase 2, 1.69 (1.46–1.97) for phase 3 and 1.54 (1.29–2.00) for the fourth and final lockdown phase. The estimated values for subsequent unlock phases are quite close to each other, starting from 1.27 (1.19–1.32) in unlock phase 1 and dropping to 1.09 (0.91–1.69) in the fifth unlock phase. The SEIR-fansy notes that the mean R ( t ) drops from 5.03 (5.01–5.04) during the first phase of lockdown, to 1.90 (1.89–1.91) during the second lockdown phase, before rising again to 2.33 (2.30–2.36) during lockdown phase 4. The estimated mean drops steadily from 1.80 (1.79–1.81) during unlock phase 2 to 0.86 (0.85–0.87) during unlock phase 5. The ICM-based mean values fluctuate, from 1.77 (1.58–1.96) during the first lockdown phase, followed by 1.22 (1.18–1.27), then dropping to 1.33 (1.28–1.38) and finally rising to 1.41 again (1.35–1.47) for the fourth phase of lockdown. Estimates from ICM during unlock phases behave like those from the SEIR-fansy model – in unlock phase 2 the estimated mean is 1.11 (1.08–1.14) and in unlock phase 5, the mean is 0.83 (0.82–0.84). In terms of agreement of reported values, SAPHIRE, SEIR-fansy and ICM report the highest mean R for phase one of the lockdown. Values reported by SAPHIRE, SEIR-fansy and ICM report a drop in intermediate lockdown phases, followed by a rise. Values during unlock period increase from phase 1 to phase 2, followed by a steady decline. SAPHIRE, SEIR-fansy and ICM report the lowest value of R for unlock phase 5.

Estimation of reported case counts

From Figs.  6 , 7 , 8 and 9 , we note that the eSIR model overestimates the count of active cases – a behavior which gets worse with time. While the observed counts decrease steadily in the test period, the eSIR model fails to capture this behaviour and returns projections which rise over time. In comparison, the SEIR-fansy model is able to replicate the decreasing behaviour but yields projections which are higher than observed counts. In terms of prediction accuracy, the SEIR-fansy model has an SMAPE value of 35.14% and an MSRPE value of 1.11. For eSIR model, those values are at 37.96% (SMAPE) and 2.28 (MSRPE).

figure 6

Comparison of projected and observed reported active cases from October 16 to December 31 for India, using training data from March 15 to October 15, 2020

figure 7

Comparison of projected and observed reported cumulative cases from October 16 to December 31 for India, using training data from March 15 to October 15, 2020

figure 8

Comparison of projected and observed reported cumulative deaths from October 16 to December 31 for India, using training data from March 15 to October 15, 2020

figure 9

Scatter plot and marginal densities of projected and observed reported active cases from October 16 to December 31 for India, using training data from March 15 to October 15, 2020

From Figs.  7 , 8 , 9 and 10 we note that while the SAPHIRE model underestimates the count of cumulative cases, the baseline, eSIR and SEIR-fansy models overestimate the count. Table 2 reveals that SAPHIRE performs the best in terms of SMAPE metric with a value of 2.25%, followed closely by SEIR-fansy (2.29%). The eSIR and baseline models perform poorly in comparison, yielding 6.59 and 6.89% respectively. The SEIR-fansy model performs best in terms of MSRPE with a value of 0.05, followed closely by SAPHIRE (0.06) . Table 3 further reveals a similar relative performance through Rel-MSPE values (all Rel-MSPE figures reported here are relative to the baseline model). The SEIR-fansy model performs the best with Rel-MSPE value of 3.27, followed by SAPHIRE (3.01), and finally, the eSIR model (1.72). All four sets of projections are highly correlated with the observed time series – with all model projections having a Pearson’s correlation coefficient of nearly 1 with the observed data. Lin’s concordance coefficient yields an ordering (from worst to best) of the eSIR model (0.48), followed by the baseline model (0.51), the SAPHIRE model (0.74) and finally, the SEIR-fansy model (0.89).

figure 10

Scatter plot and marginal densities of projected and observed cumulative cases from October 16 to December 31 for India, using training data from March 15 to October 15, 2020

Estimation of reported death counts

From Figs.  8 , 9 , 10 and 11 , we note that the eSIR and SEIR-fansy models almost always overestimate, whereas the ICM model slightly underestimates the confirmed cumulative death counts. From Table 2 and Table 3 , the SMAPE and MSRPE values, along with comparison of projections with observed data reveal that the ICM model is most accurate (SMAPE: 0.77%, MSRPE: 0.020), followed by SEIR-fansy (SMAPE: 4.74%, MSRPE: 0.12) followed by the eSIR model (SMAPE: 8.94%, MSRPE: 0.25). Relative to the eSIR model, the Rel-MSPE values of the models reveal that the SEIR-fansy model performs better (Rel-MSPE: 6.96), followed by ICM (Rel-MSPE: 3.64). Judging by values of Pearson’s correlation coefficient, all three sets of projections are highly correlated with the observed data. Lin’s concordance coefficient yields an ordering (from best to worst) of ICM (0.96), followed by SEIR-fansy (0.62) and finally eSIR (0.34).

figure 11

Scatter plot and marginal densities of projected and observed cumulative death from October 16 to December 31 for India, using training data from March 15 to October 15, 2020

Estimation of unreported case and death counts

From Table 4 , we note that the SEIR-fansy model yields underreporting factors of about 10 for active cases on October 31, November 30 and December 31. Further, we observe that the SAPHIRE model projects the maximum count of total cumulative cases on the above three dates, followed by the SEIR-fansy and then ICM. SAPHIRE returns under-reporting factors of the order of approximately 65, while SEIR-fansy and ICM return under-reporting factors which are approximately 7 and 4 respectively. For cumulative deaths, SEIR-fansy estimates underreporting factors approximately equal 3.

Uncertainty quantification of estimates and predictions

From Fig.  12 we observe that the width of 95% credible intervals associated with projections from each of the models vary significantly. While the eSIR model consistently returns the widest intervals, SEIR-fansy has the narrowest intervals. In case of cumulative counts, the ordering (best to worst) starts with SEIR-fansy, followed by the baseline, followed by SAPHIRE and finally the eSIR model. For cumulative deaths, the ordering (best to worst) starts with SEIR-fansy, followed by ICM and finally eSIR. From Table 4 , we compare projections of reported cumulative cases for each model (apart from ICM which returns projections of cumulative total cases and not cumulative reported cases) and their associated prediction intervals on October 31, November 30 and December 31, 2020. On October 31, we observe 8.18 million cumulative reported cases, while the projections (in millions) from the baseline model are 8.71 (95% credible interval: 8.63–8.80), while eSIR yields 8.35 (7.19–9.60), SAPHIRE returns 8.17 (7.90–8.52) and SEIR-fansy projects 8.51 (8.18–8.85) million cases. We do not present our projections for November 30 and December 31, 2020 here in the interest of conciseness.

figure 12

Boxplots showing width of 95% credible interval associated with projected active cases, cumulative cases and cumulative deaths from October 16 to December 31 for India, using training data from March 15 to October 15, 2020

Sensitivity analyses and performance in other countries

Sensitivity analyses for some of the discussed models have been carried out in several other publications. In the interest of conciseness, we refer to said publications and comment on what parameters are central to estimation and generating projections for the models examined here. We also include information on how these models have performed in the context of data from other countries.

Evaluation of the model results in terms of their sensitivity to initial parameter choices and under-reporting and clustering issues within the data have been discussed in the context of India in prior literature [ 53 ]. The range of scenarios considered earlier include 10-fold underreporting of cases, clustering of cases in metropolitan areas, and prior mean of R 0 ranging from 2 to 4 (See Supplementary Table S 3 ). Even though the posterior estimates and predictions changed in scale to some extent across these scenarios, they did not significantly change the broad conclusions. It is undeniable that the exact predicted case counts are sensitive to the choice of priors, but with new data coming in over a longer time frame, as seen in the results from this work, the model is capable of washing out the prior effects in the posterior outcomes.

The eSIR model has been successfully implemented and utilized in the context of COVID-19 across different geographical locations, including China [ 24 , 25 , 54 ], Poland [ 55 ], Italy [ 24 ], Bangladesh and Pakistan [ 56 ]. These countries cover a broad range in terms of socio-economic status, health infrastructure and pandemic management strategies. In each of these cases the eSIR model was seen to be successfully capturing the patterns of growth of the pandemic via estimated parameters, as well as efficiently forecasting future case counts via predictive modeling.

We conducted the sensitivity analysis (results not shown) by changing the initial parameters as 20% lower or higher than the specified values in the SAPHIRE model. The estimated R and ascertainment rates were robust to misspecification of the duration from the onset of symptoms to isolation and of the relative transmissibility of unreported versus reported cases. R estimates were positively correlated with the specified latent and infectious periods, and the estimated ascertainment rates were positively correlated with the specified ascertainment rate in the initial state. This finding is consistent with sensitivity analyses of the SAPHIRE model implemented in Wuhan [ 13 ]. The estimated ascertainment rates were positively correlated with the specified ascertainment rate in the initial state while the under-reported factors were negatively associated with initial ascertainment. The estimated under-reported factor on October 31 (see Table 4 ) decreases dramatically from 117 to 0.07 with the initial ascertainment rate increasing from 0.07 to 0.14, with an initial ascertainment rate of 0.10 providing the best fit, which is presented in this article.

The SAPHIRE model was originally developed in the context of data from China and was successfully able to delineate the transmission dynamics of COVID-19 in Wuhan [ 13 ] and in South Africa [ 57 ].

In the paper, we fix most parameters in our model and examine transmission dynamics only through β and r . It is necessary to design and implement a sensitivity analysis focusing on various combinations of the parameters that were previously fixed. The details of the sensitivity analyses are described in detail in [ 18 ]. The basic findings from the sensitivity analyses are summarized as follows. We observe that the predictions for the reported active cases ( P ) remains same for all parameter choices. The estimates for R 0 mainly differ in the first period, although some variation is noted for the second period as well. However, the estimated R  are almost the same for the later stages of the pandemic in the different models. For the untested cases, in some of the settings of our analysis, there are substantial deviations from the true numbers. The total number of active cases (which include both the unreported and the reported cases) also varies substantially with different parameter values. Consequently, we note how the estimation of unreported cases is sensitive to different choices for the parameter values. In particular, we see different values of E 0 have the most impact on our sensitivity analysis, while different choices of D E have the least impact.

The SEIR-fansy model has not been run for different countries, but it has been implemented for most Indian states separately [ 18 ] which showed that the model was able to capture the transmission dynamics of COVID-19 in most states of India quite efficiently. For instance, this model was able to match the sero-survey results of Delhi quite well [ 45 ]. For other states, the predicted reported cases came out to be quite close to the observed reported cases (with observed cases lying within the credible interval of projections).

The parameters critical to the estimation and projection methods include the infection-to-death distribution [ 32 ], infection fatality ratio [ 45 , 46 ], generation distribution [ 44 ], prior for R 0 [ 7 , 30 ] and seeding [ 7 ]. Researchers have performed sensitivity analysis for various choices of infection-to-death distribution and found the resultant projections to be robust under changes [ 7 ]. We used a range of values for our prior of IFR, with mean 1, 0.4 and 0.1%. We found that the model fits and estimated R t are more or less the same for all three choices but certainly our estimates for total infections changes. This implies the ascertainment of cases (positive results) will be affected. Sensitivity analyses towards the choice of the generation distribution was performed by other researchers [ 7 ] who found the models to be robust against various choices. It has a very minimal effect on the estimation of time varying reproduction number and total infections by the model. We used the R 0 prior suggested in both [ 7 , 30 ]. We did run sensitivity on a few other choices and found that our prior choice affected the inferred R t values for only the first few days and subsequent dynamics are the same irrespective of the choice. Finally, as discussed in [ 7 ] we validated our seeding scheme through an importance sampling leave-one-out cross validation scheme [ 58 , 59 ].

Different versions of ICM model has been applied to 11 European countries in [ 7 ]. On a subregional basis the model is used in the USA [ 60 ], Brazil [ 20 , 61 ] and Italy [ 21 ]. At a local level work the model is used for producing daily estimates for all local and regions in the UK [ 62 , 63 ]. It is also used by Scotland government [ 64 ] and New York State government [ 65 ].

In this comparative paper we have described five different models of various stochastic structures that have been used for modeling SARS-Cov-2 disease transmission in various countries across the world. We applied them to a case-study in modeling the full disease transmission of the coronavirus in India. While simulation studies are the only gold standard way to compare the accuracy of the models, here we were uniquely poised to compare the projected case-counts and death-counts against observed data on a test period. We learned several things from these models. While the estimation of the reproduction number is relatively robust across the models, the prediction of active and cumulative number of cases and cumulative deaths show variation across models. Our findings in terms of estimates of R ( t ) are reflective of the national and state-level implementations of four lockdown phases [ 66 ] which are summarized in Supplementary Table S 4 . The largest variability across models is observed in predicting the “total” number of infections including reported and unreported cases. The degree of underreporting has been a major concern in India and other countries [ 67 ]. We note from Table 4 that the underreporting factor from SAPHIRE is much higher than those reported by SEIR-fansy and ICM. This may be attributed to the fact that SEIR-fansy and ICM both fit daily reported deaths with a pre-specified death rate (which is higher than that for unreported cases), SAPHIRE does not include daily reported death counts in the likelihood function. Additionally, SEIR-fansy also considered the false positive/negative rates of tests and the selection bias in testing, which also contribute to more accurate unreported case projections along with untested infectious case counts. With a comprehensive exposition and a single beta-testing case-study we hope this paper will be useful to understand the mathematical nuance and the differences in terms of deliverables for the models.

There are several limitations to this work. First and foremost, all model estimates are based on a scenario where we assumed no change in either interventions or behavior of people in the forecast period. This is not true as there is tremendous variation in policies across Indian states in the post lockdown phase. We did observe regional lockdowns that were enacted in the forecast period. None of our models tried to capture this variability. Second, the five models we compare are a subset of a vast amount of work that has been done in this area, including models that incorporate age-specific contact network and spatiotemporal variation [ 11 , 68 ]. Third, we have not tested the models for predicting the oscillatory growth and decay behavior of the virus incidence curve, in particular, predicting the second wave. Finally, an extensive simulation study would be the best way to assess the models under different scenarios, but we have restricted our attention to India.

Availability of data and materials

Please visit https://github.com/umich-cphds/cov-ind-19 .

Abbreviations

Imperial College Model

Markov Chain-Monte Carlo

Mean squared relative prediction error

Relative mean squared prediction error

Susceptible-Exposed-Infected-Removed

Susceptible-Infected-Removed

Symmetric mean absolute prediction error

Mayo Clinic. Coronavirus disease 2019 (COVID-19)—Symptoms and causes [Internet]. 2020 [cited 2020 May 21]. Available from: https://www.mayoclinic.org/diseases-conditions/coronavirus/symptoms-causes/syc-20479963

Google Scholar  

Wikipedia. Coronavirus disease 2019. [cited 2020 Aug 3]. Available from: https://en.wikipedia.org/wiki/Coronavirus_disease_2019

Aiyar S. Covid-19 has exposed India’s failure to deliver even the most basic obligations to its people [Internet]: CNN; 2020. [cited 2020 Aug 3]. Available from: https://www.cnn.com/2020/07/18/opinions/india-coronavirus-failures-opinion-intl-hnk/index.html

Kulkarni S. India becomes third worst affected country by coronavirus, overtakes Russia Read more at: https://www.deccanherald.com/national/india-becomes-third-worst-affected-country-by-coronavirus-overtakes-russia-857442.html [Internet]. Deccan Herald. [cited 2020 Aug 3]. Available from: https://www.deccanherald.com/national/india-becomes-third-worst-affected-country-by-coronavirus-overtakes-russia-857442.html .

Basu D, Salvatore M, Ray D, Kleinsasser M, Purkayastha S, Bhattacharyya R, et al. A Comprehensive Public Health Evaluation of Lockdown as a Non-pharmaceutical Intervention on COVID-19 Spread in India: National Trends Masking State Level Variations [Internet]. Epidemiology. 2020; [cited 2020 Aug 3]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.05.25.20113043 .

IHME COVID-19 health service utilization forecasting team, Murray CJ. Forecasting COVID-19 impact on hospital bed-days, ICU-days, ventilator-days and deaths by US state in the next 4 months [Internet]. Infect Dis (except HIV/AIDS). 2020; [cited 2020 Aug 18]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.03.27.20043752 .

Imperial College COVID-19 Response Team, Flaxman S, Mishra S, Gandy A, Unwin HJT, Mellan TA, et al. Estimating the effects of non-pharmaceutical interventions on COVID-19 in Europe. Nature. 2020; [cited 2020 Aug 7]; Available from: http://www.nature.com/articles/s41586-020-2405-7 .

Tang L, Zhou Y, Wang L, Purkayastha S, Zhang L, He J, et al. A Review of Multi-Compartment Infectious Disease Models. Int Stat Rev. 2020;88:462–513. https://doi.org/10.1111/insr.12402 .

Kermack WO, McKendrick AG. Contributions to the mathematical theory of epidemics—I. Bull Math Biol. 1991;53(1–2):33–55. https://doi.org/10.1007/BF02464423 .

Article   CAS   PubMed   Google Scholar  

Song PX, Wang L, Zhou Y, He J, Zhu B, Wang F, et al. An epidemiological forecast model and software assessing interventions on COVID-19 epidemic in China. medRxiv. 2020; Available from: https://www.medrxiv.org/content/10.1101/2020.02.29.20029421v1 .

Zhou Y, Wang L, Zhang L, Shi L, Yang K, He J, et al. A Spatiotemporal Epidemiological Prediction Model to Inform County-Level COVID-19 Risk in the United States. Harv Data Sci Rev. 2020; [cited 2020 Aug 3]; Available from: https://hdsr.mitpress.mit.edu/pub/qqg19a0r .

Wu JT, Leung K, Leung GM. Nowcasting and forecasting the potential domestic and international spread of the 2019-nCoV outbreak originating in Wuhan, China: a modelling study. Lancet. 2020;395(10225):689–97. https://doi.org/10.1016/S0140-6736(20)30260-9 .

Article   CAS   PubMed   PubMed Central   Google Scholar  

Hao X, Cheng S, Wu D, Wu T, Lin X, Wang C. Reconstruction of the full transmission dynamics of COVID-19 in Wuhan. Nature. 2020; [cited 2020 Aug 18]; Available from: http://www.nature.com/articles/s41586-020-2554-8 .

Bai Y, Yao L, Wei T, Tian F, Jin D-Y, Chen L, et al. Presumed asymptomatic carrier transmission of COVID-19. JAMA. 2020;323(14):1406–7. https://doi.org/10.1001/jama.2020.2565 .

Tong Z-D, Tang A, Li K-F, Li P, Wang H-L, Yi J-P, et al. Potential Presymptomatic transmission of SARS-CoV-2, Zhejiang Province, China, 2020. Emerg Infect Dis. 2020;26(5):1052–4. https://doi.org/10.3201/eid2605.200198 .

Bertozzi AL, Franco E, Mohler G, Short MB, Sledge D. The challenges of modeling and forecasting the spread of COVID-19. Proc Natl Acad Sci. 2020;2:202006520.

Bhardwaj R. A predictive model for the evolution of COVID-19. Trans Indian Natl Acad Eng. 2020;5(2):133–40. https://doi.org/10.1007/s41403-020-00130-w .

Article   Google Scholar  

Bhaduri R, Kundu R, Purkayastha S, Kleinsasser M, Beesley LJ, Mukherjee B. Extending the susceptible-exposed-infected-removed (SEIR) model to handle the high false negative rate and symptom-based administration of COVID-19 diagnostic tests: SEIR-fansy [Internet]. Epidemiology. 2020; [cited 2021 Feb 20]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.09.24.20200238 .

Unwin HJT, Mishra S, Bradley VC, Gandy A, Mellan TA, Coupland H, et al. State-level tracking of COVID-19 in the United States [Internet]. Public Glob Health. 2020; [cited 2020 Sep 16]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.07.13.20152355 .

Mellan TA, Hoeltgebaum HH, Mishra S, Whittaker C, Schnekenberg RP, Gandy A, et al. Subnational analysis of the COVID-19 epidemic in Brazil [Internet]. Epidemiology. 2020; [cited 2020 Sep 16]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.05.09.20096701 .

Vollmer MAC, Mishra S, Unwin HJT, Gandy A, Mellan TA, Bradley V, et al. A sub-national analysis of the rate of transmission of COVID-19 in Italy [Internet]. Public Glob Health. 2020; [cited 2020 Sep 16]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.05.05.20089359 .

Lau H, Khosrawipour T, Kocbach P, Ichii H, Bania J, Khosrawipour V. Evaluating the massive underreporting and undertesting of COVID-19 cases in multiple global epicenters. Pulmonology. 2021;27(2):110–15. https://doi.org/10.1016/j.pulmoe.2020.05.015 .

Wang D, Hu B, Hu C, Zhu F, Liu X, Zhang J, et al. Clinical characteristics of 138 hospitalized patients with 2019 novel coronavirus–infected pneumonia in Wuhan, China. JAMA. 2020;323(11):1061–9. https://doi.org/10.1001/jama.2020.1585 .

Wangping J, Ke H, Yang S, Wenzhe C, Shengshu W, Shanshan Y, et al. Extended SIR prediction of the epidemics trend of COVID-19 in Italy and compared with Hunan, China. Front Med. 2020;7:169. https://doi.org/10.3389/fmed.2020.00169 .

Wang L, Zhou Y, He J, Zhu B, Wang F, Tang L, et al. An epidemiological forecast model and software assessing interventions on COVID-19 epidemic in China [Internet]. Infect Dis (except HIV/AIDS). 2020; [cited 2021 Mar 19]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.02.29.20029421 .

Bhaduri R, Kundu R, Purkayastha S, Beesley LJ, Kleinsasser M, Mukherjee B. SEIRfansy: extended susceptible-exposed-infected-recovery model [Internet]. 2020. Available from: https://CRAN.R-project.org/package=SEIRfansy

Gelman A. Bayesian data analysis. 3rd ed. Boca Raton: CRC Press; 2014. p. 661. (Chapman & Hall/CRC texts in statistical science)

R Core Team. R: A Language and Environment for Statistical Computing [Internet]. Vienna: R Foundation for Statistical Computing; 2017. Available from: https://www.R-project.org/

Butcher JC. Numerical methods for ordinary differential equations. 2nd ed. Chichester; Hoboken: Wiley; 2008. p. 463. https://doi.org/10.1002/9780470753767 .

Book   Google Scholar  

Liu Y, Gayle AA, Wilder-Smith A, Rocklöv J. The reproductive number of COVID-19 is higher compared to SARS coronavirus. J Travel Med. 2020;27(2):taaa021.

Article   PubMed   Google Scholar  

Cori A, Ferguson NM, Fraser C, Cauchemez S. A new framework and software to estimate time-varying reproduction numbers during epidemics. Am J Epidemiol. 2013;178(9):1505–12. https://doi.org/10.1093/aje/kwt133 .

Verity R, Okell LC, Dorigatti I, Winskill P, Whittaker C, Imai N, et al. Estimates of the severity of coronavirus disease 2019: a model-based analysis. Lancet Infect Dis. 2020s;20(6):669–77. https://doi.org/10.1016/S1473-3099(20)30243-7 .

Plummer M. rjags: Bayesian graphical models using MCMC. R package version 4-10. 2019.  https://CRAN.R-project.org/package=rjags .

Li R, Pei S, Chen B, Song Y, Zhang T, Yang W, et al. Substantial undocumented infection facilitates the rapid dissemination of novel coronavirus (SARS-CoV-2). Science. 2020;368(6490):489–93. https://doi.org/10.1126/science.abb3221 .

He X, Lau EHY, Wu P, Deng X, Wang J, Hao X, et al. Temporal dynamics in viral shedding and transmissibility of COVID-19. Nat Med. 2020s;26(5):672–5. https://doi.org/10.1038/s41591-020-0869-5 .

Li Q, Guan X, Wu P, Wang X, Zhou L, Tong Y, et al. Early transmission dynamics in Wuhan, China, of novel coronavirus–infected pneumonia. N Engl J Med. 2020;382(13):1199–207. https://doi.org/10.1056/NEJMoa2001316 .

Ferretti L, Wymant C, Kendall M, Zhao L, Nurtay A, Abeler-Dörner L, et al. Quantifying SARS-CoV-2 transmission suggests epidemic control with digital contact tracing. Science. 2020;368(6491):eabb6936.

Mishra V, Burma A, Das S, Parivallal M, Amudhan S, Rao G. COVID-19-hospitalized patients in Karnataka: survival and stay characteristics. Indian J Public Health. 2020;64(6):221.

Garg S, Kim L, Whitaker M, O’Halloran A, Cummings C, Holstein R, et al. Hospitalization rates and characteristics of patients hospitalized with laboratory-confirmed coronavirus disease 2019 — COVID-NET, 14 states, march 1–30, 2020. MMWR Morb Mortal Wkly Rep. 2020;69(15):458–64. https://doi.org/10.15585/mmwr.mm6915e3 .

Rahmandad H, Lim TY, Sterman J. Estimating the Global Spread of COVID-19. SSRN Electron J. 2020; [cited 2021 Mar 18]; Available from: https://www.ssrn.com/abstract=3635047 .

Diekmann O, Heesterbeek JAP, Roberts MG. The construction of next-generation matrices for compartmental epidemic models. J R Soc Interface. 2010;7(47):873–85. https://doi.org/10.1098/rsif.2009.0386 .

Robert CP, Casella G. Monte Carlo statistical methods [internet]. New York: Springer New York; 2004. [cited 2020 Aug 14]. (Springer Texts in Statistics). Available from: http://link.springer.com/10.1007/978-1-4757-4145-2

Scott J, Gandy A, Mishra S, Unwin J, Flaxman S, Bhatt S. epidemia: Modeling of Epidemics using Hierarchical Bayesian Models [Internet]. 2020. Available from: https://imperialcollegelondon.github.io/epidemia/

Bi Q, Wu Y, Mei S, Ye C, Zou X, Zhang Z, et al. Epidemiology and transmission of COVID-19 in 391 cases and 1286 of their close contacts in Shenzhen, China: a retrospective cohort study. Lancet Infect Dis. 2020;20(8):911–9. https://doi.org/10.1016/S1473-3099(20)30287-5 .

Bhattacharyya R, Bhaduri R, Kundu R, Salvatore M, Mukherjee B. Reconciling epidemiological models with misclassified case-counts for SARS-CoV-2 with seroprevalence surveys: A case study in Delhi, India [Internet]. Infect Dis (except HIV/AIDS). 2020; Aug [cited 2021 Mar 19]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.07.31.20166249 .

Murhekar MV, Bhatnagar T, Selvaraju S, Saravanakumar V, Thangaraj JWV, Shah N, et al. SARS-CoV-2 antibody seroprevalence in India, august–September, 2020: findings from the second nationwide household serosurvey. Lancet Glob Health. 2021;9(3):e257–66. https://doi.org/10.1016/S2214-109X(20)30544-1 .

Article   PubMed   PubMed Central   Google Scholar  

Walker PGT, Whittaker C, Watson OJ, Baguelin M, Winskill P, Hamlet A, Djafaara BA, Cucunubá Z, Olivera Mesa D, Green W, Thompson H, Nayagam S, Ainslie KEC, Bhatia S, Bhatt S, Boonyasiri A, Boyd O, Brazeau NF, Cattarino L, Cuomo-Dannenburg G, Dighe A, Donnelly CA, Dorigatti I, van Elsland SL, FitzJohn R, Fu H, Gaythorpe KAM, Geidelberg L, Grassly N, Haw D, Hayes S, Hinsley W, Imai N, Jorgensen D, Knock E, Laydon D, Mishra S, Nedjati-Gilani G, Okell LC, Unwin HJ, Verity R, Vollmer M, Walters CE, Wang H, Wang Y, Xi X, Lalloo DG, Ferguson NM, Ghani AC. The impact of COVID-19 and strategies for mitigation and suppression in low- and middle-income countries. Science. 2020;369(6502):413–22. https://doi.org/10.1126/science.abc0035 . Epub 2020 Jun 12.

Carpenter B, Gelman A, Hoffman MD, Lee D, Goodrich B, Betancourt M, et al. Stan : A Probabilistic Programming Language. J Stat Softw. 2017;76(1) [cited 2020 Aug 29]. Available from: http://www.jstatsoft.org/v76/i01/ .

India C-19. Coronavirus Outbreak in India [Internet]. 2020 [cited 2020 May 21]. Available from: https://www.covid19india.org

Johns Hopkins University. COVID-19 Dashboard by the Center for Systems Science and Engineering (CSSE) at Johns Hopkins University (JHU) [Internet]. 2020 [cited 2020 May 21]. Available from: https://coronavirus.jhu.edu/map.html

Lin LI-K. A concordance correlation coefficient to evaluate reproducibility. Biometrics. 1989;45(1):255–68. https://doi.org/10.2307/2532051 .

Group C-I-19 S. COVID-19 Outbreak in India [Internet]. 2020 [cited 2020 May 21]. Available from: https://umich-biostatistics.shinyapps.io/covid19/

Ray D, Salvatore M, Bhattacharyya R, Wang L, Du J, Mohammed S, et al. Predictions, Role of Interventions and Effects of a Historic National Lockdown in India’s Response to the the COVID-19 Pandemic: Data Science Call to Arms. Harv Data Sci Rev. 2020; Available from: https://hdsr.mitpress.mit.edu/pub/r1qq01kw .

Enrique Amaro J, Dudouet J, Nicolás OJ. Global analysis of the COVID-19 pandemic using simple epidemiological models. Appl Math Model. 2021;90:995–1008. https://doi.org/10.1016/j.apm.2020.10.019 .

Orzechowska M, Bednarek AK. Forecasting COVID-19 pandemic in Poland according to government regulations and people behavior [Internet]. Infect Dis (except HIV/AIDS). 2020; [cited 2021 Mar 19]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.05.26.20112458 .

Singh BC, Alom Z, Rahman MM, Baowaly MK, Azim MA. COVID-19 Pandemic Outbreak in the Subcontinent: A data-driven analysis. ArXiv200809803 Cs. 2020; [cited 2021 Mar 19]; Available from: http://arxiv.org/abs/2008.09803 .

Gu X, Mukherjee B, Das S, Datta J. COVID-19 prediction in South Africa: estimating the unascertained cases- the hidden part of the epidemiological iceberg [Internet]. Epidemiology. 2020; [cited 2021 Mar 21]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.12.10.20247361 .

Vehtari A, Gelman A, Gabry J. Practical Bayesian model evaluation using leave-one-out cross-validation and WAIC. Stat Comput. 2017;27(5):1413–32. https://doi.org/10.1007/s11222-016-9696-4 .

Bürkner P-C, Gabry J, Vehtari A. Approximate leave-future-out cross-validation for Bayesian time series models. J Stat Comput Simul. 2020;90(14):2499–523. https://doi.org/10.1080/00949655.2020.1783262 .

Unwin HJT, Mishra S, Bradley VC, Gandy A, Mellan TA, Coupland H, et al. State-level tracking of COVID-19 in the United States. Nat Commun. 2020;11(1):6189. https://doi.org/10.1038/s41467-020-19652-6 .

Candido DS, Claro IM, de Jesus JG, Souza WM, Moreira FRR, Dellicour S, et al. Evolution and epidemic spread of SARS-CoV-2 in Brazil. Science. 2020;369(6508):1255–60. https://doi.org/10.1126/science.abd2161 .

Mishra S, Scott J, Zhu H, Ferguson NM, Bhatt S, Flaxman S, et al. A COVID-19 Model for Local Authorities of the United Kingdom [Internet]. Infect Dis (except HIV/AIDS). 2020; [cited 2021 Mar 20]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.11.24.20236661 .

Gandy A, Mishra S. ImperialCollegeLondon/covid19local: Website Release for Wednesday 1tth Mar 2021, new doi for the week [Internet]. Zenodo. 2021; [cited 2021 Mar 20]. Available from: https://zenodo.org/record/4609660 .

Scottish Government. Coronavirus (COVID-19): modelling the epidemic [Internet]. Available from: https://www.gov.scot/collections/coronavirus-covid-19-modelling-the-epidemic/ .

Cuomo AM. American crisis; 2020.

Salvatore M, Basu D, Ray D, Kleinsasser M, Purkayastha S, Bhattacharyya R, et al. Comprehensive public health evaluation of lockdown as a non-pharmaceutical intervention on COVID-19 spread in India: national trends masking state-level variations. BMJ Open. 2020;10(12):e041778. https://doi.org/10.1136/bmjopen-2020-041778 .

Rahmandad H, Lim TY, Sterman J. Estimating COVID-19 under-reporting across 86 nations: implications for projections and control [Internet]. Epidemiology. 2020; [cited 2020 Sep 16]. Available from: http://medrxiv.org/lookup/doi/10.1101/2020.06.24.20139451 .

Balabdaoui F, Mohr D. Age-stratified discrete compartment model of the COVID-19 epidemic with application to Switzerland. Sci Rep. 2020;10(1):21306. https://doi.org/10.1038/s41598-020-77420-4 .

Download references

Acknowledgements

The authors would like to thank the Center for Precision Health Data Sciences at the University of Michigan School of Public Health, The University of Michigan Rogel Cancer Center and the Michigan Institute of Data Science for internal funding that supported this research. The authors are grateful to Professors Eric Fearon, Aubree Gordon and Parikshit Ghosh for useful conversations that helped formulating the ideas in this manuscript.

The authors would like to thank the Center for Precision Health Data Sciences at the University of Michigan School of Public Health, The University of Michigan Rogel Cancer Center and the Michigan Institute of Data Science. The funding bodies provided internal funding that supported this project and funded computational resources used to analyse and draw inferences from the data.

Author information

Authors and affiliations.

Department of Biostatistics, University of Michigan, Ann Arbor, MI, 48109, USA

Soumik Purkayastha, Rupam Bhattacharyya, Xuelin Gu, Maxwell Salvatore & Bhramar Mukherjee

Indian Statistical Institute, Kolkata, West Bengal, 700108, India

Ritwik Bhaduri & Ritoban Kundu

Center for Precision Health Data Science, University of Michigan, Ann Arbor, MI, 48109, USA

Xuelin Gu, Maxwell Salvatore & Bhramar Mukherjee

Department of Epidemiology, University of Michigan, Ann Arbor, MI, 48109, USA

Maxwell Salvatore & Bhramar Mukherjee

Department of Epidemiology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD, 21205, USA

Debashree Ray

Department of Biostatistics, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD, 21205, USA

School of Public Health, Imperial College London, London, W2 1PG, UK

Swapnil Mishra

You can also search for this author in PubMed   Google Scholar

Contributions

SP drafted the main paper and prepared all numerical items (Tables and Figures). RB1 and MS (eSIR), XG (SAPHIRE), RK and RB2 (SEIR-fansy) and SM (ICM) implemented the different models. DR helped with planning analysis and writing strategies to address reviewer concerns in the revised version. BM designed the study, revised the draft, provided strategic guidance and oversaw the analysis and the writing. All authors participated in writing and reviewing this manuscript. The authors read and approved the final manuscript.

Corresponding author

Correspondence to Bhramar Mukherjee .

Ethics declarations

Ethics approval and consent to participate.

Not applicable (uses publicly available data).

Consent for publication

Not Applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s note.

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

Supplementary Information

Additional file 1: supplementary table s1..

Summary of initial values and parameter settings for application of the SEIR-fansy model in the context of COVID-19 data from India. Unless mentioned otherwise, we use these parameter settings for all other models when applicable. Supplementary Table S2. Overview of projected COVID-counts for each model considered. Supplementary Table S3. Comparison of estimated projections and posterior estimates of model parameters across different sensitivity analysis scenarios under 21-day lockdown with moderate return, using observed data till April 14. Prior SD for R0 is 1.0. Reproduced from Ray et al., 2020 [ 53 ]. Supplementary Table S4. National and state-levels lockdown measures implemented over the course of COVID-19 pandemic in India. Reproduced from Salvatore et al., 2021 [ 66 ].

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ . The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Cite this article.

Purkayastha, S., Bhattacharyya, R., Bhaduri, R. et al. A comparison of five epidemiological models for transmission of SARS-CoV-2 in India. BMC Infect Dis 21 , 533 (2021). https://doi.org/10.1186/s12879-021-06077-9

Download citation

Received : 22 October 2020

Accepted : 15 April 2021

Published : 07 June 2021

DOI : https://doi.org/10.1186/s12879-021-06077-9

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Compartmental models
  • Low and middle income countries
  • Prediction uncertainty
  • Statistical models

BMC Infectious Diseases

ISSN: 1471-2334

scientific method mystery virus case study simulation

Exploring the Scientific Method

Exploring the Scientific Method

Cases and questions.

Edited by Steven Gimbel

From their grade school classrooms forward, students of science are encouraged to memorize and adhere to the “scientific method”—a model of inquiry consisting of five to seven neatly laid-out steps, often in the form of a flowchart. But walk into the office of a theoretical physicist or the laboratory of a biochemist and ask “Which step are you on?” and you will likely receive a blank stare. This is not how science works. But science does work, and here award-winning teacher and scholar Steven Gimbel provides students the tools to answer for themselves this question: What actually is the scientific method?

            Exploring the Scientific Method pairs classic and contemporary readings in the philosophy of science with milestones in scientific discovery to illustrate the foundational issues underlying scientific methodology. Students are asked to select one of nine possible fields—astronomy, physics, chemistry, genetics, evolutionary biology, psychology, sociology, economics, or geology—and through carefully crafted case studies trace its historical progression, all while evaluating whether scientific practice in each case reflects the methodological claims of the philosophers. This approach allows students to see the philosophy of science in action and to determine for themselves what scientists do and how they ought to do it.

             Exploring the Scientific Method will be a welcome resource to introductory science courses and all courses in the history and philosophy of science.        

424 pages | 4 tables | 6 x 9 | © 2011

History of Science

Philosophy of Science

Physical Sciences: History and Philosophy of Physical Sciences

  • Table of contents
  • Author Events

Related Titles

“This is a truly unique approach for a textbook. The philosophical positions that Gimbel chooses to focus on are important and the choices of primary source articles are excellent. Exploring the Scientific Method will be attractive to anyone teaching courses on the history and philosophy of science.”

Mara Harrell, Carnegie Mellon University

“The way Gimbel integrates core readings in the philosophy of science with case studies works extremely well. As far as I know, Exploring the Scientific Method is the first book that does this, and I think this is exactly the approach that is needed to orient new students in the field.”

Mathias Frisch, University of Maryland

“All things considered, Gimbel succeeds in creating an innovative textbook that combines philosophical and historical approaches to the study of scientific method. Exploring the Scientific Method is not a comprehensive introduction to philosophy of science, and it does not provide an adequate foundation for advanced study in HPS, but those are not Gimbel’s intentions. Focusing on scientific method specifically, rather than on the broader scope of philosophy of science, allows Gimbel to include an impressive variety of material while maintaining the clear themes of characterizing scientific reasoning and the structure of theories. This volume is ideal for a course geared toward students in scientific and other disciplines who wish to gain insight into scientific method, and the unique integrated approach is invaluable for students with no background in HPS.”

Brooke Abounader | Isis

Table of Contents

Introduction

How to Use This Book

Syntactic View of Theories

Deductivism

Aristotle                      from Posterior Analytics and Physics

René Descartes            from Discourse on Method

Case Studies

Inductivism

Francis Bacon             from Novum Organum

Isaac Newton               from Principia

John Stuart Mill            from System of Logic

Hypothetico-Deductivism

William Whewell           from Novum Organum Renovatum

Rudolf Carnap              “Theoretical Procedures in Science”

R. B. Braithwaite          from Scientific Explanation

Paradoxes of Evidence

David Hume                 from Enquiry

Nelson Goodman         from Fact, Fiction, and Forecast

Carl Hempel                 from “Studies in the Logic of Confirmation”

Responses to the Paradoxes of Evidence

Falsificationism

Karl Popper                 from The Logic of Scientific Discovery

Holistic View of Theories

Pierre Duhem               from Aim and Structure of Physical Theory

Thomas Kuhn               from The Structure of Scientific Revolutions

Imre Lakatos                The Methodology of Research Programmes

Semantic View of Theories

Marshall Spector          “Models and Theories”

Max Black                   “Models and Archetypes”

Ronald Giere                from Explaining Science

Critical Views of Scientific Theories

Paul Feyerabend           from Against Method

Ruth Hubbard               “Science, Facts, and Feminism”

Bruno Latour                “The Science Wars: A Dialogue”

Closing Remarks

Deductivism Case Study Readings

Astronomy       Aristotle           from On the Heavens   

Physics Epicurus           from Letter to Herodotus        

Chemistry         Paracelsus        from Hermetic and Alchemical Writings

Genetics           Aristotle           from On the Generation of Animals

Evolutionary Biology     Aristotle           from On the Generation of Animals

Geology           John Woodward           from An Essay towards a Natural History of the Earth

Psychology       Hippocrates      from The Nature of Man, The Sacred Disease

Sociology         Thomas Hobbes           from Leviathan            

Economics        Aristotle           from Politics    

Inductivism Case Study Readings

Astronomy       Ptolemy            from Almagest

Physics James Clerk Maxwell   from “Molecules”

Chemistry         Robert Boyle    from The Skeptical Chymist

Genetics           Gregor Mendel from Experiments in Plant Hybridization

Evolutionary Biology     Carolus Linnaeus          from Systema Naturae

Geology           James Hutton    from “System of the Earth”

Psychology       Heinrich Weber            from “The Sense of Touch and the Common Feeling”

Sociology         Émile Durkheim            from Suicide

Economics        François Quesnay         from “Farmers”

Bibliography

Making "Nature"

Melinda Baldwin

Philip Ball

Making Modern Science, Second Edition

Peter J. Bowler

The Changing Frontier

Adam B. Jaffe

Be the first to know

Get the latest updates on new releases, special offers, and media highlights when you subscribe to our email lists!

Sign up here for updates about the Press

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • My Account Login
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 25 February 2021

Inefficiency of SIR models in forecasting COVID-19 epidemic: a case study of Isfahan

  • Shiva Moein 1 ,
  • Niloofar Nickaeen 2 ,
  • Amir Roointan 1 ,
  • Niloofar Borhani 2 ,
  • Zarifeh Heidary 2 ,
  • Shaghayegh Haghjooy Javanmard 3 ,
  • Jafar Ghaisari 2 &
  • Yousof Gheisari 1  

Scientific Reports volume  11 , Article number:  4725 ( 2021 ) Cite this article

27k Accesses

99 Citations

27 Altmetric

Metrics details

  • Computational models
  • Epidemiology
  • Viral infection

The multifaceted destructions caused by COVID-19 have been compared to that of World War II. What makes the situation even more complicated is the ambiguity about the duration and ultimate spread of the pandemic. It is especially critical for the governments, healthcare systems, and economic sectors to have an estimate of the future of this disaster. By using different mathematical approaches, including the classical susceptible-infected-recovered (SIR) model and its derivatives, many investigators have tried to predict the outbreak of COVID-19. In this study, we simulated the epidemic in Isfahan province of Iran for the period from Feb 14th to April 11th and also forecasted the remaining course with three scenarios that differed in terms of the stringency level of social distancing. Despite the prediction of disease course in short-term intervals, the constructed SIR model was unable to forecast the actual spread and pattern of epidemic in the long term. Remarkably, most of the published SIR models developed to predict COVID-19 for other communities, suffered from the same inconformity. The SIR models are based on assumptions that seem not to be true in the case of the COVID-19 epidemic. Hence, more sophisticated modeling strategies and detailed knowledge of the biomedical and epidemiological aspects of the disease are needed to forecast the pandemic.

Similar content being viewed by others

scientific method mystery virus case study simulation

A real-world data validation of the value of early-stage SIR modelling to public health

scientific method mystery virus case study simulation

Forecasting the long-term trend of COVID-19 epidemic using a dynamic model

scientific method mystery virus case study simulation

Mathematical epidemiologic and simulation modelling of first wave COVID-19 in Malaysia

Introduction.

The Coronavirus disease 2019 (COVID-19), has severely affected different aspects of human health all over the world. The high transmissibility rate and lack of efficient therapeutics have transformed the situation into a major challenge 1 , 2 , 3 . Ranking fifth according to the outbreak, announced the first cases on February 19, 2020. The pandemic quickly impacted most regions of the country with the highest rates in Tehran, Qom and Isfahan provinces 4 , 5 .

Virus epidemics have been known as one of the major health issues causing a high mortality rate in human communities throughout history. The Spanish flue emerged in 1918, caused 50 million deaths for just over 2 years 6 . Also, since the early reports of HIV infection in 1980, more than 32 million deaths have been reported around the world due to virus infection 7 . Tragically, not only these older global epidemics but also, the local spread of SARS, MERS, and Ebola viruses in recent years have not made countries ready for such crises. The lack of preparation for the outbreak of COVID-19 beside the high rate of transmissibility, unknown complications, and inappropriate medications led to a world-wide disaster paralyzing the health care systems even in developed countries. More importantly, the future course of the disease is not yet clearly known in many affected countries and the estimates of the ultimate spread of the epidemic are not coherent. This makes policy-making and planning for the required resources very difficult for the governments 8 . In this unknown situation, forecasting approaches are of particular importance.

Since the emergence of COVID-19, different mathematical modeling approaches have been employed to simulate the disease course 9 ; the artificial intelligence-based models 10 are interesting but as they depend on many learning steps, their validity can be questioned in the absence of sufficient training datasets. Day-level forecasting based on time-series data is another approach 11 that simply follows previous patterns and cannot predict the trend of changes. Agent-based modeling is also a rational approach for predicting the disease course, which simulates the fellow of the individuals (agents) to calculate the disease spread in the community 12 . Nevertheless, such models rely on the population-level parameters such as rates of movements, distancing and virus infectivity parameters which are not yet known. Ordinary differential equation (ODE)-based models have been used for a long time to simulate the classical dynamics of epidemics 13 . This type of models was first proposed by Kermack and McKendrick in 1927 to simulate the transmission of infectious disease such as measles and rubella 14 . Such models assume susceptible (S), infective (I), and recovered (R) fractions in a close population and calculate the rate of changes in each fraction with ODEs 15 , 16 .

Over the past year, various simple and modified SIR models have been developed to predict the course of COVID-19; Anastassopoulou et al. provided a preliminary estimate of the fatality and recovery ratios for the population of Wuhan by a discrete-time SIR model 17 . Moreover, a model was developed by Giordano et al. that predicted the effect of different lockdown strategies on the epidemic in Italy. They considered different sub-groups regarding the stage and severity of the disease in the infected individuals 18 . Similarly, using an SEIR (Susceptible-Exposed-Infectious-Recovered) framework, Lin et al. predicted the effect of government policies and individual actions on the spread of the epidemic in China 19 . It should be noted that the modified SIR models require more complex data for the development, and due to the presence of little information and lack of reliable data regarding this newly emerged disease, the simple SIR model has been the choice of many investigators 20 .

Despite the simplicity and effectiveness of SIR-based models in forecasting a variety of other infectious diseases, their capability in the case of the COVID-19 pandemic is a matter of debate. In this regard, an SIR model was developed using the data of the first 3 months of COVID-19 spread in Isfahan province and the conformity and accuracy of the model was evaluated using the actual data in later months. Results of this study scrutinize the appropriateness of SIR models for the simulation and prediction of the COVID-19 epidemics.

Material and methods

In the SIR models, the population is considered to be closed and the sum of susceptible, infective or recovered fractions denoted by \(i\left( t \right)\) , \(s\left( t \right)\) , and \(r\left( t \right)\) , respectively, is equal to 1 for each \(t \ge 0\) . The model is defined by the following set of ODEs:

where \(\lambda\) is the infection rate and \(\gamma\) is the recovery rate. R 0 is known as the basic reproduction number and defined as:

R 0 is an essential determinant of outbreaks and can be interpreted as the expected number of new cases directly caused by an infectious individual before the recovery.

All simulations were performed using MATLAB R2017a 21 .

Epidemiological data

The data used in this study was provided by the medical care monitoring center (MCMC) of Isfahan University of Medical Sciences, Isfahan, Iran. This data consists of the daily hospitalized cases and deaths from all cities of Isfahan province except Kashan, Aran and Bidgol which are in the zone of Kashan University of Medical Sciences. The population of the studied region is approximately 4,632,000. Considering the low negative predictive value of PCR test for the diagnosis of COVID-19 22 , all patients admitted to hospitals with the clinical suspicion of the disease were considered as infected cases. This can describe the potential discrepancy between the actual cases assumed in this study and other reports that rely on the molecular diagnosis of the infection.

Using the epidemiologic data for the first 3 months, an SIR-based model was constructed to predict the disease course

In the SIR model, \(r\left( t \right) + i\left( t \right){ }\) determines all the individuals who are no longer susceptible. The term can be interpreted as the accumulative number of new cases in each time-point that can be recovered, died or are still infectious due to the course of disease. According to recent studies, 30% of infected individuals are asymptomatic that are not identified in the population unless in active case finding programs 23 . Out of the remaining infected cases that manifest the disease, 70% show mild to moderate symptoms and are treated in out-patient settings without any need to hospitalization. whereas, 30% of symptomatic cases have severe manifestations and must be admitted to hospitals for a while during the infection 24 , 25 . Hence, the cumulative number of new hospital-admitted cases can be estimated from Eq. ( 5 ):

In addition, the number of cumulative deaths is assumed as:

Although a mortality rate of 2–3% is reported for COVID-19 patients 26 , the actual rate depends on a variety of parameters including the average age of the population and the capacity of healthcare systems. In this regard, the parameter \(K\) is determined for each population individually based on the reported mortality rates.

To estimate the R 0 value for the COVID-19 epidemic, the model was first fitted to the cumulative new cases data reported until April 10th, 2020 from Wuhan and Italy and the epidemic final size (Fig.  1 a,b), Cumulative hospitalized cases (Fig.  1 c,d), and the Cumulative deaths (Fig.  1 e,f) were predicted. Then the model was fine-tuned based on Isfahan statistics.

figure 1

Modeling COVID-19 epidemic in Wuhan and Italy. To provide a rough estimation of the epidemic dynamics, the model was first generated with the reported data of Wuhan and Italy. ( a , b ) Epidemic final size. ( c , d ) Cumulative hospitalized cases. ( e , f ) Cumulative deaths.

An R 0 of 1.0078 could best describe the data of Italy. Also, the curves of Wuhan could be appropriately simulated when R 0 was assumed 1.0146 (Fig.  1 d). Notably, the number of new hospital-admitted cases in Italy showed a decline in days 40–50 which could be attributed to the strict social distancing policies imposed by the government (Fig.  1 c). To fit the model to the mortality data, parameter \(K\) (Eq.  6 ) was set at 0.025 for Italy. Notably, when the same value was considered for Wuhan, the simulation curve was far higher than the reported official data (Fig.  1 f). In agreement with this observation, the Chinese authorities have later declared the inaccuracy of the initial reports of deaths in Wuhan and boosted the number by 50% 27 .

The simulations for Wuhan and Italy allowed starting the modeling of the epidemic in Isfahan with a rough estimation of R 0 value. Although, each outbreak is described with a unique R 0 in classical SIR models 28 , 29 , we decided to use a set of R 0 values for different time intervals to account for the variations of the community behavior and inconsistency of social distancing regulations. At the time of model construction, the actual epidemiological data were available for the episode from Feb 14th to April 11th. The model had the best fit to the actual data when four different R 0 values were considered. Since the changes in social distancing are reflected in the hospital admission rates with a time delay of about 2 weeks 30 , the four R 0 values were corroborated with community behavior variations as represented by city traffic reports (data not shown).

To forecast the epidemic in this province, three scenarios based on the strictness level of social distancing were assumed and different R 0 values were considered in each case (Table 1 ). From Feb 14th to April 11th, the highest value of R 0 was 1.0165 (R 01 ) which is attributed to the beginning of the epidemic when people were not aware of the outbreak and thus, no restriction was imposed. In the “bad scenario”, R 0 again reached the same value after a transitional step. In addition, during the mentioned period, the smallest value of R 0 was 1.0040 (R 03 ). Hence, considering the practical issues, it is assumed that R 0 may again reach to a value as small as 1.0050 in the “good scenario”. Also, an intermediate value of 1.0095 was considered for the “feasible scenario”.

In the good scenario achieved by strict social distancing, the SIR model predicted that about 13,000 cases would be cumulatively hospitalized by the end of the epidemic and the total number of deaths would reach 800 cases. In this scenario, the curve of cumulative cases approaches a plateau on May 4 and by June 10, only 0–2 new cases would be admitted to hospitals on a daily basis (Fig.  2 a).

figure 2

Modeling COVID-19 epidemic in Isfahan province using the first 3 month data. To forecast the epidemic, Good ( a ), Feasible ( b ) and Bad ( c ) Scenarios are assumed.

The feasible scenario was attributed to closing schools, universities, cultural centers and limiting social gatherings and businesses that are not critical. In this scenario, the model predicted that total hospitalized cases and deaths would reach about 15,000 and 920, respectively. Also, the curve of cumulative cases would approach a plateau on May 24 and by July 18, it predicted that only 0–2 new cases would be admitted to hospitals on a daily basis (Fig.  2 b). In the bad scenario, a situation was assumed in which the community is back to the routine life style. In this scenario, 22,000 cases would be totally admitted to the hospitals and the number of victims would be more than 1300. The model also forecasted that a second peak is unlikely and the epidemic would last up to mid-August (Fig.  2 c). The results of the three scenarios are summarized in Table 2 .

Evaluation of model predictions with actual epidemiologic data after 9 months

The prediction patterns in the three scenarios were evaluated by their conformity to actual data over the 9 months since the start of the epidemic. Although the good-scenario predictions were far from the pattern of epidemic (data not shown), the trend of feasible scenario could appropriately match with the daily hospitalized cases until mid-June (Fig.  3 a). However, a new long-lasting huger peak was appeared at the beginning of July which was not anticipated even in the bad scenario, leading to the increase of total infections to 43,293 until October 22 (Fig.  3 b). Although the model could forecast disease course patterns in the short term, the predicted values were far from the actual data in later time intervals and the constructed model was unsuccessful at long term prediction of epidemic course.

figure 3

Evaluation of model predictions after 9 months. Actual data until October 22 are demonstrated in line with model predictions in Feasible ( a ) and Bad ( b ) scenarios. In each graph, the time-intervals for model construction and validation are indicated.

COVID-19 pandemic is emerged as a massive health burden all over the world. Mathematical modeling is a useful approach for the prediction of disease dynamics to take best decisions. Currently, various mathematical approaches have been developed for modeling the COVID-19 pandemic which mostly depend on a number of parameters for the simulation 31 . Due to the lack of access to adequate information about this newly emerged pandemic, simple strategies such as SIR-based models have been implemented by several investigators. However, their accuracy and suitability for prediction of the COVID-19 disease course is a matter of debate. In this study, using the data of the first 3 months of the COVID-19 epidemic in Isfahan, an SIR model was developed and the outcome was predicted with three scenarios that differ in terms of social distancing stringency. Then, the constructed model was evaluated for conforming the predicted values to the actual data in the six consecutive months. Although the model could appropriately simulate the patterns until the mid-June, it was unable to predict the second peak of epidemic emerged in July.

The inability of our model to predict COVID-19 pandemic in Isfahan motivated us to assess the conformity of the predictions of SIR models developed by other investigators to the actual data of different regions around the world. Although the SIR model developed by Ahmetolan et al. could predict the course of disease in the countries with controlled epidemics, such as China, Korea, and Singapore, it was unable to forecast the larger peak mainly started at the late summer in many countries located in the northern hemisphere 32 . Similarly, the modified SIR model constructed by Cooper et al., which was manually adjusted to fit actual data in a period of time 33 , was unable to predict the next surges, and the estimates of infected populations were also far from reality. For instance, the model predicted 330,000 infections at the end of September in India, but it actually turned to be around 900,000 34 . Morover, the extended SIR model of Wangping and colleagues estimated that the epidemic in Italy would be ended in August with a total number of 182,051 infected cases. Nevertheless, according to actual data, a second peak started in October with more than 700,000 total infected cases 35 . Similar inconsistencies between predictions of SIR-based models and actual data for the COVID-19 epidemic could be observed in other studies 28 , 36 , 37 , 38 .

The failure of SIR-based models to forecast the COVID-19 pandemic can be described by a variety of reasons. These over-simplified models ignore the factors that have a great effect on the course of disease. For instance, various studies demonstrate that air pollution is a main reason for the observed variations in disease contagion 39 , 40 . The other issue affecting the spread of virus is the behavioral changes considerably related to the social and cultural context of the population 41 . As mostly a single R 0 value is considered in the SIR models, the unexpected changes in the social behaviors of the population are missed and the model would be unable to follow the emerging alterations. Indeed, ceremonies and national gatherings have a great impact on disease spread. For instance, a religious meeting in Malaysia, held on February 27 to March 3 is supposed to be the source of virus spread in India and Pakistan 42 . Similar social events in Isfahan, at least partly, may describe the unpredicted peak starting from July.

Another explanation for the failure of SIR-based models to predict the COVID-19 epidemics is that the modeling is based on the assumptions that are not necessarily true; the population is considered closed in the SIR models, whereas, complete isolation was not followed in most regions, making them vulnerable to changes in the neighboring communities. In addition, the recovered individuals are assumed as immunized in the SIR models, which are no longer susceptible. This assumption contrasts with new findings suggesting there is a possibility of the reactivation of the virus or reinfection of previously infected individuals 43 , 44 , 45 . Indeed, a recent study indicates that the longevity of neutralizing antibodies in infected individuals is only around 50–60 days 46 .

Taken together, the COVID-19 pandemic features are not coherent with the SIR modeling framework and the dynamics of this outbreak is under the influence of various parameters for most of which quantitative information is not yet available. More sophisticated modeling approaches in line with more precise epidemiological and biomedical data are urgently required to make the pandemic forecasting feasible.

Liu, Y., Gayle, A. A., Wilder-Smith, A. & Rocklöv, J. The reproductive number of COVID-19 is higher compared to SARS coronavirus. J. Travel Med. 27 , 1–4. https://doi.org/10.1093/jtm/taaa021 (2020).

Article   Google Scholar  

Ferguson, N. et al. Impact of non-pharmaceutical interventions (NPIs) to reduce COVID19 mortality and healthcare demand. Imperial Coll. Lond. https://doi.org/10.25561/77482 (2020).

Yuan, J., Li, M., Lv, G. & Lu, Z. K. Monitoring transmissibility and mortality of COVID-19 in Europe. Int. J. Infect. Dis. 95 , 311–315 (2020).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Aminian, A., Safari, S., Razeghian-Jahromi, A., Ghorbani, M. & Delaney, C. P. COVID-19 outbreak and surgical practice: Unexpected fatality in perioperative period. Ann. Surg. 10 , 27–29 (2020).

Abdi, M. Coronavirus disease 2019 (COVID-19) outbreak in Iran: Actions and problems. Infect. Control Hosp. Epidemiol. https://doi.org/10.1017/ice.2020.86 (2020).

Article   PubMed   PubMed Central   Google Scholar  

Taubenberger, J. K. & Morens, D. M. 1918 Influenza: The mother of all pandemics. Rev. Biomed. 17 , 69–79 (2006).

Google Scholar  

UNAIDS. https://www.unaids.org/en/resources/fact-sheet .

Gates, B. Responding to Covid-19—A once-in-a-century pandemic?. N. Engl. J. Med. 382 , 1677–1679 (2020).

Article   CAS   PubMed   Google Scholar  

Rabajante, J. F. Insights from early mathematical models of 2019-nCoV acute respiratory disease (COVID-19) dynamics. arXiv preprint arXiv:2002.05296 (2020).

Hu, Z., Ge, Q., Jin, L. & Xiong, M. Artificial intelligence forecasting of covid-19 in china. arXiv preprint arXiv:2002.07112 (2020).

Elmousalami, H. H. & Hassanien, A. E. Day level forecasting for coronavirus disease (COVID-19) spread: Analysis, modeling and recommendations. arXiv preprint arXiv:2003.07778 (2020).

Kim, Y., Ryu, H. & Lee, S. Agent-based modeling for super-spreading events: A case study of MERS-CoV transmission dynamics in the Republic of Korea. Int. J. Environ. Res. Public Health 15 , 2369 (2018).

Article   PubMed Central   Google Scholar  

Satsuma, J., Willox, R., Ramani, A., Grammaticos, B. & Carstea, A. Extending the SIR epidemic model. Phys. A 336 , 369–375 (2004).

Bacaër, N. In A Short History of Mathematical Population Dynamics (ed Nicolas, B.) 89–96 (Springer, London, 2011).

Zhou, Y., Ma, Z. & Brauer, F. A discrete epidemic model for SARS transmission and control in China. Math. Comput. Model. 40 , 1491–1506 (2004).

Article   MathSciNet   PubMed   Google Scholar  

Britton, T. Stochastic epidemic models: A survey. Math. Biosci. 225 , 24–35 (2010).

Anastassopoulou, C., Russo, L., Tsakris, A. & Siettos, C. Data-based analysis, modelling and forecasting of the COVID-19 outbreak. PLoS ONE 15 , e0230405. https://doi.org/10.1371/journal.pone.0230405 (2020).

Giordano, G. et al. Modelling the COVID-19 epidemic and implementation of population-wide interventions in Italy. Nat. Med. https://doi.org/10.1038/s41591-020-0883-7 (2020).

Lin, Q. et al. A conceptual model for the coronavirus disease 2019 (COVID-19) outbreak in Wuhan, China with individual reaction and governmental action. Int. J. Infect. Dis. 93 , 211–216. https://doi.org/10.1016/j.ijid.2020.02.058 (2020).

Roda, W. C., Varughese, M. B., Han, D. & Li, M. Y. Why is it difficult to accurately predict the COVID-19 epidemic?. Infect. Dis. Model. 5 , 271–281. https://doi.org/10.1016/j.idm.2020.03.001 (2020).

Ben-David, U. et al. Genetic and transcriptional evolution alters cancer cell line drug response. Nature 560 , 325–330. https://doi.org/10.1038/s41586-018-0409-3 (2018).

Article   ADS   CAS   PubMed   PubMed Central   Google Scholar  

Kokkinakis, I., Selby, K., Favrat, B., Genton, B. & Cornuz, J. Covid-19 diagnosis: Clinical recommendations and performance of nasopharyngeal swab-PCR. Revue medicale suisse 16 , 699–701 (2020).

Article   PubMed   Google Scholar  

Wang, Y. et al. Clinical outcome of 55 asymptomatic cases at the time of hospital admission infected with SARS-Coronavirus-2 in Shenzhen, China. J. Infect. Dis. https://doi.org/10.1093/infdis/jiaa119 (2020).

Lauer, S. A. et al. The incubation period of coronavirus disease 2019 (COVID-19) from publicly reported confirmed cases: Estimation and application. Ann. Intern. Med. 172 , 577–582 (2020).

Liu, Y. et al. Viral dynamics in mild and severe cases of COVID-19. Lancet Infect. Dis. 20 , 656–657 (2020).

Wu, Z. & McGoogan, J. M. Characteristics of and important lessons from the coronavirus disease 2019 (COVID-19) outbreak in China: Summary of a report of 72,314 cases from the Chinese Center for Disease Control and Prevention. JAMA 323 , 1239 (2020).

Science, L. https://www.livescience.com/wuhan-coronavirus-death-toll-revised.html .

Zareie, B., Roshani, A., Mansournia, M. A., Rasouli, M. A. & Moradi, G. A model for COVID-19 prediction in Iran based on China parameters. medRxiv. https://doi.org/10.1101/2020.03.19.20038950 (2020).

Calafiore, G. C., Novara, C. & Possieri, C. A Modified SIR Model for the COVID-19 Contagion in Italy. arXiv preprint arXiv:2003.07778 (2020).

Coronavirus disease 2019 (COVID-19) pandemic: increased transmission in the EU/EEA and the UK—seventh update. https://www.ecdc.europa.eu/sites/default/files/documents/RRA-seventh-update-Outbreak-of-coronavirus-disease-COVID-19.pdf .

Tolles, J. & Luong, T. Modeling epidemics with compartmental models. JAMA 323 , 2515–2516. https://doi.org/10.1001/jama.2020.8420 (2020).

Ahmetolan, S., Bilge, A. H., Demirci, A., Peker-Dobie, A. & Ergonul, O. What can we estimate from fatality and infectious case data using the susceptible-infected-removed (SIR) model? A case study of Covid-19 pandemic. Front. Med. (Lausanne) 7 , 556366. https://doi.org/10.3389/fmed.2020.556366 (2020).

Cooper, I., Mondal, A. & Antonopoulos, C. G. Dynamic tracking with model-based forecasting for the spread of the COVID-19 pandemic. Chaos Solitons Fractals 139 , 110298. https://doi.org/10.1016/j.chaos.2020.110298 (2020).

Article   MathSciNet   PubMed   PubMed Central   Google Scholar  

Coronavirus, w. https://www.worldometers.info/coronavirus/ .

Wangping, J. et al. Extended SIR prediction of the epidemics trend of COVID-19 in Italy and compared with Hunan, China. Front. Med. (Lausanne). https://doi.org/10.3389/fmed.2020.00169 (2020).

Bastos, S. B. & Cajueiro, D. O. Modeling and forecasting the early evolution of the Covid-19 pandemic in Brazil. arXiv preprint arXiv:2003.14288 (2020).

Cintra, H. P. C. & Fontinele, F. N. Estimative of real number of infections by COVID-19 in Brazil and possible scenarios. Infect. Dis. Model. 5 , 720–736. https://doi.org/10.1016/j.idm.2020.09.004 (2020).

Guirao, A. The Covid-19 outbreak in Spain. A simple dynamics model, some lessons, and a theoretical framework for control response. Infect. Dis. Model. 5 , 652–669. https://doi.org/10.1016/j.idm.2020.08.010 (2020).

Fattorini, D. & Regoli, F. Role of the chronic air pollution levels in the Covid-19 outbreak risk in Italy. Environ. Pollut. 264 , 114732–114732. https://doi.org/10.1016/j.envpol.2020.114732 (2020).

Comunian, S., Dongo, D., Milani, C. & Palestini, P. Air pollution and Covid-19: The role of particulate matter in the spread and increase of Covid-19’s morbidity and mortality. Int. J. Environ. Res. Public Health 17 , 4487. https://doi.org/10.3390/ijerph17124487 (2020).

Article   CAS   PubMed Central   Google Scholar  

Bavel, J. J. V. et al. Using social and behavioural science to support COVID-19 pandemic response. Nat. Hum. Behav. 4 , 460–471. https://doi.org/10.1038/s41562-020-0884-z (2020).

Quadri, S. A. COVID-19 and religious congregations: Implications for spread of novel pathogens. Int. J. Infect. Dis. 96 , 219–221. https://doi.org/10.1016/j.ijid.2020.05.007 (2020).

Alizargar, J. Risk of reactivation or reinfection of novel coronavirus (COVID-19). J. Formos Med. Assoc. 119 , 1123–1123. https://doi.org/10.1016/j.jfma.2020.04.013 (2020).

Hoang, V. T., Dao, T. L. & Gautret, P. Recurrence of positive SARS-CoV-2 in patients recovered from COVID-19. J. Med. Virol. https://doi.org/10.1002/jmv.26056 (2020).

Gousseff, M. et al. Clinical recurrences of COVID-19 symptoms after recovery: Viral relapse, reinfection or inflammatory rebound?. J. Infect. https://doi.org/10.1016/j.jinf.2020.06.073 (2020).

Seow, J. et al. Longitudinal observation and decline of neutralizing antibody responses in the three months following SARS-CoV-2 infection in humans. Nat. Microbiol. https://doi.org/10.1038/s41564-020-00813-8 (2020).

Download references

Acknowledgements

This study was approved by research ethics committee of Isfahan University of Medical Sciences (IR.MUI.MED.REC.1399.044) and supported by a grant from this university (199025).

Author information

Authors and affiliations.

Regenerative Medicine Research Center, Isfahan University of Medical Sciences, Isfahan, 81746-73461, Iran

Shiva Moein, Amir Roointan & Yousof Gheisari

Department of Electrical and Computer Engineering, Isfahan University of Technology, Isfahan, 84156-83111, Iran

Niloofar Nickaeen, Niloofar Borhani, Zarifeh Heidary & Jafar Ghaisari

Department of Physiology, Applied Physiology Research Center, Isfahan Cardiovascular Research Institute, Isfahan University of Medical Sciences, Isfahan, Iran

Shaghayegh Haghjooy Javanmard

You can also search for this author in PubMed   Google Scholar

Contributions

J.G., Y.G., and S.H.J. conceptualized the main idea. S.H.J., Y.G., S.M., and A.R. gathered and interpreted the actual data. J.G., N.N., N.B. and Z.H. performed simulations. S.M. and Y.G. drafted the manuscript. J.G., N.N., A.R., and S.H.J. revised the manuscript. All authors made a substantial intellectual contribution to the work, and approved the submitted version for publication.

Corresponding authors

Correspondence to Jafar Ghaisari or Yousof Gheisari .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Additional information

Publisher's note.

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

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Moein, S., Nickaeen, N., Roointan, A. et al. Inefficiency of SIR models in forecasting COVID-19 epidemic: a case study of Isfahan. Sci Rep 11 , 4725 (2021). https://doi.org/10.1038/s41598-021-84055-6

Download citation

Received : 30 May 2020

Accepted : 11 February 2021

Published : 25 February 2021

DOI : https://doi.org/10.1038/s41598-021-84055-6

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

A study on the evolution of tripartite collaborative prevention and control under public health emergencies using covid-19 as an example.

  • Liu Mingyue

Scientific Reports (2024)

Social and economic variables explain COVID-19 diffusion in European regions

  • Christian Cancedda
  • Alessio Cappellato
  • Stefano Ceri

Epidemiology: Gray immunity model gives qualitatively different predictions

  • Milind Watve
  • Himanshu Bhisikar
  • Srashti Bajpai

Journal of Biosciences (2024)

Modeling Global Monkeypox Infection Spread Data: A Comparative Study of Time Series Regression and Machine Learning Models

  • Vishwajeet Singh
  • Saif Ali Khan
  • Yusuf Akhter

Current Microbiology (2024)

Real-time estimation and forecasting of COVID-19 cases and hospitalizations in Wisconsin HERC regions for public health decision making processes

  • Srikanth Aravamuthan
  • Juan Francisco Mandujano Reyes
  • Dörte Döpfer

BMC Public Health (2023)

By submitting a comment you agree to abide by our Terms and Community Guidelines . If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

scientific method mystery virus case study simulation

This short, targeted simulation is adapted from the full-length “Experimental Design” simulation.

How do research scientists know where to begin? In this simulation, you will learn how to design a scientific experiment. As a pharmaceutical detective, you have the chance to perform experiments with human volunteers, animals, and living human cells. Make sure that you choose the correct experimental model to design a scientifically sound experiment for testing the effect of the suspicious drug.

The scientific method

Humans have always strived to explain natural phenomena. In the first mission of the experimental design simulation, you will learn how to use the scientific method to investigate phenomena, acquire new knowledge, or correct and integrate existing knowledge. Every tool you need to use is available in this virtual lab.

Design an experiment and formulate a hypothesis

In your next mission you have the freedom to design your own experiment from scratch while examining biopsies under a microscope. Define your scientific question and choose the right model to test your hypothesis. And don’t be afraid to make mistakes - in this virtual simulation you can repeat the experiment as many times as you want.

Experimental controls

You will notice that there are many experimental variables in your experiment. Learn how to adjust them and understand why it’s so important to use experimental controls to verify your results. These controls will let you conclude  whether a suspicious compound is the reason behind the reported epidemic. You’ll get to investigate the effect of different compounds and various concentrations of the same compound.

Will you be able to set up an experiment that can prove your suspicions?

The Scientific Method | Virtual Lab

The Scientific Method

Lorem ipsum dolor sit amet, consectetur adipiscing elit, sed do eiusmod tempor incididunt ut labore et dolore magna aliqua. Ut enim ad minim veniam, quis nostrud exercitation ullamco laboris nisi ut aliquip ex ea commodo consequat. Duis aute irure dolor in reprehenderit in voluptate velit esse cillum dolore eu fugiat nulla pariatur.

Block quote

Ordered list

Unordered list

Superscript

About This Simulation

Work as a pharmaceutical detective to identify the link between a new drug and a recent epidemic. Use the scientific method to come up with a hypothesis and subsequently design an experiment that will test the validity of your hypothesis.

Learning Objectives

  • Describe the scientific method
  • Know what to consider when choosing an experimental model
  • Design an experiment and test a hypothesis
  • Select the correct experimental controls to verify your hypothesis

Preview of EX2_1.png simulation.

Lab Techniques

  • Apoptosis assay

Related Standards

  • Generic experimental design support

Learn More About This Simulation

Experience labster for yourself.

75% of students show high engagement and improved grades with Labster

Easily bolster your learning objectives with relevant, interactive content

Practice a lab procedure or visualize theory through narrative-driven scenarios

a group of people standing around a laptop computer

Related Course Packages

Bioscience lab skills, +35 other courses, related simulations.

SDS-PAGE: Separating proteins by molecular weight

SDS-PAGE: Separating proteins by molecular weight

Investigate each step of SDS-PAGE from gel selection and sample preparation to chamber assembly and what really happens when the current turns on, to separate proteins solely by molecular weight, bringing us one step closer to identifying the protein.

Animal Genetics

Animal Genetics

Learn about the Mendelian inheritance of double muscling in cattle. Find the responsible gene using linkage analysis and learn how mutation in this gene alters gene expression resulting in double muscled cattle.

Light Microscopy

Light Microscopy

Enter the virtual microscope room to see inside a tissue sample. Learn how a light microscope can magnify an image and answer biological questions.

For Science Programs Providing a Learning Advantage

Professor Margaret Brady was able to enhance student learning with A&P virtual labs.

“They did the simulation at home, then completed the in-person lab within 30 minutes, no questions asked, and passed the quiz with flying colors.”

scientific method mystery virus case study simulation

"I saw some of the students who clearly didn’t necessarily like sitting there reading a book discover they could turn on Labster and keep up with the rest of the class because it spoke to them.

scientific method mystery virus case study simulation

"Having something that's engaging for the students gives teachers that opportunity to breathe and get excited again. Because they're seeing the kids light up, they're seeing the kids engage with content."

user

"The question always is, ‘Can we demonstrate that the students are meeting course outcomes?’ Check! We can do that.”

scientific method mystery virus case study simulation

"We surveyed over 400 students. More than 90% thought Labster was easy to navigate, and that it was fun, but more importantly, most of them felt confident that they could execute the labs in person. And that confidence is a big deal."

a man in a black sweater and white shirt

“The Labster simulations get students to do things, and they're not just sitting there consuming a webinar where their mind can drift. They become an active participant in that learning experience.”

a black and white photo of a clock tower

Find answers to frequently asked questions.

Labster is hosted online, which means that students only have to login from their internet browsers once an account is created.

Labster is only available for purchase by faculty and administration at academic institutions. To procure Labster, simply reach out to us on our website. Schedule a demo, book a meeting to discuss pricing, start a free trial, or simply fill out our contact form.

Labster simulations are created by real scientists and designed with unparalleled interactivity. Unlike point and click competitors, Labster simulations immerse students and encourage mastery through active learning.

Labster supports a wide range of courses at the high school and university level across fields in biology, chemistry and physics. Some simulations mimic lab procedures with high fidelity to train foundational skills, while others are meant to bring theory to life through interactive scenarios.

Virus & Vaccines Mega Bundle (Includes COVID-19)

scientific method mystery virus case study simulation

  • Google Drive™ folder

Products in this Bundle (6)

showing 1 - 5 of 6 products

Description

This is the ultimate mega bundle that contains literally everything you need to teach your students all about viruses and vaccines, including COVID-19. Print & Digital versions are included so you can have students learn in a variety of learning modes:

  • Teacher Led Instruction with PPT/Google Slides while students complete fill-in notes
  • Student Focused Independent Learning via Peardeck slides
  • Engaging Simulated Virus Outbreak Case Study
  • Small Group Learning Activities
  • Large Group Learning Activities
  • Interactive Notebook Activities
  • Graphic Organizer Activities
  • Traditional Cut-n-Paste Activities
  • Engaging & Interactive WebQuests
  • CLOZE Reading Activities
  • Formative Assessments (Print & Google Form)
  • Summative Assessments (Print & Google Form)
  • Exit Tickets
  • and more...

Topics Covered

Virus structure

  • Bacteriophage
  • Lytic Cycle
  • Lysogenic Cycle

What a virus is

How viruses mutate

What a virus variant is

How mutations can allow viruses to evade the body's immune response

How mutations can allow viruses to enter host cells easier

Examples of Coronaviruses: SARS,MERS, COVID-19

Symptoms of a Coronavirus infection

Treatment of a Coronavirus infection

Prevention of a Coronavirus infection

How a human body cell is infected by a Coronavirus

  • Structure of a typical virus
  • Modern Cell Theory
  • Immune System Response
  • Helper T-Cells
  • Influenza Virus
  • Respiratory Sincitial Virus
  • Human Parainfluenza Virus
  • Adenoviruses
  • Coronaviruses
  • Severe Acute Respiratory Syndrome (SARS) Virus
  • Wuhan Novel Coronavirus (COVID-19)

Scientific Method

Epidemiologist

Obtaining Medical records to confirm illness symptoms

Mapping locations of confirmed cases

Establish timelines and travel histories

Developing Hypothesis for what is causing illnesses and how it is spreading

Viral or Bacterial Culture

Blood Cultures

Rapid PCR Test

Blood Test for Antibodies

Case-Definition

Case-Control Study

Longitudinal Study

Analytic Epidemiology

Vaccine Rapid Response Platforms

Moderna Nucleic Acid Vaccine Platform

Recombinant Protein

Messenger RNA

Growth Medium

Bioreactors

Viral Vaccines

Antibiotics

Chromatography

Ultrafiltration

Stabilizers

Preservatives

Freeze-dried Vaccines

Rehydration of Vaccines

What is a Corona Virus No-Prep Mini-Lesson

Updated to include coronavirus mutations/variant information. This complete No-Prep mini-lesson includes everything you need to teach your students all about the Corona virus including COVID-19!

Included in this complete No-Prep Lesson!

  • Teacher Lesson Guide
  • Interactive Notebook Vocabulary Cut and Paste Graphic Organizer
  • 10 Question MC Quiz
  • KWL Chart Activity
  • Google Slide presentation with student fill-in-the-blank notes
  • Google Slide Presentation - Peardeck Version (added 12/13/20)
  • CLOZE Reading Activity with color-coded Coronavirus Illustration
  • Optional 2nd CLOZE Reading Activity
  • Exit Ticket for a quick formative assessment
  • Answer Keys for all handouts and activities!

Google Slide Presentation with Student Fill-in-the-Blank Notes & Answer Key

Slide 2 - Explains what a Coronavirus is

Slide 3 - Explains the typical Coronavirus Structure

Slide 4 - Explains the life cycle of a Coronavirus

Slide 5 - Explains how viruses mutate and can evade antibodies/enter cells easier

Slide 6 - Explains the common symptoms of a Coronavirus infection

Slide 7 - Explains how the typical Coronavirus is spread

Slide 8 - Explains how to protect yourself from a Coronavirus infection

Slide 9 - Explains how to treat a Coronavirus infection

Includes Google Classroom version: included in your purchase are Google Doc & Google Slide versions of all of the above handouts, a google form CLOZE Reading Quiz, a google form Exit Ticket, and a google form 10 Question MC Quiz! All 3 of these assessments are auto-grading so you can import grades with the click of a mouse!

Mystery Virus Epidemiology Case Study

This highly engaging online simulation asks students to take on the role of a Public Health Director to determine the cause of a recent mystery deadly virus epidemic outbreak at a local High School and Senior Center.

In their role as Public Health Director, students are guided by an expert epidemiologist through the entire case study simulation which consists of:

  • Studying medical reports and determining which ones to follow-up on
  • Collecting information and identifying patterns of infection from patients
  • Using the Scientific Method, develop & test a hypothesis to explain the outbreak.
  • Evaluate epidemiological testing to determine what pathogen is causing the disease.
  • Use epidemiology methods to determine the source and how the pathogen spreads.
  • Make a recommendation on which patients should be isolated and quarantined.

Distance Learning: This simulation case study is appropriate for independent student work and should require minimal or no additional instructions from the teacher, The online simulation takes the student thru the entire process, step-by-step, with clear and concise directions from a virtual "expert epidemiologist".

All About mRNA Vaccines WebQuest

This web quest takes students on a fascinating journey to discover how mRNA Vaccines are made in months instead of years, and how they stimulate an immune system response to protect against sickness and infection.

This product includes a PDF version and a digital Google Slide Version that can be assigned individually, as a small or whole group activity or as an emergency substitute activity.

What is the job of a vaccine?

What are the 4 types of vaccines?

Measles Vaccine Production Process

What is the job of RNA?

What is DNA?

What is mRNA?

How do mRNA Vaccines produce an immune response?

mRNA Vaccine Development Process

mRNA Vaccine Safety

Genetic Sequencing of Viral RNA

Lipid Nanoparticles

Endocytosis

History of mRNA Vaccine Research

Potential Future Uses of mRNA Vaccines

How Vaccines are Made WebQuest

This highly engaging WebQuest takes students on a virtual journey to see first-hand how traditional vaccines and Rapid Response Platform (covid-19) vaccines are made. Each step in the process of making a vaccine is shown in a series of detailed and engaging animations. This activity can be done whole group or as independent work.

Distance Learning: This activity can be assigned to students via Google Classroom or other digital learning platform, a PDF file can be emailed to student, or a hard copy can be printed and given to students.

The entire process of making a traditional vaccine is covered from antigen production, release and isolation of the antigen, antigen purification, adjuvant addition, and finally packaging and distribution.

Covid-19 Vaccine Development Process - as of 4/8/20 there are 115 vaccine candidates in various stages of development. There are a wide range of platforms being used to develop vaccines including traditional methods and Rapid Response Platforms such as Moderna's Nucleic Acid Vaccine Platform which has been able to start clinical trials of a m-rna based vaccine in just 63 days instead of the six to ten years it takes for a traditional vaccine.

Covid-19 Vaccine Development Process

How Vaccines Work WebQuest

Highly engaging web quest takes students on a step-by-step journey through the immune system to discover how vaccines work. This web quest illustrates with engaging and colorful animations how 1) vaccines generate an immune system response to protect the body against infection 2) the immune system protects the body when a pathogen infects the human body.

After completing the web quest students can complete an online immune system response review activity. Both paper and google form quizzes are included!

Editable word, PDF, and digital versions included!

  • Antigen Presenting Cells
  • Active Killer T Cell
  • Memory Killer T Cell
  • Memory B Cell
  • Memory T Helper Cell
  • Naive B Cell
  • Naive Killer T Cell
  • Plasma B Cell
  • Vaccine Antigen
  • T Helper Cell
  • Vaccination
  • Disease Agents
  • Target Antigens
  • Attenuated Virus
  • Primary & Secondary Response

All About Viruses WebQuest

This webquest makes a great activity to introduce the topics of viruses and the body's immune response. It also is perfect for an emergency sub lesson! I've included a variety of engaging video and website resources from: Ask a Biologist, Cells Alive, Amoeba Sisters, CDC, and Live Science. - Answer Key Included!

This web quest consists of 4 pages. Students will enjoy watching engaging videos, using an interactive microscope simulation, and informative websites.

Topics Included

Tips for Customers

  • Click the Green ★ to follow our store and get notifications of freebies and new products!
  • Leave feedback to receive TpT credit for use on future purchases.
  • Questions? Please contact us in the Product Q&A Section.

Terms of Use

All rights reserved by STEM Printables

This product is intended for use by the original purchaser only.

Sharing this product with others, distributing via any means, or posting online is strictly prohibited.

Failure to comply is a copyright infringement and a violation of the Digital Millennium Copyright Act (DMCA). Clipart and elements found in this PDF are copyrighted and cannot be extracted and used outside of this file without permission or license.

Questions & Answers

Stem printables.

  • We're hiring
  • Help & FAQ
  • Privacy policy
  • Student privacy
  • Terms of service
  • Tell us what you think

COMMENTS

  1. Mystery Virus Epidemiology Case Study Simulation BIO 111

    Scientific Method: Mystery Virus Case Study Simulation Scenario: You are the Director of Public Health for a medium-sized city in the United States. There have been several reports of serious illnesses and three people have died. Your mission is to work with an epidemiologist to identify the cause of the illnesses before an epidemic breaks out.

  2. Mystery Virus: Epidemic Epidemiology Case Study Simulation

    This highly engaging online simulation asks students to take on the role of a Public Health Director to determine the cause of a recent mystery deadly virus epidemic outbreak at a local High School and Senior Center. In their role as Public Health Director, students are guided by an expert epidemiologist through the entire case study simulation ...

  3. Mystery Virus Epidemiology

    This highly engaging online simulation asks students to take on the role of a Public Health Director to determine the cause of a recent mystery killer virus epidemic outbreak at a local High School and Senior Center. In their role as Public Health Director, students are guided by an expert epidemiologist through the entire case study simulation which consists of: Studying medical reports and ...

  4. The Zombie Virus Pandemic: An Innovative Simulation Integrating

    Methods. This 2.5-hour simulation was conducted in 2018 and 2020 during students' virology course. Student teams collected and analyzed data to formulate hypotheses for the source pathogen. ... how to use epidemiological and clinical data to construct a differential diagnosis for the mystery virus; the molecular principles of viral genetic ...

  5. Mystery Virus Epidemiology Case Study Simulation.docx

    Scientific Method: Mystery Virus Case Study Simulation Scenario: You are the Director of Public Health for a medium-sized city in the United States. There have been several reports of serious illnesses and three people have died. Your mission is to work with an epidemiologist to identify the cause of the illnesses before an epidemic breaks out. ...

  6. A differential equations model-fitting analysis of COVID-19 ...

    Data and model fitting. In order to perform the present study, the data sets corresponding to daily infected, daily deceased, hospitalized and ICU patients in Portugal were obtained from the ...

  7. 1.37: Viruses and Viral Epidemic Simulation

    Introduction to Viruses. Viruses are noncellular parasitic entities that cannot be classified within any living kingdom. They can infect organisms as diverse as bacteria, plants, and animals. In fact, viruses exist in a sort of netherworld between a living organism and a nonliving entity.

  8. Mathematical epidemiologic and simulation modelling of first ...

    In Malaysia, several studies have been made by researchers such as Wong et al. 39, who modified the SIR model under vaccine intervention in several localities of Malaysia by using the simulation ...

  9. Mystery Virus Epidemiology Case Study Simulat.docx

    Scientific Method: Mystery Virus Case Study Simulation Scenario: You are the Director of Public Health for a medium-sized city in the United States. There have been several reports of serious illnesses and three people have died. Your mission is to work with an epidemiologist to identify the cause of the illnesses before an epidemic breaks out. ...

  10. 3D modelling and simulation of the dispersion of droplets and drops

    The case study we have developed herein corresponds to a public railway transport coach in which a passenger infected with the SARS-CoV-2 virus is seated and possibly contaminates other travellers ...

  11. A Review of Agent-Based Model Simulation for Covid 19 Spread

    Abstract. Containing the implications of the Covid-19 pandemic has and continues to be a priority around the globe. Pursuant to this, scientists have focused on understanding the virus spread's behavior and patterns to develop mitigation plans. Artificial Intelligence agent-based simulations (ABS) have been used by scientists to simulate ...

  12. Mystery Virus : Epidemiology Case Study Simulation

    Description. This highly engaging online simulation asks students to take on the role of a Public Health Director to determine the cause of a recent mystery killer virus epidemic outbreak at a local High School and Senior Center. In their role as Public Health Director, students are guided by an expert epidemiologist through the entire case ...

  13. Mystery Virus Epidemiology

    Distance Learning: This simulation case study is appropriate for independent student work and should require minimal or no additional instructions from the teacher, The online simulation takes the student thru the entire process, step-by-step, with clear and concise directions from a virtual "expert epidemiologist".

  14. Exam 1

    occurs when you change only one factor (variable) and keep all other conditions the same. Steps of the Scientific Method. Ask a Question. Do Background Research. Construct a Hypothesis. Test Your Hypothesis by Doing an Experiment. Analyze Your Data and Draw a Conclusion. Communicate Your Results. Results of Experiments can provide.

  15. 14.7: Activity

    Simulation of Viral Infection. The test performed in the lab is a simulation. No virus, living materials, or biohazardous reagents are used. The series of test tubes are each filled with a fluid that represents body fluid one might exchange with another individual. One tube of fluid is "positive". You cannot identify this cup by visual ...

  16. Practice the Scientific Method

    Practice the Scientific Method. You may not know it, but science is everywhere. You practice the scientific method every day as you move through the world, solving problems and learning things. Want to see for yourself how often you might use the scientific method? Play The Case of the Mystery Images. Then, learn more at Using the Scientific ...

  17. A comparison of five epidemiological models for transmission of SARS

    In this comparative paper, we describe five different models used to study the transmission dynamics of the SARS-Cov-2 virus in India. While simulation studies are the only gold standard way to compare the accuracy of the models, here we were uniquely poised to compare the projected case-counts against observed data on a test period.

  18. Exploring the Scientific Method: Cases and Questions, Gimbel

    From their grade school classrooms forward, students of science are encouraged to memorize and adhere to the "scientific method"—a model of inquiry consisting of five to seven neatly laid-out steps, often in the form of a flowchart. But walk into the office of a theoretical physicist or the laboratory of a biochemist and ask "Which step are you on?" and you will likely receive a ...

  19. Inefficiency of SIR models in forecasting COVID-19 epidemic: a case

    Using the epidemiologic data for the first 3 months, an SIR-based model was constructed to predict the disease course. In the SIR model, \(r\left( t \right) + i\left( t \right){ }\) determines all ...

  20. The Scientific Method

    The scientific method. Humans have always strived to explain natural phenomena. In the first mission of the experimental design simulation, you will learn how to use the scientific method to investigate phenomena, acquire new knowledge, or correct and integrate existing knowledge. Every tool you need to use is available in this virtual lab.

  21. Scientific Method Lab Quiz Flashcards

    Sparrows have wings, sparrows are birds. Inductive Reasoning. specific case to general principle. sparrows are birds and can fly, owls, flamingoes, eagles, and robins have wings. We can induce all birds can fly. Moorpark College, Biology 2A, Katherine Courtney Learn with flashcards, games, and more — for free.

  22. Scientific Method: Mystery Virus Case Study Simulation

    In the scientific method, a hypothesis is first proposed, then experiments are conducted to test the hypothesis, and finally the results of the experiments are analyzed to see if they support or disprove the hypothesis. In the Mystery Virus Case Study Simulation, the scientific method is used to investigate the outbreak of a new virus.

  23. Virus & Vaccines Mega Bundle (Includes COVID-19)

    This highly engaging online simulation asks students to take on the role of a Public Health Director to determine the cause of a recent mystery deadly virus epidemic outbreak at a local High School and Senior Center. In their role as Public Health Director, students are guided by an expert epidemiologist through the entire case study simulation ...