26
Available online at www.sciencedirect.com ScienceDirect Comput. Methods Appl. Mech. Engrg. 382 (2021) 113891 www.elsevier.com/locate/cma COVID-19 dynamics across the US: A deep learning study of human mobility and social behavior Mohamed Aziz Bhouri a , Francisco Sahli Costabal b , Hanwen Wang c , Kevin Linka d , Mathias Peirlinck d , Ellen Kuhl d , Paris Perdikaris a ,a Department of Mechanical Engineering and Applied Mechanics, University of Pennsylvania, Philadelphia, PA 19104, United States of America b Department of Mechanical and Metallurgical Engineering, Institute of Biological and Medical Engineering, Pontificia Universidad Católica de Chile, Chile c Applied Mathematics and Computational Science, University of Pennsylvania Philadelphia, PA 19104, United States of America d Department of Mechanical Engineering, Stanford University, Stanford, CA 94305, United States of America Received 21 September 2020; received in revised form 18 January 2021; accepted 24 April 2021 Available online 1 May 2021 Abstract This paper presents a deep learning framework for epidemiology system identification from noisy and sparse observations with quantified uncertainty. The proposed approach employs an ensemble of deep neural networks to infer the time-dependent reproduction number of an infectious disease by formulating a tensor-based multi-step loss function that allows us to efficiently calibrate the model on multiple observed trajectories. The method is applied to a mobility and social behavior-based SEIR model of COVID-19 spread. The model is trained on Google and Unacast mobility data spanning a period of 66 days, and is able to yield accurate future forecasts of COVID-19 spread in 203 US counties within a time-window of 15 days. Interestingly, a sensitivity analysis that assesses the importance of different mobility and social behavior parameters reveals that attendance of close places, including workplaces, residential, and retail and recreational locations, has the largest impact on the effective reproduction number. The model enables us to rapidly probe and quantify the effects of government interventions, such as lock-down and re-opening strategies. Taken together, the proposed framework provides a robust workflow for data-driven epidemiology model discovery under uncertainty and produces probabilistic forecasts for the evolution of a pandemic that can judiciously provide information for policy and decision making. All codes and data accompanying this manuscript are available at https://github.com/PredictiveIntelligenceLab/DeepCOVID19. c 2021 Elsevier B.V. All rights reserved. Keywords: Epidemiology model discovery; Sensitivity analysis; Neural networks; Scientific machine learning; Dynamical systems; Uncertainty quantification Corresponding author. E-mail addresses: [email protected] (M.A. Bhouri), [email protected] (F.S. Costabal), [email protected] (H. Wang), [email protected] (K. Linka), [email protected] (M. Peirlinck), [email protected] (E. Kuhl), [email protected] (P. Perdikaris). https://doi.org/10.1016/j.cma.2021.113891 0045-7825/ c 2021 Elsevier B.V. All rights reserved.

COVID-19 dynamics across the US: A deep learning study of

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: COVID-19 dynamics across the US: A deep learning study of

Available online at www.sciencedirect.com

wrcmaaorleja⃝

Kq

k(

h0

ScienceDirect

Comput. Methods Appl. Mech. Engrg. 382 (2021) 113891www.elsevier.com/locate/cma

COVID-19 dynamics across the US: A deep learning study of humanmobility and social behavior

Mohamed Aziz Bhouria, Francisco Sahli Costabalb, Hanwen Wangc, Kevin Linkad,Mathias Peirlinckd, Ellen Kuhld, Paris Perdikarisa,∗

a Department of Mechanical Engineering and Applied Mechanics, University of Pennsylvania, Philadelphia, PA 19104, United States of Americab Department of Mechanical and Metallurgical Engineering, Institute of Biological and Medical Engineering, Pontificia Universidad Católica de

Chile, Chilec Applied Mathematics and Computational Science, University of Pennsylvania Philadelphia, PA 19104, United States of America

d Department of Mechanical Engineering, Stanford University, Stanford, CA 94305, United States of America

Received 21 September 2020; received in revised form 18 January 2021; accepted 24 April 2021Available online 1 May 2021

Abstract

This paper presents a deep learning framework for epidemiology system identification from noisy and sparse observationsith quantified uncertainty. The proposed approach employs an ensemble of deep neural networks to infer the time-dependent

eproduction number of an infectious disease by formulating a tensor-based multi-step loss function that allows us to efficientlyalibrate the model on multiple observed trajectories. The method is applied to a mobility and social behavior-based SEIRodel of COVID-19 spread. The model is trained on Google and Unacast mobility data spanning a period of 66 days, and is

ble to yield accurate future forecasts of COVID-19 spread in 203 US counties within a time-window of 15 days. Interestingly,sensitivity analysis that assesses the importance of different mobility and social behavior parameters reveals that attendance

f close places, including workplaces, residential, and retail and recreational locations, has the largest impact on the effectiveeproduction number. The model enables us to rapidly probe and quantify the effects of government interventions, such asock-down and re-opening strategies. Taken together, the proposed framework provides a robust workflow for data-drivenpidemiology model discovery under uncertainty and produces probabilistic forecasts for the evolution of a pandemic that canudiciously provide information for policy and decision making. All codes and data accompanying this manuscript are availablet https://github.com/PredictiveIntelligenceLab/DeepCOVID19.c 2021 Elsevier B.V. All rights reserved.

eywords: Epidemiology model discovery; Sensitivity analysis; Neural networks; Scientific machine learning; Dynamical systems; Uncertaintyuantification

∗ Corresponding author.E-mail addresses: [email protected] (M.A. Bhouri), [email protected] (F.S. Costabal), [email protected] (H. Wang),

[email protected] (K. Linka), [email protected] (M. Peirlinck), [email protected] (E. Kuhl), [email protected]. Perdikaris).

ttps://doi.org/10.1016/j.cma.2021.113891045-7825/ c⃝ 2021 Elsevier B.V. All rights reserved.

Page 2: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

(dkdhcfccicstaep

fbcosad

Cbaac

1

afHmmsdDducc

a

1. Introduction

1.1. Background

Since its emergence in late December 2020 [1], as of early September 2020, the novel SARS-CoV-2 virusCOVID-19) has resulted in over 81 million confirmed cases globally, of which nearly 2 million people haveied [2]. After nearly months, active cases still surge in many countries and regions [2]. To date, there are nonown preventive pharmaceutical interventions [3], and numerous behavioral intervention policies, such as lock-own, mandatory self-isolation and facial covering orders [4] have been issued to reduce the transmission of such aighly infectious disease to an extent that current health and financial systems could sustain. Without these timelyountermeasures, breakdown of these systems will lead to more severe social and economic crisis in an unexpectedlyast pace, especially given the exponential rise in cases number at the beginning of the outbreak as it was thease in the US, for instance for the first few months of the COVID-19 spread [5]. Despite their effectiveness inontaining the aggressive virus spread, these attempts have also interrupted ordinary living orders, and have resultedn heavy socio-economical impacts such as exploding unemployment rates, prevailing emotional distress, and theonsequently exacerbated civil unrest [6] that has evidently encouraged super spreader events. Phased re-openingtrategy [7], as adopted by many states and cities as a compromise, seemingly becomes the only way to leveragehe casualties brought by the disease itself, and by the consequential economic interruption. Therefore, accuratelynd quantitatively assessing the cost and benefit of restricted opening gains an increasing importance nowadays,specially at a time when America saw spikes after spikes in tested positive population after some controversialolicies.

Studies [8,9] have revealed that the transmission of COVID-19 likely comes from direct contact, such as dropletsrom infected persons, or indirect contact, where virus that remained active on shared surface is brought into humanody by behaviors such as whipping face with hands [10]. The virus replicates itself in the throat, and furtherauses lung lesions. The incubation period, before human immune responses begin to intervene, could last 6 daysn average or even as long as 14 days observed [11]. During or even after the incubation period, the patients cantill remain largely asymptomatic, while being able to infect others [12–14]. The corresponding antigen detectionssays using throat wash and saliva are also reported to produce false result [15], affecting the reliability of the caseata resource.

Such pathology makes clear that close contact interactions in public space is a major cause of the spread ofOVID-19. Individual level transmission models [16] have shown that 80% of secondary transmissions may haveeen caused by a small fraction of infectious individuals (∼ 10%) from some super-spreading events, mainly largend dense gatherings. The aforementioned two challenges, long incubation period and noisy data, call for timelydjustments to existing policy plans based on robust assessments of the situation. Under such motivation, technologyompanies such as Google [17] and Apple [18] also make publicly available their privacy preserved databases.

.2. Related work

Multi-compartment models, such as SEIR and SIR [19], have been used extensively in the infectious diseasenalysis and prediction. Over the years, multi-compartment models have been also been augmented to accountor more specific population groups that are subject to different transmission rates, recovery rates, etc. [20].owever, little attention has been received towards extending multi-compartment models to account for publicobility and behavioral data that can reflect non-pharmaceutical interventions, such as government policies, dailyigration and interaction patterns, that could also contain a rich information about the disease transmission. Recent

tudies have attempted to account for spatial correlations to model the transfer of infected population betweenifferent geographic regions by utilizing human travel data from transportation and customs authorities [5,21,22].espite their novelty, such models also bring in questionable assumptions. For example, some highly culturallyirected migration may not resemble a diffusion process described by a partial differential equation model, and thenrecorded traveling by foot or by personal vehicles could bring unique biases into the error of the travel data. Inonjunction with the aforementioned works, time- and mobility-dependent parameters are also being introduced inompartment models to enable the incorporation of more diverse data sources [20].

As multi-compartment models become more sophisticated to reflect the complexity of the COVID-19 pandemic,

need for fast and reliable parameter estimation methods arises. Classical approaches to data assimilation, including

2

Page 3: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

gav

Bayesian methods [23–25] and Kalman filters [26], offer computational infrastructure for fitting multi-compartmentmodels to observational data. However, their use can be often hindered by prerequisites for model construction,such as linearity [27], and rapidly increasing computational complexity as the number of model parameters to beinferred and the size and diversity of the training data-sets increase. Therefore, there is a need for robust and scalablecomputational frameworks for epidemiology model calibration from noisy and incomplete time-series data.

1.3. Contributions of this work

Here we conduct a large-scale computational study that leverages Google’s mobility trend data [17] aroundrocery and pharmacy stores, parks, public transportation, retail and recreation spaces, residential and workplaces,long with the Unacast’s social behavior data [28] which includes information on the average distance traveled,isits to non-essential retail and services, and unique human encounters per km2 relative to national metrics pre-

COVID-19. We believe that these types of mobility and social behavior trends cover most scenarios of social mixingand can be used as predictive indicators of viral spread [29].

From a modeling perspective, the main novelty of this work stems from utilizing mobility and social behaviortrends data to model the temporal evolution of the COVID-19 infectious rate — the probability of transmittingdisease between a susceptible and an infectious individual (Fig. 1). From a methodological standpoint, this workgeneralizes the multi-step neural network method of Raissi et al. [30] to enable gradient-based optimization ofunknown epidemiology model parameters with latent variables. This is achieved by effectively back-propagating theinfluence of the unobserved variables through the disease spread dynamics with an appropriate tensorization of theloss function. This key step enables a computationally efficient treatment of multiple temporal trajectories, as wellas the concurrent training of multiple neural networks. The latter, allows us to generate a statistical ensemble overthe predicted evolution of the underlying disease spread dynamics, which enables future forecasting with quantifieduncertainty.

In summary, the specific contributions of this work can be organized as follows:

• We extend the multi-step neural network method of Raissi et al. [30] to enable dynamical system’s parametersinference with unobserved latent variables. Moreover, we tensorize the loss function formulation in orderto perform the parameters inference from multiple trajectories of different counties, while having a smallmemory footprint in terms of the size of the associated computational graph. Hence, the proposed approachis computationally efficient and enables training of multiple neural networks, which provides an uncertaintyquantification estimate of the dynamical system evolution via multiple trajectory forecasts.

• We model the effective reproduction number with a long short-term memory neural network (LSTM) to captureits temporal dependence on the mobility and social behavior variables. The LSTM architecture enables to accountfor the time lag between the change in mobility and social behavior, and the effect on the spread of the disease.

• We demonstrate the effectiveness of the proposed method by applying it to the COVID-19 spread in 203 UScounties such that the training is performed using data for a period of 66 days, and we produce reliable futureforecasts with quantified uncertainty for a period of 15 days. The lock-down efficiency is quantified by thesubsequent forecasted decrease of the effective reproduction number. We also perform a sensitivity analysisshowing that among the mobility parameters, the attendance of closed places has the highest influence on theevolution of the effective reproduction number, unlike open or transit locations. Moreover, all social behaviorparameters have relatively important sensitivities, highlighting the importance of social interactions on the spreadof the virus.

Our findings put forth a novel, flexible and robust workflow for data-driven epidemiology model discovery thatcan potentially help health authorities take a more quantitative and judicious approach to policy making in orderto slow down COVID-19 propagation. We demonstrate that well-calibrated epidemiology models can indeed returnsensible future forecasts. Such capability can directly assist with optimizing resource allocation for more precisediagnostics and treatment, as well as yield a better understanding of the disease spread dynamics. Taken together,this can be another tool in our arsenal to help us better prepare for COVID-19 and potential future pandemics, andprovide additional quantitative tools to help us minimize their spread.

This paper is organized as follows. Section 2.1 presents the mobility and social behavior-based SEIR epidemi-

ology model considered to study the spread of COVID-19. Section 2.2 details the proposed computational method

3

Page 4: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

wwiaTdHssrpaAotct

Fig. 1. Our goal is to develop a robust computational framework for calibrating COVID-19 models from noisy time-series data. This willenable the quantitative assessment of government policies (e.g., travel bans, quarantine, contact tracing, etc.), and yield reliable forecastswith quantified uncertainty that can help us manage and contain an epidemic.

for inference of the mobility and social behavior-based SEIR model parameters and the corresponding technicalingredients. In Section 3, the effectiveness of the proposed method is demonstrated by testing it on real mobilityand infections data from 203 US counties and obtaining reliable future forecasts for a period of 15 days withquantified uncertainty. Within the same section, we present a sensitivity analysis to assess the influence of thedifferent mobility and social behavior parameters considered on the spread of the disease. Finally, in Section 4 wesummarize our key findings, discuss the limitations of the proposed approach, and carve out directions for futureinvestigation.

2. Methods

2.1. Epidemiology model

We consider a mobility and social behavior-based SEIR model detailed as follows:dCdt

=d Idt

+ γ I

d Sdt

= −β(m(t)) S I

d Edt

= β(m(t)) S I − δ E

d Idt

= δ E − γ I

d Rdt

= γ I,

(1)

here S is the fraction of susceptible population, E the fraction of exposed population including individualsho have been infected but are not yet infectious themselves during the incubation period, I the fraction of

nfectious population, R the fraction of removed population that accounts for recovered and deceased individuals,nd C(t) ≡ R(t) + I (t) is the fraction of cumulative cases, which is the only variable with reliable data over time.he only available data-sets at the early stages of the COVID-19 pandemic consist of the cumulative cases C(t), theeaths and the recoveries. In the SEIR model, the removed population R accounts for the deaths and the recoveries.owever, the recoveries are never accurately measured since, for the positively diagnosed patients, only those in

erious or critical condition are closely followed by healthcare providers. Hence, we can accurately determine whenuch patients move from the I (t) to the R(t) compartment. However, such patients in serious or critical conditionepresent a very small fraction of the infected population I (t) (for instance, they only represent around 0.4% of theositively diagnosed patients these days https://www.worldometers.info/coronavirus/), while most of these patientsre asked to stay at home and then are tested after two weeks or so in order to verify if they are still infectious or not.s a consequence, the exact day of recovery for most of the infected people is not accurately known, which leaves usnly the cumulative cases as the most reliable indicator. The same reasoning applies with the disparity in reportinghe infected population in urban and rural areas, either due to undetected or asymptomatic patients. Such populationan have an influence on the model behavior when the case numbers are relatively large. However, estimations of

he undetected population based on antibody studies are currently available only at very few locations and only

4

Page 5: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

i

m

d

2

Fig. 2. Mobility and social behavior parameters for three counties randomly sampled among the data-set considered.

in urban areas [31]. For this reason, we decided to not explicitly model this population, as we would be simplyguessing its size. Note that we consider an SEIR model with variables normalized by the number of populationsuch that the resulting model (1) is universal for all regions and independent of the number of their inhabitants.

Since the spread of an infectious disease that is transmitted from human to human depends on interpersonalnteractions, we model the susceptible to infectious transition rate β in Eq. (1) as a function of mobility and

social behavior parameters denoted as m ∈ RNm , which contains the mobility and social behavior parametersm1(t), . . . , m Nm (t). This enables determining the evolution of the effective reproduction number, R0(t) ≡

β(m(t))γ

, akey parameter to characterize the evolution of an epidemic.

The proposed model is trained on data provided by Google for mobility change from baseline [17] (parametersm1, . . . , m6 as discussed below), along with the data provided by Unacast [28] for social behavior trends (parameters

7, . . . , m9 as detailed below) for all 203 US counties considered in this work. Specifically, we consider a total ofNm = 9 mobility and social behavior parameters corresponding to:

• m1: retail and recreation percent change from baseline,• m2: grocery and pharmacy percent change from baseline,• m3: parks percent change from baseline,• m4: transit stations percent change from baseline,• m5: workplaces percent change from baseline,• m6: residential percent change from baseline,• m7: average distance traveled,• m8: visits to non-essential retail and services,• m9: unique human encounters per Km2 relative to national pre-COVID-19.

Fig. 2 provides the observed mobility and social behavior trends for three counties randomly selected among theata-set considered, showing the inherent presence of noise within the trajectories.

.2. A deep learning approach to model calibration given partially observed states

In this work, γ and δ are assumed as given and equal to 1/6.5 [32,33] and 1/2.5 [11,32] respectively forCOVID-19. Additional numerical studies that aim to elucidate the model’s sensitivity on γ and δ have beenconducted. Our findings indicate that the specific values chosen for γ and δ do not have a significant impact on

the model’s predictions as shown in Appendix C). One of our main goals is to learn the dynamics of β(t), and,

5

Page 6: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

foa

wom

wsmm

sa

sado

gp

avtd

w

dti

consequently, the evolution of the solution to the SEIR model of Eq. (1). The learning task can then be formulatedas follows.

Given γ , δ and a data-set D for Nc different regions containing: (i) the initial conditions of the variable I (t = t0)or each region, (ii) the time trajectories of the observed variable C(t) for each region, and (iii) the time trajectoriesf the observed mobility and social behavior trends m(t) for each region, infer the dependence of β(t) ≡ β(m(t))nd the initial fractions of the exposed population E(t = t0) for all the different regions.

The observed data-set D can be expressed as follows:

D ≡ {I 1(t0), . . . , I Nc (t0), C1(t0), . . . , C1(tNt −1), . . . , C Nc (t0), . . . , C Nc (tNt −1),

m1(t−τ ), . . . , m1(tNt −2), . . . , mNc (t−τ ), . . . , mNc (tNt −2)},(2)

here superscripts refer to the different regions considered. Nt refers to the number of days for which the numberf cumulative cases is available. Those time instances are represented with t0, t1, . . . , tNt −1. In order to learn theapping β ≡ β(m(t)), β is modeled with a long-short term memory (LSTM) [34] neural network of the form

β = f (m, θ N N ), (3)

here θ N N refers to the weights and bias parameters defining the LSTM network. Since we are modeling thepread of an infectious disease with an incubation period, we utilize the lag structure of the LSTM network toodel β as a function of the mobility and social behavior parameters over a period of time of τ days. Hence, weodel the susceptible to infectious transition rate β(ti ) as a function of the parameters m on the τ previous days:

m(ti−τ ), m(ti−τ+1), . . . , m(ti−1). As a consequence, to model the cumulative cases from t0 to tNt −1, and hence theusceptible to infectious transition rate β from t0 to tNt −1, we need to consider the parameters m from t−τ to tNt −2s given in Eq. (2). Note that due to the incubation period and the testing delay, we considered that the parameters

m at a given day ti could not influence the propagation of the disease that is measured and reported on the exactame day ti , which justifies the one day shift in the ending day of the ranges considered for the cumulative casesnd the parameters in the data-set D needed to train the model. The LSTM takes care of capturing the temporalependence of β on the mobility and social behavior parameters m within the τ previous days. The forward passf LSTM neural network considered in this work is detailed in Appendix A.

Given the available dataset D, we formulate a multi-step loss function in order to infer the learnable parameters θ

iven in Eq. (4), which include both the LSTM parameters θ N N as well as the latent initial fractions of the exposedopulation for all Nc different regions considered,

θ ≡ (θ N N , E1(t0), . . . , E Nc (t0)). (4)

To formulate the multi-step loss function, we follow the approach put forth by Raissi et al. [30], and considern appropriate temporal discretization of Eq. (1). Given the available data containing the initial conditions of theariable I (t = t0) and the trajectories of the variable C(t), the variable I (t) can be estimated for the differentime instances t1, . . . , tNt −1 using the first equation of the SEIR model (1). For instance, using the trapezoidaliscretization rule [30,35], the first equation of the SEIR model (1) for region j gives:

I j (t0) = C j (t0) ,

[F + T ] I j= F C j

+

⎡⎢⎢⎢⎣−γ C j (t0)

20...

0

⎤⎥⎥⎥⎦,(5)

here

F =1dt

⎡⎢⎢⎢⎢⎣1

−1. . .

. . .. . .

− 1 1

⎤⎥⎥⎥⎥⎦ ∈ RNt −1×Nt −1 , T =

⎡⎢⎢⎢⎢⎣1/2

1/2. . .

. . .. . .

1/2 1/2

⎤⎥⎥⎥⎥⎦ ∈ RNt −1×Nt −1 , (6)

t is the time-step size corresponding to the period of the infections number update, and taken equal to 1 day inhis work based on the available data, and C j , I j

∈ RNt −1 are the vectors containing C j (t) and I j (t) at the time

nstances t1, . . . , tNt −1, respectively.

6

Page 7: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

uf

w

β

R

tf

Then, given the estimated I (t) from Eq. (5) and the parameters θ , we can compute S(t), then E(t) and finally I (t)sing the remaining equations of the SEIR model (1). Using the trapezoidal rule for instance, such discretizationor region j gives:

S j (t0) = 1 − E j (t0) − I j (t0),

[F + B] Sj=

⎡⎢⎢⎢⎢⎢⎣1−E j (t0)−C j (t0)

dt −β0(m j ,θ N N ) (1−E j (t0)−C j (t0)) C j (t0)

2

0...

0

⎤⎥⎥⎥⎥⎥⎦ ,

E j (t0) = E j (t0),

[F + δ T ] Ej= B S

j+

⎡⎢⎢⎢⎢⎢⎣E j (t0)

dt +β0(m j ,θ N N ) (1−E j (t0)−C j (t0)) C j (t0)

2 − δE j (t0)

2

0...

0

⎤⎥⎥⎥⎥⎥⎦ ,

I j (t0) = I j (t0),

[F + γ T ] Ij= δ T E

j+

⎡⎢⎢⎢⎢⎢⎣C j (t0)

dt + δE j (t0)

2 − γC j (t0)

2

0...

0

⎤⎥⎥⎥⎥⎥⎦ ,

(7)

here

B =

⎡⎢⎢⎢⎢⎢⎢⎢⎣

βj1 I j (t1)

2

βj1 I j (t1)

2

. . .

. . .. . .

βjNt −2 I j (tNt −2)

2

βjNt −1 I j (tNt −1)

2

⎤⎥⎥⎥⎥⎥⎥⎥⎦∈ RNt −1×Nt −1 , (8)

ˆ ji ≡ β(m j (ti−τ ), . . . , m j (ti−1), θ N N ) is the LSTM prediction of β for region j at time instance ti , and S

j, E

j, I

j∈

Nt −1 are the vectors containing S j (t), E j (t) and I j (t) at the time instances t1, . . . , tNt −1, respectively. Note thatE j (t0) appearing in the initial conditions in Eqs. (7) is part of the learnable parameters θ , as described in Eq. (4).

Notice that, given the dataset D and the parameters θ , one can estimate S(t), E(t), and I (t) in that order. Finally,he loss function for a single neural network is expressed as the difference between the variable I (t) estimated onlyrom the available data and the estimated variable I (t) which depends on the learnable parameters:

L(θ ) =1

(Nt − 1) × Nc

Nt −1∑i=1

Nc∑j=1

[I j (ti ;D) − I j (ti ; θ )]2 . (9)

One crucial point in terms of implementation consists of the possibility of expressing the finite difference equationof (5), and each of the finite difference equations of (7), as tensors of shape Nc × (Nt − 1). As a consequence, theloss L given in Eq. (9) can be expressed as the sum of the elements of a tensor of shape Nc × (Nt − 1) ensuringa minimalist computational graph, which speeds up gradient back-propagation. We emphasize that the proposedmethodology is independent of the chosen finite difference scheme, and the trapezoidal rule is chosen here for sakeof presentation and in coherence with the finite difference scheme adopted in this work as justified in Appendix A.This computational strategy allows us to consider multiple trajectories simultaneously (accounting for the differentregions considered in the data), instead of performing the learning task using observations from a single or a handful

of trajectories. Finally, the loss can be further tensorized to enable the parallel training of multiple neural networks.

7

Page 8: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

iht

3

dld

Tib4

3

dW1cedUpβ

ipr

Table 1Hyper-parameter settings employed across all numerical studies.

θ N N initialization Xavier initializationE1(t0), . . . , E Nc (t0) initialization Uniform initializationOptimization algorithm AdamLearning rate 10−3

Number of iterations 2 × 104

County batch size 60

By training an ensemble of independent networks, we can obtain a statistical characterization of model uncertaintyand quantify its effect on the resulting model predictions and forecasts. For more technical details, the interestedreader is referred to our publicly available implementation.

3. Results

All code and data presented in this section will be made publicly available at https://github.com/PredictiveIntellgenceLab/DeepCOVID19. The minimization of Eq. (9) is carried out using stochastic Adam updates [36] with theyper-parameters summarized in Table 1. Based on our experience, the reported results are robust with respect tohe numerical values chosen for the different optimization hyper-parameters.

.1. Training data and pre-processing

We remove the counties containing missing mobility or social behavior parameters for more than 3 consecutiveays or 20 days in total over the considered time period, and we use linear interpolation for missing data for 3 oress consecutive days. We also consider the USAFacts processed infections data for all US counties [37]. The twoatasets are cross-matched together such that the considered counties satisfy the following criteria:

• the infection rate C(t) is at least equal to 0.2% after t = 81 days of the spread of the disease,• the mobility and social behavior data for the τ days prior to the first day of infections is available.

For each county, the first day of infections is defined as the first day for which there are at least 10 infections.hese criteria provide a dataset with 203 US counties and a period of time of 81 days, such that the first day of

nfections for each county is between March 16, 2020 and March 24, 2020, while the last day of considered isetween June 4, 2020 and June 12, 2020. The total population of these counties is 150 million, which represents6% of the total US population.

.2. Predicted forecasts

The proposed framework is trained on the collected data corresponding to the first 66 days. For the remaining 15ays, only the mobility and social behavior data are used in order to extrapolate the fraction of infected population.1

e train an ensemble of 100 models with randomized parameter initializations. The total training time takes4.7 min on a single NVIDIA Tesla P100 GPU. Once trained, the model provides estimates of the cumulativeases fraction with a relative error equal to 4.3% on the training data (for the first 66 days), and a relative errorqual to 6.6% on the testing data, corresponding to the extrapolation carried out for the remaining 15 days of theata-set. Fig. 3 shows the distribution of relative errors averaged over the 15 days extrapolation window across allS counties considered. Evidently, the proposed model can yield accurate future forecasts of COVID-19 spread,articularly for counties having a higher fraction of cumulative cases. However, using an SEIR model with a constantover time for each county fails to explain the spread of COVID-19 for the 203 US counties considered. Fig. B.13

n Appendix B compares the relative forecasting errors averaged over the extrapolation period when using theroposed multi-step LSTM approach and a multi-step approach with a constant susceptible to infectious transition

ate β. Based on these results we conclude that enhancing the expressiveness of the SEIR model via the inclusion

8

Page 9: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

g

ofcwcrce1sfSiNhrs

dpNp

3

s

Fig. 3. Relative forecasting errors averaged over the 15 days extrapolation period in all 203 US counties considered in this work.

of mobility and social behavior-based effects in the time-depended modeling of β has a significant impact on theeneralization accuracy of the model, leading to more accurate future forecasts.

Fig. 4 provides the dependency of the relative error averaged over the 15 days extrapolation on the fractionf cumulative cases on the last day. The latter shows a better accuracy for counties with higher cumulative casesractions, which can be explained by the loss function considered (9). Indeed, counties with higher fraction ofumulative cases present a higher fraction of the infectious population. As a consequence, the corresponding termsithin the loss function are generally higher than the ones related to counties with a lower fraction of cumulative

ases. Hence the learning of the model is inherently more driven towards minimizing the errors of the estimateselated to counties with higher fraction of cumulative cases, which explains the lower errors observed compared toounties with lower fraction of cumulative cases. One possibility to remedy to this issue is to consider normalizedrrors in the loss function (9) by dividing the L2 errors [I j (ti ) − I j (ti )]2 by the values of I j (ti ), 1 ≤ i ≤ Nt − 1,≤ j ≤ Nc added with a zero or positive hyper-parameter ϵL that needs to be tuned in order to prevent significantly

mall values in the denominator. However, such approach drove the model to better approximate counties with lowerraction of cumulative cases due to the importance weight introduced by the corresponding small denominator I j (ti ).ince accurately modeling counties with higher disease propagation, and hence higher fraction of cumulative cases,

s more crucial than focusing on counties with lower infection rates, we opted for the non-normalized loss function.ote that considering the normalized loss partially remedied the disparity in accuracy for the different counties,owever the improvement observed was not significant regarding the prediction and extrapolation errors and itequired a fine tuning of the hyper-parameter ϵL , which would make the approach a problem-dependent methodince ϵL would differ depending on the data-set considered.

New York state has the highest number of counties with a fraction of infected population above 1% after 81ays of the spread of the virus. Fig. 5 shows the predictions obtained for the evolution of the fraction of infectedopulation for the major New York counties. Figs. 6 and 7 provide additional sample results for other US counties.ew Jersey and Massachusetts present the second and third highest number of counties with a fraction of infectedopulation above 1% after 81 days of the spread of the virus.

.3. Sensitivity analysis

In order to assess the influence of different mobility and social behavior parameters on the evolution of theusceptible to infectious transition rate β, a sensitivity analysis using the Morris method [38] is performed on the

1 Predictions for a longer window of 21 days are also provided in Appendix D.

9

Page 10: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

imm

m—swppnmRv

psoβ

sso

3

difdifoV

Fig. 4. Variation of model accuracy as a function of the cumulative cases fraction.

trained LSTM neural network predicting β = f (m, θ N N ). Note that the LSTM has Nm × τ inputs, such that β(tk)nference is carried out using m(tk−τ ), . . . , m(tk−1), where we remind the reader that m(t) ∈ RNm contains the

obility and social behavior parameters m1(t), . . . , m Nm (t). We use si, j to denote the modified mean of the Morrisethod sensitivity measure [39] of β(tk) to mi (tk− j ).Fig. 8 presents a heat-map of the normalized sensitivity measures si, j . As expected, in view of the Google

obility parameters (m1, . . . , m6), these results suggest that the sensitivity of β to parks attendance (parameter m3)the only open place that is considered among the six available mobility parameters — is the lowest amongst all

ix. The second least influential parameter is the transit station’s attendance (parameter m4), reflective of a locationhere people spend on average the least amount of time among the places considered in the remaining five mobilityarameters (except parameter m3). The numerical values obtained are also consistent with the COVID-19 incubationeriod estimated to be between 2 and 21 days by the United States’ CDC [40]. Indeed, the mobility parameters doot instantaneously affect β and a lag period should be taken into account in order to observe the effects of theobility parameters on the spread of the disease, which corresponds to the incubation period and the testing delay.egarding the Unacast social behavior parameters, all parameters display a significant sensitivity, with unnecessaryisits displaying the highest value as shown in Fig. 8.

We also perform a local sensitivity analysis as follows. We increase each of the mobility and social behaviorarameters by 10% for the last 21 days (each at a time), and we determine the corresponding prediction for theusceptible to infectious transition rate β. We then compare the obtained value for β for the last day to the actualne obtained with the unchanged data. Finally, we pick the parameter that corresponds to the highest change in. This result, shown in Fig. 9, allows us to locally identify the most sensitive parameter among the mobility andocial behavior ones considered for each of the 203 counties studied. Interestingly, our results suggest that thepread of COVID-19 among a population of over 60 million people is most sensitive to the closure and reopeningf workplaces, followed by changes in visits to residential places.

.4. Effects of lock-down

In Figs. 5–7 we indicated the day on which the ≥ 50 gatherings ban was issued along with the stay at home orderay (except for Navajo, Arizona for which the ≥ 50 gatherings ban was issued at day t = −4). The effect of thesenterventions on the evolution of the effective reproduction number R0(t) is significant, as the latter generally dropsollowing at least one of these two policy decisions, such that R0(t) always reaches values around 1 while theseecisions are still in place. We should note that banning ≥ 50 gatherings was sufficiently effective to decrease R0n some locations, including most of New York counties; Massachusetts counties; and Prince George’s, Maryland,or instance. However, for other counties, the drop of R0 was significant only after issuing the stay at home order asbserved for Suffolk, New York; Middlesex, New Jersey; Cook, Illinois; Jefferson, Louisiana; and Prince Williams,irginia. These observations quantify the importance of limiting gatherings in managing COVID-19 spread and the

10

Page 11: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

Fig. 5. Evolution of the fraction of cumulative cases C(t) and the effective reproduction number R0(t) across different New York statecounties. The orange and green vertical dashed lines indicate the dates corresponding to the large crowd gathering ban and stay at homeorder issued by each county. The pink vertical dashed line delineates the data used for training the model (days 0–65), and the predictedmodel forecasts (days 66–80).

11

Page 12: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

Fig. 6. Evolution of the fraction of cumulative cases C(t) and the effective reproduction number R0(t) for representative counties across theUS. The orange and green vertical dashed lines indicate the dates corresponding to the large crowd gathering ban and stay at home orderissued by each county. The pink vertical dashed line delineates the data used for training the model (days 0–65), and the predicted modelforecasts (days 66–80).

12

Page 13: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

Fig. 7. Evolution of the fraction of cumulative cases C(t) and the effective reproduction number R0(t) for representative counties across theUS. The orange and green vertical dashed lines indicate the dates corresponding to the large crowd gathering ban and stay at home orderissued by each county. The pink vertical dashed line delineates the data used for training the model (days 0–65), and the predicted modelforecasts (days 66–80).

necessity of issuing the stay at home orders in locations where gatherings bans alone are not enough to manage thespreading.

3.5. Effects of re-opening

To quantify the effect of re-opening, the retail and recreation mobility parameter of New York county (New York)is artificially altered as shown in top left plot of Fig. 10 (while the other mobility and social behavior parameters

13

Page 14: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

(

ae2ir2sda

utmtF

aiu

s

Fig. 8. si, j : Sensitivity of susceptible to infectious transition rate β to the mobility and social behavior parameters.

Fig. 9. Mobility and social behavior parameters with the greatest effect on the reproduction number, by county (left panel) and by populationright panel).

re kept the same). As shown in top right and bottom left plots of Fig. 10, such policy has a significant effect on thevolution of β(t) and by consequence on the effective reproduction number R0(t) increased beyond 1 for the last0 days of the simulation. As a result, the number of cumulative cases C(t) would significantly increase as shownn the bottom right plot of Fig. 10 and the spread of the disease would be accelerated due to the increase of theetail and recreation mobility parameter. We note that the reproduction number is significantly affected only after0 days of increasing the mobility parameter. This result is in line with the sensitivity analysis in Fig. 8, whichhows that the reproduction number is mostly affected by mobility between 17 and 21 days earlier. In practice, thiselay poses a challenge to policy makers, which can only evaluate the effects of easing and imposing restrictionsfter two to three weeks.

Moreover, the altered trajectories show a higher uncertainty than the baselines in Fig. 10. The predictivencertainty reported by the proposed model reflects the uncertainty in the sparse and noisy training data, as well ashe uncertainty stemming from the underlying modeling assumptions that define the SEIR-mobility epidemiology

odel itself (which would apply to any other epidemiology model since it is always challenging to capturehe dynamics of detected patients in hundreds of different regions using one universal epidemiology model).urthermore, the uncertainty in the estimated R0(t) is higher than the one obtained for the predictions of C(t) since

R0(t) is a latent variable, unlike the number of cumulative cases C(t), for which observational data is available. Asconsequence, it is inevitable to obtain relatively large uncertainty estimates in the predictions since the uncertainty

n the model inputs (from the training data and the SEIR-mobility model limitations) is propagated to the outputncertainty, and most likely with amplification due to nonlinear interactions in the underlying dynamics.

As a consequence, the predictions of the proposed SEIR-mobility model cannot be the sole criterion for taking

ocial and economical decisions in order to slow down the disease propagation, but can provide one additional

14

Page 15: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

Fig. 10. Effects of re-opening: Real and altered workplaces mobility parameter (top left), baseline and altered estimates of β(t) (top right),predicted effective reproduction number R(t) (bottom left), and cumulative cases percentage C(t) (bottom right).

tool at the disposal of policy makers. Nonetheless, the obtained results in Fig. 10 show that the model is able towell reflect the impact of increasing the mobility on the dynamics of the reproduction number and the number ofcumulative cases, similarly to the variations observed in Figs. 5–7 which reflect the impact of gatherings ban andstay at home orders on the disease propagation dynamics.

4. Discussion

Amidst the COVID-19 outbreak, the modeling of the virus spread takes on life and death importance, as epidemicsimulations shape national responses [41]. The reliability of such models relies heavily on their calibration on a setof incomplete and imperfect data points; a task that is often viewed as art and has been the source of scientificcontroversy [42]. In this work, we aspire to develop state-of-the-art machine learning algorithms to accuratelydetermine the dynamic parameters characterizing the evolution of an epidemic, such as the time-dependent variationof the infectious rate and the effective reproduction number. Taking a systematic approach, the proposed modelcalibration framework will provide a quantitative understanding of the effects and efficiency of various policiesdesigned to contain the spread of COVID-19 (and other infectious diseases) [43], and produce probabilistic forecastsfor the evolution of a pandemic that can judiciously provide information for policy and decision making [44,45].

Specific to this study, the proposed approach allows us to understand the most relevant mobility and socialbehavior parameters that affect the spread of COVID-19. Our results suggest that working from home policies

have the significant effect in containing the reproduction number. This is shown in both global and local sensitivity

15

Page 16: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

aibacouom

dctooeoerttrf

C

P

KtWa

D

h

A

RoSf

analyses, where the mobility parameter that quantifies the amount of trips to workplaces is the most relevant amongall mobility and social behavior parameters. Working in confined spaces has been shown to be a risk factor for thetransmission of COVID-19 [46]. On the other hand, variations in park visits showed little influence in the dynamicsof pandemic, which is in line with the reduced risk of transmission while outdoors [47]. A somehow surprising resultis the mild influence of the human encounters rate parameter on the reproduction number. This could be explainedby the use of personal protective equipment, such as face masks, which greatly reduces the risk of transmissionand by a behavior change where human encounters moved from indoor to outdoor circumstances [48]. From theseresults, it appears that not all human encounters lead to an increased transmission, but only those that occur indoorsand for a prolonged period of time, such as in workplaces and in residential visits.

Another relevant point that our analysis revealed is the delay between changes in mobility and social behaviornd changes in the reproduction number. Variations that we see today in the reproduction number are mostlynfluenced by changes in behavior that occurred between 17 and 21 days ago. This time delay can be explainedy the incubation period of the virus [40] and by the delay in obtaining and reporting positive test results. Withoutny model to relate mobility and social behavior to the pandemic dynamics, policy makers are posed with a realhallenge of making decisions that will have a measurable outcome only after 2 to 3 weeks. This is where tools likeurs can have an impact to guide and accelerate the decisions of opening or closing certain areas. We highlight thetility of our model by simulating an increase in retail and recreation mobility in New York county, for which webserve that the reproduction number starts increasing only after 20 days of the intervention. By using the proposedodel, authorities could detect the threat of increased mobility early and act in time to prevent a new outbreak.Our methodology presents some limitations that open new research opportunities. First, there are pronounced

ifferences in the testing frequency both in space and time. These disparities will affect the number of detectedases and bias the data that we use to train our model. Including the number of tests performed as part of the inputo our model could improve our predictions and assess the effect of limited testing capabilities. However, these kindsf data are currently not available at the county level. Secondly, we have not included a potentially large fractionf cases that remain undetected or asymptomatic. This population can have an influence on the model behavior,specially when the case numbers are relatively large. Nonetheless, estimations of the undetected population basedn antibody studies are currently available only at very few locations [31]. For this reason, we decided to notxplicitly model this population, as we would be simply guessing its size. Finally, our model tends to performelatively better in counties with more cases, as we can see in Fig. 4. This is intrinsically linked to the loss functionhat we chose to train the model. In our case, this function tends to penalize more counties with more cases. Wehink this behavior is desirable because it makes our predictions more accurate in regions where the pandemic isapidly evolving. Nonetheless, it would be straightforward to adjust the loss function to emphasize other criticalactors in disease transmission if needed.

RediT authorship contribution statement

Mohamed Aziz Bhouri: Conceived the computational methods, Implemented the methods, Processed the data,erformed the simulations, Wrote the manuscript. Francisco Sahli Costabal: Conceived the computational methods,

Wrote the manuscript. Hanwen Wang: Assisted with the systematic numerical studies, Wrote the manuscript.evin Linka: Proposed the epidemiology model and data, Wrote the manuscript. Mathias Peirlinck: Proposed

he epidemiology model and data, Wrote the manuscript. Ellen Kuhl: Proposed the epidemiology model and data,rote the manuscript. Paris Perdikaris: Conceived the computational methods, Wrote the manuscript, Supervised

ll aspects of this projects.

eclaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that couldave appeared to influence the work reported in this paper.

cknowledgments

This work received support from the US Department of Energy under the Advanced Scientific Computingesearch program (Grant DE-SC0019116), the Defense Advanced Research Projects Agency under the Physicsf Artificial Intelligence program, United States of America (grant HR00111890034), and the Air Force Office ofcientific Research, United States of America (grant FA9550-20-1-0060). M.P. and E.K. also acknowledge support

rom a Stanford Bio-X IIP Seed Grant, E.K. by the NIH Grant U01 HL119578, and K.L. by a DAAD Fellowship.

16

Page 17: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

vw

1

1

1

Fig. A.11. Methodology summary: Using the dataset D and the mobility and social behavior-based SEIR model, the multi-step LSTMparameters along with the initial fractions of the exposed population for the different counties are estimated by minimizing the lossfunction (9).

Appendix A. Loss function and multi-step method

The forward pass of the LSTM neural network considered in this work for one single evaluation of βi (predictedalue of β at time instance ti ) is given in Algorithm 1, where σ refers to the sigmoid function, ◦ to the elementise multiplication between two tensors of same shape and (·, ·) to the euclidean inner product between two vectors

of same size. Based on the notation used in Algorithm 1, the LSTM parameters θ N N can be expressed as follows:

θ N N ≡ (W f , W i , W s, W o, U f , U i , U s, Uo, b f , bi , bs, bo, V , c) , (A.1)

where W f , U f b f refer to the forget gate’s parameters, W i , U i bi to the input/update gate’s parameters, W s ,U s bs to the cell state gate’s parameters, W o, Uo bo to the output gate’s parameters and V , c to the output layerparameters. dh refers to the hidden state dimension and will be taken equal to 120. These parameters are initializedfollowing the Xavier method [49] (see Fig. A.11).

Algorithm 1: LSTM forward pass

1 Inputs: τ , W f , W i , W s, W o ∈ Rdh×dh , U f , U i , U s, Uo ∈ RNm×dh , b f , bi , bs, bo, V ∈ Rdh , c ∈ R,m(ti−τ+k), k = 0, . . . , τ − 1

2 Outputs: βi

3 h−1, c−1 = 04 for k = 0 : τ − 1 do5 f k = σ [hk−1W f + m(ti−τ+k)U f + b f ]6 i k = σ [hk−1W i + m(ti−τ+k)U i + bi ]7 ck = tanh[hk−1W s + m(ti−τ+k)U s + bs]8 ck = f k ◦ ck−1 + i k ◦ ck

9 ok = σ [hk−1W o + m(ti−τ+k)Uo + bo]0 hk = ok ◦ tanh(ck)1 end2 βi = 2 × σ [(hτ−1, V ) + c]

Fig. A.12 shows the convergence of the loss function of Eq (9) with the number of iterations.Other loss functions have been investigated, including:

1. a loss based on the residual of the differential equation of I (t) in the SEIR model satisfied by I (t),2. a loss based on the L2 difference between the data of C(t) and the inferred number of cumulative cases:

R(t) + I (t)3. a loss based on the L2 difference between the increase of number of cumulative cases C(ti ) − C(ti−1) and

the inferred number of such increase R(t ) + I (t ) − R(t ) − I (t )

i i i−1 i−1

17

Page 18: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

leit

Btta

Ar

γ

Fig. A.12. Convergence of the training loss for each model in the ensemble (100 models with randomized parameter initializations, differentlines correspond to different models).

The loss defined in (9) performs better than the loss mentioned in point 1 because the latter is more prone toocal minima if it is not initialized sufficiently close to the exact solution and due to the noisy data and numericalrrors, the proposed loss (9) gave considerably better results. The losses in points 2 and 3 require one additional timentegration to estimate the variable R(t). Hence, such step introduces additional numerical errors, which explainshe better performance observed with the loss defined in (9).

Various multistep methods were investigated including implicit Adams–Moulton (AM) and explicit Adams–ashforth (AB) schemes with different orders. The implicit AM schemes perform better than the explicit AB ones

hanks to their numerical stability. Higher order schemes resulted in stiffer numerical systems which sometimes ledo numerical instabilities during model training. Therefore, the first order AM scheme (trapezoidal scheme) wasdopted in this work.

ppendix B. Comparison against Bayesian calibration for a simple SEIR model with a constanteproduction number β

The same model described in by Eq. (1) is adopted to utilize the cumulative data. In this case, the values ofand δ are assumed as given and are equal to 1/6.5 [32,33] and 1/2.5 [11,32], as before. We assume that the

parameter β varies from one county to another, but is fixed over time and independently distributed with respectto some prior distribution. Specifically, let β be the vector where each entry represents β of an individual county.We assume that the entries of β follow i.i.d. truncated normal distribution with mean equal to 0.5 and a scale of0.5. Furthermore, for every county, we assume that each entry of log(C), i.e., the natural logarithm of observedcumulative case fraction, follows i.i.d. normal distributions centered at log(C(β)) predicted by the model with β,whose standard deviations follow i.i.d. log normal distributions with mean 0 and scale 10. That is,

β ∼ TruncatedNormal(0.5, 0.5), (B.1)log(C) ∼ Normal(log(C(β)), σ ), (B.2)

σ ∼ LogNormal(0, 10). (B.3)

Let the likelihood of the observations log(C) for a given set of parameters β, σ be

p(log(C)|β, σ ), (B.4)

18

Page 19: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

t

s

f

a

b

ede

Fig. B.13. Predictive error comparison: Relative forecasting errors averaged over the 15 days extrapolation period in all 203 US countiesconsidered in this work. (Left) Proposed multi-step LSTM approach modeling a time-dependent reproduction number β(m(t)), (right) amulti-step approach with a constant reproduction number β.

hen Bayes’ theorem suggests the density of the parameters β, σ given the observation log(C)

p(β, σ | log(C)) =p(log(C)|β, σ )p(β, σ )

p(log(C)). (B.5)

Due to the difficulty of integrating the normalization term in the denominator, the probability density of theparameter is estimated by pseudo random sampling using Hamiltonian Monte-Carlo method. Specifically, the modelis calibrated via Hamiltonian Markov Chain Monte Carlo sampling using the No U-Turn Sampler (NUTS) [50]implementation of the NumPyro library [51] for 4500 iterations, with a target acceptance probability of 0.85. Wethen use the mean of the last 3000 samplings as an estimator for the parameter β. The relative error averaged overthe extrapolation period obtained with such model is shown in the right plot of Fig. B.13 for the different counties.We also tried several different priors for β, but all of them gave similar results. In general, the simple SEIR modeltends to produce cumulative cases fraction variations that are convex within the period of time considered, whilethe actual trajectories generally change concavity, due to the enforced behavioral intervention policies, such aslock-down and gatherings ban. Also, the sampling time is fairly long since the sampler is rejecting the candidatepoints because of the poor accuracy, further limiting the scalability of the method.

Appendix C. Model sensitivity on γ and δ

Overall, the specific values chosen for γ and δ do not have a significant impact on our model’s predictions. Toupport this claim, we have performed a series of systematic studies, as follows.

(1) Inferring γ and δ: Here we extend the proposed framework so that γ and δ are trainable constants inferredrom the data. Hence, the trainable parameters θ are now given as:

θ ≡ (θ N N , E1(t0), . . . , E Nc (t0), γ, δ) , (C.1)

nd we restrict the range of γ and δ such that:

γ ∈ [1/8.09, 1/3.59] = [0.1236, 0.2786] ,

δ ∈ [1/6.5, 1/1.5] = [0.1538, 0.6667] ,(C.2)

ased on the range proposed by the CDC in the manual of its COVID Tracer tool [52].2

We train an ensemble of 10 models with randomized parameter initializations. Once trained, the models providestimates of the cumulative cases fraction with a relative error equal to 4.5% on the training data for the first 66ays (to be compared to 4.3% obtained with our original choice to consider γ and δ as given), and a relative errorqual to 7.7% on the testing data, corresponding to the extrapolation carried out for the remaining 15 days of the

2 https://www.cdc.gov/coronavirus/2019-ncov/downloads/php/COVID TracerManual-508.pdf.

19

Page 20: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

c

d

Fig. C.14. Convergence of the estimated γ ∗ for the ensemble of 10 models with randomized parameter initializations, different linesorrespond to different models.

ata-set (to be compared to 6.6% obtained with γ and δ considered as given). Therefore, the model is able to fitthe data even when inferring γ and δ, and also provides acceptable extrapolation errors (only 1% higher than theextrapolation error obtained with γ and δ considered as given). However, the identifiability of a unique parametersetting is not guaranteed. Indeed, Figs. C.14 and C.15 show the convergence of the estimated γ and δ for the 10models, respectively, showing that not all 10 models converge to the same final values for γ and δ. Nonetheless,the main conclusions of our sensitivity analysis study are not altered, as shown in Fig. C.16.

As a conclusion, we can assert that, thanks to the expressiveness of the LSTM network, and by modeling thesusceptible to infectious transition rate β as a function of time, the proposed framework is able to well fit the datafor numerous counties and provide reliable extrapolations in time, regardless of the chosen values for γ and δ,and whether they are considered as trainable parameters or not. However, when attempting to infer γ and δ, theuniqueness of the solution is not guaranteed, which raises an issue of identifiability.

Note that the results obtained above are consistent with the fact that we only consider the cumulative cases I + Ras a reliable data metric. As a consequence, we cannot use the available data to trust-worthily infer the durationthat a person stays in the exposed and infected compartments (which would be needed to infer γ or δ).

Also recall that the main goal of our study is to assess the influence of the different mobility and social behaviorparameters on the spread of the disease. As shown in Fig. C.16, the sensitivity results are not altered by the factof including γ and δ within the trainable parameters, and since inferring these two parameters raises an issue ofidentifiability, we consider it is safer to use fixed values for such parameters based on previous existing works,and let the model infer the dependency of the susceptible to infectious transition rate β on the mobility and socialbehavior parameters (along with the initial fractions of the exposed population E(t = t0) for all the different regionsconsidered).

(2) Choosing a different fixed δ: To further justify the conclusions drawn above, and to verify the robustnessof the sensitivity results with respect to choosing a different values combination for γ and δ, we train 10 modelswith randomized parameter initializations using δ = 0.6 instead of the value 0.4 used in the original manuscript, asthese two values are within the range proposed by the CDC in the manual of its COVID Tracer tool [52]. Again,we verify that the sensitivity analysis results are not altered, as shown in Fig. C.17. Moreover, with δ = 0.6, themodels provide estimates of the cumulative cases fraction with a relative error equal to 4.3% on the training data for

20

Page 21: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

t

Fig. C.15. Convergence of the estimated δ∗ for the ensemble of 10 models with randomized parameter initializations, different lines correspondo different models.

Fig. C.16. si, j : Sensitivity of susceptible to infectious transition rate β to the mobility and social behavior parameters with models inferringγ and δ.

the first 66 days (exactly equal to what we obtained with δ = 0.4), and a relative error equal to 7.9% on the testingdata, corresponding to the extrapolation carried out for the remaining 15 days of the data-set (to be compared to6.6% with delta = 0.4).

Overall, we can state that δ and γ parameters are disease-specific and we fixed their values based on theinformation available at the time we carried out this study. The average of the values obtained in the works that wecited corresponds to δ = 1/2.5 and γ = 1/6.5. There are also some reasons for why one may consider differentvalues:

1. More information and data are being collected and reported every single day,

21

Page 22: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

emts

A

fd

tWe(th

gAi

t

tp

Fig. C.17. si, j : Sensitivity of susceptible to infectious transition rate β to the mobility and social behavior parameters with δ = 0.6.

2. but γ can also change based on- mutations of the virus and thus different disease time courses with the virus mutation that was initiallygoing through the population- increased insights into how to treat the disease and suppress the disease faster now than at the start whennot much was known.

However, changes in the γ and δ values due to the reasons above are not expected to impact the β(t) timevolution that our model learns because we only work with the confirmed cases which include all potential virusutation types, and the number of confirmed cases is independent of the disease treatment. This is confirmed by

he sensitivity results given above, indicating that δ’s variation does not really alter the results unless there is aignificant change, which is not the case if we refer to the ranges given by the CDC.

ppendix D. Predicted forecasts for a 21-day window

The 15 day-prediction window was chosen since it is typical to most of the existing studies [53,54]. Moreover,or the considered counties, the corresponding data-set has 81 days, which only leaves 66 days for the trainingata-set if a 15-day prediction is adopted.

Nonetheless, additional simulations are performed in order to assess the robustness of the results when usingraining data from the first 60 days, while the remaining 21 days are considered for extrapolation as testing data.

e train an ensemble of 10 models with randomized parameter initializations. Once trained, the models providestimates of the cumulative cases fraction with a relative error equal to 4.9% on the training data for the first 60 daysto be compared to 4.3% obtained with a training data-set of 66 days), and a relative error equal to 11.5% on theesting data, corresponding to the extrapolation carried out for the remaining 21 days of the data-set (considerablyigher than the 6.6% obtained with a testing data-set of 15 days, as expected).

Furthermore, as it can be seen in comparison to the 15-day predictions, the uncertainty in the estimated R0(t)rows for some counties (since R0(t) is a latent variable compared to the tighter uncertainty obtained for C(t)).iming for a more extended prediction range would inevitably result in a higher uncertainty, even for C(t) as shown

n Fig. D.18.We also verify that the main conclusions of our sensitivity analysis study are not altered when considering the

raining data-set of 60 days, as shown in Fig. D.19.Finally, most COVID-19-related political decisions are made based on two week-predictions [55], which justifies

he 15 days choice. Such interval is actually chosen as the upper limit of the virus incubation period [56] since

olitical decisions taken on a given day would have an impact on the virus spread after a period of around 2 weeks.

22

Page 23: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

Fig. D.18. Evolution of the fraction of cumulative cases C(t) and the effective reproduction number R0(t) for representative counties acrossthe US using a training data-set of 60 days and a testing data-set of 21 days. The orange and green vertical dashed lines indicate the datescorresponding to the large crowd gathering ban and stay at home order issued by each county. The pink vertical dashed line delineates thedata used for training the model, and the predicted model forecasts.

Therefore we believe that our initial choice of a 15-day prediction window is reasonable, and we are optimistic thatit can be stretched to accommodate longer time forecasts if more data is used for training (as the latter becomesavailable with the evolution of the pandemic).

23

Page 24: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

Fig. D.19. si, j : Sensitivity of susceptible to infectious transition rate β to the mobility and social behavior parameters with models using atraining dataset of 60 days.

References

[1] H. Lu, C.W. Stratton, Y.-W. Tang, Outbreak of pneumonia of unknown etiology in Wuhan, China: The mystery and the miracle, 92(4) 401–402, http://dx.doi.org/10.1002/jmv.25678, https://onlinelibrary.wiley.com/doi/abs/10.1002/jmv.25678.

[2] E. Dong, H. Du, L. Gardner, An interactive web-based dashboard to track COVID-19 in real time, Lancet. Infect. Dis. 20 (5) (2020)533–534, http://dx.doi.org/10.1016/S1473-3099(20)30120-1, https://linkinghub.elsevier.com/retrieve/pii/S1473309920301201.

[3] World Health Organization, Strategic preparedness and response plan, https://www.who.int Accessed: July 8th 2020.[4] Coronavirus Government Response Tracker, https://www.bsg.ox.ac.uk/research/research-projects/coronavirus-government-response-

tracker Accessed: July 8th 2020.[5] M. Peirlinck, K. Linka, F. Sahli Costabal, E. Kuhl, Outbreak dynamics of COVID-19 in China and the united states, Biomech. Model.

Mechanobiol. (2020) http://dx.doi.org/10.1007/s10237-020-01332-5, http://link.springer.com/10.1007/s10237-020-01332-5.[6] M. Nicola, Z. Alsafi, C. Sohrabi, A. Kerwan, A. Al-Jabir, C. Iosifidis, M. Agha, R. Agha, The socio-economic implications of the

coronavirus pandemic (COVID-19): A review, 78, 185–193 http://dx.doi.org/10.1016/j.ijsu.2020.04.018, https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7162753/.

[7] Process to Reopen Pennsylvania, https://www.governor.pa.gov/process-to-reopen-pennsylvania/ Accessed: July 8th 2020.[8] X. He, E.H. Lau, P. Wu, X. Deng, J. Wang, X. Hao, Y.C. Lau, J.Y. Wong, Y. Guan, X. Tan, et al., Temporal dynamics in viral

shedding and transmissibility of COVID-19, Nat. Med. 26 (5) (2020) 672–675.[9] W.-K. Wang, S.-Y. Chen, I.-J. Liu, Y.-C. Chen, H.-L. Chen, C.-F. Yang, P.-J. Chen, S.-H. Yeh, C.-L. Kao, L.-M. Huang, P.-R. Hsueh,

J.-T. Wang, W.-H. Sheng, C.-T. Fang, C.-C. Hung, S.-M. Hsieh, C.-P. Su, W.-C. Chiang, J.-Y. Yang, J.-H. Lin, S.-C. Hsieh, H.-P.Hu, Y.-P. Chiang, J.-T. Wang, P.-C. Yang, S.-C. Chang, Detection of SARS-associated coronavirus in throat wash and saliva in earlydiagnosis 10 (7) 1213–1219 http://dx.doi.org/10.3201/eid1007.031113, https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3323313/.

[10] G. La Rosa, L. Bonadonna, L. Lucentini, S. Kenmoe, E. Suffredini, Coronavirus in water environments: Occurrence, persistenceand concentration methods - a scoping review, Water Res. 179 (2020) 115899, http://dx.doi.org/10.1016/j.watres.2020.115899, https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7187830/.

[11] S.A. Lauer, K.H. Grantz, Q. Bi, F.K. Jones, Q. Zheng, H.R. Meredith, A.S. Azman, N.G. Reich, J. Lessler, The incubation periodof coronavirus disease 2019 (COVID-19) from publicly reported confirmed cases: estimation and application, Ann. Int. Med. 172 (9)(2020) 577–582.

[12] M.M. Arons, K.M. Hatfield, S.C. Reddy, A. Kimball, A. James, J.R. Jacobs, J. Taylor, K. Spicer, A.C. Bardossy, L.P. Oakley, S.Tanwar, J.W. Dyal, J. Harney, Z. Chisty, J.M. Bell, M. Methner, P. Paul, C.M. Carlson, H.P. McLaughlin, N. Thornburg, S. Tong,A. Tamin, Y. Tao, A. Uehara, J. Harcourt, S. Clark, C. Brostrom-Smith, L.C. Page, M. Kay, J. Lewis, P. Montgomery, N.D. Stone,T.A. Clark, M.A. Honein, J.S. Duchin, J.A. Jernigan, Public Health–Seattle and King County and CDC COVID-19 Investigation Team,Presymptomatic SARS-CoV-2 infections and transmission in a skilled nursing facility, New Engl. J. Med. 382 (22) (2020) 2081–2090,http://dx.doi.org/10.1056/NEJMoa2008457.

[13] M.M. Böhmer, U. Buchholz, V.M. Corman, M. Hoch, K. Katz, D.V. Marosevic, S. Böhm, T. Woudenberg, N. Ackermann, R. Konrad,U. Eberle, B. Treis, A. Dangel, K. Bengs, V. Fingerle, A. Berger, S. Hörmansdorfer, S. Ippisch, B. Wicklein, A. Grahl, K. Pörtner,N. Muller, N. Zeitlmann, T.S. Boender, W. Cai, A. Reich, M. An der Heiden, U. Rexroth, O. Hamouda, J. Schneider, T. Veith, B.Mühlemann, R. Wölfel, M. Antwerpen, M. Walter, U. Protzer, B. Liebl, W. Haas, A. Sing, C. Drosten, A. Zapf, Investigation ofa COVID-19 outbreak in Germany resulting from a single travel-associated primary case: a case series, Lancet. Infect. Dis. (2020)http://dx.doi.org/10.1016/S1473-3099(20)30314-5.

[14] W.E. Wei, Presymptomatic transmission of SARS-CoV-2 — Singapore, january 23–march 16, 2020, in: MMWR. Morbidity andMortality Weekly Report 69, 2020, http://dx.doi.org/10.15585/mmwr.mm6914e1, https://www.cdc.gov/mmwr/volumes/69/wr/mm6914e1.

htm.

24

Page 25: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

[15] CDC, Information for Laboratories about Coronavirus (COVID-19), https://www.cdc.gov/coronavirus/2019-ncov/lab/resources/antibody-tests.html Accessed: July 8th 2020.

[16] A. Endo, S. Abbott, A.J. Kucharski, S. Funk, Centre for the Mathematical Modelling of Infectious Diseases COVID-19 WorkingGroup, Estimating the overdispersion in COVID-19 transmission using outbreak sizes outside China, Wellcome Open Res. 5 (2020)67, http://dx.doi.org/10.12688/wellcomeopenres.15842.3, https://wellcomeopenresearch.org/articles/5-67/v3.

[17] Google LLC, Google COVID-19 community mobility reports, 2020, https://www.google.com/covid19/mobility/ Accessed: June 8th2020.

[18] Apple Inc., Mobility trends reports, https://www.apple.com/covid19/mobility Accessed: July 8th 2020.[19] H.W. Hethcote, The mathematics of infectious diseases, SIAM Rev. 42 (4) (2000) 599–653.[20] A.C. Miller, N.J. Foti, J.A. Lewnard, N.P. Jewell, C. Guestrin, E.B. Fox, Mobility trends provide a leading indicator of changes in

SARS-CoV-2 transmission, MedRxiv (2020).[21] Z. Wang, X. Zhang, G.H. Teichert, M. Carrasco-Teja, K. Garikipati, System inference for the spatio-temporal evolution of infectious

diseases: Michigan in the time of COVID-19, Comput. Mech. 66 (5) (2020) 1153–1176, http://dx.doi.org/10.1007/s00466-020-01894-2,http://link.springer.com/10.1007/s00466-020-01894-2.

[22] K. Linka, M. Peirlinck, F.S. Costabal, E. Kuhl, Outbreak Dynamics of COVID-19 in Europe and the Effect of Travel Restrictions,Epidemiology, 2020, http://dx.doi.org/10.1101/2020.04.18.20071035, http://medrxiv.org/lookup/doi/10.1101/2020.04.18.20071035.

[23] A. Arnold, D. Calvetti, E. Somersalo, Parameter estimation for stiff deterministic dynamical systems via ensemble Kalman filter, InverseProblems 30 (10) (2014) 105008, http://dx.doi.org/10.1088/0266-5611/30/10/105008, https://iopscience.iop.org/article/10.1088/0266-5611/30/10/105008.

[24] P.K. Jha, L. Cao, J.T. Oden, Bayesian-based predictions of COVID-19 evolution in Texas using multispecies mixture-theoreticcontinuum models, Comput. Mech. 66 (5) (2020) 1055–1068, http://dx.doi.org/10.1007/s00466-020-01889-z, http://link.springer.com/10.1007/s00466-020-01889-z.

[25] P. Chen, K. Wu, O. Ghattas, Bayesian inference of heterogeneous epidemic models: application to COVID-19 spread accounting forlong-term care facilities, 2020, arXiv preprint arXiv:2011.01058.

[26] D. Calvetti, S. Nakkireddy, E. Somersalo, A Bayesian filtering approach to layer stripping for electrical impedance tomography, InverseProblems 36 (5) (2020) 055014, http://dx.doi.org/10.1088/1361-6420/ab6f9e, https://iopscience.iop.org/article/10.1088/1361-6420/ab6f9e.

[27] X. Zeng, R. Ghanem, Dynamics identification and forecasting of COVID-19 by switching Kalman filters, Comput. Mech. 66 (5) (2020)1179–1193, http://dx.doi.org/10.1007/s00466-020-01911-4, http://link.springer.com/10.1007/s00466-020-01911-4.

[28] Unacast, Unacast social distancing dataset, 2020, https://www.unacast.com/data-for-good. Version from June 11th 2020.[29] K. Linka, A. Goriely, E. Kuhl, Global and local mobility as a barometer for COVID-19 dynamics, medRxiv (2020) http://dx.doi.org/

10.1101/2020.06.13.20130658, https://www.medrxiv.org/content/10.1101/2020.06.13.20130658v1, 2020.06.13.20130658.[30] M. Raissi, P. Perdikaris, G.E. Karniadakis, Multistep neural networks for data-driven discovery of nonlinear dynamical systems, 2018,

arXiv preprint arXiv:1801.01236.[31] M. Peirlinck, K. Linka, F.S. Costabal, E. Bendavid, J. Bhattacharya, J. Ioannidis, E. Kuhl, Visualizing the invisible: The effect of

asymptomatic transmission on the outbreak dynamics of COVID-19, Comput. Methods Appl. Mech. Engrg. (2020) 113410.[32] Q. Li, X. Guan, P. Wu, X. Wang, L. Zhou, Y. Tong, R. Ren, K.S. Leung, E.H. Lau, J.Y. Wong, et al., Early transmission dynamics

in wuhan, China, of novel coronavirus–infected pneumonia, New Engl. J. Med. (2020).[33] S. Sanche, Y.T. Lin, C. Xu, E. Romero-Severson, N. Hengartner, R. Ke, Early release-high contagiousness and rapid spread of severe

acute respiratory syndrome coronavirus 2, 2020.[34] S. Hochreiter, J. Schmidhuber, Long short-term memory, Neural Comput. 9 (8) (1997) 1735–1780.[35] A. Iserles, A first course in the numerical analysis of differential equations, second ed., in: Cambridge Texts in Applied Mathematics,

Cambridge University Press, 2008, http://dx.doi.org/10.1017/CBO9780511995569.[36] D.P. Kingma, J. Ba, Adam: A method for stochastic optimization, 2014, arXiv preprint arXiv:1412.6980.[37] N. Altieri, R. Barter, J. Duncan, R. Dwivedi, K. Kumbier, X. Li, R. Netzorg, B. Park, C. Singh, Y.S. Tan, et al., Curating a COVID-19

data repository and forecasting county-level death counts in the United States, 2020, arXiv preprint arXiv:2005.07882.[38] M.D. Morris, Factorial sampling plans for preliminary copmputational experiments, Technometrics 33 (2) (1991) 161–174.[39] F. Campolongo, J. Cariboni, A. Saltelli, An effective screening design for sensitivity analysis of large models, Environ. Model. Softw.

22 (10) (2007) 1509–1518, http://dx.doi.org/10.1016/j.envsoft.2006.10.004.[40] U. CDC, Symptoms of coronavirus, 2020, https://www.cdc.gov/coronavirus/2019-ncov/symptoms-testing/symptoms.html Accessed: June

10th 2020.[41] World Health Organization, et al., Coronavirus Disease 2019 (COVID-19): Situation Report, 72, World Health Organization, 2020.[42] M. Enserink, K. Kupferschmidt, With COVID-19, modeling takes on life and death importance, Science (2020).[43] National Science and Technology Council, Towards epidemic prediction: federal efforts and opportunities in outbreak modeling, 2016.[44] D.E. Bloom, D. Cadarette, Infectious disease threats in the twenty-first century: strengthening the global response, Front. Immunol. 10

(2019) 549, http://dx.doi.org/10.3389/fimmu.2019.00549.[45] C.E. Walters, M.M. Meslé, I.M. Hall, Modelling the global spread of diseases: A review of current practice and capability, Epidemics

25 (2018) 1–8, http://dx.doi.org/10.1016/j.epidem.2018.05.007.[46] S. Park, Y. Kim, S. Yi, S. Lee, B. Na, C. Kim, J. Kim, H. Kim, Y. Kim, Y. Park, et al., Coronavirus disease outbreak in call center,

South Korea, Emerg. Infect. Dis. 26 (8) (2020).[47] H. Nishiura, H. Oshitani, T. Kobayashi, T. Saito, T. Sunagawa, T. Matsui, T. Wakita, M. COVID, M. Suzuki, Closed environments

facilitate secondary transmission of coronavirus disease 2019 (COVID-19), MedRxiv (2020) http://dx.doi.org/10.1101/2020.05.07.

20094441.

25

Page 26: COVID-19 dynamics across the US: A deep learning study of

M.A. Bhouri, F.S. Costabal, H. Wang et al. Computer Methods in Applied Mechanics and Engineering 382 (2021) 113891

[48] D.K. Chu, E.A. Akl, S. Duda, K. Solo, S. Yaacoub, H.J. Schünemann, A. El-harakeh, A. Bognanni, T. Lotfi, M. Loeb, et al., Physicaldistancing, face masks, and eye protection to prevent person-to-person transmission of SARS-CoV-2 and COVID-19: a systematicreview and meta-analysis, Lancet (2020).

[49] X. Glorot, Y. Bengio, Understanding the difficulty of training deep feedforward neural networks, in: Proceedings of the ThirteenthInternational Conference on Artificial Intelligence and Statistics, 2010, pp. 249–256.

[50] M.D. Hoffman, A. Gelman, The No-U-Turn sampler: adaptively setting path lengths in Hamiltonian Monte Carlo, J. Mach. Learn.Res. 15 (1) (2014) 1593–1623.

[51] D. Phan, N. Pradhan, M. Jankowiak, Composable effects for flexible and accelerated probabilistic programming in NumPyro, 2019,arXiv preprint arXiv:1912.11554.

[52] CDC, Health departments, Cent. Dis. Control Prev. (2020) https://www.cdc.gov/coronavirus/2019-ncov/php/contact-tracing/COVIDTracerTools.html.

[53] M.A. Achterberg, B. Prasse, L. Ma, S. Trajanovski, M. Kitsak, P. Van Mieghem, Comparing the accuracy of several network-basedCOVID-19 prediction algorithms, Int. J. Forecast. (2020) http://dx.doi.org/10.1016/j.ijforecast.2020.10.001, http://www.sciencedirect.com/science/article/pii/S0169207020301552.

[54] C.S. Lutz, M.P. Huynh, M. Schroeder, S. Anyatonwu, F.S. Dahlgren, G. Danyluk, D. Fernandez, S.K. Greene, N. Kipshidze, L. Liu,O. Mgbere, L.A. McHugh, J.F. Myers, A. Siniscalchi, A.D. Sullivan, N. West, M.A. Johansson, M. Biggerstaff, Applying infectiousdisease forecasting to public health: a path forward using influenza forecasting examples, BMC Public Health 19 (1) (2019) 1659,http://dx.doi.org/10.1186/s12889-019-7966-8, https://bmcpublichealth.biomedcentral.com/articles/10.1186/s12889-019-7966-8.

[55] E. Silverman, Philly plans to let indoor dining resume | Coronavirus Newsletter, https://www.inquirer.com/health/coronavirus/newsletter/covid19-coronavirus-philadelphia-indoor-dining-restrictions-herd-immunity-vaccine-20210109.html.

[56] CDC, Healthcare workers, Cent. Dis. Control Prev. (2020) https://www.cdc.gov/coronavirus/2019-ncov/hcp/clinical-guidance-management-patients.html.

26