ArticlePDF Available

Mitigation Strategies for Pandemic Influenza in the United States

Authors:

Abstract and Figures

Recent human deaths due to infection by highly pathogenic (H5N1) avian influenza A virus have raised the specter of a devastating pandemic like that of 1917–1918, should this avian virus evolve to become readily transmissible among humans. We introduce and use a large-scale stochastic simulation model to investigate the spread of a pandemic strain of influenza virus through the U.S. population of 281 million individuals for R 0 (the basic reproductive number) from 1.6 to 2.4. We model the impact that a variety of levels and combinations of influenza antiviral agents, vaccines, and modified social mobility (including school closure and travel restrictions) have on the timing and magnitude of this spread. Our simulations demonstrate that, in a highly mobile population, restricting travel after an outbreak is detected is likely to delay slightly the time course of the outbreak without impacting the eventual number ill. For R 0 < 1.9, our model suggests that the rapid production and distribution of vaccines, even if poorly matched to circulating strains, could significantly slow disease spread and limit the number ill to <10% of the population, particularly if children are preferentially vaccinated. Alternatively, the aggressive deployment of several million courses of influenza antiviral agents in a targeted prophylaxis strategy may contain a nascent outbreak with low R 0, provided adequate contact tracing and distribution capacities exist. For higher R 0, we predict that multiple strategies in combination (involving both social and medical interventions) will be required to achieve similar limits on illness rates. • antiviral agents • infectious diseases • simulation modeling • social network dynamics • vaccines
Content may be subject to copyright.
Mitigation strategies for pandemic influenza
in the United States
Timothy C. Germann*
, Kai Kadau*, Ira M. Longini, Jr.
, and Catherine A. Macken*
*Los Alamos National Laboratory, Los Alamos, NM 87545; and
Program of Biostatistics and Biomathematics, Fred Hutchinson Cancer Research Center and
Department of Biostatistics, School of Public Health and Community Medicine, University of Washington, Seattle, WA 98109
Communicated by G. Balakrish Nair, International Centre for Diarrhoeal Disease Research Bangladesh, Dhaka, Bangladesh, February 16, 2006
(received for review January 10, 2006)
Recent human deaths due to infection by highly pathogenic (H5N1)
avian influenza A virus have raised the specter of a devastating
pandemic like that of 1917–1918, should this avian virus evolve to
become readily transmissible among humans. We introduce and
use a large-scale stochastic simulation model to investigate the
spread of a pandemic strain of influenza virus through the U.S.
population of 281 million individuals for R
0
(the basic reproductive
number) from 1.6 to 2.4. We model the impact that a variety of
levels and combinations of influenza antiviral agents, vaccines, and
modified social mobility (including school closure and travel re-
strictions) have on the timing and magnitude of this spread. Our
simulations demonstrate that, in a highly mobile population,
restricting travel after an outbreak is detected is likely to delay
slightly the time course of the outbreak without impacting the
eventual number ill. For R
0
< 1.9, our model suggests that the rapid
production and distribution of vaccines, even if poorly matched to
circulating strains, could significantly slow disease spread and limit
the number ill to <10% of the population, particularly if children
are preferentially vaccinated. Alternatively, the aggressive deploy-
ment of several million courses of influenza antiviral agents in a
targeted prophylaxis strategy may contain a nascent outbreak
with low R
0
, provided adequate contact tracing and distribution
capacities exist. For higher R
0
, we predict that multiple strategies
in combination (involving both social and medical interventions)
will be required to achieve similar limits on illness rates.
antiviral agents infectious diseases simulation modeling
social network dynamics vaccines
I
t is inevitable that another influenza pandemic will occur, and
recent events sugge st that this might happen sooner rather than
later (1). A highly pathogenic H5N1 influenza A virus appears to
have become endemic in avian hosts in Asia, and it is now spreading
in migratory birds westward across eastern Europe. Human infec-
tions caused by this virus have a high case fatality rate; together with
recent genetic data that implicate direct transmission of avian-
adapted influenza virus to humans as the cause of the 1918
influenza pandemic (2), these conditions raise the specter of
another devastating pandemic. To date, H5N1 viruse s cannot
transmit readily from human to human, thus providing a window to
plan for the pandemic that will occur should the virus evolve to be
readily transmissible among humans. If the nascent pandemic is not
contained by timely intervention at its source (3, 4), international
travel could carry pandemic viruses around the globe within weeks
to months of the initiation of the outbreak, causing a worldwide
public health emergency.
Intensive pandemic planning is occurring at the national [U.S.
Department of Health and Human Services (HHS) Pandemic
Inf luenza Plan, www.hhs.govpandemicfluplan) and interna-
tional [World Health Organization (WHO) Global Inf luenza
P reparedness Plan, www.who.intcsrresourcespublications
influenzaWHOCDSCSRGIP20055enindex.html] levels.
The most pressing public health questions are: what might be the
time c ourse and geographic spread of the outbreak, and what is
the most effective utilization of available therapeutic and social
resources to min imize the impact of the outbreak? Precise
plann ing is hampered by several unknowns, most critically the
eventual human-to-human transmissibility of the human-
adapted avian strain (characterized by the basic reproductive
number R
0
, the average number of secondary infections caused
by a single typical infected individual among a completely
susceptible population), and the supply of therapeutic agents.
Manufacturers of neuraminidase inhibitors, such as oselt amivir,
have committed to considerable increases in production over the
next 3–4 years. However, the production of vaccine, the tradi-
tional first line of defense against influenza vir us infections, is
hampered by the inability to predict the antigenic details of the
evolved vir us at the time that it becomes a pandemic strain and
the consequent inability to prepare a highly effective vaccine in
advance of a pandemic outbreak. Given these uncertainties, it is
import ant to develop multiple mitigation strategies, involving
vac cination, prophylaxis with antiviral drugs, and both voluntary
and imposed changes in social patterns such as school closures
and travel restrictions.
The course of an influenza outbreak is sensitive to many factors,
particularly population mobility and the susceptibility of individuals
to the virus. Traditional mathematical models of epidemics often
take the form of deterministic SIR differential equations for the
population dynamics of susceptible (S), infectious (I), and re-
movedrecovered (R) individuals (5, 6). Such models have also
been extended to model the geographic spread of infectious dis-
eases (7, 8). However, the population-based nature of this class of
models best describes the dynamics of an epidemic when large
numbers of individuals are infected, rather than the initial or final
stage s of an outbreak, when small numbers of individuals are
involved and stochastic person-to-person transmission processes
dominate. To satisfactorily model the initial seeding and final
quenching of small community-level outbreaks requires a funda-
mentally different approach. To capture this crucial effect of
uncertainty in transmission on epidemic predictions, we develop
and use a stochastic agent-based discrete-time simulation model.
This class of model has been used to assess vaccination and antiviral
prophylaxis strategies on a local level (9–11); larger-scale versions
have recently been used to investigate strategie s at a regional level
for containing an emerging pandemic influenza strain at its source
(3, 4). Our national-level model combines an individual-level
description of influenza viral infection and transmission dynamics
with high-fidelity U.S. Census Bureau and Department of Trans-
portation data on population demographics and mobility, yielding
a massive-scale simulation model of the spatiotemporal dynamics of
spread of a pandemic strain of influenza virus among an artificial
U.S. population of 281 million people. Such an endeavor is only now
practical with modern parallel supercomputing platforms and pro-
gramming technique s.
Conflict of interest statement: No conflicts declared.
Abbreviations: TAP, targeted antiviral prophylaxis; NAI, neuraminidase inhibitor.
To whom correspondence should be addressed. E-mail: tcg@lanl.gov.
© 2006 by The National Academy of Sciences of the USA
www.pnas.orgcgidoi10.1073pnas.0601266103 PNAS
April 11, 2006
vol. 103
no. 15
5935–5940
MEDICAL SCIENCES
Results and Discussion
Simulation Model Design. The model population of 281 million
individuals is distributed among 65,334 census tracts to closely
represent the actual population distribution according to publicly
available 2000 U.S. Census data (www.census.govmainwww
cen2000.html). Each tract is in turn organized into 2,000-person
communities. The model runs in cycles of two 12-hour periods
(‘‘day’’ and ‘‘night’’), during which we identify seven contexts
(‘‘mixing groups’’) within which individuals can associate. In five of
these contexts (households, household clusters, preschools, play-
groups, schools, and work groups), relatively close person-to-person
association regularly occurs. Additionally, ‘‘neighborhoods’’ and
‘‘communities’’ provide unspecified contexts (e.g., shopping malls)
within which occasional casual person-to-person association occurs.
Because each individual may interact with any member of his or her
mixing group, the group sizes determine the numbers of people who
would be considered for antiviral prophylaxis in our socially tar-
geted strategy of mitigation (below). Daytime contacts occur in
neighborhoods and communities as well as in the age-appropriate
setting, and nighttime contacts occur only in households, household
clusters, neighborhoods, and communities. U.S. Census data on
tract-to-tract worker flow is used to model the commute of working
adults to their workplace, thus accurately capturing the short- to
medium-distance population mobility important for disease spread.
In addition, each individual takes occasional long-distance trips
(three per year on average), lasting between 1 day and 3 weeks (4.1
days on average), matching Bureau of Transportation Statistics
data (www.bts.govpublicationsnationaltransportationstatistics).
Our simple model of long-range travel could be extended to
account for different types of travel (e.g., business or leisure) or
groups of travelers (such as a family) or to explicitly incorporate the
airline network structure, as in ref. 8.
The disease transmission and natural history models are briefly
described in Materials and Methods, with further details provided in
Supporting Text, Figs. 3–5, and Tables 3–5, which are published as
supporting information on the PNAS web site. To model the
introduction of pandemic influenza into the U.S., we assume that
impenetrable borders are either prohibitively expensive or impos-
sible to create, and that international air travel is the dominant
mode of influenza introduction from outside the U.S. Conse-
quently, a small random number of incubating individuals, equiv-
alent to 0.04% of arriving international passengers, is introduced
each day at each of 14 major international airports in the conti-
nental U.S. (see Table 6, which is published as supporting infor-
mation on the PNAS web site). The simulation covers 180 days,
roughly the length of a U.S. influenza season. We assume that,
because of the uncertainty in diagnosis of influenza infections and
the sporadic nature of the early stages of an outbreak, a cumulative
number of 10,000 symptomatic individuals nationwide is required
to trigger a nationwide pandemic alert (see Supporting Text for a
sensitivity analysis of various response delays, for selected inter-
vention strategies).
Intervention Strategies. A variety of intervention strategies com-
posed of one or more of the following four actions is considered: (i)
socially targeted antiviral prophylaxis (TAP), in which symptomatic
individuals and most of their close contacts receive treatment or
prophylaxis, respectively, with antiviral drugs; (ii) dynamic mass
vaccination, either of a random selection of individuals from the
entire population or with preference for children, and with various
production and distribution rate s and starting dates; (iii) closure of
schools, including preschools and play groups; and (iv) social
distancing, as a result of legally mandated travel restriction or
quarantine programs, or voluntary changes in social behavior.
TAP (11) is triggered by the first symptomatic person to be
ascertained within a household (the index case). Because symp-
tomatic diagnosis of influenza viral infection is inaccurate, leading
either to delays in accurate diagnosis by biological assay or to
excessive use of antivirals due to false positives, we simulate several
scenarios. For the majority of our simulations, we assume that there
are 0% false positive s, and 60% of index cases are ascertained (the
rest omitted because of, for example, misdiagnosis or lack of acce ss
to health care). When an index case is ascertained, he or she is
treated, and all remaining people in this person’s household and
household cluster are offered prophylaxis. If an ascertained index
case belongs to a daycare, preschool, school, or workplace, then
100% of the people in that daycare or preschool are offered
prophylaxis, or 60% of the people in that school or workplace.
(Results for other ascertainment percentages, diagnosis delays,
false positives, or prophylaxis strategies are presented in Supporting
Text.) Due to its labor-intensive nature, TAP is likely to be feasible
only during the earliest stages of an outbreak in any particular
community, before the community health system is overwhelmed.
The dose and duration for effective treatment and prophylaxis
using neuraminidase inhibitors (NAIs) against currently circulating
strains of human influenza virus are well known (12), although a
recent study suggests that an increased dose and duration of
treatment may be needed to counteract H5N1 viral infections (13).
On the other hand, these viruses may not retain their current
unusually high growth rates if they evolve to be readily human-to-
human transmissible. In light of this uncertainty, we use the current
manufacturer’s recommended dose (10 tablets of oseltamivir for 5
days of treatment or 10 days of prophylaxis) and current estimate s
of oseltamivir efficacy in reducing infectiousne ss and susceptibility
(see Supporting Text) (3, 14). Administration of a single course of
NAI is initiated the day after the index case is ascertained,
providing therapy for the index case and prophylaxis for others. A
susceptible individual may receive subsequent courses of NAIs if
another index case occurs later in a mixing group of which he or she
is a member. We assume that 5% of people who start taking
influenza antiviral agents will stop taking them after 1 day of
treatment or prophylaxis.
Interventions involving vaccination suffer from uncertainty
about the future identity of a pandemic strain, making it impossible
to stockpile well matched pandemic vaccine s. However, prevacci-
nation based on a killed avian virus precursor to the pandemic
strain is possible, providing a perhaps poorly matched but poten-
tially efficacious vaccine. Vaccination can also be based on killed or
live attenuated emergent pandemic virus, providing a close match
to the subsequent circulating strains but available with a lag of a few
months from emergence (in nonpandemic years, vaccine manufac-
ture take s between 6 and 9 months). A ‘‘dynamic vaccination’’
scenario, in which vaccine becomes available incrementally, starting
from as early as 2 months before, to as late as 2 months after, the
first individual in the U.S. is infected, is inve stigated, with different
production rates, total production amounts, and distribution poli-
cies (either uniformly throughout the population or preferentially
to children). We compare the administration of the recommended
two doses conferring best protection levels to a strategy in which
twice as many people are given a single dose, assuming that a single
dose of vaccine confers about half the protection of two doses (15).
§
Much uncertainty exists about the societal acceptability of op-
tions for creating social distance and thereby reduction in trans-
mission. Given the importance of children in the transmission of
influenza (16), school closure is likely to be an effective (albeit
burdensome) social distancing policy. Although formally imposed
quarantine or travel restriction policies are possible, voluntary
§
In fact, efficacy of experimental vaccines against a novel pandemic strain cannot be
ascertained in the absence of actual viral challenge; immunogenicity alone can be deter-
mined. Experimental vaccines based on avian influenza virus have required much greater
amounts of antigen for acceptable levels of immunogenicity than standard human
vaccines. This discrepancy does not enter into our calculations of required doses of vaccine.
We assume that pandemic vaccines will have the same relationship between efficacy and
immunogenicity as that for standard vaccines against human influenza virus.
5936
www.pnas.orgcgidoi10.1073pnas.0601266103 Germann et al.
changes in hygienic and social behavior (including travel plans) will
undoubtedly occur. Indeed, the spontaneous public re sponse to
news of an approaching pandemic will affect social behavior in
unpredictable ways, so the social distancing strategies explored here
are hopefully realistic approximations to voluntary or imposed
distancing at three different scales: at the levels of schools, local
communities, and nationwide travel. At the local scale, this social
distancing is assumed to manife st itself in a concentration of
interactions within households and household clusters, and at
longer scales we consider uniform reductions in the amount of
long-range travel to as little as 1% of the normal frequency. (See
Supporting Text for details of implementation.) Although the social
distancing measures studied here form a necessary first step in
modeling such effects on disease transmission, further investigation
is needed into variations in contact structure that are not considered
in our model (e.g., classroom size variations with geographic region
and grade level, parents staying at home with sick children, and
other venues and mechanisms for transmission).
Simulation Results. Independent realizations of our simulations for
a given set of parameter values lead to very similar epidemic curves
(see Supporting Text, Table 7, and Figs. 6 and 7, which are published
as supporting information on the PNAS web site, for details
including the estimation of R
0
). In the absence of intervention, for
R
0
1.9, our simulated pandemic begins with sporadic outbreaks
occurring across the country in areas of dense population for 24
days before the outbreak is recognized (Table 1 and Movie 1, which
is published as supporting information on the PNAS web site). The
pandemic peaks after 85 days, with a final illne ss attack rate of 43%
(Table 2). The greatest nationwide activity is concentrated in a
2-month period when 100,000 people become ill each day,
although local areas differ in the timing and duration of their highly
active periods. This coincide s quite well with waves of past pan-
demics; the 1957–1958 influenza A (H2N2) ‘‘Asian’’ virus initially
appeared in June and July 1957, as sporadic cases in Iowa, Loui-
siana, and the West Coast, developing into local outbreaks during
August 1957 before peaking in a 60-day period covering September
and October 1957 (17). Similarly, the 1968–1969 influenza A
(H3N2) ‘‘Hong Kong’’ virus first appeared as sporadic cases along
the West Coast in July 1968, developing into local outbreaks 3
months later in October and peaking in December 1968 and
January 1969, before finally ending in March 1969 (7).
Although the dynamics of the pandemic in the absence of
mitigation are clearly sensitive to R
0
, interestingly, this sensitivity is
modest when R
0
increases beyond 1.9 compared with the effect of
increasing R
0
from 1.6 to 1.9 (Table 1). We use as our guideline for
adequate mitigation a reduction in the overall rate of illness to no
greater than that of a typical influenza epidemic, 10%. The results
presented in Fig. 2 and Table 2 suggest that as R
0
increases from 1.6
to 1.9, a transition occurs from an outbreak that can be mitigated
with moderate efforts, to one that can be mitigated only with
vigorous application of multiple strategie s. For example, several of
the single interventions that we simulated are successful for R
0
1.6, with TAP the most effective single intervention for our model
of social mobility and transmission, provided adequate antiviral
supplies exist and close contacts can be rapidly identified (see Fig.
1 and Movie 2, which is published as supporting information on the
PNAS web site). In contrast, for R
0
1.7, 10.0 million courses are
Table 1. Characteristics of simulated pandemic influenza in the
U.S. in the absence of interventions
Basic reproductive number, R
0
1.6 1.9 2.1 2.4
Rate of spread: 1,000th ill person* 14 13 12 11
10,000th ill person* 29 24 22 19
100,000th ill person* 48 37 34 29
1,000,000th ill person* 70 52 46 39
Peak of epidemic* 117 85 75 64
Daily number of new cases at peak
activity
2.3 M 4.5 M 6.0 M 7.9 M
Number of days with 100,000
new cases
86 68 60 52
Cumulative number of ill persons 92 M 122 M 136 M 151 M
M, million.
*Days after initial introduction.
Table 2. Simulated mean number of ill people (cumulative incidence per 100) and for TAP, the number
of antiviral courses required for various interventions and R
0
Intervention R
0
1.6 R
0
1.9 R
0
2.1 R
0
2.4
Baseline (no intervention) 32.6 43.5 48.5 53.7
Unlimited TAP (no. of courses)* 0.06 (2.8 M) 4.3 (182 M) 12.2 (418 M) 19.3 (530 M)
Dynamic vaccination (one-dose regimen)
†‡
0.7 17.7 30.1 41.1
Dynamic child-first vaccination
†‡
0.04 2.8 16.3 35.3
Dynamic vaccination (two-dose regimen)
‡§
3.2 33.8 41.1 48.5
Dynamic child-first vaccination
‡§
0.9 25.1 37.2 47.3
School closure
1.0 29.3 37.9 46.4
Local social distancing
25.1 39.2 44.6 50.3
Travel restrictions during entire simulation
32.8 44.0 48.9 54.1
Local social distancing and travel restictions
19.6 39.3 44.7 50.5
TAP,* school closure,** and social distancing** 0.02 (0.6 M) 0.07 (1.6 M) 0.14 (3.3 M) 2.8
††
(20 M)
Dynamic vaccination,
†‡
social distancing,
travel
restrictions,
and school closure**
0.04 0.2 0.6 4.5
TAP,* dynamic vaccination,
†‡
social distancing,
travel restrictions,
and school closure**
0.02 (0.3 M) 0.03 (0.7 M) 0.06 (1.4 M) 0.1 (3.0 M)
Dynamic child-first vaccination,
†‡
social distancing,
travel restrictions,
and school closure**
0.02 0.2 0.9 7.7
M, million.
*60% TAP, 7 days after pandemic alert, antiviral supply of 20 M courses unless stated.
10 million doses of a low-efficacy vaccine (single-dose regimen) per week.
Intervention continues for 25 weeks, beginning such that the first individuals treated develop an immune response on the date of the
first U.S. introduction.
§
10 million doses of a high-efficacy vaccine (two-dose regimen) per week.
Intervention starting 7 days after pandemic alert.
Reduction in long-distance travel, to 10% of normal frequency.
**Intervention starting 14 days after pandemic alert.
††
Exhausted the available supply of 20 M antiviral courses.
Germann et al. PNAS
April 11, 2006
vol. 103
no. 15
5937
MEDICAL SCIENCES
predicted to limit the national illness attack rate to 0.2%, but for
R
0
1.8, a prohibitively large 51 million courses would be required.
Fewer courses do not control the pandemic, and an overall attack
rate in excess of 10% ensues. An aggressive vaccine production and
distribution plan may also be succe ssful for R
0
1.9 (see Movie 3,
which is published as supporting information on the PNAS web
site), particularly if initially targeted at children (18). With the
exception of school closures for R
0
1.6, social distancing policie s
alone appear only to slow the pandemic without reducing its impact
as measured by morbidity (see Movie 4, which is published as
supporting information on the PNAS web site). Regardless of R
0
,
unless drastic travel restrictions are imposed, the extent or duration
of the pandemic is insensitive to details of the amount and
location(s) of introductions of pandemic influenza virus in our
simulations (see Figs. 8 and 9, which are published as supporting
information on the PNAS web site). Due to the highly mobile U.S.
population, the details of the introduction of the pandemic virus
only affect the precise geographic spread and timing of the epi-
demic peak.
For R
0
1.9, no single policy is predicted to be sufficient to
mitigate an outbreak. For such highly transmissible strains, a
combination of behavioral changes (to slow the spread) and ther-
apeutic and prophylactic measures is essential. Throughout the
range of R
0
tested, antivirals, provided they are available in suffi-
cient quantities and can be rapidly distributed, are a powerful tool
for management. On the other hand, combinations of behavioral
changes, together with a steady production of a low-efficacy vaccine
throughout the pandemic (dynamic vaccination), can also success-
fully control pandemics of viruses with all except the highest level
of transmissibility (Table 2 and Fig. 2). It is also important to note
that the e stimated benefits of preferentially vaccinating children are
offset by the closing of schools, so that although one measure or the
other is highly recommended, both together seem to offer no
additional protection. Based on this result, the high societal cost of
an extended closing of schools, requiring parents or grandparents
to remain home with young children, may be avoidable through
such a focused vaccination strategy. Similarly, our model suggests
that the combination of TAP, school closure, and social distanc-
ing can be succe ssful up to R
0
2.4, without any vaccination (see
Tables 8–10 and Figs. 10 and 11, which are published as supporting
information on the PNAS web site, for additional combinations of
intervention policies).
These projected major efforts necessary to mitigate pandemic
influenza in the U.S. make it obvious that, for the U.S. and other
c ountries, it would be optimal to control a potential pandemic
strain of influenza at the source. In the event that a pandemic
influenza virus does reach the U.S., according to our results, the
U.S. population could begin to experience a nation-wide pan-
demic within 1 month of the earliest introductions. Our simu-
lations indicate that the rapid imposition of a 90% reduction in
domestic travel would slow the vir us spread by only a few days
to weeks (depending on R
0
), without reducing the eventual size
Fig. 1. Two simulated pandemic influenza outbreaks
with R
0
1.9, initiated by the daily entry of a small
number of infected individuals through 14 major in-
ternational airports in the continental U.S. (beginning
on day 0). The tract-level prevalence of symptomatic
cases at any point in time is indicated on a logarithmic
color scale, from 0.03% (green) to 3% (red) of the
population. No mitigation strategies are used in the
baseline simulation (Left), resulting in a 43.5% attack
rate. (Right) A 60% TAP intervention begins at day 31,
or 7 days after the pandemic alert. At day 99, the
nationwide supply of 20 million antiviral courses is
exhausted, leading to a nationwide pandemic.
5938
www.pnas.orgcgidoi10.1073pnas.0601266103 Germann et al.
of the outbreak, unless other behavioral or medical responses are
introduced.
Conclusions
In this study, we regard strategies for mitigating pandemic
influenza in the U.S. as successful when they limit the national
att ack rate to that of annual influenza epidemics, 10% of the
U.S. population. All of our c onclusions about the suc cess of
mitigation strategies are based on a simplified model of disease
transmission and social contacts. Alternative models producing
the same R
0
may differ in quantitative details, but we expect the
following conclusions to hold qualit atively. To achieve the target
level of mitigation with antiviral agents alone, a very large
stockpile is likely to be required (10 million courses of oselti-
mavir for R
0
1.7, or 51 million courses for R
0
1.8, in our
simulations). For larger values of R
0
, the stockpile would have to
be prohibitively large, e.g., 182 million courses for R
0
1.9. Only
for R
0
1.6 is reasonable control predicted to be achievable with
the small currently available stockpile of 5 million courses. Our
articulated TAP strategy t argets sites of transmission for pro-
phylactic drug use (3, 11), consequently using much less drug
than if geographic regions or large groups, such as entire schools,
were targeted (3, 4). However, this TAP strategy requires the
identification of the effective sizes of the close-c ontact mixing
groups, which is much more difficult in practice than in our
assumed contact structure model. Consequently, the implemen-
t ation of TAP would require c onsiderable up-front preparation
or on-the-spot decision making, and its ef fectiveness may be
reduced by unforeseen sites or mechanisms of transmission that
are not included in our model. Nevertheless, we believe that,
even when antiviral stockpiles are small, the TAP strategy could
be quite effective in slowing virus spread until vaccination could
be implemented. (Of course, the potential emergence of an
antiviral-resistant strain should also be considered in any pan-
demic planning.)
When vaccine supplies are limited, our simulations indicate that,
at a population level, vaccinating n people with the recommended
two doses providing maximal protection is less effective at reducing
attack rates than vaccinating 2n people with single dose s, assuming
that a single dose confers roughly half the protection of a two-dose
regimen (which may or may not be an option, depending on the
particular vaccine). The relative benefits of single-dose vaccination
of 2n people and two-dose vaccination of n people are expected to
hold for prevaccination using poorly matched avian virus seed
stock, although benefits are expected to be less than those presented
here. The most effective single mitigation strategy would be a rapid
dynamic vaccination of the population, initiated within 2 weeks of
the pandemic alert, with a single dose of vaccine from the pandemic
virus. Specifically, for R
0
1.6, spread could potentially be con-
trolled if vaccine could be distributed nationally at the rate of 10
million doses per week for 25 weeks. For 1.9 R
0
2.4, single-dose
vaccination would likely require augmentation with some combi-
nation of TAP, social distancing measures, and travel restrictions to
be effective. Assuming that children remain major spreaders during
the early stages of a pandemic outbreak, as they are for interpan-
demic influenza (16), the preferential vaccination of school children
should be much more effective than random vaccination unless
schools are closed. If vaccination in advance of a pandemic were
possible using an avian seed virus, use of this poorly matched
vaccine could slow virus spread as much as possible until a well
matched vaccine based on the emergent human pandemic virus
could be deployed.
Based on the pre sent work, with the assumptions inherent in our
model and its parameters, we believe that a large stockpile of
avian-based vaccine with potential pandemic influenza antigens,
coupled with the capacity to rapidly make a better-matched vaccine
based on human strains, would be the be st strategy to mitigate
pandemic influenza. This effort needs to be coupled with a rapid
vaccine distribution system capable of distributing at least 10 million
vaccine doses per week to affected regions of the U.S.. For highly
transmissible strains (i.e., having R
0
1.9), social distancing
policies, including school closure andor travel restrictions, may
also be required to slow the epidemic spread sufficiently to enable
production and distribution of sufficient quantitie s of vaccine. If
Fig. 2. Epidemic curves (note the
logarithmic scale) demonstrating
the effectiveness of several differ-
ent mitigation strategies, as com-
pared to the baseline scenario with-
out any intervention, for different
values of R
0
. See Table 2 for details
of each intervention. In the case of
vaccination, results shown here are
for a uniform coverage of the en-
tire population with a single-dose
regimen.
Germann et al. PNAS
April 11, 2006
vol. 103
no. 15
5939
MEDICAL SCIENCES
antivirals were the preferred therapeutic defense, a stockpile of 20
million courses could be sufficient to effectively reduce national
spread of a virus with R
0
up to 1.7, provided extensive planning
andor on-the-spot decision making to distribute antivirals in a
timely fashion was carried out. If implemented for pandemic
planning, such infrastructure for stockpiling and rapid deployment
of therapeutics would lead to the more effective use of vaccine s (18)
and antiviral agents in annual influenza epidemics. On the other
hand, travel restrictions alone do not appear to be an effective
control strategy, due to the implausibly early and drastic measures
required to significantly reduce the large number of local outbreaks
that are likely to emerge around the country.
Although our simulation model was specifically designed for the
U.S., we believe that the qualitative conclusions reached here will
hold for other countries or regions with highly mobile populations.
However, for quantitative predictions to hold in settings other than
those explicitly studied here, it will be important to demonstrate a
robustness to various assumptions inherent in the model and its
parameters. (In the event of an actual pandemic, use of a model to
make quantitative predictions will require a rapid characterization
of the transmission dynamics, disease natural history, and vaccine
and antiviral efficacie s to e stimate these key model parameters.)
Then the computational tool introduced here, capturing both the
stochastic transmission proce sse s that dominate the initial stages
and final extinction of an outbreak and the detailed spatiotemporal
dynamics of infectious disease spread, can be applied to public
health que stions that cannot be effectively addressed with tradi-
tional mathematical models (5, 6). In particular, should avian
influenza continue to spread throughout the world, it will be
important to develop containment strategies, analogous to those
proposed for Southeast Asia (3, 4), that anticipate the possibility of
a human-to-human transmissible strain of H5N1 influenza emerg-
ing first in a highly mobile population such as Europe or the U.S.
Materials and Methods
Disease Transmission Model. Each class of mixing group is charac-
terized by its own set of age-dependent probabilities of person-to-
person contact of sufficient closeness and duration for transmission
of normal human influenza virus to plausibly occur within a 12-hour
period. Each of these contact probabilities is multiplied by the
probability of transmission given contact, a single multiplicative
constant that can be varied to model different R
0
values.
As
described in Tables 3 and 4, the contact probabilities were cali-
brated against total and age-specific illness attack rates of data in
past pandemics (3, 9, 17), although the se attack rate data alone do
not uniquely determine parameter estimate s. Infection of suscep-
tible individuals is modulated by the antiviral and vaccination
statuse s of both the infectious and susceptible persons. A suscep-
tible individual has a daily probability of becoming infected, accu-
mulated over hisher contacts within each of the mixing groups to
which heshe belongs (see Supporting Text for details). Age-
dependent distributions are used to determine individual disease
progression, whether an infected person becomes ill or remains
asymptomatic and, if symptomatic, when (if ever) the person
withdraws to household-only contacts.
Disease Natural History Model. Predictions of the model are sensitive
to the assumed disease course, but we can refer to past pandemics
for guidance. The disease course for infection with the 1957 and
1968 pandemic influenza viruse s and with post-1968 influenza A
viruses (17) has been fairly consistent, with an estimated mean
latent period of around 1.9 days and mean infectious period of
around 4.1 days in several modeling studies (7, 9, 11, 20). The mean
serial interval or generation time (i.e., average time between new
infection and transmission to another susceptible) is thus 4 days.
However, a recent reanalysis of incubation period and household
transmission data suggests a significantly shorter serial interval of
only 2.6 days, also consistent with viral shedding data from exper-
imental infection studies (4). On the other hand, H5N1 virus is quite
different from viruse s causing past pandemics (including the 1918
pandemic), bearing the distinctive molecular signature of highly
pathogenic avian influenza viruses, with possible implications for
the resulting disease course in humans. The limited clinical infor-
mation available to date on the disease course in individuals
infected with H5N1 virus suggests a longer time course (13).
Because H5N1 has not yet adapted for ready transmission among
humans, and disease presentation may change in conjunction with
this evolution, we focus on the midrange distributions in our model
(see Fig. 3b), with a generation time of 3.5 days (3).
Model Limitations. No seasonal or environmental effects or viral
evolution are modeled (although it would certainly be possible to
do so); we assume constant contact, transmission, and disease
course parameters throughout the U.S. for the entire duration of an
influenza season. Disease-related mortality was also neglected,
under the assumption that deaths would occur at the latter stages
of the infectious period and thus not significantly affect the spread
of disease. It is important to realize that, although we attempt to
make realistic estimates of model parameters, model validation in
the traditional sense is not possible due to the unpredictability of
viral evolution and the impossibility of documenting all cases of
influenza in any influenza season.
We are indebted to Norman Johnson, Peter Lomdahl, and Tim McPher-
son for several key contributions in the early stages of this work, and to
Mike Brown, Neil Ferguson, Brad Holian, Ed MacKerrow, Jef f Newman,
Gary Resnick, Tom Wehner, and Shufu Xu for their enc ouragement and
suggestions. We also thank Tony Redondo, Andy White, and the
Institutional Computing Prog ram at Los Alamos National L aboratory
for providing access to the necessary superc omputing resources. This
work was supported by the Department of Homeland Security through
program CBLA11MP (to T.C.G., K.K., and C.A.M.) and by National
Institute of General Medical Sciences MIDAS Grant U01-GM070749
(to I.M.L.). Los A lamos National Laboratory is operated by the Uni-
versity of California for the U.S. Department of Energy under Contract
W-7405-ENG-36.
1. Sto¨hr, K. & Esveld, M. (2004) Science 306, 2195–2196.
2. Tumpey, T. M., Basler, C. F., Aguilar, P. V., Zeng, H., Solo´rzano, A., Swayne, D. E., Cox,
N. J., Katz, J. M., Taubenberger, J. K., Palese, P., et al. (2005) Science 310, 77–80.
3. Longini, I. M., Nizam, A., Xu, S., Ungchusak, K., Hanshaoworakul, W., Cummings, D. A. T.
& Halloran, M. E. (2005) Science 309, 1083–1087.
4. Ferguson, N. M., Cummings, D. A. T., Cauchemez, S., Fraser, C., Riley, S., Meeyai, A.,
Iamsirithaworn, S. & Burke, D. S. (2005) Nature 437, 209–214.
5. Kermack, W. O. & McKendrick, A. G. (1927) Proc. R. Soc. London Ser. A 115, 700–721.
6. Anderson, R. M. & May, R. M. (1991) Infectious Diseases of Humans (Oxford Univ. Press,
Oxford, U.K.).
7. Rvachev, L. A. & Longini, I. M. (1985) Math. Biosci. 75, 3–22.
8. Hufnagel, L., Brockmann, D. & Geisel, T. (2004) Proc. Natl. Acad. Sci. USA 101, 15124–15129.
9. Elveback, L. R., Fox, J. P., Acker man, E., L angworthy, A., Boyd, M. & Gatewood, L. (1976)
Am. J. Epidemiol. 103, 152–165.
10. Halloran, M. E., Longini, I. M., Nizam, A. & Yang, Y. (2002) Science 298, 1428–1432.
11. Longini, I. M., Halloran, M. E., Nizam, A. & Yang, Y. (2004) Am. J. Epidemiol. 159, 623–633.
12. Moscona, A. (2005) N. Engl. J. Med. 353, 1363–1373.
13. Beigel, H., Farrar, H., Han, A. M., Hayden, F. G., Hyer, R., de Jong, M. D., Lochindarat,
S., Tien, N. T. K., Hien, N. T., Hien, T. T., et al. (2005) N. Engl. J. Med. 353, 1374 –1385.
14. Yang, Y., Longini, I. M. & Halloran, M. E. (2006) Appl. Stat., in press.
15. Ritzwoller, D. P. Bridges, C. B. Shetterly, S. Yamasaki, K. Kolczak, M. & France, E. K.
(2005) Pediatr ics 116, 153–159.
16. Brownstein, J. S., Kleinman, K. P. & Mandl, K. D. (2005) Am. J. Epidemiol. 162, 686693.
17. Kilbourne, E. D. (1975) The Influenza Viruses and Influenza (Academic, New York).
18. Longini, I. M. & Halloran, M. E. (2005) Am. J. Epidemiol. 161, 303–306.
19. Fraser, C., Riley, S., A nderson, R. M. & Ferguson, N. M. (2004) Proc. Natl. Acad. Sci . USA
101, 61466151.
20. Mills, C. E., Robins, J. M. & Lipsitch, M. (2004) Nature 432, 904–906.
R
0
is a difficult quantity to estimate during an actual epidemic, because it depends critically
upon the disease serial interval (or generation time) and to a somewhat lesser extent on the
relative durations of the latent and infectious periods (19, 20). Because our model assumes
particular values for these quantities, R
0
is a useful measure of transmissibility, but care needs
to be taken when comparing results for different models or epidemiological data.
5940
www.pnas.orgcgidoi10.1073pnas.0601266103 Germann et al.
... The national model also includes a fourth element, namely: d) longdistance travel data from Bureau of Transportation [5]. A "baseline" model of pandemic spread in the absence of any mitigation measures can be adapted to incorporate pharmaceutical and non-pharmaceutical interventions to model how they affect person-to-person contact rates or the susceptibility, infectiousness, or disease course within individuals [5,7]. The national-scale simulation model consists of 281 million individuals distributed among 65,334 census tracts to closely represent the actual population distribution according to publicly available 2000 US Census data [5]. ...
... Transmission within each contact group is described by a contact probability, which may depend on the age of both the infectious and susceptible persons. Individual contact probabilities were adjusted to be consistent with age-stratified attack rates and infection sources [5,7]. When schools are closed, the model assumes a 50% reduction in the number of childrelated contacts outside of the household and no change in the number of household contacts. ...
... If school closures occur as soon as a single child is diagnosed with pandemic influenza, diagnostic ratio/triggers of 5%, 10%, and 20% would require 20, 10, and 5 symptomatic children respectively within each geographic area. More information on daily contact probabilities and sources of infection used to calibrate the model can be found in the supplemental information accompanying the articles describing the transmission models [5,7]. ...
Article
Full-text available
Background Nonpharmaceutical interventions (NPIs) may be considered as part of national pandemic preparedness as a first line defense against influenza pandemics. Preemptive school closures (PSCs) are an NPI reserved for severe pandemics and are highly effective in slowing influenza spread but have unintended consequences. Methods We used results of simulated PSC impacts for a 1957-like pandemic (i.e., an influenza pandemic with a high case fatality rate) to estimate population health impacts and quantify PSC costs at the national level using three geographical scales, four closure durations, and three dismissal decision criteria (i.e., the number of cases detected to trigger closures). At the Chicago regional level, we also used results from simulated 1957-like, 1968-like, and 2009-like pandemics. Our net estimated economic impacts resulted from educational productivity costs plus loss of income associated with providing childcare during closures after netting out productivity gains from averted influenza illness based on the number of cases and deaths for each mitigation strategy. Results For the 1957-like, national-level model, estimated net PSC costs and averted cases ranged from $7.5 billion (2016 USD) averting 14.5 million cases for two-week, community-level closures to $97 billion averting 47 million cases for 12-week, county-level closures. We found that 2-week school-by-school PSCs had the lowest cost per discounted life-year gained compared to county-wide or school district–wide closures for both the national and Chicago regional-level analyses of all pandemics. The feasibility of spatiotemporally precise triggering is questionable for most locales. Theoretically, this would be an attractive early option to allow more time to assess transmissibility and severity of a novel influenza virus. However, we also found that county-wide PSCs of longer durations (8 to 12 weeks) could avert the most cases (31–47 million) and deaths (105,000–156,000); however, the net cost would be considerably greater ($88-$103 billion net of averted illness costs) for the national-level, 1957-like analysis. Conclusions We found that the net costs per death averted ($180,000-$4.2 million) for the national-level, 1957-like scenarios were generally less than the range of values recommended for regulatory impact analyses ($4.6 to 15.0 million). This suggests that the economic benefits of national-level PSC strategies could exceed the costs of these interventions during future pandemics with highly transmissible strains with high case fatality rates. In contrast, the PSC outcomes for regional models of the 1968-like and 2009-like pandemics were less likely to be cost effective; more targeted and shorter duration closures would be recommended for these pandemics.
... Even further, should the contact be established and should one of the individuals be infectious, the infection of the second individual is not a certainty, but rather an event that occurs with some probability. Computational epidiomiologists have implemented these stochastic contagions in all the modeling efforts and at different scales, from agent-based [6][7][8][9][10][11][12] to population-based [13][14][15][16][17] . In the case of agent-based models stochastic contagion events can be traced one by one, even though for practical purposes in computation sometimes they may be aggregated. ...
Article
Full-text available
Approximate numerical methods are one of the most used strategies to extract information from many-interacting-agents systems. In particular, numerical approximations are of extended use to deal with epidemic, ecological and biological models, since unbiased methods like the Gillespie algorithm can become unpractical due to high CPU time usage required. However, the use of approximations has been debated and there is no clear consensus about whether unbiased methods or biased approach is the best option. In this work, we derive scaling relations for the errors in approximations based on binomial extractions. This finding allows us to build rules to compute the optimal values of both the discretization time and number of realizations needed to compute averages with the biased method with a target precision and minimum CPU-time usage. Furthermore, we also present another rule to discern whether the unbiased method or biased approach is more efficient. Ultimately, we will show that the choice of the method should depend on the desired precision for the estimation of averages.
... 3 The influenza vaccine represents the most effective tool we have to mitigate these impacts, offering the potential to significantly reduce the incidence of illness, decrease the burden on healthcare systems and contribute to the maintenance of normal educational activities by keeping children healthy and in school. 4 Recent trends indicate an escalating incidence of influenza in China, thereby posing a substantial threat to public health. 5 6 According to national influenza surveillance data, each October marks the onset of the winter and spring influenza epidemic seasons across various regions in China. ...
Article
Full-text available
Introduction Influenza is a major public health threat, and vaccination is the most effective prevention method. However, vaccination coverage remains suboptimal. Low health literacy regarding influenza vaccination may contribute to vaccine hesitancy. This study aims to evaluate the effect of health education interventions on influenza vaccination rates and health literacy. Methods and analysis This cluster randomised controlled trial will enrol 3036 students in grades 4–5 from 20 primary schools in Dongguan City, China. Schools will be randomised to an intervention group receiving influenza vaccination health education or a control group receiving routine health education. The primary outcome is the influenza vaccination rate. Secondary outcomes include health literacy levels, influenza diagnosis rate, influenza-like illness incidence and vaccine protection rate. Data will be collected through questionnaires, influenza surveillance and self-reports at baseline and study conclusion. Ethics and dissemination Ethical approval has been sought from the Ethics Committee of the School of Public Health, Sun Yat-sen University. Findings from the study will be made accessible to both peer-reviewed journals and key stakeholders. Trial registration number NCT06048406.
... The COVID-19 pandemic has resulted in more than 395M confirmed cases and 5.7M deaths (up to Feb. 7th, 2022) [1] and has impacted the lives of more than 90% global population [2], [3]. Curbing the spread of a pandemic like COVID-19 continues to depend on the successful implementation of nonpharmaceutical interventions such as lockdowns, social distancing, shelter in place orders, contact tracing, isolation, and quarantine [4]- [7]. However, these interventions can also lead to substantial economic damage, motivating us to investigate the problem of curbing pandemic spread while minimizing the induced economic losses. ...
Article
Full-text available
As a common strategy of contagious disease containment, lockdowns inevitably have economic cost. The ongoing COVID-19 pandemic underscores the trade-off arising from public health and economic cost. An optimal lockdown policy to resolve this trade-off is desired. Here we propose a mathematical framework of pandemic control through an optimal fixed stabilizing non-uniform lockdown, where our goal is to reduce the economic activity as little as possible while decreasing the number of infected individuals at a prescribed rate. This framework allows us to efficiently compute the optimal stabilizing lockdown policy for general epidemic spread models, including both the classical SIS/SIR/SEIR models and a new model of COVID-19 transmissions. We demonstrate the power of this framework by analyzing publicly available data of inter-county travel frequencies to analyze a model of COVID-19 spread in the 62 counties of New York State. We find that an optimal stabilizing lockdown based on epidemic status in April 2020 would have reduced economic activity more stringently outside of New York City compared to within it, even though the epidemic was much more prevalent in New York City at that point. This finding holds for a variety of epidemic models and parameters from the literature, and is robust to errors in travel rates, different cost functions, and potential urban-rural transmission spread differentials.
... The ''all-or-nothing'' model assumes that a proportion of vaccinated individuals are completely protected from infection, while the remainder are not protected and maintain the susceptibility of an immune-naive individual (see, e.g., [1][2][3]). The ''leaky'' model assumes that each vaccinated individual has a reduced risk of becoming infected each time they are exposed (see, e.g., [4][5][6]). ...
Preprint
Full-text available
Behavioral interventions are a critical tool for managing novel and emerging pathogens. However, the dynamics of behavioral interventions have been difficult to measure and are poorly understood. In this study, we investigate the uptake, persistence, and waning, of behavioral interventions among two cohorts in Centre County, PA, focusing on the three years following the emergence of SARS-CoV-2. We detected clusters of behaviors that followed similar patterns of engagement and variations over time and identified some novel COVID-19 behavioral interventions that may have severely disrupted non-COVID-19 respiratory disease incidence. Additionally, we detected links between changes in risk perception and changes in behaviors over time. These findings can inform recommendations around behavioral interventions during outbreak management, including information dissemination and behavioral guidelines.
Conference Paper
The high quality of simulation systems depends on precise input parameters. The underlying population model is crucial for an agent-based simulation system that studies the spread of infectious diseases. To build the population model, the system requires a household structure (HSD) of the simulated area, including a household list with recorded age information for each member. Previous research has shown that changes in household structure significantly impact disease-spreading patterns. However, with the increasing frequency of severe infectious diseases like SARS (the year 2002), H1N1 (the year 2009), and COVID-19 (the year 2019), using outdated HSD data is inappropriate. This paper proposes a Monte-Carlo-based approach to approximate the HSD for a given year using aggregated information from a range of years. The validation of our algorithm shows good matches with different diseases. As a result, we obtained an HSD for 2020 to study the spread of new COVID-19 variants and future outbreaks.
Chapter
In the last two decades, social identity (SI) modeling and simulation have significantly advanced. They are building on and, in many cases, improving the over a half-century of validated SI experimental studies and theories. In this paper, observations on modeling and simulation of SI explore niches of additional opportunities based upon multiple perspectives: the evolution of social organisms, non-competitive theories of evolution, emergent properties of collective problem solving, advances in non-social computational modeling, epidemiological simulations, and complexity science. Based on these observations, specific recommendations are provided for expanding SI modeling and simulation. The main recommendation is to develop a general model of SI based on the observation that all social organisms share common traits, such as the innate drive to form SI or how individual states of uncertainty or stress trigger SI, but also recognize that complex species present more complex expressions of SI. Other recommendations are: SI models must accommodate that not all expressed SI traits have origins in or require higher fitness, all or many SI traits have triggers and maybe trigger thresholds that must be modeled, the inclusion of emergent group performance that may change SI behavior and strategies, and the development of a SI community model for research and realistic applications.
Article
Full-text available
We have performed parallel large-scale molecular-dynamics simulations on the QSC-machine at Los Alamos. The good scalability of the SPaSM code is demonstrated to-gether with its capability of efficient data analysis for enormous system sizes up to 19 000 416 964 particles. Furthermore, we introduce a newly-developed graphics package that renders in a very efficient parallel way a huge number of spheres necessary for the visualization of atomistic simulations. These abilities pave the way for future atomistic large-scale simulations of physical problems with system sizes on the µ-scale.
Article
Full-text available
Highly pathogenic H5N1 influenza A viruses are now endemic in avian populations in Southeast Asia, and human cases continue to accumulate. Although currently incapable of sustained human-to-human transmission, H5N1 represents a serious pandemic threat owing to the risk of a mutation or reassortment generating a virus with increased transmissibility. Identifying public health interventions that might be able to halt a pandemic in its earliest stages is therefore a priority. Here we use a simulation model of influenza transmission in Southeast Asia to evaluate the potential effectiveness of targeted mass prophylactic use of antiviral drugs as a containment strategy. Other interventions aimed at reducing population contact rates are also examined as reinforcements to an antiviral-based containment policy. We show that elimination of a nascent pandemic may be feasible using a combination of geographically targeted prophylaxis and social distancing measures, if the basic reproduction number of the new virus is below 1.8. We predict that a stockpile of 3 million courses of antiviral drugs should be sufficient for elimination. Policy effectiveness depends critically on how quickly clinical cases are diagnosed and the speed with which antiviral drugs can be distributed.
Article
Full-text available
Multimillion-atom molecular-dynamics simulations are used to investigate the shock-induced phase transformation of solid iron. Above a critical shock strength, many small close-packed grains nucleate in the shock-compressed body-centered cubic crystal growing on a picosecond time scale to form larger, energetically favored grains. A split two-wave shock structure is observed immediately above this threshold, with an elastic precursor ahead of the lagging transformation wave. For even higher shock strengths, a single, overdriven wave is obtained. The dynamics and orientation of the developing close-packed grains depend on the shock strength and especially on the crystallographic shock direction. Orientational relations between the unshocked and shocked regions are similar to those found for the temperature-driven martensitic transformation in iron and its alloys.
Article
This book combines mathematical models with extensive use of epidemiological and other data, to achieve a better understanding of the overall dynamics of populations of pathogens or parasites and their human hosts. The authors thus provide an analytical framework for evaluating public health strategies aimed at controlling or eradicating particular infections. With rising concern for programmes of primary health care against such diseases as measles, malaria, river blindness, sleeping sickness, and schistosomiasis in developing countries, and the advent of HIV/AIDS and other `emerging viruses', such a framework is increasingly important. Throughout, the mathematics is used as a tool for thinking clearly about fundamental and applied problems relating to infectious diseases. The book is divided into two major parts, one dealing with microparasites (viruses, bacteria, and protozoans) and the other with macroparasites (helminths and parasitic arthropods). Each part begins with simple models, developed in a biologically intuitive way, and then goes on to develop more complicated and realistic models as tools for public health planning. A major contribution by two of the leaders in the field, this book synthesizes previous work in this rapidly growing area with much new material, combining work scattered between the ecological and medical literature.
Article
A mathematical model is presented for forecasting the global spread of influenza based on information from the initial city in the transportation network to experience the disease. This model represents the natural extension of nearly 20 years of Soviet work on modeling the geographic spread of influenza in the U.S.S.R. and Bulgaria. However, the work presented in this paper is the first attempt at applying the methods on a global scale. A description of the model formulation is given along with a method for estimating critical parameters. The model is then applied to forecasting the global spread of the “Hong Kong” pandemic of 1968-1969 based on estimated parameters from Hong Kong, the initial city to report the appearance of the then new strain of influenza A. The forecast is shown to reproduce the general time-space spread of the actual epidemic as documented by World Health Organization sources.
Article
A stochastic simulation epidemic model based on discrete time intervals and appropriate for any infectious agent spread by person-to-person contacts is presented. The population is highly structured, allowing for five age groups and for subgrouping mixing in families, neighborhoods, schools, and preschool playgroups as well as total community mixing. With proper choice of relative susceptibility by age, length of latency and infectivity periods, pathogenicity and withdrawal patterns, and the relative infectiousness of silent infections, the model becomes highly agent-specific. The model includes flexible immunization routines and variable vaccine response patterns. The model is applied to the 1957 Asian and 1968 Hong Kong pandemic strains of influenza A. The results of several schedules of immunization of school children are presented and compared for the two strains.
Article
The need for a planned response to a deliberate introduction of smallpox has recently become urgent. We constructed a stochastic simulator of the spread of smallpox in structured communities to compare the effectiveness of mass vaccination versus targeted vaccination of close contacts of cases. Mass vaccination before smallpox introduction or immediately after the first cases was more effective than targeted vaccination in preventing and containing epidemics if there was no prior herd immunity (that is, no prior immunologic protection within the population). The effectiveness of postrelease targeted and mass vaccinations increased if we assumed that there was residual immunity in adults vaccinated before 1972, but the effectiveness of targeted vaccination increased more than that of mass vaccination. Under all scenarios, targeted vaccination prevented more cases per dose of vaccine than did mass vaccination. Although further research with larger-scale structured models is needed, our results suggest that increasing herd immunity, perhaps with a combination of preemptive voluntary vaccination and vaccination of first responders, could enhance the effectiveness of postattack intervention. It could also help targeted vaccination be more competitive with mass vaccination at both preventing and containing a deliberate introduction of smallpox.
Article
For the first wave of pandemic influenza or a bioterrorist influenza attack, antiviral agents would be one of the few options to contain the epidemic in the United States until adequate supplies of vaccine were available. The authors use stochastic epidemic simulations to investigate the effectiveness of targeted antiviral prophylaxis to contain influenza. In this strategy, close contacts of suspected index influenza cases take antiviral agents prophylactically. The authors compare targeted antiviral prophylaxis with vaccination strategies. They model an influenza pandemic or bioterrorist attack for an agent similar to influenza A virus (H2N2) that caused the Asian influenza pandemic of 1957-1958. In the absence of intervention, the model predicts an influenza illness attack rate of 33% of the population (95% confidence interval (CI): 30, 37) and an influenza death rate of 0.58 deaths/1,000 persons (95% Cl: 0.4, 0.8). With the use of targeted antiviral prophylaxis, if 80% of the exposed persons maintained prophylaxis for up to 8 weeks, the epidemic would be contained, and the model predicts a reduction to an illness attack rate of 2% (95% Cl: 0.2, 16) and a death rate of 0.04 deaths/1,000 persons (95% CI: 0.0003, 0.25). Such antiviral prophylaxis is nearly as effective as vaccinating 80% of the population. Vaccinating 80% of the children aged less than 19 years is almost as effective as vaccinating 80% of the population. Targeted antiviral prophylaxis has potential as an effective measure for containing influenza until adequate quantities of vaccine are available.