plague: past, present, and future by mike begon et alt

Upload: renato-sala

Post on 03-Jun-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    1/13

    Centre for Ecological and Evolutionary Synthesis

    University of Oslo

    Memo: Work by the Centre for Ecological and Evolutionary Synthesis (CEES) in Kazakhstan,

    Georgia and Azerbaijan (and China)

    Author: Nils Chr. Stenseth (Professor and Chair of CEES (http://www.cees.uio.no/); also president

    elect of the Norwegian Academy of Science and Letters;http://english.dnva.no/)

    Date: 22.04.12

    The Centre for Ecological and Evolutionary Synthesis(CEES) has during the last 10 years served as both chair

    and partner in collaborative projects on the dynamics of the plague bacterium (Yersinia pestis, the causative

    agent of plague, including the Black Death); plague continues to be a health problem and threat

    (bioterrorism) around the world, not the least in Asia and Africa but also North America). Most of ourcollaborative work on the plague system has been done out of Kazakhstan (Kazakh Scientific Centre for

    Quarantine and Zoonotic Diseases, Almaty); recently we have been extending this to China (working both

    with the Chinese Academy of Science(Beijing) and the Chinese Academy of Military Medical Sciences

    (Beijing)) as well as to the Caucasus region (Georgia and Azerbaijan)1.

    Our collaborative work within Caucusus is associated with the new state-of-the-art pathogens facility

    in Georgia (in Tbilisi), the Georgian Central Public Health Reference Laboratory(CPHRL) built by the U.S.

    Dept. of Defense, Defense Threat Reduction Agency (DTRA)2. The facility is meant to be a reference

    laboratory for strains collected in Georgia and a base for scientific work on those pathogens. CEES is at the

    international forefront in combining ecological, evolutionary and genomics work on such systems with a

    focus on the dynamics in the wild and thus is a complementary research unit to those institutions that take

    a human focus in their studies on these diseases. Being involved in the development of these facilities gives

    CEES a unique opportunity to contribute with new insights on the ecology and evolution of these bacterial

    systems3.

    Having published several highly visible publications in top international journals (see a list of

    publications at the end of this document) together with Kazakh and Chinese scientists within an open

    collaborative network, provides CEES and Norway with unique opportunities for further collaborative

    work in Central Asia and the Caucasus4. Norwegian involvement in such work in this part of the world has

    many obvious benefits, including positive visibility as well as getting first hand insight into such health and

    threat issues.

    1

    During Soviet times anti-plague stations were set-up around the country in areas endemic to the plague bacterium. These stationsmonitored and recorded ecological and meteorological data associated with plague in hopes to better understand and control the

    disease. In Kazakhstan, data was made available to us and our collaborators from the Pre-Balkash plague focus. From these data we

    were able to describe conditions that allow plague to reach epidemic proportions. Currently we are in the process of initiating

    collaborative projects to localize where the bacterium is between epizootics, a central question in eco-epidemiology: this work may

    be carried out in Kazakhstan or some of the Caucasus countries.2NC Stenseth serves as a member of the Scientific Advisory Council of CPHRL and these facilities.

    3In Azerbaijan, we are setting up similar collaborations; we will have access to the Georgian facility to conduct our experimental

    work, and also to ecological data from their three separate and unique systems. Azerbaijan is a unique region: small in area yet

    containing very unique foci for both disease strains as well as for the rodents that manifest the disease.4Building upon our work on the plague system and our general ecological and evolutionary platform, we at CEES are currently in the

    process of further developing our work on the ecology and evolution of human infectious diseases with an environmental reservoir,

    with a focus on Yersinia pestis (Plague), Bacillus anthracis(Anthrax), Borrelia burgdorferi(Lyme disease) and Francisella tularensis

    (Tularemia). Our overall objective is to develop an overarching view on the geographical distributions and genetic variations of thesebacteria. In collaboration with several other international research institutions, the CEES will focus on the eco-evolutionary dynamics

    of these sylvatic-based epidemiological systems work which will be of great value to those focusing on the human-linked part of

    h d h ll b d f h l l d l k

    http://www.cees.uio.no/http://www.cees.uio.no/http://www.cees.uio.no/http://english.dnva.no/http://english.dnva.no/http://english.dnva.no/http://english.dnva.no/http://www.cees.uio.no/
  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    2/13

    Plague-themed papers authored/co-authored by Stenseth/CEES (two of the papers are included

    at the end of the document).

    Ben Ari, T., Gershunov, A., Gage, K.L., Snll, T., Ettestad, P., Kausrud, K.L. & Stenseth, N.C. (2008) Human plague in the USA: the

    importance of regional and local climate. Biology Letters,4,737-740.

    Ben Ari, T., Gershunov, A., Rouyer, T., Cazelles, B., Gage, K. & Stenseth, N.C. (2010) Interannual Variability of Human Plague

    Occurrence in the Western United States Explained by Tropical and North Pacific Ocean Climate Variability.American

    Journal of Tropical Medicine and Hygiene.,83,624632.Ben Ari, T., Neerinckx, S., Gage, K., Kreppel, K., Laudisoit, A., Leirs, H. & Stenseth, N.C. 2011. Plague and climate: scales matter. PLoS

    Pathogen7, e1002160. (Enclosed below as Enclosure 2).

    Ben Ari, T., Neerinckx, S., Agier, L., Cazelles, B., Xu, L., Zhang, Z., Fang, X., Wang, S., Liu, Q. & Stenseth, N.C. 2012. Indentifying

    Enzootic Plague Foci in China Using Long-term Epidemiological data.Proceedings of National Academy of Science,

    Washington(in press).

    Davis, S., Begon, M., De Bruyn, L., Ageyev, V.S., Klassovskiy, N.L., Pole, S.B., Viljugrein, H., Stenseth, N.C. & Leirs, H. (2004) Predictive

    thresholds for plague in Kazakhstan. Science,304,736-738.

    Davis, S., Leirs, H., Viljugrein, H., Stenseth, N.C., De Bruyn, L., Klassovskiy, N., Ageyev, V. & Begon, M. (2007) Empirical assessment of

    a threshold model for sylvatic plague.Journal of the Royal Society Interface,4,649-657.

    Easterday, W.R., Kausrud, K.L., Star, B., Heier, L., Haley, B.J., Ageyev, V., Colwell, R.R. & Stenseth, N.C. 2012. An additional step in the

    transmission of Yersinia pestis. The ISME journal, 6, 231236.

    Frigessi, A., Holden, M., Marshall, C., Viljugrein, H., Stenseth, N.C., Holden, L., Ageyev, V. & Klassovskiy, N.L. (2005) Bayesian

    population dynamics of interacting species: great gerbils and fleas in Kazakhstan. Biometrics,61,230-238.

    Heier, L., Davis, S.A., Storvik, G.O., Viljugrein, H., Arayev, V., Klassovskaya, E. & Stenseth, N.C. 2011. Re-emergence, Spread and

    Persistence of Sylvatic Plague in Kazakhstan: Comparison of invasion vs. persistence, and reemergence vs. spread.

    Proceedings of the Royal Society of London, B 278, 2915-2923..

    Kausrud, K.L., Viljugrein, H., Frigessi, A., Begon, M., Davis, S., Leirs, H., Dubyanskiy, V. & Stenseth, N.C. (2007) Climatically driven

    synchrony of gerbil populations allows large-scale plague outbreaks. Proceedings of the Royal Society B: Biological Sciences,

    274,1963-1969.

    Kausrud, K.L., Begon, M., Ari, T.B., Viljugrein, H., Esper, J., Bntgen, U., Leirs, H., Junge, C., Yang, B., Yang, M. & Stenseth, N.C. (2010)

    Modeling the epidemiological history of plague in Central Asia: Palaeoclimatic forcing on a disease system over the past

    millennium. BMC biology,8 (112).

    Klassovskaya, E.V., Davis, S., Leirs, H., Stenseth, N.C., Begon, M., Dubyanskiy, V.M., Burdelov, L.A., Ageyev, V.S., Sapozhnikov, V.I. and

    Alipbaev, A.K. 2007. Threshold model for predicting of plague epizootics in one locality of southern Pre-Balkhash and

    testing the model in 2004-2006. Quarantine and zoonozis disease in Kazakhstan, 1-2, 18-29.

    Park, S., Chan, K.S., Viljugrein, H., Nekrassova, L., Suleimenov, B., Ageyev, V.S., Klassovskiy, N.L., Pole, S.B. & Stenseth, N.C. (2007)

    Statistical analysis of the dynamics of antibody loss to a disease-causing agent: plague in natural populations of greatgerbils as an example.Journal of the Royal Society Interface,4,57-64.

    Samia, N.I., Chan, K.-S. & Stenseth, N.C. (2007) A generalized threshold mixed model for analyzing nonnormal nonlinear time series,

    with application to plague in Kazakhstan. Biometrika,94,101-118.

    Samia,N., Kausrud, K., Heesterbeek,H., Ageyev, V., Begon, M., Chan, K.-S. & Stenseth, N.C. 2011. The dynamics of the plague-wildlife-

    human system in Central Asia is controlled by two epidemiological thresholds. Proceedings of National Academy of Science,

    Washington 108, 14527-14532.

    Snll, T., Benestad, R. & Stenseth, N. (2009) Expected future plague levels in a wildlife host under different scenarios of climate

    change. Global Change Biology,15,500-507.

    Stenseth, N.C., Samia, N.I., Viljugrein, H., Kausrud, K.L., Begon, M., Davis, S., Leirs, H., Dubyanskiy, V.M., Esper, J., Ageyev, V.S.,

    Klassovskiy, N.L., Pole, S.B. & Chan, K.S. (2006) Plague dynamics are driven by climate variation. Proceedings of the National

    Academy of Sciences,103,13110-13115.

    Stenseth, N.C. (2008) Plague Through History. Science,321,773-774.

    Stenseth, N.C., Atshabar, B.B., Begon, M., Belmain, S.R., Bertherat, E., Carniel, E., Gage, K.L., Leirs, H. & Rahalison, L. (2008) Plague:

    past, present, and future. PLoS Medicine,5,9-13. (Enclosed below as Enclosure 1)

    Stesneth, N.C. 2008. Plague and climate. In: Institute of Medicine (ed) Global Climate Change and Extreme Weather Events:

    Understanding the Contributions to the Emergence, Re-emergence, and Spread of Infectious Disease Washington, DC: The

    National Academies Press. Pp. 130-145.

    Stenseth, N.C. & Ben Ari, T. 2010. La Peste: histoire et cologie dune maladie nglige [Plague: the history and ecology of a

    neglected disease]. In: Gauthier-Clerc, M. & Thomas, F. (eds.)Ecologie de la Sant et Biodiversit, De Boeck Universit,

    Bruxelles 239-248.

    Xu, L., Qiyong, L., Stige, L.C., Ben Ari, T., Fang, X., Chan, K.-S., Wang, S., Stenseth, N.C., Zhang, Z. 2011. Nonlinear effect of climate on

    plague during the third pandemic in China. Proceedings of National Academy of Science, Washington 108, 10214-10219.

    Zhang, Z., Li, Z., Tao, Y., Chen, M., Wen, X., Xu, L., Tian, H. & Stenseth, N.C. 2007. Relationship between increase rate of human

    plague in China and global climate index as revealed by cross-spectral analysis and cross-wavelet analysis. Integrative

    Zoology2, 144-153.

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    3/13PLoS Medicine | www.plosmedicine.org 0009 January 2008 | Volume 5 | Issue 1 | e3

    Neglected Diseases

    Recent experience with SARS (severe acute respiratory

    syndrome) [1] and avian flu shows that thepublic and political response to threats from new

    anthropozoonoses can be near-hysteria. This can readily makeus forget more classical animal-borne diseases, such as plague(Box 1).

    Three recent international meetings on plague (Box 2)concluded that: (1) it should be re-emphasised that theplague bacillus (Yersinia pestis) still causes several thousandhuman cases per year [2,3] (Figure 1); (2) locally perceivedrisks far outstrip the objective risk based purely on thenumber of cases [2]; (3) climate change might increase therisk of plague outbreaks where plague is currently endemicand new plague areas might arise [2,4]; (4) remarkably little

    is known about the dynamics of plague in its natural reservoirsand hence about changing risks for humans [5]; and,therefore, (5) plague should be taken much more seriously bythe international community than appears to be the case.

    The Plague Eco-Epidemiological System

    The plague bacillus causes a rapidly progressing, seriousillness that in its bubonic form is likely to be fatal (40%70%mortality). Without prompt antibiotic treatment, pneumonicand septicaemic plague are virtually always fatal. For thesereasonsY. pestisis considered one of the most pathogenicbacteria for humans.

    Yersinia pestisis transmitted by fleas, while the other twospecies of Yersinia known to be pathogenic for humans

    (Y. enterocolitica and Y. pseudotuberculosis) are transmittedby the faecaloral route and cause intestinal symptoms ofmoderate intensity. Yersinia pestis is believed to be a cloneof Y. pseudotuberculosis that emerged within the last 1,500 to20,000 years [6,7]. This divergence was characterised by theacquisition of a few genetic elements; more particularly, twoplasmids that play a key role in flea-borne transmission [8,9].The exceptional pathogenicity of Y. pestis compared to theenteropathogenic species may be explained by its new modeof transmission. Indeed, the only means for this bacteriumto be transferred to new hosts is through septicaemia,

    which allows the bacteria present in the bloodstream to beefficiently taken up by the flea during its blood meal [10].

    Soon after Yersins identification of the plague bacillus[11], it became clear that urban outbreaks were linked totransmission among commensal rats and their fleas. In thisclassic urban-plague scenario, infected rats (transported,for example, by ships) arrive in a new city and transmit theinfection to local house rats and their fleas, which then serveas sources of human infection. Occasionally, humans developa pneumonic form of plague that is then directly transmittedfrom human to human through respiratory droplets.

    The epidemiology of plague, however, is much more

    complicated than this urban-plague scenario suggests,involving several othermore likelypathways oftransmission (Box 3 and Figure 2). This complicatedepidemiology necessitates a reconsideration of plagueecology.

    The Past

    Plague has given rise to at least three major pandemics.The first (the Justinian plague) spread around theMediterranean Sea in the 6th century AD, the second (theBlack Death) started in Europe in the 14th century andrecurred intermittently for more than 300 years, and the thirdstarted in China during the middle of the 19th century and

    spread throughout the world. Purportedly, each pandemicwas caused by a different biovar of Y. pestis, respectively,Antiqua (still found in Africa and Central Asia), Medievalis

    Plague: Past, Present, and FutureNils Chr. Stenseth*, Bakyt B. Atshabar, Mike Begon, Steven R. Belmain, Eric Bertherat, Elisabeth Carniel, Kenneth L. Gage,

    Herwig Leirs, Lila Rahalison

    Funding:This article emerged from a meeting at the Norwegian Academy ofScience and Letters in Oslo, a meeting funded by that academy, the University ofOslo, and by the European Commission INCO-COPERNICUS programme (STEPICA;ICA 2-CT2000-10046). The organizations played no further role, including no role inthe submission of this article.

    Competing Interests: NCS declares that he is the project leader of the researchcontract number ICA 2-CT2000-10046 granted from the European CommissionINCO-COPERNICUS programme for the STEPICA project that has contributedresearch on the ecology of plague in Central Asia; BBA, MB, EC, and HL participatedin the same project. MB declares that he is the project leader of the researchcontract number 063576/Z/01/Z granted from the Wellcome Trust. SRB declaresthat he is the technical coordinator and project leader of research contract number

    ICA4 2002 10056 granted from the European Commission INCO-DEV programmefor the RATZOOMAN project that has contributed to research on the ecology ofplague in southern Africa. HL declares that he participated in the same project andthat he is holding a research grant from the Fund for Scientific Research (Flanders)to investigate the distribution of plague in Africa.

    Citation: Stenseth NC, Atshabar BB, Begon M, Belmain SR, Bertherat E, et al.(2008) Plague: Past, present, and future. PLoS Med 5(1): e3. doi:10.1371/journal.pmed.0050003

    Copyright:This is an open-access article distributed under the terms of theCreative Commons Public Domain declaration, which stipulates that, once placedin the public domain, this work may be freely reproduced, distributed, transmitted,modified, built upon, or otherwise used by anyone for any lawful purpose.

    Abbreviations: DRC, Democratic Republic of Congo; WHO, World HealthOrganization

    Nils Chr. Stenseth is with the Centre for Ecological and Evolutionary Synthesis,Department of Biology, University of Oslo, Oslo, Norway. Bakyt B. Atshabar iswith the Kazakh Scientific Centre for Quarantine and Zoonotic Diseases, Almaty,Kazakhstan. Mike Begon is with the School of Biological Sciences, BiosciencesBuilding, University of Liverpool, Liverpool, United Kingdom. Steven R. Belmainis with the Natural Resources Institute, University of Greenwich, Kent, UnitedKingdom. Eric Bertherat is with Epidemic Readiness and Interventions, Departmentof Epidemic and Pandemic Alert and Response, World Health Organization, Geneva,Switzerland. Elisabeth Carniel is with the Yersinia Research Unit, Institut Pasteur,Paris, France. Kenneth L. Gage is with the Flea-Borne Diseases Laboratory, Centersfor Disease Control and Prevention, Fort Collins, Colorado, United States of America.Herwig Leirs is with the Evolutionary Ecology Group, University of Antwerp,Antwerpen, Belgium, and the Danish Pest Infestation Laboratory, Universityof Aarhus, Faculty of Agricultural Sciences, Department of Integrated PestManagement, Kongens Lyngby, Denmark. Lila Rahalison is with the LaboratoireCentral de la Peste, Institut Pasteur de Madagascar, Antananarivo, Madagascar.

    * To whom correspondence should be addressed. E-mail: [email protected]

    The Neglected Diseases section focuses attention either on a specific disease or

    describes a novel strategy for approaching neglected health issues in general.

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    4/13PLoS Medicine | www.plosmedicine.org 0010 January 2008 | Volume 5 | Issue 1 | e3

    (currently limited to Central Asia), and Orientalis (almostworldwide in its distribution) [12,13].

    The Black Death decimated Medieval Europe, andhad a major impact on the continents socio-economicdevelopment, culture, art, religion, and politics [14,15].Several high-profile critiques have questioned whetherthe Black Death was indeed caused by Y. pestis[16,17].Proponents of both sides of the debate came together at last

    years Oslo and London meetings (Box 2). It was generallyaccepted that the epidemiology of Black Death plague, asreflected in historical records, did not always match theclassical rat-flea-human plague cycle. However, the reported

    medical symptoms were very similar during each pandemic. Inaddition, the international experience of modern-day plaguerepresented at the Oslo meeting made it clear that classicalplague epidemiology is only one of several possibilities toexplain the Black Death, and may not even be the most typicalof the different plague ecology systems [18]. Discovery of Y.

    pestisgenetic material in those who died from the Black Deathand are buried in medieval graves [19] further supports the

    view that Y. pestiswas the causative agent of the Black Death.

    The Present

    Given this history, plague is often classified as a problem ofthe past. However, it remains a current threat in many parts

    of the world (Figure 1A), particularly in Africa, where boththe number of cases (Figure 1B) and the number of countriesreporting plague (Figure 1C) have increased during recentdecades. Following the reappearance of plague during the1990s in several countries, plague has been categorised as are-emerging disease [20,21].

    Plague is endemic in a variety of wildlife rodent speciesworldwide in a wide range of natural habitats, with commensalrats only sometimes playing a role as liaison hosts, carryingplague between the sylvatic reservoir and people. Variousother animal-to-human transmission pathways have beendocumented. Human plague may be contracted from (1)

    being bitten by the fleas of wildlife rodent species in ruralsettings (e.g., in the south-western United States [22,23])or of commensal rodents that move freely between villagesand the forest habitats occupied by reservoir hosts (e.g., inTanzania); rodents movements have become more frequentas human activity has fragmented the forest [24]; (2) eatinginfected animals such as guinea pigs in Peru and Ecuador[25,26] or camels that contract the disease from rodent fleasin Central Asia and the Middle East [2731]; or (3) handlingcats infected through the consumption of plague-infectedrodents in Africa or the United States [3234]. Human-to-human transmission also occurs, either directly throughrespiratory droplets or indirectly via flea bites [3537].

    doi:10.1371/journal.pmed.0050003.g001

    Figure 1.The Global Distribution of Plague(A) Map showing countries with known presence of plague in wild reservoir species (red) (after [3]). For US only the mainland below 50 N is shown. (B)Annual number of human plague cases over different continents, reported to WHO in the period 19542005. (C) Cumulative number of countries thatreported plague to WHO since 1954.

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    5/13PLoS Medicine | www.plosmedicine.org 0011 January 2008 | Volume 5 | Issue 1 | e3

    Over the last 20 years, there have been 1,000 to 5,000human cases of plague and 100 to 200 deaths reported to the

    World Health Organization (WHO) each year [38]. However,

    because of poor diagnostic facilities and underreporting,the number of cases is almost certainly much higher. Overthe years, there has been a major shift in cases from Asiato Africa (Figure 1B), with more than 90% of all casesand deaths in the last five years occurring in Madagascar,Tanzania, Mozambique, Malawi, Uganda, and the DemocraticRepublic of the Congo (DRC). Most are cases of bubonicplague contracted through contact with infected rodents andfleas. However, outbreaks of pneumonic plague still occur:the most recent large one was in October and November2006 in DRC, with hundreds of suspected cases [39], anda smaller outbreak arose just across the border in nearbyUganda in February 2007 ([40]). In December 2004 there

    was a pneumonic outbreak in a miners camp in DRC,probably imported by an infected human who had travelledfrom an endemic zone. One pneumonic case even arrivedin Kisangani, a large city several hundred kilometres away[41,42]. So even rapid-reaction medical teams may not besufficient to stop plague from spreading quickly over longdistances before it is detected and managed.

    Africa is particularly at risk for a number of reasons. Poorrural communities typically live in close proximity to rodents,

    which are widely hunted and eaten in many plague-endemicareas. Superstitions, lack of money, and distance from healthfacilities often lead to delays in seeking health care andreceiving treatment. The public health system in large parts

    of Africa is poorly organised and equipped, and political crisisand social disorganisation impede improvements. Finally,anthropogenic changes to the landscape and to patterns ofhuman mobility are increasingly favouring contact betweenplague-reservoir and peri-domestic rodents, and betweenpeople from plague-endemic and previously unaffectedregions.

    Looking to the Future

    Plague cannot be eradicated, since it is widespread inwildlife rodent reservoirs. Hence, there is a critical need tounderstand how human risks are affected by the dynamicsof these wildlife reservoirs. For example, the likelihood ofa plague outbreak in North American and Central Asianrodents, and the resulting risk to humans, is known to beaffected by climate [43,44]. Recent analysis of data fromKazakhstan [45] shows that warmer springs and wettersummers increase the prevalence of plague in its main host,the great gerbil. Such environmental conditions also seemto have prevailed during the emergence of the Second andThird Pandemics [46,47]conditions that might becomemore common in the future [48].

    Although the number of human cases of plague is relativelylow, it would be a mistake to overlook its threat to humanity,because of the diseases inherent communicability, rapidspread, rapid clinical course, and high mortality if leftuntreated. A plague outbreak may also cause widespreadpanic, as occurred in India in 1994 when a relatively smalloutbreak, with 50 deaths, was reported in the city of Surat[49]. This led to a nationwide collapse in tourism and trade,

    with an estimated cost of US$600 million [50].Outbreaks are usually tackled with a fire-fighting

    approach. Teams move into an infected area to kill fleas withinsecticides, treat human cases, and give chemoprophylaxisto exposed people. Many experts have argued that this crisis-management approach is insufficient as the outbreak is likely

    to be on the wane by the time action is taken. Informed,pre-emptive decisions about plague management andprevention before outbreaks occur would certainly be moresustainable and cost-beneficial. There has been some recentprogress, such as development of rapid diagnosis tools [51],some challenging of accepted dogma about the dynamicsof sylvatic plague in the United States [52] and in Central

    Asia [53], and the identification of predictive critical rodentabundance thresholds for plague in Kazakhstan [54]. What

    Box 1. The Plague

    The causative bacterium (Yersinia pestis) was discoveredby Yersin in 1894 [11] (see also [63]). Case-fatality ratio variesfrom 30% to 100%, if left untreated. Plague is endemic in manycountries in the Americas, Asia, and Africa. More than 90% ofcases are currently being reported from Africa.

    Clinical presentation:After an incubation period of 37 days,patients typically experience a sudden onset of fever, chills,

    headaches, body aches, weakness, vomiting, and nausea. Clinicalplague infection manifests itself in three forms, depending onthe route of infection: bubonic, septicaemic, and pneumonic.

    The bubonic form is the most common, resulting from thebite of an infected flea. The pneumonic form initially is directlytransmitted from human to human via inhalation of infectedrespiratory droplets.

    Treatment:Rapid diagnosis and treatment are essential toreduce the risk of complications and death. Streptomycin,tetracyclines, and sulfonamides are the standard treatment.Gentamicin and fluoroquinolones may represent alternativeswhen the above antibiotics are not available. Patients withpneumonic plague must be isolated to avoid respiratorytransmission.

    Challenges ahead:Biological diagnosis of plague remains achallenge because most human cases appear in remote areaswith scarce laboratory resources. So far, the main confirmationtechniques were based on the isolation of Y. pestis(requiring aminimum of 4 days). The recent development of rapid diagnostictests, now considered a confirmation method in endemicareas, opens new possibilities in terms of surveillance and casemanagement.

    Box 2. Recent International Meetings on Plague

    A meeting on plague in the present, past, and future was held

    by the Academy of Science and Letters, Oslo, Norway (http://www.cees.no/oslo-plague-meeting).

    A workshop focusing on the comparison of the Black Death andmodern plague was organised by the Wellcome Trust Centrefor the History of Medicine, University College, London (http://www.ucl.ac.uk/histmed/).

    The World Health Organization convened an expertmeeting in Antananarivo, Madagascar updating clinicalplague definitions, diagnostic methods, vaccines, andantimicrobial therapy, as well as reservoir and vectorcontrol strategies (http://www.who.int/csr/disease/plague/interregionalmeeting2006/en/index.html).

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    6/13PLoS Medicine | www.plosmedicine.org 0012 January 2008 | Volume 5 | Issue 1 | e3

    is striking, though, is our lack of understanding of this high-profile disease in even the best-studied foci, particularly in

    Africa: often, we do not even know the natural reservoir ofthe bacilli.

    The capacity of the plague bacillus to form permanentfoci under highly diverse ecological conditions attests to itsextraordinary adaptability. During its emergence in Central

    Asia, Y. pestisaccumulated copies of insertion sequencesrendering its genome highly plastic [55]. The capacity toundergo genomic rearrangements may thus be an efficientmeans for the plague bacillus to adapt to new ecologicalniches.Yersinia pestiswas, furthermore, recently shown to beable to acquire antibiotic resistance plasmids under naturalconditions [56,57], probably during its transit in the fleamidgut [58]. Of great concern is the recent observation ofthe presence of multidrug-resistant plasmids, almost identicalto those acquired by Y. pestis,in a variety of enterobacteriaisolated from retail meat products in the United States [59].This bacterial reservoir of mobile resistance determinantsis probably widespread globally and has the potential todisseminate to human and zoonotic bacterial pathogens,including Y. pestis. Obviously, the emergence and spread ofmulti-resistant strains of Y. pestiswould represent a majorthreat to human health.

    Finally, we should not overlook the fact that plague hasbeen weaponised throughout history, from catapultingcorpses over city walls, to dropping infected fleas fromairplanes, to refined modern aerosol formulations [60,61].The weaponisation research on plague carried out frombefore World War II until the 1990s fuelled a fear ofbiological warfare that may actually have stimulated researchinto surveillance and response strategies. More recently,however, fear of small-scale bioterrorism and the desire ofgovernmental authorities to more fully control all access toplague materials risks stifling the research on plague ecology,epidemiology, and pathophysiology that is required toimprove its clinical management in endemic areas. Terrorist

    use of an aerosol released in a confined space could resultin significant mortality and widespread panic [60,61], andno one would wish plague weaponisation knowledge andmaterial to fall into terrorist hands. However, the need forscientifically sound studies on the dynamics of infection,transmission, outbreak management, and improvedsurveillance and monitoring systems has never been greater.

    Plague may not match the so-called big three diseases(malaria, HIV/AIDS, tuberculosis; see for example [62])in numbers of current cases, but it far exceeds them in

    pathogenicity and rapid spread under the right conditions.It is easy to forget plague in the 21st century, seeing it as ahistorical curiosity. But in our opinion, plague should notbe relegated to the sidelines. It remains a poorly understoodthreat that we cannot afford to ignore.

    Supporting Information

    Alternative Language Text S1.

    Translation of article into French by EB

    Found at doi:10.1371/journal.pmed.0050003.sd001 (303 KB PDF)

    Alternative Language Text S2.

    Translation of article into Russian by BBA

    Found at doi:10.1371/journal.pmed.0050003.sd002 (356 KB PDF)

    References1. McLean AR, May RM, Pattison J, Weiss RA (2005) SARS. A case study in

    emerging infections. Oxford: Oxford University Press.2. World Health Organization (2003) Plague. Wkly Epidemiol Rec 78, 253-260.3. World Health Organization (2005) Plague. Wkly Epidemiol Rec 80 138-140.4. Stapp P, Antolin MF, Ball M (2004) Patterns of extinction in prairie dog

    metapopulations: plague outbreaks follow El Nino events. Front EcolEnviron 2: 235240.

    5. Gage KL, Kosoy MY (2005) Natural history of plague: perspectives frommore than a century of research. Annu Rev Entomol 50: 505528.

    6. Achtman M, Zurth K, Morelli G, Torrea G, Guiyoule A, et al. (1999)Yersinia pestis, the cause of plague, is a recently emerged clone of Yersinia

    pseudotuberculosis. Proc Natl Acad Sci U S A 96: 14043-14048.7. Achtman M, Morelli G, Zhu P, Wirth T, Diehl I, et al. (2004)

    Microevolution and history of the plague bacillus, Yersinia pestis. Proc NatlAcad Sci U S A 101: 1783717842.

    Box 3. The Plague Bacterium within the EcologicalWeb

    Maintenance of plague foci depends on a whole suite ofrodent hosts and their associated fleas (Figure 2A). Underfavourable conditions, the plague bacillus might survive inthe environment, essentially in rodent burrows [64].When aninfected flea happens to feed on a commensal rodent, thecycle continues in the latter (Figure 2B). As commensal rodentsdie, their fleas are forced to move to alternative hosts, e.g.,humans. If humans develop pneumonic plague, the infectionmay be transmitted from person to person through respiratorydroplets spread by coughing (Figure 2C). Humans may alsobecome infected through handling of infected animals (ormeat), including rodents, camels, or cats. Cats can also developpneumonic plague, passing their infection to their ownersthrough coughing. There is, finally, some evidence suggestingthat the human flea, Pulex irritans, can be involved in human-to-human transmission [37]. Mammal predators, birds of prey, andother birds that use rodent burrows for nesting may move overlarger areas than the rodents themselves, spreading the infectionover longer distances. Also, infected commensal rats or humans

    may travel over long distances.

    doi:10.1371/journal.pmed.0050003.g002

    Figure 2. Possible Transmission Pathways for the Plague Agent, Y.pestis

    These pathways include sylvatic rodent-flea cycles (A), the commensalrodent-flea cycles (B), and the pneumonic transmission in humans (C).

    The colour of the arrows indicates the mechanism (flea bites, air particles,meat consumption) through which the bacteria are transferred from

    one host to another. Dark blue arrows indicate ways in which plague canmove to other areas.

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    7/13PLoS Medicine | www.plosmedicine.org 0013 January 2008 | Volume 5 | Issue 1 | e3

    8. Sodeinde OA, Subrahmanyam YV, Stark K, Quan T, Bao Y, et al. (1992) Asurface protease and the invasive character of plague. Science 258: 1004-1007.

    9. Hinnebusch BJ, Rudolph AE, Cherepanov P, Dixon JE, Schwan TG, et al.(2002) Role of Yersinia murine toxin in survival of Y. pestis in the midgut ofthe flea vector. Science 296: 733-735.

    10. Carniel E (2003) Evolution of pathogenic Yersinia,some lights in the dark.Adv Exp Med Biol 529: 3-12.

    11. Yersin A (1894) La peste bubonique Hong-Kong. Ann Inst Pasteur 2: 428-430.

    12. Devignat R (1951) Varits de lespce Pasteurella pestis. Nouvelle hypothse.Bulletin de lOrganisation Mondiale de la Sant 4: 247-263.

    13. Guiyoule A, Grimont F, Iteman I, Grimont PAD, Lefevre M, et al. (1994)Plague pandemics investigated by ribotyping of Yersinia pestis strains. J ClinMicrobiol 32: 634-641.

    14. Twigg G (1984) The Black Death: a biological reappraisal. London:Batsform Academic and Educational.

    15. Ziegler P (1969) The Black Death. Wolfeboro Falls (NH): Alan SuttonPublishing.

    16. Scott S, Duncan CJ (2001) Biology of plagues: Evidence from historicalpopulations. Cambridge: Cambridge University Press.

    17. Cohn SK Jr (2002) The Black Death transformed. London: Arnold.18. Drancourt M (2006) Yersinia pestis as a telluric, human ectoparasite-borne

    organism. Lancet Infect Dis 6: 234-241.19. Raoult D, Aboudharam G, Crubzy E, Larrouy G, Ludes B, et al. (2000)

    Molecular identification by suicide PCR of Yersinia pestis as the agent ofMedieval Black Death. Proc Natl Acad Sci U S A 97: 1280012803.

    20. Schrag SJ, Wiener P (1995) Emerging infectious diseases: what are therelative roles of ecology and evolution? Trends Ecol Evol 10: 319-324.

    21. Duplantier JM, Duchemin JB, Chanteau S, Carniel E (2005) From therecent lessons of the Malagasy foci towards a global understanding of the

    factors involved in plague reemergence. Vet Res 36: 437-453.22. Gage KL, Ostfeld RS, Olson JG (1995) Nonviral vector-borne zoonoses

    associate with mammals in the United States. J Mammal 76: 695-715.23. Levy CE, Gage KL (1999) Plague in the United States 1995-1996. Infect Med

    16: 54-64.24. Kaoneka ARS, Solberg B (1994) Forestry-related land use in the West

    Usambara mountains, Tanzania. Agric Ecosyst Environ 49: 207-215.25. Gabastou J-M, Proano J, Vimos A, Jaramillo G, Hayes E, et al. (2000) An

    outbreak of plague including cases with pneumonic infection, Ecuador,1998. Trans R Soc Trop Med Hyg 94: 387-391.

    26. Ruiz A (2001) Plague in the Americas. Emerg Infect Dis 7: 539-540.27. Fedorov VN (1960) Plagues in camels and its prevention in the USSR. Bull

    World Health Organ 23: 275-281.28. Christie AB, Chen TH, Elberg SS (1980) Plague in camels and goats: their

    role in human epidemics. J Infect Dis 141: 724-726.29. Atshabar BB, Aikimbaev AM, Aubakirov SA, Suleimenov BM (2000)

    Epizootologic and social basis for plague human infection in 1999 inKazakhstan and their clinical-epidemiologic characteristics. Problems of theMost Dangerous Infections, Saratov: 14-21.

    30. Arbaji A, Kharabsheh S, Al Azab S, Al Kayed M, Amr ZS, et al. (2005) A12-case outbreak of pharyngeal plague following the consumption of camelmeat, in north-eastern Jordan. Ann Trop Med Parasitol 99: 789-793.

    31. Bin Saeed AA, Al-Hamdan NA, Fontaine RE (2005) Plague from eating rawcamel liver. Emerg Infect Dis 11: 1456-1457.

    32. Isaacson M (1973) Unusual cases of plague in southern Africa. S Afr Med J47: 2109-2113.

    33. Doll JM, Seitz PS, Ettestad P, Bucholtz AL, Davis T, et al. (1994) Cattransmitted fatal pneumonic plague in a person who travelled fromColorado to Arizona. Am J Trop Med Hyg 51: 109-114.

    34. Gage KL, Dennis DT, Orloski KA, Ettestad PJ, Brown TL, et al. (2000) Casesof cat-associated human plague in the Western US, 1977-1998. Clin InfectDis 30: 893-900.

    35. Pollitzer R (1954) Plague. Monogr Ser World Health Organ 22: 1-698.36. Pollitzer R (1960) A review of recent literature on plague. Bull World

    Health Organ 23: 313-400.37. Blanc G (1956) Une opinion non conformiste sur le mode de transmission

    de la peste. Revue dHygine et de Mdecine Sociale 4: 532-562.

    38. World Health Organization (2004) Human plague in 2002 and 2003. WklyEpidemiol Rec 79: 301-306.39. World Health Organization (2006) Suspected plague in the Democratic

    Republic of the Congo. Available: http://www.who.int/csr/don/2006_11_07/en/index.html. Accessed 17 December 2007.

    40. International Society for Infectious Diseases (2007 March 1) PlagueUganda (Masindi): Pneumonic. ProMED-mail. Available: http://www.promedmail.org/pls/askus/f?p=2400:1202:3632215291704747094::NO::

    F2400_P1202_CHECK_DISPLAY,F2400_P1202_PUB_MAIL_ID:X,36504.Accessed 17 December 2007.

    41. Bertherat E, Lamine KM, Formenty P, Thullier P, Mondonge V, etal. (2005) Epidmie de peste pulmonaire dans un camp minier de laRpublique Dmocratique du Congo: le rveil brutal dun vieux flau.Mdecine Tropicale 65: 511-514.

    42. Bertherat E, Kone M, Formenty P, Thullier P, Mondonge V, et al. (2006)Emergence of plague in democratic Republic of Congo: a large pneumonicoutbreak erupts in a diamond mine. 5th International Conference onEmerging Infectious Diseases; 19-22 March 2006; Atlanta, Georgia, UnitedStates of America.

    43. Parmenter RR, Yadav EP, Parmenter CA, Ettestad PJ, Gage KL (1999)Incidence of plague associated with increased winter-spring precipitation inMexico. Am J Trop Med Hyg 61: 814-821.

    44. Enscore RE, Biggerstaff BJ, Brown TL, Fulgham RF, Reynolds PJ, et al.(2002) Modeling relationships between climate and the frequency ofhuman plague cases in the southwestern United States, 1960-1997. Am JTrop Med Hyg 66: 186196.

    45. Stenseth NC, Samia NI, Viljugrein H, Kausrud K, Begon M, et al. (2006)Plague dynamics are driven by climate variation. Proc Natl Acad Sci U S A103: 13110-13115.

    46. Kausrud KL, Viljugrein H, Frigessi A, Begon M, Davis S, et al. (2007)Climatically driven synchrony of gerbil populations allows large-scale plagueoutbreaks. Proc R Soc Lond B Biol Sci 274: 1963-1969.

    47. Treydte K, Schleser GH, Helle G, Frank DL, Winiger M, et al. (2006)Twentieth century as the wettest period in northern Pakistan over the pastcentury. Nature 440: 1179-1182.

    48. Intergovernmental Panel on Climate Change (2001) Climate change 2001:The scientific basis. Houghton JT, Ding Y, Griggs DJ, Noguer M, van derLinden PJ, et al., editors. Available: http://www.grida.no/climate/ipcc_tar/

    wg1/index.htm. Accessed 13 December 2007.

    49. Ganapati M (1995) Indias pneumonic plague outbreak continues to baffle.BMJ 311: 706.

    50. Fritz CL, Dennis DT, Tipple MA, Campbell GL, McCance CR, et al.(1996) Surveillance for pneumonic plague in the United States Duringan international emergency: A model for control of imported emergingdiseases. Emerg Infect Dis 2: 30-36.

    51. Chanteau S, Rahalison L, Ralafiarisoa L, Foulon F, Ratsitorahina M, et al.(2003) Development and testing of a rapid diagnostic test for bubonic andpneumonic plague. Lancet 361: 211-216.

    52. Webb CT, Brooks CP, Gage KL, Antolin MV (2006) Classic flea-borntransmission does not drive plague epizootics in prairie dogs. Proc Natl

    Acad Sci U S A 103: 6236-6241.53. Begon M, Klassovskiy N, Ageyev V, Suleimenov B, Atshabar B, et al. (2006)

    Epizootiologic parameters for plague in Kazakhstan. Emerg Infect Dis 12: 268-273.54. Davis S, Begon M, De Bruyn L, Ageyev V, Viljugrein H, et al. (2004)

    Predictive Thresholds for plague in Kazakhstan. Science 304: 736738.55. Parkhill J, Wren BW, Thomson NR, Titball RW, Holden MT, et al. (2001)

    Genome sequence of Yersinia pestis, the causative agent of plague. Nature413: 523-527.

    56. Galimand M, Guiyoule A, Gerbaud G, Rasoamanana B, Chanteau S, et al.(1997) Multidrug resistance in Yersinia pestis mediated by a transferableplasmid. New Engl J Med 337: 677-680.

    57. Guiyoule A, Gerbaud G, Buchrieser C, Galimand M, Rahalison L, et al.(2001) Transferable plasmid-mediated resistance to streptomycin in aclinical isolate of Yersinia pestis.Emerg Infect Dis 7: 43-48.

    58. Hinnebusch BJ, Rosso ML, Schwan TG, Carniel E (2002) High-frequencyconjugative transfer of antibiotic resistance genes to Yersinia pestis in the fleamidgut. Mol Microbiol 46: 349-354.

    59. Welch TJ, Fricke WF, McDermott PF, White DG, Rosso M-L, et al. (2007)Multiple antimicrobial resistance in plague: an emerging public health risk.PLoS ONE 2: e309. doi:10.1371/journal.pone.0000309

    60. Inglesby TV, Dennis DT, Henderson DA, Bartlett JG, Ascher MS, etal. (2000) Plague as a biological weapon: medical and public healthmanagement. JAMA 283: 2281-2290.

    61. Koirala J (2006) Plague: disease, management, and recognition of act ofterrorism. Infect Dis Clin North Am 20: 273-287.

    62. Hotez PJ, Molyneux DH, Fenwick A, Ottesen E, Ehrlich Sachs S, et al.

    (2006) Incorporating a rapid-impact package for neglected tropical diseaseswith programs for HIV/AIDS, tuberculosis, and malaria. PLoS Med 3: e102.doi:10.1371/journal.pmed.0030102

    63. Simond PL (1898) La propagation de la peste. Annales de lInstitut Pasteur12: 625-687.

    64. Baltazard M, Karimi Y, Eftekhari M, Chamsa M, Mollaret HH (1963) Laconservation interpizootique de la peste en foyer invtr hypothses detravail. Bulletin de la Socit de Pathologie Exotique 56: 1230-1241.

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    8/13

    Review

    Plague and Climate: Scales Matter

    Tamara Ben Ari1,2, Simon Neerinckx1, Kenneth L. Gage3, Katharina Kreppel4, Anne Laudisoit5, Herwig

    Leirs5, Nils Chr. Stenseth1*

    1 Centre for Ecological and Evolutionary Synthesis (CEES), University of Oslo, Oslo, Norway, 2 Ecole Normale Superieure, CNRS UMR 7625, Paris, France,3 Bacterial Diseases

    Branch, Division of Vector-Borne Diseases, Center of Control and Prevention, Fort Collins, Colorado, United States of America, 4 Liverpool University Climate and Infectious

    Diseases of Animals Group (LUCINDA), Department of Veterinary Clinical Sciences, University of Liverpool, Leahurst, Great Britain, 5 Evolutionary Ecology Group,

    Department of Biology, Universiteit Antwerpen, Antwerp, Belgium

    Abstract: Plague is enzootic in wildlife populations ofsmall mammals in central and eastern Asia, Africa, Southand North America, and has been recognized recently as areemerging threat to humans. Its causative agent Yersiniapestis relies on wild rodent hosts and flea vectors for itsmaintenance in nature. Climate influences all threecomponents (i.e., bacteria, vectors, and hosts) of theplague system and is a likely factor to explain some ofplagues variability from small and regional to large scales.Here, we review effects of climate variables on plaguehosts and vectors from individual or population scales tostudies on the whole plague system at a large scale.Upscaled versions of small-scale processes are ofteninvoked to explain plague variability in time and spaceat larger scales, presumably because similar scale-inde-pendent mechanisms underlie these relationships. Thislinearity assumption is discussed in the light of recentresearch that suggests some of its limitations.

    Introduction

    Plague is a rapidly progressing infectious disease that is

    infamous for having caused the death of millions of people in

    large historic pandemics [1] as well as numerous other deadly butlocalized outbreaks [2]. Plague, caused by the pathogenic

    bacterium Y. pestis is transmitted from host to host by fleas viablood feeding, through consumption or handling of infectious host

    tissues, or through inhalation of infectious materials. Plague is

    thought to persist for long periods of time at low to very low levels

    of prevalence in so-called enzootic cycles that cause little host

    mortality and involve partially resistant rodents (often called

    enzootic or maintenance hosts). These long periods are punctuated

    by occasional outbursts or epizootics (i.e., spreading die-offs)among these hosts or epidemics, when the incidence among

    humans increases. Figure 1 illustrates these intertwined cycles.Climate has long been suspected to be a key factor in the

    alternation between quiescent and active periods of plague. Rogers

    (1928) suggested seasonal variations in temperature and humidity

    to be responsible for the seasonal patterns of human plagueincidence in India [3]. Decades later, Davis showed that human

    plague outbreaks in several African countries were less frequent

    when the weather was too hot (.27uC) or cold (,15uC) [4].

    Subsequent studies showed an increased plague incidence in

    Vietnam during the hot, dry season, when following a period of

    high seasonal rainfall [5,6]. Nowadays, several studies, as we

    report here, demonstrate climates impacts on plague incidence

    spatial and temporal variability.

    In the context of public health and wildlife conservation, we

    need an improved understanding of the mechanisms underlying

    the association between plague outbreaks and climate. As we will

    show in this review, this understanding is only partially available at

    present. There are several reasons for this. First and perhaps most

    importantly, the plague cycle is complex. It is composed of three

    components that interact with each other and all are influenced by

    climate variables with a broad range of times lags. Also, climate

    variability manifests itself at numerous temporal and spatial scales

    that condition the form of the response in plague dynamics. To

    cope with this series of difficulties, we break down the problem as

    follows: in section 1 we review individual effects of climate

    variables on each of the plague components. Our knowledge of

    these effects is primarily based on small-scale studies that are usefulbecause they provide conceptual models for how larger scale

    climate variability may force the plague system. The way these

    conceptual models are most often used raises the issue of upscaling

    conclusions by inference from the results of small scale studies, a

    subject on which we focus on in section 2. Also, in the latter we

    review likely impacts of climate change on plague incidence.

    Climate Dependence of the Flea Vectors and

    Rodent Hosts

    The plague system is the result of complex interactions between

    its components, the densities, life cycle, dynamics and geographical

    distributions all of which are individually influenced by climate

    variables. Climate variables influence the dynamics of flea vectorsand rodent hosts with responses varying considerably among

    species [7,8]. Figure 1 illustrates the plague cycle in relation to

    those climate variables known to be important (namely temper-

    ature, humidity and precipitation) to the main plague hosts and

    vectors.

    It is accepted that abundance of rodent fleas is affected by

    ambient temperature, rainfall, and relative humidity, with warm-

    moist weather providing a likely explanation for higher flea indices

    [4,5]. Indeed, temperature, rainfall, and relative humidity have

    direct effects on development and survival, as well as the behavior

    and reproduction of fleas and their populations [912]. For

    Citation: Ben Ari T, Neerinckx S, Gage KL, Kreppel K, Laudisoit A,et al. (2011) Plague and Climate: Scales Matter. PLoS Pathog 7(9): e1002160.doi:10.1371/journal.ppat.1002160

    Editor:Marianne Manchester, University of California San Diego, United States ofAmerica

    Published September 15, 2011

    This is an open-access article, free of all copyright, and may be freely reproduced,distributed, transmitted, modified, built upon, or otherwise used by anyone forany lawful purpose. The work is made available under the Creative Commons CC0public domain dedication.

    Funding: The authors received no specific funding for this study.

    Competing Interests:The authors have declared that no competing interestsexist.

    * E-mail: [email protected]

    PLoS Pathogens | www.plospathogens.org 1 September 2011 | Volume 7 | Issue 9 | e1002160

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    9/13

    example, the rate of metamorphosis of Xenopsylla cheopis (as a

    primary flea of the black rat Rattus rattus,X. cheopisis likely the main

    vector of plague in foci affecting humans), from egg to adult is

    regulated by temperature. Fleas are ecto-thermic and hence

    sensitive to temperature fluctuations; this is enhanced by the fact

    that all of the immature flea stages occur off host. Flea

    development rates increase with temperature until they reach a

    critical value; then the survival of immature stages decreases if

    high temperatures are combined with low humidity [13].

    Temperature and relative humidity impact flea survival [5]:

    survival is in fact inversely proportional to air saturation deficit at a

    constant temperature [14]. Flea larvae are susceptible to

    desiccation [15] and typically acquire water from adult excreta.

    Survival of immature stages of fleas in rodent burrows is also

    affected by soil moisture that is partly controlled by outsideprecipitation [16] even though detrimental moisture losses and

    temperature swings are reduced by living underground [9].

    Conversely, when coupled with a high organic load, excessively

    wet conditions in rodent burrows (e.g., relative humidity .95%)

    can promote the growth of destructive fungi that diminish larval

    and egg survival [5,17].

    Rodent survival and population dynamics are also affected by

    climate. A direct effect occurs when high intensity rainfall causes

    flooding of rodent burrows [5] but the effects of precipitation on

    rodent densities are mostly bottom-up [18]. Indeed, rainfall

    controls primary production which limits rodent abundances [19].

    Reproduction and recruitment periods often follow wet seasons

    when increases of primary production can be used to build up

    juvenile populations [20]. Accordingly, rodent population densities

    show clear association with annual rainfall and its seasonal

    distribution e.g., in Chile [2123], Tanzania [24], and Australia

    [25,26]. But the relation between precipitation patterns and rodent

    densities can be complex, localized, and dependent on the timing

    and the intensity of precipitation events (see also below) [8,27].

    Temperature effects on rodent populations are less clear in part

    because rodents are homeothermic and hence do not respond

    immediately to changes in ambient temperatures. In temperate

    areas, low temperatures in winter can nonetheless negatively affect

    rodent populations either directly or through low food availability

    [28]. Nevertheless, under some circumstances, conditions detri-

    mental to hosts or vectors can favor the maintenance of plague.Evidence of hibernation as a factor of prolongation or modifica-

    tion ofY. pestis infection in rodents would need further elucidation.

    The flea Citellophilus tesquorum altaicus for instance, is supposedly

    able to maintain a plague infection over the winter by feeding on

    hibernating long-tailed Siberian souslik (Citellus undulatus) [29].

    Also, populations of Tatera indica aestivating during adverse

    conditions in India presumably continue to act as hosts for

    infected fleas, thereby promoting the persistence of plague

    infection within the area [30]. Hence, the survival ofY. pestis in

    relatively plague-resistant burrowing rodents that interrupt their

    activities to hibernate through winter or aestivate in summer could

    Figure 1. Schematic of the plague cycle with small mammals as hosts and fleas as vectors. Arrows represent connections affected byclimate with a color-coding depending on the most influential climate variable on this link (i.e., precipitation, temperatures, and other variablesindirectly depending on them such as soil characteristics and soil moisture). Grey rectangles somewhat arbitrarily delimit epizootic, enzootic, andzoonotic cycles. Note that despite their location at the far end of the cycle, humans often provide the only available information on plague dynamics.doi:10.1371/journal.ppat.1002160.g001

    PLoS Pathogens | www.plospathogens.org 2 September 2011 | Volume 7 | Issue 9 | e1002160

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    10/13

    influence or prevent the transient temporal and spatial extinction

    of plague occurrence.

    Human Plague Incidence Is Not Unrelated to HumanFactors

    Human activities and behavior in plague-infected areas are also

    to be considered as important determinants of plague transmission

    to humans [8]. When occurrences of plague are due to human

    intrusions in natural plague areas, it is thus important to considerclimate as a second order variable that influences disease incidence

    through human behavior (drought, famine, war, or other events).

    In Argentina, plague transmission reportedly occurred during the

    harvesting season [31]. In Lushoto, a plague endemic region in

    Tanzania, daily and gender activities seem to impact plague levels

    [32], although plague tends to peak during the season with the

    least agricultural activity, which is a time when people usually

    gather in houses.

    What Plague Niches Reveal about PlaguesEnvironmental Preferences

    Plague foci are present over an expanded geographical range

    that includes the Western US, portions of South America, East

    and South Africa, and Southeast Asia [33]. Long-term mainte-nance of plague in defined ecological niches may inform us about

    the environmental conditions that are required for plague to

    establish in permanent foci. Unfortunately, enzootic plague is

    poorly described; in many foci, local reservoirs have yet to be

    identified. The geographical limits of plague territories are hence

    rarely defined or only by occurrence in domestic rodents and

    humans [34]. It is reasonable to assert that plague manifests itself

    under various ecological conditions [35,36]. It is noteworthy

    though, that modern plague foci in North and East Africa,

    Western North America, parts of South America, and many

    scattered regions in Asia (notably China and Kazakhstan) occur

    primarily in either semi-arid to arid areas or low humidity forest

    types of habitat, and the disease apparently fails to persist for long

    periods in humid tropical lowland areas (except from occasional

    invasion through movement of infected humans or transportation

    of infected rodents or fleas) [3739]. Also, plague is almost

    invariably absent from the hottest and driest desert regions like the

    Sahara or Sonoran deserts [37,40].

    The Complexity Introduced by Interactions

    among Scales and Other Nonlinearity

    The previous section reviewed reported effects of climate

    variables on the components of the plague cycle, indicating that

    climate affects rodents and fleas individually and their population

    dynamics and structures. Consequently, climate impacts the

    natural cycle of plague as a whole and in ways that are not likely

    to reflect simply the sum of these individual effects. In this section,

    we review studies on the effects of climate (including theenvironment) on plague at different spatial levels and occurrences

    ranging from rodent burrows to areas that form a coherent and

    apparently self-sustaining system (a single niche or focus), or even

    to larger regions that could comprise many such systems (several

    foci). We emphasize the fact that (i) scales relevant to plague

    ecology are nested within each other, as shown in Figure 2, so that

    climate effects at a given scale may not be simply extrapolated

    from those observed at a smaller scale, and (ii) processes exist that

    prevent the plague system from responding to climate fluctuations

    in a way that simply reflects the sum of the responses of its

    individual component.

    It is readily apparent that associations exist between large-scale

    climate such as climate indices and plague/plague hosts dynamics;

    associations that can be consistently explained by processes

    detailed in the first section. Effects of precipitation patterns on

    plague incidence are for instance assumed to be the results of

    climates bottom-up effects on plague hosts. In Peru, climate

    fluctuations associated with El Nino were related to a bubonic

    plague outbreak in 1999 [41]. In northern Colorado, prairie dog

    colony extinctions attributed to plague were weakly associatedwith El Nino southern oscillation [42]. In the Western US, spatial

    synchrony of periods with low and high number of human plague

    occurrences throughout the west revealed large scale synchrony

    [43]. In most foci of the southwestern part of this region, above

    normal precipitation in winter and spring was used to explain

    increases in human plague cases [44], and high summer

    temperatures, decreases of its incidence in the same area [45].

    The El Nino/Southern Oscillation (ENSO) and the Pacific

    Decadal Oscillation were similarly related to precipitation,

    temperatures, food availability, and the number of plague cases

    themselves [46]. Finally, the increase rate of human plague in

    China is associated at the province level with the Southern

    Oscillation Index and Sea Surface Temperature of the tropic

    Pacific east equator [47]. An important assumption to the above-

    cited studies is that large-scale climate variability producescoherent and synchronous effects on rodent hosts populations.

    Few studies actually investigate the reality of the invoked

    synchronous trophic forcing on plague hosts population dynamics

    so as to demonstrate such a climate induced bottom-up control. At

    this point, it is worth mentioning a counterexample provided by

    Kausrud et al. [48], who investigated the population dynamics of

    great gerbils Rhombomys opimusthe main plague host in Kazakh-stanand demonstrated spatial synchrony of their populations

    over areas larger than the ones expected by migration processes.

    Their results indicate that the observed synchrony of great gerbils

    densities was most probably due to the effects of climate and that

    similar bottom-up processes could also synchronize plague activity

    in this focus.

    In any case, there are caveats to inferring processes occurring atsmall scale as a means to explain the ones observed at larger scales.

    First, causal relations can typically only be suggested (e.g., by a

    chain of supposedly causal processes from climate indices to

    relevant climate variables and from these to plague incidence) but

    not demonstrated. More problematically, a variety of processes

    can be expected to interfere with each other in the transition from

    small to large scale. Tentatively, we propose below a classification

    of these processes into two categories. The first ones pertain to the

    complexity of the environment itself and the second ones to

    population interactions.

    An example in the first category can be provided by studies

    showing that rodent burrows could be considered a climate

    filter. Conditions in rodent burrows are subject to environmental

    factors from nearby surroundings such as temperature, humidity,

    vegetation, various soil properties (e.g., soil texture and structure,soil organic carbon content, etc.), and other landscape factors (e.g.,

    slope, orientation with respect to dominant winds, sun exposure,

    etc.). But, the underground location of most burrows implies that

    conditions inside these structures fluctuate only moderately

    [49,50]. In fact, temperatures inside and outside burrow systems

    are highly correlated, but inside humidity is a complex function of

    past rainfall and soil characteristics [51] rather than of present

    ambient humidity [52]. Note that investigations on burrow micro-

    climate have been conducted in different parts of the world

    [51,53,54], including in plague-endemic regions, but these

    measurements were generally obtained from only a small number

    PLoS Pathogens | www.plospathogens.org 3 September 2011 | Volume 7 | Issue 9 | e1002160

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    11/13

    of natural or, in some instances, artificial burrows, allowing

    investigators to draw only tentative conclusions from the results of

    these studies [49,55]. Perhaps more importantly for our present

    purposes, climate conditions inside the burrows could specifically

    influence the distribution of plague vectors so that hosts

    distribution becomes insufficient to explain vectors distribution, a

    situation that has been observed in plague areas in Madagascar

    [5658] or plague-free areas in Israel [59,60]. These examples

    emphasize the difficulty of upscaling processes by simple

    extrapolation. Among the numerous other examples of environ-

    ment-related complications are the landscape heterogeneities

    within plague areas. In particular, strong orographic forcings

    frequent in plague areas locally affect large-scale climate variability

    [61]. In fact, accurate plague prediction models seem to require

    high resolution environmental and climate representation (e.g.,

    250-m resolution images accurately predict plague distribution

    when 10-km resolution images fails in sub-Saharan Africa).

    According to the authors, plague focality can not be explained

    by fragmented environmental conditions at this coarse

    scale [35,62].

    In the second category are the complications introduced by

    interactions within and between plague host and plague vector

    populations in their response to climate variability. It is commonly

    expected that fluctuations in abundance of rodent hosts, e.g.induced by climate variability, will translate into plague prevalence

    fluctuations. However, the relationship between host abundance

    and host prevalence is complex and scale dependent. Field studies

    of rodent reservoirs show a negative correlation between host

    abundance and host prevalence at seasonal time scale. The finding

    has been explained by the juvenile dilution effect, that is, the

    arrival of numerous healthy juveniles in a population [63]; an

    effect likely to be relevant for diseases with no vertical transmission

    such as plague (or Hantavirus [64]). In contrast, a few longer field

    studies (.5 years) show a positive, although delayed relationship

    between reservoir abundance and prevalence [63]. These

    antagonist responses to variability at different time scales are

    typical of systems in which nonlinear interactions play an essential

    role. The different responses of rodent and flea populations to

    climate (fast for fleas while rodents tend to integrate environmental

    conditions over some years) provide another reason for question-

    ing the existence of a straightforward link between abundance and

    prevalence. A more complex model was proposed for the Western

    US, in which human plague incidence depends both on time-

    lagged precipitation events, which presumably increase rodent

    numbers and favor plague epizootic, and on relatively cool

    summer temperatures during the plague transmission season,

    which are favorable to infectious fleas [45]. This model has been

    coined trophic cascade hypothesis, although it has yet to be

    tested rigorously for plague under field conditions [65] (see also

    [66] for a discussion on the accuracy of the use of the term

    trophic for the cascade hypothesis). In particular, the scale at

    which the trophic cascade model is valid needs to be addressed.

    Parmenter [44], who first introduced this model, shows its

    relevance at a local scale (i.e., by demonstrating significant

    associations between plague and nearby precipitation), but could

    not extend this result to a state-wide level. Interestingly, a cascade

    relationship was recovered at an even larger scale by Ben-Ari [46],

    possibly because the integration of delayed density-dependence

    effects of large-scale climate on rodents were taken into account atdecadal time scales. The study by Stenseth et al. [67] illustrates the

    challenge that needs to be confronted when addressing cascading

    effects of climate on plague prevalence in nature. There are

    numerous relationships between climate elements (temperature

    and precipitation with or without lags) and various ways these

    elements can impact the plague components (in this particular case

    rodent density, prevalence, and flea burden). The type of dataset

    that would let us isolate/untangle the mechanisms and the spatio-

    temporal scales at which processes operate may not be available.

    These examples do not necessarily invalidate studies that scale

    up small-scale results (individual measurements, lab experiments,

    Figure 2. Illustration of the abiotic environment impact on the plague cycle as a function of spatial scale. Arrows represent connectionsaffected by climate (see Figure 1 for the meaning of color coding). Most climate variables act over a wide range of scales and only the effects wedeemed most important are represented. At the level of individuals, populations and communities hosts and vectors are influenced by climatevariability at the relevant scale (local or regional). At the smaller scale, the burrow acts as a filter on climate variables. Note that secondary hosts areplaced at the kilometer and larger scale on the basis of the type of information generally available regarding their infection.doi:10.1371/journal.ppat.1002160.g002

    PLoS Pathogens | www.plospathogens.org 4 September 2011 | Volume 7 | Issue 9 | e1002160

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    12/13

    local correlations, etc.) in the simplest way, but we believe they

    illustrate the need for more investigation on the impact of complexinteractions and environment heterogeneities at intermediate

    scales.

    Implications for Climate Change

    The need to understand disease responses in relation to climate

    variability is made particularly acute by ongoing global climate

    change. Effects of temperature rise on vector-borne diseases andnotably the ones involving free-living stages of terrestrial animals

    are expected [68]. Beyond that, a lot of uncertainty remains on

    whether or how climate change might affect pathogens and disease

    transmission, i.e., changes in population size (for vectors or hosts),

    overall prevalence, timing and seasonality, or shifts in geographical

    distribution. There are various choices to be made when

    addressing climate change impact on a disease like plague,

    particularly with respect to the degree of complexity that should be

    chosen for climate models and any associated biological models

    and the relation between them.

    Stenseth et al. [67] performed a sensitivity study to a one-degree

    increase in temperature into a statistical host vector plague model

    developed specifically for Kazakhstan. They show that this (simple)

    climate change scenario would lead to a 50% increase in plagueprevalence among great gerbils. Nakazawa et al. [69] used twoGlobal Circulation Models (GCMs) to infer mean temperature

    and precipitation changes between the present and a 30-year

    period centered around 2055. These changes were then applied to

    a higher resolution present state GCM that provided a climate

    changed forcing state, which was fed into an Ecological Niche

    Model (ENM) predicting plague occurrences in the US from a set

    of environmental variables. They show subtle shifts of plague

    habitats (generally northward). Snall et al. [70] used an elaborate

    procedure to downscale climate scenarios from several GCMs

    before using these data to force an explicit model for the joint host-

    parasite dynamics of black-tailed prairie dogs and plague in the

    Western US. A related decrease in the number of infected prairie

    dog colonies (leading to an increase in prairie dog colonies) is

    predicted, presumably, as a consequence of the negative impact ofincreased frequency of abnormally hot days on plague transmis-

    sion.

    These studies are difficult to compare with each other becauseof the specificities of their methodologies even when they lead to

    contrasting results over the same region [70]. Admittedly, moreinvestigations are required. Among the numerous possible

    approaches, the safest arguably would rely on using climate-

    forcing sensitivities of intermediate complexity where biological

    models are forced by existing modes of climate variability. This

    makes sense not only because climate change will in large part

    occur through a modification of these modes (which can be

    extracted from Intergovernmental Panel on Climate Change

    [IPCC] GCMs [61]), but also because biological data are then

    available for evaluation.

    Conclusion

    Climate impacts all components of the plague cycle (host,

    vector, and pathogen) in various ways and over a wide range of

    scales (from microindividual flea life cycleto macroa plague

    area composed of several disjoint foci). Several studies have

    established links between plague occurrence and climate factors

    that can a posteriori be justified by assuming the validity at a large

    scale of relationships that have only been observed at a small scale

    (assumption of linearity). We have reviewed both the successes and

    the limitations of this assumption. To go beyond simple inferences

    on how climate fluctuations (including long-term climate change)

    affect plague, it might be necessary to select a (or a few)

    preferential scale, on the basis of the fact that they would be the

    most informative and/or relevant for public health policies. In thisregard, intrinsic dynamics of plague hosts and vectors should be

    kept in mind as it alone greatly contributes to the entanglement of

    scales that drive the overall dynamics of plague prevalence [71].

    Further, assessing plague risks for humans at such scales may in

    fact require investigating plague dynamics on a much wider range

    of scales and presumably include a fuller description of the plague

    system in its environment, as we have tried to outline in this

    review.

    Acknowledgments

    We thank Ulf Buentgen for his help with figure preparation; Mike Begon,

    Ulf Buentgen, and Sasha Gershunov for discussions on plague and climate;

    and Xavier Capet for numerous comments and corrections on earlier

    versions of the manuscript. We also thank two anonymous reviewers forproviding helpful comments to an earlier version of this contribution.

    Author Contributions

    Conceived and designed the experiments: TBA NCS. Performed the

    experiments: TBA SN. Analyzed the data: TBA SN. Wrote the paper:

    TBA SN KLG KK AL HL NCS.

    References

    1. Stenseth NC, Atshabar BB, Begon M, Blemain SR, Bertherat E, et al. (2008)

    Plague: past, present and future. PloS Med 5: e3. doi:10.1371/journal.

    pmed.0050003.

    2. Catanach IJ (2001) The globalization of disease? India and the plague. Journal

    of World History 12: 131153.

    3. Rogers L (1928) The yearly variations in plague in India in relation to climate:

    forecasting epidemics. Proc Roy Soc Ser B 103: 4272.4. Davis DH (1953) Plague in Africa from 1935 to 1949; a survey of wild rodents in

    African territories. Bull World Health Organ 5: 665700.

    5. Cavanaugh DC, Marshall JD (1972) The influence of climate on the seasonal

    prevalence of plague in the Republic of Vietnam. J Wildl Dis 8: 8594.

    6. Cavanaugh D, Dangerfield H, Hunter D, Joy R, Marshall JDJ, et al. (1968)

    Some observations on the current plague outbreak in the Republic of Vietnam.

    Am J Public Health Nations 58: 742752.

    7. Meserve PL, Yunger JA, Gutierrez JR, Contreras LC, Milstead WB, et al. (1995)

    Heterogeneous responses of s mall mammals to an El Nino Southern Oscillation

    event in north central semiarid Chile and the importance of ecological scale.

    J Mammal 76: 580595.

    8. Gubler DJ, Reiter P, Ebi KL, Yap W, Nasci R, et al. (2001) Climate variability

    and change in the United States: potential impacts on vector- and rodent-borne

    diseases. Environ Health Persp 109: 22333.

    9. Krasnov BR, Khokhlova IS, Fielden LJ, Burdelova NV (2001) Effect of air

    temperature and humidity on the survival of pre-imaginal stages of two flea

    species (Siphonaptera: Pulicidae). J Med Entomol 38: 62937.

    10. Krasnov BR, Khokhlova IS, Fielden LJ, Burdelova NV (2001) Development

    rates of two Xenopsyllaflea species in relation to air temperature and humidity.

    Med Vet Entomol 15: 249258.

    11. Krasnov BR, Burdelova NV, Shenbrot GI, Khokhlova IS (2002) Annual cyclesof four flea species in the central Negev desert. Med Vet Entomol 16: 266276.

    12. Gage KL, Burkot TR, Eisen RJ, Hayes EB (2008) Climate and vector-borne

    diseases. Am J Prev Med 35: 43650.

    13. Gage KL, Burkot TR, Eisen RJ, Hayes EB (2008) Climate and vector-borne

    diseases. Am J Prev Med 35: 43650.

    14. Bacot AW, Martin CJ (1924) The respective influence of temperature and

    moisture upon the survival of the rat flea. J Hyg 23: 98105.

    15. Cavanaugh DC(1971) Specific effectof temperature upon transmission of theplague

    bacillus by the oriental rat flea, XenopsyllaCheopis. J Trop Med Hyg 20: 264273.

    16. Eisen RJ, Gage KL (2009) Adaptive strategies of Yersinia pestis to persist during

    inter-epizootic and epizootic periods. Vet Res 40: 01.

    17. Parmenter RR, Yadav EP, Parmenter CA, Ettestad P, Gage KL (1999)

    Incidence of plague associated with increased winter-spring precipitation in New

    Mexico. Am J Trop Med Hyg 61: 814821.

    PLoS Pathogens | www.plospathogens.org 5 September 2011 | Volume 7 | Issue 9 | e1002160

  • 8/12/2019 Plague: Past, Present, and Future by Mike Begon et alt

    13/13

    18. Meserve PL, Milstead WB, Gutierrez JR (2001) Results of a food additionexperiment in a north-central Chile small mammal assemblage: evidence for therole of bottom-up factors. Okos 94: 548556.

    19. Letnic M, Tamayo B, Dickman CR (2005) The responses of mammals to LaNina (El Nino Southern Oscillation) associated rainfall, predation, and wildfirein central Australia 86: 689703.

    20. Jaksic FM, Lima M (2003) Myths and facts on ratadas: bamboo blooms, rainfallpeaks and rodent outbreaks in South America. Aust J Ecol 28: 237251.

    21. Lima M, Jaksic FM (1999) Population rate of change in the leaf-eared mouse: therole of density-dependence, seasonality and rainfall. Aust J Ecol 24: 110116.

    22. Lima M, Keymer JE, Jaksic FM (1999) El Nino-southern oscillation-driven

    rainfall variability and delayed density dependence cause rodent outbreaks inwestern South America: linking demography and population dynamics. Am Nat153: 476491.

    23. Lima M, Marquet PA, Jaksic FM (1999) El Nino events, precipitation patterns,and rodent outbreaks are statistically associated in semiarid Chile. Ecography22: 213218.

    24. Leirs H, Verhagen R, Verheyen W, Mwanjabe P, Mbise T (1996) Forecastingrodent outbreaks in Africa: an ecological basis for Mastomys control inTanzania. J Appl Ecol 33: 937943.

    25. Dickman CR, Haythornthwaite AS, McNaught GH, Mahon PS, Tamayo B,et al. (2001) Population dynamics of three species of dasyurid marsupials in aridcentral Australia: a 10-year study. Wildl Res 28: 493506.

    26. Mills JN (2005) Regulation of rodent-borne viruses in the natural host:implications for human disease. Arch Virol 19: 4557.

    27. Brown JH, Morgan Ernest SK (2002) Rain and rodents: complex dynamics ofdesert consumers. Bioscience 52: 979987.

    28. Korslund L, Steen H (2006) Small rodent winter survival: snow conditions limitaccess to food resources. J Anim Ecol 75: 156166.

    29. Bazanova, L P, Nikitin, A Y, Popkov, A F, Maevskii MP (2007) Seasonalpeculiarities of plague agent (Yersinia pestis) transmission to the long-tailed suslikby fleas (Citellophilus tesquorum) in Tuva [in Russian]. Zoologichesky Zhurnal86: 846852.

    30. Baltazard M, Bahmanyar M (1960) Research on plague in India. Bull WorldHealth Organ 23: 169215.

    31. Macchiavello A (1946) A focus of sylvatic plague on the Peruvian-Ecuadorianfrontier. Science 104: 522.

    32. Kilonzo BS, Mvena ZS, Machangu RS, Mbise TJ (1997) Preliminaryobservations on factors responsible for long persistence and continued outbreaksof plague in Lushoto district, Tanzania. Acta tropica 68: 21527.

    33. WHO (World Health Organization) (2008) Interregional meeting on preventionand control of plague. AntananarivoMadagascar, 711 April 2006. 165.

    Available: http://www.who.int/csr/resources/publications/WHO_HSE_EPR_2008_3w.pdf. Accessed 16 August 2011.

    34. Neerinckx SB, Peterson AT, Gulinck H, Deckers J, Leirs H (2008) Geographicdistribution and ecological niche of plague in sub-Saharan Africa. Int J HealthGeogr 7: 54.

    35. Prentice MB, Rahalison L (2007) Plague. Lancet 369: 11961207.36. Dennis DT, Gage KL, Gratz NG, Poland JD, Tikhomirov E (1999) Plague

    manual: epidemiology, distribution, surveillance and control. Bull World HealthOrgan. pp 1171.

    37. Perry RD, Fetherston JD (1997) Yersinia pestis- etiologic agent of plague. ClinMicrobiol Rev 10: 3566.

    38. Gage KL, Kosoy MY (2005) Natural history of plague: perspectives from morethan a century of research. Ann Rev Entomol 50: 50528.

    39. Barnes A (1982) Surveillance and control of bubonic plague in the United States.Symp Zool Soc Lond 50: 237270.

    40. Davalos VA, Torres MA, Mauricci CO, Laguna-Torres VA, Chinarro MP(2001) Outbreak of bubonic plague in Jacocha, Huancabamba, Peru. Rev SocBras Med Trop 34: 8790.

    41. Stapp P, Antolin MF, Ball M (2004) Patterns of extinction in prairie dogmetapopulations: plague outbreaks follow El Ninno events. Front Ecol Environ2: 235240.

    42. Ben Ari T, Gershunov A, Gage KL, Snall T, Ettestad P, et al. (2008) Humanplague in the USA: the importance of regional and local climate. Biol lett 4:737740.

    43. Pollitzer R (1954) Plague. World Health Organization monograph seriesnumber 22. Geneva: World Health Organization. pp 698.

    44. Parmenter RR, Yadav EP, Parmenter C, Ettestad P, Gage KL (1999) Incidence

    of plague associated with increased winter-spring precipitation in New Mexico.Am J Trop Med Hyg 61: 814821.

    45. Enscore RE, Biggerstaff BJ, Brown TL, Fulgham RE, Reynolds PJ, et al. (2002)Modeling relationships between climate and the frequency of human plaguecases in the southwestern United States, 1960-1997. Am J Trop Med Hyg 66:186196.

    46. Ben Ari T, Gershunov A, Tristan R, Cazelles B, Gage K, et al. (2010)Interannual variability of human plague occurrence in the Western UnitedStates explained by tropical and North Pacific Ocean climate variability.

    Am J Trop Med Hyg 83: 624632.

    47. Zhang Z, Li Z, Tao Y, Chen M, Wen X, et al. (2007) Relationship betweenincrease rate of human plague in China and global climate index as revealed bycross-spectral and cross-wavelet analyses. Int Zool 2: 144153.

    48. Kausrud KL, Viljugrein H, Frigessi A, Begon M, Davis S, et al. (2007)Climatically driven synchrony of gerbil populations allows large-scale plagueoutbreaks. P Roy Soc B-Biol Sci 274: 19631969.

    49. Longanecker DS, Burroughs AL (1952) Studies of the microclimate of the

    California ground squirrel burrow and its relation to seasonal changes in the fleapopulation. 33: 488499.

    50. Sumbera R, Chitaukali WN, Elichova M, Kubova J, Burda H (2004)Microclimatic stability in burrows of an Afrotropical solitary bathyergid rodent,the silvery mole-rat (Heliophobius argenteocinereus). J Zool 263: 409416.

    51. Hall LO, Myers K (1978) Variations in the microclimate in rabbit warrens insemi-arid New South Wales. Aust J Ecol 3: 187194.

    52. Osacar-Jimenez JJ, Lucientes-Curdi J, Calvete-Margolle C (2001) Abiotic factorsinfluencing the ecology of wild rabbit fleas in north-eastern Spain. Med VetEntomol 15: 15766.

    53. Haas GE (1965) Comparative suitability of the four murine rodents of Hawaii ashosts for Xenopsylla vexabilisand X. cheopis(Siphonaptera). J Med Entomol 275-83.

    54. Shenbrot G, Krasnov B, Khokhlova I, Demidova T, Fielden L (2002) Habitat-dependent differences in architecture and microclimate of the burrows ofSundevalls jird (Meriones crassus) (Rodentia: Gerbillinae) in the Negev Desert,Israel. J Arid Environ 51: 265279.

    55. Ryckman RE (1971) Plague vector studies. II. The role of climatic factors indetermining seasonal fluctuations of flea species associated with the Californiaground squirrel. J Med Entomol 8: 541549.

    56. Klein JM (1966) Donnees ecologiques et biologiques sur Synopsyllusfonquerniei, puce de rat peridomestique dans la region de Tanarive [InFrench]. Cahiers Orstom Entomologie medicale 4: 329.

    57. Klein JM, Uilenberg G (1966) Donnees faunistiques et ecologiques sur les pucesde Madagascar (Siphonaptera) [In French]. Cahiers Orstom Entomologie 16:3160.

    58. Chanteau S (2006) Atlas de la peste a Madagascar. Paris: IRD Editions.

    59. Krasnov B, Khokhlova I, Fielden L, Burdelova NV (2002) The effect of substrateon survival and development of two species of desert fleas (Siphonaptera:Pulicidae). Parasite 9: 135142.

    60. Adjemian JCZ, Girvetz EH, Beckett L, Foley JE (2006) Analysis of GeneticAlgorithm for Rule-Set Production (GARP) modeling approach for predictingdistributions of fleas implicated as vectors of plague, Yersinia pestis, in California.

    J Med Entomol 43: 93103.

    61. IPPC IP on CC (2007) Climate Change 2007: the physical science basis -summary for policymakers. Contribution of Working Group I to the Fourth

    Assessment Report of the Intergovernmental Panel on Climate Change. Geneva:IPPC.

    62. Neerinckx S, Peterson AT, Gulinck H, Deckers S, Kimaro D, et al. (2010)

    Predicting potential risk areas of human plague for the Western Usambaramountains, Lushoto district, Tanzania. Am J Trop Med Hyg 82: 492500.

    63. Davis S, Calvet E, Leirs H (2005) Fluctuating rodent populations and risk tohumans from rodent-borne zoonoses. Vect Born Zoo Dis 5: 305314.

    64. Mills JN, Ksiazek TG, Peters CJ, Childs JE (1999) Long-term studies ofhantavirus reservoir populations in the southwestern United States: a synthesis.Emerg Infect Dis 5: 13542.

    65. Collinge SK, Johnson WC, Ray C, Matchett R, Grensten J, et al. (2005) Testingthe generality of a trophic-cascade model for plague. EcoHealth 2: 102112.

    66. Stapp P (2007) Trophic cascades and disease ecology. EcoHealth 4: 121124.

    67. Stenseth NC, Samia NI, Viljugrein H, Kausrud KL, Begon M, et al. (2006)Plague dynamics are driven by climate variation. Proc Natl Acad Sci U S A 103:1311013115.

    68. Harvell CD, Mitchell CE, Ward JR, Altizer S, Dobson AP, et al. (2002) Climatewarming and disease risks for terrestrial and marine biota. Science 296:21582162.

    69. Nakazawa Y, Williams R, Peterson AT, Mead P, Staples E, et al. (2007) Climatechange effects on plague and tularemia in the United States. Vect Born Zoo Dis7: 529540.

    70. Snall T, Benestad RE, Stenseth NC (2009) Expected future plague levels in awildlife host under different scenarios of climate change. Glob Change Biol 15:500507.

    71. Davis S, Trapman P, Leirs H, Begon M, Heesterbeek JAP (2008) Theabundance threshold for plague as a critical percolation phenomenon. Nature454: 635637.

    PLoS Pathogens | www.plospathogens.org 6 September 2011 | Volume 7 | Issue 9 | e1002160