You are on page 1of 11

| |

Received: 9 January 2018    Revised: 8 May 2018    Accepted: 27 May 2018

DOI: 10.1111/zph.12489

REVIEW

Ecosystem change and zoonoses in the Anthropocene

Barry J. McMahon1  | Serge Morand2 | Jeremy S. Gray3

1
UCD School of Agriculture & Food
Science, University College Dublin, Dublin Abstract
4, Ireland Changes in land use, animal populations and climate, primarily due to increasing
2
CNRS – CIRAD ASTRE, Faculty of
human populations, drive the emergence of zoonoses. Force of infection (FOI), which
Veterinary Technology, Kasetsart University,
Bangkok, Thailand for these diseases is a measure of the ease with which a pathogen reaches the human
3
UCD School of Biology & Environmental population, can change with specific zoonoses and context. Here, we outline three
Science, University College Dublin, Dublin
ecosystem categories—domestic, peridomestic and sylvatic, where disease ecology
4, Ireland
alters the FOI of specific zoonoses. Human intervention is an overriding effect in the
Correspondence
emergence of zoonoses; therefore, we need to understand the disease ecology and
Barry J. McMahon, UCD School of
Agriculture & Food Science, University other influencing factors of pathogens and parasites that are likely to interact differ‐
College Dublin, Belfield, Dublin 4, Ireland.
ently within ecological and cultural contexts. Planning for One Health and commu‐
Email: barry.mcmahon@ucd.ie
nity ecology, such as an ecological impact assessment, is required to prepare and
Funding information
manage the emergence and impact of zoonoses in the Anthropocene.
European Cooperation in Science and
Technology, Grant/Award Number: TD1404;
ANR FutureHealthSEA, Grant/Award KEYWORDS
Number: ANR-17-CE35-0003-01 biodiversity, infectious disease, landscape, one Health

1 |  I NTRO D U C TI O N (Hoekstra & Wiedmann, 2014). There are clear trade‐offs with the
management of land‐use changes including the regulation of infec‐
Zoonotic diseases, those transmitted from nonhuman animals to tious diseases (Foley et al., 2005; Karesh et al., 2012).
humans, are increasingly important in global public health (Jones
et al., 2008). Almost 60% of human pathogens and approximately
1.1 | Force of infection
60% emerging infectious diseases are classified as zoonotic (Jones et
al., 2008; Woolhouse & Gowtage‐Sequeria, 2005). Although many Force of infection (FOI) is defined as the rate at which susceptible
human diseases have zoonotic origins (e.g., avian influenza, HIV individuals become infected per unit time and is used to compare
AIDS, SARS and Ebola), we focus our attention here on those that different diseases or transmission in different risk groups. It may
are not normally transmissible between humans or for only a short be expressed in its simplest form as follows (modified from Davis,
period following the index case. The majority of recently emerged Calvet, & Leirs, 2005)
zoonotic diseases originated in wildlife species (Jones et al., 2008).
With an ever‐increasing global human population (Gerland et al., λ = cpv
2014) comes an associated spatial expansion of human activity (e.g.,
agriculture, urbanization and land‐use change). The result is that where λ, force of infection; c, the number of transmission contacts
the interface between humans, domesticated animals and wildlife over a particular time period; p, the level of environmental contamina‐
is likely to intensify during the Anthropocene (Jones et al., 2013), tion or prevalence of infection in the reservoir host or vector popula‐
while wildlife is undergoing a massive defaunation (Ceballos, Ehrlich, tion; and v, the probability that transmission occurs on contact.
& Dirzo, 2017). The terrestrial planet is increasingly a cultural rather For zoonoses, we propose the FOI is a measure of the ease
than a natural landscape because of the anthropogenic environ‐ with which an infection reaches the human population from an an‐
mental footprint, thus requiring management of limited resources imal origin, and can be changed by variation in several factors, the

Zoonoses Public Health. 2018;65:755–765. © 2018 Blackwell Verlag GmbH |  755


wileyonlinelibrary.com/journal/zph  
|
756       McMAHON et al.

interrelationships of which are illustrated in Figure 1. The effective‐


ness of intervention strategies (prevention and control) influences
Impacts
FOI. Thus, preintervention, the FOI for a given disease, can be re‐
ferred to as “intrinsic FOI” and postintervention as “actual FOI,” • Zoonosis emergence can be a consequence of land‐use
the difference between the two being a measure of intervention change and other forms of human intervention.
effectiveness. In practice, accurate estimation of FOI for a disease • The nature of force of infection must be appreciated to
is very difficult because of the large number of variables. For exam‐ understand zoonosis emergence within different
ple, to determine the dynamics of force of infection in Echinococcus ecosystems.
multilocularis in red foxes, a mathematical modelling approach was • We focus on ecosystem classification to categorize
required, which showed that FOI was periodic with variable ampli‐ emergence of zoonoses in order to aid the understand‐
tude, differing markedly with season and between urban and peri‐ ing and role of force of infection with reference to fu‐
urban habitats, suggesting that optimal control strategies must be ture research, public health policy and landscape
tailored to local situations (Lewis, Otero‐Abad, Hegglin, Deplazes, & planning.
Torgerson, 2014).

F I G U R E 1   Interrelationships of factors that contribute to “Force of Infection” leading to zoonotic disease (P&C = prevention and control).
In this simplified scheme, the arrows indicate effects of one factor on another—in relation to the probability of a zoonotic organism causing
disease in the human population. For example, land development and/or ecological perturbation can influence the environmental availability
of zoonotic agents, directly or via other organisms (i.e., intermediate hosts, reservoir hosts and vectors), thus contributing to the likelihood
of infection occurring (force of infection) and so to the occurrence of disease, subject to the susceptibility of the human population. The
broken lines indicate that in a limited number of cases, the zoonotic agent is affected by a predator–prey relationship between reservoir and
intermediate host, for example, foxes and rodents in the case of Echinococcus multilocularis
McMAHON et al. |
      757

dams and irrigation projects. For example in certain parts of Africa,


1.2 | Drivers of zoonotic disease emergence and
the risk of schistosomiasis is increased where such structures im‐
transmission
prove the habitat for the snail intermediate host of the parasite
The emergence of zoonoses has been linked to a range of factors (Steinmann, Keiser, Bos, Tanner, & Utzinger, 2006).
including environmental, ecological and geographical variables On a larger scale, economic globalization was implicated as
(Jones et al., 2008; Plowright et al., 2017). The transmission dynam‐ a driver in the risk of scrub typhus in Taiwan (Kuo et al., 2012).
ics of many zoonotic pathogens are complicated due to the ecology The rickettsial organism that causes scrub typhus, Orientia tsut-
of animal reservoirs and their vectors, depending on the zoonosis sugamushi, utilizes trombiculid mites as vectors and certain rodent
involved. The factors affecting the emergence and transmission of species as reservoir hosts. Both mites and rodents are strongly asso‐
zoonoses can be broadly categorized as follows. ciated with heterogenous forest and overgrown fallow land (Chaisiri,
Cosson, & Morand, 2017; Xu, Walker, Jupiter, Melby, & Arcari, 2017).
When Taiwan joined the World Trade Organization in 2001, local
1.2.1 | Human intervention
rice production became economically uncompetitive because of the
Humans influence, at some level, the emergence and/or trans‐ removal of trade barriers and agricultural subsidies, and this had
mission of practically all zoonoses, but in this section, we present profound effects on the agricultural landscape. The percentage of
some specific examples of human activity as the primary driver. uncultivated paddy fields increased from 20% before 2001 to nearly
Deforestation is a common source of zoonotic disease emergence, 45% by the end of 2004 (to finally stabilize at 40% in 2012), creating
resulting mainly from increased proximity of wild animals to the ideal conditions for both the reservoir hosts and vectors of the scrub
human population. Important pathogens to have emerged in this typhus zoonotic agent. Not only did the rodent population expand,
way over the last few decades include the highly pathogenic Ebola particularly that of the striped field mouse, Apodemus agraruis, but
virus and human immunodeficiency virus (HIV). In both cases, the the intensity of mite infestation of mice from uncultivated fields in‐
genetic relatedness of wild primates (great apes) to humans is prob‐ creased threefold (Kuo et al., 2012). The incidence of scrub typhus
ably relevant. For Ebola, the primates are not the original source of increased nearly 40%, comparing the two periods 1998–2001 and
the virus, but may function as secondary or bridging hosts to humans 2003–2007, and it is tempting to ascribe this to the changed envi‐
(Del Rio & Guarner, 2015). HIV on the other hand is derived from ronmental conditions stemming from globalization.
a simian form of the virus which adapted to its human host (Sharp More subtle effects resulting from human intervention include
& Hahn, 2011). In both cases, once the virus reached the human the rise in rabies incidence in India attributed to the widespread use
population, transmission could occur from human to human and in of the anti‐inflammatory pharmaceutical diclofenac in cattle. This
the case of HIV eventually caused the current AIDS pandemic. More proved toxic to vultures causing a rapid decline in the vulture pop‐
recent examples of deforestation resulting in new zoonotic diseases, ulation, resulting in an accumulation of cattle carcases in the land‐
again involving nonhuman primates, include the monkey malarias, scape. This in turn boosted the populations of stray dogs, the main
caused by Plasmodium knowlesi and P. cynomolgi in South‐East Asia, source of the rabies virus (Markandya et al., 2008). A more targeted
and P. simium in South America (Fornace et al., 2016). It is very likely intervention, also with unintended consequences, was the dissem‐
that other simian Plasmodium species can cause disease in humans, ination of an antirabies vaccine for foxes in Europe, resulting in an
that human Plasmodium species can infect monkeys and that such expansion of the red fox (Vulpes vulpes) population and a consequent
reciprocal infections will increase in frequency with increasing de‐ increase in the transmission of Echinococcus multilocularis, the cause
forestation (Ramasamy, 2014). of alveolar hydatid disease. This topic is addressed in more detail as
Other examples include conversion of land to agriculture or a case study below.
intensification of agricultural practices resulting in an increase in
the transmission of zoonotic diseases, exemplified by the diseases
1.2.2 | Changes in animal populations
caused by West Nile virus and Nipah virus among others (Epstein,
Field, Luby, Pulliam, & Daszak, 2006; Kilpatrick, 2011). On the con‐ As described above for schistosomiasis, scrub typhus, rabies and al‐
trary, the reversion of agricultural land to tick‐permissive woodland veolar hydatid disease, human influence can change animal numbers
appears to be behind the emergence of the tick‐borne disease Lyme (snails, mites, stray dogs and red foxes, respectively) and thus affect
borreliosis on the eastern seaboard of the United States (Matuschka the extent and timing of zoonotic disease emergence. Nonhuman
& Spielman, 1986). In the same way, an increase in tick‐borne en‐ factors can also bring about changes in animal numbers and thus
cephalitis in eastern Europe can be partially attributed to firstly the have significant effects on the dynamics of zoonotic infections. This
creation of suitable tick habitat by the abandonment of farmland in is known to be the case for perhaps the most notorious zoonotic
the economic upheavals following the collapse of the Soviet Union disease in human history, bubonic plague, caused by the bacterium
(Šumilo et al., 2008) and secondly increased use of forest by the Yersina pestis, which is acknowledged as the causal agent of the
human population for the harvesting of mushrooms and wild fruit “Black Death,” which killed as much as a third of the population of
(Randolph et al., 2008). Other more dramatic alterations to the land‐ Europe in the Middle Ages. The infection probably originated in cen‐
scape that may influence the emergence of zoonoses are man‐made tral Asia where it is still maintained by burrowing rodents, mainly
|
758       McMAHON et al.

species of gerbils, and is transmitted by fleas (Gage & Kosoy, 2005). fever was associated with periods of high temperature in Europe
Periodically, these rodents become hyperabundant, resulting even‐ (McIntyre et al., 2017; Morand, Owers, Waret‐Szkuta, McIntyre, &
tually in a Y. pestis epizootic and die‐off. The high mortality of the Baylis, 2013). Similar patterns have been observed for these diseases
rodents deprives infected fleas of their natural hosts, so that they in tropical regions (Wilson, Lush, & Baker, 2010); importantly, their
then seek other hosts, including humans, thus giving rise to plague study also revealed that zoonotic pathogens are more climate‐sen‐
epidemics (Samia et al., 2011). It is probable that climate factors are sitive than human‐ or animal‐only pathogens. Climate and weather
involved in the generation of such epizootics, by driving up both ro‐ changes have consequences in relation to habitat management, es‐
dent and flea numbers (Stenseth et al, 2006), but once the infection pecially concerning food and water resources and movement and/
spills over into other ecosystems, the epidemiology becomes more or migration of animals, including those belonging to climate refu‐
complex. The three pandemics that have been recorded (the Black gees (Gortazar et al., 2014; Pongsiri et al., 2009). All of these may
Death occurring at the start of the second), probably all had their facilitate increased contact between infected animals and suscepti‐
origin in climate‐driven rodent population dynamics in central Asia, ble humans through increased range expansion or range overlap and
and the apparent persistence of cases in areas where such rodents population growth (as described in the previous section) or aggrega‐
are absent, may in fact be due to repeated waves of infection from tion (Gortazar et al., 2014). These processes originate at a local level
the central Asian source spilling over into other rodent populations, but can have global consequences (e.g., outbreaks of avian influenza
such as that of the Black rat, Rattus rattus (Schmid et al., 2015). in humans, which usually have zoonotic origins).
Hantavirus and related pathogens also utilize rodents as reser‐
voir hosts, and there is strong evidence that major drivers of out‐
1.3 | Effects of ecosystems on zoonoses
breaks caused by these viruses can be climatic factors causing peaks
in rodent numbers, particularly in rural areas (Jonsson, Figueiredo, & Evidence indicates that the susceptibility of humans to a specific zo‐
Vapalahti, 2010). For example, Schwarz, Ranft, Piechotowski, Childs, onosis can be highly variable (Civitello et al., 2015; Jones et al., 2008;
and Brockmann (2009) found that case rates of Puumala virus infec‐ Salkeld, Padgett, Jones, & Lafferty, 2013; Wood et al., 2016), and
tion in Germany were significantly associated with mast years which exploration of epidemiological and ecological principles can provide
boost rodent numbers in China, an abundance of field mice resulting further insights. A primary ecological component of any zoonosis
from low rainfall and lack of flooding preceded a higher incidence is the habitat occupied by the reservoir host(s). Although there are
of HFRS (haemorrhagic fever with renal syndrome) in China (Guan detailed ecological descriptions of various habitats using vegetation
et al., 2009), and in America, the emergence of hantavirus has been or urban categories to obtain greater resolution within classifica‐
linked to precipitation that indirectly stimulates rodent population tion categories, for example, CORINE (Co‐ORdinated INformation
growth (Engelthaler et al., 1999; Yates et al., 2002). on the Environment), for this study, following Schwabe (Schwabe,
In contrast to disease emergence caused by animal numbers that 1984), we categorize three major host habitat types or ecosystems
have increased, there is at least one case where a decrease may have for zoonoses: “domestic,” “domiciliated” (or “peridomestic”) and “syl‐
contributed to higher disease rates. The incidence of tick‐borne en‐ vatic.” The terms “domiciliated,” “synanthropic” and “peridomestic”
cephalitis (TBE) in Sweden has risen over recent decades apparently are interchangeable, and here, “peridomestic” is used. Different eco‐
linked to a decrease in roe deer (Capreolus capreolus) populations, an systems vary in species community and composition.
important host for all stages of the vector tick, Ixodes ricinus. This is Potential exposure of humans usually decreases through the
thought to have diverted host‐seeking ticks to bank voles (Myodes ecosystem series “domestic,” “peridomestic” and “sylvatic,” whereas
glareolus), animal reservoirs of TBE, which coincidentally peaked in biodiversity can be expected to increase. In domestic ecosystems,
numbers at the same time, resulting in increased transmission to hu‐ force of infection (FOI) generally increases as a result of an increase
mans (Jaenson, Hjertqvist, Bergström, & Lundkvist, 2012). in pathogen abundance, often caused for example by systems fail‐
ures in abattoirs and by the amount of movement of the reservoir
hosts and/or mechanical vectors. In addition, animal breeding for
1.2.3 | Changes in climate and weather
certain desirable traits may compromise immunogenetics. In sylvatic
Empirical observations confirm that ongoing climate change is driv‐ ecosystems, exposure of humans is less frequent, but FOI may in‐
ing certain diseases to higher latitudes or altitudes (Altizer, Ostfeld, crease or decrease as a result of changes in human intrusion and/
Johnson, Kutz, & Harvell, 2013). However, most of the knowledge or habitat degradation. Seasonal and annual variations in reservoir
on the effects of climate change regarding the distribution and epi‐ host populations, as well as changes in species composition, are
demiology of infectious diseases comes from modelling approaches, important FOI factors in sylvatic zoonoses. The dilution effect, as
such as ecological niche modelling or climate simulation‐based risk a general concept, suggests that naturally occurring biodiversity,
modelling (Guis et al., 2012). McIntyre et al. (2017) recently investi‐ specifically species diversity, could reduce disease risk, including
gated the climate sensitivity of infectious pathogens and found that zoonoses (Keesing, Holt, & Ostfeld, 2006). For example, the infec‐
two‐thirds of them were climate‐sensitive (rainfall and tempera‐ tion prevalence of hantaviruses in wild rodent reservoir populations
ture). Enterovirus infection and shigellosis were associated with wet and reservoir population density increased where small‐mammal
summer climate and abnormally heavy rainfalls, whereas typhoid species diversity was reduced in an experimental study carried out
McMAHON et al. |
      759

in Panama (Suzán et al., 2009). The common appeal of the dilution Escherichia coli O157 (Figure 2) is a bacterial pathogen that causes
effect to public health and conservation is clear as with one initia‐ gastroenteritis, occasionally resulting in life‐threatening kidney dys‐
tive (i.e., increased biodiversity), multiple benefits could potentially function (Newell et al., 2010). The level of intensification of beef
result, such as reduced human health risk and reduced infectious dis‐ production is clearly a factor and outbreaks have been associated
ease in wildlife (Keesing et al., 2010; Pongsiri et al., 2009). with the linking of intensification and biodiversity loss, resulting in
The factors responsible for FOI in peridomestic ecosystems, increased prevalence of the pathogen in cattle, the most important
in which wild animals occur in close association with humans, for reservoir host (Donald, Green, & Heath, 2001; Jones et al., 2013).
example black rat (Rattus rattus) and Norwegian rat (R. norvegicus), While the main risk to the human population is failure of food‐
can be a mix of those in the other two ecosystems. When zoonoses processing safety systems such as appropriate application of hazard
involve arthropod vectors or parasite intermediate hosts, there are analysis and critical control points (HACCP), management at farm
more factors that might influence FOI and the system becomes more level also has a role to play because of an association between in‐
complex, so that effective intervention becomes more difficult. fection prevalence in cattle entering abattoirs and subsequent con‐
We use specific examples of zoonoses to demonstrate how the tamination levels (Loneragan & Brashears, 2005). Several preharvest
ecological requirements of zoonotic pathogens in particular ecosys‐ intervention measures, such as feeding probiotics or sodium chlorate
tems can alter the FOI of the relevant zoonoses. The public health in feed or water, have shown promise, but so far, experimental re‐
relevance of the dilution effect is discussed in relation to zoonoses sults have been inconsistent. There is no evidence that targeted an‐
prevalence and management. The information discussed could be tibiotic use reduces shedding or that antimicrobial growth promoters
used to inform monitoring, surveillance and planning for a balanced increase shedding (Sargeant, Amezcua, Rajic, & Waddell, 2007).
One Health approach. For non‐food‐borne domestic zoonoses, cultural factors in the
human population are important. For example, the prevalence of the
nematode Toxocara canis and the cestode Echinococcus granulosus,
2 | C A S E S T U D I E S
both of which utilize domestic dogs and other canids as definitive
hosts, is associated with lack of worming of dogs and the environ‐
2.1 | Domestic ecosystem
mental dissemination of dog faeces. Although sheep are interme‐
It is understood that food‐borne infections are among the most im‐ diate hosts of E. granulosus, this infection is not associated with
portant zoonoses worldwide (EFSA, 2017), and intensification of intensification of agriculture, but rather with certain extensive agri‐
food producing systems has increased the associated risk. As the cultural systems in which knowledge of the risk is low (Otero‐Abad
number of farm animals serving as reservoir hosts rises, so does the & Torgerson, 2013).
FOI, thus increasing the possibility of human infection (Pennington,
2010). In parallel with this, the industrialization of food‐processing
2.2 | Peridomestic ecosystem
systems has increased the contamination of food with such patho‐
gens as Salmonella spp., Campylobacter jejuni and certain strains of Wild animals inhabiting the domestic environment are part of
Escherichia coli. the peridomestic ecosystem and may contribute to peridomestic

F I G U R E 2   Relationships between
the pathogens, Escherichia coli, Leptospira
spp. and Echinococcus multilocularis with
their respective reservoirs hosts (cattle,
rodents, fox/rodents) and habitats
(domestic, peridomestic and sylvatic)
indicating the potential for human disease
|
760       McMAHON et al.

zoonoses. A well‐known example is leptospirosis (Figure 2), a wide‐ Europe, red fox Vulpes vulpes, as well as domestic dogs) as definitive
spread and neglected worldwide bacterial spirochaete zoonosis hosts and rodents as intermediate hosts. The abundance of foxes is
(Bharti et al., 2003), for which rodents, particularly the brown rat, a major contributor to the FOI of this pathogen. The red fox popula‐
Rattus norvegicus, are the principal reservoirs. The introduction of tion has increased in European regions over the last few decades and
an invasive mammalian species by humans can, unintentionally, alter this has been attributed in part to the success of a vaccine against
reservoir host community composition, and thus change the dynam‐ fox rabies, in conjunction with the adaptability of foxes to the urban
ics of this zoonosis (Hubálek, 2003; Nally et al., 2016). Alterations in landscape (Otero‐Abad & Torgerson, 2013). The fox population was
rodent communities (Rattus rattus, R. norvegicus and R. exculans) as a evidently held in check to some extent by the rabies virus, and when
result of the introduction of the black rat, R. rattus, to the island of this pressure was removed by the antirabies vaccination programme,
Futuna (Western Polynesia), increased leptospirosis risk in humans fox abundance increased, leading to increased transmission of the
(Theuerkauf et al., 2013), thus demonstrating how an alteration in cestode zoonosis (Combes et al., 2012; Schneider, Aspöck, & Auer,
community structure and composition of wild rodent species can in‐ 2013; Schweiger et al., 2007). In addition, generally increased risk
fluence the FOI of peridomestic zoonoses. In developing countries, is associated with humans entering sylvatic ecosystems, whereas
it is a socioeconomic disease exacerbated by urbanization, rodent specific risk factors include owning a dog that hunts game, living in
infestation and transmission via contaminated water sources asso‐ a farmhouse and being a farmer therefore increasing the likelihood
ciated with inadequate infrastructures and severe flooding events of contact with the definitive hosts (Kern et al., 2004). This is an ex‐
(Bharti et al., 2003). Elevated contact with spirochaetes within ro‐ ample of an intentional measure directed at one particular zoonosis
dent‐infested areas as a result of behaviour change in risk groups (rabies) having an unintentional effect on another (E. multilocularis)
(e.g., farmers, veterinarians and water sports enthusiasts) is another and demonstrates the complexity of the eco‐epidemiological inter‐
determining factor in leptospirosis incidence (Garvey, Connell, actions that can occur in zoonoses.
O’Flanagan, & McKeown, 2014).

2.4 | Vector‐borne zoonoses
2.3 | Sylvatic ecosystem
Vector‐borne zoonoses, similar to those involving parasite interme‐
The sylvatic cestode, Echinococcus multilocularis (Figure 2), causes in‐ diate hosts, add an additional layer of complexity to the transmission
fections that can have a poor prognosis in humans, and it appears the dynamics of the diseases they cause, because both the vectors and
dynamics of infection with this parasite has altered in recent years in the zoonotic pathogens must be maintained within the ecosystem.
parts of Europe. E. multilocularis utilizes wild and domestic canids (in The most common vector‐borne disease in the temperate northern

F I G U R E 3   Relationship between
the vector, pathogens, reservoirs hosts,
habitats and the potential for Lyme
borreliosis in humans. Vector ticks
(a) transmit the pathogens, Borrelia
burgdorferi sensu lato (b), to small
mammals and/or birds (c). The ticks also
acquire the infection from these animals
and thus pose an infection risk to humans
(d) in both sylvatic (e) and peridomestic
(f) habitats. Ticks do not acquire the
infection from large animals such as
deer (g) (in sylvatic and peridomestic
habitats), or from sheep and cattle (h) in
the domestic habitat (i), but these hosts
are essential for maintenance of the tick
population. Ticks that have fed exclusively
on such hosts do not pose an infection
risk when they infest humans (j)
McMAHON et al. |
      761

hemisphere, Lyme borreliosis (LB) (Figure 3), has attracted much at‐ 3 | M A N AG E M E NT O F ZO O N OS I S R I S K
tention in relation to the hypothesis that biodiversity can, through
the dilution effect, reduce transmission rates and thus the risk of Effective management of zoonoses depends on greater advance‐
disease (LoGiudice, Ostfeld, Schmidt, & Keesing, 2003; Ostfeld & ment in the following to understand zoonoses: delineation of key
Keesing, 2012). However, robust arguments have been made against drivers of the zoonotic outbreaks, gathering of more empirical
universal application of this hypothesis (Lafferty & Wood, 2013; data from both the field and laboratory experiments and further
Ogden & Tsao, 2009; Randolph & Dobson, 2012). LB is caused by exploration of the conceptual frameworks that develop our think‐
members of the Borrelia burgdorferi sensu lato (s.l.) genospecies com‐ ing around the complexity of these outbreaks (Johnson, Ostfeld,
plex and is transmitted by ticks of the Ixodes ricinus species complex. & Keesing, 2015). Future research should explore and harness
The reservoir hosts of the pathogens are mainly small mammals and knowledge of the disease ecology and other influencing factors
ground‐feeding passerine birds, but large mammals, though reser‐ of pathogens and parasites that are likely to interact differently
voir incompetent, are required for the ticks to complete their life cy‐ within dynamic landscapes in the Anthropocene in the coming
cles (Gray et al., 1998). Deer rather than livestock are most typically years to inform risk assessment. Combined with this, greater un‐
associated with LB, because they usually occur in the same habitats derstanding is also required for the life history traits of the zo‐
as the spirochaete reservoir hosts (Figure 3). Thus humans are at risk onosis vertebrate reservoirs, parasite intermediate hosts and
of infection when they enter woodland (sylvatic ecosystem) or when arthropod vectors, (where relevant) to provide a comprehensive
the reservoir hosts and deer occur in periurban (i.e., peridomestic) understanding of the likely risk. Linking these ecological factors
ecosystems (Rizzoli et al., 2014). Domestic livestock such as cattle with particular landscape changes will enable authorities to offset
and sheep are not generally associated with LB, mainly because they the risk of particular zoonoses in a specific set of circumstances.
are usually farmed in open habitats that do not permit prolonged As outlined in this paper, given the risks surrounding zoonoses
survival of the desiccation‐susceptible nonparasitic stages of the emergence and land‐use change (Jones et al., 2013; Kilpatrick,
tick life cycle. Nevertheless, large tick populations can be maintained 2011; McCauley et al., 2015), where large‐scale landscape change
by sheep and cattle in marginal farmland (which can feed all stages) is proposed (e.g., alteration or intensification of agricultural lands
in the virtual absence of other hosts (Milne, 1949). In such scenarios, and urbanization), we have now reached a point in resource man‐
the ticks are likely to show a low to zero prevalence of infection with agement where an ecological impact assessment, similar to an
B. burgdorferi s.l. because of the relatively small numbers of reservoir environmental impact assessment (McMahon, Wall, Fanning, &
hosts present (Figure 3) (Gray, Kahl, Janetzki, Stein, & Guy, 1995; Fahey, 2015) and in which potential disease is taken into account
Matuschka et al., 1993). This situation is in marked contrast to the (CIEEM, 2016), should be a requirement for such alterations to
generalization that there is a high risk of LB wherever there are ap‐ proceed. The specifications for an ecological impact assessment
preciable numbers of ticks (Jaenson et al., 2009). require further consideration but landscape ecology could inform
The survival of ticks in undergrazed agricultural habitats, which this process (Hartemink, Vanwambeke, Purse, Gilbert, & Dyck,
is typical of Ireland, particularly west of the Shannon River, also oc‐ 2015) using computer programs such as FRAGSTATS (McGarigal,
curs in continental Europe but to a much lesser extent. In such areas, Cushman, & Ene, 2012) to quantify landscape configuration and
an increase in biodiversity resulting from reforestation might actu‐ composition. The juxtaposition of the domestic, peridomestic and
ally increase tick infection rates, in contradiction to the situation de‐ sylvatic ecosystems influences zoonoses dynamics for both host
scribed for parts for the United States, where it has been suggested and pathogen/parasite populations (Kilpatrick, 2011; McCauley
that increased biodiversity reduces the FOI by diverting part of the et al., 2015). As intimated above, the importance of integration
tick population to hosts that are not reservoirs of the tick‐borne of epidemiology and community ecology for understanding zoon‐
pathogen B. burgdorferi s.l. (Ostfeld & Keesing, 2012). The current oses is clear due to the multispecies and multiple scales involved
low tick infection rates in farmland‐dominated landscapes, such as in (Johnson, de Roode, & Fenton, 2015) and is further highlighted in
Ireland, and the common occurrence of bird‐associated genospecies the examples outlined in this paper. Diseases, zoonoses or other‐
such as B. garinii (some strains of which are zoonotic), in north‐west‐ wise, are components of ecological systems; therefore, eradica‐
ern Europe (Kurtenbach et al., 2006; Pichon, Rogers, Egan, & Gray, tion is unlikely to result in entirely satisfactory outcomes because
2005; Rauter & Hartung, 2005; Saint Girons et al., 1998), could make the disease is a natural ecological process and other pathogens or
important contributions to a rise in infection rates, considering that parasites tend to occupy the vacated niches (Lloyd‐Smith, 2013).
afforestation and reforestation increase bird species richness and However, community ecology in conjunction with epidemiology
diversity (Graham et al., 2015). The process of reforestation in many can bring about a greater understanding of the processes involved
parts of western Europe is already in progress, and it remains to be in zoonotic outbreaks and facilitate better public health manage‐
seen whether this results in a rise in the incidence of LB in these ment (Cunningham, Scoones, & Wood, 2017; Johnson, de Roode,
areas. This issue illustrates the complexity that can exist in the dy‐ et al., 2015; Young et al., 2017). It is well accepted that there is
namics of transmission of a pathogen that circulates among several unprecedented pressure on the planet to sustain ever‐increasing
species of vertebrate host and has a vector, I. ricinus, which feeds on numbers of humans (Gerland et al., 2014). The interconnectedness
both reservoirs and nonreservoirs of the pathogens. of humans, livestock and wildlife is core to many of our current
|
762       McMAHON et al.

global challenges. It is important to understand the impacts that to a predictive framework. Science, 341, 514–519. https://doi.
changes in human populations have on pathogen exchange be‐ org/10.1126/science.1239401
Bharti, A. R., Nally, J. E., Ricaldi, J. N., Matthias, M. A., Diaz, M. M., &
tween humans, wildlife and livestock in varying ecological and cul‐
Lovett, M. A., …Peru‐United States Leptospirosis Consortium
tural contexts (Jones et al., 2013; Muehlenbein, 2016). (2003). Leptospirosis: A zoonotic disease of global importance.
Lancet Infectious Diseases, 3, 757–771. https://doi.org/10.1016/
S1473-3099(03)00830-2
Ceballos, G., Ehrlich, P. R., & Dirzo, R. (2017). Biological annihilation
4 |  CO N C LU S I O N
via the ongoing sixth mass extinction signaled by vertebrate pop‐
ulation losses and declines. Proceedings of the National Academy of
Anthropogenic changes, including agricultural specialization and in‐ Sciences of the United States of America, 114, 6089–6096. https://doi.
tensification accompanied by environmental effects, are occurring org/10.1073/pnas.1704949114
Chaisiri, K., Cosson, J.‐F., & Morand, S. (2017). Infection of rodents
constantly and so zoonotic diseases will continue to emerge (Jones
by Orientia tsutsugamushi, the agent of scrub typhus in relation to
et al., 2013). In view of the growth of the human population, it is nec‐ land use in Thailand. Tropical Medicine and Infectious Disease, 2, 53.
essary to understand the domestic, peridomestic and sylvatic eco‐ https://doi.org/10.3390/tropicalmed2040053.
systems in this context in order to predict emergence of zoonoses CIEEM (2016). Guidelines for ecological impact assessment in the UK
and Ireland: Terrestrial, freshwater and coastal, 2nd ed. Winchester,
and system resilience (Suter, 1993). Landscape planning for One
UK: CIEEM. Retrieved from https://www.cieem.net/data/files/
Health and community ecology is required to understand and man‐
Publications/EcIA_Guidelines_Terrestrial_Freshwater_and_Coastal_
age outbreaks of zoonoses (Hassell, Begon, Ward, & Fèvre, 2017) Jan_2016.pdf
incorporating a net health effect for conservation and public health Civitello, D. J., Cohen, J., Fatima, H., Halstead, N. T., Liriano, J., McMahon,
(Myers et al., 2013). Drivers of human diseases, including zoonoses, T. A., … Rohr, J. R. (2015). Biodiversity inhibits parasites: Broad evi‐
dence for the dilution effect. Proceedings of the National Academy of
are context‐ and species‐dependent demonstrating spatial and tem‐
Sciences of the United States of America, 112, 8667–8671. https://doi.
poral variability (Civitello et al., 2015; Keesing et al., 2010; Salkeld et org/10.1073/pnas.1506279112
al., 2013; Salkeld, Padgett, Jones, & Antolin, 2015). Interdisciplinary Combes, B., Comte, S., Raton, V., Raoul, F., Boué, F., Umhang, G., …
collaboration, incorporating One Health land‐use planning, must Giraudoux, P. (2012). Westward spread of Echinococcus multilocu-
laris in foxes, France, 2005–2010. Emerging Infectious Diseases, 18,
appreciate that conducting empirical studies with controls, using
2059–2062. https://doi.org/10.3201/eid1812.120219
standardization, randomization and replication of comparable ex‐ Cunningham, A. A., Scoones, I., & Wood, J. L. N. (2017). One Health
perimental units with the aim of producing results with powerful for a changing world: New perspectives from Africa. Philosophical
inference are not possible without large‐scale investment in long‐ Transactions of the Royal Society of London. Series B, Biological Sciences,
372, 20160162. https://doi.org/10.1098/rstb.2016.0162
term landscape experiments. Until the experiments are conducted,
Davis, S., Calvet, E., & Leirs, H. (2005). Fluctuating rodent popula‐
such local context that drives the emergence of zoonoses must be tions and risk to humans from rodent‐borne zoonoses. Vector
respected. Borne Zoonotic Diseases, 5, 305–314. https://doi.org/10.1089/
vbz.2005.5.305
Del Rio, C., & Guarner, J. (2015). Ebola: Implications and perspectives.
AC K N OW L E D G E M E N T S Transactions of the American Clinical and Climatological Association,
126, 93–112.
The authors would like to thank Bernard Kaye for his help prepar‐ Donald, P. F., Green, R. E., & Heath, M. F. (2001). Agricultural intensi‐
ing the figures and Jarlath Nally, Laura Espinosa, John Fry and an fication and the collapse of Europe's farmland bird populations.
Proceedings of the Royal Society B: Biological Sciences, 268, 25–29.
anonymous reviewer for comments on previous versions of the
https://doi.org/10.1098/rspb.2000.1325
manuscript. BMcM was supported by European Cooperation in
EFSA (European Food Safety Authority), & ECDC (European Centre for
Science and Technology TD1404 Network for the Evaluation of Disease Prevention and Control) (2017). The European Union sum‐
One Health, and SM was supported by ANR FutureHealthSEA mary report on trends and sources of zoonoses, zoonotic agents
(ANR‐17‐CE35‐0003‐01). and food‐borne outbreaks in 2016. EFSA Journal, 15, 5077, 228 pp.
https://doi.org/10.2903/j.efsa.2017.5077
Engelthaler, D. M., Mosley, D. G., Cheek, J. E., Levy, C. E., Komatsu, K. K.,
Ettestad, P., …Bryan, R. T. (1999). Climatic and environmental pat‐
C O N FL I C T O F I N T E R E S T
terns associated with hantavirus pulmonary syndrome, Four Corners
None declared. region, United States. Emerging Infectious Diseases, 5, 87–94. https://
doi.org/10.3201/eid0501.990110
Epstein, J. H., Field, H. E., Luby, S., Pulliam, J. R. C., & Daszak, P. (2006).
Nipah virus: Impact, origins, and causes of emergence. Current
ORCID
Infectious Disease Reports, 8, 59–65. https://doi.org/10.1007/
Barry J. McMahon  http://orcid.org/0000-0003-3143-8075 s11908-006-0036-2
Foley, J. A., DeFries, R., Asner, G. P., Barford, C., Bonan, G., Carpenter, S.
R., … Snyder, P. K. (2005). Global consequences of land use. Science,
309, 570–574. https://doi.org/10.1126/science.1111772
REFERENCES
Fornace, K. M., Abidin, T. R., Alexander, N., Brock, P., Grigg, M. J., Murphy,
Altizer, S., Ostfeld, R. S., Johnson, P. T. J., Kutz, S., & Harvell, C. D. A., … Cox, J. (2016). Association between landscape factors and spa‐
(2013). Climate change and infectious diseases: From evidence tial patterns of Plasmodium knowlesi infections in Sabah, Malaysia.
McMAHON et al. |
      763

Emerging Infectious Diseases, 22, 201–208. https://doi.org/10.3201/ Johnson, P. T. J., Ostfeld, R. S., & Keesing, F. (2015). Frontiers in research
eid2202.150656 on biodiversity and disease. Ecology Letters, 18, 1119–1133. https://
Gage, K. L., & Kosoy, M. Y. (2005). Natural history of plague: Perspectives doi.org/10.1111/ele.12479
from more than a century of research. Annual Review of Entomology, 50, Jones, B. A., Grace, D., Kock, R., Alonso, S., Rushton, J., Said, M. Y., …
505–528. https://doi.org/10.1146/annurev.ento.50.071803.130337 Udo Pfeiffer, D. (2013). Zoonosis emergence linked to agricultural
Garvey, P., Connell, J., O’Flanagan, D., & McKeown, P. (2014). intensification and environmental change. Proceedings of the National
Leptospirosis in Ireland: Annual incidence and exposures associated Academy of Sciences of the United States of America, 110, 8399–8404.
with infection. Epidemiology and Infection, 142, 847–855. https://doi. https://doi.org/10.1073/pnas.1208059110
org/10.1017/S0950268813001775 Jones, K. E., Patel, N. G., Levy, M. A., Storeygard, A., Balk, D., Gittleman,
Gerland, P., Raftery, A. E., Ševčíková, H., Li, N., Gu, D., Spoorenberg, J. L., & Daszak, P. (2008). Global trends in emerging infectious dis‐
T., … Wilmoth, J. (2014). World population stabilization unlikely eases. Nature, 451, 990–993. https://doi.org/10.1038/nature06536
this century. Science, 346, 234–237. https://doi.org/10.1126/ Jonsson, C. B., Figueiredo, L. T., & Vapalahti, O. (2010). A global per‐
science.1257469 spective on hantavirus ecology, epidemiology, and disease. Clinical
Gortazar, C., Reperant, L. A., Kuiken, T., de la Fuente, J., Boadella, M., Microbiology Reviews, 23, 412–441. https://doi.org/10.1128/
Martínez‐Lopez, B., … Keck, F. (2014). Crossing the interspecies bar‐ CMR.00062-09
rier: Opening the door to zoonotic pathogens. PLoS Path, 10, 1–5. Karesh, W. B., Dobson, A., Lloyd‐Smith, J. O., Lubroth, J., Dixon, M.
https://doi.org/10.1371/journal.ppat.1004129 A., Bennett, M., … Heymann, D. L. (2012). Ecology of zoonoses:
Graham, C. T., Wilson, M. W., Gittings, T., Kelly, T. C., Irwin, S., Quinn, J. Natural and unnatural histories. Lancet, 380, 1936–1945. https://doi.
L., & O’Halloran, J. (2015). Implications of afforestation for bird com‐ org/10.1016/S0140-6736(12)61678-X
munities: The importance of preceding land‐use type. Biodiversity Keesing, F., Belden, L. K., Daszak, P., Dobson, A., Harvell, C. D., Holt,
and Conservation, 1–21, https://doi.org/10.1007/s10531-015-0987-4 R. D., … Ostfeld, R. S. (2010). Impacts of biodiversity on the emer‐
Gray, J. S., Kahl, O., Janetzki, C., Stein, J., & Guy, E. (1995). The spa‐ gence and transmission of infectious diseases. Nature, 468, 647–652.
tial distribution of Borrelia burgdorferi‐infected Ixodes ricinus in the https://doi.org/10.1038/nature09575
Connemara region of County Galway, Ireland. Experimental & Applied Keesing, F., Holt, R. D., & Ostfeld, R. S. (2006). Effects of species di‐
Acarology, 19, 163–172. https://doi.org/10.1007/BF00046288 versity on disease risk. Ecology Letters, 9, 485–498. https://doi.
Gray, J. S., Kahl, O., Robertson, J. N., Daniel, M., Estrada‐Peña, A., org/10.1111/j.1461-0248.2006.00885.x
Gettinby, G., … Zeman, P. (1998). Lyme borreliosis habitat assessment. Kern, P., Ammon, A., Kron, M., Sinn, G., Sander, S., Petersen, L. R., …
Zentralblatt Für Bakteriologie, 287, 211–228. https://doi.org/10.1016/ Kern, P. (2004). Risk factors for alveolar echinococcosis in hu‐
S0934-8840(98)80123-0 mans. Emerging Infectious Diseases, 10, 2088–2093. https://doi.
Guan, P., Huang, D., He, M., Shen, T., Guo, J., & Zhou, B. (2009). Investigating org/10.3201/eid1012.030773
the effects of climatic variables and reservoir on the incidence of Kilpatrick, A. M. (2011). Globalization, land use, and the invasion of
hemorrhagic fever with renal syndrome in Huludao City, China: A 17‐ West Nile virus. Science, 334, 323–327. https://doi.org/10.1126/
year data analysis based on structure equation model. BMC Infectious. science.1201010
Diseases, 8, 109. https://doi.org/10.1186/1471-2334-9-109 Kuo, C.‐C., Huang, J.‐L., Shu, P.‐Y., Lee, P.‐L., Kelt, D. A., & Wang, H.‐C.
Guis, H., Caminade, C., Calvete, C., Morse, A. P., Tran, A., & Baylis, M. (2012). Cascading effect of economic globalization on human
(2012). Modelling the effects of past and future climate on the risk risks of scrub typhus and tick‐borne rickettsial diseases. Ecological
of bluetongue emergence in Europe. Journal of the Royal Society, Applications, 22, 1803–1816. https://doi.org/10.1890/12-0031.1
Interface, 9, 339–350. https://doi.org/10.1098/rsif.2011.0255 Kurtenbach, K., Hanincová, K., Tsao, J. I., Margos, G., Fish, D., & Ogden,
Hartemink, N., Vanwambeke, S. O., Purse, B. V., Gilbert, M., & Van Dyck, N. H. (2006). Fundamental processes in the evolutionary ecology of
H. (2015). Towards a resource‐based habitat approach for spatial Lyme borreliosis. Nature Reviews Microbiology, 4, 660–669. https://
modelling of vector‐borne disease risks: Resource‐based habitats for doi.org/10.1038/nrmicro1475
vector‐borne diseases. Biological Reviews, 90, 1151–1162. https://doi. Lafferty, K. D., & Wood, C. L. (2013). It's a myth that protection against
org/10.1111/brv.12149 disease is a strong and general service of biodiversity conservation:
Hassell, J. M., Begon, M., Ward, M. J., & Fèvre, E. M. (2017). Urbanization Response to Ostfeld and Keesing. Trends in Ecology & Evolution, 28,
and disease emergence: Dynamics at the wildlife–livestock–human 503–504. https://doi.org/10.1016/j.tree.2013.06.012
interface. Trends in Ecology & Evolution, 32, 55–67. https://doi. Lewis, F. I., Otero‐Abad, B., Hegglin, D., Deplazes, P., & Torgerson, P. R.
org/10.1016/j.tree.2016.09.012 (2014). Dynamics of the force of infection: Insights from Echinococcus
Hoekstra, A. Y., & Wiedmann, T. O. (2014). Humanity's unsustainable multilocularis infection in foxes. PLoS Neglected Tropical Diseases., 8,
environmental footprint. Science, 344, 1114–1117. https://doi. e2731. https://doi.org/10.1371/journal.pntd.0002731
org/10.1126/science.1248365 Lloyd‐Smith, J. O. (2013). Vacated niches, competitive release and
Hubálek, Z. (2003). Emerging human infectious diseases: Anthroponoses, the community ecology of pathogen eradication. Philosophical
zoonoses, and sapronoses. Emerging Infectious Diseases, 9, 403–404. Transactions of the Royal Society of London. Series B, Biological Sciences,
https://doi.org/10.3201/eid0903.020208 368, 20120150. https://doi.org/10.1098/rstb.2012.0150
Jaenson, T. G. T., Eisen, L., Comstedt, P., Mejlon, H. A., Lindgren, LoGiudice, K., Ostfeld, R. S., Schmidt, K. A., & Keesing, F. (2003). The
E., Bergström, S., & Olsen, B. (2009). Risk indicators for the ecology of infectious disease: Effects of host diversity and commu‐
tick Ixodes ricinus and Borrelia burgdorferi sensu lato in Sweden. nity composition on Lyme disease risk. Proceedings of the National
Medical and Veterinary Entomology, 23, 226–237. https://doi. Academy of Sciences of the United States of America, 100, 567–571.
org/10.1111/j.1365-2915.2009.00813.x https://doi.org/10.1073/pnas.0233733100
Jaenson, T. G. T., Hjertqvist, M., Bergström, T., & Lundkvist, A. (2012). Loneragan, G. H., & Brashears, M. M. (2005). Pre‐harvest interven‐
Why is tick‐borne encephalitis increasing? A review of the key factors tions to reduce carriage of E. coli O157 by harvest‐ready feed‐
causing the increasing incidence of human TBE in Sweden. Parasites lot cattle. Meat Science, 71, 72–78. https://doi.org/10.1016/j.
& Vectors, 5, 184. https://doi.org/10.1186/1756-3305-5-184 meatsci.2005.04.005
Johnson, P. T. J., de Roode, J. C., & Fenton, A. (2015). Why infectious Markandya, A., Taylor, T., Longo, A., Murty, M. N., Murty, S., &
disease research needs community ecology. Science, 349, 1259504. Dhavala, K. (2008). Counting the cost of vulture decline – An ap‐
https://doi.org/10.1126/science.1259504 praisal of the human health and other benefits of vultures in India.
|
764       McMAHON et al.

Ecological Economics, 67, 194–204. https://doi.org/10.1016/j. Neglected Tropical Diseases, 7, e2249. https://doi.org/10.1371/jour‐
ecolecon.2008.04.020 nal.pntd.0002249
Matuschka, F. R., Heiler, M., Eiffert, H., Fischer, P., Lotter, H., & Spielman, Pennington, H. (2010). Escherichia coli O157. Lancet, 376, 1428–1435.
A. (1993). Diversionary role of hoofed game in the transmission https://doi.org/10.1016/S0140-6736(10)60963-4
of Lyme disease spirochetes. The American Journal of Tropical Pichon, B., Rogers, M., Egan, D., & Gray, J. (2005). Blood‐meal analysis
Medicine and Hygiene, 48, 693–699. https://doi.org/10.4269/ for the identification of reservoir hosts of tick‐borne pathogens in
ajtmh.1993.48.693 Ireland. Vector Borne and Zoonotic Diseases, 5, 172–180. https://doi.
Matuschka, F. R., & Spielman, A. (1986). The emergence of Lyme dis‐ org/10.1089/vbz.2005.5.172
ease in a changing environment in North America and central Plowright, R. K., Parrish, C. R., McCallum, H., Hudson, P. J., Ko, A. I.,
Europe. Experimental & Applied Acarology, 2, 337–353. https://doi. Graham, A. L., & Lloyd‐Smith, J. O. (2017). Pathways to zoonotic
org/10.1007/BF01193900 spillover. Nature Reviews Microbiology, 5, 502–510. https://doi.
McCauley, D. J., Salkeld, D. J., Young, H. S., Makundi, R., Dirzo, R., org/10.1038/nrmicro.2017.45
Eckerlin, R. P., … Helgen, K. M. (2015). Effects of land use on plague Pongsiri, M. J., Roman, J., Ezenwa, V. O., Goldberg, T. L., Koren, H. S.,
(Yersinia pestis) activity in rodents in Tanzania. The American Journal of Newbold, S. C., … Salkeld, D. J. (2009). Biodiversity loss affects global
Tropical Medicine and Hygiene, 92, 776–783. https://doi.org/10.4269/ disease ecology. BioScience, 59, 945–954. https://doi.org/10.1525/
ajtmh.14-0504 bio.2009.59.11.6
McGarigal, K., Cushman, S. A., & Ene, E. (2012). FRAGSTATS v4: Spatial Ramasamy, R. (2014). Zoonotic malaria – Global overview and research
pattern analysis program for categorical and continuous maps. and policy needs. Frontiers of Public Health, 2, 123. https://doi.
Computer software program produced by the authors at the University org/10.3389/fpubh.2014.00123
of Massachusetts, Amherst. Retrieved from https://www.umass.edu/ Randolph, S. E., Asokliene, L., Avsic‐Zupanc, T., Bormane, A., Burri, C.,
landeco/research/fragstats/fragstats.html Gern, L., … Zygutiene, M. (2008). Variable spikes in tick‐borne en‐
McIntyre, K. M., Setzkorn, C., Hepworth, P. J., Morand, S., Morse, cephalitis incidence in 2006 independent of variable tick abundance
A. P., & Baylis, M. (2017). Systematic assessment of the climate but related to weather. Parasites and Vectors, 1, 44. https://doi.
sensitivity of important human and domestic animals pathogens org/10.1186/1756-3305-1-44
in Europe. Scientific Reports, 7, 7134. https://doi.org/10.1038/ Randolph, S. E., & Dobson, A. D. M. (2012). Pangloss revisited: A cri‐
s41598-017-06948-9 tique of the dilution effect and the biodiversity‐buffers‐disease
McMahon, B. J., Wall, P. G., Fanning, S., & Fahey, A. G. (2015). Targets to paradigm. Parasitology, 139, 847–863. https://doi.org/10.1017/
increase food production: One Health implications. Infection Ecology S0031182012000200
& Epidemiology, 5, 27708. https://doi.org/10.3402/iee.v5.27708 Rauter, C., & Hartung, T. (2005). Prevalence of Borrelia burgdorferi sensu
Milne, A. (1949). The ecology of the sheep tick, Ixodes ricinus L. Host lato genospecies in Ixodes ricinus ticks in Europe: A metaanalysis.
relationships of the tick: Part 2. Observations on hill and moorland Applied and Environmental Microbiology, 71, 7203–7216. https://doi.
grazings in northern England. Parasitology, 39, 173–197. https://doi. org/10.1128/AEM.71.11.7203-7216.2005
org/10.1017/S0031182000083736 Rizzoli, A., Silaghi, C., Obiegala, A., Rudolf, I., Hubálek, Z., Földvári, G.,
Morand, S., Owers, K. A., Waret‐Szkuta, A., McIntyre, K. M., & Baylis, … Kazimírová, M. (2014). Ixodes ricinus and its transmitted patho‐
M. (2013). Climate variability and outbreaks of infectious diseases gens in urban and peri‐urban areas in Europe: New hazards and rel‐
in Europe. Scientific Reports, 3, 1774. https://doi.org/10.1038/ evance for public health. Frontiers in Public Health, 2, 251. https://doi.
srep01774 org/10.3389/fpubh.2014.00251
Muehlenbein, M. P. (2016). Disease and human/animal Interactions. Saint Girons, I., Gern, L., Gray, J. S., Guy, E. C., Korenberg, E., Nuttall,
Annual Review of Anthropology, 45, 395–416. https://doi.org/10.1146/ P. A., … Postic, D. (1998). Identification of Borrelia burgdorferi sensu
annurev-anthro-102215-10003 lato species in Europe. Zentralblatt Für Bakteriologie, 287, 190–195.
Myers, S. S., Gaffikin, L., Golden, C. D., Ostfeld, R. S., Redford, K. H., https://doi.org/10.1016/S0934-8840(98)80120-5
Ricketts, T. H., … Osofsky, S. A. (2013). Human health impacts of Salkeld, D. J., Padgett, K. A., Jones, J. H., & Antolin, M. F. (2015). Public
ecosystem alteration. Proceedings of the National Academy of Sciences health perspective on patterns of biodiversity and zoonotic disease.
of the United States of America, 110, 18753–18760. https://doi. Proceedings of the National Academy of Sciences of the United States of
org/10.1073/pnas.1218656110 America, 112, E6261. https://doi.org/10.1073/pnas.1517640112
Nally, J. E., Arent, Z., Bayles, D. O., Hornsby, R. L., Gilmore, C., Regan, S., Salkeld, D. J., Padgett, K. A., Jones, J. H., & Lafferty, K. (2013). A meta‐
… McMahon, B. J. (2016). Emerging infectious disease implications analysis suggesting that the relationship between biodiversity and
of invasive mammalian species: The Greater White‐Toothed Shrew risk of zoonotic pathogen transmission is idiosyncratic. Ecology
(Crocidura russula) is associated with a novel serovar of pathogenic Letters, 16, 679–686. https://doi.org/10.1111/ele.12101
Leptospira in Ireland. PLoS Neglected Tropical Diseases, 10, https://doi. Samia, N. I., Kausrud, K. L., Heesterbeek, H., Ageyev, V., Begon, M., Chan,
org/10.1371/journal.pntd.0005174 K. S., & Stenseth, N. C. (2011). Dynamics of the plague‐wildlife‐
Newell, D. G., Koopmans, M., Verhoef, L., Duizer, E., Aidara‐Kane, A., human system in Central Asia are controlled by two epidemiologi‐
Sprong, H., … Kruse, H. (2010). Food‐borne diseases — The chal‐ cal thresholds. Proceedings of the National Academy of Sciences of the
lenges of 20 years ago still persist while new ones continue to United States of America, 108, 14527–14532. https://doi.org/10.1073/
emerge. International Journal of Food Microbiology, 139, S3–S15. pnas.1015946108
https://doi.org/10.1016/j.ijfoodmicro.2010.01.021 Sargeant, J. M., Amezcua, M. R., Rajic, A., & Waddell, L. (2007).
Ogden, N. H., & Tsao, J. I. (2009). Biodiversity and Lyme disease: Dilution Pre‐harvest interventions to reduce the shedding of E. coli
or amplification? Epidemics, 1, 196–206. https://doi.org/10.1016/j. O157 in the faeces of weaned domestic ruminants: A system‐
epidem.2009.06.002 atic review. Zoonoses and Public Health, 54, 260–277. https://doi.
Ostfeld, R. S., & Keesing, F. (2012). Effects of host diversity on infec‐ org/10.1111/j.1863-2378.2007.01059.x
tious disease. Annual Review of Ecology, Evolution, and Systematics, 43, Schmid, B. V., Büntgen, U., Easterday, W. R., Ginzler, C., Walløe, L.,
157–182. https://doi.org/10.1146/annurev-ecolsys-102710-145022 Bramanti, B., & Stenseth, N. C. (2015). Climate‐driven introduc‐
Otero‐Abad, B., & Torgerson, P. R. (2013). A systematic review of the tion of the Black Death and successive plague reintroductions
epidemiology of echinococcosis in domestic and wild animals. PLOS into Europe. Proceedings of the National Academy of Sciences of the
McMAHON et al. |
      765

United States of America, 112, 3020–3025. https://doi.org/10.1073/ prevalence. PLoS ONE, 4, e5461. https://doi.org/10.1371/journal.
pnas.1412887112 pone.0005461
Schneider, R., Aspöck, H., & Auer, H. (2013). Unexpected Increase of al‐ Theuerkauf, J., Perez, J., Taugamoa, A., Niutoua, I., Labrousse, D., Gula,
veolar echinococcosis, Austria, 2011. Emerging Infectious Diseases, 19, R., … Goarant, C. (2013). Leptospirosis risk increases with changes
475–477. https://doi.org/10.3201/eid1903.120595 in species composition of rat populations. Die Naturwissenschaften,
Schwabe, C. W. (1984). Veterinary medicine and human health, 3rd ed. 100, 385–388. https://doi.org/10.1007/s00114-013-1033-6
Baltimore, MD: Williams & Wilkins. Wilson, N., Lush, D., & Baker, M. G. (2010) Meteorological and climate
Schwarz, A. C., Ranft, U., Piechotowski, I., Childs, J. E., & Brockmann, change themes at the 2010 International Conference on Emerging
S. O. (2009). Risk factors for human infection with Puumala virus, Infectious Diseases. Eurosurveillance, 15, pii: 19627.
southwestern Germany. Emerging Infectious Diseases, 15, 1032–1039. Wood, C. L., Lafferty, K. D., DeLeo, G., Young, H. S., Hudson, P. J., &
https://doi.org/10.3201/eid1507.081413 Kuris, A. M. (2016). Does biodiversity protect humans against
Schweiger, A., Ammann, R. W., Candinas, D., Clavien, P.‐A., Eckert, J., infectious disease? Reply. Ecology, 97, 543–546. https://doi.
Gottstein, B., … Deplazes, P. (2007). Human alveolar echinococco‐ org/10.1890/15-1503.1
sis after fox population increase, Switzerland. Emerging Infectious Woolhouse, M. E. J., & Gowtage‐Sequeria, S. (2005). Host range and
Diseases, 13, 878–882. https://doi.org/10.3201/eid1306.061074 emerging and reemerging pathogens. Emerging Infectious Diseases,
Sharp, P. M., & Hahn, B. H. (2011). Origins of HIV and the AIDS pandemic. 11, 1842–1847. https://doi.org/10.3201/eid1112.050997
Cold Spring Harbor Perspectives in Medicine, 1, a006841. https://doi. Xu, G., Walker, D. H., Jupiter, D., Melby, P. C., & Arcari, C. M. (2017). A
org/10.1101/cshperspect.a006841 review of the global epidemiology of scrub typhus. PLoS Neglected
Steinmann, P., Keiser, J., Bos, R., Tanner, M., & Utzinger, J. (2006). Tropical Diseases, 11, e0006062. https://doi.org/10.1371/journal.
Schistosomiasis and water resources development: Systematic pntd.0006062
review, meta‐analysis, and estimates of people at risk. The Yates, T. L., Mills, J. N., Parmenter, C. A., Ksiazek, T. G., Parmenter, R. R.,
Lancet Infectious Diseases, 6, 411–425. https://doi.org/10.1016/ & Vandecastle, J. R. (2002). The ecology and evolutionary history of
S1473-3099(06)70521-7 an emergent disease: Hantavirus pulmonary syndrome. BioScience,
Stenseth, N. C., Samia, N. I., Viljugrein, H., Kausrud, K. L., Begon, M., 52, 989–998. https://doi.org/10.1641/0006-3568(2002)052[098
Davis, S., … Chan, K. S. (2006). Plague dynamics are driven by cli‐ 9:TEAEHO]2.0.CO;2
mate variation. Proceedings of the National Academy of Sciences of the Young, H. S., Wood, C. L., Kilpatrick, A. M., Lafferty, K. D., Nunn, C. L.,
United States of America, 103, 13110–13115. https://doi.org/10.1073/ & Vincent, J. R. (2017). Conservation, biodiversity and infectious
pnas.0602447103 disease: Scientific evidence and policy implications. Philosophical
Šumilo, D., Bormane, A., Asokliene, L., Vasilenko, V., Golovljova, I., Transactions of the Royal Society B: Biological Sciences, 372, 20160124.
Avsic‐Zupanc, T., … Randolph, S. E. (2008). Socio‐economic factors https://doi.org/10.1098/rstb.2016.0124
in the differential upsurge of tick‐borne encephalitis in Central and
Eastern Europe. Reviews in Medical Virology, 18, 81–95. https://doi.
org/10.1002/rmv.566
How to cite this article: McMahon BJ, Morand S, Gray JS.
Suter, G. W. (1993). A critique of ecosystem health concepts and indexes.
Environmental Toxicology and Chemistry, 12, 1533–1539. https://doi.
Ecosystem change and zoonoses in the Anthropocene.
org/10.1002/etc.5620120903 Zoonoses Public Health. 2018;65:755–765. https://doi.
Suzán, G., Marcé, E., Giermakowski, J. T., Mills, J. N., Ceballos, org/10.1111/zph.12489
G., Ostfeld, R. S., …Yates, T. L. (2009). Experimental evidence
for reduced rodent diversity causing increased hantavirus

You might also like