Download - The University of Liverpool Repository

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Hospital-acquired infection wikipedia , lookup

Leptospirosis wikipedia , lookup

Henipavirus wikipedia , lookup

Trichinosis wikipedia , lookup

Hepatitis B wikipedia , lookup

Sexually transmitted infection wikipedia , lookup

Rocky Mountain spotted fever wikipedia , lookup

Chagas disease wikipedia , lookup

Eradication of infectious diseases wikipedia , lookup

Visceral leishmaniasis wikipedia , lookup

Plasmodium falciparum wikipedia , lookup

Leishmaniasis wikipedia , lookup

Pandemic wikipedia , lookup

Schistosomiasis wikipedia , lookup

Onchocerciasis wikipedia , lookup

Oesophagostomum wikipedia , lookup

Cross-species transmission wikipedia , lookup

African trypanosomiasis wikipedia , lookup

Schistosoma mansoni wikipedia , lookup

Babesia wikipedia , lookup

Sarcocystis wikipedia , lookup

Parasitism wikipedia , lookup

Transcript
1
1259504
1
2
Title: Why infectious disease research needs community ecology
3
4
5
Authors: Pieter T. J. Johnson1*, Jacobus C. de Roode2, and Andy Fenton3
6
7
Affiliations:
8
1
Ecology and Evolutionary Biology, University of Colorado, Boulder, CO 80309 USA
9
2
Department of Biology, Emory University, Atlanta, GA 30322 USA
10
3
Institute of Integrative Biology, University of Liverpool, Liverpool, L69 7ZB, UK
11
12
* Corresponding author. E-mail: [email protected]
2
13
Abstract
14
Infectious diseases often emerge from interactions among multiple species and across multiple
15
levels of biological organization. Threats ranging from Ebola virus to human malaria to bat
16
white-nose syndrome illustrate the need for a mechanistic understanding of the ecological
17
interactions underlying emerging infections. We describe how recent advances in community
18
ecology can be adopted to address contemporary challenges in disease research. These analytical
19
tools can identify the factors governing complex assemblages of multiple hosts, parasites, and
20
vectors, and reveal how processes link across scales from individual hosts to regions. They can
21
also determine the drivers of heterogeneities among individuals, species, and regions to aid
22
targeting of control strategies. We provide examples where these principles have enhanced
23
disease management and illustrate how they can be further extended.
3
24
Main text:
25
Despite notable successes (1, 2), infectious diseases remain a leading source of human morbidity
26
and mortality (3) and continue to threaten wildlife conservation and food production (4-6). A
27
common factor underlying emerging diseases is the involvement of multiple host, vector, or
28
parasite species in complex ecological communities. Nearly 70% of emerging human infectious
29
diseases have wildlife hosts or vectors (7, 8), while several human parasites have spilled over to
30
cause morbidity and mortality in wildlife, such as measles in mountain gorillas and tuberculosis
31
in Asian elephants (9). The use of multiple hosts by parasites complicates control efforts that
32
target particular hosts for management; for example, Schistosoma japonicum, the primary cause
33
of human schistosomiasis in Asia, can infect 120 different species of mammals (10). Similarly,
34
more than 20 species of triatomine bugs can transmit Trypanosoma cruzi, which causes Chagas
35
disease in South America, such that efforts to control the dominant vector species alone may be
36
inadequate to achieve elimination (11). Such threats continue to grow in importance as global
37
travel and human activities increase contact with novel sources of parasites, and aid their spread
38
across the globe (6).
39
Alongside the multi-host nature of many infections, interactions among multiple co-
40
infecting parasites can alter host pathology, parasite transmission, and virulence evolution (12-
41
14). Parasites that disrupt immune function, such as HIV, have facilitated the re-emergence of
42
drug-resistant forms of tuberculosis (15), while co-infection with parasitic worms (helminths)
43
such as hookworm, can exacerbate malaria (16). Interactions between several parasite species
44
have been similarly implicated in coral reef diseases, epidemics in plants, and marine mammal
45
die offs (17-19). Because many host-parasite interactions are intimately embedded within
46
communities of organisms, management efforts are sometimes thwarted by ‘ecological surprises’
4
47
(20). Recent examples include the unexpected amplification of Middle East Respiratory
48
Syndrome coronavirus in internationally traded camels, and increased contact between badgers
49
and cattle following implementation of badger culling, ultimately leading to increased rather than
50
decreased transmission of bovine tuberculosis in the UK (21-23). Managing the challenges of
51
emerging infectious diseases thus requires a clear understanding of the full ecological context of
52
infection and transmission.
53
Our ability to understand and control infectious diseases has much to gain from the
54
discipline of community ecology, which has developed a range of analytical tools for addressing
55
complexity, species interactions and multi-level scaling (Fig. 1). These tools can be adopted to
56
improve our understanding and management of infectious diseases, both by quantifying
57
environmental and biological factors governing the structure of complex communities of
58
multiple hosts, vectors, and parasites, and also by identifying the effect and source(s) of
59
heterogeneity among individual hosts, host species, and geographic locations. These tools further
60
offer insight into interactions and feedbacks across multiple scales of organization, from within
61
hosts to across regions. We examine how the application of tools and concepts from community
62
ecology can help public health efforts to manage infectious disease threats.
63
64
Community ecology as a framework to understand infectious diseases
65
Community ecology offers a mechanistic bridge between processes unfolding at the fine scale of
66
individuals and populations and the ecological and evolutionary drivers of species distributions
67
at coarser scales. While some principles from community ecology have been applied to various
68
host-parasite systems (e.g., 24, 25-27), the ‘community ecology of disease’ remains in its relative
69
infancy, with most studies focusing on interactions between a single host and parasite species,
5
70
and often at a single scale. Data availability and quality are increasing rapidly, partly through
71
advances in sequencing technology, underscoring both the need for and the opportunity to
72
implement new methods to study infection dynamics in complex natural systems.
73
Community ecology theory tells us that, in parallel to the processes underlying
74
population genetics theory (i.e., gene flow, selection, drift, and mutation), the diversity,
75
abundance, and composition of species within a community can be understood in terms of
76
dispersal, ecological selection, ecological drift, and speciation (28). Following dispersal from the
77
regional species pool, a species’ success within a habitat is filtered by both niche- and stochastic
78
processes (29, 30). Within this framework, what needs to be understood is the degree to which
79
community structure is built predictably from niche-based effects associated with interactions
80
among species and the environment, or whether it arises through stochastic processes, such as
81
historical legacy, demographic stochasticity, and environmental fluctuations (Fig. 2).
82
Within their niche, parasites are affected by host condition, immune responses, the
83
abiotic environment and by interactions with co-infecting symbionts or associated free-living
84
organisms. If the assembly of parasite communities is predominantly deterministic, then the
85
richness and composition of parasite species will be predictable and based on characteristics of
86
the host and the environment. However, stochastic events and dispersal will also influence
87
parasite colonization-extinction dynamics. In some systems, for instance, the outcome of parasite
88
interactions depends strongly on the order of arrival within the host (14, 31). For example, long-
89
term sampling of wild field vole (Microtus agrestis) populations revealed that infection with the
90
protozoan Babesia microti reduced the probability that a host subsequently became infected with
91
the bacteria Bartonella spp. bacteria; however, if Bartonella established first, then B. microti was
92
75% less likely to invade (14). Similarly, high propagule dispersal by parasites can overcome
6
93
niche effects related to host susceptibility (32). For instance, although humans are dead-end hosts
94
with no onward transmission for many zoonotic infections, high exposure to such parasites can
95
have serious consequences for public health, such as West Nile encephalitis and late-stage Lyme
96
disease. Quantifying the relative contributions of niche-based and dispersal-based processes in
97
determining parasite community structure and individual infection risk offers an ecological
98
foundation for guiding resource investment into either defensive strategies, which focus on
99
altering niches to inhibit parasite establishment, or offensive strategies, which focus on limiting
100
dispersal (Fig. 3).
101
102
Approaches for understanding multi-level infection processes
103
Parasite metacommunities and assembly theory
104
Metacommunity theory provides a valuable toolkit for understanding the relative importance of
105
niche-based effects and dispersal-based effects in regulating the structure of parasite
106
communities (24, 33). By recognizing that landscapes support a series of ecological communities
107
connected through dispersal, metacommunity theory links interactions across local and regional
108
scales (32). For parasites, this framework can be applied to communities of parasites dispersing
109
among host individuals or across disjunct landscapes. Although rarely applied to parasite
110
communities, metacommunity-based approaches offer the potential to explore the interactive
111
roles of evolutionary history, dispersal limitation, host community composition and the abiotic
112
environment in driving parasite distributions (34) (Fig. 2). In a long-term study of 65 parasite
113
species from 15 species of desert rodents, for instance, Dallas and Presley (35) found that
114
parasite community structure was driven by niche effects associated with the ‘patch quality’ of
115
host species, including host traits such as body size, longevity, and abundance, rather than by
7
116
characteristics related to dispersal opportunities, such as host diet breadth or home range size, or
117
evolutionary history. In a study of plant parasites, Parker et al. (36) recently showed that
118
spillover risk in field experiments could be predicted by knowing the abundance of the host and
119
its phylogenetic relationships with other hosts in the community. In contrast to free-living
120
communities, parasite metacommunities do incur some unique analytical challenges, including
121
the potential for infections to sicken or kill individual hosts and thereby alter the availability of
122
habitat ‘patches’ for dispersal (26). Likewise, the co-assembly of host and parasite communities
123
needs to be examined concurrently (37), and an extra nested scale (i.e., for the within-host
124
dynamics) often needs to be included in analyses (Fig. 1).
125
Tools from network theory can be additionally valuable for understanding how
126
interactions between entire host and parasite communities vary over space and time (38). For
127
instance, Griffiths et al. (39) used network approaches to show that co-infecting parasites of
128
humans were organised into dense clusters around distinct locations in the body (e.g., organs),
129
and tended to interact with each other via shared resources within the host, rather than via the
130
immune system. Similar approaches have been applied across other scales of organization, for
131
example to define contact pathways for transmission among individual hosts (40, 41) and
132
identifying the role parasites play in structuring ecological food webs (42). While the focus of
133
network approaches thus far has often been on the patterns of links among species, emerging
134
tools allow for more explicit examination of interaction strengths which will help to forecast
135
dynamic changes in the system (43).
136
137
Infection heterogeneity and traits-based approaches
8
138
Community ecology emphasizes the importance of understanding individual and species-level
139
functional traits, thereby offering greater mechanistic and predictive power relative to simple
140
taxonomic classifications (44, 45). Although predicting the specific identities of species within
141
an assemblage is made difficult by stochastic factors, such as historical legacy, the composition
142
and frequency of functional traits may be more deterministic (46). Thus, while hosts and
143
parasites are typically defined in taxonomic terms, it may be more useful to classify them in
144
terms of functional traits that influence performance, such as transmission mode, site of
145
infection, or resource use for parasites, and body size, dispersal ability, or immune competence
146
for hosts. For instance, Han et al. (47) identified ‘trait profiles’ of known reservoir species and
147
used these to forecast candidate rodents likely to act as reservoirs for future zoonotic infections.
148
Their analysis emphasized the importance of ‘fast-paced’ species that reproduce early and often,
149
whereas taxonomic labels did a relatively poor job of classifying reservoir host status.
150
Trait-based analyses align with the long-standing recognition in disease ecology of the
151
disproportionate influence of superspreader individuals, amplification or reservoir host species,
152
or ‘hot-spot’ locations in driving transmission (22, 48, 49). Superspreading events have been
153
recorded for both wildlife and human diseases, including typhoid fever, HIV1, SARS and
154
tuberculosis (22, 50, 51), and can sometimes be linked to measurable variation in traits, such as
155
host immunity, behavior, age, diet and sex (52-54). For example, Perkins et al. (53) found that
156
large-bodied, sexually active male mice contribute 93% of potential transmission events for tick-
157
borne encephalitis virus, despite representing only ~20% of the host population. Methods to
158
partition the contributions of particular hosts, species, or locations to parasite transmission are
159
beginning to be developed (48, 55). For example, Rudge et al. (10) quantified host species
160
contributions to the number of cases generated (R0) of Schistosoma japonicum in China, for
9
161
which more than 120 host species have been identified. They showed that while bovids maintain
162
infection in marshlands, rodents are the main source of transmission in hilly areas, suggesting
163
that different control strategies are needed in the two habitats. The key challenge for
164
management is to identify how much of this heterogeneity is linked to measurable traits, and is
165
therefore predictable (niche-based), or whether it arises stochastically through unpredictable
166
temporal or spatial heterogeneity in exposure (56).
167
168
Moving across scales
169
A core principle of community ecology is the importance of scale in affecting the strength and
170
form of species interactions not only with each other but also with the environment (57)(Fig. 1).
171
Research in disease ecology often falls into one of three distinct levels: (1) within-host, which is
172
concerned with interactions with the host immune system and other parasites (13, 58); (2)
173
between-host, which is focused on parasite spread through host populations (59, 60) or, less
174
often, through host communities; or (3) on regional or biogeographical scales, which uses
175
comparative methods from macroecology to explore the drivers of parasite distributions and
176
diversity (61).
177
While studies focused on one scale often ignore, or treat as phenomenological black
178
boxes, the dynamics occurring at higher and lower scales, it appears that dynamic interactions
179
occur in both directions (41, 57). For instance, interactions among co-infecting parasites within
180
hosts can cause individual variation in susceptibility, infectiousness, behavior and survival (14,
181
62, 63), potentially with counter-intuitive consequences for transmission at the population level
182
(64). For example, African buffalo co-infected with gastrointestinal nematodes and bovine
183
tuberculosis (bTB) exhibit increased mortality, such that treating animals to reduce their worm
10
184
burdens improves individual survival but, by enabling infected hosts to live longer, is predicted
185
to increase population-level spread of bTB (63). Reciprocally, variation in host community
186
composition within a region can affect infection risk and spread at the individual and population
187
levels (10, 55). For vector-borne infections such as Lyme disease, wildlife species vary
188
considerably in their tendency to amplify the bacterium responsible and transmit it to suitable
189
tick vectors, such that regional variation in host species diversity is hypothesized to be a major
190
determinant of local infection risk for humans (65)(Box 1). However, such cross-scale processes
191
are hard to infer from observational data alone, and experimental perturbations are often needed
192
to definitively assess how processes at one scale affect those at another. In parallel with the rich
193
legacy of system manipulations from community ecology (66), disease ecologists have
194
increasingly used experimental approaches of natural systems, for example through anti-parasite
195
drug treatments (67), hormone manipulation (68), nutrient supplementation (69), and diversity
196
manipulations (70, 71). While often focused on single-host-single-parasite systems thus far,
197
implementing such experiments in more complex natural communities and at larger scales is
198
increasingly important for testing hypotheses about parasite transmission, impact and control.
199
200
How community ecology can help manage infectious diseases
201
We suggest that disease control strategies would benefit by incorporating community ecology
202
theory and approaches to explicitly account for the joint influences of dispersal and
203
environmental filters. Specifically, the “offensive” versus “defensive” concept developed for
204
invasive species can be applied to disease management (see also 72). Offensive strategies
205
allocate resources to limit the dispersal of an invader from established sites, whereas defensive
206
efforts reduce the vulnerability of uninvaded habitats to colonization (73). While this concept
11
207
parallels existing epidemiological emphasis on prevention versus control, its successful
208
application requires deeper insights into whether a parasite community is dispersal-limited,
209
niche-based, or is random in its assembly (Fig. 2). This approach can be used to strengthen
210
current methods of infectious disease management across the gamut of multi-host parasites,
211
multi-symbiont communities, and infection heterogeneities across scales (Fig. 3).
212
213
Managing multi-host parasites
214
A current pressing question is how ongoing changes in biodiversity will affect the spread and
215
severity of infectious diseases (74). When diverse communities also support species that interfere
216
with transmission, such as the presence of low susceptibility hosts, predators, or symbionts,
217
community structure can be manipulated defensively to manage infections by limiting niche
218
suitability (37). For example, zooprophylaxis (in which livestock are used as bait to divert blood-
219
feeding arthropod vectors away from people) has been proposed as a control strategy for vector-
220
borne diseases for over a century, but has had limited success in some settings because increased
221
livestock density can also increase vector abundance. However, recent models on malaria and
222
zoonotic cutaneous leishmaniasis indicate that carefully chosen livestock densities coupled with
223
insecticide treatment can effectively reduce parasite transmission to humans (75, 76). Similar
224
approaches, such as intercropping and crop rotation, have been used successfully to reduce plant
225
pests and parasites in agricultural systems (77). While evidence for such dilution effects
226
continues to grow (78), the degree to which biodiversity will regulate infection by a particular
227
parasite depends on the degree to which host assembly is deterministic, whether the parasite is
228
niche- or dispersal-limited, and how increases in richness affect host and vector abundance.
12
229
Managing host communities is also crucial to mitigating the risk of spillover events from
230
animal reservoirs to humans. To minimize the risks of spillover, there are several potential
231
offensive and defensive approaches, the choice of which will depend on the specific biology of
232
the hosts and vectors involved. The first option is to reduce infection in reservoir hosts. For
233
instance, vaccine baits have successfully eliminated rabies from several European countries
234
through their protective effects on non-human hosts (79). The second approach is to limit
235
contacts between wildlife and humans, for example by reducing bushmeat consumption and its
236
potential to introduce novel infections (80). In West Africa, for example, increasing the use of
237
alternative protein sources such as marine fish could relieve pressure on the bushmeat trade (81).
238
Such approaches require tight coordination between many parties, including medical scientists,
239
anthropologists and governments. Similarly, the use of transmission barriers can help to limit
240
contact between wildlife reservoirs and domestic animals (82, 83). The third approach is to
241
reduce the probability of infection when contact is unavoidable or unpredictable. Ongoing yet
242
unpredictable spillovers of dengue viruses from non-human primates, for instance, complicate
243
the control of human disease in Southeast Asia and Africa. One approach to control such
244
infections is through the implementation of cross-reactive vaccines, which are currently under
245
development (84). When vaccines are not available yet, as is the case for Ebola virus, reducing
246
human-human transmission through contact tracing and subsequent quarantine and treatment can
247
help to limit epidemic spread (85).
248
249
Managing symbiont communities
250
Interactions among co-infecting parasites or symbionts can also be used as niche-based
251
management tools. For example, treating patients suffering from lymphatic filariasis with the
13
252
antibiotic doxycycline eliminates essential symbiotic bacteria required by filarial worms,
253
ultimately leading to worm sterility and death (86). Restoration or augmentation of the microbial
254
community within the host can also provide protection against parasite invasion. For example,
255
transferring human-microbial communities by fecal transplants often leads to clinical resolution
256
of intestinal pathology associated with Clostridium difficile infection (87). Finally, interactions
257
among co-infecting parasites, parasite strains, or other symbionts can be manipulated to reduce
258
the spread of disease-causing organisms. Long-lived parasites, such as helminths, may
259
exacerbate disease caused by co-infecting parasites, leading to calls to incorporate de-worming
260
to improve management of HIV, malaria and TB (16, 88). In other cases, antagonistic
261
interactions between parasites or other symbionts may be used to benefit the host. For instance,
262
trials are under development to reduce the vector competence of mosquitoes by infecting them
263
with the bacterium Wolbachia, which inhibits Dengue virus and filarial worm survival and
264
transmission through a combination of immune activation, competition for cellular components,
265
and shortened mosquito lifespan (89, 90). These examples emphasize the importance of
266
understanding and predicting the outcome of multiple infections, for which community ecology
267
approaches focused on parasite traits and resource use have already offered added insights (91).
268
269
Heterogeneity and scale
270
The disproportionate roles of particular locations, particular host species, and particular
271
individual hosts in driving epidemics or epizootics raises the tantalizing promise of highly
272
efficient targeted control and treatment (22, 48, 50). In the Serengeti, for example, where rabies
273
can infect up to 12 carnivore species, domestic dogs are responsible for more than 70 percent of
274
transmission events to humans (92). Annual vaccination of 60 percent of dogs is projected to
14
275
control the virus, a target that is logistically and economically feasible (93). During the recent
276
Ebola epidemic in West Africa, close contact between deceased patients and family or friends
277
during traditional burials functioned as superspreading events (94), and implementation of
278
‘sanitary burials’ that reduced such contacts helped curb the epidemic. Thus, targeting
279
superspreading hosts or events is feasible when transmission heterogeneities are deterministic
280
and can be linked to measurable traits or characteristics.
281
Ultimately, the efficacy of offensive and defensive approaches will depend on whether
282
the scale of application is local or regional, the transmission and dispersal characteristics of the
283
parasite involved, and the point in the epidemic when the intervention is initiated and how many
284
individuals have already been infected (72). Defensive, niche-based management strategies,
285
ranging from vaccination and prophylaxis to ecological competition by probiotic symbionts, are
286
more likely to be effective when parasite dispersal is high, for parasites with high or
287
unpredictable propagule pressure, and for epidemics already underway (Fig. 3). In contrast,
288
offensive strategies that focus on reducing dispersal are more likely to succeed at community and
289
regional scales than at individual and population scales, because parasite dispersal between
290
individuals within a host population is often harder to control than dispersal between sites. For
291
instance, established populations of the Asian tiger mosquito, Aedes albopictus, recently linked
292
to a large outbreak of the viral disease chikungunya on the Indian Ocean Island La Réunion (95),
293
are almost impossible to eliminate; however, because most introductions of this vector have
294
occurred through the shipment of used tires, focused efforts to limit this trade offer the best
295
potential for containing future spread of the vector (95).
296
297
Outlook
15
298
While the disciplines of epidemiology and community ecology have developed largely
299
independently from one another, the multi-species nature of many contemporary disease threats
300
demands a community-scale approach to complement more traditional biomedical treatments.
301
Community ecology offers a theoretical framework and the analytical tools to move beyond the
302
historical emphasis on particular host-parasite interactions and consider the full suite of species
303
that influence infection dynamics. We have emphasized approaches from community ecology
304
that can advance our ability to manage infections by: (1) identifying the factors that govern the
305
structure and dynamics of communities composed of multiple hosts, vectors, and symbionts, (2)
306
identifying the drivers of heterogeneity, and (3) understanding how processes and patterns link
307
across multiple scales of biological organization. For many emerging infections, complete
308
eradication is unlikely to be successful, but a broader understanding of the ecological
309
communities in which host-parasite interactions are embedded will facilitate more effective
310
management.
311
Transforming this broader understanding into practical disease management requires tight
312
integration of surveillance, community ecology analysis, and public health implementation (96).
313
Ongoing technological advances are rapidly overcoming previous barriers in data quality and
314
quantity, highlighting emerging opportunities to incorporate approaches from community
315
ecology into existing disease research and evaluate the factors driving the structure and dynamics
316
of natural disease systems. Combining analyses of these high-resolution data with modeling
317
approaches and large-scale manipulations of host-parasite interactions – similar to the
318
foundational experiments from community ecology (66) – offer among the best opportunities for
319
developing a deeper understanding of the processes underlying disease emergence and control.
320
To date there have been some practical successes that follow this broad approach. For example,
16
321
following the observation of five dead howler monkeys - a key host for yellow fever virus - a
322
collaborative effort between the USAID PREDICT program and the Bolivian government led to
323
rapid implementation of human vaccination and mosquito control in the affected area (97).
324
Similarly, increased use of buffer zones between fruit trees and livestock housing has been
325
effectively used in Malaysia to reduce Nipah virus transmission into pigs and the risks of human
326
outbreaks (82), while electrified fences in Kruger National Park have helped limit contact
327
between bovine TB-infected wildlife and cattle in surrounding areas (83). Such scenarios
328
demonstrate how a broader appreciation for the epidemiological links among humans, domestic
329
animals, and wildlife can result in more effective control of disease risk in ecological
330
communities.
331
332
References and Notes
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
1.
2.
3.
4.
5.
6.
7.
8.
B. Jasny, L. Roberts, M. Enserink, O. Smith, What works. Science 345, 1256-1257
(2014).
J. Sepúlveda, C. Murray, The state of global health in 2014. Science 345, 1275-1278
(2014).
World Health Organization, The top 10 causes of death (2014
http://www.who.int/mediacentre/factsheets/fs310/en/ - ).
M. C. Fisher, D. A. Henk, C. J. Briggs, J. S. Brownstein, L. C. Madoff, S. L. McCraw, S.
J. Gurr, Emerging fungal threats to animal, plant and ecosystem health. Nature 484, 186194 (2012).
B. V. Purse, P. S. Mellor, D. J. Rogers, A. R. Samuel, P. P. C. Mertens, M. Baylis,
Climate change and the recent emergence of bluetongue in Europe. Nat. Rev. Microbiol.
3, 171-181 (2005).
K. F. Smith, M. Goldberg, S. Rosenthal, L. Carlson, J. Chen, C. Chen, S. Ramachandran,
Global rise in human infectious disease outbreaks. J. R. Soc. Interface 11, 20140950
(2014).
K. E. Jones, N. G. Patel, M. A. Levy, A. Storeygard, D. Balk, J. L. Gittleman, P. Daszak,
Global trends in emerging infectious diseases. Nature 451, 990-U994 (2008).
J. O. Lloyd-Smith, D. George, K. M. Pepin, V. E. Pitzer, J. R. C. Pulliam, A. P. Dobson,
P. J. Hudson, B. T. Grenfell, Epidemic dynamics at the human-animal interface. Science
326, 1362-1367 (2009).
17
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
A. M. Messenger, A. N. Barnes, G. C. Gray, Reverse zoonotic disease transmission
(zooanthroponosis): a systematic review of seldom- documented human biological threats
to animals. PLoS ONE 9, e89055 (2014).
J. W. Rudge, J. P. Webster, D.-B. Lu, T.-P. Wang, G.-R. Fang, M.-G. Basanez,
Identifying host species driving transmission of schistosomiasis japonica, a multihost
parasite system, in China. Proc. Natl. Acad. Sci. U. S. A. 110, 11457-11462 (2013).
J. E. Rabinovich, U. D. Kitron, Y. Obed, M. Yoshioka, N. Gottdenker, L. F. Chaves,
Ecological patterns of blood-feeding by kissing-bugs (Hemiptera: Reduviidae:
Triatominae). Mem. Inst. Oswaldo Cruz 106, 479-494 (2011).
S. Alizon, J. C. de Roode, Y. Michalakis, Multiple infections and the evolution of
virulence. Ecol. Lett. 16, 556-567 (2013).
A. B. Pedersen, A. Fenton, Emphasizing the ecology in parasite community ecology.
Trends Ecol. Evol. 22, 133-139 (2007).
S. Telfer, X. Lambin, R. Birtles, P. Beldomenico, S. Burthe, S. Paterson, M. Begon,
Species interactions in a parasite community drive infection risk in a wildlife population.
Science 330, 243-246 (2010).
C. K. Kwan, J. D. Ernst, HIV and Tuberculosis: a deadly human syndemic. Clin.
Microbiol. Rev. 24, 351-376 (2011).
P. Druilhe, A. Tall, C. Sokhna, Worms can worsen malaria: towards a new means to roll
back malaria? Trends Parasitol. 21, 359-362 (2005).
A. K. Gibson, S. Raverty, D. M. Lambourn, J. Huggins, S. L. Magargal, M. E. Grigg,
Polyparasitism is associated with increased disease severity in Toxoplasma gondiiinfected marine sentinel species. PLoS Negl. Trop. Dis. 5, e1142 (2011).
J. D. Voss, D. K. Mills, J. L. Myers, E. R. Remily, L. L. Richardson, Black band disease
microbial community variation on corals in three regions of the wider Caribbean. Microb.
Ecol. 54, 730-739 (2007).
H. Susi, B. Barres, P. F. Vale, A. L. Laine, Co-infection alters population dynamics of
infectious disease. Nat. Commun. 6, (2015).
D. F. Doak, J. A. Estes, B. S. Halpern, U. Jacob, D. R. Lindberg, J. Lovvorn, D. H.
Monson, M. T. Tinker, T. M. Williams, J. T. Wootton, I. Carroll, M. Emmerson, F.
Micheli, M. Novak, Understanding and predicting ecological dynamics: Are major
surprises inevitable? Ecology 89, 952-961 (2008).
J. A. Al-Tawfiq, Z. A. Memish, Middle East respiratory syndrome coronavirus:
transmission and phylogenetic evolution. Trends Microbiol. 22, 573-579 (2014).
P. J. Hudson, S. E. Perkins, I. M. Cattadori, in Infectious disease ecology : effects of
ecosystems on disease and of disease on ecosystems, R. S. Ostfeld, Keesing, F., Eviner,
V. T., Ed. (Princeton University Press, Princeton, New Jersey, 2008), pp. 347-367.
R. Woodroffe, C. A. Donnelly, H. E. Jenkins, W. T. Johnston, D. R. Cox, F. J. Bourne, C.
L. Cheeseman, R. J. Delahay, R. S. Clifton-Hadley, G. Gettinby, P. Gilks, R. G.
Hewinson, J. P. McInerney, W. I. Morrison, Culling and cattle controls influence
tuberculosis risk for badgers. Proc. Natl. Acad. Sci. U. S. A. 103, 14713-14717 (2006).
J. R. Mihaljevic, Linking metacommunity theory and symbiont evolutionary ecology.
Trends Ecol. Evol. 27, 323-329 (2012).
E. C. Rynkiewicz, A. B. Pedersen, A. Fenton, An ecosystem approach to understanding
and managing within-host parasite community dynamics. Trends Parasitol. 31, 212-221
(2015).
18
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
E. W. Seabloom, E. T. Borer, K. Gross, A. E. Kendig, C. Lacroix, C. E. Mitchell, E. A.
Mordecai, A. G. Power, The community ecology of pathogens: coinfection, coexistence
and community composition. Ecol. Lett. 18, 401-415 (2015).
R. D. Holt, A. Dobson, in Disease Ecology: Community structure and pathogen
dynamics, S. K. Collinge, C. Ray, Eds. (Oxford University Press, Oxford, UK, 2006).
M. Vellend, Conceptual synthesis in community ecology. Q. Rev. Biol. 85, 183-206
(2010).
P. B. Adler, J. HilleRisLambers, J. M. Levine, A niche for neutrality. Ecol. Lett. 10, 95104 (2007).
J. M. Chase, Drought mediates the importance of stochastic community assembly. Proc.
Natl. Acad. Sci. U. S. A. 104, 17430-17434 (2007).
A. F. Read, L. H. Taylor, The ecology of genetically diverse infections. Science 292,
1099-1102 (2001).
M. A. Leibold, M. Holyoak, N. Mouquet, P. Amarasekare, J. M. Chase, M. F. Hoopes, R.
D. Holt, J. B. Shurin, R. Law, D. Tilman, M. Loreau, A. Gonzalez, The metacommunity
concept: a framework for multi-scale community ecology. Ecol. Lett. 7, 601-613 (2004).
S. J. Presley, C. L. Higgins, M. R. Willig, A comprehensive framework for the evaluation
of metacommunity structure. Oikos 119, 908-917 (2010).
G. Suzan, G. E. Garcia-Pena, I. Castro-Arellano, O. Rico, A. V. Rubio, M. J. Tolsa, B.
Roche, P. R. Hosseini, A. Rizzoli, K. A. Murray, C. Zambrana-Torrelio, M. Vittecoq, X.
Bailly, A. A. Aguirre, P. Daszak, A. H. Prieur-Richard, J. N. Mills, J. F. Guegan,
Metacommunity and phylogenetic structure determine wildlife and zoonotic infectious
disease patterns in time and space. Ecol. Evol. 5, 865-873 (2015).
T. Dallas, S. J. Presley, Relative importance of host environment, transmission potential
and host phylogeny to the structure of parasite metacommunities. Oikos 123, 866-874
(2014).
I. M. Parker, M. Saunders, M. Bontrager, A. P. Weitz, R. Hendricks, R. Magarey, K.
Suiter, G. S. Gilbert, Phylogenetic structure and host abundance drive disease pressure in
communities. Nature 520, 542-544 (2015).
P. T. J. Johnson, D. L. Preston, J. T. Hoverman, B. E. LaFonte, Host and parasite
diversity jointly control disease risk in complex communities. Proc. Natl. Acad. Sci. U. S.
A. 110, 16916-16921 (2013).
R. Poulin, Network analysis shining light on parasite ecology and diversity. Trends
Parasitol. 26, 492-498 (2010).
E. C. Griffiths, A. B. Pedersen, A. Fenton, O. L. Petchey, Analysis of a summary network
of co-infection in humans reveals that parasites interact most via shared resources. Proc.
R. Soc. Lond. B. 281, 20132286 (2014).
S. Davis, B. Abbasi, S. Shah, S. Telfer, M. Begon, Spatial analyses of wildlife contact
networks. J. R. Soc. Interface 12, 20141004 (2015).
D. M. Tompkins, A. M. Dunn, M. J. Smith, S. Telfer, Wildlife diseases: from individuals
to ecosystems. J. Anim. Ecol. 80, 19-38 (2011).
A. R. Cirtwill, D. B. Stouffer, Concomitant predation on parasites is highly variable but
constrains the ways in which parasites contribute to food web structure. J. Anim. Ecol.
84, 734-744 (2015).
T. Poisot, E. Canard, D. Mouillot, N. Mouquet, D. Gravel, The dissimilarity of species
interaction networks. Ecol. Lett. 15, 1353-1361 (2012).
19
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
B. J. McGill, B. J. Enquist, E. Weiher, M. Westoby, Rebuilding community ecology from
functional traits. Trends Ecol. Evol. 21, 178-185 (2006).
C. T. Webb, J. A. Hoeting, G. M. Ames, M. I. Pyne, N. L. Poff, A structured and
dynamic framework to advance traits-based theory and prediction in ecology. Ecol. Lett.
13, 267-283 (2010).
E. Weiher, D. Freund, T. Bunton, A. Stefanski, T. Lee, S. Bentivenga, Advances,
challenges and a developing synthesis of ecological community assembly theory. Philos.
T. R. Soc. B. 366, 2403-2413 (2011).
B. A. Han, J. P. Schmidt, S. E. Bowden, J. M. Drake, Rodent reservoirs of future
zoonotic diseases. Proc. Natl. Acad. Sci. U. S. A. 112, 7039-7044 (2015).
S. H. Paull, S. Song, K. M. McClure, L. C. Sackett, A. M. Kilpatrick, P. T. J. Johnson,
From superspreaders to disease hotspots: linking transmission across hosts and space.
Front. Ecol. Environ. 10, 75-82 (2012).
D. G. Streicker, A. Fenton, A. B. Pedersen, Differential sources of host species
heterogeneity influence the transmission and control of multihost parasites. Ecol. Lett. 16,
975-984 (2013).
J. O. Lloyd-Smith, S. J. Schreiber, P. E. Kopp, W. M. Getz, Superspreading and the
effect of individual variation on disease emergence. Nature 438, 355-359 (2005).
Z. Shen, F. Ning, W. G. Zhou, X. He, C. Y. Lin, D. P. Chin, Z. H. Zhu, A. Schuchat,
Superspreading SARS events, Beijing, 2003. Emerg. Infect. Dis. 10, 256-260 (2004).
A. L. Graham, A. D. Hayward, K. A. Watt, J. G. Pilkington, J. M. Pemberton, D. H.
Nussey, Fitness correlates of heritable variation in antibody responsiveness in a wild
mammal. Science 330, 662-665 (2010).
S. E. Perkins, I. M. Cattadori, V. Tagliapietra, A. P. Rizzoli, P. J. Hudson, Empirical
evidence for key hosts in persistence of a tick-borne disease. Int. J. Parasitol. 33, 909917 (2003).
W. E. Stutz, O. L. Lau, D. I. Bolnick, Contrasting patterns of phenotype-dependent
parasitism within and among populations of threespine stickleback. Am. Nat. 183, 810825 (2014).
D. G. Streicker, A. Fenton, A. B. Pedersen, Differential sources of host species
heterogeneity influence the transmission and control of multihost parasites. Ecol. Lett. 16,
975-984 (2013).
J. M. Calabrese, J. L. Brunner, R. S. Ostfeld, Partitioning the aggregation of parasites on
hosts into intrinsic and extrinsic components via an extended Poisson-gamma mixture
model. PLoS ONE 6, (2011).
P. Chesson, Scale transition theory: Its aims, motivations and predictions. Ecol. Complex.
10, 52-68 (2012).
A. Fenton, S. E. Perkins, Applying predator-prey theory to modelling immune-mediated,
within-host interspecific parasite interactions. Parasitology 137, 1027-1038 (2010).
R. M. Anderson, R. M. May, Regulation and stability of host-parasite population
interactions. I. Regulatory processes. J. Anim. Ecol. 47, 219-247 (1978).
R. M. Anderson, R. M. May, The population dynamics of microparasites and their
invertebrate hosts. Philos. Trans. R. Soc. Lond., Ser. B: Biol. Sci. 291, 451-524 (1981).
R. R. Dunn, T. J. Davies, N. C. Harris, M. C. Gavin, Global drivers of human pathogen
richness and prevalence. Proc. R. Soc. Lond. B. 277, 2587-2595 (2010).
20
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
75.
76.
77.
78.
S. C. L. Knowles, A. Fenton, O. L. Petchey, T. R. Jones, R. Barber, A. B. Pedersen,
Stability of within-host-parasite communities in a wild mammal system. Proc. R. Soc.
Lond. B. 280, 20130598 (2013).
V. O. Ezenwa, A. E. Jolles, Opposite effects of anthelmintic treatment on microbial
infection at individual versus population scales. Science 347, 175-177 (2015).
A. Fenton, Dances with worms: the ecological and evolutionary impacts of deworming
on coinfecting pathogens. Parasitology 140, 1119-1132 (2013).
R. S. Ostfeld, F. Keesing, Effects of host diversity on infectious disease. Annu. Rev. Ecol.
Evol. Syst. 43, 157-182 (2012).
D. W. Schindler, Replication versus realism: The need for ecosystem-scale experiments.
Ecosystems 1, 323-334 (1998).
A. B. Pedersen, A. Fenton, The role of antiparasite treatment experiments in assessing the
impact of parasites on wildlife. Trends Parasitol. 31, 200-211 (2015).
D. A. Grear, S. E. Perkins, P. J. Hudson, Does elevated testosterone result in increased
exposure and transmission of parasites? Ecol. Lett. 12, 528-537 (2009).
A. B. Pedersen, T. J. Greives, The interaction of parasites and resources cause crashes in
a wild mouse population. J. Anim. Ecol. 77, 370-377 (2008).
P. T. J. Johnson, D. L. Preston, J. T. Hoverman, K. L. D. Richgels, Biodiversity decreases
disease through predictable changes in host community competence. Nature 494, 230233 (2013).
G. Suzan, E. Marce, J. T. Giermakowski, J. N. Mills, G. Ceballos, R. S. Ostfeld, B.
Armien, J. M. Pascale, T. L. Yates, Experimental evidence for reduced rodent diversity
causing increased hantavirus prevalence. PLoS ONE 4, e5461 (2009).
K. E. Langwig, J. Voyles, M. Q. Wilber, W. F. Frick, K. A. Murray, B. M. Bolker, J. P.
Collins, T. L. Cheng, M. C. Fisher, J. R. Hoyt, D. L. Lindner, H. I. McCallum, R.
Puschendorf, E. B. Rosenblum, M. Toothman, C. K. R. Willis, C. J. Briggs, A. M.
Kilpatrick, Context-dependent conservation responses to emerging wildlife diseases.
Front. Ecol. Environ. 13, 195-202 (2015).
K. L. S. Drury, J. D. Rothlisberger, Offense and defense in landscape-level invasion
control. Oikos 117, 182-190 (2008).
F. Keesing, L. K. Belden, P. Daszak, A. Dobson, C. D. Harvell, R. D. Holt, P. Hudson,
A. Jolles, K. E. Jones, C. E. Mitchell, S. S. Myers, T. Bogich, R. S. Ostfeld, Impacts of
biodiversity on the emergence and transmission of infectious diseases. Nature 468, 647652 (2010).
A. O. Franco, M. G. M. Gomes, M. Rowland, P. G. Coleman, C. R. Davies, Controlling
malaria using livestock-based interventions: a One Health approach. PLoS ONE 9,
e101699 (2014).
B. Kaabi, S. B. Ahmed, Assessing the effect of zooprophylaxis on zoonotic cutaneous
leishmaniasis transmission: A system dynamics approach. BioSyst. 114, 253-260 (2013).
Y. Y. Zhu, H. R. Chen, J. H. Fan, Y. Y. Wang, Y. Li, J. B. Chen, J. X. Fan, S. S. Yang,
L. P. Hu, H. Leung, T. W. Mew, P. S. Teng, Z. H. Wang, C. C. Mundt, Genetic diversity
and disease control in rice. Nature 406, 718-722 (2000).
D. J. Civitello, J. Cohen, H. Fatima, N. T. Halstead, J. Liriano, T. A. McMahon, N.
Ortega, E. L. Sauer, T. Sehgal, S. Young, J. R. Rohr, Biodiversity inhibits natural
enemies: broad evidence for the dilution effect. Proc. Natl. Acad. Sci. USA, (2015).
21
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
90.
91.
92.
93.
P. Mähl, F. Cliquet, A. L. Guiot, E. Niin, E. Fournials, N. Saint-Jean, M. Aubert, C. E.
Rupprecht, S. Gueguen, Twenty year experience of the oral rabies vaccine SAG2 in
wildlife: a global review. Vet. Res. 45, 77 (2014).
N. D. Wolfe, W. Heneine, J. K. Carr, A. D. Garcia, V. Shanmugam, U. Tamoufe, J. N.
Torimiro, A. T. Prosser, M. LeBreton, E. Mpoudi-Ngole, F. E. McCutchan, D. L. Birx, T.
M. Folks, D. S. Burke, W. M. Switzer, Emergence of unique primate T-lymphotropic
viruses among central African bushmeat hunters. Proc. Natl. Acad. Sci. U. S. A. 102,
7994-7999 (2005).
J. S. Brashares, P. Arcese, M. K. Sam, P. B. Coppolillo, A. R. E. Sinclair, A. Balmford,
Bushmeat hunting, wildlife declines, and fish supply in West Africa. Science 306, 11801183 (2004).
P. Daszak, R. Plowright, J. Epstein, J. Pulliam, S. Abdul Rahman, H. Field, C. Smith, K.
Olival, S. Luby, K. Halpin, A. D. Hyatt, A. A. Cunningham, in Disease ecology:
community structure and pathogen dynamics, S. K. Collinge, C. Ray, Eds. (2006), pp.
186-201.
A. R. Renwick, P. C. L. White, R. G. Bengis, Bovine tuberculosis in southern African
wildlife: a multi-species host-pathogen system. Epidemiol. Infect. 135, 529-540 (2007).
A. P. Durbin, S. V. Mayer, S. L. Rossi, I. Y. Amaya-Larios, J. Ramos-Castaneda, E. E.
Ooi, M. J. Cardosa, J. L. Munoz-Jordan, R. B. Tesh, W. B. Messer, S. C. Weaver, N.
Vasilakis, Emergence potential of sylvatic dengue virus type 4 in the urban transmission
cycle is restrained by vaccination and homotypic immunity. Virology 439, 34-41 (2013).
J. M. Drake, R. B. Kaul, L. W. Alexander, S. M. O'Regan, A. M. Kramer, J. T. Pulliam,
M. J. Ferrari, A. W. Park, Ebola cases and health system demand in Liberia. PLoS Biol.
13, e1002056 (2015).
M. J. Taylor, A. Hoerauf, M. Bockarie, Lymphatic filariasis and onchocerciasis. Lancet
376, 1175-1185 (2010).
E. van Nood, A. Vrieze, M. Nieuwdorp, S. Fuentes, E. G. Zoetendal, W. M. de Vos, C. E.
Visser, E. J. Kuijper, J. F. Bartelsman, J. G. Tijssen, Duodenal infusion of donor feces for
recurrent Clostridium difficile. New Engl. J. Med. 368, 407-415 (2013).
G. Harms, H. Feldmeier, Review: HIV infection and tropical parasitic diseases deleterious interactions in both directions? Trop. Med. Int. Health 7, 479-488 (2002).
A. A. Hoffmann, B. L. Montgomery, J. Popovici, I. Iturbe-Ormaetxe, P. H. Johnson, F.
Muzzi, M. Greenfield, M. Durkan, Y. S. Leong, Y. Dong, H. Cook, J. Axford, A. G.
Callahan, N. Kenny, C. Omodei, E. A. McGraw, P. A. Ryan, S. A. Ritchie, M. Turelli, S.
L. O'Neill, Successful establishment of Wolbachia in Aedes populations to suppress
dengue transmission. Nature 476, 454-U107 (2011).
Z. Kambris, P. E. Cook, H. K. Phuc, S. P. Sinkins, Immune activation by life-shortening
Wolbachia and reduced filarial competence in mosquitoes. Science 326, 134-136 (2009).
A. L. Graham, Ecological rules governing helminth-microparasite coinfection. Proc.
Natl. Acad. Sci. U. S. A. 105, 566-570 (2008).
T. Lembo, K. Hampson, D. T. Haydon, M. Craft, A. Dobson, J. Dushoff, E. Ernest, R.
Hoare, M. Kaare, T. Mlengeya, C. Mentzel, S. Cleaveland, Exploring reservoir dynamics:
a case study of rabies in the Serengeti ecosystem. J. Appl. Ecol. 45, 1246-1257 (2008).
K. Hampson, J. Dushoff, S. Cleaveland, D. T. Haydon, M. Kaare, C. Packer, A. Dobson,
Transmission dynamics and prospects for the elimination of canine rabies. PLoS Biol. 7,
462-471 (2009).
22
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
94.
608
Acknowledgments
609
For discussions and feedback helpful in shaping the manuscript, we thank S. Altizer, D. Calhoun,
610
G. Devevey, I. Doron, S. Haas, K. Hoang, B. Hoye, M. Joseph, J. Koprivnikar, T. McDevitt-
611
Galles, J. Mihaljevic, A. Pedersen, O. Petchey, A. Pierce, D. Preston, Y. Springer, W. Stutz, L.
612
Tao and S. White. PTJJ was supported by funds from NSF (DEB-1149308) and NIH
613
(R01GM109499), JCdR was supported by funds from NSF (DEB-1257160) and NIH
614
(R01GM109501), and AF was supported by funding from the Natural Environment Research
615
Council (NERC) UK (NE/G006830/1 and NE/I024038/1).
616
95.
96.
97.
98.
99.
100.
101.
102.
103.
A. Pandey, K. E. Atkins, J. Medlock, N. Wenzel, J. P. Townsend, J. E. Childs, T. G.
Nyenswah, M. L. Ndeffo-Mbah, A. P. Galvani, Strategies for containing Ebola in West
Africa. Science 346, 991-995 (2014).
E. J. Scholte, F. Schaffner, in Emerging pests and vector-borne diseases in Europe, W.
Takken, B. G. J. Knols, Eds. (Wageningen Academic Publishers, Wageningen,
Netherlands, 2007).
S. S. Morse, J. A. K. Mazet, M. Woolhouse, C. R. Parrish, D. Carroll, W. B. Karesh, C.
Zambrana-Torrelio, W. I. Lipkin, P. Daszak, Prediction and prevention of the next
pandemic zoonosis. Lancet 380, 1956-1965 (2012).
PREDICT. (http://www.vetmed.ucdavis.edu/ohi/predict/news/bolivia-successhowler-monkeys.cfm, 2014), vol. 2014.
A. D. Blackwell, M. Martin, H. Kaplan, M. Gurven, Antagonism between two intestinal
parasites in humans: the importance of co-infection for infection risk and recovery
dynamics. Proc. R. Soc. Lond. B. 280, 20131671 (2013).
R. M. Anderson, R. M. May, Population biology of infectious diseases: Part I. Nature
280, 361-367 (1979).
A. Dobson, Population dynamics of pathogens with multiple host species. Am. Nat. 164,
S64-S78 (2004).
L. Gilbert, R. Norman, K. M. Laurenson, H. W. Reid, P. J. Hudson, Disease persistence
and apparent competition in a three-host community: an empirical and analytical study of
large-scale, wild populations. J. Anim. Ecol. 70, 1053-1061 (2001).
R. Norman, R. G. Bowers, M. Begon, P. J. Hudson, Persistence of tick-horne virus in the
presence of multiple host species: Tick reservoirs and parasite mediated competition. J.
Theor. Biol. 200, 111-118 (1999).
C. L. Wood, K. D. Lafferty, Biodiversity and disease: a synthesis of ecological
perspectives on Lyme disease transmission. Trends Ecol. Evol. 28, 239-247 (2013).
23
617
Boxes and Figures
618
Box 1. The role of simple theory in disease ecology and its extension to complex communities
619
620
Fig. 0. The community ecology of disease. Coinfection by (A) nematodes increases host
621
mortality due to (B) bovine TB among (C) African buffalo (63); Tsimane villagers (D) in Bolivia
622
reveal negative correlations between (E) Giardia lamblia and (F) Ascaris lumbricoides, where
623
de-worming increased Giardia (98). For tick-borne encephalitis (G), 93% of transmission events
624
involve large-bodied, male yellow-necked mice (H), which comprise <20% of the population
625
(53). For humans, disproportionate contact among individuals (I) can lead to ‘superspreading
626
events’ for SARS (J) (50). Among-species heterogeneities can alter community-wide
627
transmission; Crayfish plague (K) introduced to Europe with red swamp crayfish (L) led to
628
native crayfish declines; white-footed mice (M) are highly competent hosts for Borrelia
629
burgdorferi (N) and influence production of infected ticks that transmit Lyme borreliosis (65).
630
Image credits: CDC, R. Grencis, Y. Krishnappa, A. Pisor, CDC, F. Dubs, Stiasny et al. (2007),
631
V. Dostál, CDC, CDC, T. Vrålstad, F. Pupin, J. Brunner and NIH.
632
633
Fig. 1. Ecological hierarchies applied to host-parasite interactions and analogous processes in
634
community ecology. Included scales range from within-host (‘parasite infracommunity’, often
635
dominated by parasite-parasite and parasite-immune system interactions) to between-host
636
(‘parasite component community’, population biology) to among species (‘parasite
637
supracommunity’, community ecology) to across regions (macroecology and disease
638
biogeography). The different colored squares represent different parasite species, while the text
639
on the right and left sides of the figure highlight the relevant processes from community ecology
24
640
and disease ecology, respectively. The potential importance for interactions and feedback across
641
these scales represents an essential research frontier in the field of disease community ecology.
642
643
Fig. 2. Parasite community assembly depends on a combination of ecological selection (A),
644
ecological drift (B), and dispersal (C). Following input via dispersal (indicated as arrows from
645
the parasite regional pool), parasite establishment depends on ecological selection: different
646
species (mice versus prairie dogs) select for different parasites based on genetics, behavior,
647
immune status and other host properties (including vaccination status/drug presence).
648
Deterministic, within-host parasite interactions (indicated by + and – signs) are an additional
649
niche-based influence on parasite communities; positive parasite interactions (facilitation) are
650
indicated by thicker arrows, while negative interactions are indicated by dashed arrows. (B)
651
Parasite community assembly is also influenced by ecological drift (stochasticity), particularly
652
when colonizing populations are small or the outcome of parasite interactions depends on their
653
order of arrival (‘priority effects’). As a result, parasite communities can appear random with
654
respect to host species or type, even if strongly affected by species interactions. (C) High rates of
655
dispersal can swamp niche effects and overwhelm stochasticity, resulting in more similar
656
parasite communities across hosts, regardless of host species. For simplicity, no feedback loops
657
are shown from the individual hosts back to the parasite pool, although understanding such
658
feedbacks is an important research frontier (Fig. 3).
659
660
Fig. 3. (A) Using community ecology-based management strategies for infectious disease.
661
Levels of ecological organization are shown in the middle, and colored arrows indicate the
662
ecological processes that connect these levels. Parasite dispersal connects scales going up
25
663
through the hierarchy, while parasite establishment connects scales moving down the hierarchy.
664
Grey shaded arrows indicate the relative importance of offensive strategies (preventing parasite
665
dispersal) and defensive strategies (preventing parasite establishment), with darker shades
666
reflecting greater importance. (B) Management strategies focused on reducing spillover from
667
wildlife to humans (zoonosis) and from humans to wildlife (anthronosis or reverse zoonosis).
668
Probability of spillover and subsequent spread of infection can be reduced through four major
669
strategies. First, control may focus on reducing disease prevalence in reservoir hosts; for
670
instance, vaccine baits have been successfully used to eliminate rabies from several European
671
countries (79). Second, contact rates can be reduced between humans and wild animals (8); for
672
example, limiting the proximity between humans and wildlife can reduce spillover of human
673
illnesses, such as measles, tuberculosis, and MRSA, to wildlife. Third, zoonotic risk can be
674
reduced by lowering the probability of infection when contact is unavoidable or unpredictable.
675
For instance, some human Dengue vaccine candidates provide cross-protection against sylvatic
676
Dengue viruses, which naturally circulate in non-human primates (84). Finally, when spillover
677
does occur, regional control strategies – including isolating infected populations, dispatching
678
medical personnel and aid, and enhanced border control – can be used to prevent disease
679
transmission across borders.
680
26
681
Box 1: The role of simple theory in disease ecology and its extension to complex
682
communities
683
The pioneering work of Anderson and May (60, 99) formalized our understanding of parasite
684
dynamics by highlighting the importance of the basic reproductive number (R0) as a measure of
685
whether a parasite will spread through a population (R0 > 1) or die out (R0 < 1). The fundamental
686
principles of these basic models – initially developed for single host-single parasite systems –
687
can provide insight into infection dynamics in more complex ecological systems. For example.
688
parasites often face a diverse community of potential host species, which differ in abundance,
689
susceptibility and infectiousness. Simple extensions of basic disease ecology theory can
690
determine the conditions under which one host species amplifies or dilutes infection risk for
691
other species in the community. For directly transmitted parasites, or even those transmitted via
692
infective stages in the environment, theory shows that the parasite's overall basic reproductive
693
number among the available host community (R0,TOT) can simply be proportional to the sum of
694
the R0 for each host species alone, provided there is equal mixing within and between host
695
species (although other relationships between the individual-level and community-level R0
696
values may occur if mixing is not equal) (10, 100). Hence, there is a clear connection between
697
this more complex scenario and the classical single-host theory.
698
This theory can be extended further for vector-borne parasites, which become
699
complicated if hosts differ in their relative competencies for the parasite and the vector. For
700
example, tick-borne parasites may involve a mammalian host species that is parasite-competent
701
but cannot support tick reproduction, and another mammalian species that is non-competent for
702
the parasite but essential for tick reproduction (Box 1, Fig. 1). Here there are three possible
703
outcomes: (1) tick-and-parasite exclusion, (2) tick persistence but parasite exclusion and (3) tick-
27
704
and-parasite persistence, depending on different combinations of the R0 values for the parasite
705
(R0,parasite) and the tick (R0,tick). Ultimately, this results in outcomes that are nonlinearly related to
706
the density of the non-competent host; initial increases in non-competent host abundance (N) can
707
cause vector amplification leading to increased parasite R0, whereas high N dilutes transmission
708
through ‘wasted’ bites on the non-competent host (101-103).
709
710
Figure legend for Box
711
Model of a tick-borne parasite system with two host species, showing potential for both
712
amplification and dilution within the same system. (A) Schematic diagram of the model, where
713
one host species (C) is parasite-competent but cannot support tick reproduction, and the other (N)
714
is non-competent but essential for tick reproduction. This system can be described by the
715
equations (modified from (101-103)):
716
d𝐶p
= (1 − 𝐶p )𝛽1 𝑇𝐼 − 𝛿𝑅 𝐶P
d𝑡
717
d𝑇
= (𝑎 𝑇 − 𝑠𝑇 𝑇)𝑇𝛽3 𝑁 − 𝛿𝑇 𝑇
d𝑡
718
d𝑇𝐼
= (𝑇 − 𝑇𝐼 )𝛽2 𝐶𝐶𝑝 − 𝛿𝑇 𝑇𝐼 − 𝑇𝐼 𝛽3 𝑁
d𝑡
719
where T is the total number of ticks, (TU in the figure is the number of uninfected ticks; TI is the
720
number of infected ticks), CP is parasite prevalence within C (CU in the figure is the number of
721
uninfected hosts; CI is the number of infected hosts), β1 is tick  C transmission rate of the
722
parasite, β2 is C  tick transmission rate of the parasite, β3 is tick  N biting rate, the δi are
723
respective mortality rates, aT is tick reproduction rate and sT is the strength of tick density
724
dependence. (B) Phase plot of Competent host (C) – Non-competent host (N) densities, showing
725
the three regions of dynamical outcome, separated by the boundaries of R0,tick=1 and R0,parasite=1,
28
𝑎𝑇 𝛽3 𝑁
𝐶𝛽1 𝛽2 (𝑎𝑇 𝛽3 𝑁−𝛿𝑇 )
726
where 𝑅0,𝑡𝑖𝑐𝑘 =
727
competent host density, showing that low host densities facilitate parasite transmission due to
728
vector amplification whereas high host densities reduce parasite transmission through wasted
729
tick bites. The vertical line marked N’ (given by the value of N at which 𝑅0,𝑡𝑖𝑐𝑘 = 1 + √1 + 𝑎 𝑇 )
730
shows the non-competent host density at which the effect on the parasite switches from
731
amplification to dilution.
𝛿𝑇
and 𝑅0,𝑝𝑎𝑟𝑎𝑠𝑖𝑡𝑒 = 𝑠
. (C) R0,parasite as a function of non-
𝑇 𝛽3 𝑁𝛿𝐶 (𝛿𝑇 +𝛽3 𝑁)