2013 Book PrimatesPathogensAndEvolution
2013 Book PrimatesPathogensAndEvolution
Jessica F. Brinkworth
Kate Pechenkina Editors
Primates,
Pathogens, and
Evolution
Developments in Primatology:
Progress and Prospects
Primates, Pathogens,
and Evolution
Editors
Jessica F. Brinkworth Kate Pechenkina
Department of Pediatrics Department of Anthropology
CHU Sainte-Justine Research Center Queens College
University of Montreal City University of New York
Montreal, QC, Canada Flushing, NY, USA
First, we must thank the 30+ authors who accepted our invitation to contribute to
this book. This volume unites researchers from a wide range of biological
fields including Anthropology, Biochemistry, Evolutionary Biology, Genetics,
Immunology, Medicine, Veterinary medicine, Virology, and Zoology. These authors
provided the collection of papers that examine the molecular interactions between
primates and pathogens within the context of evolution contained within, and did so
while juggling many other responsibilities. The publication of an edited book is a
long process. Many of the authors represented here agreed to work with us as
early as 2009, when we were first recruiting speakers for a symposium at the annual
meeting of the American Association of Physical Anthropologists. We thank these
authors for contributing their time and effort to write interesting works and for
allowing us to curate such works here. Above all, we thank them for their sup-
port and patience.
A very special thanks to our colleagues who offered anonymous review of the
chapters and who must go unnamed. Each chapter in this collection was critically
examined by 2–3 researchers. Many of our peers provided detailed reviews on
tight deadlines. Some even sent reviews from the field!
Thanks to our editor Janet Slobodien who approached us and encouraged us to
pursue this book. Melissa Higgs, editorial assistant, guided the assembly of this
book, and often helped us with the fine details of assembling figures and permis-
sions. Thanks to Lesley Poliner and Hedge Ritya, who oversaw the production and
copy editing of this volume and its supporting materials. Yurii Chinenov provided
critical comments on multiple chapters. Thank you Jeremy Sykes for editorial assis-
tance and for proofing drafts.
The initial idea to develop a collection of papers that would discuss molecular
host–pathogen interactions came to us while attending a number of talks and posters
across scattered presentation sections at the annual meeting of the American
Association of Physical Anthropologists in Columbus, Ohio, in 2008. We wanted
to provide a forum for researchers interested in the evolution of primate immunity
to meet, discuss findings and brainstorm. At the 2009 AAPA meetings, we began to
vii
viii Acknowledgements
approach potential contributors to a symposium and edited volume that would focus
on the functional outcomes of evolutionary primate–pathogen interactions. We were
very fortunate to be met with great enthusiasm by our future contributors, in particu-
lar George Armelagos who immediately suggested we include the research of
Graham Rook, Jenny Tung, and Kristin Harper. The original participants of the
“Pathogens and evolution of human and non-human primates” symposium, held in
Albuquerque, New Mexico, in April of 2010, presented some truly interesting inter-
disciplinary works that day. We thank those researchers for sharing their ideas and
enthusiasm - George Armelagos, Nels C. Elde, Harmit Malik, Cedric Feschotte,
Charlie Nunn, Kristin Harper, Jayne Raper, Jenny Tung, Susan C. Alberts, Gregory A.
Wray, Felicia Gomez, Wen-Ya Ko, Sarah Tishkoff, Ajit Varki, Melanie Martin,
Caleb Finch, Fabian Crespo, Rafael Fernandez-Botran, Manuael Casanova, and
Christopher Tilquist. Thank you to the American Association of Physical
Anthropology and the Human Biology Association who hosted this symposium at
their annual meetings in Albuquerque, New Mexico, in April 2010. A very special
thanks to the donors, particularly members of the law firm Labaton Sucharow, who
provided funds and consideration associated with the needs of this specific sympo-
sium. A special thanks to Kelly Zieman, Leslie and Sharon Brinkworth, Cheryl and
Charles Brinkworth, Deirdre O’Boy, Emerson McCallum and Michael Donnelly.
Thank you to Chris Brinkworth, Mary Wong, Eva and JoAnn Brinkworth and my
parents, Cheryl and Charlie, for help and support during key stages of this book.
Special mention to Jenny Tung, who was an early supporter of and contributor to
this project. Thanks Luis Barreiro who offered assistance at an important stage of
production and who, with Jenny, has provided me the opportunity to develop a
career working on these questions of primate evolution and immunity. My deepest
gratitude to my husband, Jeremy, for his patience, energy and unreserved enthusi-
asm for all things book, career, life. Thank you for finding symposium donors,
batting around ideas, reviewing many proposal drafts and working extra hard on all
other tasks so that I could complete this one. Very special thanks must be given to
my daughter, Jordan, who was a considerate traveling companion and made this
process rather easy. My dearest dear, thank you for all days past, present and future.
Over the course of production Brinkworth and Pechenkina’s studies were sup-
ported by the Wenner-Gren Foundation Grant (7845 and 8702, JFB), the National
Science Foundation (0752297, KP and JFB), the Réseau de Médicine Génétique
Appliquée (JFB) and the National Institutes of Health (1R01-GM102562 to Luis
Barreiro, which supports JFB). Thank you also to the City University of New York,
Queens College, City College of New York, Sophie Davis School of Biomedical
Education, the New York Consortium in Evolutionary Primatology, Centre
Hospitalier Universitaire Sainte-Justine Research Center, and the University of
Montreal for their support and the opportunity to pursue our interests in the evolu-
tion of disease and immune system function.
ix
x Contents
Introduction
Primate immune systems have evolved to interact with pathogens in different ways
(Mandl et al. 2008, 2011; Pandrea et al. 2007; Sawyer et al. 2004; Song et al. 2005;
Soto et al. 2010). Human and nonhuman primate immune systems have diverged
with some species exhibiting strong differences in immune response to certain
pathogens including immunodeficiency viruses [reviewed by Pandrea and Apetrei
(2010), Toxoplasma gondii (Epiphanio et al. 2003), herpesviruses (Estep et al. 2010;
Huang et al. 1978), and trypanosomes (Thomson et al. 2009; Welburn et al. 2001)].
Why closely related primates have evolved such divergent pathogen interaction
strategies is not well understood. Despite strong public interest in human/nonhuman
primate evolutionary history and the importance of various primate species as bio-
medical models, the current picture of interspecies differences in immunity remains
fairly incomplete. Our understanding of how primate immunity evolved is hindered
by disconnected research on primate–pathogen molecular interaction, an uneven
focus on primate coevolution with a limited number of pathogens, and the discon-
nect between research on primate molecular phylogeny and primate physiology. The
objective of the present collection of papers is to integrate research on the evolution
of primate genomes, primate immune function, primate–pathogen biochemical
interaction, and infectious disease emergence to provide a knowledge base for future
research on human and nonhuman primate speciation, immunity, and disease.
Evolution of the immune system is tied to the evolutionary history of a species and
intertwined with the evolution of physiological functions and developmental stages
of an organism. The majority of cold-blooded vertebrates appear to experience
functional shifts in their immune response depending on external temperatures,
resulting, in some cases, in impaired immune responses such as the inhibition of
immunoglobulin class switching (Jackson and Tinsley 2002). The importance of a
graft-rejection-like immune response during tadpole to adult morphogenesis in the
frog genus Xenopus and the failure of novel proteins associated with lactation to
stimulate “nonself” immune responses in mammals, for example, both suggest that
the immune system has likely evolved in parallel with the evolution of species
developmental stages (Izutsu 2009; Matzinger 1994, 2002).
To interpret variation in primate immune response within and between species,
the evolutionary forces that shaped the underlying molecular differences need to be
examined. Of these forces, pathogen-mediated natural selection has likely been the
leading factor in increasing the frequency of pathogen resistance in host popula-
tions. Factors such as an organism’s environment, diet, postural behavior, and soci-
ality led to interspecies differences in exposure to specific pathogens and the
frequency with which such pathogens were encountered. Complete or partial resis-
tance of specific human genotypes to certain pathogenic strains and the patterned
distribution of these genotypes among indigenous populations around the globe is
fairly well documented (Hamblin and Di Rienzo 2000; Hraber et al. 2007; Leffler
et al. 2013; Marmor et al. 2001; Tishkoff et al. 2001). While the genetic variability
of nonhuman primate immune factors is comparatively less well studied, specific
disease resistance genotypes in some of these species have been identified. Ecological
differences between savanna dwelling Guinea baboons (Papio papio) and chimpan-
zees (Pan troglodytes) are likely responsible for resistance on the part of the former
species to the savannah-based pathogen Trypanosoma brucei gambiensis, and high
susceptibility to fatal T. brucei-caused sleeping sickness in the latter. Baboons are
more involved in grassland ground foraging than chimpanzees, which also exploit
savannah resources but are tied to forested regions and tend to retreat to the forest
for sleep (Kageruka et al. 1991). In the course of their evolution in open habitats,
baboons have likely been continuously exposed to the low flying tsetse fly—the
vector for T. brucei (Lambrecht 1985; Welburn et al. 2001). The selective pressure
generated by T. brucei is thought to have favored fixation of two nonconserva-
tive mutations in the baboon trypanosomal lytic factor ApoLI that have been linked
to sleeping sickness resistance and are not shared with either humans or chimpan-
zees (Thomson et al. 2009). In humans, two unrelated mutations in ApoLI that
increase sleeping sickness resistance are geographically restricted to Africa and
absent in European populations. Interestingly, trypanolytic variants of ApoLI in
humans contribute to an increased risk of kidney disease, providing an example of
Primates, Pathogens and Evolution: An Introduction 3
heterozygous advantage (Genovese et al. 2010) parallel to the classic case of the
HbS variant of the HBB globin locus and other hemoglobinopathies that confer
heterozygous advantage by means of increasing resistance to malaria(Allison 1956;
Haldane 1949; Hedrick 2004; Kwiatkowski 2005; Lapoumeroulie et al. 1992; Oner
et al. 1992).
The pathogen-directed evolutionary mechanisms that contribute to the divergence
of immune systems remain hypothetical. In the simplest case, an epidemic of a highly
virulent infectious disease eliminates a large number of susceptible organisms very
quickly, thereby favoring the disproportional reproductive success of resistant indi-
viduals. In such a scenario pathogen-mediated selection is assumed to be strong.
However, such a model of host–pathogen coevolution is pertinent in extreme cases
only. Pathogens of moderate-to-low virulence can also affect host immune allele
frequencies by, for example, contributing to lowered fertility in the form of physical
inability to produce offspring after being infected or causing a lesser ability to acquire
mates as a result of decreased mobility (Cheney et al. 1988; Levin et al. 1988).
Pathogens that do not kill a host, therefore, can affect sexual selection and gene flow.
Of course, the consequences of a deadly epidemic may not be limited to removal of
the alleles that contribute to disease susceptibility, such as those of surface antigens
recruited by a pathogen to penetrate the host. Such strong selective pressure may also
affect alleles that contribute to increased susceptibility to coinfections or to overt
immune activity that leads to the development of secondary conditions (e.g., sepsis).
Pathogen pressure may also select for resistance traits and in doing so affect the
frequency of linked alleles in a population through selective sweeps.
The most popularized perception of host–pathogen evolutionary interaction is the
concept of a host–pathogen evolutionary arms race. This idea is derived from the
application of Leigh Van Valen’s 1973 Red Queen hypothesis to hosts and pathogens
(Van Valen 1973). The Red Queen hypothesis proposes that closely associated organ-
isms may coevolve so tightly that the likelihood of extinction for one or the other is
constant over geological time. Changes in one species affect the tightly coevolved
interface with another species, threatening either species with extinction. As stated by
the Red Queen in Lewis Carroll’s Through the Looking Glass: “Now, here, you see, it
takes all the running you can do, to keep in the same place.” Tightly coevolved hosts
and pathogens have to “run”—that is, evolve quickly—just to maintain a balance and
avoid extinction. Within this framework, it is tempting to characterize host–pathogen
relationships as beneficial for the host to recognize and either tolerate or eliminate a
pathogen, with a pathogen’s main recourse being to evade a host’s defensive mecha-
nisms. These interactions are thought to result in head-to-head collisions between a
host’s immune system and a pathogen, leading to selective pressure on the immune
system and the pathogen and culminating in coadaptation.
However, some pathogens actually co-opt normal mechanisms of primate
immune recognition and the subsequent responses, such as cytokine release, cell
trafficking, and tissue destruction, using them to their advantage. Mycobacterium
tuberculosis (tuberculosis) infection is enhanced by the release of host anti-
inflammatory cytokine IL-10 (Redford et al. 2011). Once consumed by a macro-
phage, Yersinia pestis (plague) migrates to its main point of dissemination, the
4 J.F. Brinkworth and K. Pechenkina
lymph nodes, while acquiring phagocytosis resistance (Oyston et al. 2000; Pujol
and Bliska 2003; Zhou et al. 2006). Similarly, HIV can disseminate to the lymph
nodes and other regions through antigen-presenting cells (Koppensteiner et al.
2012). Moreover, exaggerated and uncontrolled immune cell responses may result
in host death—such is the case of severe sepsis and septic shock triggered by strong
innate immune cell recognition of immune system insult (e.g., blood stream infec-
tions and injury) (Brown et al. 2006; Murdoch and Finn 2003; Zemans et al. 2009).
Rather than acquiring permanent and costly adaptations, some pathogens can escape
host immune defenses through transient amplification of a resistance gene that
creates tandem arrays aptly named “genomic accordions.” Poxviruses encode two
factors, E3L and K3L, which inhibit the host antiviral factor protein kinase R (PKR).
In viruses lacking E3L, K3L rapidly becomes amplified 10–15-fold via serial dupli-
cation, which increases viral fitness. However, the tradeoff for increased genome
size is less efficient replication. Remarkably, an expanded genomic array of identi-
cal resistance genes in a pathogen increases the probability of emergence, fixation,
and spreading of additional K3L mutants with improved host avoidance; these
viruses subsequently lose the K3L duplicated array but retain the novel resistance
mutation (Elde et al. 2012).
Two limiting factors for pathogen-driven evolution of the immune system are
disadvantages rendered to the host by hyper-responsiveness, leading to autoimmune
disorders, and inadvertent pressure on symbiotic and commensal organisms. An
overactive immune system is responsible for the development of chronic conditions
mediated by the immune system itself, such as systemic lupus erythematosus,
antiphospholipid syndrome, polycystic ovary syndrome, and diabetes, among oth-
ers, that negatively affect reproductive fitness and may therefore contribute to
immune system evolution (Carp et al. 2012).
Consequently, immune system divergence between species is not driven by host–
pathogen interaction alone but is profoundly affected by the living environment,
which includes a complex network of interspecies interactions between hosts, sym-
bionts, commensals, and pathogens (Klimovich 2002; Lee and Mazmanian 2010).
A species, with its associated microorganisms, can be considered a “holobiont,” an
evolutionary unit encompassing the totality of organisms involved in commensal-
istic, symbiotic, and parasitic relations (Zilber-Rosenberg and Rosenberg 2008).
Changes in the fitness of any holobiont-involved species affect the system as a
whole, rather than just the fitness of the host species. Indeed, changes in the compo-
sition of intestinal microbiota may promote outgrowth or invasion by pathogenic
microorganisms or induce an exaggerated host response that results in the onset of
various autoimmune disorders such as inflammatory bowel disease (Maynard et al.
2012). Alternatively, dietary and environmental changes may cause a shift in
species-associated microbiota that results in the appearance of new pathogens or
symbionts. Establishing whether new species-specific pathogens are effectors or
consequences of speciation is a daunting task.
Primates, Pathogens and Evolution: An Introduction 5
In areas where human settlements and nonhuman primate habitats overlap, the
potential for interspecies disease transmission increases dramatically (Chapman
et al. 2005; Daszak et al. 2000; Duval and Ariey 2012; Reynolds et al. 2012; Stothard
et al. 2012; Wheatley and Harya Putra 1994). Such transmission events can decimate
wild primate populations, as new infectious diseases may profoundly affect animal
survival, sociality, and reproduction (Berdoy et al. 2000; Nunn et al. 2008; Nunn
2012). Increased ecotourism and residential/agricultural contact has led to height-
ened transmission of anthroponotic pathogens to wild primates and subsequently to
increased mortality in primate populations [e.g., chimpanzees and Polio virus (Wallis
and Lee 1999), gorillas and respiratory disease (Palacios et al. 2011), baboons and
Schistosoma mansoni (Farah et al. 2003; Murray et al. 2000), chimpanzees and para-
myxoviruses (Kondgen et al. 2008), chimpanzees and Schistosoma mansoni (Stothard
et al. 2012), and baboons and Mycobacterium (Keet et al. 2000)]. Plasmodium falci-
parum may have been introduced anthropolonotically to the neotropical primates
during the colonial era, possibly through the forced migration of African slaves dur-
ing the slave trade. Neotropical primate P. simium has been proposed to have emerged
from Asian P. vivax during the nineteenth century, and perhaps introduced to South
America by laborers from East Asia (reviewed in Cormier 2010).
8 J.F. Brinkworth and K. Pechenkina
The transmission of pathogens from nonhuman primates to humans has also had
a profound effect on human health. The current HIV-1/AIDS pandemic likely origi-
nated with human consumption of SIV-contaminated nonhuman primate bushmeat
(Gao et al. 1999). The emergence of other diseases in humans has likewise been
attributed to human and nonhuman primate contact [i.e., Human T-Lymphotropic
virus 1 (Vandamme et al. 1998), Monkeypox (Mutombo et al. 1983; Reynolds et al.
2012), and Malaria (Liu et al. 2010)]. As humans encroach even farther onto nonhu-
man primate ranges, the need for veterinary and medical intervention will increase.
To be able to develop appropriate modes of intervention, it is very important to have
good information on the broad differences and similarities of primate immune systems
as well as the biochemical mechanics of specific primate–pathogen interactions.
Overview
The first section, Immunity and Primate Evolution, includes chapters that discuss major
elements of the primate immune system (Brinkworth and Thorn; Brinkworth and
Sterner), the use of primates as models of immune system evolution (Loisel and Tung)
Primates, Pathogens and Evolution: An Introduction 9
and the pathogen-mediated evolution of primates (Gomez et al. and Allen et al.).
The second section, Emergence and Divergent Disease Manifestation, provides data
on the emergence, biochemical mechanics, and interspecies differences in immune
response to immunodeficiency viruses, as well as a range of clinically important, but
otherwise neglected pathogens. Attention is focused on how some primate pathogens
have emerged (Harper and Knauf; Greenwood et al.), coevolved with and escaped the
defenses of their hosts (Wong et al.), and triggered divergent responses in different
primate species (Catao-Dias et al. and Greenwood et al.). The third and final section,
Primates, Pathogens, and Health, focuses on how primate–pathogen coevolution
affects the health of modern primates. Three papers on the health and evolutionary
impact of disruptions to human and microorganism association (Rook, Martin and
Blackwell, and Harper et al.) review and test the hygiene hypothesis. The volume
closes with an analysis of major human and nonhuman primate cross-species pathogen
transmission events, the social and biological factors that contributed to those events,
and what the evolutionary, public health, and conservation consequences of these
events might be (Harper et al.). We thought it a wonderful chapter with which to close.
Conclusion
Primate immune systems have been formed through complex evolutionary pro-
cesses, within which pathogens have played an important role. The evolution of
primate immunity has likely been more nuanced than natural selection driven by
coevolutionary arms races with pathogens or large pathogen-mediated selective
sweeps of hosts. Rather, the evolution of the primate immune system has likely been
considerably shaped by, amongst other possibilities, moderately virulent pathogens,
pathogen strategies that co-opt normal immunity, viruses integrated into the primate
genome, commensal microorganism maintenance, autoreactivity, and overt immune
responses, as well as the evolution of primate developmental stages. As close rela-
tives, animals within the order Primates can be comparatively studied to not only
clarify how primate immune systems have functionally diverged and highlight
therapeutic targets for disease, but also to help define how such divergence contrib-
utes to disease emergence and interspecies disease transmission. As such, mapping
the evolution of the primate immune system can improve our understanding of
primate speciation, primate conservation, and human health. This volume is one
effort to unite information on the evolutionary interactions between primate immune
systems and pathogens. The chapters in this volume represent research from a broad
range of disciplines involved in the study of primate–pathogen molecular interac-
tion, primate immune function, and primate–pathogen coevolution. The work pre-
sented here discusses primate interactions with both major and neglected pathogens,
attempts to bridge research on molecular evolution and primate immune function,
and illustrates the impact of primate–pathogen evolutionary interactions on
human and nonhuman primate health. With this effort we aim to provide a sound
base of knowledge for future investigation of human and nonhuman primate evolu-
tion, immunity, and disease.
10 J.F. Brinkworth and K. Pechenkina
References
Farah I, Borjesson A, Kariuki T, Yole D, Suleman M, Hau J, Carlsson HE (2003) Morbidity and
immune response to natural schistosomiasis in baboons (Papio anubis). Parasitol Res
91(4):344–348
Filip LC, Mundy NI (2004) Rapid evolution by positive Darwinian selection in the extracellular
domain of the abundant lymphocyte protein CD45 in primates. Mol Biol Evol 21(8):
1504–1511
Gao, F, Bailes E, Robertson DL, Chen Y, Rodenburg CM, Michael SF, Cummins LB, Arthur LO,
Peeters M, Shaw GM, Sharp PM, Hahn, BH (1999) Origin of HIV-1 in the chimpanzee Pan
troglodytes troglodytes. Nature 397:436–441
Genovese G, Friedman DJ, Ross MD, Lecordier L, Uzureau P, Freedman BI, Bowden DW,
Langefeld CD, Oleksyk TK, Uscinski Knob AL et al (2010) Association of trypanolytic ApoL1
variants with kidney disease in African Americans. Science 329(5993):841–845
Gogvadze E, Stukacheva E, Buzdin A, Sverdlov E (2009) Human-specific modulation of transcrip-
tional activity provided by endogenous retroviral insertions. J Virol 83(12):6098–6105
Greenwood EJD, Schmidt F, Heeney JL (2013) The evolution of SIV in primates and the emer-
gence of the pathogen of AIDS. In: Brinkworth JF, Pechenkina E (eds) Primates, pathogens,
and evolution. Springer, Heidelberg
Haldane JBS (1949) The rate of mutation of human genes. Heredita 35(suppl):267–273
Hamblin MT, Di Rienzo A (2000) Detection of the signature of natural selection in humans: evi-
dence from the Duffy blood group locus. Am J Hum Genet 66(5):1669–1679
Hedrick P (2004) Estimation of relative fitnesses from relative risk data and the predicted future of
haemoglobin alleles S and C. J Evol Biol 17(1):221–224
Horie M, Honda T, Suzuki Y, Kobayashi Y, Daito T, Oshida T, Ikuta K, Jern P, Gojobori T, Coffin
JM et al (2010) Endogenous non-retroviral RNA virus elements in mammalian genomes.
Nature 463(7277):84–87
Hraber P, Kuiken C, Yusim K (2007) Evidence for human leukocyte antigen heterozygote advan-
tage against hepatitis C virus infection. Hepatology 46(6):1713–1721
Huang ES, Kilpatrick B, Lakeman A, Alford CA (1978) Genetic analysis of a cytomegalovirus-like
agent isolated from human brain. J Virol 26(3):718–723
Hunter P (2010) The missing link. Viruses revise evolutionary theory. EMBO Rep 11(1):28–31
Izutsu Y (2009) The immune system is involved in Xenopus metamorphosis. Front Biosci
14:141–149
Jackson JA, Tinsley RC (2002) Effects of environmental temperature on the susceptibility of
Xenopus laevis and X. wittei (Anura) to Protopolystoma xenopodis(Monogenea). Parasitol Res
88(7):632–638
Jia B, Serra-Moreno R, Neidermyer W, Rahmberg A, Mackey J, Fofana IB, Johnson WE,
Westmoreland S, Evans DT (2009) Species-specific activity of SIV Nef and HIV-1 Vpu in
overcoming restriction by tetherin/BST2. PLoS Pathog 5(5):e1000429
Jones JL, Kruszon-Moran D, Sanders-Lewis K, Wilson M (2007) Toxoplasma gondii infection in the
United States, 1999 2004, decline from the prior decade. Am J Trop Med Hyg 77(3):405–410
Kageruka P, Mangus E, Bajyana Songa E, Nantulya V, Jochems M, Hamers R, Mortelmans J
(1991) Infectivity of Trypanosoma (Trypanozoon) brucei gambiense for baboons (Papio hama-
dryas, Papio papio). Ann Soc Belg Med Trop 71(1):39–46
Keet DF, Kriek NP, Bengis RG, Grobler DG, Michel A (2000) The rise and fall of tuberculosis in
a free-ranging chacma baboon troop in the Kruger National Park. Onderstepoort J Vet Res
67(2):115–122
Kim N, Dabrowska A, Jenner RG, Aldovini A (2007) Human and simian immunodeficiency virus-
mediated upregulation of the apoptotic factor TRAIL occurs in antigen-presenting cells from
AIDS-susceptible but not from AIDS-resistant species. J Virol 81(14):7584–7597
Kim HS, Kim DS, Huh JW, Ahn K, Yi JM, Lee JR, Hirai H (2008) Molecular characterization of
the HERV-W env gene in humans and primates: expression, FISH, phylogeny, and evolution.
Mol Cells 26(1):53–60
Klimovich VB (2002) Actual problems of evolutional immunology. Zh Evol Biokhim Fiziol
38(5):442–451
12 J.F. Brinkworth and K. Pechenkina
Kondgen S, Kuhl H, N'Goran PK, Walsh PD, Schenk S, Ernst N, Biek R, Formenty P, Matz-
Rensing K, Schweiger B et al (2008) Pandemic human viruses cause decline of endangered
great apes. Curr Biol 18(4):260–264
Koppensteiner H, Brack-Werner R, Schindler M (2012) Macrophages and their relevance in human
immunodeficiency virus type I infection. Retrovirology 9:82
Koyanagi M, Kerns JA, Chung L, Zhang Y, Brown S, Moldoveanu T, Malik HS, Bix M (2010)
Diversifying selection and functional analysis of interleukin-4 suggests antagonism-driven
evolution at receptor-binding interfaces. BMC Evol Biol 10:223
Kwiatkowski DP (2005) How malaria has affected the human genome and what human genetics
can teach us about malaria. Am J Hum Genet 77(2):171–192
Lambrecht FL (1985) Trypanosomes and Hominid evolution. Bioscience 35(10):640–646
Lapoumeroulie C, Dunda O, Ducrocq R, Trabuchet G, Mony-Lobe M, Bodo JM, Carnevale P,
Labie D, Elion J, Krishnamoorthy R (1992) A novel sickle cell mutation of yet another origin
in Africa: the Cameroon type. Hum Genet 89(3):333–337
Lee YK, Mazmanian SK (2010) Has the microbiota played a critical role in the evolution of the
adaptive immune system? Science 330(6012):1768–1773
Leffler EM, Gao Z, Pfeifer S, Segurel L, Auton A, Venn O, Bowden R, Bontrop R, Wall JD, Sella
G et al (2013) Multiple instances of ancient balancing selection shared between humans and
chimpanzees. Science 339(6127):1578–1582
Levin JL, Hilliard JK, Lipper SL, Butler TM, Goodwin WJ (1988) A naturally occurring epizootic
of simian agent 8 in the baboon. Lab Anim Sci 38(4):394–397
Liu W, Li Y, Learn GH, Rudicell RS, Robertson JD, Keele BF, Ndjango JB, Sanz CM, Morgan DB,
Locatelli S et al (2010) Origin of the human malaria parasite Plasmodium falciparum in goril-
las. Nature 467(7314):420–425
Loisel DA, Rockman MV, Wray GA, Altmann J, Alberts SC (2006) Ancient polymorphism and
functional variation in the primate MHC-DQA1 5′ cis-regulatory region. Proc Natl Acad Sci
USA 103(44):16331–16336
Maier AG, Duraisingh MT, Reeder JC, Patel SS, Kazura JW, Zimmerman PA, Cowman AF (2003)
Plasmodium falciparum erythrocyte invasion through glycophorin C and selection for Gerbich
negativity in human populations. Nat Med 9(1):87–92
Mandl JN, Barry AP, Vanderford TH, Kozyr N, Chavan R, Klucking S, Barrat FJ, Coffman RL,
Staprans SI, Feinberg MB (2008) Divergent TLR7 and TLR9 signaling and type I interferon
production distinguish pathogenic and nonpathogenic AIDS virus infections. Nat Med
14(10):1077–1087
Mandl JN, Akondy R, Lawson B, Kozyr N, Staprans SI, Ahmed R, Feinberg MB (2011) Distinctive
TLR7 signaling, type I IFN production, and attenuated innate and adaptive immune responses
to yellow fever virus in a primate reservoir host. J Immunol 186(11):6406–6416
Marmor M, Sheppard HW, Donnell D, Bozeman S, Celum C, Buchbinder S, Koblin B, Seage GR
3rd (2001) Homozygous and heterozygous CCR5-Delta32 genotypes are associated with resis-
tance to HIV infection. J Acquir Immune Defic Syndr 27(5):472–481
Matzinger P (1994) Tolerance, danger, and the extended family. Annu Rev Immunol 12:
991–1045
Matzinger P (2002) The danger model: a renewed sense of self. Science 296(5566):301–305
Maynard CL, Elson CO, Hatton RD, Weaver CT (2012) Reciprocal interactions of the intestinal
microbiota and immune system. Nature 489(7415):231–241
Moesta AK, Abi-Rached L, Norman PJ, Parham P (2009) Chimpanzees use more varied receptors
and ligands than humans for inhibitory killer cell Ig-like receptor recognition of the MHC-C1
and MHC-C2 epitopes. J Immunol 182(6):3628–3637
Murdoch C, Finn A (2003) The role of chemokines in sepsis and septic shock. Contrib Microbiol
10:38–57
Murray S, Stem C, Boudreau B, Goodall J (2000) Intestinal parasites of baboons (Papio cyno-
cephalus anubis) and chimpanzees (Pan troglodytes) in Gombe National Park. J Zoo Wildl
Med 31(2):176–178
Primates, Pathogens and Evolution: An Introduction 13
chimpanzees: evidence for a recent introduction following chimpanzee divergence. AIDS Res
Hum Retroviruses 21(5):335–342
Thomson R, Molina-Portela P, Mott H, Carrington M, Raper J (2009) Hydrodynamic gene delivery
of baboon trypanosome lytic factor eliminates both animal and human-infective African try-
panosomes. Proc Natl Acad Sci USA 106(46):19509–19514
Tishkoff SA, Varkonyi R, Cahinhinan N, Abbes S, Argyropoulos G, Destro-Bisol G, Drousiotou A,
Dangerfield B, Lefranc G, Loiselet J et al (2001) Haplotype diversity and linkage disequilib-
rium at human G6PD: recent origin of alleles that confer malarial resistance. Science
293(5529):455–462
Van Heuverswyn F, Li Y, Neel C, Bailes E, Keele BF, Liu W, Loul S, Butel C, Liegeois F, Bienvenue
Y et al (2006) Human immunodeficiency viruses: SIV infection in wild gorillas. Nature
444(7116):164
Van Valen L (1973) A new evolutionary law. Evolut Theor 1:1–30
Vandamme AM, Salemi M, Desmyter J (1998) The simian origins of the pathogenic human T-cell
lymphotropic virus type I. Trends Microbiol 6(12):477–483
Vizoso AD (1975) Recovery of herpes simiae (B virus) from both primary and latent infections in
rhesus monkeys. Br J Exp Pathol 56(6):485–488
Wallis J, Lee DR (1999) Primate conservation: the prevention of disease transmission. Int J
Primatol 20(6):803–826
Wang T, Zeng J, Lowe CB, Sellers RG, Salama SR, Yang M, Burgess SM, Brachmann RK, Haussler
D (2007) Species-specific endogenous retroviruses shape the transcriptional network of the
human tumor suppressor protein p53. Proc Natl Acad Sci USA 104(47):18613–18618
Welburn SC, Fevre EM, Coleman PG, Odiit M, Maudlin I (2001) Sleeping sickness: a tale of two
diseases. Trends Parasitol 17(1):19–24
Wheatley B, Harya Putra DK (1994) Biting the hand that feeds you: Monkeys and tourists in
Balinese monkey forests. Trop Biodivers 2:317–327
Wlasiuk G, Nachman MW (2010) Adaptation and constraint at Toll-like receptors in primates. Mol
Biol Evol 27(9):2172–2186
Wooding S, Stone AC, Dunn DM, Mummidi S, Jorde LB, Weiss RK, Ahuja S, Bamshad MJ (2005)
Contrasting effects of natural selection on human and chimpanzee CC chemokine receptor 5.
Am J Hum Genet 76(2):291–301
Yim JJ, Adams AA, Kim JH, Holland SM (2006) Evolution of an intronic microsatellite polymor-
phism in Toll-like receptor 2 among primates. Immunogenetics 58(9):740–745
Yohn CT, Jiang Z, McGrath SD, Hayden KE, Khaitovich P, Johnson ME, Eichler MY, McPherson
JD, Zhao S, Paabo S et al (2005) Lineage-specific expansions of retroviral insertions within the
genomes of African great apes but not humans and orangutans. PLoS Biol 3(4):e110
Zemans RL, Colgan SP, Downey GP (2009) Transepithelial migration of neutrophils: mechanisms
and implications for acute lung injury. Am J Respir Cell Mol Biol 40(5):519–535
Zhou D, Han Y, Yang R (2006) Molecular and physiological insights into plague transmission,
virulence and etiology. Microbes Infect 8(1):273–284
Zhu T, Korber BT, Nahmias AJ, Hooper E, Sharp PM, Ho DD (1998) An African HIV-1 sequence
from 1959 and implications for the origin of the epidemic. Nature 391(6667):594–597
Zilber-Rosenberg I, Rosenberg E (2008) Role of microorganisms in the evolution of animals and
plants: the hologenome theory of evolution. FEMS Microbiol Rev 32(5):723–735
Part I
Immunity and Primate Evolution
Vertebrate Immune System Evolution
and Comparative Primate Immunity
Introduction
Thus, the innate immune strategy of phagocytosis was discovered. The comparative
immunological approach Metchnikoff employed goes a step beyond simply applying
a representative animal model to a question of immunity. Comparative immunology
involves comparing differences and similarities in organisms’ immune responses to
draw broader conclusions about immune system function and its evolution. This
approach can be useful for the identification of immune system components respon-
sible for particular disease phenotypes, as well as the discovery of novel immune
mechanisms. The use of other vertebrate animals such as jawless vertebrates,
amphibians, and rodents as biomedical models can shed light on overriding princi-
ples of immunity. Comparisons of primate immunity in the context of vertebrate
immune system evolution can be useful for the understanding of biomedical model
use, primate evolution, and human health.
Primates are a recent addition to the comparative exploration of animal immune
responses. Direct interspecies comparisons of primate immune system function
emerged in the 1920s but did not become common until decades later. The under-
standing of comparative primate immunity mainly developed through nonhuman
primate/human xenotransplantation studies in the 1960s, as well as immunodefi-
ciency virus research from the late 1980s onwards (Benveniste et al. 1986; Daniel
et al. 1984; Hardy et al. 1964; Hitchcock et al. 1964; Murphey-Corb et al. 1986;
Reemtsma et al. 1964a, b; Starzl et al. 1964). Until the adoption of catarrhine (Old
World monkey, apes, and humans) species for HIV research, few immunological
studies using different primate species compared interspecies differences. It is
now well established that primates exhibit within-order variation and, sometimes,
biologically unique manifestations of infectious diseases (Epiphanio et al. 2003;
Ngampasutadol et al. 2008; Pandrea and Apetrei 2010; Thomson et al. 2009;
Walker 1997). Still, monkey and ape species are commonly used in immunologi-
cal research as corollaries for human disease progression due to their biochemi-
cal, physiological, and genetic similarity to humans. Because there are so many
immune system similarities between primate species, there is a tendency in non-
HIV literature to make the assumption that what is represented in one primate
species is represented in other primate species and, possibly, other nonprimate
models. By making this assumption a researcher risks overlooking aspects of
primate immune systems that are unique and using primates unnecessarily to
explore immunological traits that they share with many other model organ-
isms. Comparative information on baseline primate immunity, particularly the
anatomy and function of major immune organs and cell types, is rare and scat-
tered across many fields. In certain cases it is entirely unexplored. The goal of
this chapter is to illustrate the place of primates in immune system evolution
by (1) putting the emergence of major primate immune system components in
the context of the evolution of vertebrate immunity as a whole and (2) illustrating
how baseline primate immunity has diversified by uniting and highlighting the
available information on interspecies functional differences in baseline primate
immune system structures and components.
Vertebrate Immune System Evolution and Comparative Primate Immunity 19
As jawed vertebrates, mammals maintain an immune system that can be broken into
two major arms based on function. The innate immune response is the more ancient
of these two arms, having invertebrate origins (Leulier et al. 2003; Yoshida et al.
1986). Innate immune defenses are inherited, germline encoded, nonspecific, and
typified by barriers (e.g., mucosa, skin), antimicrobial peptides, phagocytosis (initi-
ated by cells such as macrophages and neutrophils), and inflammation (Janeway
and Medzhitov 2002; Kumar et al. 2009). This kind of immunity limits initial infec-
tions by recognizing “nonself”, and damage through a variety of sophisticated but
generalized mechanisms, including inherited pattern recognition receptors (e.g.,
Toll-like receptors, NOD-like receptors) that detect foreign material through molec-
ular patterns associated with pathogens or cellular damage. These patterns can be
shared broadly by microorganisms or may signal tissue damage. They are conven-
tionally and somewhat imprecisely referred to as pathogen- or danger- associated
molecular patterns (PAMPS or DAMPS) (Seong and Matzinger, 2004).
By contrast, adaptive immunity is highly specific, not immediate, key to immu-
nological memory, modulated by innate immunity, and acquired over a lifetime.
While phagocytosis is an important tool in innate immune defenses, the targeting
of matter bearing specific epitopes by lymphocytes (e.g., T and B cells) and the
retention of some of these target-specific lymphocytes is key to adaptive immunity.
Lymphocytes express membrane receptors (T-cell receptors for T cells and B-cell
receptors for B cells) that recognize antigens. Unlike innate immunity, which makes
use of germline encoded receptors, adaptive immunity has been traditionally viewed
as reliant on receptors and immunoglobulins that are made highly variable through
recombination activating gene (RAG)-mediated gene rearrangement/somatic
recombination that occurs during lymphocyte development. From a limited number
of receptor genes is borne a broad repertoire of specific receptors. As a result, rather
than recognizing pathogens through PAMPS, the lymphocytes and immunoglobu-
lins of the adaptive system recognize and “remember” distinct epitopes [reviewed in
(Hardy 2003)].
The simplified view of the vertebrate immune system function is one of immedi-
ate recognition of invading pathogens by the innate immune system and subsequent
initiation of a specific adaptive immune response. In mammals, for example, when
innate immune cells recognize foreign antigens, they initiate the release of reactive
signaling proteins known as cytokines. Cytokines degrade pathogens nonspecifi-
cally and initiate activation of an epitope/pathogen-specific T and B cell-mediated
adaptive immune response. T and B cells can become activated when receptors they
bear belonging to the immunoglobulin receptor superfamily, T- and B-cell receptors
(TCRs and BCRs), recognize specific epitopes that are presented to them via major
histocompatibility complexes (MHC) on phagocytic cells (e.g., macrophages and
dendritic cells). B cells can also become activated through direct encounters with
pathogens bearing these epitopes. Activated T and B cells then clonally replicate in
20 J.F. Brinkworth and M. Thorn
Fig. 1 Major lymphoid structures and their emergence in vertebrate clades over time. The sym-
bols “+” and “-” indicate the presence/occurrence of and the absence of lymphoid structure,
respectively. “*” while most birds do not appear to maintain lymph nodes, they are present ducks
(Anatidae). Molecular divergence dates from Blair and Hedges (2005)
integrity of the gut barrier (Brenchley and Douek 2008; Favre et al. 2009; Raffatellu
et al. 2008). This process does not appear to occur in AIDS-resistant/natural IV host
species such as sooty mangabeys (Cercocebus atys), suggesting that the regulation of
GALT components (including T cell subsets) has diverged since these species last
shared a common ancestor approximately 30 mya and that these differences can affect
infectious disease progression (Brenchley et al. 2008; Fabre et al. 2009; Steiper and
Young 2006).
Thymus
The thymus is the oldest primary lymphoid organ, the earliest example of which can
be found in lampreys (Lampetra planeri and Petromyzon marinus) (Bajoghli et al.
2011). While in most jawed vertebrates both B- and T-cell poiesis occur in the bone
marrow, thymocytes (immature T cells) move to the thymus to mature (reviewed in
Boehm and Bleul 2007). There, they undergo T-cell receptor rearrangement, posi-
tive selection for MHC interaction, and negative selection for self-recognition.
These processes lead to the development of an enormous repertoire of T cells that
communicate with other immune cells and bear receptors that recognize individual
and specific epitopes.
Lampreys bear a thymus precursor at the tips of their gill filaments that maintains
T-like cells and expresses genes associated with the thymus in other animals (i.e.,
FOXN4L, a FOXN1 homologue; CDA1) (Bajoghli et al. 2011). All extant jawed
vertebrates with a canonical thymus also have T cells, TCRs, and MHCs, so it is
difficult to assess how these immune components evolved together. Considerable
vertebrate extinctions over the last 650 million years give the impression that this
entire system of receptor repertoire expansion and immune cell development simply
appeared, fully formed in cartilaginous fish (class Chrondrichthyes). This complex
system did not, of course, arise in this manner, but there is very little information
about the intermediate steps in its evolution. The appearance of the canonical thy-
mus and this system of T-cell development coincides with the acquisition of two
genes responsible for immunoglobulin domain receptor recombination (RAG1 and
RAG2) as well as with two rounds of whole genome duplication preceding the diver-
gence of jawed and jawless vertebrates (round 1: 652 mya) and the divergence of
bony and cartilaginous fish (round 2: 525 mya), respectively (Blair and Hedges
2005; Flajnik and Kasahara 2010; Ohno 1970; Rast et al. 1997). Boehm and Bleul
(2007) have proposed that the acquisition of the thymus is the outcome of natural
selection for a specialized site to eliminate autoreactive lymphocytes that might be
generated through RAG-mediated Ig domain receptor recombination. The discov-
ery of thymoid tissue in lampreys that predates the emergence of the TCR/MHC
system, however, conflicts with their hypothesis (Bajoghli et al. 2011; Boehm and
Bleul 2007).
The vertebrate thymus is derived from pharyngeal arches and pouches, the num-
ber of which differs across vertebrates. Unsurprisingly, the number, anatomical
position, and structure of thymi differ between species as well. Some bony fish
(superclass Osteichthyes), for example, have one thymus composed of two lobes,
whereas other cold-blooded vertebrates, birds, and placental mammals have multi-
ple thymi (reviewed in Rodewald 2008). By the time bony fish emerge, thymi
derived from the 3rd and 4th pharyngeal pouches with an adult location ranging
from the cervical to the thoracic region, like those in extant mammals, are estab-
lished (Grevellec and Tucker 2010; Rodewald 2008). While some mammals, such
as mice, have recently been described as having more than one thymus, primates
have a single thoracic thymus (Dooley et al. 2006; Terszowski et al. 2006; Wong
et al. 2011). Unlike fish and birds, whose thymi are attached to the pharynx for the
Vertebrate Immune System Evolution and Comparative Primate Immunity 25
whole of their lives, the mammalian thymus tends to migrate during development,
introducing the possibility of intra- and interspecies variation in its location (Dooley
et al. 2006; Grevellec and Tucker 2010; Lam et al. 2002). In humans, variation in
thymus position is rare and tends to be associated with disease (Shah et al. 2001;
Sturm-O’Brien et al. 2009; Tovi and Mares 1978).
Interspecies comparisons of thymus development and function in primates are
not common, however, some differences in development have been suggested by the
findings of Buse et al. (2006). Thymus development in cynomolgus/crab-eating
macaques (Macaca fascicularis) appears to be accelerated in comparison to humans,
with macaque lymphoid cells preceding the appearance and functional maturation of
human lymphoid cells by approximately four gestational weeks (Buse et al. 2006). By
comparison, the infiltration of other cells into the thymus occurs later in cynomolgus
macaques than in humans, with macrophages appearing at week 10 in the former and
week 8 in the latter. By parturition, however, the thymi of these species appears to be
similarly developed. Some of these differences in gestational development are likely
associated with a near 120-day discrepancy in gestation between these species.
Bone Marrow
The presence of bone marrow in some species of cartilaginous fish suggests that mar-
row emerged in some species in the absence of bone (Tavassoli 1986). In mammals,
bone marrow becomes a site of lymphopoiesis and takes on the role of a primary
lymphoid organ. In other animal classes, lymphocyte development can occur in other
locations. In adult amphibians, for example, bone marrow expresses genes important
for lymphocyte development (e.g., RAG genes), but B cells mainly undergo differen-
tiation in the spleen and liver (Du Pasquier et al. 2000; Greenhalgh et al. 1993).
Amongst mammals, strong interspecies differences in haematopoiesis have been
noted between rodents and hominoid primates. A simple indication of potential dif-
ferences in bone marrow function between mice and hominoids is the disparity in
the relative proportions of circulating neutrophils between species. Neutrophils
emerge from bone marrow as terminally differentiated granulocytes and only circu-
late for a few hours. They represent 50–70 % of the circulating blood leukocytes in
hominoids such as humans, and only 10–25 % in mice, suggesting an interspecies
difference in the rate of production (Doeing et al. 2003; Mestas and Hughes 2004).
Similarly, Old World monkeys have been noted to have lower percentages of circu-
lating neutrophils (10–42 %) than humans, though these numbers vary depending
on the living conditions of the animals (reviewed in Haley 2003). More complex
differences between haematopoietic stem cell receptors have been noted between
primates and other mammals, with humans exhibiting low CD117 and high CD135
expression on stem cell progenitors and mice exhibiting the inverse pattern (Sitnicka
et al. 2003). Both of these stem cell markers are tyrosine kinase receptors.
Differences in their expression suggest interspecies differences in tyrosine kinase
activity and possibly signal transduction in these cells in the marrow.
26 J.F. Brinkworth and M. Thorn
Spleen
The spleen is a highly vascularized organ that filters dead/damaged cells and foreign
material. Its primary immune function in mammals is to protect the host from
blood-borne pathogens (Diamond 1969; Evans 1985). In the healthy primate spleen,
this is achieved through either the filtering of antigens by macrophages and den-
dritic cells located in marginal zones surrounding arterioles, migration of activated
dendritic cells to splenic T cells regions and presentation of antigen to T cells, or
trapping of activated B cells in the splenic T-cell zone and their subsequent interac-
tion with antigen-specific T helper cells and proliferation (reviewed in Cesta 2006b).
Though lampreys maintain a primordial spleen, the organ is found in cartilagi-
nous fish and first emerged as an independent structure in the last common ancestor
to extant jawed vertebrates 652–525 million years ago (reviewed in Boehm et al.
2012a)(Blair and Hedges 2005). Splenic function and structure has changed consid-
erably over evolutionary time and varies strongly from animal class to animal class.
In mammals, for example, the spleen is also a very important site of B-cell develop-
ment. Mammalian B cells initially develop in the bone marrow, but typically mature
in the spleen (Brendolan et al. 2007; Drayton et al. 2006). This is not the case for
other animal classes. In cartilaginous fish B-cell development starts in the liver and
migrates to the kidneys before the final stages occur in the spleen (Du Pasquier
1973). In bony fish, B-cell development bypasses the spleen all together and occurs
mainly in the kidneys (or pronephros). In birds, B cells spend the earliest stages of
their development in the spleen before migrating to the bursa of Fabricius, located
near the cloaca (Boehm et al. 2012a; Pickel et al. 1993). While a perpetual blood
filter, the precise role a spleen plays in the immune system of an animal varies con-
siderably across the animal kingdom.
The complex structure of the primate spleen, with red pulp with a white pulp
compartmentalized into peri-arteriolar lymphoid sheath (PALS), follicular dendritic
cell clusters, and a marginal zone with germinal centers for B-cell proliferation,
only emerged with the last common ancestor of mammals (Fig. 2) (Cesta 2006b;
Hofmann et al. 2010). The spleen of cartilaginous fish bears a familiar structure of
blood filtering and component storing red pulp containing lymphocyte organizing
white pulp, with well-defined lymphocyte zones (Rumfelt et al. 2002). The white
pulp has become more complex over evolutionary time, with the spleens of more
recently emerged animal classes taking on progressively more complex white pulp
structure and function (Jeurissen 1991; Zapata and Amemiya 2000).
In the simplest of terms, mammalian white pulp is comprised of lymphoid tissue
that surrounds arterioles, with a centralized area containing resting or proliferating
B cells surrounded by a peripheral area where T cells congregate. This is, however,
a highly derived structure. The white pulp of cartilaginous fish is structured in the
opposite fashion of more recently emerged species, with centralized zones of T
cells, dendritic cells, and immunoglobulin-producing cells, encircled by B-cell clus-
ters (Rumfelt et al. 2002). Bony fish maintain spleens with these same basic com-
ponents, but have developed end capillary structures known as ellipsoids, covered
in macrophages and involved in antigen capture. Because of the position of
Vertebrate Immune System Evolution and Comparative Primate Immunity 27
Fig. 2 Evolution of the splenic white pulp. Positioning of Reptilia, Aves, and Mammalia based on
Blair and Hedges (2005) and Shedlock and Edwards (2009). GC germinal centers, FDCs follicular
dendritic cells, PELS peri-ellipsoid lymphocyte sheath, PALS peri-arteriolar lymphoid sheath
Lymph Nodes
Lymph nodes are a relatively new immune innovation in the evolution of verte-
brates. They are found only in mammals and some birds [e.g., ducks (Anatidae)],
though reptiles maintain clusters of lymphoid nodules that bear some resemblance
to lymph nodes (Lawn and Rose 1981; Sugimura et al. 1977; Zapata and Amemiya
2000). The canonical lymph node consists of encapsulated lymphoid lobules, ves-
sels, and sinuses and acts as a local station of leukocyte responses to infection.
Lymph nodes mainly function to filter local lymph of antigens, act as sites for acti-
vated antigen-presenting cells to present antigens to naïve lymphocytes, and to host
and organize the clonal expansion of reactive lymphocytes (Kaldjian et al. 2001;
Karrer et al. 1997).
Lymph node number and location is highly variable between species. Mice, for
example, have 22 identified lymph nodes, while primates are prolific developers of
lymph nodes and maintain hundreds of these structures (e.g., humans have ~450
lymph nodes) (Van den Broeck et al. 2006; Willard-Mack 2006). Moreover, persis-
tent inflammation can lead to de novo synthesis of lymph nodes and contribute to
variation between individuals (Drayton et al. 2006).
Mammals also vary in their expression of specific types of lymph nodes. Rodents,
for example, do not exhibit tonsils (nasopharyngeal lymph nodes), unlike many
other mammals. It has been suggested that the extensive nasal-associated lymphoid
tissue (NALT) found in rodents is analogous to tonsils (Heritage et al. 1997). Many
nonhuman primates, however, maintain more extensive NALT than rodents and
maintain tonsils (Haley 2003). It is tempting to assume that many rodents, with their
heads a few millimeters from the ground, may have lost nasopharyngeal lymph
nodes due to the selection pressure of complications arising from increased lymph
node infections. However, the current information on the importance of tonsils in
Vertebrate Immune System Evolution and Comparative Primate Immunity 29
other species is sufficiently murky and incomplete that it does not support this
notion. The effect of pharyngeal tonsillectomy in both child and adult humans with
persistent pharyngeal infections has not been systematically studied and reviews of
case studies have argued that the procedure does not affect the occurrence of these
infections or lead to immune dysfunction (Blakley and Magit 2009; Burton and
Glasziou 2009; Burton et al. 2000).
The absence of tonsils in rodents does, however, suggest a significant alteration
to the embryonic structure from which they originate, the 2nd pharyngeal pouch
(Grevellec and Tucker 2010; Heritage et al. 1997). Similarly, second pouch devel-
opment has undergone modification in different primate species. Cesta (2006) has
noted that humans have four sets of tonsils (lingual, palatine, pharyngeal, and tubal),
while only three sets of tonsils have been described for other primates (lingual, pala-
tine, and pharyngeal). The acquisition of tubal tonsils in humans suggests a depar-
ture in second pouch development from that other primate species in the last 5–6
million years. The functional ramifications of the acquisition of these tonsils, how-
ever, are not known.
Monocytes/Macrophages
Phagocytic immune cells form a critical link between the mammalian innate and
adaptive immune systems by engulfing, processing, and presenting fragments of
pathogens to T cells and subsequently initiating adaptive responses. The phagocytes
include monocytes/macrophages, dendritic cells, and neutrophils (both a phagocyte
and granulocyte). Phagocytosis is likely the oldest technique of sequestering and
destroying foreign material in the interest of host defense. Indeed, phagocyte-like
cells have been identified in many lifeforms, suggesting that the phagocytes may
originate at the base of the Eukaryota domain and represent the earliest immune
cells in multicellular life (Waddell and Duffy 1986). Social acrasid amoebae, for
example, are often referenced as sharing a last common ancestor with all organisms
that either participate in or maintain cellular components that identify nonself and
phagocytose for nutrition or defense (Dzik 2010). Similarities in cellular organiza-
tion, phagocytosis, and the migration of unicellular amoebic organisms, such as
Acanthamoeba and human macrophages, have also been noted (Anderson et al.
2005). Several genera of amoebae and human macrophages share subsets of cell
receptors used for nonself detection and phagocytosis, experience manipulation of
similar cellular cascades by the same intracellular pathogens, and share a similar
mode of shifting ectoplasm and hyaloplasm to “crawl” (Al-Khodor et al. 2008;
Allen and Dawidowicz 1990; Yan et al. 2004). Particularly compelling evidence
that the origins of primate phagocytes might be rooted in single-celled eukaryotic
30 J.F. Brinkworth and M. Thorn
completely absent from this region in healthy birds and must be recruited from
circulation or tissues (Conrad 1981; Rose and Hesketh 1974; Sabet et al. 1977).
Baseline comparative studies of primate monocytes/macrophages and examina-
tions of stimulated monocytes suggest the monocyte function of closely related pri-
mates differs considerably. An examination of baseline sialic acid-recognizing
Ig-superfamily lectins (siglec) – immune cell surface molecules that can downregu-
late immune cell responses – expression across human, rhesus macaque (Macaca
mulatta), and sooty mangabey (Cercocebus atys) monocytes found marked inter-
species differences in both type and level of siglec constitutively expressed
(Jaroenpool et al. 2007). The same study also found that monocytes from SIV+
rhesus macaques more strongly expressed two siglecs (1 and 7) compared to mono-
cytes from SIV+ sooty mangabeys. The authors suggest that these differences may
contribute to the strong immune activation and pathology typically seen in SIV+
macaques and typically absent in SIV+ mangabeys. They also suggest lineage-
specific adaptation to IVs in sooty mangabeys since they shared a last common
ancestor with rhesus macaques approximately 9 million years ago (Fabre et al.
2009). Barreiro et al (2010) found that human, chimpanzee (Pan troglodytes), and
rhesus macaque monocytes stimulated with lipopolysaccharide (LPS) initiated
species-specific transcription of genes associated with immune responses to viral
infection and cancer. Specifically, human and chimpanzee responses have diverged
from one another. Human responses were enriched for genes associated with
apoptosis and cancer pathology, while the responses of chimpanzees, a natural
immunodeficiency virus host thought to also have a low incidence of cancer, were
enriched for HIV-interacting genes (Barreiro et al. 2010). Such interspecies differ-
ences in monocyte/macrophage responses suggest lineage-specific adaptation has
occurred in catarrhines since cercopithecoids and hominoids last shared a common
ancestor ~30 million years ago (Steiper and Young 2006).
Dendritic Cells
Like all phagocytes, dendritic cells (DCs) are ancient in origin, though likely
younger than macrophages. DCs maintain a tree-like shape, with multiple dendrites
as “branches.” They tend to be found in tissues that have a direct interface with the
external environment (e.g., mucosa) and lymphoid organs, where they participate in
pathogen detection and antigen presentation. Canonical dendritic cells first emerged
in teleost fish, which suggests that DCs have existed since at least the Devonian
period ~420–360 million years ago (Setiamarga et al. 2009; Wolfle et al. 2009).
Mixed leukocyte reactions in jawless fish such as hagfish and lampreys, however,
suggest that these animals maintain similar antigen-presenting cells that recognize
nonself, introducing the possibility of a much earlier emergence date for a DC pre-
cursor (Raison et al. 1987). Once emerged DC morphology and function has
remained conserved in subsequent animal classes. Mammalian and teleost DCs, for
example, are derived from similar precursor cells, maintain dendrites, express
32 J.F. Brinkworth and M. Thorn
similar cell markers, are activated by TLR ligands, participate in phagocytosis, and
are found in similar lymphoid organs (e.g., the spleen) (Bassity and Clark 2012).
In primates, DCs are a heterogeneous group of sentinel cells that arise from a
haematopoietic or peripheral blood mononucleocyte progenitor, are present through-
out body tissues and organs (skin, intestines, liver, lungs, etc.), survey for patho-
gens, and engage bacteria and viruses through a variety of receptors (reviewed in
Wu and Liu 2007). Unlike other phagocytes, however, mammalian DCs are not
directly involved in early clearance of pathogens. While the primary function of
monocytes/macrophages is to destroy nonself, the primary function of DCs appears
to be antigen presentation via MHC I and MHC II molecules. Through antigen pre-
sentation DCs are exceptional stimulators of T-cell responses (reviewed in
Nussenzweig et al. 1980; Savina and Amigorena 2007). DCs mature through this
process of antigen presentation, switching from a phagocytosing cell to an antigen-
presenting cell that migrates to secondary lymphoid tissues, presents MHC bound
antigen to T cells, and initiates T-cell responses through a two-signal system
(reviewed in Gilliet et al. 2008; Kushwah and Hu 2011).
In primates DCs represent a small proportion of leukocytes (e.g., less than 1 % of
circulating leukocytes), so most comparisons of primate DC phenotype and function
are comparisons of DCs cultured from monocytes, bone marrow, or peripheral blood
stem cells (Ashton-Chess and Blancho 2005; Gabriela et al. 2005; Ohta et al. 2008;
Prasad et al. 2010; Soderlund et al. 2000). Such in vitro culture methods can differ
across studies and species, making interspecies comparisons of these DCs difficult.
Interspecies comparisons of DC phenotypes are also frustrated by inconsistent anti-
body affinity for DC markers across species. In vivo examinations of DCs have found
that nonhuman catarrhine species maintain the two major subsets of DCs found in
humans—the monocyte-like myeloid DCs (mDCs) and plasma cell-like plasmacytoid
(pDCs) (Coates et al. 2003; Diop et al. 2008; MacDonald et al. 2002; Malleret et al.
2008; Pichyangkul et al. 2001). However, markers for these cells differ between pri-
mate species. Rhesus macaque monocyte-derived DC (moDCs) markers CD11b,
CD16, and CD56 are expressed on human monocytes and natural killer cells, rather
than DCs (Brown and Barratt-Boyes 2009; Carter et al. 1999). Such differences in
protein expression can frustrate interspecies analyses by antibody-based flow cytom-
etry and may have consequences for immune function. Moreover, antibody affinity
for human DC markers can be depressed in other primates, making the identification
of pDCs particularly difficult for more distantly related primates such as platyrrhines
(specific antibodies discussed in Jesudason et al. 2012).
There are other notable interspecies differences in catarrhine DC phenotype that
suggest the function of these cells have diverged multiple times since cercopithecoids
and hominoids last shared a common ancestor ~30 million years ago. Human, African
green monkey (Chlorocebus sp.), and cynomolgus macaque moDCs strongly express
DC-SIGN/CD209, a receptor involved in pathogen recognition, cell migration, and
T-cell activation (Geijtenbeek et al. 2000a, b, Geijtenbeek et al. 2003). The weak
expression of this receptor on rhesus macaque moDCs suggests that these activities
are mediated differently between catarrhine species (Wu et al. 2002). Information on
primate DC baseline function is very limited, with most data on interspecies
Vertebrate Immune System Evolution and Comparative Primate Immunity 33
differences in function having been collected via a challenge model. An in vitro com-
parison of moDC responses to TLR3 ligand and common vaccine adjuvant Poly I:C,
for example, found that despite similar TLR expression levels, human moDCs initiate
stronger anti-inflammatory and antiviral responses (IL-12, IFNα) to the ligand than
rhesus macaque moDCs (Ketloy et al. 2008). As Poly I:C is a common adjuvant, these
differences may have implications for vaccine studies using rhesus macaques as a
model. During in vivo infection catarrhine DCs exhibit notable differences in func-
tion. Sooty mangabeys, in particular, exhibit signs of lineage-specific adaptation of
pDC function to viral infection. Unlike the pDCs of naïve hosts, such as macaques
and humans, pDCs from natural immunodeficiency virus (IV) host sooty mangabey
infected with IVs fail to express CCR7, an important receptor for cell homing to lym-
phoid tissue. Sooty mangabey pDCs stimulated with ligands that are viral mimetics do
not migrate to the lymph nodes (Mandl et al. 2008). Given that pDCs are implicated
in the dissemination of IVs in lymph nodes, it is possible that the absence of
IV-mediated homing of pDCs to the lymph nodes in sooty mangabeys partially
explains AIDS resistance in this species. Similarly, sooty mangabey pDCs have been
found to produce less IFNα in response to yellow fever virus 17D vaccine than pDCs
isolated from rhesus macaques and humans, suggesting that antiviral responses of
pDCs differ between catarrhine species (Mandl et al. 2011).
Neutrophils
Neutrophils are the most recently emerged phagocyte, first appearing in mammals.
Neutrophils are both phagocytes and granulocytes. Granulocytes are immune cells
that maintain intracellular/membrane-bound vesicles containing biologically active
compounds that can be quickly released in response to stimulus. While neutrophils
are unique to mammals, birds, and reptiles, amphibians and fish maintain a related
and considerably more granular phagocytic cell known as a heterophil. Both cell
types engulf and destroy microbes, maintain granules that contain antimicrobial
substances, circulate in the blood, and initiate tissue repair (Benoit et al. 2008;
Robert and Ohta 2009) (reviewed in Harmon 1998; Yoder 2004). Neutrophils/het-
erophils differ considerably from other phagocytic cells in a number of functionally
important ways. Unlike macrophages and dendritic cells, neutrophils/heterophils
emerge from bone marrow myeloblasts terminally differentiated and do not differ-
entiate further when they phagocytose material or migrate to and from tissues.
While macrophages and dendritic cells tend to traverse tissues in low numbers,
neutrophils/heterophils circulate in blood and cross into tissues in very large num-
bers. Neutrophils/heterophils are also some of the shortest lived immune cells,
maintaining lifespans of hours or a few days, rather than weeks (reviewed in Amulic
et al. 2012).
The primary function of neutrophils appears to be to limit bacterial infections
(Borregaard and Cowland 1997). They are often the first innate immune cells at the site
of an infection, migrating in large numbers to sites of inflammation in the tissues by
34 J.F. Brinkworth and M. Thorn
rolling along endothelium until they are arrested and trafficked to these sites by
expressed chemokines and selectins (Ley et al. 1998; Smith et al. 2004; Zhang et al.
2001). There they amplify inflammation through the release of cytokines and che-
moattractants and are implicated in overt immune responses that lead to tissue
injury or systemic inflammation (Ear and McDonald 2008; Kobayashi 2008;
Zemans et al. 2009). It is perhaps due to the potential of these cells to cause tissue
injury that neutrophils have evolved several mechanisms that limit their antimicro-
bial activities. They undergo apoptosis after phagocytosis and can initiate suicidal
pathogen trapping. They also participate in a unique feedback mechanism that con-
trols exuberant neutrophils. Mass migration of neutrophils to a site is followed by
the recruitment of monocytes (Scapini et al. 2001; Soehnlein et al. 2008). These
monocytes differentiate into macrophages, which then limit neutrophil activities
through lipoxin-triggered phagocytosis (Godson et al. 2000).
Neutrophils participate in several antimicrobial activities that the other phago-
cytes do not. While phagocytosis in macrophages occurs through an endocytic path-
way, neutrophil phagosomes become active upon granule–phagosome fusion
(reviewed in Amulic et al. 2012; Lee et al. 2003; Savina and Amigorena 2007).
Subsequent degranulation serves as an antimicrobial mechanism, nonspecifically
destroying both pathogens and host tissue as well as enhancing phagocytosis by
other phagocytes (Soehnlein et al. 2009). Neutrophils are known to swarm lymph
nodes during parasitic infection, an immune strategy that other phagocytes do not
appear to use (Chtanova et al. 2008). In an interesting potential adaptation to large
pathogens (e.g., helminths), neutrophils participate in a kind of antimicrobial “hara-
kiri” by releasing decondensed chromatin into the environment and creating extra-
cellular traps (NETS) (Brinkmann et al. 2004; Fuchs et al. 2007). This action
inevitably kills the neutrophil, but nets and kills the pathogens as well.
Neutrophils and heterophil granular contents differ, with neutrophils relying
more heavily on an oxidative burst response during phagocytosis (Penniall and
Spitznagel 1975; Stabler et al. 1994). Heterophils and neutrophils differ in how they
contribute to lesion pathology as well. In birds and reptiles, heterophils that invade
a lesion tend to form granulomas, while in mammals neutrophils may form granu-
lomas but can also liquefy and form abscesses (Harmon 1998; Montali 1988). The
number of circulating neutrophils/heterophils differ strongly between animal classes
and orders as well, with neutrophils representing between 50 and 70 % of circulat-
ing blood leukocytes in primates, while they represent only 15–20 % of the circulat-
ing population in rodents and zebrafish (Haley 2003; Hawkey 1985; Martin and
Renshaw 2009).
Interspecies comparisons of neutrophil phenotypes are difficult to complete for
many reasons. In vitro comparisons are hard to undertake as the cells are very dif-
ficult to immortalize and primary neutrophils have a very short life span (Amulic
et al. 2012). In vivo interspecies comparisons of neutrophil activities are made dif-
ficult by how widely the proportions of neutrophils to other circulating leukocytes
differ between mammalian species. It is apparent that neutrophils phenotypically
differ within the mammalian order. Mouse neutrophils, for example, lack defensins,
while human neutrophils maintain these antimicrobial proteins (Eisenhauer and
Vertebrate Immune System Evolution and Comparative Primate Immunity 35
Lehrer 1992). Human neutrophils tend to maintain large doses of histamine, a pro-
inflammatory amine that increases blood vessel permeability, triggers cytokine
release by neighboring leukocytes, triggers smooth muscle constriction, and is
active in allergic reactions. Rabbits and guinea pigs tend to maintain lower quanti-
ties of this molecule (Haley 2003).
Even within the context of specific infections, comparative functional studies
of neutrophils in primates are rare. Through an SIV challenge model Elbim et al
(2008) found some evidence that neutrophil function differs between primate
species. Rhesus macaques exhibit increased expression of an SIV co-receptor on
neutrophils, as compared to African green monkeys. This difference in expres-
sion is correlated with increased neutrophil death in SIV+ rhesus macaques dur-
ing the early stages of infection, suggesting that neutrophil function has diverged
since these catarrhine lineages last shared a common ancestor ~9 million years
ago (Elbim et al. 2008; Fabre et al. 2009; Steiper and Young 2006). More broadly,
primates exhibit interspecies differences in the proportions of circulating neutro-
phils. Neutrophils represent 50–70 % of circulating leukocytes in humans and
other apes and 10–42 % in cercopithecoid primates (Doeing et al. 2003; Haley
2003). An explanation for the intensely neutrophil-rich blood of hominoids is
difficult to determine. A number of primate traits appear to be correlated with
increased proportions of basal neutrophils, including traits that may increase
exposure to bacterial pathogens such as greater terrestriality/body mass and
female mating promiscuity (Nunn 2002). However, there are few comparative
interspecies studies of primate neutrophil function to supplement these findings.
As such, the functional ramifications of these differences in neutrophil propor-
tions are not known.
Mast Cells
A mast cell progenitor, sharing qualities with basophils, appears to have emerged in
tunicata during the Paleozoic (de Barros et al. 2007). Mast cells have subsequently
been found in all vertebrate chordates. Mast cells are innate immune effector cells
that also modulate the activities of innate and adaptive immune cells. They are a
heterogeneous population of bone marrow-derived granulocyte cells that store bio-
logically active compounds in membrane-bound granules. Mast cell precursors
emerge from bone marrow and differentiate into mast subsets when they migrate to
mucosal and connective tissues (Galli et al. 2008; Galli et al. 2005). At these sites,
mast cells often directly interface with the environment and can live for weeks or
months (Church and Levi-Schaffer 1997). Similar to other granulocytes, mast cells
release various granular proteins upon activation, as well as proinflammatory mol-
ecules such as prostaglandins and leukotrienes, cytokines, and chemokines. The
original function of mast cells was likely inflammation and host defense against
bacteria and parasites, though they now also play a very important role in immune
regulation, tissue repair, and blood vessel development (Malaviya et al. 1996)
(reviewed in Crivellato and Ribatti 2010). In mammals, mast cell cytokine produc-
tion can also induce a subset of dendritic cells called Langerhans cells to migrate to
the local lymph nodes, where antigen presentation takes place (Jawdat et al. 2004).
Intriguingly, mast cells are major mediators of detrimental hypersensitivity reac-
tions such as anaphylaxis and yet are maintained in all vertebrate chordates
(Krishnaswamy et al. 2001). Their persistence, despite their contribution to sudden
and fatal immune reactions suggests that mast cells must also play a very beneficial
role in immune protection.
Mast cell size, granule contents, and granule numbers vary considerably between
vertebrate classes and species (Dvorak 2005). A sharp divide in histamine content in
the mast cells of warm-blooded and some cold-blood vertebrates has been well noted
(Reite 1965; Takaya 1969; Takaya et al. 1967). All descendants of early reptiles store
large amounts of histamine in mast cells. Mammalian and avian mast cells are particu-
larly prolific producers of histamine, while histamine is either absent or present in
very low levels in amphibians and most fish. The stark contrast in histamine produc-
tion between animals that emerged before and after reptiles has led to the conclusion
that mast cell granule storage of histamine emerged in the last common ancestor of
extant reptilia, possibly as early as the Permian period (300–250 mya) (Chieffi Baccari
et al. 1998; Mulero et al. 2007; Shedlock and Edwards 2009). In fact, many fish seem
unable to respond to injections of histamine, suggesting that a histamine receptor sys-
tem is absent in much of this animal class (Mulero et al. 2007; Reite 1972). Mulero
et al. (2007) have concluded that the presence of histamine in the mast cells of perci-
form fish (e.g., trout), the largest order of teleost fish, suggests mast cell storage of
histamine evolved twice in vertebrate history (Mulero et al. 2007). This view has been
contested, as some amphibians and earlier emerged classes such as ascidians (e.g., sea
squirts) have been found to maintain low levels of histamine in mast/immune cell
granules (reviewed in Crivellato and Ribatti 2010).
Vertebrate Immune System Evolution and Comparative Primate Immunity 37
Mammalian mast cells have acquired at least one trait that sets them apart from
the mast cells of other vertebrate classes. They uniquely become activated in
response to immunoglobulin E (IgE) (reviewed in Crivellato and Ribatti 2010). This
reaction appears to be very important for defense against helminth parasites. Mice
bearing an IgE deletion, for example, experience increased Schistosoma sp. worm
loads and inflammation (King et al. 1997b). It was long thought that IgE and its
receptors on mast cells represented a system of mast cell activation that newly
emerged in mammals, but this view has recently been challenged by detection of an
IgE receptor-like protein on zebrafish mast cells (Da’as et al. 2011). While the roots
of an IgE receptor-based mast cell activation system may be considerably older than
previously thought, it appears mammals alone use IgE in antiparasitic strategies.
As with many leukocyte subtypes, primate mast cells have not been well investi-
gated for baseline functional differences. Some processes, such as the cleavage of
particular granule contents (e.g., chymases) appear to be conserved in cercopithe-
coids and hominoids (Thorpe et al. 2012). There are, however, indications that the
granular contents of primate mast cells differ between species. Tryptases, for exam-
ple, are serine peptidases stored in mast cell granules and implicated in pathological
immune conditions such as anaphylaxis and asthma (Clark et al. 1995; Schwartz
et al. 1987). δ-trypase (TPSD1) is functional and active in cercopithecoids but has
undergone several nonsense mutations and truncations in hominoids over the last 30
million years, and the human–chimpanzee lineage over the last 6 million years. As
a result δ-trypase is almost nonfunctional in humans and chimpanzees (Trivedi et al.
2008). Rather β-tryptases appear to be the more active tryptase in humans and
chimpanzees, suggesting that cercopithecoid models of mast cell disorders such as
allergy and anaphylaxis may not accurately reflect human tryptase activity (Trivedi
et al. 2008, 2007).
Basophils
Basophils are exceptionally rare in most vertebrate species, with only turtles and
rabbits reported to maintain comparatively high numbers in circulation (Canfield
1998; Haley 2003). The very low circulating numbers of basophils in vertebrates
has been a significant hindrance to the comparative functional study of this cell
type. Comparative studies of primate basophil function have not been completed.
Eosinophils
of very quickly and extensively degranulizing in response to both specific and nonspe-
cific stimulation, whereas direct evidence that the eosinophils of other mammalian
species, such as mice, can do the same is very limited (Lee et al. 2012). If tissue
remodeling is a primary function of eosinophils, interspecies differences in degranula-
tion would be expected, as tissue remodeling mechanisms would likely differ between
species (Lee and Lee 2005).
Eosinophil granule content also differs markedly between mammalian species,
which may reflect differences in cell function. While many primates appear to main-
tain two types of ribonucleases of varying functions in their eosinophil granules,
mice maintain six, all of which are paralogous to primate ribonucleases and exhibit
strong ribonuclease activity (Lee et al. 2012). Eosinophil granule content differs
strongly between major primate clades, suggesting lineage-specific evolution of
eosinophil function. Catarrhines, for example, store ribonucleases ribonuclease A
family 3 (RNASE3) and ribonuclease A family 2 (RNASE2) at high levels in their
granules, while Platyrrhines only store a ribonuclease homologous to RNASE3
(Olsson et al. 1977; Rosenberg and Dyer 1995; Rosenberg et al. 1995). RNASE3 is
particularly toxic to host tissue cells. These differences in ribonuclease production
and storage suggest that ribonuclease-mediated eosinophil apoptotic, cytotoxic, and
antimicrobial activities have changed in these clades since they last shared a com-
mon ancestor 40 million years ago (Fabre et al. 2009; Steiper and Young 2006).
Natural Killer (NK) cells are granular lymphocytes. NK cells recognize “altered”
self (e.g., tumours) and nonself and have acquired their name because of their abil-
ity to destroy cells spontaneously without the need to be primed as other lympho-
cytes require (Herberman et al. 1975). Through direct contact mammalian NK cells
can target and lyse tumour cells or virally infected cells that downregulate MHC I
expression as a result of oncogenesis or infection (reviewed in Purdy and Campbell
2009). MHC I class molecules found on other immune cells are NK ligands and
interactions between MHCs and NK killer cell immunoglobulin-like receptors
(KIR) control NK effector functions (reviewed in Lanier 2008). Target cell killing
is initiated when activated NK cells release perforin and granzyme that induce
apoptosis (Lieberman 2003). For their importance, NKs are a rare cell population.
In humans, NK cells represent 10–15 % of blood lymphocytes, or approximately
2–3 % of all circulating leukocytes (Purdy and Campbell 2009).
Cells exhibiting NK-like activities, such as direct killing of foreign materials via
perforin-like proteins, have been noted in invertebrate life (Kauschke et al. 2001).
NK-like cells are present in all tunicata, suggesting that this cell type was established by
the time this clade emerged between 520 and 794 mya (Blair and Hedges 2005; Chen
et al. 2000). As the earliest evidence of MHC molecules is found in cartilaginous fish, it
seems that NK adoption of MHC class I molecules as ligands occurred long after the
40 J.F. Brinkworth and M. Thorn
emergence of NK cells (Azumi et al. 2003; Khalturin et al. 2003). Thus, the origins of
NK lymphocytes appear to predate the adaptive immune system as a whole.
The role of NK-like cells in nonmammalian classes is not well investigated but
appears to be conserved. In tunicata NK-like cells appear to mainly participate in
allorecognition and killing nonself materials via granule release of perforin and
granzyme-like proteins (Bielek 1988; Shen et al. 2004). NK-like cells with cyto-
toxic, and allogeneic and tumor-killing capabilities have been found in amphibians
and avian species as well (Goyos and Robert 2009; Horton et al. 1996; Jansen et al.
2010; Wainberg et al. 1983). While interclass comparisons of NK-cell distribution
and characteristics have led to conclusions that mammalian but not avian NK cells
can circulate in blood, it is very likely that a dearth of cross-reactive NK marker
antibodies for nonmammals has frustrated these studies (Rogers et al. 2008b).
While mammalian NK cells are the focus of intense study, the comprehensive set of
markers that can reliably identify NK cells in all mammalian species has yet to be
determined (Walzer et al. 2007). Even very closely related species such as humans
and rhesus macaques exhibit sharp differences in NK marker expression. Human
NKs strongly express CD56, a marker mainly expressed by monocytes in rhesus
macaques (Webster and Johnson 2005). The primate species for which markers are
well defined tend to be those species important to HIV research (e.g., rhesus
macaques and sooty mangabeys). Interspecies comparisons of primate NK function
also tend to be limited to these two species. Rhesus macaque and sooty mangabey
NK-cell functional comparisons do indicate the evolution of baseline interspecies
differences since these species last shared a common ancestor ~9 million years ago
(Fabre et al. 2009). The NK cells of healthy sooty mangabeys appear to be twice as
cytotoxic as rhesus macaque NK cells when activated, suggesting that sooty mang-
abeys may be able to more quickly mount a strong NK-based response than rhesus
macaques when viral infection occurs (Pereira and Ansari 2009).
An interesting feature of primate NK cells is the acquisition and rapid evolution
of primate killer-cell immunoglobulin-like receptors (KIRs). Primate NK cells
alone express KIRs. Like KIRs, the NK-cell receptors of other mammals [e.g., killer
cell lectin-like receptor, subfamily A, member 2 (KLRA2/Ly49) lectin receptors of
rodents] bind with MHC class I molecules or their homologues and activate the NK
cell (Abi-Rached and Parham 2005; Barten et al. 2001; Colucci et al. 2002; Lanier
2008). The KIR family of receptors appears to have expanded from a single gene,
killer cell immunoglobulin-like receptor three domain long cytoplasmic tail
(KIR3DL), over the last ~50 million years. In that time, KIR genotypes have rapidly
diversified between and within haplorrhine species. Humans, for example, encode
over 130 genotypes for approximately 14 KIRs, while the grey mouse lemur encodes
only one functional KIR (Averdam et al. 2009; Bimber et al. 2008; Hollenbach et al.
2010). Human KIRs exhibit lineage-specific evolution of receptors that specifically
recognize epitopes of MHC-B and MHC-C receptors, as well as an increased num-
ber of KIR “activating receptors” as compared to chimpanzees (Abi-Rached et al.
2010). The divergence of haplorrhine and strepsirrhine NK receptors and subse-
quent lineage-specific diversification of these receptors have likely affected NK
interactions with similarly rapidly evolving MHC molecules. Strepsirrhines appear
Vertebrate Immune System Evolution and Comparative Primate Immunity 41
immune system evolution. VLRS are assembled and rearranged through the activi-
ties of APOBEC-like cytidine deaminases (AID) (Rogozin et al. 2007). This system
of expanding receptor repertoires through gene rearrangement/somatic recombina-
tion bears a strong resemblance to RAG-mediated TCR and immunoglobulin rear-
rangement. It is not, however, a homologous system and appears to be outcome of
convergent evolution. VLRs are not rearranged by RAG and they are not related to
the Ig receptor superfamily to which TCRs, BCRs, and MHC molecules belong
(Rogozin et al. 2007). Canonical T and B cells emerged with the last common
ancestor of jawed vertebrates (Flajnik and Kasahara 2001). The adaptive immune
systems of jawed and jawless vertebrates bear some remarkable similarities for
potentially being the products of convergent evolution. VLR expressing lympho-
cytes, for example, can be divided into three functional categories (VLRA+,
VLRB+, and VLRC+ protolymphocytes) that maintain similar gene expression
profiles and respond to antigens in a manner similar to that of vertebrate T and B
cells (Guo et al. 2009).
B cells
B cells represent the arm of the adaptive immune system responsible for the produc-
tion of soluble antibodies, or immunoglobulins (Ig), which recognize and bind epi-
topes expressed on a variety of microbial pathogens. In mammals, naïve B cells exit
the bone marrow and migrate to the secondary lymphoid organs where an encounter
with an antigen or antigen-specific T helper cells could lead to B-cell activation,
proliferation, Ig isotype class switching, and soluble antibody secretion. Canonical
B cells first emerged with the TCR/BCR/MHC adaptive immune components in
jawed vertebrates, 652–525 mya (Boehm and Bleul 2007). However, it seems likely
that B cells emerged earlier than T cells, as T cells require the additional step of
NOTCH signaling to differentiate from the lymphocyte progenitor cell and B cells
do not (Benne et al. 2009). Of the two cell types, B cells have not been as intensely
studied in primates as T cells. However, baseline interspecies differences in B-cell
populations have been noted between primates. For example, cynomolgus macaques
(M. fascicularis) maintain different proportions of B cell subsets than humans,
which may translate as interspecies differences in immune response. Specifically,
cynomolgus macaques retain higher percentage of B cells carrying the T cell
costimulatory molecule CD80 and a lower percentage of B cells carrying CD21, a
molecule that interacts with the complement system, than humans (Vugmeyster
et al. 2004).
Much of the comparative research on baseline B cell function has focused on
immunoglobulins (antibodies). Immunoglobulins may be secreted by B cells or
associate with accessory molecules to form the membrane-bound B cell receptor
complex (BCR). Secreted Igs can recognize and neutralize an antigen, while the
BCR complex can recognize an antigen and initiate cell signaling and/or internal-
ization and presentation of antigen to T cells. The genes that encode Igs in B cells
Vertebrate Immune System Evolution and Comparative Primate Immunity 43
and TCRs in T cells are arranged in a similar way: multiple variable and constant
gene segments that recombine and assemble in a largely random manner to generate
a vast and almost limitless repertoire of antigenic specificities. The Ig molecule is
made up of two identical Ig heavy and light chains, both of which contain variable
regions. Multiple variable (V), diversity (D), and junction (J) gene segments are
rearranged by V(D)J recombination to form an Ig heavy chain, while the light
chains are formed from V and D gene segments. The diversity of the various gene
segments is further amplified by the random nucleotide additions and subtraction at
the ends of Ig gene segments before they are joined (Thai et al. 2002). This system
of immunoglobulin diversification appeared by the emergence of cartilaginous fish,
652–525 million years ago (Blair and Hedges 2005; Rast et al. 1997). It is estimated
that the human B cell repertoire is capable of generating about 1011 unique immu-
noglobulins based on these rearrangements (Schatz et al. 1992).
The immunoglobulins produced by B cells are divided into several classes distin-
guished by constant regions, which impart specific effector properties to each class.
Mammals maintain immunoglobulins belonging to the IgM, IgG, IgA, IgE, and IgD
classes and have lost the IgY found in other animals (Kapetanovic and Cavaillon
2007; Mussmann et al. 1996b; Vollmers and Brandlein 2006; Zhao et al. 2006). Of
the immunoglobulins found in primates, IgM is oldest, having emerged with the last
common ancestor to all jawed vertebrates. It is highly conserved in function and
structure. All jawed vertebrates appear to produce the IgM class of secreted immu-
noglobulin ahead of all others in response to immune system stimulation (reviewed
in Flajnik 2002). Compared to other immunoglobulins, IgM has a lower affinity for
antigens, is often polyreactive, and has the ability to bind a variety of microbial and
viral antigens (Boes 2000).
Like IgM, an IgD homolog (e.g., IgW) appears to have been present in the last
common ancestor of jawed vertebrates (Ohta and Flajnik 2006). The evolution of
IgD is enigmatic, as many different isotypes are found in jawed fish, yet IgD is vari-
ably present in bird and mammalian species (Butler et al. 1996; Zhao et al. 2002).
In primates IgD is expressed on naïve and memory B cells; however, the function of
this immunoglobulin is not well understood (Messaoudi et al. 2011). It seems to be
involved in lymphocyte and granulocyte activities. In its secreted form in bony fish,
for example, it interacts primarily with granulocytes. In humans the membrane-
bound form of IgD may be important in the activation of B cells, while the secreted
form activates basophils (Flajnik and Kasahara 2010).
IgA is found in extant reptiles, which suggests it emerged at least ~320 million
years ago (Mussmann et al. 1996a). In mammals, IgA is commonly secreted in
mucosal sites by specialized B cells, where it binds and shields bacteria to maintain
gut homeostasis (Strugnell and Wijburg 2010). It is the most commonly secreted
immunoglobulin isotype in mammals, frequently found in GALT structures, where
its production is regulated by innate and T cell pathways in germinal centers (Suzuki
et al. 2010). IgA prototypes and analogs fulfill similar functions in nonmammalian
species. In teleost fish, for example, Ig-based mucosal immunity is accomplished
through IgT, an IgA analog (Zhang et al. 2010). Amphibian mucosal Ig immunity is
fulfilled by IgX, a possible prototype of IgA (Deza et al. 2007). IgA is also an
44 J.F. Brinkworth and M. Thorn
important neutralizer of bacteria, toxins, and viruses at the mucosal barrier and may
also facilitate antigen take up by the mucosal dendritic cells (Corthesy 2007; Stubbe
et al. 2000). When IgA does not regulate gut microflora appropriately, inflammatory
responses become dysregulated. Dysfunctional IgA is implicated in variety of
inflammatory gut disorders (reviewed in Sutherland and Fagarasan 2012).
In primates, IgA function appears to be very conserved. A comparison of IgA-
encoding genes (IGHA and IGCA) and IgA molecules of closely related nonhuman
primates [baboons (Papio sp.), sooty mangabeys, rhesus macaques, and pig-tailed
macaques (M. nemestrina)] found that despite high intraspecies variation in
sequence, the IgA receptor (CD89) of macaques could readily bind with the IgA of
other primates (i.e., humans). (Rogers et al. 2008a). The conservation of gut micro-
flora “enterotypes” across distant human populations has also been interpreted as
conservation of IgA activities (Arumugam et al. 2011).
IgE and IgG diverged from the ancestral IgY in the last common ancestors of
amphibians and mammals, respectively (Flajnik and Kasahara 2010; Warr et al. 1995;
Zhao et al. 2006). IgE is a proinflammatory immunoglobulin and a potent stimulator
of granulocytes. It is associated with Th2 responses to helminthic infections and
allergens, playing a key role in the atopic hypersensitivity response (i.e., allergic
asthma and food allergies) (Broide et al. 2010; Capron et al. 1987; Hagel et al. 2004;
Stone et al. 2010). As such, it has been the subject of intense investigation.
As IgE is only found in mammals, IgE-based mechanisms at play in allergies and
helminthic infections are unique to this class and comparatively recent in the con-
text of vertebrate evolution. The essential nature of IgE, however, has been ques-
tioned as it has been maintained by mammals for over 250 million years but is
frequently suppressed in humans by modern medications without ill effect (Cooper
et al. 2008; Cruz et al. 2007; Vernersson et al. 2004, 2002).
In humans, helminthic infections and IgE-mediated allergic responses have been
assumed to have had a complicated selective affect on the species, with high levels of
IgE considered beneficial against helminthes and yet key to atopic hypersensitivity
(Hagel et al. 2004). Moreover, under the hygiene hypothesis, human IgE responses have
coevolved with certain helminth species which are thought to dampen IgE-mediated
atopic hypersensitivity reactions (see Rook 2013; Martin et al. 2013 for reviews of this
hypothesis). Loss of these species in the modern human microbiota, due to increased
hygiene/use of antihelminthic medications, is thought to contribute to overt IgE
responses to allergens and subsequent hypersensitivity (Cardoso et al. 2012). However,
a recent finding of cross-reactivity between allergenic extracts from the helminth Ascaris
lumbricoides and mite allergens suggests the interplay between IgE responses to hel-
minthes and allergens may be more complicated and sometimes involve cases of “mis-
taken identity” (Acevedo and Caraballo 2011; Acevedo et al. 2009).
In primates, IgE appears to be functionally diverse. Analysis of membrane-
bound IgE gene-encoding regions suggests that tarsiers do not produce membrane-
bound IgE (Wu et al. 2012). Wu et al. (2012) note that mouse strains unable to
produce membrane-bound IgE exhibit very low IgE production during parasitic
infections, which suggests a functional difference between tarsier IgE and the IgE
of other primates (Achatz et al. 2008). While catarrhines produce both long and
Vertebrate Immune System Evolution and Comparative Primate Immunity 45
short isoforms of IgE, platyrrhines do not appear to produce a short isoform, and
lemurs, lorises and nonprimate mammals do not appear to produce a long. This
suggests that the ancestral condition for IgE is the short isoform and that the long
isoform was acquired sometime after the divergence of haplorrhines and strepsir-
rhines ~77 mya, but before the divergence of platyrrhines from catarrhines 43 mya
(Steiper and Young 2006 isoform; Wu et al. 2012).
IgE-mediated atopic hypersensitivity reactions appear to be increasing across the
globe (Rorke and Holgate 2004). Using data collected through the International
Study of Asthma and Allergies in Childhood (ISAAC), Asher et al (2010) have
noted a positive correlation between ecological factors associated with developed
nations, such as gross national product per capita, trans fatty acids and tylenol use,
and the incidence of asthma, eczema, and rhinoconjunctivitis (Asher et al. 2010).
Other studies have noted a positive association between lower socioeconomic status
and asthma/allergic reactions (Hagel et al. 2004; Litonjua et al. 1999; Von Behren
et al. 1999). It is worth noting, however, that an association between these factors
and conditions may be an outcome of increased “hygiene” in developed nations, but
may also be influenced by increased disease surveillance in wealthier countries.
While potential data bias should be considered, these data propose an interesting
connection between human cultural behaviors and human IgE expression.
IgG is the most abundant antibody type found in mammalian serum and has high
affinity and specificity for target antigens. The key function of IgG is to bind to
target antigens and activate effector cells, such as NK cells or monocytes, to destroy
immunoglobulin-coated targets (Schroeder and Cavacini 2010). Complement-
mediated cytotoxicity is another function modulated by IgG that depends on the
C1q complement component binding to the constant portion of IgG antibody bound
to a target (Schroeder and Cavacini 2010). IgG and IgG receptor interactions are
known to differ between primate species, with some cynomolgus macaque IgG sub-
classes exhibiting stronger effector function and greater binding affinity for IgG
receptors, known as Fc receptors, than human IgG (Warncke et al. 2012).
Additionally, the IgG receptor CD16 binds to different IgG subclasses in sooty
mangabeys than it binds in humans (Rogers et al. 2006). These results suggest that
IgG function has diverged in the 25–30 million years since these species shared a
last common ancestor (Fabre et al. 2009; Steiper and Young 2006).
T Cells
T cells, so named because they mature in the thymus, are lymphocytes that contrib-
ute to adaptive immune defense through cytotoxic activities and the regulation of
other immune cells. In jawed vertebrates, T cells are activated indirectly through
antigen presentation by MHC I and MHC II molecules of professional antigen-
presenting cells, such as DCs and macrophages, to the T-cell receptors (TCRs) of T
cells. This activity constitutes a bridge between innate immune responses and adap-
tive immunity, and initiates the adaptive immune response. Both TCRs and MHC
46 J.F. Brinkworth and M. Thorn
molecules are highly variable and capable of binding a wide variety of both host and
pathogen-derived antigenic peptides. In humans, for example, human lymphocyte
antigen (HLA) encodes about 2,000 MHC alleles according to the ImMunoGeneTics
HLA database (http://www.ebi.ac.uk/imgt/hal/atats.html) (MHC/HLA diversity in
Old World primates is discussed in Loisel and Tung 2013). It was previously thought
that TCRs, MHCs, and T and B lymphocytes emerged in the immunological equiva-
lent of a “big bang”, all at once, with the emergence of jawed vertebrates. With the
discovery of VLR+ lymphocytes in jawless fish, it appears that the origins of T-cell
lymphocytes are considerably more ancient than the Ig receptor superfamily
system.
The discovery of specialized lymphocytes in jawless vertebrates significantly
challenged a long-held wisdom that Ig receptors such as T-cell receptors (TCR)
drove the divergence of T and B cells. Specialized VLRs (VLRA, VLRB, and
VLRC) are not related to Ig receptors, but they are found in jawless fish on lympho-
cyte-like cells that are already proficient in different functions reminiscent of T and
B cells. VLRB molecules are mainly secreted by and found in the membrane of
cells whose functions and other receptors best resemble those of B cells (Alder et al.
2008; Herrin et al. 2008). Cells bearing VLRA and VLRC tend to respond to similar
stimuli and produce the same cytokines and receptors as T cells (Guo et al. 2009).
This suggests that lymphocyte subsets became specialized before the “big bang” of
Ig receptors. Moreover, lamprey have a TCR-like gene, along with VLRs, suggest-
ing that the foundations of both systems of somatic rearrangement/recombination
may have existed in some earlier animals (Pancer et al. 2004). As Hsu (2011) has
pointed out, these receptors suggest that rather than lymphocyte differentiation
being based on the inherent functions of the receptor, lymphocyte receptors have
been selected for based on the benefit they brought to lymphocyte function.
Within jawed vertebrates, both TCR gene organization and T-cell development are
highly conserved (Flajnik and Kasahara 2010; Rast et al. 1997). TCR gene segment
rearrangement takes place in the thymus of all jawed vertebrates and produces a
unique antigen-specific TCR for each T cell. Task-specific TCR subtypes known in
mammals (alpha/beta and gamma/delta) have been found in cartilaginous fish, which
suggests that TCRs had already differentiated into subtypes by the the time that jawed
vertebrates emerged (Hirano et al. 2011; Kreslavsky et al. 2010). Despite conserva-
tion, TCRs do differ across vertebrate classes and species. Two unique TCR genes
arose in marsupials and sharks (TCR mu and NAR-TCR) and are expressed as atypical
TCRs in these animals (Criscitiello et al. 2006; Parra et al. 2007).
In mammals, T cells can be divided into task-specific categories, the two largest
subsets being CD8+ cytotoxic T cells and CD4+ helper T cells (Th). Cytotoxic T
cells can directly kill infected cells, while helper T cells differentiate into Th subsets
(e.g., Th1, Th2, and Th17), after interacting with antigen-presenting cells, to regu-
late other T cells, B-cell antibody production, and Ig class switching. It is perhaps
out of a historical emphasis on adaptive immune responses that many subsets of T
cells have been defined and characterized [e.g., Th1, Th2, Th17, regulatory T cells
(Tregs), and additional within group subsets]. When adaptive immune features such
as Th cells or memory T-cell retention emerged in vertebrate immunity is very
Vertebrate Immune System Evolution and Comparative Primate Immunity 47
difficult to determine. In the most catholic interpretation, these traits appear with the
TCR/MHC/lymphocyte system in jawed vertebrates. Whether or not progenitors of
such T-cell subtypes exist in the repertoire of lymphocyte-like cells found in earlier
lifeforms is matter for further investigation.
Almost all comparative studies of primate T cells are based on a challenge model.
Under these experimental circumstances, primate T cells do exhibit unique charac-
teristics. Human T cells have been described as “overreactive” to stimulation in
comparison to chimpanzee T cells, a functional difference that Soto et al. (2010)
attribute to increased levels of inhibitory sialec acid-recognizing Ig-superfamily
lectin 5 (Siglec 5) on chimpanzee T and B cells. Siglec 5 can suppress immune cell
activation (Soto et al. 2010).
As nonhuman primates are important models of immunodeficiency virus infec-
tion and T-cell activities are key to HIV infection progression, most comparative
information on primate T cells has been captured in the context of SIV/HIV research.
As the primary targets of immunodeficiency virus infection, CD4+ T cells have
been particularly well examined for interspecies functional differences. Available
information suggests that lineage-specific adaptations have evolved in this subset of
T cells. The proportions of CD4+ T cells readily infected by HIV/SIV, that is those
CD4+ T cells that also express CCR5, differ between catarrhine species that are
natural and naïve hosts of IVs. Healthy sooty mangabeys, for example, maintain
fewer CD4+ CCR5+ T cells than humans and macaques. Moreover, these cells do
not upregulate CCR5 when stimulated and appear to be less susceptible to IV infec-
tion as a result (Paiardini et al. 2011). This trait may help explain sooty mangabey
resistance to IV pathogenesis. CD4+ T cells also proliferate more quickly in IV
infected humans and macaques than they do in IV natural hosts such as mandrills
(Mandrillus sphinx) and sooty mangabeys, suggesting that the haemopoietic mech-
anisms leading to the production T cells differ between species (Chan et al. 2010;
Engram et al. 2010). Healthy macaques have been found to maintain more CD4+ T
cells that produce a protease, granzyme B, that induces apoptosis in virus-infected
cells (granzyme B) in the lamina propria of the gut than African green monkeys, a
natural IV host (Hutchison et al. 2011). Granzyme B is a highly potent protease that
may contribute to the translocation of microbes from within the gut into the perito-
neal cavity during IV infection in naïve hosts. Taken together, these differences
suggest pathogen-mediated selection of CD4+ traits in natural IV hosts over the 30
million years since all catarrhines last shared a common ancestor.
Conclusions
The major components of the primate immune system represent hundreds of mil-
lions of years of evolution. Here, we have outlined how the vertebrate immune sys-
tem evolved and how primates have become specialized within that system, providing
the first summary of functional differences in primate baseline immunity within the
context of vertebrate immune system evolution. Where possible, we have offered a
48 J.F. Brinkworth and M. Thorn
References
Abi-Rached L, Parham P (2005) Natural selection drives recurrent formation of activating killer
cell immunoglobulin-like receptor and Ly49 from inhibitory homologues. J Exp Med
201(8):1319–1332
Abi-Rached L, Moesta AK, Rajalingam R, Guethlein LA, Parham P (2010) Human-specific evolu-
tion and adaptation led to major qualitative differences in the variable receptors of human and
chimpanzee natural killer cells. PLoS Genet 6(11):e1001192
Acevedo N, Caraballo L (2011) IgE cross-reactivity between Ascaris lumbricoides and mite
allergens: possible influences on allergic sensitization and asthma. Parasite Immunol
33(6):309–321
Acevedo N, Sanchez J, Erler A, Mercado D, Briza P, Kennedy M, Fernandez A, Gutierrez M, Chua
KY, Cheong N et al (2009) IgE cross-reactivity between Ascaris and domestic mite allergens:
the role of tropomyosin and the nematode polyprotein ABA-1. Allergy 64(11):1635–1643
Achatz G, Lamers M, Crameri R (2008) Membrane bound IgE: the key receptor to restrict high IgE
levels. Open Immunol J 1:25–32
Adachi S, Yoshida H, Kataoka H, Nishikawa S (1997) Three distinctive steps in Peyer’s patch
formation of murine embryo. Int Immunol 9(4):507–514
Vertebrate Immune System Evolution and Comparative Primate Immunity 49
Alder MN, Rogozin IB, Iyer LM, Glazko GV, Cooper MD, Pancer Z (2005) Diversity and function
of adaptive immune receptors in a jawless vertebrate. Science 310(5756):1970–1973
Alder MN, Herrin BR, Sadlonova A, Stockard CR, Grizzle WE, Gartland LA, Gartland GL,
Boydston JA, Turnbough CL Jr, Cooper MD (2008) Antibody responses of variable lympho-
cyte receptors in the lamprey. Nat Immunol 9(3):319–327
Al-Khodor S, Price CT, Habyarimana F, Kalia A, Abu Kwaik Y (2008) A Dot/Icm-translocated
ankyrin protein of Legionella pneumophila is required for intracellular proliferation within
human macrophages and protozoa. Mol Microbiol 70(4):908–923
Allen PG, Dawidowicz EA (1990) Phagocytosis in Acanthamoeba: I. A mannose receptor is
responsible for the binding and phagocytosis of yeast. J Cell Physiol 145(3):508–513
Amulic B, Cazalet C, Hayes GL, Metzler KD, Zychlinsky A (2012) Neutrophil function: from
mechanisms to disease. Annu Rev Immunol 30:459–489
Anderson IJ, Watkins RF, Samuelson J, Spencer DF, Majoros WH, Gray MW, Loftus BJ (2005)
Gene discovery in the Acanthamoeba castellanii genome. Protist 156(2):203–214
Arumugam M, Raes J, Pelletier E, Le Paslier D, Yamada T, Mende DR, Fernandes GR, Tap J, Bruls
T, Batto JM et al (2011) Enterotypes of the human gut microbiome. Nature
473(7346):174–180
Asher MI, Stewart AW, Mallol J, Montefort S, Lai CK, Ait-Khaled N, Odhiambo J (2010) Which
population level environmental factors are associated with asthma, rhinoconjunctivitis and
eczema? Review of the ecological analyses of ISAAC Phase One. Respir Res 11:8
Ashton-Chess J, Blancho G (2005) An in vitro evaluation of the potential suitability of peripheral
blood CD14(+) and bone marrow CD34(+)-derived dendritic cells for a tolerance inducing
regimen in the primate. J Immunol Methods 297(1–2):237–252
Averdam A, Petersen B, Rosner C, Neff J, Roos C, Eberle M, Aujard F, Munch C, Schempp W,
Carrington M et al (2009) A novel system of polymorphic and diverse NK cell receptors in
primates. PLoS Genet 5(10):e1000688
Azumi K, De Santis R, De Tomaso A, Rigoutsos I, Yoshizaki F, Pinto MR, Marino R, Shida K,
Ikeda M, Arai M et al (2003) Genomic analysis of immunity in a Urochordate and the emer-
gence of the vertebrate immune system: “waiting for Godot”. Immunogenetics
55(8):570–581
Bajoghli B, Guo P, Aghaallaei N, Hirano M, Strohmeier C, McCurley N, Bockman DE, Schorpp
M, Cooper MD, Boehm T (2011) A thymus candidate in lampreys. Nature 470(7332):90–94
Barreiro LB, Marioni JC, Blekhman R, Stephens M, Gilad Y (2010) Functional comparison of
innate immune signaling pathways in primates. PLoS Genet 6(12):e1001249
Barten R, Torkar M, Haude A, Trowsdale J, Wilson MJ (2001) Divergent and convergent evolution
of NK-cell receptors. Trends Immunol 22(1):52–57
Bassity E, Clark TG (2012) Functional identification of dendritic cells in the teleost model, rain-
bow trout (Oncorhynchus mykiss). PLoS One 7(3):e33196
Benne C, Lelievre JD, Balbo M, Henry A, Sakano S, Levy Y (2009) Notch increases T/NK poten-
tial of human hematopoietic progenitors and inhibits B cell differentiation at a pro-B stage.
Stem Cells 27(7):1676–1685
Benoit M, Desnues B, Mege JL (2008) Macrophage polarization in bacterial infections. J Immunol
181(6):3733–3739
Benveniste RE, Arthur LO, Tsai CC, Sowder R, Copeland TD, Henderson LE, Oroszlan S (1986)
Isolation of a lentivirus from a macaque with lymphoma: comparison with HTLV-III/LAV and
other lentiviruses. J Virol 60(2):483–490
Bernard D, Six A, Rigottier-Gois L, Messiaen S, Chilmonczyk S, Quillet E, Boudinot P,
Benmansour A (2006) Phenotypic and functional similarity of gut intraepithelial and systemic
T cells in a teleost fish. J Immunol 176(7):3942–3949
Bielek E (1988) Ultrastructural analysis of leucocyte interaction with tumour targets in a teleost,
Cyprinus carpio L. Dev Comp Immunol 12(4):809–821
Bimber BN, Moreland AJ, Wiseman RW, Hughes AL, O’Connor DH (2008) Complete character-
ization of killer Ig-like receptor (KIR) haplotypes in Mauritian cynomolgus macaques: novel
50 J.F. Brinkworth and M. Thorn
insights into nonhuman primate KIR gene content and organization. J Immunol 181(9):
6301–6308
Blair JE, Hedges SB (2005) Molecular phylogeny and divergence times of deuterostome animals.
Mol Biol Evol 22(11):2275–2284
Blakley BW, Magit AE (2009) The role of tonsillectomy in reducing recurrent pharyngitis: a sys-
tematic review. Otolaryngol Head Neck Surg 140(3):291–297
Blumbach B, Diehl-Seifert B, Seack J, Steffen R, Müller IM, Müller WEG (1999) Cloning and
expression of novel receptors belonging to the immunoglobulin superfamily from the marine
sponge Geodia cydonium. Immunogenetics 49:751–763.
Boehm T (2011) Design principles of adaptive immune systems. Nat Rev Immunol
11(5):307–317
Boehm T, Bleul CC (2007) The evolutionary history of lymphoid organs. Nat Immunol
8(2):131–135
Boehm T, Hess I, Swann JB (2012a) Evolution of lymphoid tissues. Trends Immunol
33(6):315–321
Boehm T, Iwanami N, Hess I (2012b) Evolution of the immune system in the lower vertebrates.
Annu Rev Genomics Hum Genet 13:127–149
Boes M (2000) Role of natural and immune IgM antibodies in immune responses. Mol Immunol
37(18):1141–1149
Borregaard N, Cowland JB (1997) Granules of the human neutrophilic polymorphonuclear leuko-
cyte. Blood 89(10):3503–3521
Brenchley JM, Douek DC (2008) The mucosal barrier and immune activation in HIV pathogene-
sis. Curr Opin HIV AIDS 3(3):356–361
Brenchley JM, Paiardini M, Knox KS, Asher AI, Cervasi B, Asher TE, Scheinberg P, Price DA,
Hage CA, Kholi LM et al (2008) Differential Th17 CD4 T-cell depletion in pathogenic and
nonpathogenic lentiviral infections. Blood 112(7):2826–2835
Brendolan A, Rosado MM, Carsetti R, Selleri L, Dear TN (2007) Development and function of the
mammalian spleen. Bioessays 29(2):166–177
Brinkmann V, Reichard U, Goosmann C, Fauler B, Uhlemann Y, Weiss DS, Weinrauch Y, Zychlinsky
A (2004) Neutrophil extracellular traps kill bacteria. Science 303(5663):1532–1535
Broide DH, Finkelman F, Bochner BS, Rothenberg ME (2010) Advances in mechanisms of
asthma, allergy, and immunology in 2010. J Allergy Clin Immunol 127(3):689–695
Brown KN, Barratt-Boyes SM (2009) Surface phenotype and rapid quantification of blood den-
dritic cell subsets in the rhesus macaque. J Med Primatol 38(4):272–278
Burton MJ, Glasziou PP (2009) Tonsillectomy or adeno-tonsillectomy versus non-surgical treat-
ment for chronic/recurrent acute tonsillitis. Cochrane Database Syst Rev (1):CD001802
Burton MJ, Towler B, Glasziou P (2000) Tonsillectomy versus non-surgical treatment for chronic/
recurrent acute tonsillitis. Cochrane Database Syst Rev (2):CD001802
Buse E, Habermann G, Vogel F (2006) Thymus development in Macaca fascicularis (Cynomolgus
monkey): an approach for toxicology and embryology. J Mol Histol 37(3–4):161–170
Butler JE, Sun J, Navarro P (1996) The swine Ig heavy chain locus has a single JH and no identifi-
able IgD. Int Immunol 8(12):1897–1904
Butterworth AE (1977) The eosinophil and its role in immunity to helminth infection. Curr Top
Microbiol Immunol 77:127–168
Campbell TW (2004) Hematology of birds. In: Thrall MA, Weiser G, Allison R, Campbell TW
(eds) Veterinary hematology and clinical chemistry. Wiley-Blackwell, New York, pp 238–276
Campbell JL, Eisemann JH, Williams CV, Glenn KM (2000) Description of the gastrointestinal
tract of five lemur species: Propithecus tattersalli, Propithecus verreauxi coquereli, Varecia
variegata, Hapalemur griseus, and Lemur catta. Am J Primatol 52(3):133–142
Canfield PJ (1998) Comparative cell morphology in the peripheral blood film from exotic and
native animals. Aust Vet J 76(12):793–800
Capron A, Dessaint JP, Capron M, Ouma JH, Butterworth AE (1987) Immunity to schistosomes:
progress toward vaccine. Science 238(4830):1065–1072
Vertebrate Immune System Evolution and Comparative Primate Immunity 51
Cardoso LS, Oliveira SC, Araujo MI (2012) Schistosoma mansoni antigens as modulators of the
allergic inflammatory response in asthma. Endocr Metab Immune Disord Drug Targets
12(1):24–32
Carter DL, Shieh TM, Blosser RL, Chadwick KR, Margolick JB, Hildreth JE, Clements JE, Zink
MC (1999) CD56 identifies monocytes and not natural killer cells in rhesus macaques.
Cytometry 37(1):41–50
Cesta MF (2006) Normal structure, function, and histology of mucosa-associated lymphoid Tissue.
Toxicol Pathol 34:599. doi:10.1080/01926230600865531
Cesta MF (2006a) Normal structure, function, and histology of mucosa-associated lymphoid
tissue. Toxicol Pathol 34(5):599–608
Cesta MF (2006b) Normal structure, function, and histology of the spleen. Toxicol Pathol
34(5):455–465
Chan ML, Petravic J, Ortiz AM, Engram J, Paiardini M, Cromer D, Silvestri G, Davenport MP
(2010) Limited CD4+ T cell proliferation leads to preservation of CD4+ T cell counts in SIV-
infected sooty mangabeys. Proc Biol Sci 277(1701):3773–3781
Chen JY, Oliveri P, Li CW, Zhou GQ, Gao F, Hagadorn JW, Peterson KJ, Davidson EH (2000)
Precambrian animal diversity: putative phosphatized embryos from the Doushantuo Formation
of China. Proc Natl Acad Sci USA 97(9):4457–4462
Chen G, Zhuchenko O, Kuspa A (2007) Immune-like phagocyte activity in the social amoeba.
Science 317(5838):678–681
Chieffi Baccari G, de Paulis A, Di Matteo L, Gentile M, Marone G, Minucci S (1998) In situ char-
acterization of mast cells in the frog Rana esculenta. Cell Tissue Res 292(1):151–162
Chivers DJ, Hladik CM (1980) Morphology of the gastrointestinal tract in primates: comparisons
with other mammals in relation to diet. J Morphol 166(3):337–386
Chtanova T, Schaeffer M, Han SJ, van Dooren GG, Nollmann M, Herzmark P, Chan SW, Satija H,
Camfield K, Aaron H et al (2008) Dynamics of neutrophil migration in lymph nodes during
infection. Immunity 29(3):487–496
Church MK, Levi-Schaffer F (1997) The human mast cell. J Allergy Clin Immunol 99(2):
155–160
Clark JM, Abraham WM, Fishman CE, Forteza R, Ahmed A, Cortes A, Warne RL, Moore WR,
Tanaka RD (1995) Tryptase inhibitors block allergen-induced airway and inflammatory
responses in allergic sheep. Am J Respir Crit Care Med 152(6 Pt 1):2076–2083
Coates PT, Barratt-Boyes SM, Zhang L, Donnenberg VS, O’Connell PJ, Logar AJ, Duncan FJ,
Murphey-Corb M, Donnenberg AD, Morelli AE et al (2003) Dendritic cell subsets in blood and
lymphoid tissue of rhesus monkeys and their mobilization with Flt3 ligand. Blood
102(7):2513–2521
Colucci F, Di Santo JP, Leibson PJ (2002) Natural killer cell activation in mice and men: different
triggers for similar weapons? Nat Immunol 3(9):807–813
Conrad RE (1981) Induction and collection of peritoneal exudate macrophages. Dekker, New York
Cooper EL (2010) Evolution of immune systems from self/not self to danger to artificial immune
systems (AIS). Phys Life Rev 7(1):55–78
Cooper PJ, Ayre G, Martin C, Rizzo JA, Ponte EV, Cruz AA (2008) Geohelminth infections: a
review of the role of IgE and assessment of potential risks of anti-IgE treatment. Allergy
63(4):409–417
Cornes JS (1965) Number, size, and distribution of Peyer’s patches in the human small intestine:
Part I The development of Peyer’s patches. Gut 6(3):225–229
Corthesy B (2007) Roundtrip ticket for secretory IgA: role in mucosal homeostasis? J Immunol
178(1):27–32
Criscitiello MF, Saltis M, Flajnik MF (2006) An evolutionarily mobile antigen receptor variable
region gene: doubly rearranging NAR-TcR genes in sharks. Proc Natl Acad Sci USA
103(13):5036–5041
Crivellato E, Ribatti D (2010) The mast cell: an evolutionary perspective. Biol Rev Camb Philos
Soc 85(2):347–360
52 J.F. Brinkworth and M. Thorn
Crivellato E, Nico B, Gallo VP, Ribatti D (2010) Cell secretion mediated by granule-associated
vesicle transport: a glimpse at evolution. Anat Rec (Hoboken) 293(7):1115–1124
Cruz AA, Lima F, Sarinho E, Ayre G, Martin C, Fox H, Cooper PJ (2007) Safety of anti-
immunoglobulin E therapy with omalizumab in allergic patients at risk of geohelminth infec-
tion. Clin Exp Allergy 37(2):197–207
Da’as S, Teh EM, Dobson JT, Nasrallah GK, McBride ER, Wang H, Neuberg DS, Marshall JS, Lin
TJ, Berman JN (2011) Zebrafish mast cells possess an FcvarepsilonRI-like receptor and par-
ticipate in innate and adaptive immune responses. Dev Comp Immunol 35(1):125–134
Dalquest WW, Werner HJ, Robert JH, Richmond ND, Roslund HR, Voge M, Bern HA, Wilber CG,
Sealander JA, Conaway CH et al (1952) General notes. J Mammal 33:102–118
Daniel MD, King NW, Letvin NL, Hunt RD, Sehgal PK, Desrosiers RC (1984) A new type D
retrovirus isolated from macaques with an immunodeficiency syndrome. Science
223(4636):602–605
Davies B, Chattings LS, Edwards SW (1991) Superoxide generation during phagocytosis by
Acanthamoeba castellanii: similarities to the respiratory burst of immune phagocytes.
Microbiology 137(3):705–710
Davis MM, Chien Y (2003) T-cell antigen receptors. In: Paul WE (ed) Fundamental immunology.
Lippincott, Philadelphia, pp 227–258
de Barros CM, Andrade LR, Allodi S, Viskov C, Mourier PA, Cavalcante MC, Straus AH,
Takahashi HK, Pomin VH, Carvalho VF et al (2007) The hemolymph of the ascidian Styela
plicata (Chordata-Tunicata) contains heparin inside basophil-like cells and a unique sulfated
galactoglucan in the plasma. J Biol Chem 282(3):1615–1626
Deza FG, Espinel CS, Beneitez JV (2007) A novel IgA-like immunoglobulin in the reptile
Eublepharis macularius. Dev Comp Immunol 31(6):596–605
Diamond LK (1969) Splenectomy in childhood and the hazard of overwhelming infection.
Pediatrics 43(5):886–889
Diop OM, Ploquin MJ, Mortara L, Faye A, Jacquelin B, Kunkel D, Lebon P, Butor C, Hosmalin A,
Barre-Sinoussi F et al (2008) Plasmacytoid dendritic cell dynamics and alpha interferon
production during Simian immunodeficiency virus infection with a nonpathogenic outcome.
J Virol 82(11):5145–5152
Doeing DC, Borowicz JL, Crockett ET (2003) Gender dimorphism in differential peripheral blood
leukocyte counts in mice using cardiac, tail, foot, and saphenous vein puncture methods. BMC
Clin Pathol 3(1):3
Dooley J, Erickson M, Gillard GO, Farr AG (2006) Cervical thymus in the mouse. J Immunol
176(11):6484–6490
Drayton DL, Liao S, Mounzer RH, Ruddle NH (2006) Lymphoid organ development: from ontog-
eny to neogenesis. Nat Immunol 7(4):344–353
Du Pasquier L (1973) Ontogeny of the immune response in cold-blooded vertebrates. Curr Top
Microbiol Immunol 61:37–88
Du Pasquier L, Robert J, Courtet M, Mussmann R (2000) B-cell development in the amphibian
Xenopus. Immunol Rev 175:201–213
Dvorak AM (2005) Ultrastructural studies of human basophils and mast cells. J Histochem
Cytochem 53(9):1043–1070
Dzik JM (2010) The ancestry and cumulative evolution of immune reactions. Acta Biochim Pol
57(4):443–466
Ear T, McDonald PP (2008) Cytokine generation, promoter activation, and oxidant-independent
NF-kappaB activation in a transfectable human neutrophilic cellular model. BMC Immunol
9:14
Ehlers D, Zosel B, Mohrig W, Kauschke E, Ehlers E (1992) Comparison of an in vivo and in vitro
phagocytosis in Galleria mellonella L. Parasitol Res 78:354–359
Eisenhauer PB, Lehrer RI (1992) Mouse neutrophils lack defensins. Infect Immun
60(8):3446–3447
Vertebrate Immune System Evolution and Comparative Primate Immunity 53
Elbim C, Monceaux V, Mueller YM, Lewis MG, Francois S, Diop O, Akarid K, Hurtrel B,
Gougerot-Pocidalo MA, Levy Y et al (2008) Early divergence in neutrophil apoptosis between
pathogenic and nonpathogenic simian immunodeficiency virus infections of nonhuman pri-
mates. J Immunol 181(12):8613–8623
Engram JC, Cervasi B, Borghans JA, Klatt NR, Gordon SN, Chahroudi A, Else JG, Mittler RS,
Sodora DL, de Boer RJ et al (2010) Lineage-specific T-cell reconstitution following in vivo
CD4+ and CD8+ lymphocyte depletion in nonhuman primates. Blood 116(5):748–758
Epiphanio S, Sinhorini IL, Catao-Dias JL (2003) Pathology of toxoplasmosis in captive new world
primates. J Comp Pathol 129(2–3):196–204
Evans DI (1985) Postsplenectomy sepsis 10 years or more after operation. J Clin Pathol
38(3):309–311
Fabre PH, Rodrigues A, Douzery EJ (2009) Patterns of macroevolution among Primates inferred
from a supermatrix of mitochondrial and nuclear DNA. Mol Phylogenet Evol 53(3):808–825
Fautin DG, Mariscal RN (1991) Cnidaria: anthozoa. In: Harrison FW, Westfall JA (eds) Microscopic
anatomy of invertebrates, placozoa, porfera, cnidaria and ctenophora. Wiley-Liss, Inc., New
York, pp 267–358
Favre D, Lederer S, Kanwar B, Ma ZM, Proll S, Kasakow Z, Mold J, Swainson L, Barbour JD,
Baskin CR et al (2009) Critical loss of the balance between Th17 and T regulatory cell popula-
tions in pathogenic SIV infection. PLoS Pathog 5(2):e1000295
Fisher RE (2000) The primate appendix: a reassessment. Anat Rec 261(6):228–236
Flajnik MF (2002) Comparative analyses of immunoglobulin genes: surprises and portents. Nat
Rev Immunol 2(9):688–698
Flajnik MF, Kasahara M (2001) Comparative genomics of the MHC: glimpses into the evolution
of the adaptive immune system. Immunity 15(3):351–362
Flajnik MF, Kasahara M (2010) Origin and evolution of the adaptive immune system: genetic
events and selective pressures. Nat Rev Genet 11(1):47–59
Fuchs TA, Abed U, Goosmann C, Hurwitz R, Schulze I, Wahn V, Weinrauch Y, Brinkmann V,
Zychlinsky A (2007) Novel cell death program leads to neutrophil extracellular traps. J Cell
Biol 176(2):231–241
Gabriela D, Carlos PL, Clara S, Elkin PM (2005) Phenotypical and functional characterization of
non-human primate Aotus spp. dendritic cells and their use as a tool for characterizing immune
response to protein antigens. Vaccine 23(26):3386–3395
Galli SJ, Kalesnikoff J, Grimbaldeston MA, Piliponsky AM, Williams CM, Tsai M (2005) Mast
cells as “tunable” effector and immunoregulatory cells: recent advances. Annu Rev Immunol
23:749–786
Galli SJ, Grimbaldeston M, Tsai M (2008) Immunomodulatory mast cells: negative, as well as
positive, regulators of immunity. Nat Rev Immunol 8(6):478–486
Geijtenbeek TB, Kwon DS, Torensma R, van Vliet SJ, van Duijnhoven GC, Middel J, Cornelissen
IL, Nottet HS, KewalRamani VN, Littman DR et al (2000a) DC-SIGN, a dendritic cell-specific
HIV-1-binding protein that enhances trans-infection of T cells. Cell 100(5):587–597
Geijtenbeek TB, Torensma R, van Vliet SJ, van Duijnhoven GC, Adema GJ, van Kooyk Y, Figdor
CG (2000b) Identification of DC-SIGN, a novel dendritic cell-specific ICAM-3 receptor that
supports primary immune responses. Cell 100(5):575–585
Geijtenbeek TB, Van Vliet SJ, Koppel EA, Sanchez-Hernandez M, Vandenbroucke-Grauls CM,
Appelmelk B, Van Kooyk Y (2003) Mycobacteria target DC-SIGN to suppress dendritic cell
function. J Exp Med 197(1):7–17
Gilliet M, Cao W, Liu YJ (2008) Plasmacytoid dendritic cells: sensing nucleic acids in viral infec-
tion and autoimmune diseases. Nat Rev Immunol 8(8):594–606
Godson C, Mitchell S, Harvey K, Petasis NA, Hogg N, Brady HR (2000) Cutting edge: lipoxins
rapidly stimulate nonphlogistic phagocytosis of apoptotic neutrophils by monocyte-derived
macrophages. J Immunol 164(4):1663–1667
Gorgollon P (1978) The normal human appendix: a light and electron microscopic study. J Anat
126(Pt 1):87–101
54 J.F. Brinkworth and M. Thorn
Goyos A, Robert J (2009) Tumorigenesis and anti-tumor immune responses in Xenopus. Front
Biosci 14:167–176
Greenhalgh P, Olesen CE, Steiner LA (1993) Characterization and expression of recombination
activating genes (RAG-1 and RAG-2) in Xenopus laevis. J Immunol 151(6):3100–3110
Grevellec A, Tucker AS (2010) The pharyngeal pouches and clefts: development, evolution, struc-
ture and derivatives. Semin Cell Dev Biol 21(3):325–332
Guo P, Hirano M, Herrin BR, Li J, Yu C, Sadlonova A, Cooper MD (2009) Dual nature of the adap-
tive immune system in lampreys. Nature 459(7248):796–801
Hagel I, Di Prisco MC, Goldblatt J, Le Souef PN (2004) The role of parasites in genetic
susceptibility to allergy: IgE, helminthic infection and allergy, and the evolution of the human
immune system. Clin Rev Allergy Immunol 26(2):75–83
Haley PJ (2003) Species differences in the structure and function of the immune system. Toxicology
188(1):49–71
Hamann KJ, Barker RL, Ten RM, Gleich GJ (1991) The molecular biology of eosinophil granule
proteins. Int Arch Allergy Appl Immunol 94(1–4):202–209
Hardy RR (2003) B-lymphocyte development and biology. In: Paul WE (ed) Fundamental immu-
nology. Lippincott, Philadelphia, pp 159–194
Hardy JD, Kurrus FD, Chavez CM, Neely WA, Eraslan S, Turner MD, Fabian LW, Labecki TD
(1964) Heart transplantation in man. Developmental studies and report of a case. JAMA
188:1132–1140
Harmon BG (1998) Avian heterophils in inflammation and disease resistance. Poult Sci
77(7):972–977
Hawkey CM (1985) Analysis of hematologic findings in healthy and sick adult chimpanzees (Pan
troglodytes). J Med Primatol 14(6):327–343
Herberman RB, Nunn ME, Holden HT, Lavrin DH (1975) Natural cytotoxic reactivity of mouse
lymphoid cells against syngeneic and allogeneic tumors. II. Characterization of effector cells.
Int J Cancer 16(2):230–239
Heritage PL, Underdown BJ, Arsenault AL, Snider DP, McDermott MR (1997) Comparison of
murine nasal-associated lymphoid tissue and Peyer’s patches. Am J Respir Crit Care Med
156(4 Pt 1):1256–1262
Herrin BR, Alder MN, Roux KH, Sina C, Ehrhardt GR, Boydston JA, Turnbough CL Jr, Cooper
MD (2008) Structure and specificity of lamprey monoclonal antibodies. Proc Natl Acad Sci
USA 105(6):2040–2045
Hida S, Tadachi M, Saito T, Taki S (2005) Negative control of basophil expansion by IRF-2 critical
for the regulation of Th1/Th2 balance. Blood 106(6):2011–2017
Hirano M, Das S, Guo P, Cooper MD, Frederick WA (2011) The evolution of adaptive immunity
in vertebrates. Adv Immunol 109:125–157
Hitchcock CR, Kiser JC, Telander RL, Seljeskog EL (1964) Baboon renal grafts. JAMA
189:934–937
Hofmann J, Greter M, Du Pasquier L, Becher B (2010) B-cells need a proper house, whereas
T-cells are happy in a cave: the dependence of lymphocytes on secondary lymphoid tissues
during evolution. Trends Immunol 31(4):144–153
Hogan SP, Rosenberg HF, Moqbel R, Phipps S, Foster PS, Lacy P, Kay AB, Rothenberg ME (2008)
Eosinophils: biological properties and role in health and disease. Clin Exp Allergy
38(5):709–750
HogenEsch H, Felsburg PJ (1992) Immunohistology of Peyer’s patches in the dog. Vet Immunol
Immunopathol 30(2–3):147–160
HogenEsch H, Hahn FF (2001) The lymphoid organs: anatomy, development, and age-related
changes. In: Mohr U, Carlton WW, Dungworth DL, Benjamin SA, Capen CC, Hahn FF (eds)
Pathobiology of the aging dog. Iowa State University Press, Ames, pp 127–135
Hollenbach JA, Meenagh A, Sleator C, Alaez C, Bengoche M, Canossi A, Contreras G, Creary L,
Evseeva I, Gorodezky C et al (2010) Report from the killer immunoglobulin-like receptor
(KIR) anthropology component of the 15th International Histocompatibility Workshop: world-
Vertebrate Immune System Evolution and Comparative Primate Immunity 55
wide variation in the KIR loci and further evidence for the co-evolution of KIR and HLA.
Tissue Antigens 76(1):9–17
Hooper LV (2004) Bacterial contributions to mammalian gut development. Trends Microbiol
12(3):129–134
Horton TL, Ritchie P, Watson MD, Horton JD (1996) NK-like activity against allogeneic tumour
cells demonstrated in the spleen of control and thymectomized Xenopus. Immunol Cell Biol
74(4):365–373
Hsu E (1998) Mutation, selection, and memory in B lymphocytes of exothermic vertebrates.
Immunol Rev 162:25–36
Huang G, Xie X, Han Y, Fan L, Chen J, Mou C, Guo L, Liu H, Zhang Q, Chen S et al (2007) The
identification of lymphocyte-like cells and lymphoid-related genes in amphioxus indicates the
twilight for the emergence of adaptive immune system. PLoS One 2(2):e206
Hutchison AT, Schmitz JE, Miller CJ, Sastry KJ, Nehete PN, Major AM, Ansari AA, Tatevian N,
Lewis DE (2011) Increased inherent intestinal granzyme B expression may be associated with
SIV pathogenesis in Asian non-human primates. J Med Primatol 40(6):414–426
Janeway CA Jr, Medzhitov R (2002) Innate immune recognition. Annu Rev Immunol
20:197–216
Jansen CA, van de Haar PM, van Haarlem D, van Kooten P, de Wit S, van Eden W, Viertlbock BC,
Gobel TW, Vervelde L (2010) Identification of new populations of chicken natural killer (NK)
cells. Dev Comp Immunol 34(7):759–767
Jaroenpool J, Rogers KA, Pattanapanyasat K, Villinger F, Onlamoon N, Crocker PR, Ansari AA
(2007) Differences in the constitutive and SIV infection induced expression of Siglecs by
hematopoietic cells from non-human primates. Cell Immunol 250(1–2):91–104
Jawdat DM, Albert EJ, Rowden G, Haidl ID, Marshall JS (2004) IgE-mediated mast cell activation
induces Langerhans cell migration in vivo. J Immunol 173(8):5275–5282
Jenkins MK (2003) Peripheral T-lymphocyte responses and function. In: Paul WE (ed) Fundamental
immunology. Lippincott, Philadelphia, pp 303–320
Jesudason S, Collins MG, Rogers NM, Kireta S, Coates PT (2012) Non-human primate dendritic
cells. J Leukoc Biol 91(2):217–228
Jeurissen SH (1991) Structure and function of the chicken spleen. Res Immunol 142(4):352–355
Kaldjian EP, Gretz JE, Anderson AO, Shi Y, Shaw S (2001) Spatial and molecular organization of
lymph node T cell cortex: a labyrinthine cavity bounded by an epithelium-like monolayer of
fibroblastic reticular cells anchored to basement membrane-like extracellular matrix. Int
Immunol 13(10):1243–1253
Kapetanovic R, Cavaillon JM (2007) Early events in innate immunity in the recognition of micro-
bial pathogens. Expert Opin Biol Ther 7(6):907–918
Karasuyama H, Mukai K, Obata K, Tsujimura Y, Wada T (2011) Nonredundant roles of basophils
in immunity. Annu Rev Immunol 29:45–69
Karrer U, Althage A, Odermatt B, Roberts CW, Korsmeyer SJ, Miyawaki S, Hengartner H,
Zinkernagel RM (1997) On the key role of secondary lymphoid organs in antiviral immune
responses studied in alymphoplastic (aly/aly) and spleenless (Hox11(-)/-) mutant mice. J Exp
Med 185(12):2157–2170
Kaspers B, Kothlow S, Butter C (2008) Avian antigen presenting cells. In: Davidson F, Kaspers B,
Schat KA (eds) Avian immunology. Academic, London, pp 183–202
Kauschke E, Komiyama K, Moro I, Eue I, Konig S, Cooper EL (2001) Evidence for perforin-like
activity associated with earthworm leukocytes. Zoology (Jena) 104(1):13–24
Kelenyi G, Nemeth A (1969) Comparative histochemistry and electron microscopy of the eosino-
phil leucocytes of vertebrates. I. A study of avian, reptile, amphibian and fish leucocytes. Acta
Biol Acad Sci Hung 20(4):405–422
Ketloy C, Engering A, Srichairatanakul U, Limsalakpetch A, Yongvanitchit K, Pichyangkul S,
Ruxrungtham K (2008) Expression and function of Toll-like receptors on dendritic cells and
other antigen presenting cells from non-human primates. Vet Immunol Immunopathol
125(1–2):18–30
56 J.F. Brinkworth and M. Thorn
Khalturin K, Becker M, Rinkevich B, Bosch TC (2003) Urochordates and the origin of natural
killer cells: identification of a CD94/NKR-P1-related receptor in blood cells of Botryllus. Proc
Natl Acad Sci USA 100(2):622–627
King BK, Li J, Kudsk KA (1997a) A temporal study of TPN-induced changes in gut-associated
lymphoid tissue and mucosal immunity. Arch Surg 132(12):1303–1309
King CL, Xianli J, Malhotra I, Liu S, Mahmoud AA, Oettgen HC (1997b) Mice with a targeted
deletion of the IgE gene have increased worm burdens and reduced granulomatous inflammation
following primary infection with Schistosoma mansoni. J Immunol 158(1):294–300
Kiyono H, Fukuyama S (2004) NALT- versus Peyer’s-patch-mediated mucosal immunity. Nat Rev
Immunol 4(9):699–710
Kobayashi Y (2008) The role of chemokines in neutrophil biology. Front Biosci 13:2400–2407
Kreslavsky T, Gleimer M, Garbe AI, von Boehmer H (2010) alphabeta versus gammadelta fate
choice: counting the T-cell lineages at the branch point. Immunol Rev 238(1):169–181
Krishnaswamy G, Kelley J, Johnson D, Youngberg G, Stone W, Huang SK, Bieber J, Chi DS (2001)
The human mast cell: functions in physiology and disease. Front Biosci 6:D1109–D1127
Kumar H, Kawai T, Akira S (2009) Pathogen recognition in the innate immune response. Biochem
J 420(1):1–16
Kurtz J (2002) Phagocytosis by invertebrate hemocytes: causes of individual variation in Panorpa
vulgaris scorpionflies. Microsc Res Tech 57(6):456–468
Kushwah R, Hu J (2011) Complexity of dendritic cell subsets and their function in the host immune
system. Immunology 133(4):409–419
Kwa SF, Beverley P, Smith AL (2006) Peyer’s patches are required for the induction of rapid Th1
responses in the gut and mesenteric lymph nodes during an enteric infection. J Immunol
176(12):7533–7541
Kyriazis AA, Esterly JR (1971) Fetal and neonatal development of lymphoid tissues. Arch Pathol
91(5):444–451
Lam SH, Chua HL, Gong Z, Wen Z, Lam TJ, Sin YM (2002) Morphologic transformation of the
thymus in developing zebrafish. Dev Dyn 225(1):87–94
Lanier LL (2008) Up on the tightrope: natural killer cell activation and inhibition. Nat Immunol
9(5):495–502
Laurin M, Everett ML, Parker W (2011) The cecal appendix: one more immune component with a
function disturbed by post-industrial culture. Anat Rec (Hoboken) 294(4):567–579
Lawn AM, Rose ME (1981) Presence of a complete endothelial barrier between lymph and lym-
phoid tissue in the lumbar lymph nodes of the duck (Anas platyrhynchos). Res Vet Sci
30(3):335–342
Lee JJ, Lee NA (2005) Eosinophil degranulation: an evolutionary vestige or a universally destruc-
tive effector function? Clin Exp Allergy 35(8):986–994
Lee WL, Harrison RE, Grinstein S (2003) Phagocytosis by neutrophils. Microbes Infect
5(14):1299–1306
Lee JJ, Jacobsen EA, Ochkur SI, McGarry MP, Condjella RM, Doyle AD, Luo H, Zellner KR,
Protheroe CA, Willetts L et al (2012) Human versus mouse eosinophils: “that which we call an
eosinophil, by any other name would stain as red”. J Allergy Clin Immunol 130(3):572–584
Leulier F, Parquet C, Pili-Floury S, Ryu JH, Caroff M, Lee WJ, Mengin-Lecreulx D, Lemaitre B
(2003) The Drosophila immune system detects bacteria through specific peptidoglycan recog-
nition. Nat Immunol 4(5):478–484
Ley K, Allietta M, Bullard DC, Morgan S (1998) Importance of E-selectin for firm leukocyte adhe-
sion in vivo. Circ Res 83(3):287–294
Lieberman J (2003) The ABCs of granule-mediated cytotoxicity: new weapons in the arsenal. Nat
Rev Immunol 3(5):361–370
Lind JP, Wolff PL, Petrini KR, Keyley CW, Olson DE, Redig PT (1990) Morphology of the eosino-
phil in raptors. J Assoc Avian Veterinarians 4:33–38
Litonjua AA, Carey VJ, Weiss ST, Gold DR (1999) Race, socioeconomic factors, and area of resi-
dence are associated with asthma prevalence. Pediatr Pulmonol 28(6):394–401
Vertebrate Immune System Evolution and Comparative Primate Immunity 57
Loisel DA, Tung J (2013) Genetic variation in the immune system of Old World monkeys: func-
tional and selective effects. In: Brinkworth JF, Pechenkina E (eds) Primates, pathogens, and
evolution. Springer, Heidelberg
MacDonald KP, Munster DJ, Clark GJ, Dzionek A, Schmitz J, Hart DN (2002) Characterization of
human blood dendritic cell subsets. Blood 100(13):4512–4520
Malaviya R, Ikeda T, Ross E, Abraham SN (1996) Mast cell modulation of neutrophil influx and
bacterial clearance at sites of infection through TNF-alpha. Nature 381(6577):77–80
Malleret B, Karlsson I, Maneglier B, Brochard P, Delache B, Andrieu T, Muller-Trutwin M, Beaumont
T, McCune JM, Banchereau J et al (2008) Effect of SIVmac infection on plasmacytoid and
CD1c+ myeloid dendritic cells in cynomolgus macaques. Immunology 124(2):223–233
Mandl JN, Barry AP, Vanderford TH, Kozyr N, Chavan R, Klucking S, Barrat FJ, Coffman RL, Staprans
SI, Feinberg MB (2008) Divergent TLR7 and TLR9 signaling and type I interferon production
distinguish pathogenic and nonpathogenic AIDS virus infections. Nat Med 14(10):1077–1087
Mandl JN, Akondy R, Lawson B, Kozyr N, Staprans SI, Ahmed R, Feinberg MB (2011) Distinctive
TLR7 signaling, type I IFN production, and attenuated innate and adaptive immune responses
to yellow fever virus in a primate reservoir host. J Immunol 186(11):6406–6416
Marjanovic D, Laurin M (2007) Fossils, molecules, divergence times, and the origin of lissamphib-
ians. Syst Biol 56(3):369–388
Marr S, Morales H, Bottaro A, Cooper M, Flajnik M, Robert J (2007) Localization and differential
expression of activation-induced cytidine deaminase in the amphibian Xenopus upon antigen
stimulation and during early development. J Immunol 179(10):6783–6789
Martin JS, Renshaw SA (2009) Using in vivo zebrafish models to understand the biochemical basis
of neutrophilic respiratory disease. Biochem Soc Trans 37(Pt 4):830–837
Martin M, Blackwell AD, Gurven M, Kaplan H (2013) Make new friends and keep the old?
Parasite coinfection and comorbidity in Homo sapiens. In: Brinkworth JF, Pechenkina E (eds)
Primates, pathogens, and evolution. Springer, Heidelberg
Mayer WE, Uinuk-Ool T, Tichy H, Gartland LA, Klein J, Cooper MD (2002) Isolation and char-
acterization of lymphocyte-like cells from a lamprey. Proc Natl Acad Sci USA
99(22):14350–14355
Mebius RE, Kraal G (2005) Structure and function of the spleen. Nat Rev Immunol 5(8):606–616
Mehandru S, Poles MA, Tenner-Racz K, Horowitz A, Hurley A, Hogan C, Boden D, Racz P,
Markowitz M (2004) Primary HIV-1 infection is associated with preferential depletion of CD4+
T lymphocytes from effector sites in the gastrointestinal tract. J Exp Med 200(6):761–770
Messaoudi I, Estep R, Robinson B, Wong SW (2011) Nonhuman primate models of human immu-
nology. Antioxid Redox Signal 14(2):261–273
Mestas J, Hughes CC (2004) Of mice and not men: differences between mouse and human immu-
nology. J Immunol 172(5):2731–2738
Metchnikoff E (1887) Uber den Kampf der Zellen gegen Erysipelkokken. Archiv fur Pathologische
Anatomie und Physiologie und fur Klinische Medicin 107:209–249
Metchnikoff E (1893) Lecon sur la pathologie comparee de inflammation. Ann Inst Pasteur
7:348–357
Mitchell PC (1916) Further observations on the intestinal tract of mammals. Proc Zool Soc Lond
86(1):183–252
Moita LF, Wang-Sattler R, Michel K, Zimmermann T, Blandin S, Levashina EA, Kafatos FC
(2005) In vivo identification of novel regulators and conserved pathways of phagocytosis in A.
gambiae. Immunity 23(1):65–73
Montali RJ (1988) Comparative pathology of inflammation in the higher vertebrates (reptiles,
birds and mammals). J Comp Pathol 99(1):1–26
Mosser DM, Edwards JP (2008) Exploring the full spectrum of macrophage activation. Nat Rev
Immunol 8(12):958–969
Mulero I, Sepulcre MP, Meseguer J, Garcia-Ayala A, Mulero V (2007) Histamine is stored in mast
cells of most evolutionarily advanced fish and regulates the fish inflammatory response. Proc
Natl Acad Sci USA 104(49):19434–19439
58 J.F. Brinkworth and M. Thorn
Murata H (1959) Comparative studies of the spleen in submammalian vertebrates. II. Minute struc-
ture of the spleen, with special reference to the periarterial lymphoid sheath. Bull Yamaguchi
Med Sch 6:83–105
Murphey-Corb M, Martin LN, Rangan SR, Baskin GB, Gormus BJ, Wolf RH, Andes WA, West M,
Montelaro RC (1986) Isolation of an HTLV-III-related retrovirus from macaques with simian
AIDS and its possible origin in asymptomatic mangabeys. Nature 321(6068):435–437
Mussmann R, Du Pasquier L, Hsu E (1996a) Is Xenopus IgX an analog of IgA? Eur J Immunol
26(12):2823–2830
Mussmann R, Wilson M, Marcuz A, Courtet M, Du Pasquier L (1996b) Membrane exon sequences
of the three Xenopus Ig classes explain the evolutionary origin of mammalian isotypes. Eur J
Immunol 26(2):409–414
Nagata T, Suzuki T, Ohta Y, Flajnik MF, Kasahara M (2002) The leukocyte common antigen
(CD45) of the Pacific hagfish, Eptatretus stoutii: implications for the primordial function of
CD45. Immunogenetics 54(4):286–291
Neefjes J, Jongsma ML, Paul P, Bakke O (2011) Towards a systems understanding of MHC class
I and MHC class II antigen presentation. Nat Rev Immunol 11(12):823–836
Ngampasutadol J, Tran C, Gulati S, Blom AM, Jerse EA, Ram S, Rice PA (2008) Species-
specificity of Neisseria gonorrhoeae infection: do human complement regulators contribute?
Vaccine 26(Suppl 8):I62–I66
Nunn CL (2002) A comparative study of leukocyte counts and disease risk in primates. Evolution
56(1):177–190
Nussenzweig MC, Steinman RM, Gutchinov B, Cohn ZA (1980) Dendritic cells are accessory
cells for the development of anti-trinitrophenyl cytotoxic T lymphocytes. J Exp Med
152(4):1070–1084
Ohnmacht C, Voehringer D (2009) Basophil effector function and homeostasis during helminth
infection. Blood 113(12):2816–2825
Ohno S (1970) Evolution by gene duplication. Springer, New York, p 160
Ohta Y, Flajnik M (2006) IgD, like IgM, is a primordial immunoglobulin class perpetuated in most
jawed vertebrates. Proc Natl Acad Sci USA 103(28):10723–10728
Ohta S, Ueda Y, Yaguchi M, Matsuzaki Y, Nakamura M, Toyama Y, Tanioka Y, Tamaoki N, Nomura
T, Okano H et al (2008) Isolation and characterization of dendritic cells from common marmo-
sets for preclinical cell therapy studies. Immunology 123(4):566–574
Olah I, Vervelde L (2008) Structure of the avian lymphoid system. In: Davison F, Kaspers B, Schat
KA (eds) Avian immunology. Elsevier, London, pp 13–50
Olsson I, Venge P (1974) Cationic proteins of human granulocytes. II. Separation of the cationic
proteins of the granules of leukemic myeloid cells. Blood 44(2):235–246
Olsson I, Venge P, Spitznagel JK, Lehrer RI (1977) Arginine-rich cationic proteins of human eosin-
ophil granules: comparison of the constituents of eosinophilic and neutrophilic leukocytes. Lab
Invest 36(5):493–500
Owen RL, Piazza AJ, Ermak TH (1991) Ultrastructural and cytoarchitectural features of lympho-
reticular organs in the colon and rectum of adult BALB/c mice. Am J Anat 190(1):10–18
Paiardini M, Cervasi B, Reyes-Aviles E, Micci L, Ortiz AM, Chahroudi A, Vinton C, Gordon SN,
Bosinger SE, Francella N et al (2011) Low levels of SIV infection in sooty mangabey central
memory CD4(+) T cells are associated with limited CCR5 expression. Nat Med
17(7):830–836
Pancer Z, Mayer WE, Klein J, Cooper MD (2004) Prototypic T cell receptor and CD4-like core-
ceptor are expressed by lymphocytes in the agnathan sea lamprey. Proc Natl Acad Sci USA
101(36):13273–13278
Pandrea I, Apetrei C (2010) Where the wild things are: pathogenesis of SIV infection in African
nonhuman primate hosts. Curr HIV/AIDS Rep 7(1):28–36
Parra ZE, Baker ML, Schwarz RS, Deakin JE, Lindblad-Toh K, Miller RD (2007) A unique T cell
receptor discovered in marsupials. Proc Natl Acad Sci USA 104(23):9776–9781
Paul WE (2003) The immune system: an introduction. In: Paul WE (ed) Fundamental immunol-
ogy. Lippincott, Philadelphia, pp 1–22
Vertebrate Immune System Evolution and Comparative Primate Immunity 59
Pearson C, Uhlig HH, Powrie F (2012) Lymphoid microenvironments and innate lymphoid cells in
the gut. Trends Immunol 33(6):289–296
Pech LL, Strand MR (1996) Granular cells are required for encapsulation of foreign targets by
insect haemocytes. J Cell Sci 109(Pt 8):2053–2060
Penniall R, Spitznagel JK (1975) Chicken neutrophils: oxidative metabolism in phagocytic cells
devoid of myeloperoxidase. Proc Natl Acad Sci USA 72(12):5012–5015
Pereira LE, Ansari AA (2009) A case for innate immune effector mechanisms as contributors to
disease resistance in SIV-infected sooty mangabeys. Curr HIV Res 7(1):12–22
Peterson KJ, Butterfield NJ (2005) Origin of the Eumetazoa: testing ecological predictions of
molecular clocks against the Proterozoic fossil record. Proc Natl Acad Sci USA
102(27):9547–9552
Pichyangkul S, Saengkrai P, Yongvanitchit K, Limsomwong C, Gettayacamin M, Walsh DS,
Stewart VA, Ballou WR, Heppner DG (2001) Isolation and characterization of rhesus blood
dendritic cells using flow cytometry. J Immunol Methods 252(1–2):15–23
Pickel JM, McCormack WT, Chen CH, Cooper MD, Thompson CB (1993) Differential regulation
of V(D)J recombination during development of avian B and T cells. Int Immunol
5(8):919–927
Pospisil R, Mage RG (1998a) B-cell superantigens may play a role in B-cell development and
selection in the young rabbit appendix. Cell Immunol 185(2):93–100
Pospisil R, Mage RG (1998b) Rabbit appendix: a site of development and selection of the B cell
repertoire. Curr Top Microbiol Immunol 229:59–70
Prasad S, Kireta S, Leedham E, Russ GR, Coates PT (2010) Propagation and characterisation of
dendritic cells from G-CSF mobilised peripheral blood monocytes and stem cells in common
marmoset monkeys. J Immunol Methods 352(1–2):59–70
Purdy AK, Campbell KS (2009) Natural killer cells and cancer: regulation by the killer cell Ig-like
receptors (KIR). Cancer Biol Ther 8(23):2211–2220
Raffatellu M, Santos RL, Verhoeven DE, George MD, Wilson RP, Winter SE, Godinez I, Sankaran
S, Paixao TA, Gordon MA et al (2008) Simian immunodeficiency virus-induced mucosal inter-
leukin-17 deficiency promotes Salmonella dissemination from the gut. Nat Med 14(4):421–428
Raison RL, Gilbertson P, Wotherspoon J (1987) Cellular requirements for mixed leucocyte reactiv-
ity in the cyclostome, Eptatretus stoutii. Immunol Cell Biol 65(Pt 2):183–188
Randal Bollinger R, Barbas AS, Bush EL, Lin SS, Parker W (2007) Biofilms in the large bowel
suggest an apparent function of the human vermiform appendix. J Theor Biol
249(4):826–831
Rast JP, Anderson MK, Strong SJ, Luer C, Litman RT, Litman GW (1997) Alpha, beta, gamma,
and delta T cell antigen receptor genes arose early in vertebrate phylogeny. Immunity
6(1):1–11
Reemtsma K, McCracken BH, Schlegel JU, Pearl M (1964a) Heterotransplantation of the kidney:
two clinical experiences. Science 143(3607):700–702
Reemtsma K, McCracken BH, Schlegel JU, Pearl MA, Pearce CW, Dewitt CW, Smith PE, Hewitt RL,
Flinner RL, Creech O Jr (1964b) Renal heterotransplantation in man. Ann Surg 160:384–410
Reite OB (1965) A phylogenetical approach to the functional significance of tissue mast cell hista-
mine. Nature 206(991):1334–1336
Reite OB (1972) Comparative physiology of histamine. Physiol Rev 52(3):778–819
Rhodes CP, Ratcliffe NA, Rowley AF (1982) Presence of coelomocytes in the primitive chordate
amphioxus (Branchiostoma lanceolatum). Science 217(4556):263–265
Robert J, Ohta Y (2009) Comparative and developmental study of the immune system in Xenopus.
Dev Dyn 238(6):1249–1270
Rodewald HR (2008) Thymus organogenesis. Annu Rev Immunol 26:355–388
Rogers KA, Scinicariello F, Attanasio R (2006) IgG Fc receptor III homologues in nonhuman
primate species: genetic characterization and ligand interactions. J Immunol
177(6):3848–3856
Rogers KA, Jayashankar L, Scinicariello F, Attanasio R (2008a) Nonhuman primate IgA: genetic
heterogeneity and interactions with CD89. J Immunol 180(7):4816–4824
60 J.F. Brinkworth and M. Thorn
Rogers SL, Viertlboeck BC, Gobel TW, Kaufman J (2008b) Avian NK activities, cells and
receptors. Semin Immunol 20(6):353–360
Rogozin IB, Iyer LM, Liang L, Glazko GV, Liston VG, Pavlov YI, Aravind L, Pancer Z (2007)
Evolution and diversification of lamprey antigen receptors: evidence for involvement of an
AID-APOBEC family cytosine deaminase. Nat Immunol 8(6):647–656
Rook GAW (2013) Microbial exposures and other early childhood influences on the subsequent
function of the immune system. In: Brinkworth JF, Pechenkina E (eds) Primates, pathogens,
and evolution. Springer, Heidelberg
Rorke S, Holgate ST (2004) The atopy phenotype revisited. Revue francaise d’allergologie e
d’immunologie clinique 44:436–444
Rose ME, Hesketh P (1974) Fowl peritoneal exudate cells: collection and use for the macrophage
migration inhibition test. Avian Pathol 3(4):297–300
Rosenberg HF, Domachowske JB (2001) Eosinophils, eosinophil ribonucleases, and their role in
host defense against respiratory virus pathogens. J Leukoc Biol 70(5):691–698
Rosenberg HF, Dyer KD (1995) Eosinophil cationic protein and eosinophil-derived neurotoxin.
Evolution of novel function in a primate ribonuclease gene family. J Biol Chem 270(37):
21539–21544
Rosenberg HF, Dyer KD, Tiffany HL, Gonzalez M (1995) Rapid evolution of a unique family of
primate ribonuclease genes. Nat Genet 10(2):219–223
Rowley AF, Page M (1985) Ultrastructural, cytochemical and functional studies on the eosino-
philic granulocytes of larval lampreys. Cell Tissue Res 240(3):705–709
Rumfelt LL, McKinney EC, Taylor E, Flajnik MF (2002) The development of primary and second-
ary lymphoid tissues in the nurse shark Ginglymostoma cirratum: B-cell zones precede den-
dritic cell immigration and T-cell zone formation during ontogeny of the spleen. Scand J
Immunol 56(2):130–148
Sabet T, Wen-Cheng H, Stanisz M, El-Domeiri A, Van Alten P (1977) A simple method for obtain-
ing peritoneal macrophages from chickens. J Immunol Methods 14(2):103–110
Sailendri K, Muthukkaruppan V (1975) Morphology of lymphoid organs in a cichlid teleost,
Tilapia mossambica (Peters). J Morphol 147(1):109–121
Savina A, Amigorena S (2007) Phagocytosis and antigen presentation in dendritic cells. Immunol
Rev 219:143–156
Scapini P, Laudanna C, Pinardi C, Allavena P, Mantovani A, Sozzani S, Cassatella MA (2001)
Neutrophils produce biologically active macrophage inflammatory protein-3alpha (MIP-
3alpha)/CCL20 and MIP-3beta/CCL19. Eur J Immunol 31(7):1981–1988
Schatz DG, Oettinger MA, Schlissel MS (1992) V(D)J recombination: molecular biology and
regulation. Annu Rev Immunol 10:359–383
Schroeder HW Jr, Cavacini L (2010) Structure and function of immunoglobulins. J Allergy Clin
Immunol 125(2 Suppl 2):S41–S52
Schwartz LB, Metcalfe DD, Miller JS, Earl H, Sullivan T (1987) Tryptase levels as an indicator of
mast-cell activation in systemic anaphylaxis and mastocytosis. N Engl J Med 316(26):
1622–1626
Secombes CJ, Manning MJ (1980) Comparative studies on the immune system of fishes and
amphibians: antigen localisation in carp Cyprinus carpio L. J Fish Dis 3:399
Seong SY, Matzinger P (2004) Hydrophobicity: an ancient damage-associated molecular pattern
that initiates innate immune responses. Nat Rev Immunol 4(6):469–478
Setiamarga DH, Miya M, Yamanoue Y, Azuma Y, Inoue JG, Ishiguro NB, Mabuchi K, Nishida M
(2009) Divergence time of the two regional medaka populations in Japan as a new time scale
for comparative genomics of vertebrates. Biol Lett 5(6):812–816
Shah SS, Lai SY, Ruchelli E, Kazahaya K, Mahboubi S (2001) Retropharyngeal aberrant thymus.
Pediatrics 108(5):E94
Shamri R, Xenakis J, Spencer L (2011) Eosinophils in innate immunity: an evolving story. Cell
Tissue Res 343(1):57–83
Shedlock AM, Edwards SV (2009) Amniotes (Amniota). In: Hedges SB, Kumar S (eds) The
timetree of life. Oxford University Press, New York, pp 375–379
Vertebrate Immune System Evolution and Comparative Primate Immunity 61
Shen L, Stuge TB, Bengten E, Wilson M, Chinchar VG, Naftel JP, Bernanke JM, Clem LW, Miller
NW (2004) Identification and characterization of clonal NK-like cells from channel catfish
(Ictalurus punctatus). Dev Comp Immunol 28(2):139–152
Shikina T, Hiroi T, Iwatani K, Jang MH, Fukuyama S, Tamura M, Kubo T, Ishikawa H, Kiyono H
(2004) IgA class switch occurs in the organized nasopharynx- and gut-associated lymphoid tis-
sue, but not in the diffuse lamina propria of airways and gut. J Immunol 172(10):6259–6264
Siddiqui R, Khan NA (2012) Acanthamoeba is an evolutionary ancestor of macrophages: a myth
or reality? Exp Parasitol 130(2):95–97
Sitnicka E, Buza-Vidas N, Larsson S, Nygren JM, Liuba K, Jacobsen SE (2003) Human CD34+
hematopoietic stem cells capable of multilineage engrafting NOD/SCID mice express flt3:
distinct flt3 and c-kit expression and response patterns on mouse and candidate human hema-
topoietic stem cells. Blood 102(3):881–886
Smith ML, Olson TS, Ley K (2004) CXCR2- and E-selectin-induced neutrophil arrest during
inflammation in vivo. J Exp Med 200(7):935–939
Smith HF, Fisher RE, Everett ML, Thomas AD, Bollinger RR, Parker W (2009) Comparative
anatomy and phylogenetic distribution of the mammalian cecal appendix. J Evol Biol
22(10):1984–1999
Soderlund J, Nilsson C, Ekman M, Walther L, Gaines H, Biberfeld G, Biberfeld P (2000)
Recruitment of monocyte derived dendritic cells ex vivo from SIV infected and non-infected
cynomolgus monkeys. Scand J Immunol 51(2):186–194
Soehnlein O, Kai-Larsen Y, Frithiof R, Sorensen OE, Kenne E, Scharffetter-Kochanek K, Eriksson
EE, Herwald H, Agerberth B, Lindbom L (2008) Neutrophil primary granule proteins HBP and
HNP1-3 boost bacterial phagocytosis by human and murine macrophages. J Clin Invest
118(10):3491–3502
Soehnlein O, Zernecke A, Weber C (2009) Neutrophils launch monocyte extravasation by release
of granule proteins. Thromb Haemost 102(2):198–205
Sokol CL, Chu NQ, Yu S, Nish SA, Laufer TM, Medzhitov R (2009) Basophils function as antigen-
presenting cells for an allergen-induced T helper type 2 response. Nat Immunol
10(7):713–720
Soto PC, Stein LL, Hurtado-Ziola N, Hedrick SM, Varki A (2010) Relative over-reactivity of
human versus chimpanzee lymphocytes: implications for the human diseases associated with
immune activation. J Immunol 184(8):4185–4195
Spencer J, MacDonald TT, Finn T, Isaacson PG (1986) The development of gut associated lym-
phoid tissue in the terminal ileum of fetal human intestine. Clin Exp Immunol 64(3):536–543
Stabler JG, McCormick TW, Powell KC, Kogut MH (1994) Avian heterophils and monocytes:
phagocytic and bactericidal activities against Salmonella enteritidis. Vet Microbiol 38(4):
293–305
Starzl TE, Marchioro TL, Peters GN, Kirkpatrick CH, Wilson WE, Porter KA, Rifkind D, Ogden
DA, Hitchcock CR, Waddell WR (1964) Renal heterotransplantation from baboon to man:
experience with 6 cases. Transplantation 2:752–776
Steiper ME, Young NM (2006) Primate molecular divergence dates. Mol Phylogenet Evol
41(2):384–394
Stone KD, Prussin C, Metcalfe DD (2010) IgE, mast cells, basophils, and eosinophils. J Allergy
Clin Immunol 125(2 Suppl 2):S73–S80
Strugnell RA, Wijburg OL (2010) The role of secretory antibodies in infection immunity. Nat Rev
Microbiol 8(9):656–667
Stubbe H, Berdoz J, Kraehenbuhl JP, Corthesy B (2000) Polymeric IgA is superior to monomeric
IgA and IgG carrying the same variable domain in preventing Clostridium difficile toxin A
damaging of T84 monolayers. J Immunol 164(4):1952–1960
Sturm-O’Brien AK, Salazar JD, Byrd RH, Popek EJ, Giannoni CM, Friedman EM, Sulek M,
Larrier DR (2009) Cervical thymic anomalies–the Texas Children’s Hospital experience.
Laryngoscope 119(10):1988–1993
Sugimura M, Hashimoto Y, Nakanishi YH (1977) Thymus- and bursa-dependent areas in duck
lymph nodes. Jpn J Vet Res 25(1–2):7–16
62 J.F. Brinkworth and M. Thorn
Sutherland DB, Fagarasan S (2012) IgA synthesis: a form of functional immune adaptation extend-
ing beyond gut. Curr Opin Immunol 24(3):261–268
Suzuki K, Kawamoto S, Maruya M, Fagarasan S (2010) GALT: organization and dynamics leading
to IgA synthesis. Adv Immunol 107:153–185
Takaya K (1969) The relationship between mast cells and histamine in phylogeny with special
reference to reptiles and birds. Arch Histol Jpn 30(4):401–420
Takaya K, Fujita T, Endo K (1967) Mast cells free of histamine in Rana catasbiana. Nature
215(5102):776–777
Tanaka S, Miura S, Tashiro H, Serizawa H, Hamada Y, Yoshioka M, Tsuchiya M (1991)
Morphological alteration of gut-associated lymphoid tissue after long-term total parenteral
nutrition in rats. Cell Tissue Res 266(1):29–36
Tavassoli M (1986) Bone marrow in boneless fish: lessons of evolution. Med Hypotheses 20(1):9–15
Teixeira MM, Talvani A, Tafuri WL, Lukacs NW, Hellewell PG (2001) Eosinophil recruitment into
sites of delayed-type hypersensitivity reactions in mice. J Leukoc Biol 69(3):353–360
Terszowski G, Muller SM, Bleul CC, Blum C, Schirmbeck R, Reimann J, Pasquier LD, Amagai T,
Boehm T, Rodewald HR (2006) Evidence for a functional second thymus in mice. Science
312(5771):284–287
Thai TH, Purugganan MM, Roth DB, Kearney JF (2002) Distinct and opposite diversifying activi-
ties of terminal transferase splice variants. Nat Immunol 3(5):457–462
Thomson R, Molina-Portela P, Mott H, Carrington M, Raper J (2009) Hydrodynamic gene delivery
of baboon trypanosome lytic factor eliminates both animal and human-infective African try-
panosomes. Proc Natl Acad Sci USA 106(46):19509–19514
Thorpe M, Yu J, Boinapally V, Ahooghalandari P, Kervinen J, Garavilla LD, Hellman L (2012)
Extended cleavage specificity of the mast cell chymase from the crab-eating macaque (Macaca
fascicularis): an interesting animal model for the analysis of the function of the human mast
cell chymase. Int Immunol 24:771–782
Tovi F, Mares AJ (1978) The aberrant cervical thymus. Embryology, pathology, and clinical impli-
cations. Am J Surg 136(5):631–637
Trivedi NN, Tong Q, Raman K, Bhagwandin VJ, Caughey GH (2007) Mast cell alpha and beta
tryptases changed rapidly during primate speciation and evolved from gamma-like transmem-
brane peptidases in ancestral vertebrates. J Immunol 179(9):6072–6079
Trivedi NN, Raymond WW, Caughey GH (2008) Chimerism, point mutation, and truncation dra-
matically transformed mast cell delta-tryptases during primate evolution. J Allergy Clin
Immunol 121(5):1262–1268
Turner RJ, Manning MJ (1973) Response of the toad, Xenopus laevis, to circulating antigens.
Cellular changes in the spleen. J Exp Zool 183(1):21–34
Uhm TG, Kim BS, Chung IY (2012) Eosinophil development, regulation of eosinophil-specific
genes, and role of eosinophils in the pathogenesis of asthma. Allergy Asthma Immunol Res
4(2):68–79
Van den Broeck W, Derore A, Simoens P (2006) Anatomy and nomenclature of murine lymph
nodes: descriptive study and nomenclatory standardization in BALB/cAnNCrl mice. J Immunol
Methods 312(1–2):12–19
Veazey RS, DeMaria M, Chalifoux LV, Shvetz DE, Pauley DR, Knight HL, Rosenzweig M,
Johnson RP, Desrosiers RC, Lackner AA (1998) Gastrointestinal tract as a major site of CD4+
T cell depletion and viral replication in SIV infection. Science 280(5362):427–431
Veiga-Fernandes H, Coles MC, Foster KE, Patel A, Williams A, Natarajan D, Barlow A, Pachnis
V, Kioussis D (2007) Tyrosine kinase receptor RET is a key regulator of Peyer’s patch organo-
genesis. Nature 446(7135):547–551
Vernersson M, Aveskogh M, Munday B, Hellman L (2002) Evidence for an early appearance of
modern post-switch immunoglobulin isotypes in mammalian evolution (II); cloning of IgE,
IgG1 and IgG2 from a monotreme, the duck-billed platypus, Ornithorhynchus anatinus. Eur J
Immunol 32(8):2145–2155
Vernersson M, Aveskogh M, Hellman L (2004) Cloning of IgE from the echidna (Tachyglossus
aculeatus) and a comparative analysis of epsilon chains from all three extant mammalian lin-
eages. Dev Comp Immunol 28(1):61–75
Vertebrate Immune System Evolution and Comparative Primate Immunity 63
Vollmers HP, Brandlein S (2006) Natural IgM antibodies: the orphaned molecules in immune
surveillance. Adv Drug Deliv Rev 58(5–6):755–765
Von Behren J, Kreutzer R, Smith D (1999) Asthma hospitalization trends in California, 1983–
1996. J Asthma 36(7):575–582
von Krogh C (1936) The morphology of the primate spleen. Anthropol Anz 13:89–100
Vugmeyster Y, Howell K, Bakshi A, Flores C, Hwang O, McKeever K (2004) B-cell subsets in
blood and lymphoid organs in Macaca fascicularis. Cytometry A 61(1):69–75
Waddell DR, Duffy KT (1986) Breakdown of self/nonself recognition in cannibalistic strains of
the predatory slime mold, Dictyostelium caveatum. J Cell Biol 102(1):298–305
Wainberg MA, Beaupre S, Beiss B, Israel E (1983) Differential susceptibility of avian sarcoma
cells derived from different periods of tumor growth to natural killer cell activity. Cancer Res
43(10):4774–4780
Walker CM (1997) Comparative features of hepatitis C virus infection in humans and chimpan-
zees. Springer Semin Immunopathol 19(1):85–98
Walzer T, Jaeger S, Chaix J, Vivier E (2007) Natural killer cells: from CD3(-)NKp46(+) to post-
genomics meta-analyses. Curr Opin Immunol 19(3):365–372
Warncke M, Calzascia T, Coulot M, Balke N, Touil R, Kolbinger F, Heusser C (2012) Different
adaptations of IgG effector function in human and nonhuman primates and implications for
therapeutic antibody treatment. J Immunol 188(9):4405–4411
Warr GW, Magor KE, Higgins DA (1995) IgY: clues to the origins of modern antibodies. Immunol
Today 16(8):392–398
Webster RL, Johnson RP (2005) Delineation of multiple subpopulations of natural killer cells in
rhesus macaques. Immunology 115(2):206–214
Willard-Mack CL (2006) Normal structure, function, and histology of lymph nodes. Toxicol
Pathol 34(5):409–424
Wolfle U, Martin S, Emde M, Schempp C (2009) Dermatology in the Darwin anniversary. Part 2:
evolution of the skin-associated immune system. J Dtsch Dermatol Ges 7(10):862–869
Wong ES, Papenfuss AT, Heger A, Hsu AL, Ponting CP, Miller RD, Fenelon JC, Renfree MB,
Gibbs RA, Belov K (2011) Transcriptomic analysis supports similar functional roles for the
two thymuses of the tammar wallaby. BMC Genomics 12:420
Wood W, Jacinto A (2007) Drosophila melanogaster embryonic haemocytes: masters of multitask-
ing. Nat Rev Mol Cell Biol 8(7):542–551
Wu L, Liu YJ (2007) Development of dendritic-cell lineages. Immunity 26(6):741–750
Wu L, Bashirova AA, Martin TD, Villamide L, Mehlhop E, Chertov AO, Unutmaz D, Pope M,
Carrington M, KewalRamani VN (2002) Rhesus macaque dendritic cells efficiently transmit
primate lentiviruses independently of DC-SIGN. Proc Natl Acad Sci USA 99(3):
1568–1573
Wu PC, Chen JB, Kawamura S, Roos C, Merker S, Shih CC, Hsu BD, Lim C, Chang TW (2012)
The IgE gene in primates exhibits extraordinary evolutionary diversity. Immunogenetics
64(4):279–287
Yan L, Cerny RL, Cirillo JD (2004) Evidence that hsp90 is involved in the altered interactions of
Acanthamoeba castellanii variants with bacteria. Eukaryot Cell 3(3):567–578
Yoder JA (2004) Investigating the morphology, function and genetics of cytotoxic cells in bony
fish. Comp Biochem Physiol C Toxicol Pharmacol 138(3):271–280
Yoon J, Ponikau JU, Lawrence CB, Kita H (2008) Innate antifungal immunity of human eosino-
phils mediated by a beta 2 integrin, CD11b. J Immunol 181(4):2907–2915
Yoshida H, Ochiai M, Ashida M (1986) Beta-1,3-glucan receptor and peptidoglycan receptor are
present as separate entities within insect prophenoloxidase activating system. Biochem Biophys
Res Commun 141(3):1177–1184
Yoshimoto T, Yasuda K, Tanaka H, Nakahira M, Imai Y, Fujimori Y, Nakanishi K (2009) Basophils
contribute to T(H)2-IgE responses in vivo via IL-4 production and presentation of peptide-
MHC class II complexes to CD4+ T cells. Nat Immunol 10(7):706–712
Yousefi S, Gold JA, Andina N, Lee JJ, Kelly AM, Kozlowski E, Schmid I, Straumann A,
Reichenbach J, Gleich GJ et al (2008) Catapult-like release of mitochondrial DNA by eosino-
phils contributes to antibacterial defense. Nat Med 14(9):949–953
64 J.F. Brinkworth and M. Thorn
Zapata A, Amemiya CT (2000) Phylogeny of lower vertebrates and their immunological structures.
Curr Top Microbiol Immunol 248:67–107
Zemans RL, Colgan SP, Downey GP (2009) Transepithelial migration of neutrophils: mechanisms
and implications for acute lung injury. Am J Respir Cell Mol Biol 40(5):519–535
Zhang XW, Liu Q, Wang Y, Thorlacius H (2001) CXC chemokines, MIP-2 and KC, induce
P-selectin-dependent neutrophil rolling and extravascular migration in vivo. Br J Pharmacol
133(3):413–421
Zhang YA, Salinas I, Li J, Parra D, Bjork S, Xu Z, LaPatra SE, Bartholomew J, Sunyer JO (2010)
IgT, a primitive immunoglobulin class specialized in mucosal immunity. Nat Immunol
11(9):827–835
Zhao Y, Kacskovics I, Pan Q, Liberles DA, Geli J, Davis SK, Rabbani H, Hammarstrom L (2002)
Artiodactyl IgD: the missing link. J Immunol 169(8):4408–4416
Zhao Y, Pan-Hammarstrom Q, Yu S, Wertz N, Zhang X, Li N, Butler JE, Hammarstrom L (2006)
Identification of IgF, a hinge-region-containing Ig class, and IgD in Xenopus tropicalis. Proc
Natl Acad Sci USA 103(32):12087–12092
Genetic Variation in the Immune System
of Old World Monkeys: Functional
and Selective Effects
Introduction
The selective pressures exerted by pathogens and parasites have played a significant
role in human and nonhuman primate evolution, especially in shaping the form and
function of immune defenses (Barreiro and Quintana-Murci 2010; Nunn and Altizer
2006; Stearns and Koella 2008). While the evolutionary processes that historically
influenced the primate immune system cannot be directly observed today, the exten-
sive phenotypic diversity observed in contemporary primate populations reflects
these changes. Specifically, although many aspects of basic biology are conserved
among primates, both pathogen susceptibility and disease progression upon infec-
tion can greatly vary among species, subspecies, and populations. The adaptive and
mechanistic origins of this variation are therefore of great interest, including as a
source of insight into mechanisms of disease in humans.
Recent advances in genetic and genomic techniques, as well as steadily decreas-
ing costs, have enabled functional and evolutionary studies of intraspecies and inter-
species genetic variation in the primate immune system that would have been
D.A. Loisel
Department of Human Genetics, University of Chicago, 920 E 58th Street,
Chicago, IL 60637, USA
Department of Biology, Saint Michael’s College, Colchester, VT 05439
J. Tung (*)
Department of Evolutionary Anthropology, Duke University, Box 90383,
Durham, NC 27708, USA
Department of Human Genetics, University of Chicago, 920 E 58th Street,
Chicago, IL 60637, USA
Duke Institute for Population Research, Duke University, Box 90420,
Durham, NC 27708, USA
e-mail: [email protected]
prohibitive only a few decades ago. Thus, while interest in immune genetic varia-
tion is long-standing, research on this topic has become increasingly sophisticated.
It is now possible to not only identify which immune loci may have evolved under
selection (Bustamante et al. 2005; Haygood et al. 2007; Nielsen et al. 2005) but also
infer the type of selection that occurred, the time scale on which it occurred, and the
type of functional change (e.g., coding or regulatory) that was its target (Haygood
et al. 2010; Kosiol et al. 2008). At the same time, it is becoming increasingly routine
to use functional data to investigate the mechanistic and adaptive consequences of
sequence-level genetic variation.
The purpose of this chapter is to highlight recent development in these areas,
focusing specifically on research in Old World monkeys (OWMs). In what follows,
we briefly discuss the rationale for studying genetic variation in the OWM immune
system. We then review recent research on natural selection and functional genetic
variation in this clade pertaining to OWM immune function.
Many OWMs exhibit an unusually large geographic range relative to other primates
(macaques and baboons cover the largest geographic range of any primate genera,
other than humans) and also exhibit a high level of taxonomic diversity. As a result,
the OWM clade exhibits substantial ecological and phenotypic diversity. This diver-
sity extends to variation in susceptibility to infectious disease, including malaria
Genetic Variation in the Immune System of Old World Monkeys… 67
(Schmidt et al. 1977) and tuberculosis (Langermans et al. 2001), and is perhaps best
illustrated by the case of SIV/HIV. Surveys across OWM species indicate major
differences between the response to SIV in its natural African hosts and the response
to SIV in species that are not naturally exposed. Sooty mangabeys (Cercocebus
atys), a natural host for SIVsm (the strain from which HIV-2 in humans arose: Lemey
et al. 2003; Sharp et al. 1995), do not develop AIDS-like symptoms, despite robust
replication of the virus in the host upon infection. Comparisons between sooty
mangabeys and much more susceptible hosts, such as rhesus macaques, suggest that
the key to this difference may be reduced immune activation and/or reduced reli-
ance on T-cell-mediated immunity in the mangabeys (Silvestri et al. 2007). Infected
rhesus macaques also exhibit more dramatic T-cell depletion and SIV-specific anti-
body production in response to SIV infection than members of their sister species,
cynomolgus macaques (M. fascicularis) (Monceaux et al. 2007; Trichel et al. 2002).
Finally, variation in the response to SIV extends to the population level as well:
Chinese-origin rhesus macaques exhibit a slower rate of progression and a reduced
degree of immune activation compared to Indian-origin rhesus macaques (Ling
et al. 2002; Reimann et al. 2005).
OWMs are important behavioral and evolutionary models for humans and other spe-
cies that exhibit highly sophisticated social systems. Consequently, detailed longitudi-
nal and comparative data about the social structure, life history, and ecology of OWMs
in natural populations are available for many OWM species (for a partial summary of
these field studies, see Tung et al. 2010). These data permit the comparative study of
immune system evolution and its genetic effects in an ecological context, for example,
by relating socioecological characteristics that may influence exposure to, and spread
of, disease with molecular changes at immune loci. For instance, Nunn et al. (2000)
used such an approach in OWMs and other primates to show that species with more
promiscuous female mating patterns tended to have higher basal white blood cell
counts. They interpreted this result as a consequence of selection on the primate
immune system due to sexually transmitted disease. Indeed, in a follow-up study of 15
immune genes, Wlasiuk and Nachman (2010) identified stronger signatures of selec-
tion on branches leading to more promiscuous primate species, most notably for the
subset of genes whose function involves direct interactions with pathogens. Other
studies have focused on variation in the incidence of parasites and pathogens at differ-
ent geographic locations (Fooden 1994; Phillips-Conroy et al. 1988; Tung et al. 2009),
including how genetic effects contribute to these differences (Tung et al. 2009). In
humans, immune loci exhibit high Fst levels (a measure of the degree of overall genetic
variation explained by between-population differences) relative to the genome-wide
average, suggesting a history of local adaptation and selection (Akey et al. 2002;
Barreiro et al. 2008). Understanding when and whether this holds true in OWMs
could provide insight into how ecological differences between populations spur the
process of genetic differentiation between geographically separated groups.
Parasites and pathogens have long been considered to be primary drivers of evolu-
tion. Indeed, parasite or pathogen infection can have severe consequences for the
health of wild primates and can exert strong selective effects (Keele et al. 2009;
Knauf et al. 2012; Sapolsky and Else 1987). Thus, genetic variation that increases
an individual’s ability to evade, overcome, or minimize the consequences of infec-
tion is also likely to be adaptive. Genes involved in the immune system, which often
evolve rapidly (including in, but not only in, primates: Hughes et al. 2005; Schlenke
and Begun 2003), are a natural place to look for such variants.
In primates, this idea has primarily been supported by studies of a small number
of candidate loci, such as those in the major histocompatibility complex (MHC)
Genetic Variation in the Immune System of Old World Monkeys… 69
(see Table 1 for examples of these genes). In the past decade, however, the avail-
ability of genome sequences for multiple primate species, including the rhesus
macaque, has facilitated larger-scale genome-wide studies that provide insight into
the general importance of adaptive change in the primate immune system. Typically,
these studies attempt to infer the presence of lineage-specific positive selection by
comparing rates of synonymous and nonsynonymous site evolution in a gene’s cod-
ing sequence. When more than two species are included in such a comparison,
mutations in genes that appear to have evolved under selection can be localized to
specific branches within the species tree (Yang 1997). Two recent studies—a three-
species analysis of selection in primates (human, chimpanzee, and macaque: Gibbs
et al. 2007) and a six-species analysis of selection in these three primates and three
additional mammals (mouse, rat, and dog: Kosiol et al. 2008)—demonstrated that
positive selection has indeed played an unusually important role, relative to other
functional groups of genes, in the evolution of immune response genes in OWMs,
as well as in other mammals. Genes involved in the “immune response” were sig-
nificantly enriched among positively selected genes in both studies relative to other
classes of genes, whether considering selection over the entire three-species primate
tree (Gibbs et al. 2007; Kosiol et al. 2008) or on the individual branches leading to
each species, which included rhesus macaque (Kosiol et al. 2008). Many of these
genes fell into subclasses that have been the target of previous candidate gene stud-
ies, including those whose protein products interact with MHC class I (e.g., LILRB1,
LAIR) and those involved in T-cell-mediated immunity (e.g., CD3E, TCRA) (Gibbs
et al. 2007; Kosiol et al. 2008). Interestingly, the overall rate of nonsynonymous to
synonymous changes on the macaque lineage appears to be somewhat slower than
the comparable rate in humans and chimpanzees (Gibbs et al. 2007), perhaps reflect-
ing more efficient negative selection for genes evolving under constraint. The abil-
ity of selection to influence phenotypic variation is correlated with effective
population size, which is larger in macaques than in humans or chimpanzees
(Hernandez et al. 2007). Potentially, therefore, positive selection may also be more
efficient in macaques and other OWMs with large effective population sizes. If true,
such a scenario would increase power to detect selective events in OWMs once a
larger number of genome sequences become available.
Interestingly, genes that have undergone selection somewhere in the primate tree
often have evolved under selection across more than one lineage. Sequencing of the
rhesus macaque genome, for example, identified 67 genes that carried signatures of
positive selection on all three branches (rhesus macaque, human, and chimpanzee:
Gibbs et al. 2007). Similarly, Kosiol et al. (2008) estimated that the majority of the
genes they identified as positively selected in any lineage (including rodents and
dogs) were identified as positively selected in more than one lineage. For instance,
glycophorin C (GYPC), an erythrocytic cell surface antigen, is believed to have
evolved under recent selection in humans (Wilder et al. 2009) in response to patho-
gen pressure from the malarial parasite Plasmodium falciparum. However, GYPC
also appears to have been positively selected on all examined branches of the pri-
mate tree, suggesting that this gene has been the target of pathogen-mediated adap-
tive evolution for a much longer period of time (Kosiol et al. 2008). Several
70 D.A. Loisel and J. Tung
Table 1 (continued)
Immune Timescale of
Gene Function function Type of selection selection Reference
Tetherin Antiviral Innate Positive selection Across Old Liu et al. (2010);
restriction World Lim et al.
factor Monkeys (2010c)
TLR4, TLR5, Pathogen Innate Positive selection Across Old Nakajima et al.
and other recognition World (2008);
TLR genes Monkeys Wlasiuk
et al. (2009);
Wlasiuk and
Nachman
(2010)
TRIM5 Antiviral Innate Balancing selection Rhesus Newman et al.
restriction macaques (2006)
factor and Sooty
mangbeys
Evidence from genome-wide analysis and studies of individual genes and gene
families suggests that positive selection has helped shape the evolution of gene
protein-coding sequences in OWMs. However, protein-coding sequence accounts
for only a small percentage of the genome. Both theoretical arguments and empiri-
cal evidence suggest that other components of the genome, especially those involved
in gene regulation, may also have been targets of selection (King and Wilson 1975;
Wray 2007; Wray et al. 2003). In particular, a recent meta-analysis of six scans for
selection (all of which included human, chimpanzee, and macaque data; some
included sequence from additional mammalian species as well) revealed a signifi-
cant enrichment for selection on genes involved in T-cell-mediated immunity in
both the coding and noncoding regions of these genes (Haygood et al. 2010; see also
Torgerson et al. 2009).
72 D.A. Loisel and J. Tung
Balancing
selection
Diversification
via selection
No selection
(neutral)
No changes Changes in Changes in
in gene cis-regulatory coding
sequence sequence sequence
(continued)
Genetic Variation in the Immune System of Old World Monkeys… 73
Box 1 (continued)
2004; Semple et al. 2003), in contrast to many other innate immune system
genes (Mukherjee et al. 2009). In primates, evidence of positive selection has
been observed in genes from both the alpha and beta defensin families
(Boniotto et al. 2003; Das et al. 2010; Lynn et al. 2004; Semple et al. 2003),
and changes at positively selected sites have been shown to alter antimicrobial
activity in vitro (Antcheva et al. 2004).
Killer Cell Immunoglobulin-Like Receptors (KIRs): KIR genes encode
cell-surface receptors that interact with MHC class I genes to modulate natu-
ral killer cell and T-cell function, thus playing a central role in both innate and
adaptive immunity. Radiation of the KIR family, as well as genetic diversity
within specific KIR genes, may have evolved in response to either direct
interactions with pathogens or as a result of coevolution with their MHC class
I ligands (Hao and Nei 2005; Parham 1997). In OWMs, KIR family genes are
marked by extensive intraspecific allelic diversity and substantial interspecific
differences in KIR gene number and haplotype structure (Bimber et al. 2008;
Hershberger et al. 2005; Kruse et al. 2010; Palacios, et al. 2011).
Toll-Like Receptors (TLRs): TLRs are innate immune system genes that
are necessary for recognizing conserved patterns associated with pathogen
infection (e.g., lipopolysaccharide (LPS) in gram-negative bacteria). Although
TLRs share a general functional role, only some primate TLRs are associated
with strong signature of selection. In OWMs, for example, interspecific com-
parisons of TLR sequences in 10 genes indentified evidence for positive
selection at only three loci (TLR4, TLR1, and TLR8: Nakajima et al. 2008;
Wlasiuk and Nachman 2010), echoing variation in the selective histories of
this family in human populations (Barreiro et al. 2009).
TNF, a cytokine involved in acute inflammation (Baena et al. 2007). Recently, stud-
ies in cell lines derived from various primate species demonstrated widespread dif-
ferences in gene regulatory responses after LPS stimulation, some of which may be
a consequence of selection on gene regulation (Barreiro et al. 2010). Investigating
the role of regulatory sequence in the adaptive evolution of the immune system,
including how regulatory and coding changes may act in concert, remains an impor-
tant topic for future studies. Similarly, although maintenance of genetic diversity at
immune loci by means of balancing selection or frequency-dependent selection is
likely to be important (Charlesworth 2006; Dean et al. 2002; Garrigan and Hedrick
2003), we know relatively little about the extent to which this and other alternative
modes of selection play a role in OWM immune system evolution (genome-wide
studies have focused largely on positive selection; however, see Andres et al. 2009).
The generation of large-scale sequence data sets will be useful for addressing these
gaps in our current knowledge. For instance, relatively recent selective sweeps can be
localized using information about linkage disequilibrium (LD: the degree to which
genetic variants segregate non-independently due to physical linkage) across the
genome: recently selected loci exhibit more extensive LD with neighboring regions than
is typical genome-wide (Sabeti et al. 2006, 2007; Voight et al. 2006). Similarly, local
adaptation can be investigated by identifying genetic markers that show unusually high
levels of genetic differentiation between populations, relative to the genomic distribu-
tion of the same metric (Akey et al. 2004; Barreiro et al. 2008; Sabeti et al. 2006, 2007;
Voight et al. 2006). Both of these methods have identified a large number of candidates
for selection among human populations, including many immune-related loci (Barreiro
and Quintana-Murci 2010). Selective change on these time scales may also be common
in OWMs, especially among geographically widespread species. Indeed, functionally
important population-specific differences at MHC class I genes in rhesus macaques
(Otting et al. 2007) and at MHC class I (Kita et al. 2009) and class II genes in cynomol-
gus macaques (Ling et al. 2011; Sano et al. 2006) have already been described.
Additionally, in a broader comparison including a number of OWMs, Garamszegi and
Nunn (2011) found that both levels of MHC DRB allelic variation and rates of nonsyn-
onymous substitutions at this locus were correlated with parasite species richness
(defined as the total number of parasite species per host species) in 41 primates.
A signature of selection implies the presence, either in the past or in contemporary popu-
lations, of functional genetic variation at the same locus. Thus, data that establish the
molecular- and organism-level effects of selected variants act as important corroborating
evidence for the adaptive history of a given region. Perhaps more importantly, they also
provide biological insight into the fitness-related effects of specific allelic variants.
Although establishing the functional relevance of specific genetic variants
remains challenging, a number of methods are now well established, particularly
those that focus on the molecular mechanisms that act as intermediate links between
genotype and organism-level phenotypes (Box 2). The combination of selection and
Genetic Variation in the Immune System of Old World Monkeys… 75
variant
T
of one of the 4 A* HPILMFRSDAGMIL
A alleles (allelic
imbalance) Assess functional consequences
mRNA of ancestral sequence
(continued)
76 D.A. Loisel and J. Tung
Box 2 (continued)
behavior may not always reflect in vivo behavior (particularly when using
immortalized cells), reporter constructs are evaluated outside of the context of
normal chromatin architecture, cell lines that match the tissue and/or species
of interest may not be available, assays are low throughput.
Measuring Cis-Regulatory Effects via Allele-Specific Gene Expression
Assays In Vivo: Allele-specific expression assays quantify the relative abun-
dance of mRNA from the two alleles of a gene found within the same indi-
vidual (panel B) (Pastinen 2010; Yan et al. 2002). Differences in the expression
levels of these alleles imply the presence of variation in the cis-regulatory
region controlling that gene, which can be localized by testing for an associa-
tion between putative regulatory variants and the magnitude of allele-specific
gene expression across individuals (e.g., Tung et al. 2011).
Strengths: In vivo measurements reflect the natural behavior of cis-regula-
tory functional variants, amenable to studies of natural populations, environ-
mental and genetic background effects are controlled because comparisons
are conducted between alleles within individuals, potentially feasible on a
genome-wide scale (Fontanillas et al. 2010; Pastinen 2010). Limitations:
Putative regulatory variants are identified by association and are not experi-
mentally isolated, RNA samples may be difficult to obtain for the population
or tissue of interest, the magnitude of allele-specific effects may be modified
by epistatic or gene-environment interactions (de Meaux et al. 2005; Tung
et al. 2011; von Korff et al. 2009).
Reconstructing the Ancestral Forms and Functional Properties of
Modern Genes: Ancestral state reconstruction, which focuses on the infer-
ence and synthesis of sequences no longer found in contemporary popula-
tions (panel C), can be used to compare the functional consequences of
evolutionary changes on a longer time scale (Harms and Thornton 2010;
Thornton 2004). This approach permits the recreation of likely mutational
pathways leading to extant genes. See, for example, Goldschmidt et al. 2008;
Zhang and Rosenberg 2002.
Strengths: Permits experimental tests of the relationship between ancient
sequence changes and phenotypic change, series of reconstructed proteins can
indicate the probable sequence of mutations necessary to evolve novel func-
tions. Limitations: Requires gene-appropriate functional tests that do not nec-
essarily generalize well (i.e., assays for steroid hormone binding affinity work
only for steroid hormone receptors: Bridgham et al. 2009), dependent on the
accuracy of the inference method used to determine ancestral states, does not
take into account dependence of the focal gene’s function on other genes in the
ancestral genome or on the ancestral environment (Harms and Thornton 2010).
Genetic Variation in the Immune System of Old World Monkeys… 77
Genes of the MHC function in the presentation of self and nonself peptides to T
cells and thus play a central role in the recognition of and response to pathogens and
parasites. Extensive study of MHC sequence diversity suggests that selection has
often contributed to shaping genetic variation at these loci (reviewed in Apanius
et al. 1997; Garrigan and Hedrick 2003; Hughes and Yeager 1998; Klein 1986;
Meyer and Thomson 2001). Evidence for adaptive evolution at OWM MHC loci
takes several different forms. At the population level, researchers have observed
allele frequency distributions that are inconsistent with those predicted by neutral
evolution (cynomolgus macaques: Bonhomme et al. 2007). On a deeper time scale,
comparisons of closely related species have revealed that some MHC polymor-
phisms seem to have been maintained since prior to divergence from the species’
common ancestor. These “trans-species” polymorphisms are unusual among dis-
tinct species but are relatively common among both class I (Otting et al. 2007) and
class II MHC genes (Doxiadis et al. 2006; Huchard et al. 2006; Loisel 2007), sug-
gesting that selection has acted to maintain segregating variation over long periods
of time (Klein et al. 2007). Finally, molecular evolution studies have shown that the
exons encoding the antigen-binding regions in MHC proteins consistently exhibit
elevated nonsynonymous to synonymous substitution ratios; this hallmark of strong
positive selection has been observed at several MHC class II genes in baboons
(Huchard et al. 2008; Loisel 2007) and in rhesus macaque class I genes (Urvater
et al. 2000). As MHC diversity on all three of these levels may contribute to
adaptively relevant phenotypic variation, these signatures of selection may harbor
important clues to the location of functionally important genetic variants.
For instance, Loisel et al. (2006) demonstrated that MHC population genetic
diversity is associated with functional consequences for gene regulation. The pro-
moter region of the classical MHC class II gene DQA1 is highly polymorphic in
many primate species, including yellow baboons, rhesus macaques, and pigtail
macaques, and shows evidence of trans-species polymorphism (Loisel et al. 2006).
As a result, the phylogenetic relationships among alleles for this region (the “gene
tree”) are not congruent with the phylogenetic relationships between OWM species
(the “species tree”), a pattern that is particularly unusual over such extended periods
of evolutionary time (~27 million years; Fig. 1). To assess the functional effects of
DQA1 promoter region variation, Loisel et al. (2006) cloned 12 of the alleles found
78 D.A. Loisel and J. Tung
Fig. 1 Evolutionary relationships of MHC-DQA1 exon 2 coding sequence in Old World monkeys.
(a) The allelic genealogy for MHC-DQA1 exon 2 sequences (containing the antigen-binding
region) in 12 species of Old World monkeys (OWMs) is characterized by the long-term mainte-
nance of ancestral allelic lineages across speciation events, a phenomenon known as trans-species
polymorphism. The incongruity between the DQA1 exon 2 gene tree and the OWM species tree
(shown in b.) reflects the presence of this trans-species polymorphism, in contrast to the typical
expected relationship between gene and species trees (shown in inset c.). The relationship between
DQA1 exon 2 sequences was inferred using neighbor-joining with evolutionary distances com-
puted using the Tamura-Nei method. The percentage of replicate trees in which the associated taxa
clustered together in the bootstrap test (1,000 replicates) is shown next to the branches. DQA1 exon
2 sequences were obtained from the IPD-MHC nonhuman primate database (Robinson et al. 2003)
Genetic Variation in the Immune System of Old World Monkeys… 79
primates (Gibbs et al. 2007; Han et al. 2007), suggesting that retroviruses have
exerted significant selective pressures on primate evolution for a very long time.
The evolution of the antiretroviral gene, tripartite motif protein 5 (TRIM5), illus-
trates the possible effects of these pressures. TRIM5 encodes a cytosolic protein
that restricts infection by HIV-1 and other retroviruses by interacting with retro-
virus capsid lattice (Stremlau et al. 2006). Through this interaction, TRIM5
restricts retroviral infection by acting as a pattern recognition receptor for the
capsid lattice and activating general innate immune signaling pathways (Pertel
et al. 2011). The TRIM5 gene is marked by numerous signatures of ancient and
recurrent natural selection within primates (Liu et al. 2005; Ortiz et al. 2006;
Sawyer et al. 2005, 2007), including both lineage-specific positive selection
(Johnson and Sawyer 2009) and long-term balancing selection, indicated by the
presence of trans-specific polymorphism between sooty mangabeys and rhesus
macaques (Newman et al. 2006) and between humans and chimpanzees (Cagliani
et al. 2010).
Research on natural selection at TRIM5 has been instrumental for understanding
the function of genetic diversity in the gene, which is high in both OWMs and
humans (Cagliani et al. 2010; Dietrich et al. 2010; Newman et al. 2006). For
instance, efforts to localize the specific nucleotide targets of positive selection in
TRIM5 indicated that two specific regions of protein, known as the B30.2 and
coiled-coil domains, were likely essential to its antiviral activity (Sawyer et al.
2005; Song et al. 2005b). Indeed, the ability of the human TRIM5α protein to
restrict the replication of retroviruses (including HIV-1, SIVagm, and murine leuke-
mia viruses) in vitro was significantly enhanced by mutating only a few amino
acids in these regions to the version found in rhesus macaques (Sawyer et al. 2005;
Yap et al. 2005). Thus, detailed analysis of selection at TRIM5 identified at least
some of the sites likely associated with species-specific differences in its ability to
restrict retroviral infection (Sawyer et al. 2005; Song et al. 2005a). Indeed, long-
term maintenance of TRIM5 alleles in rhesus macaques and sooty mangabeys may
have roots in retroviral resistance: allelic variation segregating within these species
also explains differences in the ability to restrict in vitro retroviral infection
(Newman et al. 2006; Wilson et al. 2008). This effect likely explains the finding
that specific TRIM5 alleles in rhesus macaques are associated with differences in
plasma viral load and disease progression following experimental SIV infection
(Lim et al. 2010a, b).
Finally, immune-related genetic variation may also have selective and functional
consequences observable within or between contemporary primate populations.
For instance, a recent study of a natural baboon population in Kenya identified a
significant association between genetic variation at the FY (DARC) locus and the
probability of infection by the hematoprotozoan parasite Hepatocystis (a close
Genetic Variation in the Immune System of Old World Monkeys… 81
Conclusions
Taken together, studies thus far on the genetics of OWM immunity reinforce a gen-
eral theme in evolutionary immunogenetics: adaptive change in the immune system
is pervasive, is recurrent, and takes a variety of forms. Specifically, studies of indi-
vidual immune-related genes reveal that selective regimes include both directional
selection and selection for maintenance of genetic variation (sometimes at the same
locus); that functional and/or selective divergence are apparent on multiple levels of
comparisons, including within populations, between populations, between species,
and between larger taxonomic groups; and that functional and selective change can
target both coding and noncoding regions.
We anticipate that newly developed high-throughput sequencing approaches will
make important contributions to extending this work, particularly functional anno-
tation of selectively important genetic changes. These methods allow data on
sequence variation and quantitative measurements of genome function, including
82 D.A. Loisel and J. Tung
References
Burton PR, Clayton DG, Cardon LR, Craddock N, Deloukas P, Duncanson A, Kwiatkowski DP,
McCarthy MI, Ouwehand WH, Samani NJ et al (2007) Genome-wide association study of 14,000
cases of seven common diseases and 3,000 shared controls. Nature 447(7145): 661–678
Bustamante CD, Fledel-Alon A, Williamson S, Nielsen R, Hubisz MT, Glanowski S, Tanenbaum
DM, White TJ, Sninsky JJ, Hernandez RD et al (2005) Natural selection on protein-coding
genes in the human genome. Nature 437(7062):1153–1157
Cagliani R, Fumagalli M, Biasin M, Piacentini L, Riva S, Pozzoli U, Bonaglia MC, Bresolin N,
Clerici M, Sironi M (2010) Long-term balancing selection maintains trans-specific polymor-
phisms in the human TRIM5 gene. Hum Genet 128(6):577–588
Cain C, Bleckman R, Marioni JC, Gilad Y (2011) Gene expression differences among primates are
associated with changes in a histone epigenetic modification. Genetics 187(4):1225–1234
Charlesworth D (2006) Balancing selection and its effects on sequences in nearby genomic
regions. PLoS Genet 2(4):e64
Chorley BN, Wang X, Campbell MR, Pittman GS, Noureddine MA, Bell DA (2008) Discovery
and verification of functional single nucleotide polymorphisms in regulatory genomic regions:
current and developing technologies. Mutat Res 659(1–2):147–157
Comuzzie AG, Cole SA, Martin L, Carey KD, Mahaney MC, Blangero J, VandeBerg JL (2003)
The baboon as a nonhuman primate model for the study of the genetics of obesity. Obes Res
11(1):75–80
Das S, Nikolaidis N, Goto H, McCallister C, Li JX, Hirano M, Cooper MD (2010) Comparative
genomics and evolution of the alpha-defensin multigene family in primates. Mol Biol Evol
27(10):2333–2343
Daza-Vamenta R, Glusman G, Rowen L, Guthrie B, Geraghty DE (2004) Genetic divergence of
the rhesus macaque major histocompatibility complex. Genome Res 14:1501–1515
de Meaux J, Goebel U, Pop A, Mitchell-Olds T (2005) Allele-specific assay reveals functional
variation in the chalcone synthase promoter of Arabidopsis thaliana that is compatible with
neutral evolution. Plant Cell 17(3):676–690
Dean M, Carrington M, O’Brien SJ (2002) Balanced polymorphism selected by genetic versus
infectious human disease. Annu Rev Genomics Hum Genet 3:263–292
Dietrich EA, Jones-Engel L, Hu SL (2010) Evolution of the antiretroviral restriction factor
TRIMCyp in Old World Primates. PLoS One 5(11):e14019
Doxiadis GGM, Rouweler AJM, de Groot NG, Louwerse A, Otting N, Verschoor EJ, Bontrop RE
(2006) Extensive sharing of MHC class II alleles between rhesus and cynomolgus macaques.
Immunogenetics 58(4):259–268
Elde NC, Child SJ, Geballe AP, Malik HS (2009) Protein kinase R reveals an evolutionary model
for defeating viral mimicry. Nature 457(7228):485–488
Fellay J, Shianna KV, Ge DL, Colombo S, Ledergerber B, Weale M, Zhang KL, Gumbs C,
Castagna A, Cossarizza A et al (2007) A whole-genome association study of major determi-
nants for host control of HIV-1. Science 317(5840):944–947
Fernandez S, Wassmuth R, Knerr I, Frank C, Haas JP (2003) Relative quantification of HLA-
DRA1 and -DQA1 expression by real-time reverse transcriptase-polymerase chain reaction
(RT-PCR). Eur J Immunogenet 30:141–148
Fernando MMA, Stevens CR, Walsh EC, De Jager PL, Goyette P, Plenge RM, Vyse TJ, Rioux JD
(2008) Defining the role of the MHC in autoimmunity: a review and pooled analysis. PLoS
Genet 4(4):e1000024
Filip LC, Mundy NI (2004) Rapid evolution by positive Darwinian selection in the extracellular
domain of the abundant lymphocyte protein CD45 in primates. Mol Biol Evol 21: 1504–1511
Fontanillas P, Landry CR, Wittkopp PJ, Russ C, Gruber JD, Nusbaum C, Hartl DL (2010) Key
considerations for measuring allelic expression on a genomic scale using high-throughput
sequencing. Mol Ecol 19:212–227
Fooden J (1994) Malaria in macaques. Int J Primatol 15(4):573–596
Garamszegi LZ, Nunn CL (2011) Parasite-mediated evolution of the functional part of the MHC
in primates. J Evol Biol 24(1):184–195
Gardner MB, Luciw PA (2008) Macaque models of human infectious disease. ILAR J 49(2):220–255
Genetic Variation in the Immune System of Old World Monkeys… 85
Keele BF, Jones JH, Terio KA, Estes JD, Rudicell RS, Wilson ML, Li YY, Learn GH, Beasley TM,
Schumacher-Stankey J et al (2009) Increased mortality and AIDS-like immunopathology in
wild chimpanzees infected with SIVcpz. Nature 460(7254):515–519
King M-C, Wilson ACC (1975) Evolution at two levels in humans and chimpanzees. Science
188(4184):107–116
Kita YF, Hosomichi K, Kohara S, Itoh Y, Ogasawara K, Tsuchiya H, Torii R, Inoko H, Blancher A,
Kulski JK et al (2009) MHC class I A loci polymorphism and diversity in three southeast Asian
populations of cynomolgus macaque. Immunogenetics 61(9):635–648
Klein J (1986) Natural history of the major histocompatibility complex. Wiley, New York, NY
Klein J, Sato A, Nikolaidis N (2007) MHC, TSP, and the origin of species: from immunogenetics
to evolutionary genetics. Annu Rev Genet 41:281–304
Knauf S, Batamuzi EK, Mlengeya T, Kilewo M, Lejora IA, Nordhoff M, Ehlers B, Harper KN,
Fyumagwa R, Hoare R et al (2012) Treponema infection associated with genital ulceration in
wild baboons. Vet Pathol 49(2):292–303
Knight JC (2003) Functional implications of genetic variation in non-coding DNA for disease
susceptibility and gene regulation. Clin Sci 104(5):493–501
Kosiol C, Vinar T, da Fonseca RR, Hubisz MJ, Bustamante CD, Nielsen R, Siepel A (2008)
Patterns of positive selection in six mammalian genomes. PLoS Genet 4(8):e1000144
Kruse PH, Rosner C, Walter L (2010) Characterization of rhesus macaque KIR genotypes and
haplotypes. Immunogenetics 62(5):281–293
Kulski JK, Anzai T, Shiina T, Inoko H (2004) Rhesus macaque class I duplicon structures, organi-
zation, and evolution within the alpha block of the major histocompatibility complex. Mol Biol
Evol 21(11):2079–2091
Lalani AS, Masters J, Zeng W, Barrett J, Pannu R, Everett H, Arendt CW, McFadden G (1999) Use
of chemokine receptors by poxviruses. Science 286(5446):1968–1971
Langermans JAM, Andersen P, van Soolingen D, Vervenne RAW, Frost PA, van der Laan T, van
Pinxteren LAH, van den Hombergh J, Kroon S, Peekel I et al (2001) Divergent effect of bacil-
lus Calmette-Guerin (BCG) vaccination on Mycobacterium tuberculosis infection in highly
related macaque species: implications for primate models in tuberculosis vaccine research.
Proc Natl Acad Sci U S A 98(20):11497–11502
Lehrer RI (2004) Primate defensins. Nat Rev Microbiol 2(9):727–738
Lemey P, Pybus OG, Wang B, Saksena NK, Salemi M, Vandamme AM (2003) Tracing the origin
and history of the HIV-2 epidemic. Proc Natl Acad Sci U S A 100(11):6588–6592
Leuchte N, Berry N, Kohler B, Almond N, LeGrand R, Thorstensson R, Titti F, Sauermann U
(2004) MhcDRB-sequences from cynomolgus macaques (Macaca fascicularis) of different
origin. Tissue Antigens 63(6):529–537
Lim SY, Chan T, Gelman RS, Whitney JB, O’Brien KL, Barouch DH, Goldstein DB, Haynes BF,
Letvin NL (2010a) Contributions of Mamu-a*01 status and TRIM5 allele expression, but Not
CCL3L copy number variation, to the control of SIVmac251 replication in Indian-origin rhesus
monkeys. PLoS Genet 6(6):e1000997
Lim SY, Rogers T, Chan T, Whitney JB, Kim J, Sodroski J, Letvin NL (2010b) TRIM5 alpha
modulates immunodeficiency virus control in rhesus monkeys. PLoS Pathog 6(1):e1000738
Lim ES, Malik HS, Emerman M (2010c) Ancient adaptive evolution of tetherin shaped the func-
tions of Vpu and Nef in human immunodeficiency virus and primate lentiviruses. J Virol 84:
7124–7134
Ling BH, Veazey RS, Luckay A, Penedo C, Xu KY, Lifson JD, Marx PA (2002) SIVmac pathogen-
esis in rhesus macaques of Chinese and Indian origin compared with primary HIV infections in
humans. AIDS 16(11):1489–1496
Ling F, Wei LQ, Wang T, Wang HB, Zhuo M, Du HL, Wang JF, Wang XN (2011) Characterization
of the major histocompatibility complex class II DOB, DPB1, and DQB1 alleles in cynomol-
gus macaques of Vietnamese origin. Immunogenetics 63(3):155–166
Liu HF, Wang YQ, Liao CH, Kuang YQ, Zheng YT, Su B (2005) Adaptive evolution of primate
TRIM5 alpha, a gene restricting HIV-1 infection. Gene 362:109–116
Genetic Variation in the Immune System of Old World Monkeys… 87
Liu Y, Helms C, Liao W, Zaba LC, Duan S, Gardner J, Wise C, Miner A, Malloy MJ, Pullinger CR
et al (2008) A genome-wide association study of psoriasis and psoriatic arthritis identifies new
disease loci. PLoS Genet 4(4):e1000041
Liu J, Chen K, Wang J-H, Zhang C (2010) Molecular evolution of the primate antiviral restriction
factor tetherin. PLoS ONE 5:e11904
Loisel DA (2007) Evolutionary genetics of immune system genes in a wild primate population.
Duke University, Durham, NC
Loisel DA, Rockman MV, Wray GA, Altmann J, Alberts SC (2006) Ancient polymorphism and
functional variation in the primate MHC-DQA1 5′ cis-regulatory region. Proc Natl Acad Sci U
S A 103:16331–16336
Lynn DJ, Lloyd AT, Fares MA, O’Farrelly C (2004) Evidence of positively selected sites in mam-
malian alpha-defensins. Mol Biol Evol 21(5):819–827
McDade TW (2003) Life history theory and the immune system: steps toward a human ecological
immunology. Am J Phys Anthropol 46:100–125
Messier W, Stewart CB (1997) Episodic adaptive evolution of primate lysozymes. Nature
385(6612):151–154
Meyer D, Thomson G (2001) How selection shapes variation of the human major histocompatibil-
ity complex: a review. Ann Hum Genet 65:1–26
Monceaux V, Viollet L, Petit F, Cumont MC, Kaufmann GR, Aubertin AM, Hurtrel B, Silvestri G,
Estaquier J (2007) CD4(+) CCR5(+) T-cell dynamics during simian immunodeficiency virus
infection of Chinese rhesus macaques. J Virol 81(24):13865–13875
Morzycka-Wroblewska E, Munshi A, Ostermayer M, Harwood JI, Kagnoff MF (1997) Differential
expression of HLA-DQA1 alleles associated with promoter polymorphism. Immunogenetics
45:163–170
Mukherjee S, Sarkar-Roy N, Wagener DK, Majumder PP (2009) Signatures of natural selection
are not uniform across genes of innate immune system, but purifying selection is the dominant
signature. Proc Natl Acad Sci U S A 106(17):7073–7078
Mummidi S, Bamshad M, Ahuja SS, Gonzalez E, Feuillet PM, Begum K, Galvis MC, Kostecki V,
Valente AJ, Murthy KK et al (2000) Evolution of human and non-human primate CC chemo-
kine receptor 5 gene and mRNA - potential roles for haplotype and mRNA diversity, differen-
tial haplotype-specific transcriptional activity, and altered transcription factor binding to
polymorphic nucleotides in the pathogenesis of HIV-1 and simian immunodeficiency virus.
J Biol Chem 275(25):18946–18961
Nakajima T, Ohtani H, Satta Y, Uno Y, Akari H, Ishida T, Kimura A (2008) Natural selection in the
TLR-related genes in the course of primate evolution. Immunogenetics 60:727–735
Neil S, Bieniasz P (2009) Human immunodeficiency virus, restriction factors, and interferon.
J Interferon Cytokine Res 29(9):569–580
Newman RM, Hall L, Connole M, Chen GL, Sato S, Yuste E, Diehl W, Hunter E, Kaur A, Miller
GM et al (2006) Balancing selection and the evolution of functional polymorphism in Old
world monkey TRIM5 alpha. Proc Natl Acad Sci U S A 103(50):19134–19139
Nielsen R, Bustamante C, Clark AG, Glanowski S, Sackton TB, Hubisz MJ, Fledel-Alon A,
Tanenbaum DM, Civello D, White TJ et al (2005) A scan for positively selected genes in the
genomes of humans and chimpanzees. PLoS Biol 3(6):976–985
Nunn CL, Altizer SM (2006) Infectious diseases in primates: behavior, ecology and evolution.
Oxford University Press, Oxford
Nunn CL, Gittleman JL, Antonovics J (2000) Promiscuity and the primate immune system.
Science 290(5494):1168–1170
Ober C (1999) Studies of HLA, fertility and mate choice in a human isolate. Hum Reprod Update
5(2):103–107
Ober C, Hyslop T, Elias S, Weitkamp LR, Hauck WW (1998) Human leukocyte antigen matching
and fetal loss: results of a 10 year prospective study. Hum Reprod 13(1):33–38
OhAinle M, Kerns JA, Malik HS, Emerman M (2006) Adaptive evolution and antiviral activity of
the conserved mammalian cytidine deaminase APOBEC3H. J Virol 80(8):3853–3862
88 D.A. Loisel and J. Tung
Sapolsky RM, Else JG (1987) Bovine tuberculosis in a wild baboon population - epidemiologic
aspects. J Med Primatol 16(4):229–235
Sawyer SL, Emerman M, Malik HS (2004) Ancient adaptive evolution of the primate antiviral
DNA-editing enzyme APOBEC3G. PLoS Biol 2(9):1278–1285
Sawyer SL, Wu LI, Emerman M, Malik HS (2005) Positive selection of primate TRIM5 alpha
identifies a critical species-specific retroviral restriction domain. Proc Natl Acad Sci U S A
102(8):2832–2837
Sawyer SL, Emerman M, Malik HS (2007) Discordant evolution of the adjacent antiretroviral
genes TRIM22 and TRIM5 in mammals. PLoS Pathog 3(12):1918–1929
Schad J, Ganzhorn JU, Sommer S (2005) Parasite burden and constitution of major histocompati-
bility complex in the Malagasy mouse lemur, Microcebus murinus. Evolution 59:439–450
Schaner P, Richards N, Wadhwa A, Aksentijevich I, Kastner D, Tucker P, Gumucio D (2001)
Episodic evolution of pyrin in primates: human mutations recapitulate ancestral amino acid
states. Nat Genet 27(3):318–321
Schlenke TA, Begun DJ (2003) Natural selection drives drosophila immune system evolution.
Genetics 164(4):1471–1480
Schmidt LH, Fradkin R, Harrison J, Rossan RN (1977) Differences in virulence of Plasmodium
knowlesi for macaca-irus (Fascicularis) of Philippine and Malayan origins. Am J Trop Med
Hyg 26(4):612–622
Schwensow N, Fietz J, Dausmann KH, Sommer S (2007) Neutral versus adaptive genetic variation
in parasite resistance: importance of major histocompatibility complex supertypes in a free-
ranging primate. Heredity 99(3):265–277
Semple CAM, Rolfe M, Dorin JR (2003) Duplication and selection in the evolution of primate
beta-defensin genes. Genome Biol 4(5):R31
Sharp PM, Robertson DL, Hahn BH (1995) Cross-species transmission and recombination of aids
viruses. Philos Trans R Soc Lond B Biol Sci 349(1327):41–47
Shiina T, Hosomichi K, Inoko H, Kulski JK (2009) The HLA genomic loci map: expression, inter-
action, diversity and disease. J Hum Genet 54(1):15–39
Silva JC, Kondrashov AS (2002) Patterns in spontaneous mutation revealed by human-baboon
sequence comparison. Trends Genet 18(11):544–547
Silvestri G, Paiardini M, Pandrea I, Lederman MM, Sodora DL (2007) Understanding the benign
nature of SIV infection in natural hosts. J Clin Invest 117(11):3148–3154
Sommer S (2005) The importance of immune gene variability (MHC) in evolutionary ecology and
conservation. Front Zool 2:16
Song B, Javanbakht H, Perron M, Park DH, Stremlau M, Sodroski J (2005a) Retrovirus restriction
by TRIM5 alpha variants from old world and new world primates. J Virol 79(7):3930–3937
Song BW, Gold B, O’hUigin C, Javanbakht H, Li X, Stremlau M, Winkler C, Dean M, Sodroski J
(2005b) The B30.2(SPRY) domain of the retroviral restriction factor TRIM5 alpha exhibits line
age-specific length and sequence variation in primates. J Virol 79(10):6111–6121
Stearns SC, Koella JC (eds) (2008) Evolution in health and disease, 2nd edn. Oxford University
Press, New York
Stremlau M, Perron M, Lee M, Li Y, Song B, Javanbakht H, Diaz-Griffero F, Anderson DJ,
Sundquist WI, Sodroski J (2006) Specific recognition and accelerated uncoating of retroviral
capsids by the TRIM5 alpha restriction factor. Proc Natl Acad Sci U S A 103(14):
5514–5519
Thornton JW (2004) Resurrecting ancient genes: experimental analysis of extinct molecules. Nat
Rev Genet 5(5):366–375
Torgerson DG, Boyko AR, Hernandez RD, Indap A, Hu XL, White TJ, Sninsky JJ, Cargill M,
Adams MD, Bustamante CD et al (2009) Evolutionary processes acting on candidate cis-
regulatory regions in humans inferred from patterns of polymorphism and divergence. PLoS
Genet 5(8):e1000592
Trichel AM, Rajakumar PA, Murphey-Corb M (2002) Species-specific variation in SIV disease
progression between Chinese and Indian subspecies of rhesus macaque. J Med Primatol
31(4–5):171–178
90 D.A. Loisel and J. Tung
Tung J, Primus A, Bouley AJ, Severson TF, Alberts SC, Wray GA (2009) Evolution of a malaria
resistance gene in wild primates. Nature 460(7253):388–391
Tung J, Alberts SC, Wray GA (2010) Evolutionary genetics in wild primates: combining genetic
approaches with field studies of natural populations. Trends Genet 26:353–362
Tung J, Akinyi MY, Mutura S, Altmann J, Wray GA, Alberts SC (2011) Allele-specific gene
expression in a wild nonhuman primate population. Mol Ecol 20(4):725–739
Urvater JA, Otting N, Loehrke JH, Rudersdorf R, Slukvin II, Piekarczyk MS, Golos TG, Hughes
AL, Bontrop RE, Watkins DI (2000) Mamu-I: a novel primate MHC class I B-related locus
with unusually low variability. J Immunol 164:1386–1398
Vallender EJ, Priddy CM, Chen GL, Miller GM (2008a) Human expression variation in the mu-
opioid receptor is paralleled in rhesus macaque. Behav Genet 38(4):390–395
Vallender EJ, Priddy CM, Hakim S, Yang H, Chen GL, Miller GM (2008b) Functional variation in
the 3′ untranslated region of the serotonin transporter in human and rhesus macaque. Genes
Brain Behav 7(6):690–697
Voight BF, Kudaravalli S, Wen XQ, Pritchard JK (2006) A map of recent positive selection in the
human genome. PLoS Biol 4(3):446–458
von Korff M, Radovic S, Choumane W, Stamati K, Udupa SM, Grando S, Ceccarelli S, Mackay I,
Powell W, Baum M et al (2009) Asymmetric allele-specific expression in relation to develop-
mental variation and drought stress in barley hybrids. Plant J 59(1):14–26
Wilder JA, Hewett EK, Gansner ME (2009) Molecular evolution of GYPC: evidence for recent
structural innovation and positive selection in humans. Mol Biol Evol 26(12):2679–2687
Wilson DE, Reeder DM (eds) (2005) Mammal species of the world: a taxonomic and geographic
reference, 3rd edn. The Johns Hopkins University Press, Baltimore, MD
Wilson SJ, Webb BLJ, Maplanka C, Newman RM, Verschoor EJ, Heeney JL, Towers GJ (2008) Rhesus
macaque TRIM5 alleles have divergent antiretroviral specificities. J Virol 82(14):7243–7247
Wiseman RW, Karl JA, Bimber BN, O’Leary CE, Lank SM, Tuscher JJ, Detmer AM, Bouffard P,
Levenkova N, Turcotte CL et al (2009) Major histocompatibility complex genotyping with
massively parallel pyrosequencing. Nat Med 15(11):1322–1326
Wittkopp P, Haerum B, Clark A (2004) Evolutionary changes in cis and trans gene regulation.
Nature 430:85–88
Wittkopp PJ, Haerum BK, Clark AG (2008) Independent effects of cis- and trans-regulatory varia-
tion on gene expression in drosophila melanogaster. Genetics 178(3):1831–1835
Wlasiuk G, Khan S, Switzer WM, Nachman MW (2009) A history of recurrent positive selection
at the Toll-Like Receptor 5 in primates. Mol Biol Evol 26:937–949
Wlasiuk G, Nachman MW (2010) Adaptation and constraint at toll-like receptors in primates. Mol
Biol Evol 27(9):2172–2186
Wray GA (2007) The evolutionary significance of cis-regulatory mutations. Nat Rev Genet
8(3):206–216
Wray GA, Hahn MW, Abouheif E, Balhoff JP, Pizer M, Rockman MV, Romano LA (2003) The
evolution of transcriptional regulation in eukaryotes. Mol Biol Evol 20(9):1377–1419
Yan H, Yuan WS, Velculescu VE, Vogelstein B, Kinzler KW (2002) Allelic variation in human
gene expression. Science 297(5584):1143
Yang Z (1997) PAML: a program package for phylogenetic analysis by maximum likelihood.
Comput Appl Biosci 13:555–556
Yap MW, Nisole S, Stoye JP (2005) A single amino acid change in the SPRY domain of human
Trim5 alpha leads to HIV-1 restriction. Curr Biol 15(1):73–78
Zhang JZ, Rosenberg HF (2002) Complementary advantageous substitutions in the evolution of an
antiviral RNase of higher primates. Proc Natl Acad Sci U S A 99(8):5486–5491
Zhang JZ, Webb DM (2004) Rapid evolution of primate antiviral enzyme APOBEC3G. Hum Mol
Genet 13(16):1785–1791
Toll-Like Receptor Function and Evolution
in Primates
Introduction
The immune system of vertebrate animals can be divided into the innate and the
adaptive immune responses. The innate immune system is an ancient, genetically
inherited and nonspecific response that forms the host’s first line of defense against
immune insult. The innate immune response is immediate, functions continuously,
requires no previous antigen exposure to be activated and is characterized by gener-
alized immune tactics such as inflammation, barriers (e.g., skin and mucosa), and
phagocytosis (reviewed in Janeway and Medzhitov 2002; Kumar et al. 2009). As the
immediate response to invading antigens and a modulator of adaptive immune
Toll-Like Receptor Function and Evolution in Primates 95
responses, innate immunity appears to exert great control over the course of infec-
tion. When excessive, the innate immune response can cause lethal pathologies in
the host (e.g., septic shock) (reviewed in Murdoch and Finn 2003). When weak or
disabled, initial infections grow unabated and adaptive responses exhibit dysfunc-
tion (Hagberg et al. 1984). As such, the innate immune system is an important
determinant of immune function and host survival.
Toll-like receptors represent one type of innate immune receptor and play a key
role in regulating innate immunity. The number of TLR types varies across animal
classes. Like some mammals, primates have ten functional TLRs. TLRs are horse-
shoe shaped, type 1 membrane proteins that contain three major functional domains:
an ecto- (extracellular) domain, a transmembrane domain, and a cytoplasmic
domain (Fig. 1). The ectodomain interacts directly with TLR ligands through a
series of leucine-rich repeats that facilitate protein–protein binding (Matsushima
et al. 2007). As type 1 membrane proteins, TLRs are bound to plasma membranes
and traverse either the cellular membrane (referred to as extracellular TLRs and
include TLR1, TLR2, TLR4, TLR5, TLR6, and TLR10) or endosomal membranes
(referred to as intracellular TLRs and include TLR3, TLR7, TLR8, and TLR9). The
cytoplasmic region of all TLRs, however, is located in the cytoplasm and contains
an ultra-conserved region referred to as the Toll/IL-1 receptor (TIR) domain that is
important for intracellular signaling.
The recognition of self from nonself is a primary function of normal mammalian
immunity. TLRs are immune system sentinels that enable phagocytic cells (i.e.,
moncytes/macrophages, neutrophils, and dendritic cells) to detect nonself or
foreign antigens and respond accordingly. Specifically, TLRs recognize many
molecular motifs (referred to as pathogen-associated molecular patterns or PAMPS)
that are generally absent in vertebrate hosts but broadly shared across microbial
organisms (Jin and Lee 2008; Kawai and Akira 2007). Although TLRs have been
traditionally viewed as specialized sensors of broadly shared microbial molecules
(Table 2), some receptors appear to be more promiscuous and use co-receptors to
recognize a wider range of motifs. This is particularly true of extracellular TLRs.
For example, when TLR2 and TLR4 form heterodimers with either other TLRs or
other co-receptors (e.g., LY96) they can recognize motifs (e.g., lipomannan from
Mycobacterium or lipopolysaccharide from Escherichia coli) that they cannot rec-
ognize as homodimers or single receptors (Gioannini et al. 2004; Shimazu et al.
1999; Zahringer et al. 2008).
TLRs become activated upon TLR–ligand binding. Once activated, the TIR
domain of the cytoplasmic region recruits adaptor proteins and triggers signaling
cascades that initiate downstream transcription factor binding (e.g., NF-κB and
IFR7) (Kawai and Akira 2007, 2010). These transcription factors then induce the
production of cytokines and costimulatory molecules (reviewed in Kawai and Akira
2007; Palm and Medzhitov 2009). Cytokines nonspecifically degrade antigens as
well as initiate the activation of the epitope-specific T and B cell-mediated adaptive
immune response. The engagement of the adaptive immune response generally
occurs after the 4th hour of infection, with clonal expansion of T and B cells initi-
ated more than 96 hours post-infection. TLR signaling pathways appear to help
96 J.F. Brinkworth and K.N. Sterner
Fig. 1 Toll-like receptor and co-receptor cellular locations in primates. TLR2 is shown here het-
erodimerized with co-receptors TLR1 and TLR6, TLR4 with co-receptor LY96. Some (i.e., TLR2
and TLR4), possibly all TLRs, habitually exist as homodimers
regulate this arm of adaptive immunity (e.g., T-helper 1 and 2 type responses)
(Dabbagh et al. 2002; Krieg 2007). TLRs, therefore, are not only key components
in the early recognition of microbial invaders but also play a role in modulating the
adaptive immune response. As a result, they have the potential to strongly influence
and shape disease outcomes.
Table 2 Toll-like receptor location, ligands, and representative organisms
Cellular location TLR PAMP type PAMP Organisms
Cell membrane TLR1/2 heterodimer Lipoproteins/lipopeptides/GPI anchors/ Pam3Cys lipopeptides (Aliprantis et al. bacteria, mycobacteria,
lipoglycans 1999), Lipomannan (Quesniaux et al. Gram-positive bacteria
2004; Vignal et al. 2003),
TLR2 Peptidoglycan, Gram-positive bacteria,
glycosylphosphatidylinositol-anchored trypanosomes, yeast
proteins, lipoteichoic acid, porins,
zymosan (Campos et al. 2001; Massari
et al. 2002; Schwandner et al. 1999;
Takeuchi et al. 1999; Underhill et al.
1999)
TLR2/6 heterodimer Pam2CYs lipopeptides, MALP-2, PSM, Mycobacteria, Borrelia
(Bulut et al. 2001; Buwitt-Beckmann burgdorferi, Gram-
et al. 2005) positive bacteria
TLR4 Lipopolysaccharide (LPS), fungal Gram-negative bacteria
Toll-Like Receptor Function and Evolution in Primates
Intra- and interspecies differences in TLR pathway function may explain within
and between species differences in susceptibility to TLR-detected pathogens.
Primates exhibit interspecies differences in immune response to a wide selection
of TLR-detected pathogens that are major agents of human diseases (Table 1).
Our current understanding of primate immune responses to such TLR-detected
pathogens is biased towards inferences drawn from DNA sequence data
(e.g., clinical association studies and phylogenetic analyses) and clinical/veteri-
narian observation, when available. Relatively few studies have directly exam-
ined primate TLR pathway function through controlled immunological challenge
experiments. Due to the clinical importance of such diseases as bacterial sepsis
and HIV/AIDS, most of these studies investigate primate TLR2, TLR4, and
TLR7. Moreover, controlled experiments almost entirely use catarrhine species
with studies biased towards the use of macaque, sooty mangabey, common chim-
panzee, baboon, and human models. Despite these limitations, however, current
data suggest both species- and family-level divergence in TLR pathway function.
Interestingly, a combination of controlled experimentation and veterinary/clinical
observations suggests a significant divergence between hominoid and cercopi-
thecoid species in response to TLR-detected pathogens in particular (Barreiro
et al. 2010; Brinkworth et al. 2012). Divergence in immune response has been
observed for a subset of TLR2-, TLR4-, and TLR7-detected pathogens dis-
cussed below.
Intraspecific Variation
Taken together, these studies suggest that differences observed in primate immune
responses may be influenced by divergent TLR pathway signaling, but further con-
trolled studies of TLR-mediated immune function are needed in this area. It is also
important to consider that intraspecific variation in TLRs and TLR signaling plays an
important role in mediating immune responses between individuals (Table 3). Host
genetics have been shown to affect infection susceptibility, strength of inflammation
and infectious disease progression (Allikmets et al. 1996; Hawn et al. 2007; Jallow
et al. 2009; Kang et al. 2002; Ogus et al. 2004). Polymorphisms in both TLR genes
(Table 3) and TLR pathway genes (e.g., Orange and Geha 2003; Puel et al. 2004) have
been associated with alterations in primate infectious disease manifestation, although
most association studies connecting infectious disease progression and TLR signaling
focus on human TLR genes and their co-receptors specifically. While these data are
invaluable, there are less comparable data for nonhuman primates. One recent excep-
tion examined the impact of TLR7 polymorphisms on the degree of AIDS-like symp-
toms in immunized and unimmunized SIV-infected rhesus macaques and found that
unimmunized rhesus macaques that were carriers for two TLR7 polymorphisms (c.13G
and c.-17C alleles) survived SIV infection longer than unimmunized noncarriers
(Siddiqui et al. 2011). These findings suggest TLR7 polymorphisms may influence not
only disease progression but also vaccine efficacy in rhesus macaques and humans.
Genes that encode proteins involved in the immune system are considered strong
targets of natural selection because of their relationship to an individual’s reproduc-
tive fitness and mortality (Barreiro and Quintana-Murci 2010; Nielsen et al. 2005).
102 J.F. Brinkworth and K.N. Sterner
Because TLRs play an integral part in the innate immune response, amino acid
substitutions that are deleterious to receptor function and decrease a host’s fitness
should be selected against, resulting in strong signals of purifying selection at the
DNA level (Roach et al. 2005). Using population genetics and evolutionary genom-
ics, it is possible to identify patterns of sequence divergence within and between
species, respectively, and infer what role evolutionary processes (e.g., natural selec-
tion, genetic drift, and gene flow) have played in shaping these differences. A num-
ber of recent studies have examined the protein-coding sequences of primate
Toll-like receptors for evidence of adaptive evolution (Areal et al. 2011; Nakajima
et al. 2008; Ortiz et al. 2008; Wlasiuk and Nachman 2010a). Briefly, statistical tests
of neutrality are used to examine if patterns of genetic variation observed between
species deviate from those expected given more neutral models of sequence change
(Barreiro and Quintana-Murci 2010). In order to do this, a ratio of nonsynonymous
(dN) to synonymous (dS) nucleotide substitution rates (dN/dS) is used as an indica-
tor of selective pressure acting on a protein-coding gene. Using maximum
likelihood-based approaches, it is possible to test alternative models that allow
dN/dS to vary across sites and/or lineages (Delport et al. 2010; Kumar et al. 2009;
Pond and Frost 2005; Yang 2007). Once indicated, natural selection can be classi-
fied as either purifying selection (i.e., selection for the removal of deleterious
alleles), or positive selection (i.e., selection for the fixation of advantageous alleles).
When dN/dS is equivalent to 1, neutral evolution is assumed. Purifying or positive
selection is indicated by dN/dS < 1 and dN/dS > 1, respectively. These types of inter-
specific tests are useful for detecting older evolutionary events (such as those rele-
vant to differences observed between species), whereas intraspecific tests of
neutrality (reviewed in Barreiro and Quintana-Murci 2010) are better able to detect
more recent selective events. Both approaches have been used to study the molecu-
lar evolution of primate Toll-like receptors.
Although primate TLR genes are generally conserved and show functional
constraint, individual TLR domains show different degrees of underlying genetic
variation (Areal et al. 2011; Nakajima et al. 2008; Ortiz et al. 2008; Wlasiuk
et al. 2009; Wlasiuk and Nachman 2010a). It is important to note these studies
include different primate species, which may explain some inconsistencies in
their findings. That being said, patterns emerge from these and similar studies
that further our understanding of TLR molecular evolution in primates. For
example, the TIR domain of TLRs is highly conserved across primates and is
under strong purifying selection due to its role in intracellular signaling (Areal
et al. 2011; Nakajima et al. 2008; Ortiz et al. 2008; Wlasiuk et al. 2009; Wlasiuk
and Nachman 2010a). The extracellular or ectodomain domain of many TLRs,
on the other hand, is more divergent due to its role in ligand binding (Areal et al.
2011; Wlasiuk et al. 2009; Wlasiuk and Nachman 2010a). These findings are not
necessarily primate specific and may reflect more general trends of TLR evolu-
tion observed throughout mammalian (Areal et al. 2011) and avian (Alcaide and
Edwards 2011) evolution.
When more lineage-specific models of sequence evolution are applied, the distri-
bution of positively selected sites does not vary significantly across primates except
104 J.F. Brinkworth and K.N. Sterner
In this chapter we review our current understanding of TLR function and evolution
in primates and discuss how an evolutionary perspective can help explain why some
primate species differ in susceptibility to certain pathogen-driven diseases. As we
move forward, it is important to combine studies that address how TLR-dependent
responses differ between primates with those that seek to understand the genetic and
environmental context in which these disparities exist. Currently, there is little evi-
dence that directly links species-specific immune responses (e.g., those described
above) to fixed, species-specific TLR sequence differences identified as adaptive in
certain primate lineages. While this may suggest that sequence-based approaches
lack power to detect adaptive substitutions, it may also suggest that a more holistic,
pathway-based approach is needed. This is especially important when trying to
determine how and why the human innate immune response can become dysregu-
lated following TLR activation.
References
Abe K, Kagei N, Teramura Y, Ejima H (1993) Hepatocellular carcinoma associated with chronic
Schistosoma mansoni infection in a chimpanzee. J Med Primatol 22(4):237–239
Agnese DM, Calvano JE, Hahm SJ, Coyle SM, Corbett SA, Calvano SE, Lowry SF (2002) Human
Toll-like receptor 4 mutations but not CD14 polymorphisms are associated with an increased
risk of gram-negative infections. J Infect Dis 186(10):1522–1525
Alcaide M, Edwards SV (2011) Molecular evolution of the Toll-like receptor multigene family in
birds. Mol Biol Evol 28(5):1703–1715
Alexopoulou L, Holt AC, Medzhitov R, Flavell RA (2001) Recognition of double-stranded RNA
and activation of NF-kappaB by Toll-like receptor 3. Nature 413(6857):732–738
Aliprantis AO, Yang RB, Mark MR, Suggett S, Devaux B, Radolf JD, Klimpel GR, Godowski P,
Zychlinsky A (1999) Cell activation and apoptosis by bacterial lipoproteins through Toll-like
receptor-2. Science 285(5428):736–739
Allikmets R, Buchbinder SP, Carrington M, Dean M, Detels R, Donfield S, Goedert JJ, Gomperts
E, Huttley GA, Kaslow R et al (1996) Genetic restrictions of HIV-1 infection and progression
to AIDS by a deletion allele of the CKR5 structural gene. Science 273:1856–1862
Arbour NC, Lorenz E, Schutte BC, Zabner J, Kline JN, Jones M, Frees K, Watt JL, Schwartz DA
(2000) TLR4 mutations are associated with endotoxin hyporesponsiveness in humans. Nat
Genet 25(2):187–191
Areal H, Abrantes J, Esteves PJ (2011) Signatures of positive selection in Toll-like receptor (TLR)
genes in mammals. BMC Evol Biol 11:368
Arko RJ (1989) Animal models for pathogenic Neisseria species. Clin Microbiol Rev
2(Suppl):S56–59
Bafica A, Santiago HC, Goldszmid R, Ropert C, Gazzinelli RT, Sher A (2006) Cutting edge: TLR9
and TLR2 signaling together account for MyD88-dependent control of parasitemia in
Trypanosoma cruzi infection. J Immunol 177(6):3515–3519
Barber RC, Chang LY, Arnoldo BD, Purdue GF, Hunt JL, Horton JW, Aragaki CC (2006) Innate
immunity SNPs are associated with risk for severe sepsis after burn injury. Clin Med Res
4(4):250–255
108 J.F. Brinkworth and K.N. Sterner
Barreiro LB, Quintana-Murci L (2010) From evolutionary genetics to human immunology: how
selection shapes host defence genes. Nat Rev Genet 11(1):17–30
Barreiro LB, Ben-Ali M, Quach H, Laval G, Patin E, Pickrell JK, Bouchier C, Tichit M, Neyrolles
O, Gicquel B et al (2009) Evolutionary dynamics of human Toll-like receptors and their differ-
ent contributions to host defense. PLoS Genet 5(7):e1000562
Barreiro LB, Marioni JC, Blekhman R, Stephens M, Gilad Y (2010) Functional comparison of
innate immune signaling pathways in primates. PLoS Genet 6(12):e1001249
Barrett RD, Hoekstra HE (2011) Molecular spandrels: tests of adaptation at the genetic level. Nat
Rev Genet 12(11):767–780
Ben-Ali M, Corre B, Manry J, Barreiro LB, Quach H, Boniotto M, Pellegrini S, Quintana-Murci L
(2011) Functional characterization of naturally occurring genetic variants in the human TLR1-
2-6 gene family. Hum Mutat 32(6):643–652
Benveniste RE, Arthur LO, Tsai CC, Sowder R, Copeland TD, Henderson LE, Oroszlan S (1986)
Isolation of a lentivirus from a macaque with lymphoma: comparison with HTLV-III/LAV and
other lentiviruses. J Virol 60(2):483–490
Bettauer RH (2010) Chimpanzees in hepatitis C virus research: 1998–2007. J Med Primatol
39(1):9–23
Biraben JN (1975) Les hommes et la peste en France et dans les pays européens et méditerranéens.
Mouton, Paris
Blackwell TS, Christman JW (1996) Sepsis and cytokines: current status. Br J Anaesth 77(1):
110–117
Bochud P-Y, Hersberger M, Taffe P, Bochud M, Stein CM, Rodrigues SD, Calandra T, Francioli P,
Telenti A, Speck RF et al (2007) Polymorphisms in Toll-like receptor 9 influence the clinical
course of HIV-1 infection. AIDS 21(4):441–446, 410.1097/QAD.1090b1013e328012b328018ac
Bochud PY, Hawn TR, Siddiqui MR, Saunderson P, Britton S, Abraham I, Argaw AT, Janer M,
Zhao LP, Kaplan G et al (2008) Toll-like receptor 2 (TLR2) polymorphisms are associated with
reversal reaction in leprosy. J Infect Dis 197(2):253–261
Bosinger SE, Sodora DL, Silvestri G (2011) Generalized immune activation and innate immune
responses in simian immunodeficiency virus infection. Curr Opin HIV AIDS 6(5):411–418
Bouer A, Werther K, Machado RZ, Nakaghi AC, Epiphanio S, Catao-Dias JL (2010) Detection of
anti-Toxoplasma gondii antibodies in experimentally and naturally infected non-human primates
by Indirect Fluorescence Assay (IFA) and indirect ELISA. Rev Bras Parasitol Vet 19(1):26–31
Brinkworth J, Pechenkina E, Silver J, Goyert S (2012) Innate immune responses to TLR2 and TLR4
agonists differ between baboons, chimpanzees and humans. J Med Primatol 41:388–393
Brown WJ, Lucas CT, Kuhn US (1972) Gonorrhoea in the chimpanzee. Infection with laboratory-
passed gonococci and by natural transmission. Br J Vener Dis 48(3):177–178
Bulut Y, Faure E, Thomas L, Equils O, Arditi M (2001) Cooperation of Toll-like receptor 2 and 6
for cellular activation by soluble tuberculosis factor and Borrelia burgdorferi outer surface
protein A lipoprotein: role of Toll-interacting protein and IL-1 receptor signaling molecules in
Toll-like receptor 2 signaling. J Immunol 167(2):987–994
Buwitt-Beckmann U, Heine H, Wiesmuller KH, Jung G, Brock R, Akira S, Ulmer AJ (2005) Toll-
like receptor 6-independent signaling by diacylated lipopeptides. Eur J Immunol 35(1):282–289
Campos MA, Almeida IC, Takeuchi O, Akira S, Valente EP, Procopio DO, Travassos LR, Smith
JA, Golenbock DT, Gazzinelli RT (2001) Activation of Toll-like receptor-2 by glycosylphos-
phatidylinositol anchors from a protozoan parasite. J Immunol 167(1):416–423
Catão-Dias JL, Epiphanio S, Martins Kierulff MC (2013) Neotropical primates and their suscepti-
bility to Toxoplasma gondii: new insights for an old problem. In: Brinkworth JF, Pechenkina E
(eds) Primates, pathogens, and evolution. Springer, Heidelberg
Cerf-Bensussan N, Gaboriau-Routhiau V (2010) The immune system and the gut microbiota:
friends or foes? Nat Rev Immunol 10(10):735–744
Dabbagh K, Dahl ME, Stepick-Biek P, Lewis DB (2002) Toll-like receptor 4 is required for optimal
development of Th2 immune responses: role of dendritic cells. J Immunol 168(9):4524–4530
Daniel MD, King NW, Letvin NL, Hunt RD, Sehgal PK, Desrosiers RC (1984) A new type D retro-
virus isolated from macaques with an immunodeficiency syndrome. Science 223(4636):602–605
Toll-Like Receptor Function and Evolution in Primates 109
Gage KL, Kosoy MY (2005) Natural history of plague: perspectives from more than a century of
research. Annu Rev Entomol 50:505–528
Galvani AP, Slatkin M (2003) Evaluating plague and smallpox as historical selective pressures for
the CCR5-Delta 32 HIV-resistance allele. Proc Natl Acad Sci USA 100(25):15276–15279
Garcia MA, Yee J, Bouley DM, Moorhead R, Lerche NW (2004) Diagnosis of tuberculosis in
macaques, using whole-blood in vitro interferon-gamma (PRIMAGAM) testing. Comp Med
54(1):86–92
Gioannini TL, Teghanemt A, Zhang D, Coussens NP, Dockstader W, Ramaswamy S, Weiss JP
(2004) Isolation of an endotoxin-MD-2 complex that produces Toll-like receptor 4-dependent
cell activation at picomolar concentrations. Proc Natl Acad Sci USA 101(12):4186–4191
Gottfried R (1983) The black death: natural and human disaster in medieval Europe. Free Press,
New York, p 203
Gutierrez MC, Brisse S, Brosch R, Fabre M, Omais B, Marmiesse M, Supply P, Vincent V (2005)
Ancient origin and gene mosaicism of the progenitor of Mycobacterium tuberculosis. PLoS
Pathog 1(1):e5
Haensch S, Bianucci R, Signoli M, Rajerison M, Schultz M, Kacki S, Vermunt M, Weston DA,
Hurst D, Achtman M et al (2010) Distinct clones of Yersinia pestis caused the black death.
PLoS Pathog 6(10):e1001134
Hagberg L, Hull R, Hull S, McGhee JR, Michalek SM, Svanborg EC (1984) Difference in suscep-
tibility to gram-negative urinary tract infection between C3H/HeJ and C3H/HeN mice. Infect
Immun 46(3):839–844
Haudek SB, Natmessnig BE, Furst W, Bahrami S, Schlag G, Redl H (2003) Lipopolysaccharide
dose response in baboons. Shock 20(5):431–436
Hawn TR, Verbon A, Lettinga KD, Zhao LP, Li SS, Laws RJ, Skerrett SJ, Beutler B, Schroeder L,
Nachman A et al (2003) A common dominant TLR5 stop codon polymorphism abolishes fla-
gellin signaling and is associated with susceptibility to legionnaires’ disease. J Exp Med
198(10):1563–1572
Hawn TR, Verbon A, Janer M, Zhao LP, Beutler B, Aderem A (2005) Toll-like receptor 4 polymor-
phisms are associated with resistance to Legionnaires’ disease. Proc Natl Acad Sci USA
102(7):2487–2489
Hawn TR, Misch EA, Dunstan SJ, Thwaites GE, Lan NT, Quy HT, Chau TT, Rodrigues S,
Nachman A, Janer M et al (2007) A common human TLR1 polymorphism regulates the innate
immune response to lipopeptides. Eur J Immunol 37(8):2280–2289
Hayashi F, Smith KD, Ozinsky A, Hawn TR, Yi EC, Goodlett DR, Eng JK, Akira S, Underhill DM,
Aderem A (2001) The innate immune response to bacterial flagellin is mediated by Toll-like
receptor 5. Nature 410(6832):1099–1103
Heil F, Hemmi H, Hochrein H, Ampenberger F, Kirschning C, Akira S, Lipford G, Wagner H,
Bauer S (2004) Species-specific recognition of single-stranded RNA via Toll-like receptor 7
and 8. Science 303(5663):1526–1529
Hemmi H, Kaisho T, Takeuchi O, Sato S, Sanjo H, Hoshino K, Horiuchi T, Tomizawa H, Takeda
K, Akira S (2002) Small anti-viral compounds activate immune cells via the TLR7 MyD88-
dependent signaling pathway. Nat Immunol 3(2):196–200
Hoshino K, Takeuchi O, Kawai T, Sanjo H, Ogawa T, Takeda Y, Takeda K, Akira S (1999) Cutting
edge: Toll-like receptor 4 (TLR4)-deficient mice are hyporesponsive to lipopolysaccharide:
evidence for TLR4 as the Lps gene product. J Immunol 162(7):3749–3752
Jallow M, Teo YY, Small KS, Rockett KA, Deloukas P, Clark TG, Kivinen K, Bojang KA, Conway
DJ, Pinder M et al (2009) Genome-wide and fine-resolution association analysis of malaria in
West Africa. Nat Genet 41(6):657–665
Janeway CA Jr, Medzhitov R (2002) Innate immune recognition. Annu Rev Immunol 20:197–216
Jin MS, Lee J-O (2008) Structures of the Toll-like receptor family and its ligand complexes.
Immunity 29(2):182–191
Johnson CM, Lyle EA, Omueti KO, Stepensky VA, Yegin O, Alpsoy E, Hamann L, Schumann RR,
Tapping RI (2007) Cutting edge: a common polymorphism impairs cell surface trafficking and
functional responses of TLR1 but protects against leprosy. J Immunol 178(12):7520–7524
Toll-Like Receptor Function and Evolution in Primates 111
Kang TJ, Lee SB, Chae GT (2002) A polymorphism in the Toll-like receptor 2 is associated with
IL-12 production from monocyte in lepromatous leprosy. Cytokine 20(2):56–62
Kawai T, Akira S (2007) TLR signaling. Semin Immunol 19(1):24–32
Kawai T, Akira S (2010) The role of pattern-recognition receptors in innate immunity: update on
Toll-like receptors. Nat Immunol 11(5):373–384
Keele BF, Jones JH, Terio KA, Estes JD, Rudicell RS, Wilson ML, Li Y, Learn GH, Beasley TM,
Schumacher-Stankey J et al (2009) Increased mortality and AIDS-like immunopathology in
wild chimpanzees infected with SIVcpz. Nature 460(7254):515–519
Krieg AM (2007) Antiinfective applications of Toll-like receptor 9 agonists. Proc Am Thorac Soc
4(3):289–294
Kumar H, Kawai T, Akira S (2009) Pathogen recognition in the innate immune response. Biochem
J 420(1):1–16
Kurt-Jones EA, Popova L, Kwinn L, Haynes LM, Jones LP, Tripp RA, Walsh EE, Freeman MW,
Golenbock DT, Anderson LJ et al (2000) Pattern recognition receptors TLR4 and CD14 medi-
ate response to respiratory syncytial virus. Nat Immunol 1(5):398–401
Kurt-Jones EA, Chan M, Zhou S, Wang J, Reed G, Bronson R, Arnold MM, Knipe DM, Finberg
RW (2004) Herpes simplex virus 1 interaction with Toll-like receptor 2 contributes to lethal
encephalitis. Proc Natl Acad Sci USA 101(5):1315–1320
Langermans JA, Andersen P, van Soolingen D, Vervenne RA, Frost PA, van der Laan T, van
Pinxteren LA, van den Hombergh J, Kroon S, Peekel I et al (2001) Divergent effect of bacillus
Calmette-Guerin (BCG) vaccination on Mycobacterium tuberculosis infection in highly related
macaque species: implications for primate models in tuberculosis vaccine research. Proc Natl
Acad Sci USA 98(20):11497–11502
Lorenz E, Mira JP, Frees KL, Schwartz DA (2002) Relevance of mutations in the TLR4 receptor
in patients with gram-negative septic shock. Arch Intern Med 162(9):1028–1032
Lucas CT, Chandler F Jr, Martin JE Jr, Schmale JD (1971) Transfer of gonococcal urethritis from
man to chimpanzee. An animal model for gonorrhea. JAMA 216(10):1612–1614
Lucotte G (2001) Distribution of the CCR5 gene 32-basepair deletion in West Europe. A hypoth-
esis about the possible dispersion of the mutation by the Vikings in historical times. Hum
Immunol 62(9):933–936
Lund JM, Alexopoulou L, Sato A, Karow M, Adams NC, Gale NW, Iwasaki A, Flavell RA (2004)
Recognition of single-stranded RNA viruses by Toll-like receptor 7. Proc Natl Acad Sci USA
101(15):5598–5603
Ma X, Liu Y, Gowen BB, Graviss EA, Clark AG, Musser JM (2007a) Full-exon resequencing
reveals Toll-like receptor variants contribute to human susceptibility to tuberculosis disease.
PLoS One 2(12):e1318
Ma Y, Haynes RL, Sidman RL, Vartanian T (2007b) TLR8: an innate immune receptor in brain,
neurons and axons. Cell Cycle 6(23):2859–2868
Major ME, Dahari H, Mihalik K, Puig M, Rice CM, Neumann AU, Feinstone SM (2004) Hepatitis
C virus kinetics and host responses associated with disease and outcome of infection in chim-
panzees. Hepatology 39(6):1709–1720
Mandl JN, Barry AP, Vanderford TH, Kozyr N, Chavan R, Klucking S, Barrat FJ, Coffman RL,
Staprans SI, Feinberg MB (2008) Divergent TLR7 and TLR9 signaling and type I interferon
production distinguish pathogenic and nonpathogenic AIDS virus infections. Nat Med
14(10):1077–1087
Mandl JN, Akondy R, Lawson B, Kozyr N, Staprans SI, Ahmed R, Feinberg MB (2011) Distinctive
TLR7 signaling, type I IFN production, and attenuated innate and adaptive immune responses
to yellow fever virus in a primate reservoir host. J Immunol 186(11):6406–6416
Martin R (2003) Earth history, disease, and the evolution of primates. In: Greenblatt C, Spigelmann
M (eds) Emerging pathogens: archaeology, ecology and evolution of infectious disease. Oxford
University Press, New York
Massari P, Henneke P, Ho Y, Latz E, Golenbock DT, Wetzler LM (2002) Cutting edge: immune
stimulation by neisserial porins is Toll-like receptor 2 and MyD88 dependent. J Immunol
168(4):1533–1537
112 J.F. Brinkworth and K.N. Sterner
Orange JS, Geha RS (2003) Finding NEMO: genetic disorders of NF-[kappa]B activation. J Clin
Invest 112(7):983–985
Ortiz M, Kaessmann H, Zhang K, Bashirova A, Carrington M, Quintana-Murci L, Telenti A (2008)
The evolutionary history of the CD209 (DC-SIGN) family in humans and non-human primates.
Genes Immun 9(6):483–492
Ozinsky A, Underhill DM, Fontenot JD, Hajjar AM, Smith KD, Wilson CB, Schroeder L, Aderem A
(2000) The repertoire for pattern recognition of pathogens by the innate immune system is defined
by cooperation between Toll-like receptors. Proc Natl Acad Sci USA 97(25): 13766–13771
Palm NW, Medzhitov R (2009) Pattern recognition receptors and control of adaptive immunity.
Immunol Rev 227(1):221–233
Pandrea I, Apetrei C (2010) Where the wild things are: pathogenesis of SIV infection in African
nonhuman primate hosts. Curr HIV/AIDS Rep 7(1):28–36
Park BS, Song DH, Kim HM, Choi BS, Lee H, Lee JO (2009) The structural basis of lipopolysac-
charide recognition by the TLR4-MD-2 complex. Nature 458(7242):1191–1195
Payne KS, Novak JJ, Jongsakul K, Imerbsin R, Apisitsaowapa Y, Pavlin JA, Hinds SB (2011)
Mycobacterium tuberculosis infection in a closed colony of rhesus macaques (Macaca mulatta).
J Am Assoc Lab Anim Sci 50(1):105–108
Plantinga TS, Ioana M, Alonso S, Izagirre N, Hervella M, Joosten LA, van der Meer JW, de la Rua
C, Netea MG (2012) The evolutionary history of TLR4 polymorphisms in Europe. J Innate
Immun 4(2):168–175
Pollitzer R (1951) Plague studies. 1. A summary of the history and survey of the present distribu-
tion of the disease. Bull World Health Organ 4(4):475–533
Poltorak A, He X, Smirnova I, Liu MY, Van Huffel C, Du X, Birdwell D, Alejos E, Silva M,
Galanos C et al (1998) Defective LPS signaling in C3H/HeJ and C57BL/10ScCr mice: muta-
tions in Tlr4 gene. Science 282(5396):2085–2088
Pond SL, Frost SD (2005) Datamonkey: rapid detection of selective pressure on individual sites of
codon alignments. Bioinformatics 21(10):2531–2533
Puel A, Picard C, Ku CL, Smahi A, Casanova JL (2004) Inherited disorders of NF-kappaB-
mediated immunity in man. Curr Opin Immunol 16(1):34–41
Quesniaux VJ, Nicolle DM, Torres D, Kremer L, Guerardel Y, Nigou J, Puzo G, Erard F, Ryffel B
(2004) Toll-like receptor 2 (TLR2)-dependent-positive and TLR2-independent-negative regula-
tion of proinflammatory cytokines by mycobacterial lipomannans. J Immunol 172(7):4425–4434
Qureshi ST, Lariviere L, Leveque G, Clermont S, Moore KJ, Gros P, Malo D (1999) Endotoxin-
tolerant mice have mutations in Toll-like receptor 4 (Tlr4). J Exp Med 189(4):615–625
Rayner JC, Liu W, Peeters M, Sharp PM, Hahn BH (2011) A plethora of Plasmodium species in
wild apes: a source of human infection? Trends Parasitol 27(5):222–229
Redl H, Bahrami S, Schlag G, Traber DL (1993) Clinical detection of LPS and animal models of
endotoxemia. Immunobiology 187(3–5):330–345
Roach JC, Glusman G, Rowen L, Kaur A, Purcell MK, Smith KD, Hood LE, Aderem A (2005) The
evolution of vertebrate Toll-like receptors. Proc Natl Acad Sci USA 102(27):9577–9582
Rutz M, Metzger J, Gellert T, Luppa P, Lipford GB, Wagner H, Bauer S (2004) Toll-like receptor
9 binds single-stranded CpG-DNA in a sequence- and pH-dependent manner. Eur J Immunol
34(9):2541–2550
Sadun EH, von Lichtenberg F, Cheever AW, Erickson DG (1970) Schistosomiasis mansoni in the
chimpanzee. The natural history of chronic infections after single and multiple exposures. Am
J Trop Med Hyg 19(2):258–277
Samson M, Libert F, Doranz BJ, Rucker J, Liesnard C, Farber CM, Saragosti S, Lapoumeroulie C,
Cognaux J, Forceille C et al (1996) Resistance to HIV-1 infection in caucasian individuals
bearing mutant alleles of the CCR-5 chemokine receptor gene. Nature 382(6593):722–725
Sapolsky RM, Else JG (1987) Bovine tuberculosis in a wild baboon population: epidemiological
aspects. J Med Primatol 16(4):229–235
Schott E, Witt H, Neumann K, Taube S, Oh DY, Schreier E, Vierich S, Puhl G, Bergk A, Halangk J
et al (2007) A Toll-like receptor 7 single nucleotide polymorphism protects from advanced inflam-
mation and fibrosis in male patients with chronic HCV-infection. J Hepatol 47(2):203–211
114 J.F. Brinkworth and K.N. Sterner
Schroder NW, Diterich I, Zinke A, Eckert J, Draing C, von Baehr V, Hassler D, Priem S, Hahn K,
Michelsen KS et al (2005) Heterozygous Arg753Gln polymorphism of human TLR-2 impairs
immune activation by Borrelia burgdorferi and protects from late stage Lyme disease.
J Immunol 175(4):2534–2540
Schuenemann VJ, Bos K, Dewitte S, Schmedes S, Jamieson J, Mittnik A, Forrest S, Coombes BK,
Wood JW, Earn DJ et al (2011) From the cover: targeted enrichment of ancient pathogens
yielding the pPCP1 plasmid of Yersinia pestis from victims of the Black Death. Proc Natl Acad
Sci USA 108(38):E746–752
Schuring RP, Hamann L, Faber WR, Pahan D, Richardus JH, Schumann RR, Oskam L (2009)
Polymorphism N248S in the human Toll-like receptor 1 gene is related to leprosy and leprosy
reactions. J Infect Dis 199(12):1816–1819
Schwandner R, Dziarski R, Wesche H, Rothe M, Kirschning CJ (1999) Peptidoglycan- and lipotei-
choic acid-induced cell activation is mediated by Toll-like receptor 2. J Biol Chem
274(25):17406–17409
Shi Q, Wang J, Wang XL, VandeBerg JL (2004) Comparative analysis of vascular endothelial cell
activation by TNF-alpha and LPS in humans and baboons. Cell Biochem Biophys 40(3):289–303
Shi Q, Cox LA, Glenn J, Tejero ME, Hondara V, Vandeberg JL, Wang XL (2010) Molecular path-
ways mediating differential responses to lipopolysaccharide between human and baboon arte-
rial endothelial cells. Clin Exp Pharmacol Physiol 37(2):178–184
Shimazu R, Akashi S, Ogata H, Nagai Y, Fukudome K, Miyake K, Kimoto M (1999) MD-2, a
molecule that confers lipopolysaccharide responsiveness on Toll-like receptor 4. J Exp Med
189(11):1777–1782
Siddiqui RA, Krawczak M, Platzer M, Sauermann U (2011) Association of TLR7 variants with
AIDS-like disease and AIDS vaccine efficacy in rhesus macaques. PLoS One 6(10):e25474
Sing A, Rost D, Tvardovskaia N, Roggenkamp A, Wiedemann A, Kirschning CJ, Aepfelbacher M,
Heesemann J (2002) Yersinia V-antigen exploits Toll-like receptor 2 and CD14 for interleukin
10-mediated immunosuppression. J Exp Med 196(8):1017–1024
Stephens JC, Reich DE, Goldstein DB, Shin HD, Smith MW, Carrington M, Winkler C, Huttley
GA, Allikmets R, Schriml L et al (1998) Dating the origin of the CCR5-Delta32 AIDS-
resistance allele by the coalescence of haplotypes. Am J Hum Genet 62(6):1507–1515
Sterner KN, Weckle A, Chugani HT, Tarca AL, Sherwood CC, Hof PR, Kuzawa CW, Boddy AM,
Abbas A, Raaum RL et al (2012) Dynamic gene expression in the human cerebral cortex dis-
tinguishes children from adults. PLoS One 7(5):e37714
Tada H, Nemoto E, Shimauchi H, Watanabe T, Mikami T, Matsumoto T, Ohno N, Tamura H,
Shibata K, Akashi S et al (2002) Saccharomyces cerevisiae- and Candida albicans-derived
mannan induced production of tumor necrosis factor alpha by human monocytes in a CD14-
and Toll-like receptor 4-dependent manner. Microbiol Immunol 46(7):503–512
Takeuchi O, Hoshino K, Kawai T, Sanjo H, Takada H, Ogawa T, Takeda K, Akira S (1999)
Differential roles of TLR2 and TLR4 in recognition of Gram-negative and gram-positive bacte-
rial cell wall components. Immunity 11(4):443–451
Tarara R, Suleman MA, Sapolsky R, Wabomba MJ, Else JG (1985) Tuberculosis in wild olive
baboons, Papio cynocephalus anubis (Lesson), in Kenya. J Wildl Dis 21(2):137–140
Termeer C, Benedix F, Sleeman J, Fieber C, Voith U, Ahrens T, Miyake K, Freudenberg M,
Galanos C, Simon JC (2002) Oligosaccharides of Hyaluronan activate dendritic cells via Toll-
like receptor 4. J Exp Med 195(1):99–111
Thomson M, Nascimbeni M, Havert MB, Major M, Gonzales S, Alter H, Feinstone SM, Murthy KK,
Rehermann B, Liang TJ (2003) The clearance of hepatitis C virus infection in chimpanzees may
not necessarily correlate with the appearance of acquired immunity. J Virol 77(2):862–870
Triantafilou M, Uddin A, Maher S, Charalambous N, Hamm TS, Alsumaiti A, Triantafilou K
(2007) Anthrax toxin evades Toll-like receptor recognition, whereas its cell wall components
trigger activation via TLR2/6 heterodimers. Cell Microbiol 9(12):2880–2892
Underhill DM, Ozinsky A, Smith KD, Aderem A (1999) Toll-like receptor-2 mediates mycobacteria-
induced proinflammatory signaling in macrophages. Proc Natl Acad Sci USA
96(25):14459–14463
Toll-Like Receptor Function and Evolution in Primates 115
van der Kleij D, Latz E, Brouwers JF, Kruize YC, Schmitz M, Kurt-Jones EA, Espevik T, de Jong
EC, Kapsenberg ML, Golenbock DT et al (2002) A novel host-parasite lipid cross-talk.
Schistosomal lyso-phosphatidylserine activates Toll-like receptor 2 and affects immune polar-
ization. J Biol Chem 277(50):48122–48129
van der Poll T, Levi M, van Deventer SJ, ten Cate H, Haagmans BL, Biemond BJ, Buller HR, Hack
CE, ten Cate JW (1994) Differential effects of anti-tumor necrosis factor monoclonal antibod-
ies on systemic inflammatory responses in experimental endotoxemia in chimpanzees. Blood
83(2):446–451
Vasl J, Prohinar P, Gioannini TL, Weiss JP, Jerala R (2008) Functional activity of MD-2 polymor-
phic variant is significantly different in soluble and TLR4-bound forms: decreased endotoxin
binding by G56R MD-2 and its rescue by TLR4 ectodomain. J Immunol 180(9):6107–6115
Vignal C, Guerardel Y, Kremer L, Masson M, Legrand D, Mazurier J, Elass E (2003) Lipomannans,
but not lipoarabinomannans, purified from Mycobacterium chelonae and Mycobacterium kan-
sasii induce TNF-alpha and IL-8 secretion by a CD14-Toll-like receptor 2-dependent mecha-
nism. J Immunol 171(4):2014–2023
Vitone N, Altizer S, Nunn CL (2004) Body size, diet and sociality influence the species richness of
parasitic worms in anthropoid primates. Evol Ecol Res 6:183–199
Vodros D, Fenyo EM (2004) Primate models for human immunodeficiency virus infection.
Evolution of receptor use during pathogenesis. Acta Microbiol Immunol Hung 51(1–2):1–29
von Bulow GU, Puren AJ, Savage N (1992) Interleukin-1 from baboon peripheral blood mono-
cytes: altered response to endotoxin (lipopolysaccharide) and Staphylococcus aureus stimula-
tion compared with human monocytes. Eur J Cell Biol 59(2):458–463
Walker CM (1997) Comparative features of hepatitis C virus infection in humans and chimpan-
zees. Springer Semin Immunopathol 19(1):85–98
Walsh GP, Tan EV, Dela Cruz EC, Abalos RM, Villahermosa LG, Young LJ, Cellona RV, Nazareno
JB, Horwitz MA (1996) The Philippine cynomolgus monkey (Macaca fasicularis) provides a
new nonhuman primate model of tuberculosis that resembles human disease. Nat Med
2(4):430–436
Walsh DS, Dela Cruz EC, Abalos RM, Tan EV, Walsh GP, Portaels F, Meyers WM (2007) Clinical and
histologic features of skin lesions in a cynomolgus monkey experimentally infected with mycobac-
terium ulcerans (Buruli ulcer) by intradermal inoculation. Am J Trop Med Hyg 76(1):132–134
Werner H, Janitschke K, Kohler H (1969) Observations on marmoset monkeys of the species
Saguinus (Oedipomidas) oedipus following oral and intraperitoneal infection by different cyst-
forming Toxoplasma strains of varying virulence. I. Clinical, pathological anatomical and
parasitological findings. Zentralbl Bakteriol Orig 209(4):553–569
World Health Organization (2004) Manual for the monitoring of yellow fever virus infection.
Immunization VaB Geneva, World Health Organization, Switzerland, p 68
Wlasiuk G, Nachman MW (2010a) Adaptation and constraint at Toll-like receptors in primates.
Mol Biol Evol 27(9):2172–2186
Wlasiuk G, Nachman MW (2010b) Promiscuity and the rate of molecular evolution at primate
immunity genes. Evolution 64(8):2204–2220
Wlasiuk G, Khan S, Switzer WM, Nachman MW (2009) A history of recurrent positive selection
at the Toll-like receptor 5 in primates. Mol Biol Evol 26(4):937–949
Woodall JP (1968) The reaction of a mangabey monkey (Cercocebus galeritus agilis Milne-
Edwards) to inoculation with yellow fever virus. Ann Trop Med Parasitol 62(4):522–527
Yang Z (2007) PAML 4: phylogenetic analysis by maximum likelihood. Mol Biol Evol 24(8):
1586–1591
Yarovinsky F, Zhang D, Andersen JF, Bannenberg GL, Serhan CN, Hayden MS, Hieny S,
Sutterwala FS, Flavell RA, Ghosh S et al (2005) TLR11 activation of dendritic cells by a pro-
tozoan profilin-like protein. Science 308(5728):1626–1629
Yim JJ, Ding L, Schaffer AA, Park GY, Shim YS, Holland SM (2004) A microsatellite polymor-
phism in intron 2 of human Toll-like receptor 2 gene: functional implications and racial differ-
ences. FEMS Immunol Med Microbiol 40(2):163–169
116 J.F. Brinkworth and K.N. Sterner
Introduction
F. Gomez
Department of Genetics and Biology, School of Medicine and School of Arts and Sciences,
University of Pennsylvania, Philadelphia, PA 19104, USA
Department of Anthropology, Hominid Paleobiology Doctoral Program, The George
Washington University, Washington, DC 20052, USA
Department of Anthropology, Center for the Advanced Study of Hominid Paleobiology,
The George Washington University, Washington, DC 20052, USA
Division of Biostatistics and Statistical Genomics, Washington University School
of Medicine in St Louis, 4444 Forest Park Blvd St. Louis, MO 63108, USA
W.-Y. Ko
Department of Genetics and Biology, School of Medicine and School of Arts and Sciences,
University of Pennsylvania, Philadelphia, PA 19104, USA
CIBIO, Research Center in Biodiversity and Genetic Resources, University of Porto,
4485–661 Vairão, Portugal
A. Davis
Department of Genetics and Biology, School of Medicine and School of Arts and Sciences,
University of Pennsylvania, Philadelphia, PA 19104, USA
Department of Biology, Swarthmore College, Swarthmore, PA 19081, USA
Graduate Program in Biological and Biomedical Sciences, Harvard Medical School,
Boston, MA 02115, USA
S.A. Tishkoff (*)
Department of Genetics and Biology, School of Medicine and School of Arts and Sciences,
University of Pennsylvania, Philadelphia, PA 19104, USA
e-mail: [email protected]
J.F. Brinkworth and K. Pechenkina (eds.), Primates, Pathogens, and Evolution, 117
Developments in Primatology: Progress and Prospects,
DOI 10.1007/978-1-4614-7181-3_5, © Springer Science+Business Media New York 2013
118 F. Gomez et al.
Fig. 1 P. falciparum life cycle in a human and Anopheles mosquito. A Plasmodium-infected mos-
quito injects parasites into the human host during a bite that initiates a blood meal. Following the
initial bite, the parasites migrate to the liver where they invade hepatic cells. The parasites develop
within hepatic cells and are then released into the bloodstream, where they invade erythrocytes.
During the erythrocytic stage of infection, the parasites develop and replicate. Some parasites will
go on to infect other red blood cells (merozoite form). Other intraerythrocytic parasites will
develop into male and female gametes. Gametes are taken up by a mosquito during feeding and
will fuse in the mosquito gut to become a zygote. The zygote develops within the mosquito and
eventually migrates to the mosquito salivary gland, from which it will be injected into a human
host during the next blood meal (adapted from Cowman and Crabb 2006)
In the sections to follow, we discuss specific genes that play potential roles in resis-
tance to malaria. We describe the genes, discuss the patterns of genetic variation at
these loci, and describe evidence for natural selection at each locus. We also review
some of the suggested mechanisms for protection or resistance that are associated
with genetic variants at these genes.
Hemoglobin Variants
Adult hemoglobin is comprised of four globin chains (two α and two β chains). These
chains are encoded by the α-globin genes, HBA1 and HBA2, and by the β-globin
gene, HBB. Hundreds of polymorphisms that cause hemoglobinopathies (disorders
of hemoglobin) have been identified worldwide. There are two broad categories of
hemoglobinopathies. These categories include pathologies that affect the structure of
hemoglobin, namely HbS, HbC, and HbE, and pathologies that are related to the
production of hemoglobin or the regulation of hemoglobin production (i.e., thalas-
semias, which will not be discussed here). Hemoglobinopathies were among the first
genetic disorders to be associated with malarial disease because of the strong correla-
tion between the global distribution of many of these diseases (and the genetic vari-
ants that cause them) and the current or historic distribution of malarial disease.
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation 121
HbS
HbS is the mutation that causes sickle-cell anemia, which is one of the most com-
mon hemoglobinopathies associated with protection from malaria. HbS is caused
by a point mutation in the HBB gene that results in a glutamic acid to valine substi-
tution at codon 6. Erythrocytes with this form of hemoglobin are prone to deform
into a characteristic sickle shape. When this mutation is in its homozygous form
(HbSS), the formation of sickled red blood cells is quite serious and can often be
fatal or extremely debilitating (Ashley-Koch et al. 2000). However, in the heterozy-
gous form, HbAS, there are very limited clinical symptoms. These individuals are
relatively healthy and benefit from protection from malaria. Indeed, studies have
confirmed that HbAS is protective against severe and lethal malaria (Aidoo et al.
2002; Williams et al. 2005; Williams 2006).
The HbS allele is widespread over areas of the world where malaria is, or was,
historically endemic. It is especially common in Sub-Saharan Africa, where it
reaches frequencies of 15–20 % (Weatherall 2001; Allison 2009; Verra et al. 2009).
The mechanisms by which HbS confers protection against P. falciparum malaria are
not well understood. Some common hypotheses postulate that HbS hemoglobin may
interfere with parasite growth in RBCs (Nagel and Roth 1989) or that HbS could
enhance splenic clearance of PRBCs (Shear et al. 1993; Roberts and Williams 2003;
Kwiatkowski 2005). It has also been suggested that HbS is protective against malaria
because the mutation causes an increase in the concentration of heme and conse-
quently heme oxygenase-1 (HO-1). HO-1 is an enzyme that catabolizes free heme
and has been shown to be protective against a number of diseases including cerebral
malaria (Pamplona et al. 2007; Ferreira et al. 2011). Recently, using a mouse model,
Ferreira et al. (2011) confirmed the protective effects of HO-1 against malaria and
showed that the protective effect of HbS occurs irrespective of parasite load. In
another recent study, Glushakova et al. (2010) showed that sickled RBCs may be
protective against malaria because they do not support parasite replication. In their
study, they showed that infected HbS RBCs have inefficient parasite egress or an
aborted parasite life cycle. It has also been hypothesized that the protective mecha-
nism of the HbS allele may involve the disruption of cytoadherence and sequestra-
tion. Compared to HbAA-infected erythrocytes, HbAS-infected erythrocytes show a
54 % reduction in adherence to vascular endothelium (Cholera et al. 2008).
The HbS mutation is associated with five “classical” haplotypes that are defined
by different RFLP (restriction fragment length polymorphism) patterns across a
70-kb region surrounding HBB. These haplotypes have different geographic distri-
butions and are named accordingly—Benin, Bantu (central Africa), Cameroon,
Senegal, and Arab (Hanchard et al. 2007). Because the HbS mutation is known to
occur on several haplotype backgrounds, some have suggested that this mutation
may have arisen multiple times (Pagnier et al. 1984; Nagel et al. 1985; Chebloune
et al. 1988). However, recombination and gene conversion have also been suggested
as alternative explanations for the occurrence of HbS on multiple haplotype back-
grounds (Orkin et al. 1982; Webster et al. 2003; Hanchard et al. 2007).
Aside from the well-documented correlation between the distribution of HbS
and P. falciparum malaria (Fig. 2), there is additional evidence of recent positive
122 F. Gomez et al.
Fig. 2 A global map of the spatial limits and endemic levels of P. falciparum malaria and the
geographic distribution of malaria susceptibility alleles. Hyper-holoendemic=areas where childhood
infection prevalence is 50 % or more of population; mesoendemic = areas where childhood infection
prevalence ranges between 11 and 50 %; hypoendemic = areas where childhood infection prevalence
is ≤10 %. Areas where only P. vivax is prevalent and the spatial limits of malaria transmission are
also shown. (Adapted from Snow et al. 2005.) Allelic distribution of HbS, HbC, HbE, Fy*O allele,
Southeast Asian Ovalocytosis (SAO), and Pyruvate Kinase (PK) deficiency overlay the map of
malaria endemicity
HbC
HbC is another mutation that causes a structural change in hemoglobin. The HbC
allele causes a glutamic acid to lysine substitution in codon 6 of HBB, the same
codon at which the HbS substitution occurs (Table 1). This mutation is primarily
found in the malaria-endemic regions of West Africa and is much less common than
Table 1 Genetic polymorphisms, adaptive consequences, and associated pathologies
Sequence alteration/ Adaptive Associated pathology
Alleles haplotype consequences and severity References
Erythrocyte
Hemoglobin structural variants
HBB
HbS Glutamic acid/valine HbS heterozygotes HbS homozygotes Kwiatkowski (2005),
substitution at codon 6 have >90 % have sickle-cell Williams et al.
protection lethal disease, which can (2005), Williams
malaria (study be fatal and (2006)
conducted in debilitating; HbS
Kenya) heterozygotes have
mild to minimal
pathologies
HbC Glutamic acid/lysine HbC heterozygotes HbC homozygotes Modiano et al. (2001b),
substitution at codon 6 have 29 % have mild clinical Diallo et al. (2004)
reduction in risk symptoms; HbC
for clinical malaria; heterozygotes are
HbC homozygotes relatively healthy
have 93 %
reduction in risk
(study conducted in
Burkina Faso)
HbE Glutamic acid/lysine HbE heterozygotes HbE homozygotes Rees et al. (1998),
substitution at codon have less parasitic usually have Chotivanich et al.
26 invasion than HbE symptomless (2002)
homozygotes and anemia; HbE
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation
Table 1 (continued)
Sequence alteration/ Adaptive Associated pathology
Alleles haplotype consequences and severity References
Erythrocyte surface variants
Spectrin (SPTA1) Many alleles Missense mutations; In vitro studies show Elliptocytosis—elliptic Palek (1987), Gallagher
insertion/deletion mutation SPTA1 in RBCs that can and Forget (1996),
polymorphisms; that cause result in hemolysis Dhermy et al.
frameshift mutations elliptocytosis (2007)
resulted in 42 %
reduction in
parasite invasion
SLC4A1 SAO (Southeast 27-base pair deletion In vitro studies show Ovalocytosis—uncom- Cortes et al. (2004),
Asian (SLC4A1D27) causes reduced parasite mon form of Cortes et al. (2005)
ovalocytosis) a nine amino acid invasion; also elliptocytosis;
deletion protection from thought to be lethal
cerebral malaria
DARC (Duffy FY*A/FY*B; Glycine/aspartic acid Fy*O prevents P. vivax Occasional clinical Miller et al. (1976),
antigen Promoter SNP substitution at codon erythrocyte significance but no Mercereau-Puijalon
receptor for 42; T/C SNP at −33 invasion significant and Menard (2010)
chemokines) disrupts erythroid pathologies
expression (Fynul)
ABO O (O01/O02) 1-Base deletion in exon 6 Rosetting is signifi- No significant Carlson and Wahlgren
(D261) shared in O01 cantly reduced in pathologies (1992), Olsson and
a O02;O01 and O02 type O individuals Chester (1996),
differ by 10 mutations Rowe et al. (2007),
in exons 6 and 7 and Calafell et al.
F. Gomez et al.
intron 6 (2008)
Sequence alteration/ Adaptive Associated pathology
Alleles haplotype consequences and severity References
Erythrocyte enzyme variants
G6PD A- Asparagine/aspartic acid In vitro studies show A− variant cause a Cappellini and Fiorelli
at codon 156; valine/ that parasite 10–15 % reduction (2008), Johnson
methionine at codon98 growth is slower in in G6PD enzyme et al. (2009)
G6PD deficient activity, which
RBCs reduces RBC
ability to counteract
oxidative stress
PKLR Many alleles SNPs, insertion/deletion In vitro studies show RBC PKLR deficiency Zanella et al. (2005),
polymorphisms; frame reduced invasion of can cause Ayi et al. (2008)
shift; splice site PKLR deficient hemolytic anemia;
mutations RBCs there is a wide
variety in clinical
severity of PKLR
deficiency
Endothelial receptors for cytoadhesion and sequestration
ICAM-1 ICAMKilifi/rs5491 Lysine/methionine at In vitro studies show No significant Fernandez-Reyes et al.
codon 29 that ICAM-1Kilifi can pathologies (1997), Adams
alter the binding et al. (2000), Tse
ability of some et al. (2004)
strains of P.
falciparum;
association studies
are inconsistent
(studies in Africa
and Asia)
(continued)
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation
125
126
Table 1 (continued)
Sequence alteration/ Adaptive Associated pathology
Alleles haplotype consequences and severity References
rs5498 Lysine/glutamic acid at Association studies are No significant Amodu et al. (2005),
codon 469 inconsistent pathologies; Ma et al. (2006),
possible associa- Puthothu et al.
tions with Type 1 (2006)
diabetes and
asthma
CD36 T1264G/ Tyrosine/stop Association studies are CD36 deficiency Aitman et al. (2000),
rs3211938 inconsistent Fry et al. (2009)
Immune system variants
HLA
HLA-B (MHC- Bw53 haplotype Associated with No significant Hill et al. (1991), Hill
class I) reduced risk for pathologies et al. (1992)
severe malaria
(study conducted in
The Gambia)
HLA-DR (MHC DRB1*1302 haplotype Associated with No significant Hill et al. (1991)
class II) reduced risk for pathologies
severe malaria
(study conducted in
F. Gomez et al.
The Gambia)
Sequence alteration/ Adaptive Associated pathology
Alleles haplotype consequences and severity References
TNF
−238 G/A SNP at −238 Multiple association Associations with Knight et al. (1999),
studies with many autoimmune Ubalee et al.
conflicting results; (e.g., rheumatoid (2001), Bayley
some suggest arthritis, asthma, et al. (2004), Flori
protection some and multiple et al. (2005)
−308 G/A SNP at −308 suggest increased sclerosis) and McGuire et al. (1994),
susceptibility infectious diseases Wattavidanage et al.
but no consistency (1999), Aidoo et al.
across studies (2001), Ubalee
et al. (2001), Meyer
et al. (2002),
Bayley et al.
(2004), Flori et al.
(2005)
−376 G/A SNP at −376 Knight et al. (1999),
Bayley et al. (2004)
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation
127
128 F. Gomez et al.
HbS. As described by Diallo et al. (2004), HbC is associated with phenotypes that
are much less severe than HbS. HbCC homozygotes are described as having mild
anemia and heterozygotes (HbAC) are relatively healthy. Like HbS, this structural
variant of hemoglobin is thought to provide protection from malaria. In a large
case–control study conducted in the Mossi of Burkina Faso, ~90 % of HbCC indi-
viduals and ~29 % of HbAC individuals showed a reduction in clinical malaria
(Modiano et al. 2001b; Min-Oo and Gros 2005). Other studies conducted in Ghana
(Mockenhaupt et al. 2004) and Mali (Agarwal et al. 2000) support the protective
effect of the HbC allele against malarial infection (Min-Oo and Gros 2005). It has
been suggested that HbC may inhibit the invasion or growth of the parasite in HbC
RBCs (Fairhurst et al. 2003). Fairhurst et al. (2005) also reported that HbC-positive
PRBCs showed 25 % reduced adhesion to vascular endothelium. They also showed
rosetting occurred 33 % less often in HbAC PRBCs and they showed rosetting
rarely occurred in HbCC PRBCs.
There have been a limited number of studies that focus on evidence for natural
selection on HbC haplotypes. Wood et al. (2005) estimated the age and strength of
selection associated with the HbC allele by sequencing a ~5.2 kb region of the HBB
gene and analyzing the extent of LD (linkage disequilibrium) (Table 2). Based on
allele frequencies, they estimated that the HbC allele is <5,000 years old and that
the selection coefficient at HbC is 0.04–0.09. However, despite this relatively strong
selection, they observed very little LD upstream from the HbC allele. They sug-
gested that either recombination or gene conversion may have weakened signatures
of recent selection. Similar results were also reported by Modiano et al. (2008).
They demonstrated that HbC haplotypes displayed more evidence of recombination
than did HbS Benin haplotypes and suggested that HbC is likely to be older than the
HbS Benin haplotype.
HbE
An additional variant that affects the structure of hemoglobin is HbE, which is prev-
alent in Southeast Asia. This variant is caused by a glutamic acid to lysine substitu-
tion in codon 26 of HBB (Table 1). Like HbC, HbE is relatively benign. HbEE
individuals usually have symptomless anemia, and HbAE individuals are generally
asymptomatic (Rees et al. 1998). A retrospective study of malaria patients con-
ducted in Thailand found that severe malaria complications were less acute and less
prevalent in individuals with the HbE allele (Hutagalung et al. 1999). In vitro exper-
iments suggest that HbAE RBCs showed reduced parasitic invasion compared to
HbEE RBCs, normal RBCs, and RBCs with other hemoglobinopathies (Chotivanich
et al. 2002).
To examine whether natural selection has affected the pattern of LD on haplo-
types that contain the HbE variant, Ohashi et al. (2004) sequenced a portion of the
HBB locus in Thai patients with mild malaria. They reported that there is extended
LD associated with the HbE variant extending >100 kb. Additionally, their haplo-
type analysis showed that the HbE variant was found on one haplotype background
in the Thai patients they surveyed. They suggested that the age of the HbE variant
Table 2 Age of selected polymorphisms described here, selection coefficients, and corresponding evidence of selection
Gene and Age
protective allele Generations Years Selection coefficient Evidence for natural selection References
HBB
S 45–70 1,125–1,750 0.152* Extended LD (>60 kB); long Currat et al. (2002),
range haplotype similarity Hanchard et al.
(>400 kb); evidence of (2007), Liu et al.
positive directional selection (2009), Hedrick
(2011)
10–28 250–700 – Modiano et al. (2008),
Hedrick (2011)
C 75–150 1,875–3,750 0.04–0.09 Evidence for selection is not Wood et al. (2005)
38–120 950–3,000 – strong; very little LD is Modiano et al. (2008),
associated with HbC perhaps Hedrick (2011)
due to age, recombination or
gene conversion
E 100.3 2,006(1,240–4,440) 0.079 (0.035–0.099) Extended LD (>100 kb); positive Ohashi et al. (2004),
(62–222) directional selection Hedrick (2011)
DARC
ES (null) 1323 33,075 (6,500–97,200) – No clear cut pattern; fixation in Hamblin and Di Rienzo
Africa implies a selective (2000)
ES (null) 490/310 12,250(4,250–26,500)/7,750 – advantage; reduced haplotype Seixas et al. (2002),
(3,625–13,125) variability in Africa also Hedrick (2011)
suggests positive selection in
Africa
ABO
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation
O01 1.15 million – Strong population differentiation; Calafell et al. (2008), Fry
O02 2.5 million – significantly positive Tajima's et al. (2008b)
D and Fu and Li's F suggests
balancing selection
(continued)
129
Table 2 (continued)
130
is between 1,240 and 4,440 years old (Table 2). The results from Ohashi et al. (2004)
support the conclusion that the HbE mutation occurred recently and rose to high
frequency because of selection due to malarial disease.
As a whole, the population genetic data described here suggest that variants
involved in changing the structure of hemoglobin are relatively new mutations
(Table 2). All of the variants discussed here are all likely to have risen to high fre-
quency within the past 10,000 years due to strong selective pressures from falci-
parum malaria. Because these are new mutations and the selection pressures on
these mutations were strong, most loci exhibit long range LD on the haplotypes that
contain the functional alleles (i.e., a classic selective sweep). One exception is the
HBC haplotypes with shorter tracts of LD due to a recombination hotspot. These
haplotypes, however, show other signatures of recent strong selection.
Alleles that cause erythrocyte membrane abnormalities are another class of RBC
polymorphisms that confer resistance to P. falciparum malaria. These mutations
likely perturb the complex process of parasite invasion into erythrocytes, and there-
fore play an important role in providing protection from disease.
One type of RBC membrane abnormality that can affect malaria pathogenesis is
elliptocytosis. Just as it sounds, elliptocytosis is a condition in which RBCs are
elliptical, rather than round, in shape. This condition is considered pathologic
because individuals affected by this condition suffer from anemia and hemolysis
(Palek 1987). The most common form of elliptocytosis is hereditary elliptocytosis.
Hereditary elliptocytosis (HE) includes a heterogeneous group of disorders and a
number of variants in different genes are associated with these disorders (Roux
et al. 1989).
Well-known mutations that cause HE are polymorphisms in the α and β chains of
spectrin, a structural protein that supports the cell membrane and regulates move-
ments of membrane proteins in erythrocytes. A number of these mutations have
been reviewed in Gallagher and Forget (1996). Within African populations there are
two common spectrin mutations, SpαI/65 and SpαI/46. These mutations result in
abnormal 46- or 65-kDa amino acid variants in Spectrin αI (SPTA1). Studies show
that in vitro invasion and growth of P. falciparum is reduced in both the SpαI/65 and
SpαI/46 RBCs (Dhermy et al. 2007).
In addition to spectrin mutations, deletions in the gene SLC4A1 have been stud-
ied extensively and may confer resistance to P. falciparum malaria. The primary
mutation that confers protection from malaria is a 27-bp deletion that causes what
is commonly known as Southeast Asian Ovalocytosis, or SAO (Table 1). As with
132 F. Gomez et al.
the spectrin mutations, this deletion causes the formation of oval-shaped RBCs. The
oval shape of the RBC is thought to be the result of modifications to band 3, a major
anion transporter of erythrocytes (Liu et al. 1990). This deletion is generally not
found in the homozygous state, and it is therefore thought to be lethal when an indi-
vidual inherits two copies (Genton et al. 1995; Allen et al. 1999; Patel et al. 2004;
Williams 2006). SAO is found at high frequencies (up to 30 %) in many parts of
Southeast Asia and in Papua New Guinea (Verra et al. 2009). Like the spectrin
mutations, it is thought that SAO confers protection from malaria because the struc-
tural change in the RBC membrane reduces parasite invasion. It has also been shown
that SAO is associated with reduced in vitro invasion by some strains of P. falci-
parum (Cortes et al. 2004). Upon further investigation, Cortes et al. (2005) showed
a marked increase in SAO RBCs infected with P. falciparum that adhere to CD36,
an adhesion receptor that is generally not associated with cerebral malaria. They
suggested that this increased adhesion to CD36 may prevent the parasite from
adhering to neurovascular adhesion molecules, such as ICAM-1, thereby protecting
the carriers of the SAO-causing deletion from cerebral malaria.
The population genetics of the SLC4A1 deletion in Japan, Taiwan, and Indonesia
were studied by Wilder et al. (2009), who addressed whether SAO chromosomes
differ from wild-type SLC4A1 chromosomes and whether or not these chromosome
show a pattern of variation that is consistent with natural selection. They rese-
quenced ~5 kb of the SLC4A1 gene and typed all individuals for the SAO deletion.
They found significantly negative values of Fay and Wu’s H in all populations,
which implies a larger than expected number of high-frequency derived alleles in
these populations. They also showed that the SAO deletion is on one haplotype
background and the non-SAO haplotype that is closest to the SAO haplotype is rare
and found only in Japan and Indonesia. The Indonesian sample showed the highest
level of diversity, and this was the only population in which the SAO deletion was
polymorphic. From these data, they concluded that the SAO deletion is relatively
recent and that it occurred on a rare haplotype that segregates in Asian
populations.
pattern of selection is consistent with the decoy hypothesis, which argues that
surface proteins like glycophorin A can act as decoy receptors to attract various
pathogens to nonnucleated cells (Gagneux and Varki 1999). Wang et al. (2003)
observed only accelerated rates of protein evolution at GYPA and suggested that
these patterns might have evolved adaptively to escape pathogen recognition during
erythrocyte invasion (evasion hypothesis).
Ko et al. (2011) studied sequence evolution of GYPA, GYPB, and another homo-
logue (GYPE) across diverse African ethnic groups residing within a broad geo-
graphic range. They observed signatures of balancing and positive selection on
different extracellular domains of GYPA. Ko et al. (2011) observed accelerated rates
of adaptive protein evolution at exons 3 and 4, which encode O-sialoglycan-poor
protein domains that are likely to be selected for maintaining the stability of the
protein structure. They also observed a skewed frequency spectrum toward an
excess of intermediate-frequency alleles (evidence of balancing selection) at exon
2, which encodes an O-sialoglycan-rich domain that can modify the binding affinity
of P. falciparum ligands. Ko et al. (2011) further showed that the magnitude of
skewness in the frequency spectrum at exon 2 is significantly correlated with malaria
exposure. Ko et al. (2011) suggested that the observed selection patterns at GYPA
can be better explained by the evasion hypothesis rather than the decoy hypothesis
because the decoy hypothesis usually predicts both fast evolutionary rates and high
genetic diversity at a genetic region, which Ko et al. (2011) did not observe.
Additionally, because the glycophorin genes are highly homologous to each
other at the nucleotide level (>95 %), high levels of gene conversion have been
observed among these homologues (Blumenfeld and Huang 1995; Wang et al. 2003;
Ko et al. 2011). In particular, Ko et al. (2011) showed that two GYPA nonsynony-
mous SNPs that code for the N allele of the MN blood group polymorphism were
introduced from GYPB through gene conversion. These two SNPs were also identi-
fied as candidate variants targeted by balancing selection. Thus, gene conversion
may be a mechanism for introducing high levels of variation upon which natural
selection may act (Wang et al. 2003; Ko et al. 2011).
(erythroid silent) or FY*Bnull. An FY*Anull allele called FY*AES has been identified
in Papua New Guinea (Zimmerman et al. 1999); however, these haplotypes are gen-
erally rare (Mercereau-Puijalon and Menard 2010).
There are strong differences in the global geographic distribution patterns of the
Duffy alleles. For example, the FY*O allele is almost fixed in West and Central
Africa; thus, most Africans do not express the Duffy antigen (Hamblin and Di
Rienzo 2000; Fig. 2). The FY*BES allele is rare among Europeans, American
Indians, and Asian populations. Recently, Howes et al. (2011) used an extensive
collection of literature-based surveys of FY allele frequencies and a Bayesian geo-
statistical model to generate a continuous surface map of the FY alleles across the
globe. Their model showed that the FY*BES allele is fixed in West, Central, and East
Africa and at high frequencies in Madagascar and the Arabian Peninsula. They also
demonstrate that the most prevalent allele outside of Africa is FY*A.
The driving force behind fixation of the Duffy null allele in Africa is a long-
standing question within the fields of anthropology and human genetics. In a land-
mark study, Miller et al. (1976) reported that African Americans who do not express
the Duffy antigen on their RBCs are resistant to malarial infections caused by
Plasmodium vivax. Miller et al. (1976) suggested that because Duffy negativity
offers a selective advantage to people of African descent, the alleles that are respon-
sible for Duffy negativity have reached extremely high frequencies in African popu-
lations. It is now well understood that Duffy-negative individuals are resistant to
malaria caused by P. vivax because the Duffy antigen is the receptor that P. vivax
uses to enter the RBC and is necessary to begin the blood stage of a P. vivax
infection.
Studies of the FY (DARC) locus in a population genetics context loosely support
the idea that directional selection has influenced the patterns of genetic variation
and LD at this locus in African populations (Hamblin and Di Rienzo 2000; Hamblin
et al. 2002). In 2000, Hamblin and Di Rienzo showed that the DARC locus is two- to
threefold more diverse in Europeans than in Africans, which is expected if positive
directional selection has occurred in Africa. However, they also observed two major
FY*O haplotypes, which is not expected if Duffy negativity recently evolved and
quickly gained a selective advantage. Additionally, Hamblin et al. (2002) observed
patterns of genetic variation that departed from neutral evolution on chromosomes
with FY*A alleles in Italian and Chinese populations, suggesting that selection has
acted at this locus in regions other than Africa.
Using sequence data from the FY locus, Hamblin and DiRienzo proposed that
the time of fixation of the FY* O allele is 33,000 years ago, with a confidence inter-
val of 6,500–97,200 years (Table 2), which is inconsistent with recent selection
(Hamblin and Di Rienzo 2000; Hamblin et al. 2002). However, when Seixas et al.
(2002) used microsatellites to date the fixation of the FY*O allele, they suggested a
date of either 7,750 or 12,250 years ago, depending on the haplotype used for esti-
mation (see Hedrick 2011), which is more consistent with recent selection at DARC.
Because P. vivax infections are generally considered to be benign, or at a mini-
mum less serious than infection caused by P. falciparum, some have wondered
whether the strength of selection from P. vivax infection is sufficient to cause the
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation 135
FY*O allele to fix across the African continent (Livingstone 1984). When global
variation within the genome of P. vivax has been examined, an Asian origin for this
parasite has been suggested (Garnham 1966; Carter 2003; Escalante et al. 2005;
Cornejo and Escalante 2006). Additionally, using samples obtained from South and
Central America, Africa, Southeast Asia, and Melanesia, Mu et al. (2005) estimated
that the time to the most recent common ancestor (TMRCA) of P. vivax is 53,000–
265,000 years ago. Although these observations do not necessarily remove P. vivax
as a possible selective agent underlying the FY*O allele fixation in Africa, an old
TMRCA combined with an Asian origin of P. vivax does suggest that this parasite
was not present in Africa with enough time to cause a single SNP to sweep across
West and Central Africa. Thus, it has been suggested that some other selective pres-
sure besides P. vivax may have caused the FY*O allele to fix across most of Sub-
Saharan Africa (Mu et al. 2005).
ABO
Human blood group antigens have long been considered one of the many human
polymorphisms that are associated with resistance to malaria (Athreya and Coriell
1967; Martin et al. 1979; Kassim and Ejezie 1982; Cavalli-Sforza and Feldman
2003). The ABO blood antigens are coded for by one gene (ABO) on chromosome
9 (Table 1). There are four amino acid changes—Arg176Gly, Gly235Ser, Leu266-
Met, and Gly268Ala—that determine whether or not the A or B antigens are
expressed on the red blood cell surface (Type A—Arg176, Gly235, Leu266, and
Gly268; Type B—Gly176, Ser235, Met266, and Ala268; Seto et al. 1997). Type O
is the result of a single nucleotide deletion that causes a reading frame shift, which
creates a premature stop codon, resulting in the expression of a shortened protein
that lacks a necessary catalytic site for determining a blood cell surface antigen
(Yamamoto et al. 1990; Daniels 2005). There are two common O haplotypes, O01
and O02, which differ at exons 6 and 7 (Yamamoto et al. 1990; Olsson and Chester
1996; Roubinet et al. 2004; Calafell et al. 2008).
ABO blood types are differently distributed in global human populations—an
intriguing pattern that has made this a classic human genetic polymorphism of inter-
est to geneticists and anthropologists (reviewed in Uneke 2007). Over the years, a
number of studies have shown that there is a high prevalence of the O allele in tropi-
cal areas of the world and in places where malaria was historically prevalent—and
that blood group A is typically found in colder parts of the world, where malaria is
not historically prevalent (Cserti and Dzik 2007). In addition, numerous case–con-
trol studies have investigated the relationship between P. falciparum malaria and
ABO blood groups. These studies are comprehensively reviewed in Cserti and Dzik
(2007) and Cserti-Gazdewich et al. (2010). For example, Fry et al. (2008b) con-
ducted a case–control study with patients from The Gambia, Kenya, and Malawi
and showed that A, B, and AB individuals appear to be at significantly greater risk
for severe malaria than are individuals with blood group O. They also suggest that
individuals with type B have a slightly lower risk than blood group A, and blood
136 F. Gomez et al.
group AB has the greatest risk of the non-O blood types. Therefore, it has been
hypothesized that type O confers protection from P. falciparum malaria, the A group
confers a disadvantage for protection from malarial disease, and the B antigen has
an intermediate effect (Cserti and Dzik 2007).
The mechanism by which the O blood group confers protection from malaria is
not well understood. It has been shown that red blood cells with A, B, and O pheno-
types show differing amounts of cytoadherence to endothelial receptors and unin-
fected RBCs (Cserti and Dzik 2007; Cserti-Gazdewich et al. 2010). Chen et al.
(2000) reviewed the data that describe differences in rosette formation among the
blood groups and suggested that because A and B antigens are found in abundance
on the surface of non-O RBCs, they are important rosetting receptors. Carlson and
Wahlgren (1992) showed that rosettes can be formed with O RBCs, but they found
the greatest rosetting among A, B, and AB RBCs. Recently, Rowe et al. (2007)
conducted a matched case–control study of 567 Malian children and showed that
type O was least frequent among the patients with severe malaria. They were also
able to show that rosetting was significantly reduced in patients with type O RBCs,
suggesting that the protective effect of type O may be related to a reduced level of
rosetting.
The ABO locus has long been considered a target of natural selection (Akey et al.
2004; Sabeti et al. 2006), and polymorphisms at ABO are considered to be main-
tained due to balancing selection that preceded the divergence of modern humans
(Saitou and Yamamoto 1997; Calafell et al. 2008; Fry et al. 2008b). Thus, there may
be ancient selection maintaining genetic diversity at the ABO locus. In fact, in addi-
tion to malaria, ABO blood groups have also been implicated in resistance to other
infections, including Escherichia coli (Blackwell et al. 2002), Helicobacter pylori
(Boren et al. 1993), Campylobacter jejuni (Ruiz-Palacios et al. 2003), and Norwalk
virus (Lindesmith et al. 2003). Infections from these organisms could also have
acted as selective forces influencing variation at ABO (Calafell et al. 2008). Fry
et al. (2008b) analyzed SNP data from the International HapMap Project and
observed significantly low levels of population differentiation, measured by FST,
when the YRI (Yoruban population), ASN (combined Chinese and Japanese popu-
lation), and CEU (northwestern European population) were compared across the
ABO locus. Their observations are consistent with a model of long-standing balanc-
ing selection at this locus across all HapMap populations. In another study of pat-
terns of genetic diversity at ABO, Calafell et al. (2008) also found evidence for
balancing selection at this locus. They estimated the age of the two primary O lin-
eages (O01 and O02) to be >1 million years old, consistent with a signature of long-
standing balancing selection (Table 2). Thus, it is likely that genetic variation at
ABO has been maintained due to multiple selective forces acting over different time
periods.
Studies of genetic variation at loci that code for erythrocyte membrane proteins
and surface antigens show differing patterns of variation. The SLC4A1 locus shows
evidence of recent positive selection, while GYPA and DARC show mixed signa-
tures of balancing and positive selection depending on the gene region that is exam-
ined. By contrast, the ABO locus shows evidence of long-standing balancing
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation 137
selection. Although most of these loci are likely to participate in the process of
Plasmodium entry into the RBC, they play fundamentally different roles in normal
host biology, which affects the patterns of natural selection and evolution observed
at each locus.
G6PD
inferred the A- variant to be between 2,500 and 3,750 years old. To examine SNP
variation at G6PD in African populations, Verrelli et al. (2002) sequenced a total of
5.2 kb of the G6PD locus in eight African and five non-African populations. Similar
to what Tishkoff et al. (2001) reported, Verrelli et al.’s (2002) results show that there
is less variation associated with A- haplotypes than is expected by chance alone.
However, Verrelli et al. (2002) also suggested that there is an excess of amino acid
polymorphism and that the A- variant is much older than 2,000 years, which may be
consistent with a model of balancing selection maintaining diversity (Table 2).
Pyruvate Kinase
Pyruvate kinase (PK) is an enzyme that is important for the reactions involved in
glucose metabolism. It catalyzes the conversion of phosphoenolpyruvate to pyru-
vate, which results in the generation of one molecule of ATP. This reaction is con-
sidered to be the rate-limiting step of glycolysis, which is critically important for
energy production in red blood cells because they lack mitochondria and rely exclu-
sively on glycolysis for the production of ATP (Ayi et al. 2008; Berghout et al.
2012). First identified in 1960 (Valentine et al. 1961; Zanella et al. 2005), RBC PK
deficiency is known to be the most frequently inherited enzymatic disorder of the
glycolytic pathway and is one of the most common causes of nonspherocytic hemo-
lytic anemia (Ayi et al. 2008; Durand and Coetzer 2008; Tekeste and Petros 2010).
Similar to G6PD, many mutations in the human PKLR gene cause PK deficiency.
For example, Zanella et al. (2005) reported that there are at least 158 known muta-
tions in PKLR that cause enzyme deficiency. However, unlike G6PD, mutations that
cause PK deficiency are generally not in areas where malaria is prevalent (Fig. 1).
Although, PK-deficiency mutations do not show a general pattern of global dis-
tribution characteristic of genetic variation that is correlated with malarial disease
(Fig. 2), some data suggest that variation at this gene could confer protection against
malaria. The first line of evidence in this regard has come from mouse models.
Min-Oo et al. (2003) identified PKL (mouse PLKR gene) using QTL mapping in
mouse strains resistant to malaria. At PKL, Min-Oo et al. (2003) showed that a non-
synonymous SNP that is known to cause PK deficiency in humans, is significantly
associated with decreased parasitemia, and is associated with malarial infection sur-
vival in mice. (Min-Oo et al. 2003; Min-Oo et al. 2004; Min-Oo and Gros 2005).
Since this discovery in the mouse, further studies in humans have suggested that PK
deficiency is related to protection from malaria (Ayi et al. 2008; Durand and Coetzer
2008). Ayi et al. (2008) showed that invasion of RBCs by P. falciparum is signifi-
cantly lower in subjects with homozygous mutations for PK deficiency and phago-
cytosis of PRBCs in patients who are homozygous for PK-deficiency mutations is
markedly higher than phagocytosis in control PRBCs.
Patterns of genetic diversity at PKLR in humans are also suggestive of selection
acting at this locus. In a study of SNPs at the PKLR and neighboring GBA loci,
Mateu et al. (2002) showed that there is strong LD over about 90 kb in this region.
Additionally, Machado et al. (2010) examined SNP and STR variation in
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation 139
individuals from Angola, Mozambique, and Portugal. In this study they showed,
using FST, that there is strong population differentiation between Africans and
Portuguese for SNPs at PKLR. These data suggest selection could have resulted in
significantly different alleles frequencies among Africans and Europeans.
In a recent resequencing study of PKLR in 387 individuals from the Human
Genome Diversity Project (HGDP; Berghout et al. 2012), seven nonsynonymous
SNPs were described, three of which have been reported in PK-deficient individu-
als. Using the algorithms in the computer program SIFT (Henn et al. 2012), Berghout
et al. (2012) showed that these sites are likely to affect the function of PK. The
authors also identified potential signatures of recent positive selection at the PKLR
locus in a pooled population from Pakistan and Sub-Saharan Africa; they observed
significantly negative value of Tajima’s D in the Pakistani population and signifi-
cantly negative values of Fu and Li’s D* and F* in the combined Sub-Saharan popu-
lation and the population from Pakistan. It should be noted that there is likely to be
some substructure among the pooled populations examined by Berghout et al.
(2012), which can also result in negative values of Tajima’s D and Fu and Li’s D*
and F* (Simonsen et al. 1995; Ptak and Przeworski 2002).
ICAM-1
The ICAM-1Kilifi allele is located in the exon that codes for the Ig-like domain that
interacts with PfEMP-1. Experiments have shown that proteins containing the ICAM-
1Kilifi variant amino acid (29M) are less adherent to certain laboratory strains of P. falci-
parum. Adams et al. (2000) showed, under static and flow conditions, that the P.
falciparum strain A4 is dramatically less adherent to proteins with the 29M mutation
than the reference protein. However, they also showed that adherence to the ItG P. falci-
parum strain was less affected by the ICAM-1Kilifi mutation. These data suggest that the
changes in adhesion efficacy due to the ICAM-1Kilifi variant may be strain specific.
CD36
population and is at lower frequencies in other African groups. They also observed
unusually long extended haplotypes for the Yoruba but not for the populations in the
HGDP panel and the Gambia. Overall, Fry et al. (2009) did not find strong evidence
for an association between SNPs at CD36 and malarial phenotypes but suggested
that there was a nonsignificant trend toward heterozygous advantage for the 1264 G
allele. In addition, Bhatia et al. (2011) examined levels of population differentiation
across the genomes of African Americans, Nigerians, and Gambians and identified
strong signal of excessive population differentiation at CD36, which could be inter-
preted as a signal of natural selection at this locus.
These population genetics results are supported by functional studies that sug-
gest CD36 plays a role in malaria susceptibility. McGilvray et al. (2000) examined
whether CD36 may be involved in monocyte/macrophage-mediated malaria clear-
ance. Through several in vitro experiments, they showed that CD36 plays a signifi-
cant role in nonopsonic (nonmediated) phagocytosis of PRBCs and therefore is an
important component leading to infection clearance. Using a mouse model, Patel
et al. (2007) showed that CD36 contributes significantly to the success of an innate
inflammatory response to Plasmodium infections and that CD36 also influences the
duration and severity of malarial disease. These studies imply that CD36 has a com-
plex involvement in Plasmodium infections; it can facilitate infection and pathogen-
esis through adherence to PfEMP-1, but it may also play a critical role in mounting
an efficient immune response to disease.
The patterns of genetic variation at ICAM-1 and CD36 do not show a consistent
signature of strong recent natural selection. Although cytoadherence is an important
and fatal component of malarial disease, there are multiple host receptors that the
parasite exploits. This is an important advantage for the parasite, because it creates
flexibility in the ways in which cytoadherence can be achieved and limits the host’s
ability to protect itself. Thus, interpreting patterns of genetic variation and identify-
ing a classic “selective sweep” at these loci can be difficult. However, at ICAM-1
Gomez et al. (2013) showed that the frequency of the ICAM-1Kilifi allele is correlated
with malaria endemicity. These data suggest correlation analyses can be used to
detect potentially functional alleles in the absence of other signatures of natural
selection. With regard to CD36, it is difficult to explain the inconsistent signatures
of selection at this locus across Africa. Further investigation of sequence variation
at this locus (including regulatory regions) in diverse African populations may help
to reveal the evolutionary history of this locus and its potential role in malaria sus-
ceptibility (Gomez 2012).
Hill et al. (1991) conducted a very influential study of HLA variation and malaria
in The Gambia. In this study of over 2,000 malaria cases and controls, they showed
that the HLA-Bw53 allele and the DRB1*1302-DQBB1*0501 haplotype (an HLA
class II gene) are associated with reduced susceptibility to severe malaria. Hill et al.
(1991) also highlight the fact that these alleles are common in West Africans and are
less common in other global populations, suggesting that natural selection may have
played a role in shaping the allele frequency distribution at these genes.
Although these studies and those reviewed by Ghosh et al. (2008) show that
variation at HLA is potentially important for protection against malaria, the extent
and relevance of this protection is unclear. Once the parasite has entered a human
host, it spends some time in liver cells, but it causes the most serious symptoms
and spends the remainder of its life cycle in the red blood cell. This is an important
point because erythrocytes do not express HLA molecules after enucleation,
which severely limits the ability of the erythrocyte to present antigens to the
immune system (de Villartay et al. 1985; Cserti-Gazdewich et al. 2010).
Furthermore, when Jepson et al. (1997) compared the humoral and cellular
immune response to malaria antigens in dizygotic and haploidentical dizygotic
twins, they found that genes in the HLA class II region did not significantly con-
tribute to the heritable component of an immune response to malaria. This result
suggests that variance in HLA genes does not greatly contribute to variance in
immune response to malarial antigens.
Many other genes involved in immune function, especially those that encode cyto-
kines, have been suggested to harbor variants that influence susceptibility to malaria.
Unfortunately, many of the associations uncovered at cytokine genes have inconsis-
tent results depending on study location and the malarial phenotypes investigated.
One locus that is consistently shown to have polymorphisms that influence a num-
ber of different malarial phenotypes is the gene that encodes tumor necrosis factor
(TNFα), an important proinflammatory cytokine. TNFα is among the initial inflam-
matory cytokines that are released in response to infection. The gene, TNF, which
encodes this cytokine has several promoter polymorphisms (TNF-308, TNF-238,
and TNF-376) that have been shown in several ethnic groups to be related to suscep-
tibility to severe malaria, symptomatic reinfections with P. falciparum, and parasite
density (McGuire et al. 1994; Meyer et al. 2002; Flori et al. 2005). However, despite
some evidence that genetic variation at TNF may confer protection against some
malarial phenotypes, the exact function of these promoter SNPs and how those
functions relate to malaria susceptibility remains unclear (reviewed in Bayley et al.
2004; Kwiatkowski 2005; Smith and Humphries 2009).
Additionally, a number of other cytokine genes have been shown to contain
mutations that may affect malaria susceptibility. These include IL1A (Walley et al.
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation 145
2004), IL1B (Walley et al. 2004), IL4 (Gyan et al. 2004), IL12B (Morahan et al.
2002; Marquet et al. 2008), and IL10 (Wilson et al. 2005; Ouma et al. 2008).
However, as discussed above, many of these associations require replication and
further study to understand the functional mechanisms of these genes that are related
to malaria.
When compared to neighboring populations, the Fulani people, who inhabit North
Sudan, Central Africa, and much of West Africa, demonstrate markedly lower
susceptibility to malarial infection despite facing similar exposure levels com-
pared to neighboring populations (Modiano et al. 1995; Modiano et al. 1996;
Luoni et al. 2001; Dolo et al. 2005). Thus, the Fulani are speculated to have dis-
tinct genetic traits that confer increased protection against P. falciparum infection.
However, many of the known genetic traits that influence infection rate (e.g., HbS,
HbC, and G6PD deficiency) are not overly represented in the Fulani (Modiano
et al. 2001a). It has been proposed that the functional basis of the increased pro-
tection is the result of a more effective immune response (Luoni et al. 2001; Dolo
et al. 2005). When Torcia et al. (2008) studied the expression profiles of a panel
of genes involved in immune response in the Fulani and the Mossi (a neighboring
ethnic group), it was shown that peripheral blood mononuclear cells (PBMCs)
from the Fulani have higher expression of genes related to Th1 and Th2 immune
response compared to the Mossi. Additionally, they also showed a decrease in the
expression of genes related to T-cell regulatory activity in both PBMCs and T
regulatory cells. The authors suggest these data point to decreased activity of T
regulatory cells as the underlying factor mediating the Fulani’s immunity to
malaria.
More recent studies continue to demonstrate that, when compared to other neigh-
boring ethnic groups, the Fulani show differences in immune response to malaria
infection and different patterns of genetic variation at genes involved in the immune
response to malaria (Cserti and Dzik 2007; Senga et al. 2007; Carvalho et al. 2010;
Arama et al. 2011; Henn et al. 2011). For example, Israelsson et al. (2011) showed
that the Fulani have a higher frequency of SNPs associated with a proinflammatory
immune response to malaria compared to neighboring groups. Specifically, their
study showed Fulani have a higher frequency of SNPs that increase the expression
of IL1B (a proinflammatory cytokine) and SNPs that decrease the expression of
IL10 (an inhibitory cytokine). Additionally, Israelsson et al. (2011) showed that
when compared to other ethnic groups, the Fulani have a higher frequency of SNPs
in TNF that are associated with decreased severity of malarial infection (Israelsson
et al. 2011). It should be noted that cultural practices (i.e., pastoralism) and lifestyle
have also been suggested to contribute to the observed difference in malarial disease
in Fulani communities (Wallace and Wallace 2002).
146 F. Gomez et al.
Human genetic studies have identified a number of genetic variants that play a role
in malaria resistance and susceptibility. Because malaria is a strong selection pres-
sure, mutations that are potentially deleterious can be maintained in populations and
will evolve adaptively if they confer protection from malarial infection. Genes that
carry these adaptive mutations will exhibit signatures of natural selection that vary
depending on the age and functional consequences of a particular variant.
Several of the examples discussed above include malaria-protective variants that
have arisen relatively recently and have strong deleterious effects or result in obvious
148 F. Gomez et al.
clinical abnormalities (Table 2). The HbS mutation at HBB and the A- allele at G6PD
are examples of mutations that fall into this category. These genes are generally char-
acterized by distinct haplotypes on which the protective mutation occurred and the
adaptive haplotypes tend to have low haplotype variability and high LD. Other loci,
such as ABO, GYPA, ICAM-1, and CD36 have pleiotropic effects and have complex
signatures of selection resulting from different selective forces acting over different
time periods. These loci may not exhibit classic signatures of recent positive selec-
tion. However, studies that examine the correlation of patterns of variation at these
loci with malaria endemicity (e.g., GYPA and ICAM-1) may be informative for iden-
tifying signatures of selection that are associated with malaria susceptibility.
Looking forward, as the cost of genomic sequencing becomes more affordable
and population genomic studies become a reality, we will be able to examine many
diverse African populations with varying risk for malaria. We can use these data to
identify new candidate loci that influence susceptibility to malaria and test whether
genetic variation is correlated with malaria endemicity. These data may help to
explain the prevalence of deleterious genetic variants in specific ethnic groups and
populations of recent African descent. Low-cost whole-genome sequencing will
also create the opportunity to study genomic variation in Plasmodium genomes.
These data combined with human genomic sequences will provide the means to
explore host–pathogen coevolution and will help us better characterize malaria as a
foundational selective pressure in human evolution.
In summary, malaria has been and continues to be an important selective pres-
sure in modern human evolution. When we are able to combine our improved tech-
nology with access to all human populations at risk for malaria, we will better
understand how our genes play a role in susceptibility to malaria and the role malaria
has played in shaping the human genome.
Acknowledgments We would like to thank the two anonymous reviewers for their critiques and
helpful suggestions. We also thank Dr. Katrina Van Heest for her editorial assistance. S.A. Tishkoff
is supported by a National Institutes of Health grants R01GM076637 and DP1-OD-006445-01,
and NSF Hominid grant (BCS0827436). A Doctoral Dissertation Improvement Grant from the US
National Science Foundation (NSF) (BCS0925802) was given to F. Gomez An NSF IGERT grant
(9987590) to F. Gomez and S.A. Tishkoff supported this research. F. Gomez was also supported by
a Ford Foundation Pre-doctoral fellowship, a Cosmos Club research award, a Sigma Xi (GWU)
Grant-in-Aid of Research (GIAR), and an American Anthropological Association Minority
Dissertation Writing Fellowship.
References
Abbas A, Lichtman A (2005) Cellular and molecular immunology. Elsevier Saunders, Philadelphia
Abecasis GR, Auton A, Brooks LD, DePristo MA, Durbin RM, Handsaker RE, Kang HM, Marth
GT, McVean GA (2012) An integrated map of genetic variation from 1,092 human genomes.
Nature 491(7422):56–65
Adams S, Turner GD, Nash GB, Micklem K, Newbold CI, Craig AG (2000) Differential binding
of clonal variants of Plasmodium falciparum to allelic forms of intracellular adhesion molecule
1 determined by flow adhesion assay. Infect Immun 68(1):264–269
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation 149
Agarwal A, Guindo A, Cissoko Y, Taylor JG, Coulibaly D, Kone A, Kayentao K, Djimde A, Plowe
CV, Doumbo O, Wellems TE, Diallo D (2000) Hemoglobin C associated with protection from
severe malaria in the Dogon of Mali, a West African population with a low prevalence of hemo-
globin S. Blood 96(7):2358–2363
Aidoo M, McElroy PD, Kolczak MS, Terlouw DJ, ter Kuile FO, Nahlen B, Lal AA, Udhayakumar
V (2001) Tumor necrosis factor-alpha promoter variant 2 (Tnf2) is associated with pre-term
delivery, infant mortality, and malaria morbidity in western Kenya: Asembo Bay Cohort Project
Ix. Genet Epidemiol 21(3):201–211
Aidoo M, Terlouw DJ, Kolczak MS, McElroy PD, ter Kuile FO, Kariuki S, Nahlen BL, Lal AA,
Udhayakumar V (2002) Protective effects of the sickle cell gene against malaria morbidity and
mortality. Lancet 359(9314):1311–1312
Aitman TJ, Cooper LD, Norsworthy PJ, Wahid FN, Gray JK, Curtis BR, McKeigue PM,
Kwiatkowski D, Greenwood BM, Snow RW, Hill AV, Scott J (2000) Malaria susceptibility and
Cd36 mutation. Nature 405(6790):1015–1016
Akey JM, Eberle MA, Rieder MJ, Carlson CS, Shriver MD, Nickerson DA, Kruglyak L (2004)
Population history and natural selection shape patterns of genetic variation in 132 genes. PLoS
Biol 2(10):e286
Albrechtsen A, Nielsen FC, Nielsen R (2010) Ascertainment biases in Snp chips affect measures
of population divergence. Mol Biol Evol 27(11):2534–2547
Allen SJ, O’Donnell A, Alexander ND, Mgone CS, Peto TE, Clegg JB, Alpers MP, Weatherall DJ
(1999) Prevention of cerebral malaria in children in Papua New Guinea by Southeast Asian
ovalocytosis band 3. Am J Trop Med Hyg 60(6):1056–1060
Allison AC (2009) Genetic control of resistance to human malaria. Curr Opin Immunol 21(5):
499–505
Amodu OK, Gbadegesin RA, Ralph SA, Adeyemo AA, Brenchley PE, Ayoola OO, Orimadegun
AE, Akinsola AK, Olumese PE, Omotade OO (2005) Plasmodium falciparum malaria in
South-West Nigerian children: is the polymorphism of Icam-1 and E-selectin genes contribut-
ing to the clinical severity of malaria? Acta Trop 95(3):248–255
Arama C, Giusti P, Bostrom S, Dara V, Traore B, Dolo A, Doumbo O, Varani S, Troye-Blomberg
M (2011) Interethnic differences in antigen-presenting cell activation and Tlr responses in
Malian children during Plasmodium falciparum malaria. PLoS One 6(3):e18319
Armesilla AL, Vega MA (1994) Structural organization of the gene for human Cd36 glycoprotein.
J Biol Chem 269(29):18985–18991
Ashley-Koch A, Yang Q, Olney RS (2000) Sickle hemoglobin (Hbs) allele and sickle cell disease:
a huge review. Am J Epidemiol 151(9):839–845
Athreya BH, Coriell LL (1967) Relation of blood groups to infection. I. A survey and review of
data suggesting possible relationship between malaria and blood groups. Am J Epidemiol
86(2):292–304
Ayi K, Min-Oo G, Serghides L, Crockett M, Kirby-Allen M, Quirt I, Gros P, Kain KC (2008)
Pyruvate kinase deficiency and malaria. N Engl J Med 358(17):1805–1810
Baum J, Ward RH, Conway DJ (2002) Natural selection on the erythrocyte surface. Mol Biol Evol
19(3):223–229
Bayley JP, Ottenhoff TH, Verweij CL (2004) Is there a future for Tnf promoter polymorphisms?
Genes Immun 5(5):315–329
Bellamy R, Kwiatkowski D, Hill AV (1998) Absence of an association between intercellular adhesion
molecule 1, complement receptor 1 and interleukin 1 receptor antagonist gene polymorphisms
and severe malaria in a West African population. Trans R Soc Trop Med Hyg 92(3):312–316
Berghout J, Higgins S, Loucoubar C, Sakuntabhai A, Kain KC, Gros P (2012) Genetic diversity in
human erythrocyte pyruvate kinase. Genes Immun 13(1):98–102
Beutler E (1994) G6pd deficiency. Blood 84(11):3613–3636
Bhatia G, Patterson N, Pasaniuc B, Zaitlen N, Genovese G, Pollack S, Mallick S, Myers S, Tandon
A, Spencer C, Palmer CD, Adeyemo AA, Akylbekova EL, Cupples LA, Divers J, Fornage M,
Kao WH, Lange L, Li M, Musani S, Mychaleckyj JC, Ogunniyi A, Papanicolaou G, Rotimi
CN, Rotter JI, Ruczinski I, Salako B, Siscovick DS, Tayo BO, Yang Q, McCarroll S, Sabeti P,
150 F. Gomez et al.
Lettre G, De Jager P, Hirschhorn J, Zhu X, Cooper R, Reich D, Wilson JG, Price AL (2011)
Genome-wide comparison of African-ancestry populations from care and other cohorts reveals
signals of natural selection. Am J Hum Genet 89(3):368–381
Blackwell CC, Dundas S, James VS, Mackenzie DA, Braun JM, Alkout AM, Todd WT, Elton RA,
Weir DM (2002) Blood group and susceptibility to disease caused by Escherichia Coli O157.
J Infect Dis 185(3):393–396
Blumenfeld OO, Huang CH (1995) Molecular genetics of the glycophorin gene family, the anti-
gens for Mnss blood groups: multiple gene rearrangements and modulation of splice site usage
result in extensive diversification. Hum Mutat 6(3):199–209
Boren T, Falk P, Roth KA, Larson G, Normark S (1993) Attachment of Helicobacter pylori to
human gastric epithelium mediated by blood group antigens. Science 262(5141):1892–1895
Calafell F, Roubinet F, Ramirez-Soriano A, Saitou N, Bertranpetit J, Blancher A (2008)
Evolutionary dynamics of the human Abo gene. Hum Genet 124(2):123–135
Cappellini MD, Fiorelli G (2008) Glucose-6-phosphate dehydrogenase deficiency. Lancet
371(9606):64–74
Carlson J, Wahlgren M (1992) Plasmodium falciparum erythrocyte rosetting is mediated by pro-
miscuous lectin-like interactions. J Exp Med 176(5):1311–1317
Carter R (2003) Speculations on the origins of Plasmodium vivax malaria. Trends Parasitol
19(5):214–219
Carvalho DB, de Mattos LC, Souza-Neiras WC, Bonini-Domingos CR, Cosimo AB, Storti-Melo
LM, Cassiano GC, Couto AA, Cordeiro AJ, Rossit AR, Machado RL (2010) Frequency of Abo
blood group system polymorphisms in Plasmodium falciparum malaria patients and blood
donors from the Brazilian Amazon region. Genet Mol Res 9(3):1443–1449
Cavalli-Sforza LL, Feldman MW (2003) The application of molecular genetic approaches to the
study of human evolution. Nat Genet 33(Suppl):266–275
Chakraborty S, Craig AG (2004) The role of Icam-1 in Plasmodium falciparum cytoadherence.
Eur J Cell Biol 84:15–27
Chebloune Y, Pagnier J, Trabuchet G, Faure C, Verdier G, Labie D, Nigon V (1988) Structural
analysis of the 5′ flanking region of the beta-globin gene in African sickle cell anemia patients:
further evidence for three origins of the sickle cell mutation in Africa. Proc Natl Acad Sci USA
85(12):4431–4435
Chen Q, Schlichtherle M, Wahlgren M (2000) Molecular aspects of severe Malaria. Clin Microbiol
Rev 13(3):439–450
Cholera R, Brittain NJ, Gillrie MR, Lopera-Mesa TM, Diakite SA, Arie T, Krause MA, Guindo A,
Tubman A, Fujioka H, Diallo DA, Doumbo OK, Ho M, Wellems TE, Fairhurst RM (2008)
Impaired cytoadherence of Plasmodium falciparum infected erythrocytes containing sickle
hemoglobin. Proc Natl Acad Sci USA 105(3):991–996
Chotivanich K, Udomsangpetch R, Pattanapanyasat K, Chierakul W, Simpson J, Looareesuwan S,
White N (2002) Hemoglobin E: a balanced polymorphism protective against high parasitemias
and thus severe P falciparum malaria. Blood 100(4):1172–1176
Cornejo OE, Escalante AA (2006) The origin and age of Plasmodium vivax. Trends Parasitol
22(12):558–563
Corpeleijn E, van der Kallen CJ, Kruijshoop M, Magagnin MG, de Bruin TW, Feskens EJ, Saris
WH, Blaak EE (2006) Direct association of a promoter polymorphism in the Cd36/fat fatty
acid transporter gene with type 2 diabetes mellitus and insulin resistance. Diabet Med
23(8):907–911
Cortes A, Benet A, Cooke BM, Barnwell JW, Reeder JC (2004) Ability of Plasmodium falciparum
to invade Southeast Asian ovalocytes varies between parasite lines. Blood 104(9):2961–2966
Cortes A, Mellombo M, Mgone CS, Beck HP, Reeder JC, Cooke BM (2005) Adhesion of
Plasmodium falciparum infected red blood cells to Cd36 under flow is enhanced by the cere-
bral malaria-protective trait South-East Asian ovalocytosis. Mol Biochem Parasitol 142(2):
252–257
Cowman AF, Crabb BS (2006) Invasion of red blood cells by malaria parasites. Cell
124(4):755–766
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation 151
Craig A, Scherf A (2001) Molecules on the surface of the Plasmodium falciparum infected eryth-
rocyte and their role in malaria pathogenesis and immune evasion. Mol Biochem Parasitol
115(2):129–143
Crosnier C, Bustamante LY, Bartholdson SJ, Bei AK, Theron M, Uchikawa M, Mboup S, Ndir O,
Kwiatkowski DP, Duraisingh MT, Rayner JC, Wright GJ (2011) Basigin is a receptor essential
for erythrocyte invasion by Plasmodium falciparum. Nature 480(7378):534–537
Cserti CM, Dzik WH (2007) The Abo blood group system and Plasmodium falciparum malaria.
Blood 110(7):2250–2258
Cserti-Gazdewich CM, Mayr WR, Dzik WH (2010) Plasmodium falciparum malaria and the
immunogenetics of Abo, Hla, and Cd36 (Platelet Glycoprotein Iv). Vox Sang 100(1):99–111
Currat M, Trabuchet G, Rees D, Perrin P, Harding RM, Clegg JB, Langaney A, Excoffier L (2002)
Molecular analysis of the beta-globin gene cluster in the Niokholo Mandenka population
reveals a recent origin of the beta(S) senegal mutation. Am J Hum Genet 70(1):207–223
Daniels G (2005) The molecular genetics of blood group polymorphism. Transpl Immunol
14(3–4):143–153
de Villartay JP, Rouger P, Muller JY, Salmon C (1985) Hla antigens on peripheral red blood cells:
analysis by flow cytofluorometry using monoclonal antibodies. Tissue Antigens 26(1):12–19
Dhermy D, Schrevel J, Lecomte MC (2007) Spectrin-based skeleton in red blood cells and malaria.
Curr Opin Hematol 14(3):198–202
Diallo DA, Doumbo OK, Dicko A, Guindo A, Coulibaly D, Kayentao K, Djimde AA, Thera MA,
Fairhurst RM, Plowe CV, Wellems TE (2004) A comparison of anemia in hemoglobin C and
normal hemoglobin a children with Plasmodium falciparum malaria. Acta Trop 90(3):295–299
Dolo A, Modiano D, Maiga B, Daou M, Dolo G, Guindo H, Ba M, Maiga H, Coulibaly D, Perlman
H, Blomberg MT, Toure YT, Coluzzi M, Doumbo O (2005) Difference in susceptibility to
malaria between two sympatric ethnic groups in Mali. Am J Trop Med Hyg 72(3):243–248
Doolan DL, Dobano C, Baird JK (2009) Acquired immunity to malaria. Clin Microbiol Rev
22(1):13–36, Table of Contents
Duah NO, Weiss HA, Jepson A, Tetteh KK, Whittle HC, Conway DJ (2009) Heritability of anti-
body isotype and subclass responses to Plasmodium falciparum antigens. PLoS One
4(10):e7381
Durand PM, Coetzer TL (2008) Pyruvate kinase deficiency protects against malaria in humans.
Haematologica 93(6):939–940
Escalante AA, Cornejo OE, Freeland DE, Poe AC, Durrego E, Collins WE, Lal AA (2005) A
Monkey’s tale: the origin of Plasmodium vivax as a human malaria parasite. Proc Natl Acad
Sci USA 102(6):1980–1985
Fairhurst RM, Fujioka H, Hayton K, Collins KF, Wellems TE (2003) Aberrant development of
Plasmodium falciparum in hemoglobin Cc red cells: implications for the malaria protective
effect of the homozygous state. Blood 101(8):3309–3315
Fairhurst RM, Baruch DI, Brittain NJ, Ostera GR, Wallach JS, Hoang HL, Hayton K, Guindo A,
Makobongo MO, Schwartz OM, Tounkara A, Doumbo OK, Diallo DA, Fujioka H, Ho M,
Wellems TE (2005) Abnormal display of Pfemp-1 on erythrocytes carrying haemoglobin C
may protect against malaria. Nature 435(7045):1117–1121
Fernandez-Reyes D, Craig AG, Kyes SA, Peshu N, Snow RW, Berendt AR, Marsh K, Newbold CI
(1997) A high frequency African coding polymorphism in the N-terminal domain of Icam-1
predisposing to cerebral malaria in Kenya. Hum Mol Genet 6(8):1357–1360
Ferreira A, Marguti I, Bechmann I, Jeney V, Chora A, Palha NR, Rebelo S, Henri A, Beuzard Y,
Soares MP (2011) Sickle hemoglobin confers tolerance to Plasmodium infection. Cell
145(3):398–409
Flori L, Kumulungui B, Aucan C, Esnault C, Traore AS, Fumoux F, Rihet P (2003) Linkage and
association between Plasmodium falciparum blood infection levels and chromosome 5q31-
Q33. Genes Immun 4(4):265–268
Flori L, Delahaye NF, Iraqi FA, Hernandez-Valladares M, Fumoux F, Rihet P (2005) Tnf as a
malaria candidate gene: polymorphism-screening and family-based association analysis of
mild malaria attack and parasitemia in Burkina Faso. Genes Immun 6(6):472–480
152 F. Gomez et al.
Fry AE, Auburn S, Diakite M, Green A, Richardson A, Wilson J, Jallow M, Sisay-Joof F, Pinder
M, Griffiths MJ, Peshu N, Williams TN, Marsh K, Molyneux ME, Taylor TE, Rockett KA,
Kwiatkowski DP (2008a) Variation in the Icam1 gene is not associated with severe malaria
phenotypes. Genes Immun 9(5):462–469
Fry AE, Griffiths MJ, Auburn S, Diakite M, Forton JT, Green A, Richardson A, Wilson J, Jallow
M, Sisay-Joof F, Pinder M, Peshu N, Williams TN, Marsh K, Molyneux ME, Taylor TE,
Rockett KA, Kwiatkowski DP (2008b) Common variation in the Abo glycosyltransferase is
associated with susceptibility to severe Plasmodium falciparum malaria. Hum Mol Genet
17(4):567–576
Fry AE, Ghansa A, Small KS, Palma A, Auburn S, Diakite M, Green A, Campino S, Teo YY, Clark
TG, Jeffreys AE, Wilson J, Jallow M, Sisay-Joof F, Pinder M, Griffiths MJ, Peshu N, Williams
TN, Newton CR, Marsh K, Molyneux ME, Taylor TE, Koram KA, Oduro AR, Rogers WO,
Rockett KA, Sabeti PC, Kwiatkowski DP (2009) Positive selection of a Cd36 nonsense variant
in sub-Saharan Africa, but no association with severe malaria phenotypes. Hum Mol Genet
18(14):2683–2692
Gagneux P, Varki A (1999) Evolutionary considerations in relating oligosaccharide diversity to
biological function. Glycobiology 9(8):747–755
Gallagher PG, Forget BG (1996) Hematologically important mutations: spectrin variants in heredi-
tary elliptocytosis and hereditary pyropoikilocytosis. Blood Cells Mol Dis 22(3):254–258
Garcia A, Marquet S, Bucheton B, Hillaire D, Cot M, Fievet N, Dessein AJ, Abel L (1998) Linkage
analysis of blood Plasmodium falciparum levels: interest of the 5q31-Q33 chromosome region.
Am J Trop Med Hyg 58(6):705–709
Garnham PC (1966) Malaria parasites and other haemosporidia. Blackwell Scientific, Oxford
Garrigan D, Hedrick PW (2003) Perspective: detecting adaptive molecular polymorphism: lessons
from the Mhc. Evolution 57(8):1707–1722
Gelhaus A, Scheding A, Browne E, Burchard GD, Horstmann RD (2001) Variability of the Cd36
gene in West Africa. Hum Mutat 18(5):444–450
Genton B, al-Yaman F, Mgone CS, Alexander N, Paniu MM, Alpers MP, Mokela D (1995)
Ovalocytosis and cerebral malaria. Nature 378(6557):564–565
Ghosh K (2008) Evolution and selection of human leukocyte antigen alleles by Plasmodium
falciparum infection. Hum Immunol 69(12):856–860
Glushakova S, Humphrey G, Leikina E, Balaban A, Miller J, Zimmerberg J (2010) New stages in
the program of malaria parasite egress imaged in normal and sickle erythrocytes. Curr Biol
20(12):1117–1121
Gomez F (2012) Genetic variation at Icam-1 and Cd36: a study of malaria resistance candidate loci
in diverse global human populations. PhD Thesis The George Washington University
Gomez F, Tomas G, Ko WY, Ranciaro A, Froment A, Ibrahim M, Lema G, Nyambo TB, Omar SA,
Wambebe C, Hirbo JB, Rocha J, Tishkoff SA (2013) Patterns of nucleotide and haplotype diversity
at Icam-1 across global human populations with varying levels of malaria exposure. Hum Genet.
Epub date: 2013/04/24
Guerra CA, Gikandi PW, Tatem AJ, Noor AM, Smith DL, Hay SI, Snow RW (2008) The limits and
intensity of Plasmodium falciparum transmission: implications for malaria control and elimi-
nation worldwide. PLoS Med 5(2):e38
Guindo A, Fairhurst RM, Doumbo OK, Wellems TE, Diallo DA (2007) X-linked G6pd deficiency
protects hemizygous males but not heterozygous females against severe malaria. PLoS Med
4(3):e66
Gyan BA, Goka B, Cvetkovic JT, Kurtzhals JL, Adabayeri V, Perlmann H, Lefvert AK, Akanmori
BD, Troye-Blomberg M (2004) Allelic polymorphisms in the repeat and promoter regions of
the interleukin-4 gene and malaria severity in Ghanaian children. Clin Exp Immunol
138(1):145–150
Haldane JBS (1949) The rate of mutation in human genes. Proc VIII Int Cong Genet Hereditas
35:267–273
Hamblin MT, Di Rienzo A (2000) Detection of the signature of natural selection in humans:
evidence from the Duffy blood group locus. Am J Hum Genet 66(5):1669–1679
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation 153
Hamblin MT, Thompson EE, Di Rienzo A (2002) Complex signatures of natural selection at the
Duffy blood group locus. Am J Hum Genet 70(2):369–383
Hanchard N, Elzein A, Trafford C, Rockett K, Pinder M, Jallow M, Harding R, Kwiatkowski D,
McKenzie C (2007) Classical sickle beta-globin haplotypes exhibit a high degree of long-range
haplotype similarity in African and Afro-caribbean populations. BMC Genet 8:52
Hay SI, Guerra CA, Gething PW, Patil AP, Tatem AJ, Noor AM, Kabaria CW, Manh BH, Elyazar
IR, Brooker S, Smith DL, Moyeed RA, Snow RW (2009) A world malaria map: Plasmodium
falciparum endemicity in 2007. PLoS Med 6(3):e1000048
Hedrick PW (2011) Population genetics of malaria resistance in humans. Heredity (Edinb)
107(4):283–304
Henn BM, Gignoux CR, Jobin M, Granka JM, Macpherson JM, Kidd JM, Rodriguez-Botigue L,
Ramachandran S, Hon L, Brisbin A, Lin AA, Underhill PA, Comas D, Kidd KK, Norman PJ,
Parham P, Bustamante CD, Mountain JL, Feldman MW (2011) Hunter-gatherer genomic diver-
sity suggests a Southern African origin for modern humans. Proc Natl Acad Sci USA 108(13):
5154–5162
Henn BM, Botigue LR, Gravel S, Wang W, Brisbin A, Byrnes JK, Fadhlaoui-Zid K, Zalloua PA,
Moreno-Estrada A, Bertranpetit J, Bustamante CD, Comas D (2012) Genomic ancestry of
North Africans supports back-to-Africa migrations. PLoS Genet 8(1):e1002397
Hill AV, Allsopp CE, Kwiatkowski D, Anstey NM, Twumasi P, Rowe PA, Bennett S, Brewster D,
McMichael AJ, Greenwood BM (1991) Common West African Hla antigens are associated
with protection from severe malaria. Nature 352(6336):595–600
Hill AV, Elvin J, Willis AC, Aidoo M, Allsopp CE, Gotch FM, Gao XM, Takiguchi M, Greenwood
BM, Townsend AR et al (1992) Molecular analysis of the association of Hla-B53 and resis-
tance to severe malaria. Nature 360(6403):434–439
Howes RE, Patil AP, Piel FB, Nyangiri OA, Kabaria CW, Gething PW, Zimmerman PA, Barnadas
C, Beall CM, Gebremedhin A, Menard D, Williams TN, Weatherall DJ, Hay SI (2011) The
global distribution of the Duffy blood group. Nat Commun 2:266
Hutagalung R, Wilairatana P, Looareesuwan S, Brittenham GM, Aikawa M, Gordeuk VR (1999)
Influence of hemoglobin E trait on the severity of falciparum malaria. J Infect Dis 179(1):
283–286
Imai M, Tanaka T, Kintaka T, Ikemoto T, Shimizu A, Kitaura Y (2002) Genomic heterogeneity of
type Ii Cd36 deficiency. Clin Chim Acta 321(1–2):97–106
Israelsson E, Maiga B, Kearsley S, Dolo A, Homann MV, Doumbo OK, Troye-Blomberg M,
Tornvall P, Berzins K (2011) Cytokine gene haplotypes with a potential effect on susceptibility
to malaria in sympatric ethnic groups in Mali. Infect Genet Evol 11(7):1608–1615
Jallow M, Teo YY, Small KS, Rockett KA, Deloukas P, Clark TG, Kivinen K, Bojang KA, Conway
DJ, Pinder M, Sirugo G, Sisay-Joof F, Usen S, Auburn S, Bumpstead SJ, Campino S, Coffey
A, Dunham A, Fry AE, Green A, Gwilliam R, Hunt SE, Inouye M, Jeffreys AE, Mendy A,
Palotie A, Potter S, Ragoussis J, Rogers J, Rowlands K, Somaskantharajah E, Whittaker P,
Widden C, Donnelly P, Howie B, Marchini J, Morris A, SanJoaquin M, Achidi EA, Agbenyega
T, Allen A, Amodu O, Corran P, Djimde A, Dolo A, Doumbo OK, Drakeley C, Dunstan S,
Evans J, Farrar J, Fernando D, Hien TT, Horstmann RD, Ibrahim M, Karunaweera N, Kokwaro
G, Koram KA, Lemnge M, Makani J, Marsh K, Michon P, Modiano D, Molyneux ME, Mueller
I, Parker M, Peshu N, Plowe CV, Puijalon O, Reeder J, Reyburn H, Riley EM, Sakuntabhai A,
Singhasivanon P, Sirima S, Tall A, Taylor TE, Thera M, Troye-Blomberg M, Williams TN,
Wilson M, Kwiatkowski DP (2009) Genome-wide and fine-resolution association analysis of
malaria in West Africa. Nat Genet 41(6):657–665
Jepson A, Sisay-Joof F, Banya W, Hassan-King M, Frodsham A, Bennett S, Hill AV, Whittle H
(1997) Genetic linkage of mild malaria to the major histocompatibility complex in Gambian
children: study of affected sibling pairs. BMJ 315(7100):96–97
Johnson MK, Clark TD, Njama-Meya D, Rosenthal PJ, Parikh S (2009) Impact of the method of
G6pd deficiency assessment on genetic association studies of malaria susceptibility. PLoS One
4(9):e7246
154 F. Gomez et al.
Patel SS, King CL, Mgone CS, Kazura JW, Zimmerman PA (2004) Glycophorin C (Gerbich
antigen blood group) and band 3 polymorphisms in two malaria holoendemic regions of Papua
New Guinea. Am J Hematol 75(1):1–5
Patel SN, Lu Z, Ayi K, Serghides L, Gowda DC, Kain KC (2007) Disruption of Cd36 impairs
cytokine response to Plasmodium falciparum glycosylphosphatidylinositol and confers sus-
ceptibility to severe and fatal malaria in vivo. J Immunol 178(6):3954–3961
Phimpraphi W, Paul R, Witoonpanich B, Turbpaiboon C, Peerapittayamongkol C, Louicharoen C,
Casademont I, Tungpradabkul S, Krudsood S, Kaewkunwal J, Sura T, Looareesuwan S,
Singhasivanon P, Sakuntabhai A (2008) Heritability of P. falciparum and P. vivax malaria in a
karen population in Thailand. PLoS One 3(12):e3887
Piazza A, Mayr WR, Contu L, Amoroso A, Borelli I, Curtoni ES, Marcello C, Moroni A, Olivetti
E, Richiardi P et al (1985) Genetic and population structure of four Sardinian villages. Ann
Hum Genet 49(Pt 1):47–63
Potts WK, Wakeland EK (1993) Evolution of Mhc genetic diversity: a tale of incest, pestilence and
sexual preference. Trends Genet 9(12):408–412
Prugnolle F, Manica A, Charpentier M, Guegan JF, Guernier V, Balloux F (2005) Pathogen-driven
selection and worldwide Hla class I diversity. Curr Biol 15(11):1022–1027
Ptak SE, Przeworski M (2002) Evidence for population growth in humans is confounded by fine-
scale population structure. Trends Genet 18(11):559–563
Puthothu B, Krueger M, Bernhardt M, Heinzmann A (2006) Icam1 amino-acid variant K469e is
associated with paediatric bronchial asthma and elevated Sicam1 levels. Genes Immun
7(4):322–326
Rac ME, Safranow K, Poncyljusz W (2007) Molecular basis of human Cd36 gene mutations. Mol
Med 13(5–6):288–296
Rees DC, Styles L, Vichinsky EP, Clegg JB, Weatherall DJ (1998) The hemoglobin E syndromes.
Ann NY Acad Sci 850:334–343
Register TC, Burdon KP, Lenchik L, Bowden DW, Hawkins GA, Nicklas BJ, Lohman K, Hsu FC,
Langefeld CD, Carr JJ (2004) Variability of serum soluble intercellular adhesion molecule-1
measurements attributable to a common polymorphism. Clin Chem 50(11):2185–2187
Rihet P, Traore Y, Abel L, Aucan C, Traore-Leroux T, Fumoux F (1998) Malaria in humans:
Plasmodium falciparum blood infection levels are linked to chromosome 5q31-Q33. Am J
Hum Genet 63(2):498–505
Roberts LS, Janovy J (2005) Foundations of parasitology. McGraw Hill Higher Education, Boston
Roberts DJ, Williams TN (2003) Haemoglobinopathies and resistance to malaria. Redox Rep
8(5):304–310
Roubinet F, Despiau S, Calafell F, Jin F, Bertranpetit J, Saitou N, Blancher A (2004) Evolution of
the O alleles of the human Abo blood group gene. Transfusion 44(5):707–715
Roux AF, Morle F, Guetarni D, Colonna P, Sahr K, Forget BG, Delaunay J, Godet J (1989)
Molecular basis of Sp Alpha I/65 hereditary elliptocytosis in North Africa: insertion of a Ttg
triplet between codons 147 and 149 in the alpha-spectrin gene from five unrelated families.
Blood 73(8):2196–2201
Rowe JA, Handel IG, Thera MA, Deans AM, Lyke KE, Kone A, Diallo DA, Raza A, Kai O, Marsh
K, Plowe CV, Doumbo OK, Moulds JM (2007) Blood group O protects against severe
Plasmodium falciparum malaria through the mechanism of reduced rosetting. Proc Natl Acad
Sci USA 104(44):17471–17476
Rowe JA, Claessens A, Corrigan RA, Arman M (2009) Adhesion of Plasmodium falciparum-
infected erythrocytes to human cells: molecular mechanisms and therapeutic implications.
Expert Rev Mol Med 11:e16
Ruiz-Palacios GM, Cervantes LE, Ramos P, Chavez-Munguia B, Newburg DS (2003)
Campylobacter jejuni binds intestinal H(O) antigen (Fuc Alpha 1, 2gal Beta 1, 4glcnac), and
fucosyloligosaccharides of human milk inhibit its binding and infection. J Biol Chem
278(16):14112–14120
Ruwende C, Hill A (1998) Glucose-6-phosphate dehydrogenase deficiency and malaria. J Mol
Med (Berl) 76(8):581–588
158 F. Gomez et al.
Ruwende C, Khoo SC, Snow RW, Yates SN, Kwiatkowski D, Gupta S, Warn P, Allsopp CE, Gilbert
SC, Peschu N et al (1995) Natural selection of hemi- and heterozygotes for G6pd deficiency in
Africa by resistance to severe malaria. Nature 376(6537):246–249
Ryan AW, Mapp J, Moyna S, Mattiangeli V, Kelleher D, Bradley DG, McManus R (2006) Levels
of interpopulation differentiation among different functional classes of immunologically
important genes. Genes Immun 7(2):179–183
Sabeti PC, Reich DE, Higgins JM, Levine HZ, Richter DJ, Schaffner SF, Gabriel SB, Platko JV,
Patterson NJ, McDonald GJ, Ackerman HC, Campbell SJ, Altshuler D, Cooper R, Kwiatkowski
D, Ward R, Lander ES (2002) Detecting recent positive selection in the human genome from
haplotype structure. Nature 419(6909):832–837
Sabeti PC, Schaffner SF, Fry B, Lohmueller J, Varilly P, Shamovsky O, Palma A, Mikkelsen TS,
Altshuler D, Lander ES (2006) Positive natural selection in the human lineage. Science
312(5780):1614–1620
Saitou N, Yamamoto F (1997) Evolution of primate Abo blood group genes and their homologous
genes. Mol Biol Evol 14(4):399–411
Sakuntabhai A, Ndiaye R, Casademont I, Peerapittayamongkol C, Rogier C, Tortevoye P, Tall A,
Paul R, Turbpaiboon C, Phimpraphi W, Trape JF, Spiegel A, Heath S, Mercereau-Puijalon O,
Dieye A, Julier C (2008) Genetic determination and linkage mapping of Plasmodium
falciparum malaria related traits in Senegal. PLoS One 3(4):e2000
Saunders MA, Hammer MF, Nachman MW (2002) Nucleotide variability at G6pd and the signa-
ture of malarial selection in humans. Genetics 162(4):1849–1861
Seixas S, Ferrand N, Rocha J (2002) Microsatellite variation and evolution of the human Duffy
blood group polymorphism. Mol Biol Evol 19(10):1802–1806
Senga E, Loscertales MP, Makwakwa KE, Liomba GN, Dzamalala C, Kazembe PN, Brabin BJ
(2007) Abo blood group phenotypes influence parity specific immunity to Plasmodium
falciparum malaria in Malawian women. Malar J 6:102
Seto NO, Palcic MM, Compston CA, Li H, Bundle DR, Narang SA (1997) Sequential interchange
of four amino acids from blood group B to blood group a glycosyltransferase boosts catalytic
activity and progressively modifies substrate recognition in human recombinant enzymes. J
Biol Chem 272(22):14133–14138
Shear HL, Roth EF Jr, Fabry ME, Costantini FD, Pachnis A, Hood A, Nagel RL (1993) Transgenic
mice expressing human sickle hemoglobin are partially resistant to rodent malaria. Blood
81(1):222–226
Sherman IW, Eda S, Winograd E (2003) Cytoadherence and sequestration in Plasmodium
falciparum: defining the ties that bind. Microbes Infect 5(10):897–909
Sim BK, Chitnis CE, Wasniowska K, Hadley TJ, Miller LH (1994) Receptor and ligand domains
for invasion of erythrocytes by Plasmodium falciparum. Science 264(5167):1941–1944
Simonsen KL, Churchill GA, Aquadro CF (1995) Properties of statistical tests of neutrality for
DNA polymorphism data. Genetics 141(1):413–429
Sjoberg K, Lepers JP, Raharimalala L, Larsson A, Olerup O, Marbiah NT, Troye-Blomberg M,
Perlmann P (1992) Genetic regulation of human anti-malarial antibodies in twins. Proc Natl
Acad Sci USA 89(6):2101–2104
Smith AJ, Humphries SE (2009) Cytokine and cytokine receptor gene polymorphisms and their
functionality. Cytokine Growth Factor Rev 20(1):43–59
Stirnadel HA, Beck HP, Alpers MP, Smith TA (1999) Heritability and segregation analysis of
immune responses to specific malaria antigens in Papua New Guinea. Genet Epidemiol
17(1):16–34
Stirnadel HA, Al-Yaman F, Genton B, Alpers MP, Smith TA (2000) Assessment of different
sources of variation in the antibody responses to specific malaria antigens in children in Papua
New Guinea. Int J Epidemiol 29(3):579–586
Tanaka T, Nakata T, Oka T, Ogawa T, Okamoto F, Kusaka Y, Sohmiya K, Shimamoto K, Itakura K
(2001) Defect in human myocardial long-chain fatty acid uptake is caused by Fat/Cd36 muta-
tions. J Lipid Res 42(5):751–759
Impact of Natural Selection Due to Malarial Disease on Human Genetic Variation 159
Tekeste Z, Petros B (2010) The Abo blood group and Plasmodium falciparum malaria in Awash,
Metehara and Ziway Areas, Ethiopia. Malar J 9:280
Timmann C, Evans JA, Konig IR, Kleensang A, Ruschendorf F, Lenzen J, Sievertsen J, Becker C,
Enuameh Y, Kwakye KO, Opoku E, Browne EN, Ziegler A, Nurnberg P, Horstmann RD (2007)
Genome-wide linkage analysis of malaria infection intensity and mild disease. PLoS Genet
3(3):e48
Timmann C, Thye T, Vens M, Evans J, May J, Ehmen C, Sievertsen J, Muntau B, Ruge G, Loag W,
Ansong D, Antwi S, Asafo-Adjei E, Nguah SB, Kwakye KO, Akoto AO, Sylverken J, Brendel
M, Schuldt K, Loley C, Franke A, Meyer CG, Agbenyega T, Ziegler A, Horstmann RD (2012)
Genome-wide association study indicates two novel resistance loci for severe malaria. Nature
489(7416):443–446
Tishkoff SA, Williams SM (2002) Genetic analysis of African populations: human evolution and
complex disease. Nat Rev Genet 3(8):611–621
Tishkoff SA, Varkonyi R, Cahinhinan N, Abbes S, Argyropoulos G, Destro-Bisol G, Drousiotou A,
Dangerfield B, Lefranc G, Loiselet J, Piro A, Stoneking M, Tagarelli A, Tagarelli G, Touma EH,
Williams SM, Clark AG (2001) Haplotype diversity and linkage disequilibrium at human
G6pd: recent origin of alleles that confer malarial resistance. Science 293(5529):455–462
Torcia MG, Santarlasci V, Cosmi L, Clemente A, Maggi L, Mangano VD, Verra F, Bancone G,
Nebie I, Sirima BS, Liotta F, Frosali F, Angeli R, Severini C, Sannella AR, Bonini P, Lucibello
M, Maggi E, Garaci E, Coluzzi M, Cozzolino F, Annunziato F, Romagnani S, Modiano D
(2008) Functional deficit of T regulatory cells in Fulani, an ethnic group with low susceptibility
to Plasmodium falciparum malaria. Proc Natl Acad Sci USA 105(2):646–651
Traherne JA (2008) Human Mhc architecture and evolution: implications for disease association
studies. Int J Immunogenet 35(3):179–192
Tse MT, Chakrabarti K, Gray C, Chitnis CE, Craig A (2004) Divergent binding sites on intercel-
lular adhesion molecule-1 (Icam-1) for variant Plasmodium falciparum isolates. Mol Microbiol
51(4):1039–1049
Turner GD, Morrison H, Jones M, Davis TM, Looareesuwan S, Buley ID, Gatter KC, Newbold CI,
Pukritayakamee S, Nagachinta B et al (1994) An immunohistochemical study of the pathology
of fatal malaria. Evidence for widespread endothelial activation and a potential role for inter-
cellular adhesion molecule-1 in cerebral sequestration. Am J Pathol 145(5):1057–1069
Ubalee R, Suzuki F, Kikuchi M, Tasanor O, Wattanagoon Y, Ruangweerayut R, Na-Bangchang K,
Karbwang J, Kimura A, Itoh K, Kanda T, Hirayama K (2001) Strong association of a tumor necrosis
factor-alpha promoter allele with cerebral malaria in Myanmar. Tissue Antigens 58(6):407–410
Uneke CJ (2007) Plasmodium falciparum malaria and Abo blood group: is there any relationship?
Parasitol Res 100(4):759–765
Valentine WN, Tanaka KR, Miwa S (1961) A specific erythrocyte glycolytic enzyme defect
(Pyruvate Kinase) in three subjects with congenital non-spherocytic hemolytic anemia. Trans
Assoc Am Physicians 74:100–110
Van den Eede P, Van HN, Van Overmeir C, Vythilingam I, Duc TN, Hungle X, Manh HN, Anne J,
D’Alessandro U, Erhart A (2009) Human Plasmodium knowlesi infections in young children
in central Vietnam. Malar J 8:249
van Hellemond JJ, Rutten M, Koelewijn R, Zeeman AM, Verweij JJ, Wismans PJ, Kocken CH, van
Genderen PJ (2009) Human Plasmodium knowlesi infection detected by rapid diagnostic tests
for malaria. Emerg Infect Dis 15(9):1478–1480
Verra F, Mangano VD, Modiano D (2009) Genetics of susceptibility to Plasmodium falciparum:
from classical malaria resistance genes towards genome-wide association studies. Parasite
Immunol 31(5):234–253
Verrelli BC, McDonald JH, Argyropoulos G, Destro-Bisol G, Froment A, Drousiotou A, Lefranc
G, Helal AN, Loiselet J, Tishkoff SA (2002) Evidence for balancing selection from nucleotide
sequence analyses of human G6pd. Am J Hum Genet 71(5):1112–1128
Vijgen L, Van Essche M, Van Ranst M (2003) Absence of the Kilifi mutation in the rhinovirus-
binding domain of Icam-1 in a Caucasian population. Genet Test 7(2):159–161
160 F. Gomez et al.
Introduction
Hearing the word “lice” will immediately terrify parents and cause school nurses to
spring into action. Pediculosis (a louse infestation) is not a new problem—lice have
coevolved with humans over millions of years, and at this moment in human history,
head lice are a worldwide epidemic. Although they can infect anyone, they are most
common among children aged 3–12 and are widely spread throughout our school
systems. According to the World Health Organization, it is thought that around
10–20 % of children are infested worldwide. In the USA alone, approximately 6–12
million infestations occur every year (Frankowski and Weiner 2002). Parents have
attacked this problem using every method from shaving their child’s head to cover-
ing the entire scalp with petroleum jelly, vinegar, and even toxic chemicals like
kerosene (Meinking 1999; Frankowski and Weiner 2002). Even though we have
been evolving with lice for millions of years, we are still struggling to understand
and eradicate these parasites.
J.F. Brinkworth and K. Pechenkina (eds.), Primates, Pathogens, and Evolution, 161
Developments in Primatology: Progress and Prospects,
DOI 10.1007/978-1-4614-7181-3_6, © Springer Science+Business Media New York 2013
162 J.M. Allen et al.
Although it is the most common, the head louse (Pediculus humanus capitis) is
just one of three types of lice that infest humans. Most similar to head lice are body
lice (Pediculus humanus humanus), perhaps more aptly called “clothing lice”
because they live principally in the clothing. The third and more distantly related
louse is the pubic louse (Pthirus pubis), which lives primarily in the pubic region.
Taxonomically, pubic lice belong to a different louse family, Phthiridae, whereas
head and clothing lice belong to the family Pediculidae. These insects are host-
specific obligate parasites that do not spend any part of their life cycle off their host.
They live around 30 days and females attach eggs to the base of a hair shaft, laying
around 3–5 eggs per day. Eggs hatch in 5–9 days, and in about 10 days nymphs
become reproductively active. These insects feed on the blood of their hosts several
times a day (Buxton 1947). Because lice have secondarily lost their wings, they can-
not fly. Instead, lice move by climbing up and down hair shafts. Their tibiae are
modified with claws that are adapted specifically for holding onto hair (Fig. 1). If a
louse is removed from its host for a long period of time, it does not survive. In fact,
most lice become so dehydrated that they cannot move after 21 h off the host
(Burgess 2004).
Although we have been studying lice for hundreds of years (Darwin 1871; Hooke
1665), research slowed significantly in the 1940s when DDT was introduced and
seemed likely to eradicate lice. In the 1990s, however, prevalence of lice increased
worldwide due to pesticide resistance, and in 1997 there was a massive outbreak of
typhus, which is transmitted by clothing lice (Raoult and Roux 1999), and louse
research found a new beginning (Burgess 2004). This research reached a new peak
in 2010 when the first louse genome (Pediculus humanus humanus) was sequenced
(Kirkness et al. 2010). The genome work revealed a number of fascinating charac-
teristics about lice. For example, clothing lice have the smallest insect genome
sequenced to date (only 108 megabases), and they do not have many of the genes
related to environmental sensing, which may be a result of their highly specialized
lifestyle as an obligate parasite. Here we review the biology and latest research on
head, clothing, and pubic lice including what these parasites can teach us about our
own evolutionary history.
Parasitic Lice Help to Fill in the Gaps of Early Hominid History 163
Head lice have long been considered an economical and societal problem rather
than a dangerous infectious disease (Hansen and O’Havier 2004; Stafford 2008).
The estimated cost of head louse infestation ranges from $367 million to $1 billion
per year. Spending on over-the-counter chemical treatments, loss of school days,
and loss of workdays for parents contribute to this high cost (Hansen and O’Havier
2004; Frankowski and Bocchini 2010). There are a number of myths and stigmas
associated with head lice that cause undue stress to parents and family members of
children infested with lice (Frankowski and Weiner 2002; Gordon 2007; Frankowski
and Bocchini 2010). For example, head lice can be found in all social classes and
are not correlated with cleanliness (Frankowski and Bocchini 2010). Head lice do
not appear to spread disease in natural populations. Although, some studies have
detected infectious bacteria in head lice (Sasaki et al. 2006a,b; Bonilla et al. 2009;
Angelakis et al. 2011 but see Parola et al. 2006), and other research has suggested
that head lice even have the capacity to transmit the infectious bacteria (Goldberger
and Anderson 1912; Murray and Torrey 1975) to date, there has not been a known
outbreak of disease caused by head louse transmission. The most common symp-
tom from a head louse infestation is pruritis (itching) caused by a bite from the
louse. In extreme cases, scratching the bite can cause an infection from common
skin bacteria (Meinking 1999), but for the most part head lice are a mere annoyance
rather than a dangerous parasite.
Manual removal of lice is a common means of treating head louse infestations,
an activity which has shown up in artwork for centuries (the term “nit-picking”
actually refers to the physical removal of lice). Recently, the most popular treatment
of head lice has been chemical. Unfortunately, pesticide resistance has increased
tremendously over the last few decades, (Burgess 2004) and in some places lice are
even resistant to more than one chemical. Currently no pesticide is 100 % effective
(Frankowski and Weiner 2002; Tebruegge et al. 2011). Other nonchemical methods
(such as manual removal) have been suggested; however, these methods have been
met with mixed success likely due to the effort required to remove all the lice
(Frankowski and Bocchini 2010). One new method of louse eradication focuses on
the temperature sensitivity of the lice. This sensitivity to temperature has been
known for some time (Buxton 1947); in fact studies have shown that lice are likely
to leave a person with a fever for a healthy person (Lloyd 1919 cited in Buxton
1947), and it has been suggested for some time that hot temperatures may be a way
to kill lice. Following this literature, a modified prototype of the LouseBuster™ has
recently been released. This appliance uses hot air to kill the lice by dehydrating
them at high temperatures. In clinical trials, this method had a 94 % success rate in
killing lice and 100 % success killing eggs on infected individuals. Thus, the
LouseBuster™ may prove to be a faster, more effective, nonchemical method for
treating head lice (Bush et al. 2011).
Head lice are most commonly transmitted from human to human by direct con-
tact (Canyon and Speare 2010). It has been long suggested that lice can be
164 J.M. Allen et al.
transmitted via fomites, objects such as combs, and pillows (Burkhart and Burkhart
2007). However, this idea has been strongly challenged, and little to no evidence of
fomite transmission has been found (Canyon and Speare 2010). Furthermore, head
lice removed from the head die quickly due to dehydration (Burgess 2004), making
fomite transmission unlikely. Unfortunately, methods to eradicate lice from schools
have been difficult due to these types of misinformed ideas about how lice move
from child to child.
An extremely controversial societal issue with lice is the “no nit” policies adopted
by some schools. These policies require children to stay out of school until they are
free of detectable nits (louse eggs). One issue with this policy is that empty nit cas-
ings (those from which the louse has already hatched) can remain on the hair long
after a louse outbreak has ended. Furthermore, eggs are incubated by body heat so
nits more than 1 cm from the scalp are unlikely viable (Frankowski and Weiner
2002). This means that nits that are farther from the scalp are likely empty and left
over from a cured infestation and may remain in the hair until it grows out. Even
more problematic, there is evidence that cases of head lice are frequently misdiag-
nosed (Pollack et al. 2000). Because of this, many children, particularly those with
longer hair, miss school unnecessarily (Gordon 2007), which inflates the total cost
of pediculosis per year. Not only are these children missing valuable class time, they
are likely to face bullying by their classmates upon return due to the stigma associ-
ated with head lice.
Clothing lice (Pediculus humanus humanus), which are also called body lice, live in
clothing fibers where they attach eggs to cloth fibers rather than hair (Buxton 1947).
They look very similar to head lice, although there are some important differences
between them. Clothing lice are generally larger than head lice and can consume a
larger blood meal (Busvine 1978; Meinking 1999; Reed et al. 2004). Clothing lice
only go to the skin to feed a few times a day, much less than head lice and likely why
they consume a larger blood meal. Additionally, unlike head lice, clothing lice are
associated with conditions of poor hygiene and are commonly found on those forced
to live in crowded situations where they lack the ability to change or wash clothes
regularly such as refugees, the homeless, soldiers, and victims of war or natural
disasters (Meinking 1999).
It has been unclear for quite some time whether head lice and clothing lice are
one species or two, and many studies have found conflicting results (Busvine 1978;
Amevigbe et al. 2000; Burgess 2004; Leo and Barker 2005; Leo et al. 2005; Light
et al. 2008). However, recent studies with more comprehensive sampling are finding
that clothing lice and head lice are in fact the same species and that clothing lice
evolve from head lice in certain conditions (such as in situations where individuals
have a considerable head louse infestation and poor hygiene; Li et al. 2010). It is
now clear that throughout our history clothing lice have opportunistically evolved
Parasitic Lice Help to Fill in the Gaps of Early Hominid History 165
repeatedly from head lice to fill a different ecological niche from that of their head
louse counterparts (Reed et al. 2004; Light et al. 2008; Li et al. 2010).
Finally, and perhaps most importantly, clothing lice are the known vectors of three
human pathogens: Rickettsia prowazekii (agent of epidemic typhus; Andersson and
Andersson 2000), Borrelia recurrentis, and Bartonella quintana (the agents of relapsing
fever and trench fever, respectively; Buxton 1947). These transmissions occur when
louse feces are unintentionally rubbed into an open wound caused by the louse bite,
most generally occurring when scratching the area of the bite (Buxton 1947). These
diseases have had a devastating impact throughout human history. For example, epi-
demic typhus may have been largely responsible for the demise of Napoleon’s Grand
Army around 1812 (Raoult et al. 2006). These diseases have been a problem not only
historically but also recently. In 1999, a large outbreak of endemic typhus broke out in
Burundi, infecting more than 100,000 people. With the growing worldwide problem of
lice due to pesticide resistance, these diseases will need to be carefully monitored.
The third type of louse that parasitizes humans is Pthirus pubis, commonly known
as the “crab” or pubic louse. Pubic lice are a sexually transmitted disease (STD),
and their presence has often been found in combination with other STD infections
(Anderson and Chaney 2009). Pubic lice are also a worldwide phenomenon.
Although it is much more difficult to calculate the level of prevalence because infec-
tions are not often reported, recent estimates suggest that 2 % of the world’s adult
population are infected with pubic lice (Anderson and Chaney 2009). Pubic lice are
found in all levels of society and among all ethnic groups (Meinking 1999).
Pubic lice live primarily in the androgenic hairs (hair that begins to grow at sex-
ual maturity) around the groin; however, these lice have been found on children in
the eyelashes, eyebrows, and edges of the hairline. In rare cases the presence of
Pthirus on children has alerted authorities to incidents of child abuse; however, this
is not common (Chosidow 2000). It is thought that Pthirus prefer hair that is more
widely spaced due to the wider spacing of their claws (Waldeyer 1900; Nuttall
1918; Fisher and Morton 1970). Interestingly, it is thought that fomite transmission
is more important in pubic lice, which may explain how children get an infestation
in their eyebrows and eyelashes without sexual contact (Meinking 1999). Similar to
human head lice, pesticides and manual removal are considered to be the primary
treatment for pubic lice (Orion et al. 2004).
The genus Pthirus has two species of lice: the human pubic louse and the gorilla
louse, Pthirus gorillae. The host associations of this genus of louse have puzzled
researchers for some time: why are humans and gorillas, but not chimpanzees, para-
sitized by Pthirus? Furthermore, although lice are very host specific, why do humans
have two genera of lice (Pediculus and Pthirus), whereas chimpanzees and gorillas
each have one (Pediculus and Pthirus, respectively; Fig. 2)? In 2007, Reed et al.
conducted molecular dating analyses on gorilla and human pubic lice and found that
166 J.M. Allen et al.
Fig. 2 The coevolutionary history of humans (Homo), chimpanzees (Pan), and gorillas (Gorilla)
and their lice. The primate lineages are indicated by thin black lines and black boxes that depict the
longevity (box height) and the species richness (box width) of the primate genera known from
physical evidence (either the fossil record or extant species). Parasite lineages are indicated with
thick gray bars with the lighter gray representing Pediculus and the darker gray Pthirus. Dotted
lines indicate possible coevolutionary scenarios that remain unclear due to lack of data. The most
recent estimated divergence between gorillas and the lineage leading to humans and chimpanzees
is shown at 9.2 mya; however, we also show the possible divergence (and consequently the cospe-
ciation event) between these lineages at 13 mya incorporating the gorilla-like fossil Chororapithecus
abyssinicus (dashed lines at 13 mya). We further show the host switch event 3–4 mya by lice in the
genus Pthirus from the gorilla lineage onto the hominin lineage. The extinct hominin genus
Paranthopus and its possible association with both Pediculus and Pthirus are shown at approxi-
mately 2 mya. Events on the right represent our current knowledge of hominin history including
time ranges for the origin of clothing use by Homo sapiens, the first putative shelters, the earliest
known butcher marks, and the development of bipedality (as indicated from the louse data).
Asterisks mark isolated fossil finds. Species abbreviations are as follows: Ar. ka = Ardipithecus
kaddaba, Ar. ra = Ardipithecus ramidus, Au. ana = Australopithecus anamensis, Au.
afa = Australopithecus afarensis, and Au. afr = Australopithecus africanus. O = Orrorin tugenensis,
S = Sahelanthropus tchadensis, K = Kenyanthropus platyops, and A = Australopithecus bahrel-
ghazali. (1) McBrearty and Jablonski (2005). (2) Foley (2002). (3) Pickford et al. (1988). (4) Suwa
et al. (2007). (5) Strait et al. (1997). (6) Leakey et al. (2001). (7) Brunet et al. (1995). (8) Kimbel
et al. (2006). (9) WoldeGabriel et al. (2009). (10) Haile-Selassie (2001). (11) Pickford et al. (2002).
(12) Brunet et al. (2002). (13) Toups et al. (2011). (14) Rantala (1999). (15) McPherron et al.
(2010). (16) Reed et al. (2007). (17) Pickford et al. (2002)
Parasitic Lice Help to Fill in the Gaps of Early Hominid History 167
these two species were sister taxa. Additionally, these two taxa diverged only 3–4
million years ago (mya), much more recently than the gorilla/human-chimp split,
which is estimated at around 9 mya (Wilkinson et al. 2011). This finding was
extremely interesting as it shed light onto two details of human evolutionary history
that were previously unknown. Here we go into more detail about the biology of
Pthirus to discuss what this host switch tells us about our ancestors 3–4 mya.
Sucking lice have been coevolving with their hosts for at least the last 65 million
years and likely much longer (Light et al. 2010). Due to their obligate nature and the
fact that they mostly move between hosts via direct contact, these blood-feeding lice
have been used to give us clues about their hosts’ evolutionary history not easily
gleaned from the fossil record (e.g., behavior). Interestingly, this idea dates back to
Darwin (1871) who in On the Decent of Man wrote:
“…and the fact of races of man being infested with parasites which appear to be specifically
distinct might fairly be urged as an argument that the races themselves ought to be classed
as distinct species.”
Although the idea that races of humans represent different species is no longer
entertained and even the concept of race has dramatically changed since Darwin’s
time, the idea that lice can tell us about our evolutionary history is now well accepted
and gaining momentum (Whiteman and Parker 2005; Hypsa 2006; Nieberding and
Olivieri 2007; Reed et al. 2009).
In 2003, it became apparent upon genetic examination of recently collected
human head and clothing lice that there were several distinct lineages of lice (Kittler
et al. 2003). Reed et al. (2004) used fossil calibrations for the split between humans
and chimpanzees (5.6 mya) and the split between great apes and Old World mon-
keys (22.5 mya) to estimate the divergence time between these human louse lin-
eages. They found that the youngest of these lineages splits around 1.18 mya, which
is similar in age to the ancestor of Homo sapiens and H. neanderthalensis. The other
louse lineage is even older, suggesting its origin may date back to a H. erectus-like
host. Because head lice are primarily transmitted through direct contact, this finding
suggests that modern humans came into contact with archaic hominin species and
picked up distinct lineages of head lice. Although we do not know which archaic
humans our ancestors came into contact with, the timing of the divergence of the
ancient head louse lineages is consistent with contact with H. neanderthalensis and
H. erectus. Furthermore, while the type of contact between different hominin spe-
cies is unknown, recent work sequencing the H. neanderthalensis genome found
evidence of interbreeding between H. neanderthalensis and H. sapiens (Green et al.
2010; Yotova et al. 2011). Similar types of contact between H. sapiens and H. erec-
tus would have been sufficient for the transfer of lice and may explain the existence
of ancient lineages of head lice on modern humans.
168 J.M. Allen et al.
Because clothing lice live exclusively in the clothing, they are thought to have
evolved only after humans began to wear clothes, and it has long been proposed that
dating the origin of clothing lice could give us a date by which H. sapiens must have
been wearing clothing (Kittler et al. 2003, 2004). Previous estimates of when
humans started wearing clothing were based on the emergence of eyed needles
(which suggests complex clothing had already been developed) around 40,000 mya
(Delson et al. 2000) and sometime after the loss of body hair as late as 1.2 mya
(Rogers et al. 2004; based on molecular evidence) and as early as 3 mya (Reed et al.
2007; detailed below). Recent molecular evidence from clothing lice suggests that
clothing use originated between 83,000 and 170,000 years ago, which is earlier than
previously proposed, and suggests that clothing use by H. sapiens likely originated
before they moved out of Africa (Toups et al. 2011). This clothing use may have
enabled modern humans to more readily move into colder climates as they migrated
out of Africa and eventually throughout the world.
The coevolutionary relationships between great apes and their lice have been
worked out morphologically and molecularly and are illustrated in Fig. 2. Humans
and chimpanzees share lice in the genus Pediculus as sister taxa, and humans and
gorillas share lice in the genus Pthirus (Fig. 2). The split between human and chim-
panzee lice was estimated to be 5–7 mya (Reed et al. 2004, using mitochondrial
genes; Light et al. 2008, using a multigenic approach with mitochondrial and
nuclear genes), strongly suggesting cospeciation between these two lice and their
primate hosts (humans and chimpanzees also are believed to have diverged at this
time; Wilkinson et al. 2011). On the other hand, Reed et al. (2007) found that the
Pthirus and Pediculus are sister taxa and they diverged ~13 mya, long before the
presumed 7 mya split between gorillas and the other African apes (Fig. 2). Reed
et al. (2007) hypothesized an evolutionary scenario in which there was a louse
duplication (or speciation) event on the African ape common ancestor to humans,
chimpanzees, and gorillas. In other words, one louse species would have diverged
into two on the ape common ancestor. More recent refinement of the somewhat
troublesome great ape molecular clock has pushed the gorilla divergence back to
9.2 mya (Wilkinson et al. 2011). Additionally, a recent fossil find of the gorilla-like
ape Chororapithecus abyssinicus dating to 10–11 mya (Suwa et al. 2007) suggests
that the great ape molecular clock may still be underestimating the divergence of the
gorillas by millions of years. If true, the presumed louse duplication event ~13 mya
suggested by Reed et al. (2007) could have actually been a cospeciation event, sug-
gesting that the divergence time between gorillas and the other African apes occurred
~13 mya, far earlier than currently thought (Fig. 2).
Reed et al. (2007) also examined the history of the two species of Pthirus, one on
gorillas and the other on humans, which diverged 3–4 mya. Reed et al. (2007)
hypothesized that after Pthirus and Pediculus diverged ~13 mya, the Pthirus lineage
remained on ancestral gorillas but went extinct on the common ancestor of humans
and chimpanzees, and the Pediculus lineage remained on the common ancestor of
humans and chimpanzees but went extinct on ancestral gorillas. Then, approxi-
mately 3–4 mya, the Pthirus lineage from ancestral gorillas switched to a human
ancestor (Fig. 2). This type of host switch is not uncommon; there have been a
Parasitic Lice Help to Fill in the Gaps of Early Hominid History 169
Reed et al. (2007) postulated that a Pthirus-type louse switched hosts from archaic
gorillas to hominins approximately 3–4 mya. It is interesting to consider what was
necessary for this host switch to have been successful: there had to have been a
niche for Pthirus to occupy. Studies of chewing lice have found that their mouth-
parts (which grip the hair) are highly adapted and specialized to the hair of their
hosts (Reed et al. 2000), so it is likely that hair type is similarly important to suck-
ing lice. Rather than using their mouthparts to grip hair, sucking lice use their
claws, which are also highly specialized. Pthirus, in particular, is highly adapted
to hairs in the pubic regions as these hairs are more widely spaced, which match
the wider spacing of their claws (see below). As stated previously, pubic hair is a
type of androgenic hair—hair that grows in response to increased levels of andro-
gens circulating in the human body at sexual maturity (Randall 2008). Proposed
functions of pubic hair (pheromonal and visual signaling; Randall 2008) could
have only come into play after the loss of typical ape body hair. Additionally, the
invasion of a new host would have been far more likely if Pediculus had already
been confined to the head by the loss of functional body hair, leaving competitor-
free regions available to Pthirus. We hypothesize that the loss or reduction of
body hair as well as the development of androgenic hair would have facilitated the
success of this host switch and therefore suggest that human hair loss and the gain
of androgenic hair had occurred by 3–4 mya, a date that is much older than other
predictions.
Among primates, humans are unique in their apparent nakedness. Humans,
however, are not actually hairless. They have a similar number and density of
hair follicles as other great apes, but the hairs are much finer (i.e., smaller in
diameter) and shorter and offer little protection or insulation (Kushlan 1985;
Amaral 1996; Rantala 1999). There is a great variety of hypotheses ranging from
the bizarre to the pedestrian as to why humans had such a drastic reduction in
body hair (reviewed in Rantala 2007). Many of these hypotheses are directly
related to the Pthirus host switch because they either incorporate habitat (as dis-
cussed below) and thermodynamics (which is closely tied to habitat) or attempt
to establish the timing of hair reduction. Habitat and thermodynamics are impor-
tant to the cooling device, bipedality, hunting, vestiary, allometry, and other
hypotheses of why humans lost their body hair. The timing of hair loss is incor-
porated to some extent in any hair loss hypothesis, but it is particularly important
in the clothing, vestiary, and ectoparasite hypotheses.
170 J.M. Allen et al.
However, the effective hairlessness of humans is not just a result of hair density
but also hair size. In contrast to humans, other similar-sized and larger apes (orang-
utans and gorillas) have substantial body hair. Thus, hominins appear to have exag-
gerated the typical primate strategy of shedding metabolic heat via reduced
insulative effectiveness of body hair (the heavy sweating of humans and patas mon-
keys does not appear to be typical of primates; Amaral 1996) by reducing hair size.
It seems likely that this was in answer to an additional metabolic heat load beyond
that experienced by typical apes. Although it is impossible to say with any degree of
certainty what this additional heat load was, the development of bipedalism, a more
energetically efficient mode of locomotion than knuckle walking (Sockol et al.
2007), hints that increased daily travel may have been important to the basal hom-
inin niche and that hair loss may have occurred very early on. The extra metabolic
heat produced by travel through forests could have been shed by decreasing body
hair insulation without the costs of nakedness associated with savanna environ-
ments. Based on the timing of the Pthirus host switch, the habitat in which this
switch likely occurred (see below), and the problems of hairlessness in open habi-
tats, hair loss in the hominin line almost certainly occurred in a forested habitat and
was complete and effectively irreversible by the time savanna habitats were fully
utilized. The human dependence on sweating as a cooling mechanism likely
occurred long after hair loss to deal with the additional heat loads in open habitats
as sweating is less effective in the humid still air of forests than in drier more open
habitats (Newman 1970; Montagna 1972).
Other than lice, the only line of evidence that helps establish the timing of hair
loss in the hominin line is genetic. The human melanocortin 1 receptor (MC1R)
gene is involved in human skin coloration. By looking at the neutral variation in this
gene, Rogers et al. (2004) estimated that human skin has been exposed to strong
sunlight for at least 1.2 my. Therefore, based on the MC1R data, human ancestors
became both hairless and began living in the savanna between 1.2 mya and 6–7 mya
(the chimpanzee/hominin split). The lice data are consistent with the MC1R data
but give a narrower range of 3–4 mya from the Pthirus switch to the 6–7 million
year split between the human and chimpanzee lineages.
The timing of hair loss is particularly central to the clothing, vestiary, and ecto-
parasite hypotheses. The clothing hypothesis (Glass 1966) is similar to the vestiary
hypothesis (Kushlan 1985) in that they both posit that clothing superseded the insu-
lative value of body hair and hair loss occurred with or after the invention of cloth-
ing. However, while the vestiary hypothesis holds (erroneously, as discussed above)
that the loss of body hair was advantageous during the hot days and that clothing
replaced the need for body hair during the cool nights, the clothing hypothesis
argues that after clothing was invented, body hair disappeared as it was no longer
needed (Glass 1966; however, Glass does not propose a reason for the invention and
use of clothing by hominins with functional coats of body hair). The louse and
MC1R data estimates for both hairlessness (Pthirus, >3–4 mya in Reed et al. 2007;
MC1R, >1.2 mya in Rogers et al. 2004) and the invention of clothing (Pediculus,
0.08–0.17 mya in Toups et al. 2011) indicate that clothing had nothing to do with
the evolution of hairlessness in hominins.
172 J.M. Allen et al.
The ectoparasite hypothesis states that when hominins first established long-term
habitations, they were beset with new types of ectoparasites, such as fleas, that com-
pleted their life cycles in the living space but off the body of the host (Rantala 1999).
Thus, the loss of body hair was a defense against increased parasite loads encouraged
by the establishment of a home base. In apparent conflict with the adaptation-
against-ectoparasites hypothesis is the presence of pubic hair (Pagel and Bodmer
2003). Pubic hair, however, may play an important role in sexual selection and thus
may have been selected for in spite of its ability to shelter ectoparasites. The moist
and humid environment of the pubic region (due to an increased density of sweat
glands; Stoddart 1990) is favorable to pheromonal signaling (Guthrie 1976), and
pubic hair could have initially functioned in pheromonal signaling (Randall 1994),
visual signaling (Randall 2008), or both. Rantala (1999) associates the beginnings
of long-term settlements with an increase in cooperative hunting and places both
developments at ~1.8 mya based on excavations of Homo habilis artifacts at Olduvai
Gorge. While this estimate is consistent with the hairlessness range provided by the
MC1R gene (>1.2 mya), it is far later than the hairlessness estimate provided by the
Pthirus host switch.
The two genera of human lice (Pediculus and Pthirus) occupy distinct niches on the
body (head/clothing and pubic region, respectively). Many researchers have won-
dered why Pediculus and Pthirus do not co-occur and are apparently isolated to
these different regions especially given their similar biology (Howlett 1917).
Hypotheses have included Pthirus having a preference for darkness and moist areas
(Nuttall 1918; however, this idea is not accepted as Pthirus survives on eyelashes
and eyebrows) and that differences in hair spacing have geographically restricted
these lice because Pediculus cannot adequately grasp the hairs of the pubic region,
and Pthirus cannot grasp the hairs on the head. Both genera of lice have been found
occasionally occupying and surviving in other regions of the body (see below), but
they do not seem to be successful in these areas. Of these hypotheses, hair spacing
seems to be the most likely explanation for restricting these two types of lice to their
respective habitats.
Schwartz and Rosenblum (1981) found that there are fewer hairs per unit of body
surface in larger primates compared to smaller primates. If the Reed et al. (2007)
hypothesis that the human pubic louse (Pthirus pubis) is a descendent of gorilla lice
is true, then based on Schwartz and Rosenblum’s (1981) findings, we can postulate
that Pthirus was adapted to living among widely spaced hairs because gorillas are
the largest extant primate. Early studies of Pthirus support this idea. Pthirus uses its
second and third pair of legs to cling to host hair, and these legs, when stretched
apart, span a distance of 2 mm (Waldeyer 1900; Nuttall 1918; Fisher and Morton
1970). It just so happens that hairs in the pubic region are also distributed 2 mm
apart (Waldeyer 1900; Nuttall 1918). Furthermore, the number of hairs present in
Parasitic Lice Help to Fill in the Gaps of Early Hominid History 173
the pubic region (34 hairs/cm2) is significantly less than the head (220 hairs/cm2;
Waldeyer 1900; Payot 1920). All in all, Pthirus appears to prefer body regions with
widely spaced hairs for better grasping as well as for ease of flattening itself against
the skin (Burgess et al. 1983; Burgess 1995; Nuttall 1918; Buxton 1947; Fisher and
Morton 1970). According to measurements made by Schultz (1931), chimpanzee
hair density is more similar to humans than to gorillas. That, in addition to the lack
of pubic-type hair on chimpanzees, may help explain why there are no Pthirus spe-
cies currently parasitizing chimpanzees.
The Pthirus preference for widely spaced hairs is likely why this genus can
be occasionally found in other sparsely haired areas on the human body, such as
the margins of the scalp, eyebrows, eyelashes, and areas of the trunk such as the
chest, stomach, and thighs (if sufficient body hair is present; Burgess 1995, and
references therein; Buxton 1941, 1947; Elgart and Higdon 1973). Pediculus, in
comparison, is rarely found in the pubic region of humans (Busvine 1944).
Pediculus schaeffi, the louse on chimpanzees louse, is more catholic in habitat
choice than Pediculus humanus and can be found almost anywhere on a chim-
panzee host but favors the groin, underarms, and head (D. Cox, pers. comm.). It
is likely that the Pediculus found on hominins before the loss of body hair was
similarly widely spread but was restricted to the head region during the hominin
denudation and subsequently prevented from spreading to androgenic hair
because the larger spacing of pubic hair made movement from one hair to
another difficult.
Differences in mobility also may prevent Pediculus from traveling to the pubic
region as often as Pthirus appears to move to other parts of the body. Although
several studies have found that Pediculus moves faster than Pthirus when displaced
from the body (Nuttall 1918; Busvine 1944), Pthirus does tend to wander more
(Burgess et al. 1983). Furthermore, head lice are recognized as being rather picky in
how they move from hair to hair, suggesting that they are unlikely to move readily
to foreign objects or fomites (Canyon et al. 2002). Closer examination of the first
tarsal claws (which may facilitate movement when lice are not in contact with hair)
of both Pediculus and Pthirus reveals why there may be differences in mobility
between these two genera (Ubelaker et al. 1973). The inner surface of the first tarsal
claw in Pthirus is serrated, allowing for traction even on smooth surfaces, whereas
in Pediculus the inner surface of the claw is smooth and the lice are unable to move
without hair follicles or roughened surfaces (Nuttall 1918; Ubelaker et al. 1973;
Burkhart and Burkhart 2000). This simple difference, along with preferential move-
ment patterns, may restrict Pediculus from moving easily on smooth, non-haired
substrates. Pthirus, on the other hand, with their serrated first tarsal claws, may be
able to move much more easily on non-haired substrates, thus allowing them to
reach other parts of the body such as the perimeter of the scalp. The biology of these
two parasites supports the idea that human ancestors had not only lost their body
hair by the host switch 3–4 mya (isolating Pediculus in the head region) but that
early hominins had also developed androgenic hair, providing a suitable environ-
ment for Pthirus.
174 J.M. Allen et al.
Given the fossil species currently known, the most parsimonious scenario explain-
ing the appearance of Pthirus in the human lineage is a host switch directly from
gorilla ancestors to human ancestors. This scenario strongly implies that the two
ancestral host species came into repeated and close contact, which further implies
significant overlap in habitat. By combining data from the fossil record, paleocli-
mate, extant species, and the Pthirus host switch, we can augment the current think-
ing of the habitat and habits of human ancestors.
The fossil record of nonhuman African apes is abysmal to say the least. There
is currently only one known chimpanzee fossil, which lived 0.5 mya (McBrearty
and Jablonski 2005); one gorilla fossil from 5 to 6 mya (Pickford et al. 1988);
and the very gorilla-like Chororapithecus abyssinicus from 10 to 11 mya (Suwa
et al. 2007; Fig. 2). The reasons for the dearth of these fossils are likely due to
several factors (Cote 2004). For one, apes were an uncommon component of the
fauna in any region, and African fossil sites commonly produce fewer speci-
mens than Eurasian fossil sites. Therefore, a site must produce a large number
of fossils if any apes are expected to be represented in the first place. Second,
with the exceptions of Samburupithecus kiptalami and Nakalipithecus nakay-
ami, which might have been adapted to drier forests (Kunimatsu et al. 2007),
nonhuman African apes are, and appear to always have been, tightly associated
with moist tropical forests. The wet acidic soil in these types of habitats is much
more conducive to quick bone decomposition than fossilization, so it is likely
that very few specimens were fossilized to begin with (Kingston 2007). Added
to those problems is the fact that sites currently under moist tropical forests are
seldom found or excavated, partly due to a lack of exposed strata and partly due
to the political insecurity that often inflames those regions, reducing safety for
international teams and handicapping intranational capacity for, and interest in,
research.
It is generally thought that the nonhuman African apes are conservative in body
form and habits contrasting with the hominins that stumbled upon a new behavior/
body form (bipedalism) that led to their subsequent radiation into a speciose and
relatively diverse group. That assumption is supported by the dearth of non-hominin
ape fossils in Africa, which indicates they were restricted to the wet forests that are
particularly hostile to fossil formation (Kingston 2007), and by the nature of the few
fossils that have been found. Additionally, the morphology and wear of the
Chororapithecus abyssinicus fossil from 11 to 10 mya (Suwa et al. 2007) and the
gorilla tooth from 5 to 6 mya (Pickford et al. 1988), as well as their respective faunal
assemblage contexts, suggest that gorillas have been conservative in diet and habitat
over the period of time during which the hominin group was rapidly developing
novel traits and habits.
Molecular work has also indicated the conservatism of the gorilla lineage.
Thalmann et al. (2007) have estimated that eastern and western gorillas (Gorilla
beringei and G. gorilla, respectively) diverged 0.9–1.6 mya with very little subse-
quent gene flow between those two species. Thalmann et al. (2007) also suggest that
Parasitic Lice Help to Fill in the Gaps of Early Hominid History 175
the genetic divergence between the eastern and western forms is small enough to
unite the two into a single species. The gene flow between the two groups is low
enough that it is likely not the cause of their genetic similarity but a result of it. With
the apparent conservatism of the gorilla lineage in mind, we can cautiously use the
natural history of extant gorillas to inform us of the probable habits of their ances-
tors and examine how that information fits into and expands the understanding of
our ancestors.
Gorillas today range across forested tropical Africa (with a large interruption
between the eastern and western species) from lowland rainforest to high-altitude
montane forest. In spite of these wide longitudinal and altitudinal ranges, both spe-
cies share similar diets based on succulent herbaceous vegetation (Kingdon 1974).
The diet often incorporates more fruit in areas where fruit is available, but herba-
ceous vegetation still forms a large portion of the diet, retaining its primary impor-
tance especially as a fallback food in times of fruit scarcity (Yamagiwa and Basabose
2006).
The importance of fibrous herbaceous foods makes gorillas more independent of
often unreliably fruiting trees than the two chimpanzee species (Pan troglodytes and
P. paniscus); however, it also limits their available habitat to moist forests with
enough sunlight penetrating the canopy to support a rank herbaceous understory
(Schaller 1963). While swidden agriculture produces ample areas of lush secondary
growth, prior to agriculture, suitable gorilla foraging areas would have been limited
to montane forests, river edges, treefall gaps, elephant tramples, and the like
(Schaller 1965a), with feeding and use by gorillas likely slowing succession and
extending the usable life and possibly the size of temporary clearings (Plumptre
1994). Even if the gorilla lineage had significantly different dietary preferences than
extant gorillas during the host switch of Pthirus 3–4 mya, it is unlikely that the dif-
ferences would have a meaningful impact on our analysis as the conservatism of
body and tooth form and lack of fossils indicate a folivorous/frugivorous diet in a
moist forest.
During the 3–4 mya range given for the Pthirus host switch, there were 1–4
hominin species present (Fig. 2) depending on the validities of species identifica-
tions: Australopithecus anamensis, A. afarensis, A. bahrelghazali, and Kenyanthropus
platyops (Fig. 2). K. platyops (3.5 mya) is known from only one locality and the
skull upon which the identification is based is severely fragmented and distorted
(Leakey et al. 2001). Therefore, the identity of the specimen as a new genus
(Kenyanthropus), a new species within Australopithecus, or another A. afarensis
specimen is controversial (White 2003; Spoor et al. 2010). However, if K. platyops
is a valid species, it is potentially ancestral to both Homo and Paranthropus (robust
australopithecines) and lived in a well-watered forest or woodland (Leakey et al.
2001; Strait and Grine 2004).
Australopithecus bahrelghazali is another species known only from a single fos-
sil from 3.5 mya (Brunet et al. 1995; Brunet 2010) and, like K. platyops, is contro-
versial as to whether it is a separate species or an unusual A. afarensis (Kimbel et al.
2006; Guy et al. 2008). This is a unique find because it is the only australopithecine
found in Chad rather than East or Southern Africa. Unfortunately, the state of the
176 J.M. Allen et al.
australopithecines could presumably still utilize warm moist forests, use of montane
forests of australopithecines is not likely given the loss of body hair by this point
and the cold temperatures experienced in these forests. Unfortunately, fossil evi-
dence of wet forest use by hominins will likely be as difficult to come by as fossil
evidence of the other great apes that are restricted to wet forests. As mentioned
before, the fossil record is almost silent even on the subject of common chimpan-
zees, which venture into drier habitats (such as woodlands and scrublands) than
gorillas but remain largely tied to closed canopy forests. Thus, if restricted to fossil
evidence, our picture of hominin evolution and habits is severely limited by the
taphonomic processes in the moist forests that form the origin of African ape
diversity.
Our suggestion that australopithecines expanded their habitat from drier wooded
areas into wet tropical forests introduces several more possibilities. Habitat use
could have been seasonal with australopithecines predictably moving from drier
habitats to wet forest areas and back. Alternately, a widely spread generalist species,
as A. afarensis has been proposed to be, might have populations permanently inhab-
iting entirely different habitats. In this case, the question becomes which habitat was
preferred, i.e., which, if any, was able to support higher densities of hominins. A
third possibility is that one of the habitats was primary with the other being utilized
only in times of drought, etc. The importance of moist forests with poor fossiliza-
tion conditions to hominins raises the possibility that it was not the dry woodlands
that were the center of hominin radiation; rather the radiation occurred in the rain
forests with a minority of the species expanding out into more xeric areas to be fos-
silized and finally found. If this is true, we may still be missing a large portion of
hominin diversity.
Of course, for Pthirus to have switched hosts successfully, sharing habitats
would have been insufficient for transfer—far more intimate contact would have
been required. However, the contact would not have to be as intimate as most people
seem to gleefully assume. Although hybrids between different guenon monkey spe-
cies are known (Struhsaker et al. 1988; de Jong and Butynski 2010), there are far
more likely scenarios than two species as divergent as archaic gorillas and australo-
pithecines having sexual contact.
Pthirus are known to be transmitted via fomites (Meinking 1999) and therefore
could have potentially switched hosts if a hominin used an abandoned louse-infested
archaic gorilla nest (Reed et al. 2007). All extant great apes including orangutans
fashion nests (Schaller 1965b), so it is reasonable to assume that all apes 3–4 mya
also made nests. While this scenario is possible, it seems unlikely. Great ape nests
(aside from those of humans) are constructed swiftly for a single use and are simple
rudimentary structures. This is especially true for gorilla nests made on the ground,
which have better rims than bottoms, provide little if any padding, and are typically
constructed in 1 min (though the minority that are made in the trees are more sub-
stantial; Schaller 1965a). Unfortunately (but not unexpectedly), there is no surviv-
ing evidence of australopithecine nests or nest use. However, if they practiced the
single-use pattern typical of great apes, there would have been little reason for
individuals who did not reuse their own nests to reuse the nests of another species.
Parasitic Lice Help to Fill in the Gaps of Early Hominid History 179
On the other hand, if australopithecines reused nests like at least some of the later
Homo spp., the reuse would probably have been motivated by increased effort
required to build more complex and functional structures. Again, the reuse of old
slipshod gorilla-type nests would have been unlikely.
A more probable scenario for contact between human and gorilla ancestors is the
existence of mixed foraging groups. Though rare compared to multiple species
associations in monkeys (Yamagiwa and Basabose 2006), mixed foraging groups
have been observed containing gorillas and chimpanzees. Interactions from avoid-
ance (Yamagiwa et al. 1996) to obliviousness (Kuroda et al. 1996) have been seen
taking place in mixed groups of apes (even from a distance of 3 m; Stanford 2006).
The nature of the interaction likely depends on food availability and level of com-
petition as well as the individual personalities of those involved. Habituated gorilla
and chimpanzee troops tend to ignore human observers, but some physical interac-
tions (e.g., playing, bluffing, and testing) have occurred, normally with young ani-
mals. In primate mixed foraging groups, interactions are most frequent between
young animals and can involve play.
Mixed foraging groups of gorilla and human ancestors would have given oppor-
tunities for the Pthirus switch through play or grooming; however, primate interspe-
cific interactions are rare even where multispecies associations are common.
Aggressive interactions are the most numerous interactions, with play being rela-
tively unusual and grooming being extremely rare and of short duration (Ihobe
1990; Heymann and Buchanan-Smith 2000). Because juveniles are typically the age
group involved in interspecific interactions and young hominins would have not yet
developed the pubic hair that forms the current habitat of Pthirus on humans, the
ancestral lice would have to be passed to an adult host before they could become
established on the new host species. The possibility of a host switch through mixed
foraging groups does exist, but because aggression and play rarely involve physical
contact (Rose 1977), which occurs quickly and generally between juveniles who
would then have to pass Pthirus to an adult host, social interaction is a less likely
route of host switching than the last possibility: the consumption of archaic gorilla
meat by human ancestors.
Other than the gorillas, all the African great apes hunt and consume meat,
although only humans have managed to prey on animals of similar and larger
body sizes through the use of relatively sophisticated tools. Given the body size of
A. afarensis (♀, ~29 kg; ♂, ~45 kg; McHenry 1994) compared to that of gorillas
(♀, ~80 kg; ♂, ~169 kg; Smith and Jungers 1997) as well as the dangerous nature
of enraged gorillas and relatively simple tools used by the australopithecines, it
seems unlikely that hunting archaic gorillas by early hominins would be a particu-
larly effective or common food acquisition strategy. It is far more likely that scav-
enging on gorilla carcasses led to the kind of contact most conducive to a louse
host switch: repeated close contact over a substantial period of time. Additionally,
lice are extremely sensitive to environmental conditions and readily abandon dead
hosts. The desperate situation of lice on a dead or dying host makes the switch to
any available host, even one of the incorrect species, much more likely than casual
contact between a living native host and a potential novel host.
180 J.M. Allen et al.
The presence of likely butcher marks on the bones of large mammals contem-
porary with A. afarensis indicates that scavenging was likely an important and
effective component in their feeding repertoire (McPherron et al. 2010, 2011; but
see Domínguez-Rodrigo et al. 2010, 2011). Although the validity of these butcher
marks does not determine the level of carnivory by A. afarensis (Domínguez-
Rodrigo et al. 2010), simple tools would have enabled both more efficient pro-
cessing of meat than allowed by primate dentition alone and the transportation of
meat away from the main carcass to a location with less predation danger. Thus,
the most likely scenario is that a hominin habitually used moist tropical forests far
more than previously realized or shown by the fossil record and opportunistically
scavenged meat from gorilla carcasses. This feeding resulting in contact with
Pthirus that was frequent enough to establish a population of Pthirus on hominins
3–4 mya.
Conclusion
There has been much research into the biology, epidemiology, and the evolution-
ary history of lice. These parasites have bedeviled human and nonhuman primates
alike for millions of years, and the increase in louse prevalence over the last 20
years suggests they will continue to parasitize humans for some time yet. Due to
their obligate host-specific nature, we can use these parasites to inform us about
human evolutionary history and gather information that is not available in the host
fossil record, providing an unexpected benefit to an otherwise bothersome
parasite.
Although most great ape lice have strictly cospeciated with their great ape hosts,
Pthirus pubis (the human pubic louse) has a different evolutionary history. Pthirus
switched to the human lineage 3–4 mya from an archaic gorilla. The biology of
Pthirus suggests that for this host switch to have occurred, suitable habitat had to be
available, which indicates that hominins had not only lost their body hair but also
developed androgenic hair by 3–4 mya.
Because gorillas are conservative in their habitats and diet, we postulate that
this likely means that these archaic hominins were using similar habitat as goril-
las (moist forest habitat). The best candidate for this host switch was
Australopithecus afarensis (based on our current knowledge). A. afarensis pos-
sibly butchered large mammal carcasses during this time, presenting a scenario
of A. afarensis scavenging archaic gorilla meat and Pthirus likely switching to
A. afarensis from a dead gorilla host. Pthirus then continued to evolve with the
hominin lineage as a sexually transmitted disease due to their placement on the
body, explaining the presence of two genera of lice (Pthirus and Pediculus) on
extant humans today.
Parasitic Lice Help to Fill in the Gaps of Early Hominid History 181
References
Amaral LQ (1996) Loss of body hair, bipedality and thermoregulation: comments on recent papers
in the journal of human evolution. J Hum Evol 30:357–366
Amevigbe MDD, Ferrer A, Champorie S, Monteny N, Deunff J, Richard-Lenoble D (2000)
Isoenzymes of human lice: Pediculus humanus and P. capitis. Med Vet Entomol 14:419–425
Anderson AL, Chaney E (2009) Pubic lice (Pthirus pubis): history, biology and treatment vs.
knowledge and beliefs of US college students. Int J Environ Res Public Health 6:592–600
Andersson JO, Andersson SG (2000) A century of typhus, lice and Rickettsia. Res Microbiol
151:143–150
Angelakis E, Rolain JM, Raoult D, Brouqui P (2011) Bartonella Quintana in head louse nits.
FEMS Immunol Med Mic 62(2):244–246
Bonilla DL, Kabeya H, Henn J, Kramer VL, Kosoy MY (2009) Bartonella quintana in body lice
and head lice from homeless persons, San Francisco, California, USA. Emerg Infect Dis
15(6):912–915
Bonnefille R (2010) Cenozoic vegetation, climate changes and hominid evolution in tropical
Africa. Glob Planet Change 72:390–411
Bonnefille R, Potts R, Chalié F, Jolly D, Peyron O (2004) High-resolution vegetation and climate
change associated with Pliocene Australopithecus afarensis. Proc Natl Acad Sci
101:12125–12129
Brace CL, Montagu A (1977) Human evolution, 2nd edn. Macmillan, New York
Brunet M (2010) Two new Mio-Pliocene Chadian hominids enlighten Charles Darwin’s 1871 pre-
diction. Phil Trans Roy Soc B 365:3315–3321
Brunet M, Beauvilain A, Coppens Y, Heintz E, Moutaye AHE, Pilbeam D (1995) The first australo-
pithecine 2,500 kilometers west of the Rift Valley (Chad). Nature 378:273–275
Brunet M, Guy F, Pilbeam D, Mackaye HT, Likius A, Ahounta D, Beauvilain A, Blondel C,
Bocherensk H, Boisserie J-R, De Bonis L, Coppens Y, Dejax J, Denys C, Duringer P, Eisenmann
V, Fanone G, Fronty P, Geraads D, Lehmann T, Lihoreau F, Louchart A, Mahamat A, Merceron
G, Mouchelin G, Otero O, Campomanes PP, De Leon MP, Rage J-C, Sapanetkk M, Schusterq
M, Sudrek J, Tassy P, Valentin X, Vignaud P, Viriot L, Zazzo A, Zollikofer C (2002) A new
hominid from the Upper Miocene of Chad, Central Africa. Nature 418:145–151
Burgess IF (1995) Human lice and their management. Adv Parasitol 36:271–342
Burgess IF (2004) Human lice and their control. Annu Rev Entomol 49:457–481
Burgess I, Maunder JW, Myint TT (1983) Maintenance of the crab louse, Pthirus pubis, in the
laboratory and behavioral studies using volunteers. Community Med 5:238–241
Burkhart CN, Burkhart CG (2000) The route of head lice transmission needs enlightenment for
proper epidemiologic evaluations. Int J Dermatol 39(11):878–879
Burkhart CN, Burkhart CG (2007) Fomite transmission in head lice. J Am Acad Dermatol
56:1044–1047
Bush SE, Rock AN, Jones SL, Malenke JR, Clayton DH (2011) Efficacy of the LouseBuster, a new
medical device for treating head lice (Anoplura: Pediculidae). J Med Entomol 48(1):67–72
Busvine JR (1944) Simple experiments on the behaviour of body lice (Siphunculata). Proc Roy Ent
Soc Lond 19:22–26
Busvine JR (1978) Evidence from double infestations for the specific status of human head lice
and body lice (Anoplura). Syst Entomol 3:1–8
Buxton PA (1941) On the occurrence of the crab-louse (Phthirus pubis: Anoplura) in the hair of
the head. Parasitology 33:117–118
Buxton P (1947) The louse an account of the lice which infest man, their medical importance and
control. Edward-Arnold, London
Canyon DV, Spearce R, Muller R (2002) Spatial and kinetic factors for the transfer of head lice
(Pediculus capitis) between hairs. J Invest Dermatol 119:629–631
Canyon DV, Speare R (2010) Indirect transmission of head lice via inanimate objects. Open
Dermatol J 4:72–76
182 J.M. Allen et al.
Hooke R (1665) Micrographia: or some physiological description of minute bodies made by mag-
nifying glasses with observations and inquiries thereupon. Council Royal Society of London
for Improving of Natural Knowledge, London
Howlett FM (1917) Notes on head- and body-lice and upon temperature reactions of lice and mos-
quitoes. Parasitology 10:186–188
Hypsa V (2006) Parasite histories and novel phylogenetic tools: alternative approaches to inferred
parasite evolution from molecular markers. Int J Parasitol 36:141–155
Ihobe H (1990) Interspecific interactions between wild pygmy chimpanzees (Pan paniscus) and
red colobus (Colobus badius). Primates 31:109–112
Kimbel WH, Delezene LK (2009) “Lucy” redux: a review of research on Australopithecus afaren-
sis. Yearbk Phys Anthropol 52:2–48
Kimbel W, Lockwood CA, Ward CV, Leakey MG, Rak Y, Johanson DC (2006) Was Australopithecus
anamensis ancestoral to A. afarensis? A case of anagenesis in the hominin fossil record. J Hum
Evol 51:134–152
Kingdon J (1974) East African mammals: an altas of evolution in Africa. University of Chicago
Press, Chicago
Kingston JD (2007) Shifting adaptive landscapes: progress and challenges in reconstructing early
hominid environments. Yearbk Phys Anthropol 50:20–58
Kirkness EF, Haas BJ, Sun W, Braig HR, Perotti MA, Clark JM et al (2010) Genome sequences of
the human body louse and its primary endosymbiont provide insights into the permanent para-
sitic lifestyle. Proc Natl Acad Sci 107(27):12168–12173
Kittler R, Kayser M, Stoneking M (2003) Molecular evolution of Pediculus humanus and the ori-
gin of clothing. Curr Biol 13:1414–1417
Kittler R, Kayser M, Stoneking M (2004) Erratum molecular evolution of Pediculus humanus and
the origin of clothing. Curr Biol 14:2309
Kunimatsu Y, Nakatsukasa M, Sawada Y, Sakai T, Hyodo M, Hyodo H, Itaya T, Nakaya H, Saegusa
H, Mazurier A, Saneyoshi M, Tsujikawa H, Yamamoto A, Mbua E (2007) A new late Miocene
great ape from Kenya and its implications for the origins of African great apes and humans.
Proc Natl Acad Sci 104:19220–19225
Kuroda S, Nishihara T, Suzuki S, Oko RA (1996) Sympatric chimpanzees and gorillas in the Ndoki
Forest, Congo. In: McGrew WC, Marchant LF, Nishida T (eds) Great ape societies. Cambridge
University Press, Cambridge, pp 71–81
Kushlan JA (1985) The vestiary hypothesis of human hair reduction. J Hum Evol 14:29–32
Leakey MG, Spoor F, Brown FH, Gathogo PN, Kairie C, Leakey LN, McDougall I (2001) New hom-
inin genus from eastern Africa shows diverse middle Pliocene lineages. Nature 410:433–440
Leo NP, Barker SC (2005) Unravelling the origins of the head lice and body lice of humans.
Parasitol Res 98:44–47
Leo NP, Hughes JM, Yang X, Poudel SKS, Brogdon WG, Barker SC (2005) The head and body
lice of humans are genetically distinct (Insecta: Phthiraptera: Pediculidae): evidence from dou-
ble infestations. Heredity 95:34–40
Li W, Ortiz G, Fournier P-E, Gimenez G, Reed DL, Pittendrigh B et al (2010) Genotyping of
human lice suggests multiple emergences of body lice from local head louse populations. PLoS
Neglect Trop Dis 4(3):e641
Light JE, Toups MA, Reed DL (2008) What’s in a name: the taxonomic status of human head and
body lice. Mol Phylogenet Evol 47(3):1203–1216
Light JE, Smith VS, Allen JM, Durden LA, Reed DL (2010) Evolutionary history of mammalian
sucking lice (Phthiraptera: Anoplura). BMC Evol Biol 10:292
Lloyd L (1919) Lice and their menace to man. Frowde, Oxford
Luca F, Perry GH, Di Rienzo A (2010) Evolutionary adaptations to dietary changes. Annu Rev
Nutr 30:291–314
Macho GA, Shimizu D (2010) Kinematic parameters inferred from enamal microstructure: new
insights into the diet of Australopithecus anamensis. J Hum Evol 58:23–32
Mahoney SA (1980) Cost of locomotion and heat balance during rest and running from 0 to 55°C
in a patas monkey. J Appl Physiol: Respirat Environ Exercise Physiol 49:789–800
184 J.M. Allen et al.
Reed DL, Toups MA, Light JE, Allen JM, Flannigan S (2009) Lice and other parasites as markers
of primate evolutionary history. In: Huffman M, Chapman C (eds) Primate parasite ecology:
the dynamics and study of host-parasite relationships. Cambridge University Press, Cambridge,
pp 231–250
Richmond BG, Jungers WL (2008) Orrorin tugenensis femoral morphology and the evolution of
hominin bipedalism. Science 319:1662–1665
Rogers AR, Iltis D, Wooding S (2004) Genetic variation at the MC1R locus and the time since loss
of body hair. Curr Anthropol 45:105–108
Rose MD (1977) Interspecific play between free ranging guerezas (Colobus guereza) and vervet
monkeys (Cercopithecus aethiops). Primates 18:957–964
Sasaki T, Poudel SKS, Isawa H, Hayashi T, Seki N, Tomita T, Sawabe K, Kobayashi M (2006a)
First molecular evidence of Bartonella quintana in Pediculus humanus capitis (Phthiraptera:
Pediculidae, collected from Nepalese children. J Med Entomol 43(1):110–112
Sasaki T, Poudel SKS, Isawa H, Hayashi T, Seki N, Tomita T, Sawabe K, Kobayashi M (2006b)
First molecular evidence of Bartonella quintana in Pediculus humanus capitis (Phthiraptera:
Pediculidae), collected from Nepalese children. J Med Entomol 43(5):788
Schaller GB (1963) The Mountain Gorilla: ecology and behavior. University of Chicago Press,
Chicago
Schaller GB (1965a) The behavior of the Mountain Gorilla. In: DeVore I (ed) Primate behavior:
field studies of monkeys and apes. Holt, Rinehart, and Winston, New York, pp 324–367
Schaller GB (1965b) Behavioral comparisons of the apes. In: DeVore I (ed) Primate behavior: field
studies of monkeys and apes. Holt, Rinehart, and Winston, New York, pp 474–481
Schultz AH (1931) The density of hair in primates. Hum Biol 3:303–321
Schultz AH (1969) The life of primates. Universe Books, New York
Schwartz GG, Rosenblum LA (1981) Allometry of hair density and the evolution of human hair-
lessness. Am J Phys Anthropol 55:9–12
Smith RJ, Jungers WL (1997) Body mass in comparative primatology. J Hum Evol 32:523–559
Sockol MD, Raichlen DA, Pontzer H (2007) Chimpanzee locomotor energetics and the origin of
human bipedalism. Proc Natl Acad Sci 104:12265–12269
Spoor F, Leakey MG, Leakey LN (2010) Hominin diversity in the Middle Pliocene of eastern
Africa: the maxilla of KNM-WT 40000. Proc Roy Soc Lond B 365:3377–3388
Stanford CB (2006) The behavioral ecology of sympatric African apes: implications for under-
standing fossil hominoid ecology. Primates 47:91–101
Stafford (2008) Head lice: evidence-based guidelines based on the Stafford Report 2008 Update.
Public Health Med Environ Group
Stoddart DM (1990) The scented ape: the biology and culture of human odour. Cambridge
University Press, Cambridge
Strait DS, Grine FE (2004) Inferring hominoid and early hominid phylogeny using craniodental
characters: the role of fossil taxa. J Hum Evol 47:399–452
Strait DS, Grine FE, Moniz MA (1997) A reappraisal of early hominid phylogeny. J Hum Evol
32:17–82
Struhsaker TT, Butynski TM, Lwanga JS (1988) Hybridization between redtail (Cercopithecus
ascanius schmidti) and blue (C. mitis stuhlmanni) monkeys in the Kibale Forest, Uganda. In:
Gautier-Hion A, Bourlière F, Gautier JP, Kingdon J (eds) A primate radiation: evolutionary
biology of the African Guenons. Cambridge University Press, Cambridge, pp 477–497
Suwa G, Kono RT, Katoh S, Asfaw B, Beyene Y (2007) A new species of great ape from the late
Miocene epoch in Ethiopia. Nature 448:921–924
Teaford MF, Ungar PS (2000) Diet and the evolution of the earliest human ancestors. Proc Natl
Acad Sci 97:13506–13511
Tebruegge M, Pantazidou A, Curtis N (2011) What’s bugging you? An update on the treatment of
head lice infestation. Arch Dis Child Educ Pract Ed 96:2–8
Thalmann O, Fischer A, Lankester F, Pääbo S, Vigilant L (2007) The complex evolutionary history
of gorillas: Insights from genomic data. Mol Biol Evol 24:146–158
186 J.M. Allen et al.
Toups MA, Kitchen A, Light JE, Reed DL (2011) Origin of clothing lice indicates early clothing
use by anatomically modern humans in Africa. Mol Biol Evol 28(1):29–32
Ubelaker JE, Payne E, Allison VF, Moore DV (1973) Scanning electron microscopy of the human
pubic louse, Pthirus pubis (Linnaeus, 1758). J Parasitol 59(5):913–919
Ungar P (2004) Dental topography and diets of Australopithecus afarensis and early Homo. J Hum
Evol 46:605–622
Ungar PS, Scott RS, Grine FE, Teaford MF (2010) Molar microwear textures and the diets of
Australopithecus anamensis and Australopithecus afarensis. Proc Roy Soc Lond B Biol
365:3345–3354
Verhaegen M, Puech P-F, Munro S (2002) Aquarboreal ancestors? Trends Ecol Evol 17:212–217
Waldeyer L (1900) Ein Fall von Phthirius pubis im Bereiche des behaarten Kopfes. Charite-
Annalen, Berlin, XXV:494–499
Wheeler PE (1992) The influence of the loss of functional body hair on the water budgets of early
hominids. J Hum Evol 23:379–388
White T (2003) Early hominids–diversity or distortion? Science 299:1994–1997
Whiteman NK, Parker PG (2005) Using parasites to infer host population history: a new rationale
for parasite conservation. Anim Cons 8:175–181
Wilkinson RD, Steiper ME, Soligo C, Martin RD, Yang ZH, Tavare S (2011) Dating primate diver-
gences through an integrated analysis of palaeontological and molecular data. Syst Biol
60(1):16–32
WoldeGabriel G, White TD, Suwa G, Renne P, de Heinzelin J, Hart WK, Heiken G (1994)
Ecological and temporal placement of early Pliocene hominids at Aramis, Ethiopia. Nature
371:330–333
WoldeGabriel G, Ambrose SH, Barboni D, Bonnefille R, Bremond L, Currie B, DeGusta D, Hart
WK, Murray AM, Renne PR, Jolly-Saad MC, Stewart KM, White TD (2009) The geological,
isotopical, botanical, invertebrate and lower vertebrate surroundings of Ardipithecus ramidus.
Science 326(5949):65e1–65e5
Wolfe ND, Dunavan CP, Diamond J (2007) Origins of major human infectious diseases. Nature
447(17):279–283
Yamagiwa J, Basabose AK (2006) Effects of fruit scarcity on foraging strategies of sympatric
gorillas and chimpanzees. In: Hohmann G, Robbins M, Boesch C (eds) Feeding ecology in
apes and other primates. Cambridge University Press, Cambridge, pp 73–96
Yamagiwa J, Maruhashi T, Yumoto T, Mwanza N (1996) Dietary and ranging overlap in sympatric
gorillas and chimpanzees in Kahuzi-Biega National Park, Zaire. In: McGrew WC, Marchant
LF, Nishida T (eds) Great ape societies. Cambridge University Press, Cambridge, pp 82–98
Yotova V, Lefebvre JF, Moreau C, Gbeha E, Hovhannesyan K, Bourgeois S, Bédarida S, Azevedo
L, Amorim A, Sarkisian T, Avogbe P, Chabi N, Dicko MH, Amouzou ESKS, Sanni A, Roberts-
Thomson J, Boettcher B, Scott RJ, Labuda D (2011) An X-linked haplotype of Neanderthal
origin is present among all non-African populations. Mol Biol Evol 28(7):1957–1962
Part II
Emergence and Divergent Disease
Manifestation
Treponema pallidum Infection in Primates:
Clinical Manifestations, Epidemiology,
and Evolution of a Stealthy Pathogen
Introduction
J.F. Brinkworth and K. Pechenkina (eds.), Primates, Pathogens, and Evolution, 189
Developments in Primatology: Progress and Prospects,
DOI 10.1007/978-1-4614-7181-3_7, © Springer Science+Business Media New York 2013
190 K.N. Harper and S. Knauf
T. pallidum subsp. pallidum, the causative agent of syphilis, is the only treponemal
subspecies that is primarily transmitted via sex. An estimated 12 million new cases
of syphilis occur every year and can be found throughout the world (WHO 2001).
Certain socio-ecological risk factors (e.g., multiple sex partners and immunosup-
pression) are associated with a greater risk of contracting and transmitting the dis-
ease (Karp et al. 2009; Rolfs et al. 1990; Yahya-Malima et al. 2008). Yaws, caused
by subsp. pertenue, is a nonvenereal disease that is usually acquired during early
childhood. Transmission appears to occur primarily via skin-to-skin contact (Perine
et al. 1984), although flies may also serve as vectors (Cousins 1972; Kumm and
Turner 1936; Satchell and Harrison 1953). Yaws was once common in tropical
regions throughout the world. Although it has never been eradicated, it grew
increasingly rare after a WHO-sponsored eradication campaign mid-century (Arya
and Bennett 1976; Guthe et al. 1953, 1972). Surveillance is typically poor in the
areas most likely to be affected; accurate figures on the infection’s current preva-
lence are not available. However, recent foci of infection in the Republic of Congo
and the Central African Republic (Salomone 1999), the Democratic Republic of
Congo (Gersti et al. 2009), Papua New Guinea (Mitjà et al. 2011), East Timor
(Satter and Tokarz 2010), and Vanuatu (Fegan et al. 2010) have been documented.
The disease appears to have been eradicated in India (Lahariya and Pradhan 2007)
and has not been reported recently in South America either. Finally, bejel, caused
by subsp. endemicum, was once common in arid regions such as the Middle East
and the Balkans, but it has not been reported in the literature since the 1990s
(Yakinci et al. 1995). Like yaws, it is typically acquired during childhood and is
thought to be transmitted primarily via fomites such as utensils and drinking vessels
(Perine et al. 1984).
All three diseases are chronic, multistage infections in humans that are easily
treated by antibiotics in most cases, but they can be deadly if neglected (Table 1). It
should be noted that increasing macrolide resistance has begun to complicate treat-
ment worldwide, and antibiotic-resistant T. pallidum could one day represent a sub-
stantial health challenge (Stamm 2010). Syphilis typically begins with a hard,
painless chancre at the site where the bacterium entered the body. Untreated infec-
tion leads to a secondary stage, often characterized by a rash that appears weeks
later. The rash associated with secondary syphilis may take several forms, but it
generally affects the palms and soles and does not itch (Richens and Mabey 2009).
Mild fever, patchy hair loss, and weight loss are also common clinical signs in
humans. If left untreated, tertiary-stage disease may occur, though sometimes not
until decades after the infection was initially contracted. Destructive lesions called
gummata may appear in virtually any organ of the body; neurosyphilis may result in
psychiatric disorders, while cardiovascular syphilis killed many people in the pre-
antibiotic era (Holmes et al. 2007).
The disease progression is similar for infection by all three of the T. pallidum
subspecies, although a number of important differences have been reported (Table 1).
Table 1 Characteristics of natural infection with T. pallidum in different primate species
Humans Nonhuman primatesa
Baboons-West
Syphilis Yaws Bejel Africa Baboons-Tanzania Gorillas
Genetic characteris- T. pallidum subsp. pallidum T. pallidum subsp. pallidum T. pallidum subsp. Genetic Genetic sequences Unknown
tics of strain pallidum sequences from strains at
responsible from one Lake Manyara
strain National Park
collected in show most
Guinea similarity to
show most subsp.
similarity pertenue
to subsp.
pertenue
Primary-stage Hard, painless chancre Papillomatous “mother yaw” Lesions at site where When present Moderate to Scabby, dry raised
lesions develops at site where develops at site where bacterium enters the at all: mild, severe genital lesions
bacterium entered the bacterium entered the body are rare. When small, ulceration; primarily
body (usually anogeni- body (most often legs) present, they resemble keratotic enlargement of affecting lips,
tal region) chancres lesions and inguinal lymph nose, eyes, and
Secondary-stage Non-itchy rash affecting Crustopapillomatous skin Angular stomatitis, mucous ulcers nodes cheeks. Large
lesions palms and soles of feet; lesions, polydactylitis, patches in mouth, rashes affecting destructive
alopecia, mild fever, osteoperiostitis, nodules, with hypertrophic muzzle, lesions
weight loss, enlarge- plaques, and papules on lesions, pigmentary eyelids, primarily
ment of lymph nodes the skin changes, osteoperiostitis armpits affecting nose
Tertiary-stage Gummata can develop in Hyperkeratosis, ulceration Severe osteoperiostitis, None reported and mouth
lesions virtually any organ of around the nasal and sometimes resulting in areas; lesions
the body, neurological maxillary areas (gan- saber tibia, gummata of also found on
and cardiovascular gosa), saber tibia, gondou the skin, palate, and wrists and
involvement (hypertrophic osteitis of nasal septum, gangosa ankles
nasal process of maxilla)
(continued)
Table 1 (continued)
Humans Nonhuman primatesa
Baboons-West
Syphilis Yaws Bejel Africa Baboons-Tanzania Gorillas
Primary transmis- Sexual Skin to skin Thought to be mouth to Unknown Lesions targeting Unknown
sion mode mouth. Fomites such the genitals
as utensils and and involve-
drinking vessels may ment of only
be important sexually active
animals
suggests
sexual
transmission
Typical age of Sexually mature Childhood Childhood; 66 % acquired Unknown Sexually mature Lesions have been
infection individuals before the age of 16 in animals reported in
one study of 3,507 cases almost 5 % of
in Iraq infants in the
Parc National
d’Odzala-
Kokoua
(Republic of
Congo); peak
incidence
between infant
and juvenile
stages
Congenital Yes Rarely, if ever Rarely, if ever Unknown Unknown Unknown
transmission
Geographic range Global Once very common in hot, Once very common in hot, West Africa: Tanzania West Africa:
humid regions of the arid regions such as the Guinea, Republic of
world. Still reported in Middle East and the Senegal, Congo,
some countries in Eastern Mediterranean. Cameroon Cameroon,
Western Africa and the Last reported case in Democratic
South Pacific Turkey, during the 1990s Republic of
Congo
Sources Perine et al. (1984), Salazar Noordhoek et al. (1991), Yakinci et al. (1995), Pace Baylet et al. Wallis and Lee Cousins (1984),
et al. (2002a, b) Perine et al. (1984) and Csonka (1984), (1971), (1999), Karesh (2000),
Csonka (1953), Perine Fribourg- Mlengeya Levréro et al.
et al. (1984) Blanc and (2004), Knauf (2007)
Mollaret et al. (2012)
(1969),
Harper
et al.
(2008),
Smajs et al.
(2011)
a
Infection in NHPs is not well characterized enough to determine whether there are primary, secondary, and tertiary-stage lesions. Therefore, all lesions are described in one
category
194 K.N. Harper and S. Knauf
First, only syphilis is regularly transmitted congenitally. While there are several
reports of possible yaws infection via in utero transmission in the literature
(Engelhardt 1959; Wilson and Mathis 1930), it appears likely that congenital infec-
tion does not occur in the nonvenereal treponematoses (Antal et al. 2002). It has
been hypothesized that this may be because yaws and bejel are typically acquired
early in life, so active infection at sexual maturity is virtually nonexistent (Willcox
1955). It is also commonly believed that only subsp. pallidum, the agent of syphilis,
affects the central nervous system (CNS). However, in-depth study of the progres-
sion of subsp. pertenue and endemicum infections has not been carried out, and one
of the rare studies focusing on CNS involvement in yaws infection reported the pres-
ence of spirochetes in ocular fluid (Smith et al. 1971). In addition, Román and
Román (1986) draw attention to a possible association between CNS complications
and yaws infection. Finally, although for the most part syphilis transmission occurs
via a venereal route and the other treponemal diseases are nonsexually transmitted,
the potential for transmission via alternate routes is known for all T. pallidum infec-
tions. Cases of extragenital syphilis resulting from nonsexual contact have been
reported several times (Luger 1972; Taylor 1954), for example, by human bites (Oh
et al. 2008) or by mouth-to-mouth feeding in infants (Zhou et al. 2009). Similarly,
infectious yaws and bejel lesions have been reported on the genitalia (Turner and
Hollander 1957b; Wilson and Mathis 1930), suggesting that the opportunity for
sexual transmission might be present to some extent in all of the
T. pallidum subspecies. In summary, because so few in-depth studies have been
performed on the nonsexually transmitted T. pallidum subspecies, the exact nature
of the similarities and differences between the three diseases remains ambiguous.
Only recently have we gained the tools to differentiate between the three subspe-
cies. For decades, a diagnosis of syphilis, yaws, or bejel was given based on the clini-
cal presentation of the disease, the age of the patient, and the area of the world in
which they lived. In areas where multiple treponemal diseases were present, the dif-
ferential diagnosis could be quite puzzling (Lagarde et al. 2003). In the 1990s, single-
nucleotide polymorphisms that reliably differentiated subsp. pallidum strains from
the two nonsexually transmitted subspecies were identified for the first time
(Centurion-Lara et al. 1998; Noordhoek et al. 1990). Subsequently, patterns of genetic
variation that could be used to separate all three subspecies were identified (Centurion-
Lara et al. 2006; Harper et al. 2008). Thus, while we still have no serological test that
can differentiate between the three diseases, we can now use genetics to do so.
The multiple subspecies of T. pallidum present in humans raise the question:
which selective pressures led to the emergence of multiple transmission modes? In
the past, researchers have speculated that subsp. pertenue may represent the ances-
tral, nonsexually transmitted pathogen, perhaps infecting the earliest hominids
(Cockburn 1967; Harper et al. 2008). In this view, subsp. pallidum is a relatively
recent pathogen. It could have emerged in the Old World, as large cities began to
appear and novel sexual behavior fostered sexual transmission (Hudson 1963). Or,
according to a more conventionally held theory supported by recent genetic evi-
dence (Harper et al. 2008), it may have arisen in the New World during Pre-
Columbian times, introduced into the Old World by Columbus and his men during
Treponema pallidum Infection in Primates: Clinical Manifestations, Epidemiology… 195
the voyages of discovery (Crosby 1969). Some researchers believe that the skeletal
evidence indicates that subsp. pallidum may not have emerged until the voyages of
discovery themselves, when Columbus and his crew took a nonvenereal form of
T. pallidum back to the Old World, where it rapidly gained the ability to be sexually
transmitted in its new environment (Baker and Armelagos 1988). Alternatively, one
team of researchers has posited that all three T. pallidum subspecies are relatively
recent additions to the family of human pathogens and arose at similar times in
human history (Gray et al. 2006). Luckily, novel data have recently allowed us to
reject yet another theory. The “Unitarian hypothesis” states that all human trepone-
mal infections are caused by a single, protean pathogen with an opportunistic trans-
mission mode, resulting in dramatically different clinical manifestations depending
upon factors such as climate (Hudson 1963). Given the morphological and overall
genetic similarity between the T. pallidum subspecies, this view did not seem unrea-
sonable in the past. Recent genetic evidence stemming from intensive sequencing
efforts by multiple groups has convincingly shown, however, that the three trepone-
mal diseases, syphilis, yaws, and bejel, are caused by genetically distinct pathogens
(Centurion-Lara et al. 2006; Harper et al. 2008; Smajs et al. 2011). Since an essen-
tial tenet of the Unitarian hypothesis is that the strains that cause the three diseases
are interchangeable and thus cannot be distinguished, we can now discard this evo-
lutionary explanation for T. pallidum’s multiple transmission modes. As described
above, however, many competing hypotheses remain.
In the remainder of this chapter, we will investigate what is known about T. pal-
lidum in NHPs and will conclude by discussing what insights NHP infections may
provide in understanding the history of human infection.
Thanks to observational and serological surveys of wild NHPs, we now know that
Treponema infection occurs naturally in many species. Based on the available
sequences from the Fribourg-Blanc strain, collected from a wild baboon captured in
Guinea in the 1960s (Fribourg-Blanc et al. 1966), we also know that the NHP strains
responsible are very closely related to human T. pallidum strains. This baboon strain
harbors 16S ribosomal subunit DNA that is 100 % identical to that found in T. pal-
lidum human strains,1 a level of identity which exceeds the threshold of 97–99 %
often used to distinguish between bacterial species (Gevers et al. 2005). However,
we do not yet know whether NHP strains represent a sister subspecies to pertenue
or whether human subsp. pertenue and NHP strains represent one large clade.
1
Sequences from the 16S ribosomal subunit (2nd operon) available from GenBank were obtained
and aligned in ClustalX. They included the subsp. pallidum strains Dallas-1 (NC016844.1),
Chicago (CP001752.1), Nichols (NC000919.1), Street Strain 14 (NC010741.1), and Mexico A
(HM585252.1), the subsp. pertenue strains CDC-2 (NC016848.1), Gauthier (NC016843.1),
Samoa D (NC016842.1), and the Fribourg-Blanc strain (HM165231.1).
196 K.N. Harper and S. Knauf
African Monkeys
NHP treponemal infection has probably been studied best in wild baboons (Papio
spec.). Beginning in the 1960s, serological surveys demonstrated that treponemal dis-
ease was present at high prevalence in the baboons of Equatorial Guinea, Senegal, and
Cameroon (Table 2) (Baylet et al. 1971; Fribourg-Blanc and Mollaret 1969). As
described previously, one T. pallidum strain, the Fribourg-Blanc strain, was isolated
from an infected baboon during the course of these surveys (Fribourg-Blanc et al. 1966).
Table 2 T. pallidum infection in African monkeys
Seroprevalence in
Species Origin sample Test Source
Mangabey (Cercocebus, Laboratory animal caught in West Africa 0/1 (0.0 %) FTA-ABS Felsenfeld and Wolf (1971)
species not determined) Bangui region, Central African Republic 0/2 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Green monkey Laboratory animals caught in West Africa 0/3 (0.0 %) FTA-ABS Felsenfeld and Wolf (1971)
(Chlorocebus sabaeus)
Vervet monkey Laboratory animals caught in Kenya 1/2 (50.0 %), + result FTA-ABS Felsenfeld and Wolf (1971)
(Chlorocebus equivocal
pygerythrus)
Chlorocebus, species not Casamance region of Senegal 3/8 (37.5 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
determined Fouta Toro Fleuve region of Senegal 28/45 (73.0 %) FTA-ABS Baylet et al. (1971)
Niamey, Niger 0/1 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Bangui, Central African Republic 0/3 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Brazzaville region, Democratic Republic 0/3 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
of Congo
Laboratory animals caught in East Africa 2/7 (28.6 %), + FTA-ABS Felsenfeld and Wolf (1971)
results equivocal
Colobus monkey (Colobus, Casamance region of Senegal 1/1 (100.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
species not determined)
Patas monkey (Erythrocebus Laboratory animals caught in West Africa 1/13 (7.7 %) FTA-ABS Felsenfeld and Wolf (1971)
patas) Casamance region of Senegal 6/26 (23.1 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Mali 0/18 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Bobo-Dioulasso region of Burkina Faso 0/17 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Niamey, Niger 0/1 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Fort-Lamy, Chad 1/14 (7.1 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Laboratory animals caught in East Africa 1/10 (10.0 %), + FTA-ABS Felsenfeld and Wolf (1971)
Treponema pallidum Infection in Primates: Clinical Manifestations, Epidemiology…
result equivocal
Fouta Toro Fleuve region, Senegal 1/4 (25.0 %) FTA-ABS Baylet et al. (1971)
Koussanar region of Senegal 1/38 (2.6 %) FTA-ABS Baylet et al. (1971)
197
(continued)
Table 2 (continued)
198
Seroprevalence in
Species Origin sample Test Source
Yellow baboon (Papio Kindia region of Guinea 164/216 (75.9 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
cynocephalus) Casamance region of Senegal 6/10 (60.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Kaolack region of Senegal 80/171 (46.8 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Bobo-Dioulasso region of Burkina Faso 0/159 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Yaounde, Cameroon 3/3 (100.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Kenya 0/276 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
Olive baboon (Papio Casamance region of Senegal 49/82 (59.8 %) FTA-ABS Baylet et al. (1971)
anubis) Koussanar region of Senegal 15/55 (27.3 %) FTA-ABS Baylet et al. (1971)
Falémé region of Senegal 0/111 (0.0 %) FTA-ABS Baylet et al. (1971)
Lake Manyara National Park, Tanzania 43/57 (75.4 %) Serodia TP*PA Knauf et al. (2012)
Baboon (Papio, species not Brazzaville region, Democratic Republic 0/2 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
determined) of Congo
K.N. Harper and S. Knauf
Treponema pallidum Infection in Primates: Clinical Manifestations, Epidemiology… 199
Experiments have shown that this strain is capable of causing infection in humans
(Smith et al. 1971), and genetically it is closely related to, but possibly distinct from,
human subsp. pertenue strains (Centurion-Lara et al. 2006; Harper et al. 2008; Smajs
et al. 2011). Clinical signs in the affected populations were described as mild and
included small keratotic lesions and ulcers on the muzzle, eyelids, and armpits (Baylet
et al. 1971), though most infected animals did not appear to display any lesions at all
(Table 1). It is interesting that in these same surveys, not a single animal from countries
farther east, such as Burkina Faso (n = 159) or Kenya (n = 276), was found to be positive
(Fribourg-Blanc and Mollaret 1969).
In the late 1980s, however, a form of treponemal disease with strikingly different
clinical signs was described among the olive baboons (P. anubis) at Gombe Stream
National Park (Table 1). “Penelope,” an adult female baboon at Gombe, was the first
recorded case displaying what would come to be recognized as the genital-associated
lesions typical of the infection (Collins et al. 2011). Subsequently, researchers
noticed the skin disease in several baboon troops in the national park (Wallis 2000b;
Wallis and Lee 1999). At that time, laboratory analysis (dark field illumination) of
skin lesions confirmed the presence of T. pallidum. However, DNA-based verifica-
tion of the subspecies of T. pallidum involved was not performed, nor had other
pathogens that cause genital ulceration in humans and NHPs been ruled out defini-
tively. Because of the infection’s association with genital lesions and the observa-
tion that it appeared in sexually mature animals, it was hypothesized that the disease
might be sexually transmitted (Wallis and Lee 1999). Moreover, unlike the mild
lesions described in West Africa, it was reported that lesions in a small proportion
of the individuals affected at Gombe became so severe that urinary flow was
obstructed and death resulted, with autopsy of one young male revealing wide-
spread sepsis within the urogenital tract (Wallis and Lee 1999).
In the 1990s, similar lesions (Fig. 1) were reported for the first time in olive
baboons at Lake Manyara National Park, also in Tanzania but 700 km away from
Gombe (Mlengeya 2004). In contrast to other studies examining treponemal infec-
tion in NHPs, Knauf et al. (2012) were able to describe in detail the macroscopic
clinical manifestations and histological findings that characterize T. pallidum infec-
tion in baboons. In addition, for the first time, they were able to demonstrate simian
T. pallidum strains in situ, using immunohistochemistry (Fig. 2). Molecular biologi-
cal tests such as qualitative and quantitative PCR were performed using skin tissue
samples, and they demonstrated the presence of T. pallidum in infected animals
while ruling out other pathogens. Qualitative and quantitative serological tests,
including the Serodia TP*PA test, offered still more proof that T. pallidum infection
was responsible for the lesions observed. Four informative genetic polymorphisms
were used to demonstrate that, despite their predilection for causing anogenital
lesions similar to those found in human syphilis, the simian strains collected from
baboons at Lake Manyara National Park were most closely related to nonvenereal
human T. pallidum subsp. pertenue strains. Finally, the study showed that many
animals that looked healthy in the field had positive PCR, serological, or histologi-
cal results for T. pallidum. For example, of 20 baboons with no clinical signs of
infection, 15 showed histological abnormalities, of which six tested PCR positive
200 K.N. Harper and S. Knauf
Fig. 1 Photos depicting clinical signs associated with genitotropic T. pallidum infection in
baboons. The genitals of a severely affected female (left) and male (right) are shown
Fig. 2 Photos depicting T. pallidum in a skin tissue sample from the genitals of an Olive baboon
at Lake Manyara National Park, Tanzania. Immunohistochemistry utilizing rabbit polyclonal
antibodies against T. pallidum, was performed with epithelial cells counterstained using Mayer’s
hematoxylin. Bar 10 μm
for T. pallidum. Seven of the 20 clinically unaffected animals had anti-T. pallidum
antibody titers ≥1:80 (Knauf et al. 2012). This indicates that the prevalence of the
disease is much higher than originally suspected from field observations (Mlengeya
2004).
That this infection is prevalent in the baboons of Tanzania is also reinforced by a
small survey we performed of baboons imported into the United States by labora-
tory suppliers (Harper, data not published). In 2006, we assayed 17 serum samples
from olive baboons (Papio anubis) imported from Tanzania to the United States
using the Sero-DIA TPPA test (Fujirebio Diagnostics, Malvern, PA), a highly
Treponema pallidum Infection in Primates: Clinical Manifestations, Epidemiology… 201
sensitive and specific serological treponemal test. Sixteen samples came from adult
males imported in 2003 (n = 6) and 2005 (n = 10). Pooled serum from six adult
females imported in 2001 was also tested. Four of the 16 samples from adult males
tested positive for T. pallidum antibodies, as did the pooled serum from the adult
females. It is not clear where in Tanzania these animals were captured, but the
relatively high prevalence of infection in these samples suggests that the infection
has become established outside of Gombe and Lake Manyara National Parks.
Serological evidence of T. pallidum infection has also been found in other African
monkey species (Table 2), though no gross-pathological lesions that appear to result
from infection have yet been described. Antibodies have been detected in the
Chlorocebus species [vervet (C. pygerythrus) and green monkeys (C. sabaeus)],
patas monkeys (Erythrocebus patas), and in one Colobus monkey (Colobus spec.).
Gorillas
Reports that yaws infection is frequent in wild gorillas (Gorilla gorilla) in West
Africa have appeared for some time. For example, stories from indigenous people
as well as reports from the Service de Chasse in the former French Equatorial Africa
region told of gorillas in the Ewo, Kelle, and Mekambo regions that suffered from a
leprosy-like disease (Cousins 1984). In the 1950s, two researchers got the opportu-
nity to examine four young gorillas captured in this area that displayed the “lep-
rous” lesions described in earlier reports. The raised lesions, scabby and dry,
primarily affected the lips, nose, eyes, and cheeks, with one animal also exhibiting
lesions on the shoulder and forearm (Table 1). It should be noted that leprosy has
been confirmed in an Asian wild-born macaque (Valverde et al. 1998), two African
wild-born sooty mangabeys (Gormus et al. 1991), and wild-born chimpanzees from
West Africa (Hubbard et al. 1991; Leininger et al. 1978; Suzuki et al. 2010). It is not
yet clear whether these laboratory animals were infected while living in the wild or
via contact with infected handlers while being captured and cared for prior to export
in leprosy-endemic areas. However, in the case of the wild gorillas with “leprous”
lesions, the researchers reported that the infection was successfully cleared up in 8
days with penicillin injections, which was consistent with T. pallidum infection
rather than leprosy (Cousins 1984).
Mid-century, the scientist Armand Denis shared his observations on yaws in the
wild gorillas of the Republic of Congo (ROC). He described the first case he saw,
affecting an adult male, thus:
It was like a mask eaten into by some flesh-consuming disease. The lips were gone. The
nostrils were eaten almost away and the fangs of teeth were blackened and askew in what
remained of the creature’s lower jaw. Only the eyes were untouched… Whatever the disease
was the wretched animal had caught I had no idea… (Denis 1963, p 181).
Although the clinical signs described again appear to overlap with those caused
by leprosy, Denis learned that this disease was thought by the natives to be yaws, the
same disease that was common in human populations that lived along the coast.
202 K.N. Harper and S. Knauf
Reports of yaws were not limited to gorillas living in the ROC. Lesions consistent
with yaws were also described in the young gorillas of Rio Muni, in Equatorial Guinea
(Cousins 1984). Similarly, yaws was diagnosed in two out of five young gorillas
imported from the French Cameroons (modern day Cameroon and Nigeria) to the
Lincoln Park Zoo, mid-century (Cousins 1972), though the method of T. pallidum
confirmation is not given in the report. Sensitive and specific serological tests have
confirmed that gorillas do indeed come into contact with T. pallidum (Table 3). Four
blood samples from solitary gorillas in the Parc National d’Odzala-Kokoua in the
Republic of Congo, two of which had skin lesions consistent with yaws, tested posi-
tive for treponemal antibodies (Karesh 2000).
Skeletal evidence consistent with (but not specific to) treponemal infection has
also been found in the remains of wild gorillas (Lovell et al. 2000). Lovell and col-
leagues examined 126 gorilla skeletons collected from the Democratic Republic of
Congo (DRC), the ROC, and Cameroon in the early twentieth century. Eighteen
percent of gorillas were found to exhibit possible signs of treponemal disease, with
active lesions found almost entirely among subadults. In addition, cranial deformi-
ties involving massive osseous tumors have been documented in gorillas, and some
researchers have speculated that these may correspond to goundou, a rare manifes-
tation of tertiary-stage yaws in humans (Cousins 2008). However, whereas in
humans the lesions grow out of the nose and upper jaw, in gorillas, growth of
goundou-like lesions seems to occur primarily over the cheekbones (Cousins 2008).
Recently, an evaluation of skin lesions in 377 gorillas living in the Parc National
d’Odzala-Kokoua in the ROC gave us a much better understanding of treponemal
disease in this population (Levréro et al. 2007). The macroscopic lesions identified
were found to be similar to yaws in humans (Fig. 3) and affected 17 % of animals
examined. However, since some animals that are infected do not display lesions
(Karesh 2000), the actual prevalence of the infection is likely to be higher. Infection
Treponema pallidum Infection in Primates: Clinical Manifestations, Epidemiology… 203
Fig. 3 Photos depicting skin lesions suspected to result from T. pallidum infection in gorillas.
Skin lesions are shown on (a) an adult male, (b) a juvenile female, and (c) an adult female gorilla
(Source: Levréro et al. 2007)
in the gorillas studied began early, with almost 5 % of infants displaying skin lesions
compatible with yaws and peak incidence occurring between the infant and juvenile
stages. Some animals were found to exhibit destruction of the nose and/or lips or to
display deep lesions on their wrists or ankles. Possible social consequences of
severe infection were hinted at in the observations of Levréro et al. One adult female
with a nose that was completely destroyed was forced to leave her group; during the
animal’s seven subsequent visits to the clearing before disappearing, the researchers
observed the males from groups she approached behaving antagonistically towards
her. As the researchers note, the consequences of such rejection must certainly have
implications for survival. Although lesions were also reported in gorillas at nearby
sites, including Maya, Moba, and Lossi, infection may not be universal, since no
such lesions were observed at the Mbeli clearing, also nearby.
Treponemal infection in gorillas has been identified with yaws, due to its gross-
pathological manifestations and also the fact that the areas in which the gorillas
reside tend to be places in which the human prevalence of this disease was quite high
historically, affecting the vast majority of some human groups inhabiting the rain
forest (Hackett 1953; Pampiglione and Wilkinson 1975) and still infecting more
than 10 % of residents in some areas (Gersti et al. 2009). Unfortunately, not a single
report of treponemal infection has been confirmed via molecular biological tests,
such as T. pallidum-specific PCR or immunohistochemistry, capable of demonstrating
the spirochetes in situ. Obtaining the samples needed to perform such tests in endan-
gered great apes is extremely challenging, both politically and technically, and this
explains why they have not yet been performed. However, without the information
these tests can provide, it is impossible to determine with certainty that T. pallidum
is responsible for the lesions and, if so, to which human subspecies the responsible
strains are most closely related.
Although most reports of treponemal infection in wild gorillas have described a
yaws-like disease, descriptions of animals with diseased sexual organs that may be
linked to treponemal infection have also surfaced. For example, in Cameroon, one
adult female exhibited genitalia covered with running sores, as well as deforming
sores on the hands and wrists (Cousins 1984). This animal was said by the nearby
residents to be suffering from “marjal” or “mebata,” a local word for yaws. As with
204 K.N. Harper and S. Knauf
the yaws-like cases, providing an explanation for the etiology of such lesions is
challenging, especially in the context of other sexually transmitted infections that
can cause genital ulceration in NHPs. Definitively assigning an etiological agent in
such cases will only be possible with advanced molecular biological techniques.
Chimpanzees
Asian Monkeys
Some of the first yaws experiments were performed in macaques (Macaca spec.).
However, it appears that natural infection of wild macaques is very rare, if it occurs
at all; over 1,000 animals have been tested without one robust seropositive reaction
(Table 4). Similarly, though Thivolet et al. found treponemal antibodies in 46 of 415
African monkeys, not a single one of the 152 Asian monkeys this research group
tested were positive (Kuhn 1970).
Only one serologically reactive animal captured in Asia has been reported. Kuhn
(1970) describes finding a seropositive reaction in a Celebes crested macaque
(Macaca nigra), from Sulawesi, Indonesia. Given the paucity of positive infections
among Asian monkeys, however, one must wonder whether this result represents
one of the rare false positives that can stem from the FTA-ABS test or an infection
acquired during the animal’s time in captivity vs. during its life in the wild.
Additional sampling of Asian NHPs will help answer this question.
Table 4 T. pallidum infection in Asian and South American monkeys
Species Origin Seroprevalence in sample Tests Source
Asia
Macaques (Macaca, species Laboratory animals caught in 1/6 (16.7 %), + results equivocal FTA-ABS Felsenfeld and Wolf (1971)
not determined) Southeast Asia
Crab-eating Macaque Phnom-Penh region of Cambodia 0/1236 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
(Macaca fascicularis)
Crab-eating Macaque Asia 0/5 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
(Macaca fascicularis)
Rhesus Macaque (Macaca Laboratory animals caught in 0/22 (0.0 %) FTA-ABS Felsenfeld and Wolf (1971)
mulatta) Southeast Asia
Rhesus Macaque (Macaca India 0/5 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
mulatta)
Rhesus Macaque (Macaca Asia 0/30 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
mulatta)
Pig-tailed Macaque (Macaca, Asia 0/5 (0.0 %) TPI and FTA-ABS Fribourg-Blanc and Mollaret (1969)
species not determined)
South America
Owl monkey (Aotus, species Laboratory animals, origin not 2/84 (2.4 %), + results equivocal FTA-ABS Levine et al. (1970)
not determined) specified 0/9 TPI
Owl monkey (Aotus Laboratory animals caught in South 0/3 (0.0 %) FTA-ABS Felsenfeld and Wolf (1971)
trivirgatus) America
Red-bellied titi (Callicebus Laboratory animals caught in South 5/25 (20.0 %), + results equivocal FTA-ABS Felsenfeld and Wolf (1971)
moloch) America
Squirrel Monkey (Saimiri Laboratory animals caught in South 0/18 (0.0 %) FTA-ABS Felsenfeld and Wolf (1971)
sciureus) America
Squirrel monkey (Saimiri Laboratory animals, origin not 4/63 (6.3 %), + results equivocal FTA-ABS Levine et al. (1970)
sciureus) specified 0/10 TPI
Marmoset (Callithrix spec.) Laboratory animals, origin not 0/24 (0.0 %) FTA-ABS Levine et al. (1970)
specified 0/23 TPI
206 K.N. Harper and S. Knauf
To date, not a single wild South American NHP has been found with a definite sero-
positive reaction to T. pallidum, although artificial infection of owl (Aotus) and
squirrel (Saimiri) monkeys is possible (Elsas et al. 1968; Smith 1969). However, as
is clear from Tables 2 and 4, sampling in African monkeys has been much more
thorough. In order to conclude definitively that natural infection of South American
monkeys does not occur, more extensive sampling would have to be performed.
Sampling areas of Central and South America the prevalence of treponemal disease
in human indigenous groups was once very high, such as French Guiana, Brazil,
Venezuela, and Colombia (Black 1975; Hopkins and Flórez 1977; St John 1985),
would be especially important.
NHPs played a pivotal role in the discovery of T. pallidum. In the early years of the
twentieth century, Metchnikoff and Roux (1903, 1904, 1905) demonstrated that
syphilis could be transferred from humans to chimpanzees and from one chimpan-
zee to another. Thus, apes became the first reliable animal model for the study of T.
pallidum infection. Around the same time, Castellani (1907) demonstrated that
monkeys could be infected with T. pallidum subsp. pertenue. Even so, our under-
standing of the differences in how infection progresses in humans versus NHPs is
rudimentary. Here, we review what is known about NHP responses to experimental
infection with T. pallidum, focusing on between-species similarities and differences
(Table 5).
Turner and Hollander (1957b) did a series of experiments on rhesus macaques
and African green monkeys. After inoculating animals in the thighs, eyebrows, and
genital regions, the animals were followed for 14–17 months before postmortem
examination; no gross changes suggesting syphilitic infection were identified, and a
number of the animals remained seronegative throughout, though their organs con-
tained infective spirochetes. On this basis, Turner and Hollander suggested that
NHPs did not provide a suitable animal model for understanding T. pallidum patho-
genesis. Similarly, although humans typically develop a primary chancre or “mother
yaw” at the site where the bacterium enters the body, in a study utilizing ten
macaques, Sepetjian et al. (1972) reported that only two animals developed signifi-
cant lesions at the site of inoculation—the others exhibited no lesions at all or mild
macular discoloration—and these lesions disappeared quickly. Like Turner and
Hollander, Sepetjian et al. (1969) also observed no visceral lesions upon necroscopy
after 4 months, although in half the animals treponemes were present in various
organs.
Table 5 Comparison of NHP host response to experimental T. pallidum subsp. pallidum infection and human response to natural infection with syphilis
Humans Macaques African Green Monkeys Owl Monkeys (Aötus trivirgatus)
Geographic distribution Global Asia Africa South America
of host
Lesion at site of Yes Small, mild, and quick to Only in half the animals studied; No
inoculation disappear when present at small and disappeared within
all after subsp. pallidum several weeks
infectiona
Treponemes disseminate Yes Yes Yes Yes
through body
Lesions remote from Yes None observed None observed Yes
inoculation site
Serologic response to Yes Sometimes but not always; Appears rare (1 out of 4 animals Sometimes but not always
infection response is often slow to seroconverted during study
develop period)
Robust CD4+ T-cell Yes Yes Unknown Unknown
response
Sources Salazar et al. (2002a, Sepetjian et al. (1969, 1972), Turner and Hollander (1957b) Elsas et al. (1968), Smith et al. (1965)
b), Perine et al. Turner and Hollander
(1984) (1957b), Marra et al.
(1998)
a
Significant lesions at the site of inoculation were observed when the animals were injected with subsp. pertenue and Fribourg-Blanc strain (Sepetjian et al.
1969)
Treponema pallidum Infection in Primates: Clinical Manifestations, Epidemiology…
207
208 K.N. Harper and S. Knauf
Other studies suggest that the NHP response to infection may offer parallels to
that observed in humans, however. For example, experiments have demonstrated
that owl monkeys and macaques develop a chronic, systemic infection in response
to inoculation with strains of human-derived T. pallidum (Elsas et al. 1968; Marra
et al. 1998). In some of these animals, a long-lasting serological response develops
(Sepetjian et al. 1972), though, as Turner and Hollander and others have described,
others remain seronegative for months on end, despite harboring infective trepo-
nemes (Smith et al. 1965; Turner and Hollander 1957b; Wells and Smith 1967).
Similar to humans, NHPs can develop T. pallidum-related lesions weeks or months
after infection, distant from the site of inoculation (Elsas et al. 1968; Sepetjian et al.
1969; Smith et al. 1965). Clark and Yobs (1968) have opined that the variation in
terms of lesion development and serological response in NHP hosts such as the owl
monkey might be viewed as typical of the varied response found in humans; in their
view, the lesions observed are even consistent with primary and secondary stages of
disease, as in humans.
Immunological studies suggest further similarities between the human and NHP
response to T. pallidum infection. In humans, humoral, antibody, and CD8+ cyto-
toxic T-cell responses have been shown to be relatively ineffective at clearing syphi-
litic infection or curbing the progression of lesions (Carlson et al. 2011). Instead,
delayed-type hypersensitivity, which is mediated by CD4+ T cells, appears to play a
role of particular importance (Carlson et al. 2011). Similar findings have been
reported in macaques infected with a subsp. pallidum strain. As in humans, CD4+
T cells appear to be responsible for clearing T. pallidum from the central nervous
system during early infection (Marra et al. 1998). In terms of the components of the
bacterium that stimulate an immune response, T. pallidum lacks lipopolysaccha-
rides (LPS) in its cell wall and therefore does not cause an inflammation cascade by
activating the toll-like receptor (TLR) 4 (Schroder et al. 2008). TLR4 activation is a
common pathway induced by the LPS of Gram-negative bacteria and to a certain
extent by the lipoteichoic acids of Gram-positive bacteria. This mechanism repre-
sents a major immunological host defense against bacterial infection, which is virtu-
ally absent in T. pallidum infection. Instead, the T. pallidum infected host responds
to infection via the activation of TLR 2/1 by immunostimulatory lipoproteins in the
outer membrane of the spirochete (Lien et al. 1999; Schroder et al. 2008). There is
evidence that natural simian and human infections are similar in this respect. For
example, in our laboratory, we performed immunoblots upon sera from wild baboons
in Lake Manyara National Park infected with T. pallidum. The sera were tested for anti-
T. pallidum IgM and IgG antibodies to recombinant T. pallidum proteins Tp15, Tp17,
and Tp47, and Treponema membrane protein A (TmpA) obtained from human-infect-
ing T. pallidum strains. The results consistently showed that the animals were producing
antibodies against these proteins (Knauf unpublished data). On this basis, we predict
that simian T. pallidum strains express immunostimulatory lipoproteins similar to the
Tp15, Tp17, Tp47, and TmpA proteins of known human strains and that NHPs are
responding to these antigens in a similar way to humans.
In humans, the relative distribution of the four immunoglobulin G subclasses
changes over the course of infection (Baughn et al. 1988; Moskophidis 1989;
Treponema pallidum Infection in Primates: Clinical Manifestations, Epidemiology… 209
Salazar et al. 2002a, b). Whether similar changes occur in NHPs is not known, but
field research on animals displaying signs of early- and late-stage disease will help
answer this question.
Most of the results described above feature inoculation of NHPs with strains of T.
pallidum obtained from human infections. In one study, however, five macaques were
infected with the baboon-derived Fribourg-Blanc strain (Sepetjian et al. 1969). The
lesions developed by the macaques were characteristic of yaws rather than syphilis.
Rather than an indurated chancre, they displayed vegetative, hyperkeratotic papular
lesions. Interestingly, the TPI serology was weak in one animal and nonreactive in the
other three tested, one of which died 5 weeks into the study; the single weak reaction
did not even begin to develop until the eleventh week of infection. FTA-ABS serology
was similar, weak and slow to develop in two animals (one of which had the weak TPI
test) and nonreactive in the other two. It remains unclear whether the fact that the
antibody response of these macaques to an NHP-derived strain was so weak is signifi-
cant. Is the NHP response to T. pallidum strains native to their own species similarly
slow to their own species similarly slow to develop? At least for baboons naturally
infected with T. pallidum, Knauf et al. (unpublished data) were able to demonstrate
that anti-T. pallidum antibody titers in infected animals were generally high, with
mean anti-T. pallidum IgM+IgG titers of 1:2.94E+04 ± SE9.87E+03 in the initial clin-
ical stage of disease, 1:2.17E+05 ± SE1.83E+05 in the moderate stage, and
1.78E+06 ± SE1.38E+06 in the severe stage of infection. Moreover, comparison to a
PCR-based test demonstrated that sensitivity (Sen) and positive predictive values (PP) for
the antibody-based gelatin particle agglutination assay (Sen 100 %, PP 95 %), the FTA-
ABS IgG (Sen 100 %, PP 98 %), and the immunoblot IgG (Sen 100 %, PP 93 %) serologi-
cal tests in the baboons were reliable, although the specificity and negative prediction
values need more testing. Of course in studies of wildlife, the major disadvantage is the
paucity of information regarding the timing of infection. Thus, results regarding the
serological response of wild NHPs to infection need to be considered carefully.
There is one report of inoculating human “volunteers” with a strain of T. palli-
dum obtained from an NHP infection. In The Caracas Project, a study of late-stage
yaws and pinta carried out in 1969, five people were inoculated with the Fribourg-
Blanc strain (Smith et al. 1971). Two of the five patients developed reactive FTA-
ABS/TPI serology, while three had nonreactive tests. Furthermore, two of the three
with nonreactive serology had elevated IgG levels in their cerebrospinal fluid, indi-
cating that four of the five patients mounted some type of immunological response
to infection. Finally, abnormal ophthalmological results consistent with late-stage
yaws were found in three of the five patients, although it is not entirely clear that
these abnormalities were due to infection with the Fribourg-Blanc strain. Thus, the
Caracas Project showed that an NHP-derived T. pallidum strain could establish an
infection in humans and possibly cause ocular abnormalities. Unfortunately,
whether or not these infections yielded additional clinical signs in humans is
unknown, as the project focused only on neuro-ophthalmologic manifestations. Nor
was the length of experimental infection or a discussion of the ethics of infecting
healthy adults with T. pallidum and allowing them to develop late-stage infection
provided.
210 K.N. Harper and S. Knauf
As described above, even though T. pallidum infections have not been documented
in species such as macaques, owl, and squirrel monkeys in the wild (Table 4), labo-
ratory experiments show that these animals can be infected with T. pallidum. If
these species are susceptible to infection, what prevents them from serving as hosts
in the wild? Which is more important in explaining why NHP species in Africa but
not Asia and South America are infected with T. pallidum: host geography (Fig. 4)
or phylogeny (Fig. 5)? Is the reason for this inconsistency across primate species
rooted in biological differences, the varied environments they inhabit, or chance
playing out over evolutionary time? The genital ulcerative disease caused by T. pal-
lidum in olive baboons, which has been reported exclusively in Papio anubis (Knauf
et al. 2012; Wallis and Lee 1999), poses a similar puzzle. Olive and yellow baboons
are known to hybridize (Alberts and Altmann 2001), which means that a significant
level of sexual interaction occurs between the two species. However, even at
Treponema hot spots in Tanzania where olive and yellow baboon subspecies overlap
and transmission opportunities should arise frequently, yellow baboons have never
been observed with genital lesions consistent with T. pallidum infection. Could
Treponema pallidum Infection in Primates: Clinical Manifestations, Epidemiology… 211
some constellation of host genetics in yellow baboons provide them with physiolog-
ical protection against the disease? Or are differences in behavior between olive and
yellow baboons responsible for the presence of the infection in one subspecies but
not the other? The factors that determine whether or not an NHP clade serves as a
natural host remain mysterious at this point.
Parallel questions exist about the importance of strain vs. host characteristics in
determining clinical manifestations. For example, is the recently described infection
in Tanzanian baboons, characterized by genital ulceration, caused by a strain or
strains which are genetically distinct from the strains that cause milder, non-genitally
ulcerating infections in the baboons of West Africa? Or is some characteristic of the
Tanzanian baboons, or their environment, responsible for these novel manifesta-
tions? Previously, it has been demonstrated that climatic factors such as temperature
and humidity play an important role in modulating clinical lesions in T. pallidum
infection (Turner and Hollander 1957a). Could this important finding help explain
the disease dynamics observed in some of the naturally occurring NHP epidemics
in East Africa (Collins et al. 2011)? Detailed surveys of T. pallidum manifestations
in different NHP groups may help begin to answer some of these questions, allow-
ing us to begin to disentangle the effects of the evolution of host and pathogen from
environment.
212 K.N. Harper and S. Knauf
Fig. 5 Species distribution of T. pallidum infection in nonhuman primates. Drawn from sources
cited in Tables 2–4; only species in which more than ten animals had been tested were included
Conclusion
Surveys of T. pallidum infection in NHPs performed thus far have provided fasci-
nating insights into the ecology of treponemal disease in the wild. It appears that T.
pallidum infection in wild NHPs is common in many parts of Africa, but very rare,
if it occurs at all, in Asia and South America. Robust evidence of seropositive ani-
mals has only been found in Old World species of NHPs (Figs. 4 and 5). However,
more comprehensive testing of Asian and South American NHPs must be performed
before solid conclusions regarding the geographical and phylogenetic distribution
of infection may be drawn.
Because the history of T. pallidum infection in humans remains controversial,
studies of NHP strains may help elucidate the origin of our own. Thus far, only two
NHP strains have been characterized genetically. These two strains are both from
baboons, one from Guinea and the other from Tanzania, and both are closely related
Treponema pallidum Infection in Primates: Clinical Manifestations, Epidemiology… 213
to human T. pallidum subsp. pertenue strains (Harper et al. 2008; Knauf et al. 2012),
which are responsible for the disease yaws. However, it is not clear whether the
NHP strains are merely closely related to subsp. pertenue or whether they are actu-
ally members of the subspecies. Sampling strains from other NHP species and geo-
graphic regions and sequencing their entire genomes using next-generation methods
should help us better understand how NHP strains are related to human strains and
to one another. If T. pallidum strains tend to diverge in the same order as their host
species, this would indicate that the infection is ancient in primates, with each spe-
cies possessing its own genetically distinct type of pathogen. It could also indicate
that T. pallidum infected our earliest hominid ancestors. In contrast, if T. pallidum
strains diverge in patterns that do not reflect their hosts’ phylogenies, host switching
may be implicated—and it is possible that T. pallidum infection may be a more
recent addition to the human disease-scape.
The clinical signs that have been described in association with T. pallidum infec-
tion in NHPs are primarily consistent with mild, nonvenereal infection. However,
the T. pallidum infection described in baboons in Tanzania is interesting in that it
appears to be associated with the genitals and possibly spread via sexual transmis-
sion. Our field observations of hundreds of baboons at Lake Manyara (Knauf et al.
2012; Mlengeya 2004) echo the findings reported from Gombe (Wallis 2000a;
Wallis and Lee 1999), in that the clinical lesions we observed are almost exclusively
found in the anogenital region and appear to be limited to sexually mature individu-
als. Further epidemiological characterization of this infection should clarify the
means of spread.
Given the fact that the human T. pallidum subspecies, transmitted via different
modes, are genetically distinct (Centurion-Lara et al. 2006; Harper et al. 2008;
Smajs et al. 2011), the same may be true of NHP strains transmitted via alternate
routes. If future studies reveal that some NHP strains are consistently sexually trans-
mitted while others are not, then comparison of closely related strains transmitted
via different routes may help generate hypotheses about the genetic polymorphisms
underlying sexual transmission. It is possible that comparison of genetic polymor-
phisms in T. pallidum associated with sexual transmission in both humans and
NHPs could identify genetic regions that are especially important in adaptation to
this mode of transmission. In addition, the selective pressures that favor sexual
transmission of the pathogen in some NHPs but not others could be studied;
environmental, social, immunological, and other factors might all be investi-
gated as possible driving forces. In short, the spectrum of different transmission
routes and types of clinical manifestations in NHPs may provide a unique
opportunity to better understand the pathogenicity and evolution of T. pallidum
in our own species.
Current efforts to study T. pallidum infection in wild NHPs are hampered by the dif-
ficulty of gathering adequate biological samples, especially in endangered species such
as the great apes. A solid diagnosis of treponemal disease can only be obtained by dem-
onstrating that the spirochete is present, and carefully documenting microscopic and
macroscopic signs of infection, including molecular biological proof, is essential.
Serology can be used to screen large numbers of wild NHPs for T. pallidum infection,
214 K.N. Harper and S. Knauf
but it requires that blood samples be taken. Moreover, because serology cannot aid in
differentiating between the T. pallidum subspecies, nor can it provide information about
clinical manifestations, disease progression, or even whether an infection is active,
serum samples are not sufficient for a truly comprehensive examination. Though obtain-
ing specimens from NHPs for histological and molecular biological tests is often very
difficult, in order to reach a definitive diagnosis using the methods currently available,
access to these types of samples is a necessity.
In the future, it is hoped that noninvasive means of studying treponemal infection
in wild NHPs may be developed, which could greatly facilitate diagnosis and sur-
veillance. Antibody-based tests have been successfully used on urine and fecal
samples to study NHP pathogens such as simian immunodeficiency virus (SIV)
(Santiago et al. 2003), simian foamy virus (SFV) (Liu et al. 2008), hepatitis B
(Makuwa et al. 2005), and Cryptosporidium and Giardia (Salzer et al. 2007). In
addition, urine and fecal samples have been used to obtain DNA from NHP patho-
gens such as SIV (Santiago et al. 2003), SFV (Liu et al. 2008), respiratory viruses
(Köndgen et al. 2010), and Plasmodium spp. (Liu et al. 2010), paving the way for a
flurry of recent phylogeographic studies of unprecedented size. Because it is not
clear whether urine or feces are reasonable samples for T. pallidum antibody-based
tests, or whether pathogen DNA is shed in urine or feces at detectable levels during
the various stages of infection, the potential of noninvasive samples to revolutionize
studies of T. pallidum in NHPs is unclear and currently under investigation (Knauf
et al., unpublished data).
Understanding treponemal infection in NHPs is important for several reasons.
First, it is possible that morbidity and mortality associated with T. pallidum infec-
tion has important conservation implications, especially for species such as the
great apes (Levréro et al. 2007). Second, a dialogue about reinitiating the yaws
eradication campaign that began mid-century is currently active (Asiedu et al. 2008;
Rinaldi 2008). Before the feasibility of such a campaign may be assessed, we must
understand the relevant disease ecology. Frequent transmission from an animal res-
ervoir would have serious ramifications for an attempt to eradicate infection in
humans, and phylogenetic studies can help determine whether or not this occurs.
Third, understanding the evolution and transmission of T. pallidum among diverse
primate species may help us better understand the forces that mold its evolution and
transmission dynamics within our own species. Further study may shed light on the
geographic distribution, the mode of transmission, and the NHP species affected by
T. pallidum infection, as well as the relationship between human and NHP strains.
References
Alberts SC, Altmann J (2001) Immigration and hybridization patterns of yellow and anubis
baboons in and around Amboseli, Kenya. Am J Primatol 53(4):139–154
Antal G, Lukehart SA, Meheus AZ (2002) The endemic treponematoses. Microbes Infect
4:83–94
Arya OP, Bennett FJ (1976) Role of the medical auxiliary in the control of sexually transmitted
disease in a developing country. Br J Vener Dis 52(2):116–121
Asiedu K, Amouzou B, Dhariwal A, Karam M, Lobo D, Patnaik S, Meheus A (2008) Yaws eradica-
tion: past efforts and future perspectives. Bull World Health Organ 86(7):499–500
Baker B, Armelagos G (1988) The origin and antiquity of syphilis: paleopathological diagnosis
and interpretation. Curr Anthropol 29(5):703–737
Baughn RE, Jorizzo JL, Adams CB, Musher DM (1988) Ig class and IgG subclass responses to
Treponema pallidum in patients with syphilis. J Clin Immunol 8(2):128–139
Baylet R, Thivolet J, Sepetjian M, Nouhouay Y, Baylet M (1971) La tréponématose naturelle
ouverte du singe Papio papio en Casamance. Bulletin de la Sociêtê de Pathologie Exotique et
de ses Filiales 64(6):842–846
Binnicker M, Jespersen D, Rollins L (2011) Treponema-specific tests for serodiagnosis of syphilis:
comparative evaluation of seven assays. J Clin Microbiol 49(4):1313–1317
Black F (1975) Infectious diseases in primitive societies. Science 187:515–518
Carlson JA, Dabiri G, Cribier B, Sell S (2011) The immunopathobiology of syphilis: the manifes-
tations and course of syphilis are determined by the level of delayed-type hypersensitivity. Am
J Dermatopathol 33(5):433–460
Castellani A (1907) Experimental investigations on framboesia tropica (yaws). J Hyg 7(4):558–569
Centurion-Lara A, Castro C, Castillo R, Shaffer J, Wv V, Lukehart S (1998) The flanking region
sequences of the 15-kDa lipoprotein gene differentiate pathogenic treponemes. J Infect Dis
177(4):1036–1040
Centurion-Lara A, Molini B, Godornes C, Sun E, Hevner K, Wv V, Lukehart S (2006) Molecular
differentiation of Treponema pallidum subspecies. J Clin Microbiol 44(9):3377–3380
Clark J, Yobs A (1968) Observations of the pathogenesis of syphilis in Aötus trivirgatus. Br J Vener
Dis 44:208–215
Cockburn T (1967) The evolution of human infectious diseases. In: Cockburn T (ed) Infectious
diseases: their evolution and eradication. Charles C Thomas, Springfield, IL, pp 84–107
Collins A, Sindimwo A et al (2011) Reproductive disease in olive baboons (Papio anubis) of
Gombe National Park: Outbreak, time-course and attempts to limit recurrence. In: The 8th
TAWIRI Scientific Conference. Tanzania Wildlife Research Institute, Arusha, Tanzania
Cousins D (1972) Diseases and injuries in wild and captive gorillas. International Zoo Yearbook
12:211–218
Cousins D (1984) Notes on the occurrence of skin infections in gorillas. Der Zoologische Garten
54(4–5):333–338
Cousins D (2008) Possible goundou in gorillas. Gorilla J 37:22–24
Crosby A (1969) The early history of syphilis: a reappraisal. Am Anthropol 71(2):218–227
Csonka GW (1953) Clinical aspects of bejel. Br J Vener Dis 29:95–103
Denis A (1963) On safari: the story of my life. Collins, London
Diaz de Isla R (1539) Treatise on the Serpentine Malady, which in Spain is commonly called
Bubas, which was drawn up in the Hospital of All Saints in Lisbon.
Edroma E, Rosen N, Miller P (1997) Conserving the Chimpanzees of Uganda
Elsas F, Smith J, Israel C, Gager W (1968) Late syphilis in the primate. Br J Vener Dis
44:267–273
Engelhardt H (1959) A study of yaws (does congenital yaws occur?). J Trop Med Hyg
62:238–240
Fegan D, Glennon M, Thami Y, Pakoa G (2010) Resurgence of yaws in Tanna, Vanuatu: time for a
new approach? Trop Doct 40:68–69
216 K.N. Harper and S. Knauf
Felsenfeld O, Wolf R (1971) Serological reactions with treponemal antigens in nonhuman pri-
mates and the natural history of treponematosis in man. Folia Primatol 16:294–305
Fribourg-Blanc A, Mollaret H (1969) Natural treponematosis of the African primate. Primates
Med 3:113–121
Fribourg-Blanc A, Niel G, Mollaret H (1966) Confirmation serologique et microscopique de la
treponemose du cynocephale de guinee. Bulletin de la Societe de Pathologie Exotique et de Ses
Filiales 59(1):54–59
Gersti S, Kiwila G, Dhorda M, Lonlas S, Myatt M, Ilunga B, Lemasson D, Szumilin E, Guerin P,
Ferradini L (2009) Prevalence study of yaws in the Democratic Republic of Congo using the
lot quality assurance sampling method. PLoS One 4(7):e6338
Gevers D, Cohan F, Lawrence J, Spratt B, Coenye T, Feil E, Stackebrandt E, Peer YD, Vandamme
P, Thompson F et al (2005) Re-evaluating prokaryotic species. Nat Rev Microbiol 3:733–739
Gormus B, Xu K, Alford P, Lee D, Hubbard G, Eichberg J, Meyers W (1991) A serologic study of
naturally acquired leprosy in chimpanzees. Int J Lepr 59(3):450–457
Gray R, Mulligan C, Molini B, Sun E, Giacani L, Godornes C, Kitchen A, Lukehart S, Centurion-
Lara A (2006) Molecular evolution of the tprC, D, I, K, G and J genes in the pathogenic genus
Treponema. Mol Biol Evol 23(11):2220–2233
Guthe T, Reynolds FW, Krag P, Willcox RR (1953) Mass treatment of treponemal diseases, with
particular reference to syphilis and yaws. Br Med J 1(4810):594–598
Guthe T, Ridet J, Vorst F, D’Costa J, Grab B (1972) Methods for the surveillance of endemic trepo-
nematoses and sero-immunological investigations of “disappearing” disease. Bull World
Health Organ 46(1):1–14
Hackett C (1953) Extent and nature of the yaws problem in Africa. Bull World Health Organ
8(1–3):127–182
Harper K, Ocampo P, Steiner B, George R, Silverman M, Bolotin S, Pillay A, Saunders N,
Armelagos G (2008) On the origin of the treponematoses: a phylogenetic approach. PLoS
NTDs 2(1):e148
Harper K, Zuckerman M, Harper M, Kingston J, Armelagos G (2011) The origin and antiquity of
syphilis revisited: an appraisal of Old World Pre-Columbian evidence for treponemal infection.
Yearbook of Physical Anthropology 54:99–133
Herring A, Ballard R, Mabey D, Peeling R (2006) Evaluation of rapid diagnostic tests: syphilis.
Nat Rev Microbiol 4(12):S33–S40
Holmes K, Sparling P, Stamm W, Piot P, Wasserheit J, Corey L, Cohen M (2007) Sexually trans-
mitted diseases. McGraw Hill, New York
Hopkins D, Flórez D (1977) Pinta, yaws, and venereal syphilis in Colombia. Int J Epidemiol
6(4):349–355
Hubbard G, Lee D, Eichberg J, Gormus B, Xu K, Meyers W (1991) Spontaneous leprosy in a
chimpanzee (Pan troglodytes). Vet Pathol 28:546–548
Hudson E (1963) Treponematosis and anthropology. Ann Intern Med 58:1037–1048
Hunter E, Russell H, Farshy C, Sampson J, Larsen S (1986) Evaluation of sera from patients with
lyme disease in the fluorescent treponemal antibody-absorption test for syphilis. Sex Transm
Dis 13(4):232–236
Jowi J, Gathua S (2005) Lyme disease: report of two cases. East Afr Med J 82(5):267–269
Karesh W (2000) Suivi de la santé des gorilles au Nord-Congo. Canopée 18:16–17
Karp G, Schlaeffer F, Jotkowitz A, Riesenberg K (2009) Syphilis and HIV co-infection. Eur J
Intern Med 20(1):9–13
Knauf S, Batamuzi E, Mlengeya T, Kilewo M, Lejora I, Nordhoff M, Ehlers B, Harper K,
Fyumagwa R, Hoare R et al (2012) Treponema infection associated with genital ulceration in
wild baboons. Vet Pathol 49(2):292–303
Köndgen S, Schenk S, Pauli G, Boesch C, Leendertz F (2010) Noninvasive monitoring of respira-
tory viruses in wild chimpanzees. Ecohealth 7:332–341
Kuhn U (1970) The treponematoses. The Chimpanzee 3:71–81
Kumm H, Turner T (1936) The transmission of yaws from man to rabbits by an insect vector,
Hippelates pallipes loew. Am J Trop Med 16(3):245–271
Treponema pallidum Infection in Primates: Clinical Manifestations, Epidemiology… 217
Pampiglione S, Wilkinson A (1975) A study of yaws among pygmies in Cameroon and Zaire. Br J
Vener Dis 51:165–169
Perine P, Hopkins D, Niemel P, John RS, Causse G, Antal G (1984) Handbook of endemic
Treponematoses: yaws, endemic syphilis, and pinta. World Health Organization, Geneva
Richens J, Mabey CW (2009) Sexually transmitted infections (Excluding HIV). In: Cook GC,
Zumla AI (eds) Manson’s tropical diseases, 22nd edn. Saunders-Elsevier, Philadelphia, pp
403–434
Rinaldi A (2008) Yaws: a second (and maybe last?) chance for eradication. PLoS NTDs 2(8):e275
Rolfs R, Goldberg M, Sharrar R (1990) Risk factors for syphilis: cocaine use and prostitution. Am
J Public Health 80(7):853–857
Román G, Román L (1986) Occurrence of congenital, cardiovascular, visceral, neurologic, and
neuro-ophthalmologic complications in late yaws: a theme for future research. Rev Infect Dis
8(5):760–770
Rothschild B, Turnbull W (1987) Treponemal infection in a Pleistocene bear. Nature 329:61–62
Rothschild B, Hershkovitz I, Rothschild C (1995) Origin of yaws in the Pleistocene. Nature
378:343–344
Russell H, Sampson J, Schmid G, Wilkinson H, Plikaytis B (1984) Enzyme-linked immunosorbent
assay and indirect immunofluorescence assay for Lyme disease. J Infect Dis 149(3):465–470
Salazar J, Hazlett KRO, Radolf JD (2002) The immune response to infection with Treponema pal-
lidum, the stealth pathogen. Microbes Infect 4:1133–1140
Salomone G (1999) Le pian chez les peuples de la forêt équatoriale du Nord-Congo et du sud de la
République Centrafricaine. In: Bahuchet S, Bley D, Pagezy H, Vernazza-Licht N (eds)
L’Homme et la Forêt Tropicale. University of Provence Press, Marseilles, pp 675–688
Salzer J, Rwego I, Goldberg T, Kuhlenschmidt M, Gillespie T (2007) Giardia sp. and
Cryptosporidium sp. infections in primates in fragmented and undisturbed forest in Western
Uganda. J Parasitol 93(2):439–440
Santiago M, Lukasik M, Kamenya S, Li Y, Bibollet-Ruche F, Bailes E, MUller M, Emery M,
Goldenberg D, Lwanga J et al (2003) Foci of endemic simian immunodeficiency virus infection
in wild-living Eastern chimpanzees (Pan troglodytes schweinfurthii). J Virol 77(13):7545–7562
Satchell G, Harrison R (1953) Experimental observations on the possibility of transmission of yaws
by wound-feeding Diptera, in Western Samoa. Trans R Soc Trop Med Hyg 47(2):148–153
Satter E, Tokarz V (2010) Secondary yaws: an endemic treponemal infection. Pediatr Dermatol
27(4):364–367
Schroder NW, Eckert J, Stubs G, Schumann RR (2008) Immune responses induced by spirochetal
outer membrane lipoproteins and glycolipids. Immunobiology 213(3–4):329–340
Seña A, White B, Sparling P (2010) Novel Treponema pallidum serologic tests: a paradigm shift in
syphilis screening for the 21st century. Clin Infect Dis 51(6):700–708
Sepetjian M, Guerraz F, Salussola D, Thivolet J, Monier J (1969) Contribution à l’étude du trépo-
nème isolé du singe par A. Fribourg-Blanc. Bull WHO 40:141–151
Sepetjian M, Thivolet J, Salussola D, Guerraz F, Monier J (1972) Étude comparative du FTA et du
FTA ABS quantitatifs chez l'homme et l'animal a différents stades de la syphilis. Patholologie
Biologie 20(9):449–455
Smajs D, Norris S, Weinstock G (2011) Genetic diversity in Treponema pallidum: implications for
pathogenesis, evolution and molecular diagnostics of syphilis and yaws. Infect Genet Evol
12(2):191–202
Smith J (1969) Late ocular syphilis: transfer of infection from man to experimental animals. Trans
Am Opthalmol Soc 67:658–697
Smith J, Singer J, Reynolds D, Moore M, Yobs A, Clark J (1965) Experimental ocular syphilis and
neurosyphilis. Br J Vener Dis 41:15–23
Smith J, David N, Indgin S, Israel C, Levine B, Justice J, McCrary J, Medina R, Paez P, Santana E
et al (1971) Neuro-ophthalmological study of late yaws and pinta II. The Caracas Project. Br J
Vener Dis 47:226–251
St John R (1985) Yaws in the Americas. Rev Infect Dis 7(S2):S266–S272
Treponema pallidum Infection in Primates: Clinical Manifestations, Epidemiology… 219
Introduction
Herpesviruses are large double-stranded DNA viruses that can establish life-long
infection in their respective hosts and can undergo two different phases in their life
cycle: lytic or latent (Fig. 1a) (Pellett and Roizman 2007). Lytic replication is char-
acterized by the expression of most viral genes in an ordered cascade (immediate
early, early, and late), leading to the production of infectious virions. Latency is
marked by minimal viral gene expression and the maintenance of the viral genome
in the nucleus. Reactivation to lytic replication from latency can be triggered by
multiple factors, such as stress or chemical reagents. The ability of herpesviruses to
utilize these two very distinct modes of replication is an excellent survival strategy
as establishment of latency with periodic reactivation may facilitate persistent infec-
tion in the host while allowing evasion from the immune system.
Herpesviruses are prevalent in nature, with most animal species being infected
by at least one herpesvirus (Knipe and Howley 2007). Out of more than 200 herpes-
viruses identified to date, only eight are endemic to humans and are grouped into
three subfamilies (α, β, γ) based on their structural and biological properties
(Table 1). γ-Herpesviruses are lymphotropic and are further divided into two sub-
groups: γ-1 (Lymphocryptovirus) and γ-2 (Rhadinovirus) (Fig. 2). Epstein-Barr
virus (EBV) was the first γ-herpesvirus to be discovered and is the prototype mem-
ber of the γ-1 group (Diehl et al. 1968). Kaposi’s sarcoma-associated herpesvirus
(KSHV) is so far the only human virus assigned to the Rhadinovirus family, which
also includes herpesvirus saimiri (HVS), rhesus rhadinovirus (RRV), and mouse
herpesvirus 68 (mHV68) (Table 2) (Blaskovic et al. 1980; Chang et al. 1994;
Desrosiers et al. 1997; Melendez et al. 1968; Moore et al. 1996).
L.-Y. Wong (*) • Z. Toth • K.F. Brulois • K.-S. Inn • S.H. Lee • H.-R. Lee • J.U. Jung
Department of Molecular Microbiology and Immunology, University of Southern California,
Keck School of Medicine, Los Angeles, CA 90033, USA
e-mail: [email protected]
J.F. Brinkworth and K. Pechenkina (eds.), Primates, Pathogens, and Evolution, 221
Developments in Primatology: Progress and Prospects,
DOI 10.1007/978-1-4614-7181-3_8, © Springer Science+Business Media New York 2013
222 L.-Y. Wong et al.
Fig. 1 (a) Diagram showing the two phases of a herpesvirus life cycle. Upon infection, the virus
can enter latency where the viral genome is typically kept in an episomal form in the cell nucleus
and only a few viral genes are expressed. Alternatively, the virus may undergo lytic replication to
produce infectious virus that go on to infect new cells. Viral genes are expressed in an ordered
cascade during lytic replication, starting with immediate-early genes, early genes, and late genes.
Common functions of the proteins encoded by these lytic genes are listed underneath each group.
Late gene expression only occurs after viral DNA replication. (b) A burst of lytic replication
occurs after infection by a herpesvirus, after which the virus will establish latency in the now
infected host. Periodically the virus will reactivate to produce new virus, followed by another
round of latency where no virus is produced
Fig. 2 Phylogenetic tree showing the three subfamilies of herpesviruses (α, β, γ). The tree was
constructed using sequences of the viral DNA polymerase catalytic subunit. The program
MUSCLE was used for alignment and PhyML for phylogeny analysis (Dereeper et al. 2008)
KSHV was first discovered in 1994 as the etiological agent in Kaposi’s sarcoma
(KS) lesions and has since been further implicated in primary effusion lymphoma
(PEL) and multicentric Castleman’s disease (MCD) (Cesarman et al. 1995; Chang
et al. 1994; Soulier et al. 1995). KSHV DNA has been detected in all cases of KS
using multiple assays such as polymerase chain reaction and immunohistochemis-
try (Boshoff et al. 1995; Dupin et al. 1999). The first case of KS was described in
1872 by Dr. Moritz Kaposi as an indolent tumor affecting elderly men of
Mediterranean and Jewish origins (Iscovich et al. 2000; Kaposi 1872). This later
became known as classical KS after three other epidemiological forms were
224 L.-Y. Wong et al.
(Aluigi et al. 1996; Ganem 2006; Salahuddin et al. 1988). These spindle cells secrete
many proinflammatory and angiogenic factors, which contribute to the inflammatory
and neovascular characteristics of KS (Salahuddin et al. 1988). The driving force
behind spindle cells has largely been attributed to the expression of a viral latent pro-
tein; FLICE-inhibitory protein (vFLIP); an antiapoptotic, anti-autophagy factor; and a
strong inducer of the nuclear factor kappa-light-chain-enhancer of activated B-cell
(NF-κB) pathway (Grossmann et al. 2006; Lee et al. 2009).
KSHV is also associated with PEL and MCD, rare lymphoproliferative disorders
of B-cell origin usually diagnosed in AIDS patients. Coinfection with EBV is com-
monly found in cases of PEL, an aggressive subtype of non-Hodgkin B-cell lym-
phoma. PEL is characterized by proliferation of cells that lack most B-cell markers
but have features reminiscent of plasmablasts (Jenner et al. 2003). MCD is a rare,
polyclonal lymphocyte hyperplasia in which KSHV-infected cells are found to be
IgM-λ positive B cells located in the mantle zone in the lymph node (Du et al.
2001). In vitro infection of human tonsillar B cells with KSHV revealed that the
virus has a propensity to infect IgM-λ positive B cells and subsequently acquire
characteristics similar to infected cells in MCD (Hassman et al. 2011). It is unclear
why KSHV shows favoritism towards a specific subtype of B cells or how this
specificity is linked to the pathogenesis seen in MCD.
In nature, KSHV infection is restricted to humans. In an experimental setting,
KSHV could infect a variety of primary cells and cell lines from different species
such as human, mouse, and rat (Lagunoff et al. 2002; McAllister and Moses 2007;
Renne et al. 1998). Despite the promiscuous ability of KSHV to infect multiple cell
types, these infected cells are not immortalized or transformed and, without any
selective pressure such as antibiotic selection, tend to lose the viral genome after
several passages (Lagunoff et al. 2002). Recently (2012), Jones et al. successfully
developed a KSHV-induced tumorigenesis in vitro model using primary rat mesen-
chymal cell (Jones et al. 2012).
The in vivo study of KSHV and its malignancies uses mostly nude, transgenic,
or humanized mouse model systems (Dittmer et al. 1999; Wu et al. 2006). KSHV-
positive PEL cells are transplantable into nude mice lacking functional B and T
cells, but the virus does not spread to murine cells (Picchio et al. 1997). The study
of vFLIP using transgenic mice showed increased incidence of lymphomas, and
when this viral protein expression is restricted to B cells, the mice acquired pheno-
types similar to MCD abnormalities (Ballon et al. 2011; Chugh et al. 2005). The
first nonhuman primate model of KSHV infection was established using common
marmosets (Callithrix jacchus) (Chang et al. 2009). Infection with recombinant
KSHV led to persistent infection in these animals, and one marmoset showed devel-
opment of KS-like lesion, which shared similar histopathological characteristics as
human AIDS-associated KS lesions (such as presence of spindle cells with detect-
able expression of KSHV proteins). It would be interesting to determine if the
KS-like lesion progresses similarly in this common marmoset animal model
compared to human KS.
226 L.-Y. Wong et al.
Sequencing of the complete KSHV genome revealed that KSHV encodes multiple
ORFs with noticeable homology to cellular genes (Neipel et al. 1997). These pirated
genes are involved in multiple pathways such as antiapoptosis, autocrine/paracrine
signaling, immune responses, and cell cycle regulation. Most of these “pilfered”
genes are also found in other Rhadinoviruses at the equivalent genomic position,
suggesting that the piracy event may have happened early during the virus evolu-
tion. KSHV v-cyclin, a latent gene, is a type D cyclin with 53 % similarity to both
HVS and cellular cyclin (Li et al. 1997). In addition, some of the genes first identi-
fied as unique to KSHV turned out to have cellular equivalents that were discovered
after the characterization of the viral counterparts. Two notable examples are the
KSHV K3 and K5 genes, also known as the modulator of immune recognition
MIR-1 and MIR-2, respectively. K3 and K5 are ubiquitin E3 ligases and potent
immune dysregulators that downregulate a variety of surface immune receptors,
such as MHC class I (Coscoy and Ganem 2000; Ishido et al. 2000). Studies on K3
and K5 led to the identification of a family of cellular ubiquitin E3 ligases called
membrane-associated RING CH domain (MARCH) proteins which share a com-
mon enzymatic motif and domain organization (Goto et al. 2003). A list of pirated
viral genes and their cellular homologues is provided in Table 3.
While the few examples of viral molecular mimicry discussed so far contain
cellular homologues that are conserved not only in function but also in sequence,
the γ-2-herpesviruses also contain unique genes that have no cellular counterparts
but functionally mimic a cellular protein. The first and last ORFs of these viruses
(but the first and second ORFs for HVS) encode transmembrane proteins involved
in modulating lymphocyte activation events and are unique to the respective viruses
(K1 and K15 for KSHV, R1 and R15 for RRV, Tip and STP for HVS) (Fig. 3).
Molecular Mimicry by γ-2 Herpesviruses to Modulate Host Cell Signaling Pathways 227
Fig. 3 Genomic location of the viral transmembrane proteins and viral pre-miRNAs of KSHV,
RRV, and HVS. In HVS strains A and B, STP-A and STP-B are the first ORFs of the genome, but
in HVS subgroup C, Tip is the first ORF followed by STP-C as the second ORF. The pre-miRNAs
are located in the latency-associated region. The U-rich RNAs of HVS are located at the left end
region of the viral genome. Dihydrofolate reductase (DHFR)
KSHV K1
K1, the first ORF of KSHV, is located at the 5′ end of the KSHV genome after the
terminal repeats (Fig. 2) (Lagunoff and Ganem 1997; Lee et al. 1998b; Russo et al.
1996). K1 is expressed at low levels during latency, and its presence has been
228 L.-Y. Wong et al.
Fig. 4 Structures of the transmembrane signaling proteins of KSHV (K1 and K15), RRV (R1 and
RK15), and HVS (Tip, STP-C and STP-A). The pathways depicted by squares indicate they are
activated by the respective viral proteins, while those in octagons indicate inhibition of signaling.
P-Y denotes phosphorylated tyrosine residues
detected in PEL, MCD, and KS tumors (Lagunoff and Ganem 1997; Lee et al. 2003;
Samaniego et al. 2001). K1 encodes a type-I transmembrane glycoprotein that mim-
ics the signaling activity of the cellular B-cell receptor (BCR) (Fig. 4) (Lee et al.
1998a). Its extracellular domain is highly divergent between different KSHV strains,
yet they all maintain the ability to oligomerize via this region (Lagunoff et al. 1999;
Lee et al. 1998b). On the other hand, the cytoplasmic tails of K1 variants are highly
conserved and contain an immunoreceptor tyrosine-based activation (ITAM) motif,
which is involved in cell activation signals (Cambier 1995; Lee et al. 1998a). The
structure of K1 is similar to the BCR, and indeed K1 expression has been shown to
generate a signaling profile reminiscent of BCR activation. At the same time, K1
deregulates BCR by retaining the newly synthesized μ chain of the BCR complex in
the endoplasmic reticulum (ER) as well as promoting the internalization of surface
BCRs via clathrin-dependent endocytosis (Lee et al. 2000; Tomlinson and Damania
2008). Thus, K1 can be localized to the cell surface, ER, and endosomes (early and
recycling) (Tomlinson and Damania 2008). The presence of K1 in recycling endo-
somes strongly suggests that K1 is recycled after endocytosis; however, direct evi-
dence for this scenario has yet to be presented (Table 4).
Through its ITAM motif, K1 induces calcium mobilization and increases tyro-
sine phosphorylation of cellular proteins leading to the activation of nuclear factor
for activated T cells (NFAT) and NF-κB, events reminiscent of BCR activation
(Lagunoff et al. 1999; Lee et al. 1998a). A prerequisite step for initiation of BCR
signaling is the clustering of multiple BCRs, which only occurs after ligand bind-
ing. In contrast, K1 is constitutively active and does not require extracellular sig-
nals. It has been postulated that this intrinsic activation of K1 is due to its extracellular
domain, which mediates multimerization even in the absence of ligand binding
(Lagunoff et al. 1999). K1 interacts with a variety of cellular signal transduction
Table 4 Summary of the ORFs discussed
KSHV K1 KSHV K15 RRV R1 RRV RK15 HVS STP HVS Tip
Interaction partners Syk, Lyn, Vav, Src, Hck, Lyn, Fyn, Syk Src, TRAFs, Ras Lck, p80, STAT-1 and
PLCγ2, Grab, Yes (STP-C), STAT3 -3
PTK-1 and −2
Signaling motifs ITAM—phosphory- SH2—phosphory- ITAM—phosphory- SH3 only TRAF-binding motif SH3 and CSKH
lated by Src lated by Src lated by Syk (STP-A and (phosphorylated by
kinases, SH3, STP-C), Lck)
TRAF-binding collagen-like
motif repeats (STP-A
and STP-C), SH2
(STP-A and
STP-B)
Ligand dependent No. Multimerization No. Oligomerization
of extracellular of collagen-like
domain repeats
Ca2+ mobilization Increased upon αK1 Inhibited upon BCR Increased upon Decreased
stimulation stimulation stimulation
Cellular pTyr level Increased Decreased Increased Decreased
Activated pathways NFAT, NF-κB, AP-1, NF-κB (SH2 NFAT NF-κB (SH2 NF-κB, NFAT, NF-κB, NFAT
VEGF, PI3K/Akt dependent), dependent), MAPK, AP-1
AP-1 (K15P JNK
only), MAPK
(TRAF and SH2
dependent), JNK
(K15M only),
OncomiRs (only
Molecular Mimicry by γ-2 Herpesviruses to Modulate Host Cell Signaling Pathways
K15M tested)
Downregulate BCR surface levels BCR signaling and Lck activity, TCR and
surface levels CD4 surface levels
(continued)
229
Table 4 (continued)
230
KSHV K1 KSHV K15 RRV R1 RRV RK15 HVS STP HVS Tip
Transformation Yes. Rat-1 fibroblasts Yes. Rat-1 fibroblast STP-C—T cells from
and common and common rabbit, primates,
marmoset T cells marmoset T cells and human—
induces lym- induces lym-
phoma in phoma in rhesus
transgenic mice monkeys. STP-A
–lymphocytes
from common
marmosets
Consequences Cell survival, Angiogenesis, Antiapoptosis, Blocks TCR signaling
inflammatory metastasis IL-2-independent
cytokines, growth
angiogenesis,
blocks BCR
signaling, and
suppressing lytic
replication
L.-Y. Wong et al.
Molecular Mimicry by γ-2 Herpesviruses to Modulate Host Cell Signaling Pathways 231
KSHV K15
K15 is another gene unique to KSHV and is found at the 3′-end of the KSHV
genome, adjacent to the terminal repeats (Fig. 3). This gene contains eight exons
and multiple splicing variants with the dominant spliced form encompassing all the
exons to express a protein with 12 transmembrane domains followed by a short
cytoplasmic tail (Fig. 4) (Brinkmann et al. 2003; Choi et al. 2000b; Glenn et al.
1999). While K15 is widely considered to be a lytic gene, low levels of expression
can be detected in KSHV-positive PEL cell lines explanted from patients (Sharp
et al. 2002; Wong and Damania 2006). It is uncertain if this low level of K15 expres-
sion is due to lytic reactivation in a minority of infected cells (1–5 %) or if the
latently infected cells periodically express a low amount of K15. Two variants of
K15 exist: the predominant (P) form and the minor (M) form, termed K15P and
K15M, respectively (Glenn et al. 1999; Poole et al. 1999). These two variants show
low sequence identity at the protein level, but are almost structurally (splicing pat-
tern and protein domains) and functionally identical (Glenn et al. 1999; Poole et al.
1999). The cytoplasmic tails of the K15 variants are highly conserved, suggesting
that loss of function in this domain is deleterious to the virus. So far, most studies
have been done with the K15P form.
K15P
Expression of K15P interferes with BCR signaling but induces the activation of
NF-κB, activator protein-1 (AP-1), and mitogen-activated protein kinase (MAPK)
pathways, which ultimately leads to expression of genes involved in cell survival,
proliferation, and inflammation. Specifically, K15P activates the extracellular
signal-regulated kinases (ERK), c-Jun amino-terminal kinases (JNK), but not the
p38 cascade of the MAPK superfamily (Brinkmann et al. 2003, 2007; Cho et al.
2008; Pietrek et al. 2010, Tsai et al. 2009; Wang et al. 2009). The cytoplasmic
domain of K15 contains multiple motifs important in cellular signaling cascades,
including a tumor necrosis factor (TNF), receptor-associated factors (TRAF)-
interacting site, a Src homology 2 (SH2) domain, and a proline-rich SH3 motif
(Choi et al. 2000b). These domains mediate interactions with a variety of cellular
signaling molecules to modulate the pathways mentioned above. The SH2 and SH3
motifs of K15 interact with multiple Src kinases such as Src, Hck, Lck, Fyn, and
Yes, with the Y481 in the SH2 motif being constitutively phosphorylated by tyro-
sine kinases (Brinkmann et al. 2003). This phosphorylation is required for activa-
tion of NF-κB and MAPK (Brinkmann et al. 2003).
The K15 cytoplasmic tail is sufficient to downregulate BCR signaling, as shown
in a study using a chimeric protein composed of the K15 cytoplasmic portion fused
to the CD8α transmembrane domain and expressed in a B-cell line (Choi et al.
2000b). The inhibition of the calcium cascade upon BCR stimulation requires intact
SH2 and SH3 motifs of K15 (Choi et al. 2000b). It is hypothesized that the
Molecular Mimicry by γ-2 Herpesviruses to Modulate Host Cell Signaling Pathways 233
recruitment of Lyn by K15 sequesters this major B-cell kinase from BCR signaling
events (Cho et al. 2008). Indeed, mutating the SH3 motif of K15 interferes with its
ability to deregulate BCR signaling (Cho et al. 2008). Furthermore, K15 increases
the rate of BCR internalization from the cell surface; thus, it is likely that K15 uses
multiple mechanisms to inhibit BCR activation (Lim et al. 2007).
Despite the conservation of signaling motifs in the cytoplasmic tails of the K15P
and K15M variants, there are phenotypical differences between these two variants.
For one, even though both are predominantly cytoplasmic, the P variant contains a
mitochondria localization sequence and can thus localize to the host mitochondria
and other organelles such as ER and Golgi, while K15M is found on lysosomal
membranes (Choi et al. 2000b; Sharp et al. 2002; Tsai et al. 2009; Wang et al. 2007).
Although both K15 variants activate NF-κB to the same degree, only the P form can
positively regulate AP-1 activity (Tsai et al. 2009). Whether this distinct effect is
due to the difference in localization remains to be tested.
The importance of the SH2 motif in K15 to initiate the expression of cellular
cytokines involved in angiogenesis and inflammation (e.g., IL-8, IL-6, CXCL2) also
varies slightly between the two K15 variants. While mutation of the amino acids
tyrosine-glutamic acid-glutamic acid-valine at position 481–484 (Y481EEV) of the
K15P SH2 domain abolishes this phenotype, the K15M Y490EEV mutant retained
residual activity in cytokine upregulation, which may be due to the ability of K15M
to activate the JNK pathway independently of its SH2 motif (Pietrek et al. 2010;
Wang et al. 2007). Furthermore, K15M can upregulate miR-21 and miR-23, cellular
cancer-related miRNAs (oncomiRs) that are involved in tumor cell proliferation,
metastasis, and angiogenesis (Tsai et al. 2009). This ability is dependent on the
tyrosine residue in the SH2 domain. It is unclear whether K15P, which also has an
SH2 motif, can upregulate the same oncomiRs.
Despite the higher prevalence of the K15P isoform in KSHV isolates, the two
variants seem to function equivalently in most of the in vitro assays used to study
them (Pietrek et al. 2010; Wang et al. 2007). Whether the P variant has an in vivo
survival advantage over the M isolates (due to activation of AP-1 pathway) remains
to be tested.
Rhesus Rhadinovirus
RRV was first discovered in rhesus monkeys (Macaca mulatta) housed at the New
England Primate Research Center (NEPRC) (Desrosiers et al. 1997). The animals
reacted to HVS antigens in serology tests, hinting that they were infected by an
agent related to γ-herpesviruses. A further survey after the characterization of RRV
revealed that at least 90 % of rhesus monkeys housed at both the NEPRC and
234 L.-Y. Wong et al.
Oregon Regional Primate Research Center (ORPRC) were seropositive for RRV
(Mansfield et al. 1999; Wong et al. 1999). This uniformly high prevalence of infec-
tion at both centers suggested that RRV is endemic in rhesus monkeys. Sequencing
of its viral genome revealed that RRV is a close cousin of KSHV with high sequence
conservation. Indeed, both these Old World Rhadinoviruses share homologous
genes that are not found in New World Rhadinoviruses like HVS (e.g., viral inter-
leukin 6 (vIL-6)) (Alexander et al. 2000; Searles et al. 1999). RRV has been detected
in B cells of rhesus macaques, suggesting that, similar to KSHV, lymphocytes are
the major reservoir of viral persistence (Bergquam et al. 1999; Desrosiers et al.
1997; Wong et al. 1999). Unlike KSHV, which primarily enters latency upon de
novo infection, RRV can undergo lytic replication and produce high viral titer in
primary rhesus fibroblasts (Desrosiers et al. 1997).
To recapitulate the pathogenesis of KSHV often seen in HIV-coinfected patients,
two groups from NEPRC and ORPRC have infected rhesus monkeys with both
RRV and simian immunodeficiency viruses (SIVmac239, a virus used to model
HIV-like infection in rhesus monkeys) (Mansfield et al. 1999; Wong et al. 1999).
One group reported that transient lymphadenopathy followed by persistent infection
has been detected in the rhesus macaques infected with RRV alone (Mansfield et al.
1999). At 3 months postinfection, these pathologies disappeared as the immune
systems of the rhesus macaques were able to dampen the pathogenesis of the virus
(Mansfield et al. 1999). However, when the animals were coinfected with both RRV
and SIVmac239, they showed weaker antibody response to both viruses and had
shorter survival times compared to monkeys infected with only one of the agents.
The other group found that coinfected animals displayed symptoms of lymphopro-
liferative disorder such as splenomegaly and hypergammaglobulinemia in coin-
fected animals, symptoms reminiscent of MCD induced by KSHV in humans
(Wong et al. 1999). These results demonstrate the utility of RRV/SIVmac239 coin-
fection in monkeys as a model to recapitulate the KSHV-related lymphoprolifera-
tive disorders in HIV-positive patients.
RRV R1
Like K1, the R1 gene encodes a transmembrane protein with an extracellular immu-
noglobulin domain. However, the cytoplasmic tail of R1 contains multiple potential
ITAM motifs and is considerably longer than K1 (171 and 38 amino acids for R1
and K1, respectively) (Fig. 4) (Damania et al. 1999, 2000). Nonetheless, R1 is able
to transduce a signaling profile similar to that of K1-activated B cells (Damania
et al. 2000). R1 specifically interacts with the tyrosine kinase Syk and is itself a
substrate of Syk (Damania et al. 2000). Fusion of the cytoplasmic tail of R1 to the
extracellular and transmembrane domains of CD8 results in increased cellular tyro-
sine phosphorylation, intracellular calcium influx, and NFAT activation upon stimu-
lation with an αCD8 antibody, events that are reminiscent of B-cell activation
(Damania et al. 2000). In fact, expression of R1 is sufficient to constitutively
Molecular Mimicry by γ-2 Herpesviruses to Modulate Host Cell Signaling Pathways 235
activate NFAT, a key family of transcription factors important for lymphocyte acti-
vation, without any stimulation (Damania et al. 2000). Thus R1 is a close cousin of
K1 with respect to its ability to control B-cell signaling independently of BCR
(Fig. 4). Interestingly, the duration of calcium signaling was longer and stronger
with R1 than K1 upon stimulation with αCD8 antibody (Damania et al. 2000).
Whether this difference in activity strength is due to the longer cytoplasmic tail with
more potential ITAM motifs is unknown.
Likewise, R1 displays oncogenic traits similar to K1 (Damania et al. 1999).
When expressed in Rat-1 cells, R1 can induce foci formation in vitro and tumors
in vivo when inoculated into nude mice (Damania et al. 1999). Furthermore, recom-
binant HVS with R1, expressed in place of STP, could still immortalize T cells from
common marmosets, highlighting the ability of R1 to activate cellular pathways
leading to cell growth and survival, even in the absence of exogenous signals
(Damania et al. 1999). The authors did not mention if there were any differences in
the transforming abilities of HVS/K1 or HVS/R1 versus wild-type HVS (Damania
et al. 1999).
RRV RK15
The RK15 gene is situated at the same genomic location as KSHV K15 and is the
homologue of K15 in RRV (Fig. 3) (Alexander et al. 2000). It is termed RK15 as not
to cause confusion with an RRV gene already named R15 of strain 17577 that is the
RRV homologue of cellular OX2 and KSHV ORF74 (Pratt et al. 2005; Wang et al.
2009). Similar to K15, RK15 displays a complex splicing pattern with the most
common transcript encoding all the 12 transmembranes (Wang et al. 2009). While
RK15 does not have the SH2 YEEV motif that is important for the signaling activi-
ties of K15, it retains the SH3 motif, albeit at a different location (K15P–P387PLP,
K15M–P396PLP versus RK15–P492PLP) (Brinkmann et al. 2003; Wang et al. 2007,
2009). Nonetheless, RK15 can still activate the NF-κB and JNK pathways to a simi-
lar degree as K15, but only weakly activates the ERK cascade compared to K15.
This suggests that RK15 utilizes an SH2-independent mechanism to target NF-κB
and JNK activity. Furthermore this also implies that RK15 require an SH2 domain
to robustly activate the ERK pathway. A study by Wang et al. using microarray
analysis to compare the spectrum of cellular genes regulated by K15 and RK15
found that there was little overlap between the cellular genes regulated by these two
viral proteins (Wang et al. 2009). While both induced the expression of proinflam-
matory cytokines and chemokines such as IL-8 and IL-6, the extent of induction
was more robust in the presence of K15 compared to RK15. On the other hand,
RK15 preferentially upregulates cellular factors such as fibroblast growth factor 21
(FGF21), Kruppel-like factor 15 (KLF15), and complement component 5a receptor
1 (C5aR1), genes whose expression is unchanged in the presence of K15 (Wang
et al. 2009). This differential display of gene induction activities between K15 and
RK15 could be due to the inability of RK15 to activate the ERK pathway, leading to
236 L.-Y. Wong et al.
Herpesvirus Saimiri
In HVS subgroups A and B, the first ORF encodes a protein named STP A and B
(STP-A and STP-B), respectively (Fig. 3) (Murthy et al. 1989). In HVS subgroup C
(strain C488), however, STP-C is the second ORF from the left end of the genome,
Molecular Mimicry by γ-2 Herpesviruses to Modulate Host Cell Signaling Pathways 237
as the first ORF encodes a protein called tyrosine-interacting protein (Tip) (Fig. 3)
(Duboise et al. 1998a). STP-C and Tip are translated from a bicistronic mRNA;
thus, both genes are very likely to be expressed at the same time (Biesinger et al.
1995). The STPs and Tip are not required for viral replication but are essential for
the transformation of lymphocytes in vitro and the induction of lymphoma in vivo.
Expression of STP-A or STP-C alone in rat fibroblasts results in transformation of
the cells and tumor formation following inoculation in nude mice (Jung et al. 1991).
Even though the STPs are structurally similar, they do not share high sequence
homology, with STP-B being 28 % and 22 % identical to STP-A and STP-C, respec-
tively. STP-A and STP-C both have collagen-like repeats (glycine-X-Y with X or Y
being proline or glutamine) after a highly acidic amino terminus, while STP-B lacks
these repeats. STP-C has 18 of these repeats arranged in tandem, while STP-A only
has 9 repeats that are scattered throughout the protein. The importance of these
collagen-like repeats for STP oligomerization and ultimately the oncogenicity of
HVS has been shown in several studies. Indeed, a chimeric version of STP-B con-
taining the 18 copies of repeats from STP-C is now able to transform cells (Choi
et al. 2000a). Oligomerization of STP-C through its collagen repeats may mimic a
ligand-independent, constitutively active receptor, which then activates the NF-κB
pathway to induce expression of genes important for survival and transformation.
Not surprisingly, mutating the repeats in STP-C abolishes its oncogenic activity
(Choi et al. 2000a).
STP-C was the first reported virus-encoded protein to interact with cellular rat
sarcoma (Ras), a molecular switch for MAPK cascade important for oncogenesis
and cell growth (Jung and Desrosiers 1995). Expression of STP-C activates the Ras
signaling cascade, and disrupting the STP-C-Ras interaction interferes with the
transformation of cells by this viral oncogene (Fig. 4). In fact, a recombinant virus
that encodes Ras in lieu of STP-C can still transform lymphocytes, albeit with lower
efficiency compared to the wild-type virus (Jung and Desrosiers 1995). Thus, Ras is
an important player in the transformation of lymphocytes by STP-C.
STPs also contain other motifs involved in cellular signaling cascades, such as a
TRAF-interacting domain and an SH2 motif (in STP-A and STP-B only). TRAFs
are a family of cellular proteins involved in relaying signaling from initiation by the
external stimuli to downstream molecules and thus regulate diverse functions. The
TRAF-binding site of STP-C is required for its interaction with TRAF-1, TRAF-2,
and TRAF-3 and subsequent activation of NF-κB (Lee et al. 1999). Mutation of the
TRAF-binding motif abolishes the ability of STP-C to immortalize primary human
T cells in vitro. However, this STP-C TRAF-binding mutant protein can still trans-
form lymphocytes of common marmoset in vivo and in vitro, suggesting that while
host TRAFs are important for the oncogenicity of HVS STP-C, this multifunctional
protein can utilize other mechanisms to achieve the same result (Lee et al. 1999).
Residue Y115 in the SH2 motif of STP-A is responsible for binding to Src and is
itself a substrate for this tyrosine kinase (Chung et al. 2004; Garcia et al. 2007). This
interaction leads to strong activation of the transcription factors, AP-1 and NFAT,
while the binding of STP-A to TRAF-6 elicits a vigorous NF-kB response. The acti-
vation of all three pathways contributes to IL-2 promoter activity (Garcia et al. 2007).
238 L.-Y. Wong et al.
Tyrosine-Interacting Protein
Tip is encoded by the first ORF of HVS subgroup C (Fig. 3) (Biesinger et al. 1995).
Like all the viral proteins discussed so far, Tip also contains multiple motifs for
interacting with cellular proteins. First and foremost is the interaction of Tip with
Lck, the major Src tyrosine kinase in T cells (Biesinger et al. 1995). This binding is
mediated by the SH3 motif and the C-terminal Src-related kinase homology (CSKH)
domain of Tip (Jung et al. 1995a). The interaction between Tip and Lck results in
the phosphorylation of Tip by the kinase as well as deregulation of the downstream
T-cell receptor (TCR) signaling cascade, likely by sequestering Lck away from the
TCR (Cho et al. 2004).
In addition, Tip expression also decreases the cell surface level of TCR and CD4
molecules, which is dependent on the localization of Tip to the membrane rafts via
its transmembrane and an amphipathic helix motif just preceding the transmem-
brane domain (Cho et al. 2006; Min et al. 2008). Tip also interacts with a cellular
endosomal protein, p80, through its serine-rich (SR) domain (Park et al. 2002). This
interaction in turn mediates formation of large vesicles that help traffic the internal-
ized TCR and CD4 molecules to the lysosome for degradation (Park et al. 2003).
The functional consequences of Tip binding to Lck are controversial. In
certain contexts, expression of Tip suppresses the activation of Lck. For exam-
ple, expression of Tip in the human Jurkat T-cell line results in lowered levels of
tyrosine phosphorylation, while in fibroblasts Tip is able to inhibit the activity of
an oncogenic mutant of Lck (F505) (Jung et al. 1995b). On the other hand, infec-
tion of human peripheral blood T lymphocytes (PBL) with a Tip-deleted mutant
virus results in a significant decrease in Lck activation compared to wild-type
virus, strongly suggesting that the presence of Tip increases the activity of Lck
(Lund et al. 1997). The enzymatic activity of Lck is also enhanced by Tip in an
in vitro assay (Kjellen et al. 2002; Lund et al. 1997; Wiese et al. 1996). Duboise
et al. mutated the Lck-binding SH3 motif (proline to alanine mutations) of Tip
in the context of the virus and surprisingly found that this mutant virus can still
immortalize common marmoset lymphocytes in vivo and in vitro (Duboise et al.
1998b). This data suggests that deregulation of Lck by Tip is not required for its
Molecular Mimicry by γ-2 Herpesviruses to Modulate Host Cell Signaling Pathways 239
ability to immortalize. Interestingly, this mutant virus was even more pathogenic
compared to the wild-type virus, with increased lymphoid infiltration upon nec-
ropsy of the infected animals (Duboise et al. 1998b). However, it was reported
that mutating just one of the dual Lck-interacting sites in Tip does not com-
pletely abrogate the interaction between Tip and Lck (Heck et al. 2006). In sum-
mary, the oncogenic ability of HVS and the modulation of Tip on Lck are
complex and multifaceted.
Another function of Tip is its activation of several transcription factors including
NF-κB, NFAT, STAT1, and STAT3 (Merlo and Tsygankov 2001). The modulation
of STATs by Tip is Lck dependent, but STAT activation is dispensable for the trans-
formative function of Tip (Heck et al. 2005).
The studies on these six ORFs from KSHV, RRV, and HVS suggest conservation of
functions despite the lack of sequence homology and divergence of cell tropism. K1
and R1 both activate pathways that are also induced by BCR activation and both
have transforming potentials. In addition, the abilities of K1 and R1 to replace STP
in recombinant HVS and still immortalize T cells from common marmosets suggest
that these three viral proteins have a conserved capability to activate the same path-
ways leading to cell survival and transformation (Damania et al. 1999; Lee et al.
1998b). Comparison of K15 and RK15 revealed a weaker propensity for RK15 to
activate inflammatory pathways, perhaps reflecting the more cytokine-dependent
nature of KSHV pathogenesis as compared with RRV. Nonetheless, all of these viral
proteins activate cellular pathways important for cell survival. The in vivo contribu-
tion of K1, K15, R1, and RK15 to virus-induced tumorigenesis is still unknown and
hopefully will be elucidated in the near future.
γ-2-Herpesvirus miRNAs
KSHV encodes 12 pre-miRNAs clustered between ORF K12 and ORF71, an area
termed the latency-associated region (Fig. 3) (Cai et al. 2005; Grundhoff et al. 2006;
Pfeffer et al. 2005; Samols et al. 2005). From these 12 pre-miRNAs, 17–25 mature
miRNAs can be predicted to form following maturation (Abend et al. 2010; Lin
et al. 2010, 2011). As these miRNAs are located within the major latency-associated
region of the viral genome, it is not surprising that these miRNAs are highly enriched
in latent PEL cells (Cai et al. 2005; Samols et al. 2005). With the exception of miR-
K10 and miR-K12, all other miRNAs levels remain constant during lytic replica-
tion, indicating that KSHV miRNAs primarily function during viral latency (Cai
et al. 2005; Grundhoff et al. 2006).
KSHV miRNAs have been shown to target both viral and cellular genes, exerting
control over both the viral life cycle and host immune responses. miR-K12-11 is an
ortholog of cellular miR-155, an important regulator of lymphocyte differentiation
and innate immunity (Faraoni et al. 2009; Gottwein et al. 2007; Skalsky et al. 2007).
miR-155 is a potent antiapoptotic miRNA when expressed in breast cancer cells and
has been implicated in tumorigenesis of lymphoid and myeloid cancers (Ovcharenko
et al. 2007). PEL cells lack the expression of miR-155, but express high levels of
miR-K12-11 (Skalsky et al. 2007). Since it has an identical seed sequence and regu-
lates the same set of genes as miR-155, miR-K12-11 may contribute to the KSHV-
induced oncogenesis (Gottwein et al. 2007). A recent study (2011) showed that
miR-K12-11 is indeed the viral mimic of the cellular miR-155 (Boss et al. 2011).
Ectopic expression of either miRNAs in humanized NSG (NOD-scid interleukin-2
receptor γ-chain null (IL2Rγnull)) mouse model led to hyperproliferation of human
B cells in the spleen. This phenotype could be traced to the dysregulation of CCAAT
enhancer-binding protein β (C/EBPβ), a regulator of IL-6 transcription, by these
miRNAs (Boss et al. 2011). Since IL-6 is a key driver of inflammation and prolifera-
tion of lymphocytes, activation of IL-6 by miR-K12-11 is likely to be an important
aspect of KSHV pathogenesis.
The example of miR-K1, which targets multiple signaling pathways, highlights
the versatility of miRNA-mediated regulation of gene expression. First, miR-K1 has
been shown to be an important regulator of latency by positively regulating the
NF-κB pathway through reducing the level of the inhibitor IκBα to maintain the
viral latency state (Lei et al. 2010). Furthermore, by manipulating this key pathway,
Molecular Mimicry by γ-2 Herpesviruses to Modulate Host Cell Signaling Pathways 241
KSHV can suppress host immunity and enhance cell survival. Second, miR-K1
specifically prevents the expression of p21, an important regulator of cell cycle
(Gottwein and Cullen 2010).
Lytic replication is critical for gammaherpesviruses to maintain a disseminated
latent infection in vivo but has unwanted consequences such as immune activation
and cell death to release infectious virion. Thus, tight control of lytic reactivation
may allow gammaherpesviruses to maintain persistent infection in the presence of
host immune surveillance. Rta, encoded by ORF50, is the master transcriptional
regulator of lytic replication in KSHV and is directly targeted by at least two viral
miRNAs (miR-K9, miR-K12-7) (Bellare and Ganem 2009; Lin et al. 2011; Lu et al.
2010b). As a consequence of miR-K3 targeting the nuclear factor I/B (NFIB), a
transcriptional activator for the Rta promoter, miR-K3 also has a role in repressing
expression of Rta and thus contributes to the maintenance of latency (Lu et al.
2010a). The number of viral miRNAs devoted to targeting Rta, either directly or
indirectly through inhibition of other mRNAs, suggests that KSHV tightly main-
tains latency to prevent inappropriate induction of lytic replication.
Sequencing of KSHV miRNAs from PEL cell lines and tissue samples from
patients with different KSHV-induced diseases revealed high conservation among
most of the viral miRNAs, suggesting the importance of miRNAs in KSHV biology
(Marshall et al. 2007). Interestingly, a few miRNAs (e.g., miR-K12) showed poly-
morphisms between the different isolates (Marshall et al. 2007). Since it has been
shown that a single nucleotide polymorphism (SNP) in a viral miRNA could affect
its maturation and function, it is tempting to speculate that these “minor” changes in
viral miRNA sequences may impact the virus biology in a significant manner
(Gottwein et al. 2006). These viral miRNA SNPs could reflect a relatively rapid
adaptation strategy by the virus to outwit the host response.
RRV also encodes 15 miRNAs processed from 7 pre-miRNAs located in the
same genomic region as the KSHV miRNAs (Fig. 3) (Schafer et al. 2007; Umbach
et al. 2010). Umbach et al. (2010) used a deep sequencing assay to analyze the
expression of viral miRNAs in RRV-induced tumors. They showed that all but one
RRV miRNA lack sequence homology with KSHV miRNAs. The targets and func-
tions of RRV miRNAs have yet to be characterized. The observation that the miR-
NAs of KSHV and RRV are located in the same viral genomic region but do not
share sequence homology suggests that these viruses may have adapted the use of
miRNA before evolutionary diverging, but have since altered the seed sequence to
target their respective host.
Analysis of the HVS genome failed to predict with sufficient certainty the pres-
ence of miRNAs (Walz et al. 2010). Several potential miRNA candidates are pre-
dicted to be encoded at the location where miRNAs are found in the KSHV and
RRV genomes, but have yet to be verified (Walz et al. 2010). HVS, however, does
encode seven viral U-rich RNAs called HSURs in its 3′ terminal L-DNA region
(Fig. 3) (Biesinger et al. 1990; Lee et al. 1988). These noncoding RNAs are not
required for viral replication in cell culture (Ensser et al. 1999). Steiz and colleagues
found that HSUR-1 and HSUR-2 bind to cellular miR-27, miR-142-3p, and miR-
16, leading to degradation of miR-27 while not affecting the other two miRNAs
242 L.-Y. Wong et al.
Conclusion
References
Abend JR, Uldrick T, Ziegelbauer JM (2010) Regulation of tumor necrosis factor-like weak
inducer of apoptosis receptor protein (TWEAKR) expression by Kaposi’s sarcoma-associated
herpesvirus microRNA prevents TWEAK-induced apoptosis and inflammatory cytokine
expression. J Virol 84:12139–12151
Ablashi DV, Schirm S, Fleckenstein B, Faggioni A, Dahlberg J, Rabin H, Loeb W, Armstrong G,
Peng JW, Aulahk G et al (1985) Herpesvirus saimiri-induced lymphoblastoid rabbit cell line:
growth characteristics, virus persistence, and oncogenic properties. J Virol 55:623–633
Molecular Mimicry by γ-2 Herpesviruses to Modulate Host Cell Signaling Pathways 243
Alexander L, Denekamp L, Knapp A, Auerbach MR, Damania B, Desrosiers RC (2000) The pri-
mary sequence of rhesus monkey rhadinovirus isolate 26–95: sequence similarities to Kaposi’s
sarcoma-associated herpesvirus and rhesus monkey rhadinovirus isolate 17577. J Virol
74:3388–3398
Aluigi MG, Albini A, Carlone S, Repetto L, De Marchi R, Icardi A, Moro M, Noonan D, Benelli
R (1996) KSHV sequences in biopsies and cultured spindle cells of epidemic, iatrogenic and
Mediterranean forms of Kaposi’s sarcoma. Res Virol 147:267–275
Ambroziak JA, Blackbourn DJ, Herndier BG, Glogau RG, Gullett JH, McDonald AR, Lennette
ET, Levy JA (1995) Herpes-like sequences in HIV-infected and uninfected Kaposi’s sarcoma
patients. Science 268:582–583
Ballon G, Chen K, Perez R, Tam W, Cesarman E (2011) Kaposi sarcoma herpesvirus (KSHV)
vFLIP oncoprotein induces B cell transdifferentiation and tumorigenesis in mice. J Clin Invest
121:1141–1153
Bellare P, Ganem D (2009) Regulation of KSHV lytic switch protein expression by a virus-
encoded microRNA: an evolutionary adaptation that fine-tunes lytic reactivation. Cell Host
Microbe 6:570–575
Benelli R, Albini A, Parravicini C, Carlone S, Repetto L, Tambussi G, Lazzarin A (1996) Isolation
of spindle-shaped cell populations from primary cultures of Kaposi’s sarcoma of different
stage. Cancer Lett 100:125–132
Bergquam EP, Avery N, Shiigi SM, Axthelm MK, Wong SW (1999) Rhesus rhadinovirus estab-
lishes a latent infection in B lymphocytes in vivo. J Virol 73:7874–7876
Berkova Z, Wang S, Wise JF, Maeng H, Ji Y, Samaniego F (2009) Mechanism of Fas signaling
regulation by human herpesvirus 8 K1 oncoprotein. J Natl Cancer Inst 101:399–411
Biesinger B, Trimble JJ, Desrosiers RC, Fleckenstein B (1990) The divergence between two onco-
genic Herpesvirus saimiri strains in a genomic region related to the transforming phenotype.
Virology 176:505–514
Biesinger B, Muller-Fleckenstein I, Simmer B, Lang G, Wittmann S, Platzer E, Desrosiers RC,
Fleckenstein B (1992) Stable growth transformation of human T lymphocytes by herpesvirus
saimiri. Proc Natl Acad Sci U S A 89:3116–3119
Biesinger B, Tsygankov AY, Fickenscher H, Emmrich F, Fleckenstein B, Bolen JB, Broker BM
(1995) The product of the herpesvirus saimiri open reading frame 1 (tip) interacts with T cell-
specific kinase p56lck in transformed cells. J Biol Chem 270:4729–4734
Blaskovic D, Stancekova M, Svobodova J, Mistrikova J (1980) Isolation of five strains of herpes-
viruses from two species of free living small rodents. Acta Virol 24:468
Boshoff C, Schulz TF, Kennedy MM, Graham AK, Fisher C, Thomas A, McGee JO, Weiss RA,
O’Leary JJ (1995) Kaposi’s Sarcoma-associated herpesvirus infects endothelial and spindle
cells. Nat Med 1:1274–1278
Boss IW, Plaisance KB, Renne R (2009) Role of virus-encoded microRNAs in herpesvirus biol-
ogy. Trends Microbiol 17:544–553
Boss IW, Nadeau PE, Abbott JR, Yang Y, Mergia A, Renne R (2011) A Kaposi’s sarcoma-associated
herpesvirus-encoded ortholog of microRNA miR-155 induces human splenic B-cell expansion
in NOD/LtSz-scid IL2Rgammanull mice. J Virol 85:9877–9886
Brinkmann MM, Glenn M, Rainbow L, Kieser A, Henke-Gendo C, Schulz TF (2003) Activation
of mitogen-activated protein kinase and NF-kappaB pathways by a Kaposi’s sarcoma-
associated herpesvirus K15 membrane protein. J Virol 77:9346–9358
Brinkmann MM, Pietrek M, Dittrich-Breiholz O, Kracht M, Schulz TF (2007) Modulation of host
gene expression by the K15 protein of Kaposi’s sarcoma-associated herpesvirus. J Virol
81:42–58
Cai X, Lu S, Zhang Z, Gonzalez CM, Damania B, Cullen BR (2005) Kaposi’s sarcoma-associated
herpesvirus expresses an array of viral microRNAs in latently infected cells. Proc Natl Acad
Sci U S A 102:5570–5575
Cambier JC (1995) Antigen and Fc receptor signaling. The awesome power of the immunoreceptor
tyrosine-based activation motif (ITAM). J Immunol 155:3281–3285
244 L.-Y. Wong et al.
Diehl V, Henle G, Henle W, Kohn G (1968) Demonstration of a herpes group virus in cultures of
peripheral leukocytes from patients with infectious mononucleosis. J Virol 2:663–669
Dittmer D, Stoddart C, Renne R, Linquist-Stepps V, Moreno ME, Bare C, McCune JM, Ganem D
(1999) Experimental transmission of Kaposi’s sarcoma-associated herpesvirus (KSHV/HHV-
8) to SCID-hu Thy/Liv mice. J Exp Med 190:1857–1868
Dourmishev LA, Dourmishev AL, Palmeri D, Schwartz RA, Lukac DM (2003) Molecular genetics
of Kaposi’s sarcoma-associated herpesvirus (human herpesvirus-8) epidemiology and patho-
genesis. Microbiol Mol Biol Rev 67:175–212
Du MQ, Liu H, Diss TC, Ye H, Hamoudi RA, Dupin N, Meignin V, Oksenhendler E, Boshoff C,
Isaacson PG (2001) Kaposi sarcoma-associated herpesvirus infects monotypic (IgM lambda)
but polyclonal naive B cells in Castleman disease and associated lymphoproliferative disor-
ders. Blood 97:2130–2136
Duboise SM, Guo J, Czajak S, Desrosiers RC, Jung JU (1998a) STP and Tip are essential for her-
pesvirus saimiri oncogenicity. J Virol 72:1308–1313
Duboise SM, Lee H, Guo J, Choi JK, Czajak S, Simon M, Desrosiers RC, Jung JU (1998b)
Mutation of the Lck-binding motif of Tip enhances lymphoid cell activation by herpesvirus
saimiri. J Virol 72:2607–2614
Dupin N, Fisher C, Kellam P, Ariad S, Tulliez M, Franck N, van Marck E, Salmon D, Gorin I,
Escande JP, Weiss RA, Alitalo K, Boshoff C (1999) Distribution of human herpesvirus-8
latently infected cells in Kaposi’s sarcoma, multicentric Castleman’s disease, and primary effu-
sion lymphoma. Proc Natl Acad Sci U S A 96:4546–4551
Ensoli B, Nakamura S, Salahuddin SZ, Biberfeld P, Larsson L, Beaver B, Wong-Staal F, Gallo RC
(1989) AIDS-Kaposi’s sarcoma-derived cells express cytokines with autocrine and paracrine
growth effects. Science 243:223–226
Ensoli B, Sturzl M, Monini P (2000) Cytokine-mediated growth promotion of Kaposi’s sarcoma
and primary effusion lymphoma. Semin Cancer Biol 10:367–381
Ensser A, Pfinder A, Muller-Fleckenstein I, Fleckenstein B (1999) The URNA genes of herpesvi-
rus saimiri (strain C488) are dispensable for transformation of human T cells in vitro. J Virol
73:10551–10555
Falk LA, Wolfe LG, Deinhardt F (1972) Isolation of Herpesvirus saimiri from blood of squirrel
monkeys (Saimiri sciureus). J Natl Cancer Inst 48:1499–1505
Faraoni I, Antonetti FR, Cardone J, Bonmassar E (2009) miR-155 gene: a typical multifunctional
microRNA. Biochim Biophys Acta 1792:497–505
Fleckenstein B, Ensser A (2007) Gammaherpesviruses of New World primates. Human herpesvi-
ruses: biology, therapy, and immunoprophylaxis. Cambridge: Cambridge University Press.
Foreman KE, Bacon PE, Hsi ED, Nickoloff BJ (1997) In situ polymerase chain reaction-based
localization studies support role of human herpesvirus-8 as the cause of two AIDS-related
neoplasms: Kaposi’s sarcoma and body cavity lymphoma. J Clin Invest 99:2971–2978
Friedman-Kien AE (1981) Disseminated Kaposi’s sarcoma syndrome in young homosexual men.
J Am Acad Dermatol 5:468–471
Ganem D (2006) KSHV infection and the pathogenesis of Kaposi’s sarcoma. Annu Rev Pathol
1:273–296
Gao SJ, Kingsley L, Li M, Zheng W, Parravicini C, Ziegler J, Newton R, Rinaldo CR, Saah A,
Phair J, Detels R, Chang Y, Moore PS (1996) KSHV antibodies among Americans, Italians and
Ugandans with and without Kaposi’s sarcoma. Nat Med 2:925–928
Garcia MI, Kaserman J, Chung YH, Jung JU, Lee SH (2007) Herpesvirus saimiri STP-A oncopro-
tein utilizes Src family protein tyrosine kinase and tumor necrosis factor receptor-associated
factors to elicit cellular signal transduction. J Virol 81:2663–2674
Garzon R, Calin GA, Croce CM (2009) MicroRNAs in cancer. Annu Rev Med 60:167–179
Glenn M, Rainbow L, Aurade F, Davison A, Schulz TF (1999) Identification of a spliced gene from
Kaposi’s sarcoma-associated herpesvirus encoding a protein with similarities to latent mem-
brane proteins 1 and 2A of Epstein-Barr virus. J Virol 73:6953–6963
246 L.-Y. Wong et al.
Kaposi M (1872) Idiopathic multiple pigmented sarcoma of the skin. Archiv fur Dertmatologie
und Syphilis 4:265–273
Kasolo FC, Mpabalwani E, Gompels UA (1997) Infection with AIDS-related herpesviruses in
human immunodeficiency virus-negative infants and endemic childhood Kaposi’s sarcoma in
Africa. J Gen Virol 78(Pt 4):847–855
Kedes DH, Operskalski E, Busch M, Kohn R, Flood J, Ganem D (1996) The seroepidemiology of
human herpesvirus 8 (Kaposi’s sarcoma-associated herpesvirus): distribution of infection in
KS risk groups and evidence for sexual transmission. Nat Med 2:918–924
Kjellen P, Amdjadi K, Lund TC, Medveczky PG, Sefton BM (2002) The herpesvirus saimiri tip484
and tip488 proteins both stimulate lck tyrosine protein kinase activity in vivo and in vitro.
Virology 297:281–288
Knipe DM, Howley PM (eds) (2007) Field virology. Lippincott Williams & Wilkins, Philadelphia,
PA
Koomey JM, Mulder C, Burghoff RL, Fleckenstein B, Desrosiers RC (1984) Deletion of DNA
sequence in a nononcogenic variant of Herpesvirus saimiri. J Virol 50:662–665
Krol J, Loedige I, Filipowicz W (2010) The widespread regulation of microRNA biogenesis, func-
tion and decay. Nat Rev Genet 11:597–610
Lagunoff M, Ganem D (1997) The structure and coding organization of the genomic termini of
Kaposi’s sarcoma-associated herpesvirus. Virology 236:147–154
Lagunoff M, Majeti R, Weiss A, Ganem D (1999) Deregulated signal transduction by the K1 gene
product of Kaposi’s sarcoma-associated herpesvirus. Proc Natl Acad Sci USA 96:
5704–5709
Lagunoff M, Lukac DM, Ganem D (2001) Immunoreceptor tyrosine-based activation motif-
dependent signaling by Kaposi’s sarcoma-associated herpesvirus K1 protein: effects on lytic
viral replication. J Virol 75:5891–5898
Lagunoff M, Bechtel J, Venetsanakos E, Roy AM, Abbey N, Herndier B, McMahon M, Ganem D
(2002) De novo infection and serial transmission of Kaposi’s sarcoma-associated herpesvirus
in cultured endothelial cells. J Virol 76:2440–2448
Lee SI, Murthy SC, Trimble JJ, Desrosiers RC, Steitz JA (1988) Four novel U RNAs are encoded
by a herpesvirus. Cell 54:599–607
Lee H, Guo J, Li M, Choi JK, DeMaria M, Rosenzweig M, Jung JU (1998a) Identification of an
immunoreceptor tyrosine-based activation motif of K1 transforming protein of Kaposi’s
sarcoma-associated herpesvirus. Mol Cell Biol 18:5219–5228
Lee H, Veazey R, Williams K, Li M, Guo J, Neipel F, Fleckenstein B, Lackner A, Desrosiers RC,
Jung JU (1998b) Deregulation of cell growth by the K1 gene of Kaposi’s sarcoma-associated
herpesvirus. Nat Med 4:435–440
Lee H, Choi JK, Li M, Kaye K, Kieff E, Jung JU (1999) Role of cellular tumor necrosis factor
receptor-associated factors in NF-kappaB activation and lymphocyte transformation by herpes-
virus Saimiri STP. J Virol 73:3913–3919
Lee BS, Alvarez X, Ishido S, Lackner AA, Jung JU (2000) Inhibition of intracellular transport of
B cell antigen receptor complexes by Kaposi’s sarcoma-associated herpesvirus K1. J Exp Med
192:11–21
Lee BS, Connole M, Tang Z, Harris NL, Jung JU (2003) Structural analysis of the Kaposi’s
sarcoma-associated herpesvirus K1 protein. J Virol 77:8072–8086
Lee BS, Lee SH, Feng P, Chang H, Cho NH, Jung JU (2005) Characterization of the Kaposi’s
sarcoma-associated herpesvirus K1 signalosome. J Virol 79:12173–12184
Lee JS, Li Q, Lee JY, Lee SH, Jeong JH, Lee HR, Chang H, Zhou FC, Gao SJ, Liang C,
Jung JU (2009) FLIP-mediated autophagy regulation in cell death control. Nat Cell Biol
11:1355–1362
Lei X, Bai Z, Ye F, Xie J, Kim CG, Huang Y, Gao SJ (2010) Regulation of NF-kappaB inhibitor
IkappaBalpha and viral replication by a KSHV microRNA. Nat Cell Biol 12:193–199
Li M, Lee H, Yoon DW, Albrecht JC, Fleckenstein B, Neipel F, Jung JU (1997) Kaposi’s sarcoma-
associated herpesvirus encodes a functional cyclin. J Virol 71:1984–1991
248 L.-Y. Wong et al.
Lim CS, Seet BT, Ingham RJ, Gish G, Matskova L, Winberg G, Ernberg I, Pawson T (2007) The
K15 protein of Kaposi’s sarcoma-associated herpesvirus recruits the endocytic regulator inter-
sectin 2 through a selective SH3 domain interaction. Biochemistry 46:9874–9885
Lin YT, Kincaid RP, Arasappan D, Dowd SE, Hunicke-Smith SP, Sullivan CS (2010) Small RNA
profiling reveals antisense transcription throughout the KSHV genome and novel small RNAs.
RNA 16:1540–1558
Lin X, Liang D, He Z, Deng Q, Robertson ES, and Lan K (2011) miR-K12-7-5p encoded by
Kaposi’s sarcoma-associated herpesvirus stabilizes the latent state by targeting viral ORF50/
RTA. PLoS One 6:e16224
Lu CC, Li Z, Chu CY, Feng J, Feng J, Sun R, Rana TM (2010a) MicroRNAs encoded by Kaposi’s
sarcoma-associated herpesvirus regulate viral life cycle. EMBO Rep 11:784–790
Lu F, Stedman W, Yousef M, Renne R, Lieberman PM (2010b) Epigenetic regulation of Kaposi’s
sarcoma-associated herpesvirus latency by virus-encoded microRNAs that target Rta and the
cellular Rbl2-DNMT pathway. J Virol 84:2697–2706
Lund T, Medveczky MM, Medveczky PG (1997) Herpesvirus saimiri Tip-484 membrane protein
markedly increases p56lck activity in T cells. J Virol 71:378–382
Mansfield KG, Westmoreland SV, DeBakker CD, Czajak S, Lackner AA, Desrosiers RC (1999)
Experimental infection of rhesus and pig-tailed macaques with macaque rhadinoviruses. J
Virol 73:10320–10328
Marshall V, Parks T, Bagni R, Wang CD, Samols MA, Hu J, Wyvil KM, Aleman K, Little RF,
Yarchoan R, Renne R, Whitby D (2007) Conservation of virally encoded microRNAs in Kaposi
sarcoma–associated herpesvirus in primary effusion lymphoma cell lines and in patients with
Kaposi sarcoma or multicentric Castleman disease. J Infect Dis 195:645–659
McAllister SC, Moses AV (2007) Endothelial cell- and lymphocyte-based in vitro systems for
understanding KSHV biology. Curr Top Microbiol Immunol 312:211–244
Medveczky P, Szomolanyi E, Desrosiers RC, Mulder C (1984) Classification of herpesvirus sai-
miri into three groups based on extreme variation in a DNA region required for oncogenicity. J
Virol 52:938–944
Melendez LV, Daniel MD, Hunt RD, Garcia FG (1968) An apparently new herpesvirus from pri-
mary kidney cultures of the squirrel monkey (Saimiri sciureus). Lab Anim Care 18:374–381
Melendez LV, Daniel MD, Garcia FG, Fraser CE, Hunt RD, King NW (1969) Herpesvirus saimiri.
I. Further characterization studies of a new virus from the squirrel monkey. Lab Anim Care
19:372–377
Merlo JJ, Tsygankov AY (2001) Herpesvirus saimiri oncoproteins Tip and StpC synergistically
stimulate NF-kappaB activity and interleukin-2 gene expression. Virology 279:325–338
Min CK, Bang SY, Cho BA, Choi YH, Yang JS, Lee SH, Seong SY, Kim KW, Kim S, Jung JU,
Choi MS, Kim IS, Cho NH (2008) Role of amphipathic helix of a herpesviral protein in mem-
brane deformation and T cell receptor downregulation. PLoS Pathog 4:e1000209
Moore PS, Gao SJ, Dominguez G, Cesarman E, Lungu O, Knowles DM, Garber R, Pellett PE,
McGeoch DJ, Chang Y (1996) Primary characterization of a herpesvirus agent associated with
Kaposi’s sarcomae. J Virol 70:549–558
Murthy SC, Trimble JJ, Desrosiers RC (1989) Deletion mutants of herpesvirus saimiri define an
open reading frame necessary for transformation. J Virol 63:3307–3314
Neipel F, Albrecht JC, Fleckenstein B (1997) Cell-homologous genes in the Kaposi’s sarcoma-
associated rhadinovirus human herpesvirus 8: determinants of its pathogenicity? J Virol
71:4187–4192
Niemi M, Mustakallio KK (1965) The fine structure of the spindle cell in Kaposi’s sarcoma. Acta
Pathol Microbiol Scand 63:567–575
Ovcharenko D, Kelnar K, Johnson C, Leng N, Brown D (2007) Genome-scale microRNA and
small interfering RNA screens identify small RNA modulators of TRAIL-induced apoptosis
pathway. Cancer Res 67:10782–10788
Park J, Lee BS, Choi JK, Means RE, Choe J, Jung JU (2002) Herpesviral protein targets a cellular WD
repeat endosomal protein to downregulate T lymphocyte receptor expression. Immunity 17:221–233
Molecular Mimicry by γ-2 Herpesviruses to Modulate Host Cell Signaling Pathways 249
Park J, Cho NH, Choi JK, Feng P, Choe J, Jung JU (2003) Distinct roles of cellular Lck and p80
proteins in herpesvirus saimiri Tip function on lipid rafts. J Virol 77:9041–9051
Pellett PE, Roizman B (2007) Fields virology. Lippincott-Raven, Philadelphia, PA
Pfeffer S, Sewer A, Lagos-Quintana M, Sheridan R, Sander C, Grasser FA, van Dyk LF, Ho CK,
Shuman S, Chien M, Russo JJ, Ju J, Randall G, Lindenbach BD, Rice CM, Simon V, Ho DD,
Zavolan M, Tuschl T (2005) Identification of microRNAs of the herpesvirus family. Nat
Methods 2:269–276
Picchio GR, Sabbe RE, Gulizia RJ, McGrath M, Herndier BG, Mosier DE (1997) The KSHV/
HHV8-infected BCBL-1 lymphoma line causes tumors in SCID mice but fails to transmit virus
to a human peripheral blood mononuclear cell graft. Virology 238:22–29
Pietrek M, Brinkmann MM, Glowacka I, Enlund A, Havemeier A, Dittrich-Breiholz O, Kracht M,
Lewitzky M, Saksela K, Feller SM, Schulz TF (2010) Role of the Kaposi’s sarcoma-associated
herpesvirus K15 SH3 binding site in inflammatory signaling and B-cell activation. J Virol
84:8231–8240
Poole LJ, Zong JC, Ciufo DM, Alcendor DJ, Cannon JS, Ambinder R, Orenstein JM, Reitz MS,
Hayward GS (1999) Comparison of genetic variability at multiple loci across the genomes of
the major subtypes of Kaposi’s sarcoma-associated herpesvirus reveals evidence for recombi-
nation and for two distinct types of open reading frame K15 alleles at the right-hand end. J
Virol 73:6646–6660
Prakash O, Tang ZY, Peng X, Coleman R, Gill J, Farr G, Samaniego F (2002) Tumorigenesis and
aberrant signaling in transgenic mice expressing the human herpesvirus-8 K1 gene. J Natl
Cancer Inst 94:926–935
Prakash O, Swamy OR, Peng X, Tang ZY, Li L, Larson JE, Cohen JC, Gill J, Farr G, Wang S,
Samaniego F (2005) Activation of Src kinase Lyn by the Kaposi sarcoma-associated herpesvi-
rus K1 protein: implications for lymphomagenesis. Blood 105:3987–3994
Pratt CL, Estep RD, Wong SW (2005) Splicing of rhesus rhadinovirus R15 and ORF74 bicistronic
transcripts during lytic infection and analysis of effects on production of vCD200 and vGPCR.
J Virol 79:3878–3882
Regamey N, Tamm M, Wernli M, Witschi A, Thiel G, Cathomas G, Erb P (1998) Transmission of
human herpesvirus 8 infection from renal-transplant donors to recipients. N Engl J Med
339:1358–1363
Renne R, Blackbourn D, Whitby D, Levy J, Ganem D (1998) Limited transmission of Kaposi’s
sarcoma-associated herpesvirus in cultured cells. J Virol 72:5182–5188
Roizmann B, Desrosiers RC, Fleckenstein B, Lopez C, Minson AC, Studdert MJ (1992) The fam-
ily Herpesviridae: an update. The Herpesvirus Study Group of the International Committee on
Taxonomy of Viruses. Arch Virol 123:425–449
Russo JJ, Bohenzky RA, Chien MC, Chen J, Yan M, Maddalena D, Parry JP, Peruzzi D, Edelman
IS, Chang Y, Moore PS (1996) Nucleotide sequence of the Kaposi sarcoma-associated herpes-
virus (HHV8). Proc Natl Acad Sci U S A 93:14862–14867
Salahuddin SZ, Nakamura S, Biberfeld P, Kaplan MH, Markham PD, Larsson L, Gallo RC (1988)
Angiogenic properties of Kaposi’s sarcoma-derived cells after long-term culture in vitro.
Science 242:430–433
Samaniego F, Pati S, Karp JE, Prakash O, and Bose D (2001) Human herpesvirus 8 K1-associated
nuclear factor-kappa B-dependent promoter activity: role in Kaposi’s sarcoma inflammation? J
Natl Cancer Inst Monogr (28):15–23
Samols MA, Hu J, Skalsky RL, Renne R (2005) Cloning and identification of a microRNA cluster
within the latency-associated region of Kaposi’s sarcoma-associated herpesvirus. J Virol
79:9301–9305
Schafer A, Cai X, Bilello JP, Desrosiers RC, Cullen BR (2007) Cloning and analysis of microR-
NAs encoded by the primate gamma-herpesvirus rhesus monkey rhadinovirus. Virology
364:21–27
Schalling M, Ekman M, Kaaya EE, Linde A, Biberfeld P (1995) A role for a new herpes virus
(KSHV) in different forms of Kaposi’s sarcoma. Nat Med 1:707–708
250 L.-Y. Wong et al.
Wong SW, Bergquam EP, Swanson RM, Lee FW, Shiigi SM, Avery NA, Fanton JW, Axthelm MK
(1999) Induction of B cell hyperplasia in simian immunodeficiency virus-infected rhesus
macaques with the simian homologue of Kaposi’s sarcoma-associated herpesvirus. J Exp Med
190:827–840
Wu W, Vieira J, Fiore N, Banerjee P, Sieburg M, Rochford R, Harrington W Jr, Feuer G (2006)
KSHV/HHV-8 infection of human hematopoietic progenitor (CD34+) cells: persistence of
infection during hematopoiesis in vitro and in vivo. Blood 108:141–151
Zhong W, Wang H, Herndier B, Ganem D (1996) Restricted expression of Kaposi sarcoma-
associated herpesvirus (human herpesvirus 8) genes in Kaposi sarcoma. Proc Natl Acad Sci
USA 93:6641–6646
Neotropical Primates and Their Susceptibility
to Toxoplasma gondii: New Insights for an Old
Problem
Introduction
J.F. Brinkworth and K. Pechenkina (eds.), Primates, Pathogens, and Evolution, 253
Developments in Primatology: Progress and Prospects,
DOI 10.1007/978-1-4614-7181-3_9, © Springer Science+Business Media New York 2013
254 J.L. Catão-Dias et al.
New World primates (NWPs, Platyrrhini) are distributed from South and Central
America to North America, in Mexico. Table 1 presents a distribution of NWPs
according to their diet and forest strata preferences. They are all small- to medium-
sized primates, weighing from 120 g to approximately 10 kg. The different species
live in sympatry throughout most of their ranges with up to 13 species at some
Amazonian sites. One of the characteristics of Platyrrhini is the absence of ter-
restrial species, and few species occasionally forage on the ground or travel short
distances between trees, but none spend most of the day feeding on the ground
(Feagle 1999). New World primates include species that feed on gums, fruits,
leaves, seeds, a variety of invertebrates, and small vertebrates. Some of the smaller
species rely heavily on nectar during the dry periods of the year. Some NWPs
show morphological adaptations for more specialized diets, while others are
omnivorous, and most species show preference for one of the arboreal strata,
where they stay most part of the time.
Perelman et al. (2011) divided Platyrrhini into three families and listed 17 gen-
era, but did not include the two monotypic genera: Callibella (Aguiar and Lacher
2003; Van Roosmalen and Van Roosmalen 2003) and Oreonax (Groves 2001),
although some authors do not separate Oreonax from Lagothrix (Rosenberger and
Matthews 2008). According to Rylands and Mittermeier (2009), there are 19 genera
and 199 species and subspecies of primates in the Neotropics.
Pitheciidae (Cacajao, uakaris; Calicebus, titis; Chiropotes, bearded sakis;
Pithecia, sakis) have an unusual dental specialization for processing fruits and seeds
encased in a hard outer covering that are generally too hard for other monkeys to
bite through (Kinzey 1992). Compared to Cacajao and Chiropotes, Pithecia seems
to eat fruits with relatively softer outer covering, has a more diverse diet, and con-
centrates its activity in the middle and upper canopy but can use the lower canopy
for foraging (Feagle 1999).
Chiropotes prefers high rain forests, being usually in the middle and upper levels
of the main canopy (Feagle 1999). They feed on hard, often unripe fruits and on
seeds with very hard shells, which they open with their large canines. They occa-
sionally feed on insects (Norconk et al. 2009). Cacajao are also specialized on fruits
with hard outer shells and immature seeds and include in their diet small amounts
Table 1 Diet and forest strata preferences by Neotropical primates
Species and
Family Genusa subspeciesb Common names Diet Forest strata preference References
Cebidae Mico 14 Amazonian Fruits, arthropods, Middle of the canopy to the Bicca-Marques et al. (2011),
marmosets flowers, exudates ground, lower levels of the Feagle (1999)
forest
Cebuella 2 Pygmy marmoset Exudates, arthro- Middle of the canopy to the Norconk et al. (2009), Feagle
pods, fruits, and ground, lower levels of the (1999), Soini (1988),
small vertebrates forest Bicca-Marques et al. (2011).
Callithrix 6 Marmosets Exudates, arthro- Middle of the canopy to the Norconk et al. (2009), Feagle
pods, fruits, ground, lower levels of the (1999), Bicca-Marques et al.
small verte- forest (2011)
brates, eggs,
seeds, molluscs
Callimico 1 Goeldi’s monkey Arthropods, fruits, Middle of the canopy to the Norconk et al. (2009), Feagle
fungi, exudates ground, lower levels of the (1999)
forest
Leontopithecus 4 Lion tamarins Fruits, arthropods, Middle of the canopy to the Norconk et al. (2009), Feagle
exudates, ground, lower levels of the (1999), Kierulff et al. (2002)
flowers, small forest
vertebrates
Saguinus 33 Tamarins Arthropods, fruits, Middle of the canopy to the Norconk et al. (2009), Feagle
exudates, young ground, lower levels of the (1999), Snowdon and Soini
leaves, small forest (1988), Bicca-Marques et al.
Neotropical Primates and Their Susceptibility to Toxoplasma gondii…
vertebrates (2011)
Aotus 12 Owl monkeys Fruits, young leaves, No preference for a particular Norconk et al. (2009), Feagle
flowers, canopy level (1999), Cunha (2008),
arthropods Bicca-Marques et al. (2011)
Saimiri 10 Squirrel monkeys Arthropods, fruits, Middle and lower levels of the Norconk et al. (2009), Feagle
young leaves, forest, eventually come down (1999), Defler (2005), Baldwin
flowers, seeds, to the ground. and Baldwin (1981), Boinski
small verte- (1987), Ingberman et al. (2008),
255
Table 1 (continued)
Species and
Family Genusa subspeciesb Common names Diet Forest strata preference References
Cebus 26 Capuchin monkeys Fruits, arthropods, Main canopy levels but Norconk et al. (2009), Izawa
young leaves, frequently come down to the (1979), Freese and
seeds, flowers, understory or to the ground Oppehheimer (1981), Fragaszy
small verte- during both travel and et al. (2004)
brates, eggs. In feeding
the wild, use
large stones as
tools to open
hard palm fruits
Atelidae Lagothrix 5 Woolly monkeys Fruits, insects, Upper levels of the main canopy, Norconk et al. (2009), Feagle
young leaves, rarely coming down to the (1999), Bicca-Marques et al.
flowers, seeds ground (2011)
Brachyteles 2 Muriquis Young leaves, fruits Upper levels of the main canopy, Norconk et al. (2009), Feagle
(pulp), flowers, rarely coming down to the (1999), Mendes et al. (2010)
mature leaves, ground
seeds
Ateles 15 Spider monkeys Fruits (pulp), young Highest levels of the forest but Norconk et al. (2009), Feagle
leaves, flowers eventually come down to the (1999), Zanon et al. (2008),
ground Bicca-Marques et al. (2011).
Alouatta 19 Howler monkeys Young leaves, fruits Most species prefer the main Norconk et al. (2009), Feagle
(pulp), mature canopy and emergent levels; (1999), Bicca-Marques et al.
leaves, flowers species from dry areas (2011)
regularly come down to the
ground
Pitheciidae Cacajao 6 Uakaris Seeds, fruits (pulp), Upper parts of the forest but Norconk et al. (2009), Rickli and
flowers eventually come down to the Reis (2008)
ground to forage
J.L. Catão-Dias et al.
Species and
Family Genusa subspeciesb Common names Diet Forest strata preference References
Chiropotes 5 Bearded saki Seeds, fruit (pulp), Middle and upper levels of the Norconk et al. (2009), Feagle
monkeys arthropods, main canopy, rarely coming (1999), Bicca-Marques et al.
flowers down to the ground (2011).
Pithecia 9 Saki monkeys Seeds, fruits (pulp, Middle and upper canopy, Norconk et al. (2009), Feagle
whole, arils), eventually come down to the (1999)
young leaves, ground to forage
arthropods, and
flowers
Callicebus 29 Titi monkeys Fruits (pulp), seeds, Main canopy to the understory Norconk et al. (2009), Bordignon
young leaves, and rarely coming down to et al. (2008), Feagle (1999),
flowers, the ground Bicca-Marques et al. (2011)
arthropods
a
Taxonomy of Perelman et al. (2011)
b
Number of species and subspecies according to Rylands and Mittermeier (2009)
Neotropical Primates and Their Susceptibility to Toxoplasma gondii…
257
258 J.L. Catão-Dias et al.
of leaves, flowers, nectar, and insects (Rickli and Reis 2008). They use more
frequently the upper parts of the forest but can eventually come down to the ground
to forage (Rickli and Reis 2008).
Calicebus have very short canine teeth in comparison with other species of the
family Pitheciidae. They are mainly frugivorous but can supplement their diet with
leaves, seeds, flowers, and insects (Bordignon et al. 2008; Norconk et al. 2009). The
different species are distributed in different habitats from mature forest to the dry
scrub forest (caatinga) or bamboo thickets, where they use the main canopy or the
understory or low levels in the forest (Feagle 1999; Bordignon et al. 2008).
The family Atelidae includes two predominantly folivorous genera (Alouatta,
howler monkey, and Brachyteles, muriqui) that supplement their diet with fruits, flow-
ers, and seeds and two frugivorous genera, the spider (Ateles) and woolly monkeys
(Lagothrix and Oreonax), that feed mainly on fruits but also leaves, buds, flowers, and
insects (Groves 2001; Zanon et al. 2008; Norconk et al. 2009; Bicca-Marques et al.
2011). All atelines have a long, prehensile tail, and spider monkeys, woolly monkeys,
and muriquis are largely restricted to high primary rain forests where they prefer the
upper levels of the main canopy, rarely coming down to the ground (Mendes et al.
2010; Feagle 1999). Alouatta are found in a variety of habitats, including primary and
secondary forest, dry deciduous forest, and habitats containing patches of relatively
low trees in open savannah. Most species seem to prefer the main canopy and emer-
gent levels, but some species that live in drier areas (A. caraya) regularly come down
to the ground and cross-open areas between patches of forest (Feagle 1999).
The family Cebidae is composed of four subfamilies: Cebinae that includes
capuchin monkeys (Cebus); Saimirinae, represented by squirrel monkeys (Saimiri);
Callitrichinae, the smallest and most distinctive New World primates, separated in
Goeldi’s monkey (Callimico), tamarins (Saguinus and Leontopithecus), and mar-
mosets (Mico, Callithrix, Cebuella, Callibella); and the subfamily Aotinae (owl
monkeys), the only nocturnal Platyrrhini (Fernandez-Duque 2006; Norconk et al.
2009; Perelman et al. 2011).
Aotus are found in a variety of forest habitats, and there are no indications that
they prefer any particular canopy level. They are primarily frugivorous with a diet
that is supplemented by flowers, leaves, insects, and occasionally small vertebrates
and eggs (Wright 1981; Feagle 1999; Cunha 2008). Callitrichines use different
types of forest but seem to be characterized by the ability to exploit marginal and
disturbed habitat, and their diet is composed of fruits, flower, arthropods (mainly
insects), exudates, fungus, and small vertebrates (lizards, birds, frogs, and small
rodents). All species spend most of the time in the middle levels of the forest and
forage for fruits and insects in the middle of the canopy to the ground in the lower
levels of the forest (Snowdon and Soini 1988; Soini 1988; Feagle 1999; Kierulff
et al. 2002; Bicca-Marques et al. 2011).
Capuchin monkeys (Cebus) and squirrel monkeys (Saimiri) are the two most
omnivorous Platyrrhini (Feagle 1999). Saimiri occupy a variety of rain forest habi-
tats but seem to prefer riverine and secondary forests, where they are commonly
found in the middle and lower levels (Feagle 1999; Defler 2005). They are frugivo-
res and insectivores and supplement their diet with leaves, seeds, small vertebrates,
nectar, and eggs (Baldwin and Baldwin 1981; Boinski 1987; Feagle 1999; Defler
Neotropical Primates and Their Susceptibility to Toxoplasma gondii… 259
2005; Ingberman et al. 2008). The insect component of their diets is the highest of
any non- Callitrichinae, but the differences between these genera lie in mandibular
and dental robusticity, which is much stronger in Cebus and could determine differ-
ences in their diets (Norconk et al. 2009).
Capuchin monkeys are divided into tufted (Cebus apella, C. macrocephalus, C.
libidinosus, C. nigritus, C. robustus, C. cay, C. flavius, C. xanthosternos) and
untufted (C. albifrons, C. olivaceus, C. kaapori) (Bicca-Marques et al. 2011)1.
Cebus are found in virtually all types of neotropical forest including humid and dry
forest, swamp forests, seasonally flooded forest, as well as more open vegetation
types in savannahs and caatingas, where rainfall is absent for 5–6 months each year
(Fragaszy et al. 2004; Bicca-Marques et al. 2011). They seem to prefer the main
canopy levels but frequently come down to the understory or to the ground during
both travel and feeding (Feagle 1999).
Their diet includes many types of fruits and other vegetal parts and animal matter
(invertebrates and small vertebrates), and they are considered omnivores and also
classified as frugivore–insectivore (Izawa 1979; Freese and Oppehheimer 1981;
Feagle 1999). Fragaszy et al. (2004) characterized them as innovative and extreme
foragers due to their ability to acquire sustenance from a variety of potentially dan-
gerous sources that require special foraging skills and also for trying to eat almost
anything remotely edible. It gives them three types of adaptive advantages: first, the
flexibility to switch from more accessible foods such as fruits to more inaccessible
ones at time of food scarcity; second, the capacity to exploit habitats with different
structure and phenological characteristics (such as secondary or disturbed forests);
and third, it reduces the degree of dietary overlap between capuchins and other
arboreal vertebrates, mainly other primate species, more specialized in fruit or
insects (Brown and Zunino 1990; Fragaszy et al. 1990, 2004).
Toxoplasma gondii
1
Recent phylogeographic analysis has shown that capuchins contain two well-supported mono-
phyletic clades, the morphologically distinct “gracile” (or untufted) and “robust” (or tufted)
groups, and placed the age of the split at 6.7 Ma (95 % highest posterior density 4.1–9.4 Ma)
(Alfaro et al. 2012a). Morphological and behavioral–ecological data also support a division of
capuchins into the same two distinct groups. As a consequence Alfaro et al. (2012b) have argued
for a division of capuchin monkeys into two genera: Sapajus Kerr, 1792, for the robust capuchins
and Cebus Erxleben, 1777, for the gracile capuchins.
260 J.L. Catão-Dias et al.
The life cycle of T. gondii was elucidated throughout the 1960s and 1970s of the
twentieth century with the identification of the parasite’s sexual cycle (Frenkel et al.
1969, 1970; Hutchison et al. 1970; Ferguson et al. 1974). Figure 1 summarizes the
T. gondii life cycle (Gardiner et al. 1998). Three infective forms are recognized:
sporozoites in sporulated oocysts and asexual fast- and slow-replicating forms
tachyzoites and bradyzoites. Felines are the only known natural hosts and become
infected through meat consumption or intake of water and/or other food contami-
nated with infective forms. However, there is a marked difference in competence
Neotropical Primates and Their Susceptibility to Toxoplasma gondii… 261
Fig. 1 Toxoplasma gondii life cycle (Adapted from Gardiner et al. 1998)
between the various types of zoites in terms of how effectively they infect suscep-
tible felines. In this sense, bradyzoites are exceptionally efficient when compared to
tachyzoites and sporozoites. The ingestion of only one bradyzoite is sufficient
to cause infection in felines, while the intake of at least 1,000 oocysts is required to
produce the same effect (Dubey 2009). Besides the domestic cat (Felis catus), 17
species of wild felines, including five New World species (Puma concolor, cougar;
Puma yagouaroundi, jaguarundi; Leopardus geoffroyi, Geoffroy’s cat; L. pardalis,
ocelot; and L. colocolo, pampas cat) have been identified as natural hosts (Silva
2007; Elmore et al. 2010).
Once ingested by cats, the infective form invades the enterocyte of the small
intestine and undergoes asexual reproduction cycles, followed by sexual reproduc-
tion with the formation of non-sporulated oocyst with eight sporozoites. These
oocysts are excreted in feces and, depending on environmental conditions, undergo
sporulation and become infective. Fecal excretion of non-sporulated oocysts by
immunocompetent felines occurs 1–2 weeks after infection and, once established,
generates large amounts of oocysts. However, recently published data showed that
the incidence of immunosuppressive events can promote new episodes of oocyst
excretion in affected animals (Malmasi et al. 2009). Sporulated oocysts are very
resistant to environmental conditions such as desiccation and freezing, and to the
action of disinfectants, and can survive in the environment for many months. It is
262 J.L. Catão-Dias et al.
believed that the favorable conditions of humidity and temperature, as those found
in several tropical biomes, such as Amazon and Atlantic forest, greatly improve the
viability of sporulated oocysts (Silva 2007).
The infection of intermediate hosts may occur in many ways, either through meat
consumption/ingestion of prey infected by asexual zoites, the consumption of other
food and water contaminated with sporulated oocysts, or congenitally (Dubey et al.
1998). However, there are reports of infection caused by blood transfusions, trans-
plantation, and laboratory accidents in humans (Dubey and Jones 2008). The pos-
sibility of transmission occurring through inhalation of aerosols or ingestion of
secretions containing T. gondii in Neotropical primates has also been suggested
(Furuta et al. 2001; Carme et al. 2009). In most vertebrate homeotherm hosts, zoites
multiply asexually within the infected cells through endodyogeny and tend to form
perennial cysts viable for many years in multiple tissues and organs, particularly in
skeletal and cardiac muscle, and in the central nervous system (Dubey 2009).
In summary, T. gondii life cycle is characterized by sexually reproducing and
developing oocysts in the small intestine of a natural feline host, before transmitting
to an intermediate homeotherm host. Within the intermediate host, the parasite
develops, forming cysts in various tissues.
Although host immune responses to T. gondii have been studied since the pathogen’s
discovery more than a century ago, their interactions are not fully understood
(Boothroyd 2009; Tait and Hunter 2009). Experimental studies have shown that more
than 1,000 host genes are modulated in Toxoplasma-infected cells, among them genes
encoding proteins implicated in several processes including inflammation, apoptosis,
metabolism, and cell growth and differentiation (Blader and Saeij 2009).
The development of toxoplasmosis in mice and humans is determined by mecha-
nisms involving the pathogenicity of the parasite strain, in addition to the host
immune status (Hill et al. 2011; Pifer and Yarovinsky 2011). Surface antigen 1
(SAG1), the major surface protein of T. gondii, has been recognized as an essential
target of adaptive immune response. However, the parasite has developed strategies
to avoid the powerful immune response (Buzoni-Gatel and Werts 2006). In addition,
it is believed that three main T. gondii genotypes responsible for infection in humans
(types I, II, and III) may induce different patterns of the disease (Saeij et al. 2005).
Clonal lineages differ in growth, migration, and transmigration. In laboratory mice,
it is known that type I strains are very virulent (lethal dose – LD100 of one parasite),
while type II and III strains are much less virulent (lethal dose – LD50 ~103 and
~105). In humans, type I strains are frequently associated with postnatally acquired
ocular infections, whereas type II strains are more related with congenital infections
and toxoplasmic encephalitis (Blader and Saeij 2009; Sibley et al. 2009).
Typically, the natural and intermediate host response to T. gondii infection is
capable of holding back parasite dissemination, reducing mortality rates. The early
Neotropical Primates and Their Susceptibility to Toxoplasma gondii… 263
It is known that both humoral and cellular immune responses are involved in the reso-
lution of acute toxoplasmosis, but the cellular response is considered the most impor-
tant mechanism responsible for the host defense (Däubener and Hadding 1997; Innes
1997). When the parasites invade the intestinal mucosa of intermediate hosts (the
main infection route), the infected enterocytes suffer morphological and physiological
changes and secret chemokines and cytokines that attract polymorphonuclear leuko-
cytes, macrophages, and dendritic cells (DCs) (Buzoni-Gatel and Werts 2006).
Proinflammatory cytokines produced by lymphocytes, macrophages, DCs, and
neutrophils are crucial for controlling T. gondii. In the first stages of infection,
T. gondii activates cells such as macrophages, DCs, and neutrophils to produce high
levels of IL-12 (Gazzinelli et al. 1993b; Johnson and Sayles 1997). Mitogen-
activated protein kinase p38 is required for IL-12 production by macrophages in
response to soluble tachyzoite antigen (STAg) (Mason et al. 2004).
Many studies have demonstrated that IFN-γ, the hallmark of the inflammatory
response, is the major defense component against T. gondii infection that inhibits
parasite replication in various human and mouse cells (Suzuki et al. 1988; Sharma
1990; Däubener and Hadding 1997; Buzoni-Gatel and Werts 2006). Additionally,
IFN-γ, which is produced by natural killer cells (NK cells) in response to IL-12
secretion, contributes to the differentiation of lymphocytes into the Th1 phenotype.
IFN-γ, produced by activated T lymphocytes, natural killer cells, and natural killer
T cells, stimulates macrophages to produce reactive oxygen intermediates (ROI),
leading to the death of T. gondii (Nathan et al. 1983). Hence, IFN-γ is an essential
cytokine for resistance to acute and chronic Toxoplasma infections (Suzuki et al.
1988; Gazzinelli et al. 1993a). Similar to IFN-γ, TNF-α, IL-6, and IL-1 have syner-
gistic effects on the induction of an adequate immune response against T. gondii
(Lang et al. 2007).
In addition to IFN-γ, other cytokines, such as TNF-α, IL-2, and lymphotoxin-α, are
cofactors important to host responses to T. gondii infection. Natural killer cells and
CD4+ and CD8+ T cells are three lymphocyte subsets that produce these cytokines and
have been suggested to influence T. gondii immunity in mice and humans (reviewed in
Blanchard et al. 2008). IL-2, produced by CD4+ T cells, is an important T cell mitogen
(reviewed in Tait and Hunter 2009). CD4+ and CD8+ T cell activation prevent reactiva-
tion of infection, probably by IFN-γ production (Gazzinelli et al. 1992). CD8+ T cells
are known to be crucial for protection against the intracellular parasite T. gondii, because
they are involved in the capacity to induce apoptosis in infected cells (Däubener and
Hadding 1997). CD8+ T cells have also been reported to directly kill the extracellular
264 J.L. Catão-Dias et al.
Humoral Immunity
The balance between the production of proinflammatory (IFN-γ, TNF-α, IL-6, IL-1)
and anti-inflammatory (TGF- β and IL-10) cytokines appears to be decisive for the out-
come of T. gondii infection (reviewed by Lang et al. 2007). IL-10, an anti-inflammatory
cytokine, traditionally inhibits proinflammatory responses, controlling cytokine and
chemokine production. Neutralization of IL-10 in murine models increases central ner-
vous system inflammation. IL-10-deficient mice lose control of their immune response
and die in the acute phases of toxoplasmosis, after uncontrolled IFN-gamma and TNF-α
266 J.L. Catão-Dias et al.
Anti-apoptotic effect
Supress NO production
transcriptional level by reduction of mRNA and protein levels of iNOS. The reduc-
tion of NO production in serum has been implicated in triggering a conversion stage
in the parasite and protecting the host from immunopathological effects of infection
(reviewed in Lang et al. 2007).
In addition, the parasite can suppress the host immune response through induction
of apoptosis in CD4+ T cells while inhibiting the apoptosis of infected cells by inter-
fering with the mitochondrial cytochrome c protein, decreasing caspase proteolytic
activation, and exposing of phosphatidylserine, among others (reviewed by Denkers
and Butcher 2005; Lang et al. 2007; Luder et al. 2009; Santos et al. 2011). Finally,
T. gondii can avoid the fusion of lysosomes with the phagocytic vacuole and form
cysts, thus escaping the host immune system (Seed 1996; Denkers and Butcher 2005).
The tachyzoite, due to its rapid replication, is the protozoan form involved in
triggering Toxoplasma-induced tissue necrosis. Thus, the host’s ability to prevent,
inhibit, or minimize the spread of tachyzoites will determine, in large part, the clinical
changes observed in toxoplasmosis. In most hosts, particularly in immunocompe-
tent adult mammalians, T. gondii infection has no clinically relevant implications,
268 J.L. Catão-Dias et al.
The first report of toxoplasmosis in New World primates was made in 1916, affecting
a specimen of Stentor seniculus (Alouatta seniculus) (reviewed by Nery-Guimaraes
et al. 1971). Since that time, many other reports of this disease in NWP have been
made, predominantly affecting captive animals (Hessler et al. 1971; Anderson and
McClure 1982; Borst and Vanknapen 1984; Cunningham et al. 1992; Dietz et al.
1997; Pertz et al. 1997; Juan-Salles et al. 1998; Bouer et al. 1999; Epiphanio et al.
2000, 2001, 2003; Andrade et al. 2007; Carme et al. 2009; Cedillo-Pelaez et al. 2011).
Although it has been known that NWPs are susceptible to toxoplasmosis for almost
100 years, very little is known about the immune response induced by T. gondii in
these animals, and only serological data are available in the literature.
Table 2 presents a comparison between major clinical and pathological manifesta-
tions of toxoplasmosis in NWP, felids, and humans. In general, toxoplasmosis in
NWPs is a disease with a hyperacute clinical course. In most cases, the animals are
found dead without prior clinical history. When present, the main clinical findings
reported are prostration, dyspnea, hypothermia, nasal foamy serum–bloody exuda-
tion, anorexia, and vomiting. At necropsy, in most animals, the macroscopic changes
are diverse and occur in multiple organs and tissues and are characterized by severe
pulmonary edema and congestion, hepatic congestion with hepatomegaly, spleno-
megaly, spleen lymphoid hyperplasia, mesenteric and mediastinal fibrin-hemorrhagic
Table 2 Major clinical and pathological manifestations of toxoplasmosis in New Word Primates, felids, and humans
New Word Primates Felids Humans
Pattern II (most
Pattern I Cebidae and Pattern III Immune
(Callitrichinae) Atelidae) (Cebus) competent Immune suppressed Immune competent Immune suppressed Congenitally infected
Clinical Hyperacute, Acute, severe, Subacute, Subclinical Congenitally or Most subclinical; Fever, headache, Asymptomatic to
manifes- markedly and mild, or mild lactationally few develop myalgia, anorexia, fatal. Prematurity,
tations severe; mortality with self- infected offspring; fever and fatigue, abdominal intrauterine
mortality from 20 % very limiting stillborn or lethargy lymphadenopa- pain, vomiting, growth retarda-
close to to 80 %. low diarrhea depression, thy. nausea, dyspnea, tion, debilitating
100 %. Prostration, mor- hypothermia, ascites, Debilitating arthralgia. ocular disease,
Prostration, anorexia, tality hepatomegaly, ocular disease Lymphadenopathy strabismus,
dyspnea, hypother- rate chorioretinitis, and has been psychomotor
nasal foamy mia, sudden death reported in few impairment,
exudation, dyspnea, Older cats with cases prostration,
anorexia vomiting immunosuppressive microcephalus,
processes: anorexia, hydrocephalus,
lethargy, dyspnea, convulsion,
persistent/intermit- icterus, hypotonia,
tent fever and hepatomegaly
Pathological Severe Severe Mild and Nonspecific Moderate to severe Retinochoroiditis Encephalitis Retinochoroiditis,
findings fibrin- fibrin- non- necrotizing hepatitis, anterior uveitis,
hemorrhagic hemorrhagic spe- lymphadenitis, and and encephalitis
necrotizing necrotizing cific pneumonia.
pneumonia, pneumonia, Neuronal necrosis
hepatitis, hepatitis, and meningitis
splenitis, splenitis,
enteritis, and enteritis, and
lymphadeni- lymphadeni-
tis tis
Based on Anderson and McClure (1982), Elmore et al. (2010) Epiphanio et al. (2003), Silva (2007), Epiphanio and Catão-Dias (unpublished data)
270 J.L. Catão-Dias et al.
2
The capuchin monkeys are particularly complex in their taxonomy. For many years, taxonomic
arrangements reduced all tufted capuchin monkeys to just one species, Cebus apella, with 11 and
16 (Hill 1960) subspecies. The most recent revisions, by Groves (2001) and Silva (2001), both
based on morphology, differently recognized, as species, the following: apella and macrocephalus
in the Amazon and libidinosus, nigritus, robustus, cay, and xanthosternos to the south. Groves
(2001) presented an alternative as follows: Amazon forms C. apella apella, C. a. fatuellus, C. a.
macrocephalus, C. a. peruanus, and C. a. tocantinus and southern forms C. libidinosus libidinosus,
C. l. pallidus, C. l. paraguayanus, C. l. juruanus (Amazonian), C. nigritus nigritus, C. n. robustus,
C. n. cucullatus, and C. xanthosternos (see Fragaszy et al. (2004) and Rylands et al. (2005)). In
Brazil, captive capuchins (independently of origin) are generally named as Cebus apella, and in
many occasions and in different institutions, individuals of different subspecies or species are kept
together generating hybrid groups (M.C.M. Kierulff, personal observation). Because of all these
problems, we decided to maintain the original names used for Cebus species cited in the refer-
ences, mostly named just as Cebus apella with no distinction to subspecies, even known that it
refers to species other than the Amazonians.
Table 3 Seroprevalence of Toxoplasma gondii in New World Primates
Genus/species Local Animal tested (n) Test Captive/free ranging Positive test (%) Reference
S. oedipus oedipus South 100 SFR FR 0 Werner et al. (1969), Nery-Guimaraes and Franken
America (1971)
S. sciureus Brazil 17 SFR Captive/FR 17,6 Nery-Guimaraes and Franken (1971)
C. apella 26 15,3
A. belzebuth 1 0
Aotus sp. 1 0
A. geoffroyi 1 0
C. jacchus 2 0
C. penicillata 5 0
L. lagotricha 1 0
C. apella Brazil 5 SFR FR 60 Sogorb et al. (1972)
A. fusca 12 42,1
Saimiri spp. Brazil 49 IHT FR 63,3 Ferraroni and Marzochi (1980)
C. apella Colombia 10 NA Captive 0/0 Cadavid et al. (1991)
C. capucinus 15 13,3
C. albifrons 22 40,9
Saimiri sciureus England 4a IFA Captive 100/75 Cunningham et al. (1992)
11# 100/54,5
A. seniculus French 50 DA FR 4 De Thoisy et al. (2003)
Neotropical Primates and Their Susceptibility to Toxoplasma gondii…
S. midas Guiana 50 0
C. apella Brazil 43 MAT FR 30,2 Garcia et al. (2005)
A. caraya 17 17,6
L. lagotricha USA 2a LA Captive 100 Gyimesi et al. (2006)
13 IHT 0/0/0
MAT
C. apella Brazil 14 IFAT Captive 28.7 Leite et al. (2008)
13 MAT 30.8
(continued)
271
272
Table 3 (continued)
Genus/species Local Animal tested (n) Test Captive/free ranging Positive test (%) Reference
Alouatta caraya USA 1 MAT Captive 0 de Camps et al. (2008)
Ateles geoffroyi 4 0
Callicebus moloch 5 0
donacophilus 3 0
Callimico goeldii 1 0
Callithrix kuhlii 4 0
L. lagotricha 1 0
L. chrysomelas 3 0
L. rosalia 8 0
Pithecia pithecia 6 0
Saguinus
geoffroyi
S. sciureus Israel 24a MAT Captive 83,3 Salant et al. (2009)
C. apella Brazil 105 IFAT Captive 79 Bouer et al. (2010)
Callithrix sp. 42 26,2
Alouatta sp. 20 50
Leontopithecus 15 20
sp. Ateles sp. 7 57,14
Saimiri sp. 6 33,33
Saguinus sp. 5 0
Aotus sp. 3 66,66
J.L. Catão-Dias et al.
Lagothrix sp. 3 0
Genus/species Local Animal tested (n) Test Captive/free ranging Positive test (%) Reference
C. jacchus Brazil 25 LA Captive 0 Epiphanio & Catão-Dias, unpublised data
C. penicillata 18 0
C. geoffroyi 8 0
C. aurita 2 0
C. kuhlii 1 0
Callithrix sp. 1 0
S. bicolor 2 0
S. midas 3 0
S. niger 2 0
L. chrysomelas 46 2,2
L. rosalia 1 0
L. chrysopygus 1 0
C. apella 100 73
A. marginatus 2 100
A. belzebuth 1 0
A. paniscus 5 20
Ateles sp. 3 100
Saimiri sciureus 6 0
L. lagotricha 2 0
A. caraya 7 42,9
A. fusca 5 0
Alouatta sp. 5 0
Aotus sp. 2 50
IFAT Indirect immunofluorescence, MAT Modified Agglutination Test, IHT indirect heamagglution test, LA latex agglutination, PCR polymerase chain reac-
Neotropical Primates and Their Susceptibility to Toxoplasma gondii…
Moreover, the data linking the evolution of the serological profile with the occur-
rence of toxoplasmosis outbreaks in NWPs can provide interesting information for
understanding this process. At the London Zoo, one-third of the S. sciureus colony
died of toxoplasmosis, and serological surveys showed that most animals had titers
indicative of recent infection. IgG was detected in 11 surviving animals and IgM in
six individuals from this group (Cunningham et al. 1992). In another outbreak that
caused the death of 24 S. sciureus in Israel, 83.3 % of the animals had positive serol-
ogy, suggesting that humoral immunity is not an effective defense mechanism for
sudden toxoplasmosis or reinfection in this NWP species (Salant et al. 2009).
Similarly, Lagothrix lagotricha individuals who died of sudden toxoplasmosis
showed positive serology for anti-Toxoplasma gondii antibodies in three distinct
types of tests (latex agglutination, indirect hemagglutination, modified agglutina-
tion), besides identification of T. gondii by PCR (Gyimesi et al. 2006).
Experimental infections have been performed in an attempt to better understand
the role of the humoral response in the development of toxoplasmosis in NWPs. In
one experiment, 28 Saguinus sp. were infected and died within few days, without
the detection of specific antibodies anti-Toxoplasma gondii (Werner et al. 1969). In
another study, five C. apella individuals were infected with T. gondii (N strain, type
II tachyzoite suspension—1 × 105 mL—by intraperitoneal route) and exhibited non-
specific and mild clinical signs for only 3 days postinfection and were euthanized
102 days postinfection. Macroscopic and microscopic lesions observed were mild
and were not correlated with toxoplasmosis. However, anti-Toxoplasma gondii IgG
titers were detected by IFA and ELISA 9 days postinfection and lasted until the end
of the investigation (Bouer et al. 2010). In another experiment, S. sciureus orally
infected died approximately 1 week postinfection; however, anti-Toxoplasma
immunoglobulin titers were not detected by immunoblot (Furuta et al. 2001).
from those seen in most other platyrrhines. In Cebus, the infection tends to induce
high and persistent IgG titers, the animals rarely die, and the morphological changes
seen in experimental cases are mild and nonspecific (Bouer et al. 2010).
The variable prevalence of anti-Toxoplasma gondii antibodies in NWPs, and
especially the very low frequency of antibodies in Callitrichinae, may explain the
very high susceptibility of these animals to infection. Callitrichinae, specifically, die
rapidly during the acute phase of infection, before IgG production is initiated. On
the other hand, the high IgG titers observed in Cebus, associated with rare reports of
death in this genus, suggests that these animals are able to establish an efficient
antibody immune response against the infection.
Over the course of 20 years at the Laboratory of Comparative Pathology of Wildlife
(LAPCOM—Laboratório de Patologia Comparada de Animais Selvagens) of the
FMVZ–USP, the authors have witnessed at least six major toxoplasmosis outbreaks in
NWP. These outbreaks mainly occurred among the genera Callithrix, Leontopithecus,
Lagothrix, Alouatta, Saimiri, and Saguinus. In these outbreaks, all individuals who
exhibited clinical signs subsequently died. In at least two of the outbreaks, diseased
animals (Leontopithecus chrysopygus, L. rosalia, L. chrysomelas) were submitted to
recommended therapeutic procedures (Osborn and Lowenstine 1998), but did not sur-
vive. Significantly, Cebus specimens were not affected in any of the outbreaks fol-
lowed by LAPCOM, although these animals constitute the large majority of the
NWPs populations kept in captivity in the zoos involved.
The clinical and serological aspects of toxoplasmosis (described above) for most
NWPs, especially Callitrichinae, resemble those reported for the mountain hare
(Lepus timidus). These Eurasian animals are especially susceptible to the disease
and develop severe acute symptoms that are often fatal. Experimental studies have
shown that mountain hares have very low antibody titers and inefficient prolifera-
tion of T lymphocytes when exposed to T. gondii, compared with those observed in
domestic rabbits (Oryctolagus cuniculus) (Gustafsson et al. 1997), suggesting that
the hares are incompetent to establish an efficient adaptive immune response against
the parasite.
Other aspects that deserve to be addressed in understanding differing susceptibil-
ity of NWP species to T. gondii include the variability of strains, infective doses,
and the possibility of recrudescence of cysts. The recent characterization of differ-
ent T. gondii strains shows that there is significant genetic diversity that was previ-
ously unknown (Dubey 2009; Dubey and Su 2009). In addition, it is known that the
main three known strains can induce different patterns of the disease in humans
(Saeij et al. 2005).
Furthermore, studies investigating the types of strains involved in cases of toxo-
plasmosis in NWPs are rare and are restricted to four outbreaks in Saimiri: two in
French Guiana, with the identification of strains II and III and atypical alleles
(Carme et al. 2009); one in Israel involving strain III (Salant et al. 2009); and one in
Mexico involving strain I (Cedillo-Pelaez et al. 2011). The high susceptibility of
Saimiri to the three main T. gondii strains suggests that, at least for this NWP spe-
cies, the genetic variability of the parasite may not be a determining factor for dis-
ease manifestation.
276 J.L. Catão-Dias et al.
T. gondii, by more efficiently controlling the spread of pathogens and allowing the
establishment of a more balanced host–parasite relationship similar to the vast major-
ity of warm-blooded hosts infected with T. gondii. Moreover, the inadequate prolifera-
tive response observed in L. chrysomelas would prevent the emergence of an efficient
cellular response against the infection, favoring the development of fatal infections.
Our results, associated with those published by other researchers, raise several
questions about the mechanisms involved in the unsatisfactory cellular immune
response shown by the vast majority of NWPs against T. gondii, such as:
1. Would the inflammatory response induced by T. gondii in NWPs be intensified to
the point of leading to the overproduction of IFN-γ, TNF-α, IL-6, IL-12, or IL-1β?
2. Could there be overstimulation of TLR, with consequent amplification of the
expression of the transcription factor NF-kB?
3. Could regulatory mechanisms, such as anti-inflammatory cytokines TGF-β,
IL-10, and lipoxins, be downregulated?
4. Could there be disturbances in the fusion of the parasitophorous vacuole with
lysosomes, favoring the survival of T. gondii in cells infected by T. gondii in
NWPs?
5. Could the apoptosis mechanisms of Toxoplasma-infected cells be deficient?
6. Could the mechanisms of NO production be lacking?
Currently, many sequences of genes encoding proteins related to the immune
system of platyrrhines are known (Table 4). Hopefully in the near future, these tools
can be used to clarify some of the many unanswered questions regarding the toxo-
plasmosis immunology in NWPs.
Approximately three million years ago, Panamanian land bridge formed and allowed
the immigration of many intermittent “invaders” from the North America into the
South America, including carnivores (e.g., felids, canids, and mustelids) over the
course of the Pleistocene (Webb 1976; Simpson 1980; Marshall 1988). At that time,
T. gondii may also have migrated into the continent with carnivore species (reviewed
in Sibley et al. 2009). In fact, T. gondii strains from North and South America share
a common ancestry, and it was estimated that they last shared a common ancestor
one million years ago (Sibley et al. 2009).
Platyrrhines, however, have a long evolutionary history preceding the migration
of carnivore species into South America. The early platyrrhine fossils come from
the Late Oligocene in Bolivia (24–28 million years ago). Indeed, species related to
the living Neotropical primates were present in Colombia, during the Middle to
Late Miocene, suggesting a common platyrrhine ancestor in the Late Oligocene or
Early Miocene (14–24 million years ago) (Fleagle and Tejedor 2002). Recent
Table 4 List of selected genes related to immune response in New World Primates
Species Gene GenBank accession
Aotus infulatus IL-12B DQ989359.1
Aotus lemurinus TNF-a AF097329
Aotus lemurinus IFN-γ AF097327.1
Aotus nancymaae CD4 FJ623078.1
Aotus nancymaae TLR9 AY788894.1
Aotus nancymaae IFN-γ AF014512.1
Aotus nigriceps TNF-a AF097328
Aotus trivirgatus CD40 ligand AF344860.1
Aotus trivirgatus MHCI AB113205.1
Aotus vociferans TNF-a AF014508
Aotus vociferans IL-10 AAD01532.1
Aotus vociferans IFN-γ AF014507.1
Ateles belzebuth TLR4 AB446521
Ateles belzebuth MHCI AB113112
Ateles geoffroyi IL-10 ABM65916.1
Ateles geoffroyi TLR4 AB446522
Callicebus moloch MHCII AF197231.1
Callithrix jacchus IL-12B AB539805.1
Callithrix jacchus iNOS AM712438
Callithrix jacchus CD4 AF452616.1
Callithrix jacchus CD8 DQ189217
Callithrix jacchus IL-1a AB539804
Callithrix jacchus TNF-a DQ520835
Callithrix jacchus TLR4 AB446516
Callithrix jacchus TLR9 XM_002758237
Callithrix jacchus IL-27 XM_002756059.1
Callithrix jacchus CD40 DQ189221.1
Callithrix jacchus p47 GTPase XM_002762275
Callithrix jacchus MyD88 XM_002759734
Callithrix jacchus P2X(7) XM_002753098
Callithrix jacchus MHCII AF197230.1
Cebus apella TLR4 AB446520.1
Leontopithecus rosalia TLR4 AB446518
Saguinus imperator TLR5 FJ542217
Saguinus labiatus MHCII JF414576.1
Saguinus mystax IFN-γ FJ598592.1
Saguinus oedipus TNF AY091968
Saguinus oedipus TLR4 AB446517.1
Saguinus oedipus TLR2 EU488857.1
Saguinus oedipus MHCII AF197226.1
Saimiri sciureus IL-12B DQ989358.1
Saimiri sciureus CD4 AF452617
Saimiri sciureus CD8 AJ130819
Saimiri sciureus IL-1b AF294754
Saimiri sciureus TNF-a AJ437697
Saimiri sciureus TNF DQ989365
Saimiri sciureus TLR4 AB446519
Saimiri sciureus IL-10 Q8MKG9.1
Saimiri sciureus IFN-γ AF414102.1
Neotropical Primates and Their Susceptibility to Toxoplasma gondii… 279
phylogenetic analysis using DNA samples of primate species showed that Platyrrhini
diverged from a last common ancestor with Catarrhini 43.5 million years ago during
the Eocene. The common ancestor to Pitheciidae originated 20.2 million years ago,
and the Cebidae radiation initiated with the emergence of Cebinae and Saimirinae
approximately 20 million years ago (Perelman et al. 2011). These data indicate that
nonhuman primates were already in South America at the time of the felid (possibly
infected with T. gondii) invasion, from the north hemisphere.
The high susceptibility of NWPs to toxoplasmosis has been known for almost
100 years. The main hypotheses proposed to explain this condition can be summa-
rized as follows: (a) NWPs have evolved for over 20 million years without the pres-
ence of felids and therefore would not have acquired adaptations to the pathogen
over this time, especially cellular responses to T. gondii (Cunningham et al. 1992),
and (b) even after the arrival of felids (possibly infected with T. gondii) to the
Neotropics, the arboreal habits of NWPs would have restricted their contact with the
feces of felids infected by the protozoan oocysts, limiting the development of effi-
cient immune response (Innes 1997). The available data compiled in this review
corroborates many of the assumptions covered above to justify the high susceptibil-
ity of some NWPs to toxoplasmosis. On the other hand, they do not explain the
significant differences observed between groups of NWPs, in particular the high
resistance reported for Cebus.
To some extent, it is possible to assess how a particular animal explores and
occupies an environment in relation to the intensity and diversity of the parasite load
it carries. A study conducted in Costa Rica showed C. capucinus with higher para-
site infestation in fecal samples than Alouatta and Ateles (Stuart et al. 1998).
Fragaszy et al. (2004) suggested that at least three characteristics of Cebus behavior
may lead to higher parasite infestation in these monkeys than in sympatric Alouatta
and Ateles: Cebus drinks from water holes, frequently forages on the ground, and
eats a wider variety of foods. These behaviors may bring Cebus into contact with
greater variety of parasites. Considering the epidemiological characteristics of
T. gondii, it is possible to speculate that the aspects described above could justify a
diverse exposure of NWPs to the protozoan.
Contaminated water can be a source of toxoplasmosis, and this behavior has
already been recorded for other Alouatta species that drink water in holes in
branches or trunks or in bromeliads (Glander 1978; Gilbert and Stouffer 1989;
Bicca-Marques 1992; Giudice and Mudry 2000) or go to the ground to drink water
(Almeida-Silva et al. 2005). Other Platyrrhini such as Aotus (Wright 1981), Saimiri
(Baldwin and Baldwin 1981), Cebuella (Soini 1988), Saguinus (Snowdon and Soini
1988), and Brachyteles (Mendes et al. 2010) have also been observed drinking on
the banks of streams and/or rivers during the dry season.
Cebus sp. comes down to the ground more frequently in comparison to other
NWPs where they may be more frequently exposed to excreted T. gondii. Recent
studies have reported the use of tools (stones used as hammers and anvils) to open
hard nuts by wild Cebus in places where the groups (from different species) use the
ground more frequently (Fragaszy et al. 2004; Canale et al. 2009). Robinson (1984)
found that many of the invertebrates consumed by C. olivaceus were found on the
280 J.L. Catão-Dias et al.
ground. They visually search the leaf litter and sweep the leaves to reveal the insects
hidden underneath (Fragaszy et al. 2004).
Additionally, in our opinion, an important behavioral characteristic of Cebus is
that it is the most carnivorous of the platyrrhines. It has been often noted that the
consumption of invertebrates and small warm-blooded vertebrates plays a role in
the transmission of T. gondii to NWPs (Epiphanio et al. 2003; Carme et al. 2009;
Salant et al. 2009). Most of the protein in Cebus diets comes from invertebrates
(insects and other arthropods), but while hunting for invertebrates, they sometimes
find vertebrates that they capture and consume. It has been widely reported that they
capture and consume a variety of relatively large vertebrates that may weigh up to
one third the Cebus body weight and may constitute up to 3 % of their feeding time
(Fragaszy et al. 2004). Along with chimpanzees (Pan troglodytes), Cebus are one of
the few nonhuman primate species that have been reported to hunt vertebrate prey
in more than an occasional, incidental manner (Fragaszy et al. 2004).
The types of vertebrate prey that Cebus has been reported to consume include
birds and their nestlings and eggs, lizards, frogs, rodents, bats, squirrels (Sciurus
variegatoides), coati pups (Nasua narica), and infant titi monkeys (Callicebus
moloch) (Izawa 1978; Terborgh 1983; Fedigan 1990; Galetti 1990; Rose 1997;
Sampaio and Ferrari 2005). Cebus xanthosternos from the Una Reserve, Bahia,
Brazil, have been seen preying upon bristle-spined rat pups (Chaetomys subspi-
nosus) on four occasions (Priscilla G. Suscke, personal communication) (Fig. 4).
In a forest in Rio de Janeiro, a bamboo rat (Kannabateomys amblyonyx) was
found preyed upon immediately following a passage of a Cebus nigritus group
Neotropical Primates and Their Susceptibility to Toxoplasma gondii… 281
Most NWPs are very susceptible to toxoplasmosis, and according to the authors’
experience, this is possibly the most important cause of acute death of infectious
origin affecting Platyrrhini in captivity in Brazil. However, the susceptibility of
NWPs to T. gondii is variable, with Callitrichinae (Callithrix, Saguinus, and
Leontopithecus) showing mortality rate close to 100 %, Atelidae and some Cebidae
(Saimiri, Aotus) showing variable mortality patterns, and genus Cebus showing
high resistance, with rare deaths reported due to toxoplasmosis.
The reasons for the high susceptibility of most NWPs to T. gondii are not clear.
We believe that it may be due, at least in part, to ecological and behavioral charac-
teristics of different NWPs that led to different degrees of exposure to T. gondii over
evolutionary time. It is possible to speculate that such variable exposure to the pro-
tozoa may have led, along the evolutionary process of NWPs, to differentiated
immunological features culminating in the relative ability to resist the infection.
Naturally, there are many unanswered questions to investigate and novel areas to
research regarding Toxoplasma–NWP interactions. To better understand
Toxoplasmosis manifestation in NWPs, we believe certain studies are very impor-
tant including further research on NWP cellular (proliferative assays, role and mea-
surement of cytokines) and humoral immune responses to the pathogen (more
comprehensive serological surveys, both in captivity and in the wild; use and valida-
tion of different techniques) as well as molecular epidemiology of T. gondii (char-
acterization of strains and their environmental distribution). Obtaining new
information in these areas will certainly help clarify questions about NWP–T. gon-
dii interactions.
282 J.L. Catão-Dias et al.
Finally, considering our laboratory has witnessed on several occasions the devas-
tating effect that toxoplasmosis can have on ex situ conservation programs for
NWPs, we would like to emphasize the importance of curatorial/zoological institu-
tions adopting the best management practices. We see such policies as the only
effective option, currently, for the prevention of new outbreaks that can, otherwise,
decimate genetically invaluable populations of NWPs.
References
Aguiar JM, Lacher TE Jr (2003) On the morphological distinctiveness of Callithrix humilis Van
Roosmalen et al. 1998. Neotrop Primates 11:11–18
Alfaro JWL, Boubli JP, Olson LE, Di Fiore A, Wilson B, Gutierrez-Espeleta GA, Chiou KL,
Schulte M, Neitzel S, Ross V, Schwochow D, Nguyen MTT, Farias I, Janson CH, Alfaro ME
(2012a) Explosive Pleistocene range expansion leads to widespread Amazonian sympatry
between robust and gracile capuchin monkeys. J Biogeogr 39:272–288
Alfaro JWL, Silva JS Jr, Rylands AB, Boubli JP (2012b) How different are robust and gracile
capuchin monkeys? An argument for the use of Sapajus and Cebus. Am J Primatol 74(4):273–
286. doi:10.1002/ajp.22007:1-14
Aliberti J (2005) Host persistence: exploitation of anti-inflammatory pathways by Toxoplasma
gondii. Nat Rev Immunol 5:162–170
Aliberti J, Hieny S, Reis e Sousa C, Serhan CN, Sher A (2002a) Lipoxin-mediated inhibition of
IL-12 production by DCs: a mechanism for regulation of microbial immunity. Nat Immunol
3:76–82
Aliberti J, Serhan C, Sher A (2002b) Parasite-induced lipoxin A4 is an endogenous regulator of
IL-12 production and immunopathology in Toxoplasma gondii infection. J Exp Med
196:1253–1262
Almeida-Silva B, Guedes PG, Boubli JP, Strier KB (2005) Deslocamento terrestre e o comporta-
mento de beber em um grupo de barbados (Alouatta guariba clamitans Cabrera, 1940) em
Minas Gerais, Brasil. Neotrop Primates 13:1–3
Anderson DC, McClure HM (1982) Acute disseminated fatal toxoplasmosis in a squirrel monkey.
J Am Vet Med Assoc 181:1363–1366
Andrade MCR, Coelho JMCO, Amendoeira MRR, Vicente RT, Cardoso CVP, Ferreira PCBF,
Marchevsky RS (2007) Toxoplasmosis in squirrel monkeys: histological and immunohisto-
chemical analysis. Ciência Rural 37:1724–1727
Baldwin JD, Baldwin JI (1981) The squirrel monkeys, genus Saimiri. In: Coimbra-Filho AF,
Mittermeir RA (eds) Ecology and behavior of neotropical primates. Academia Brasileira de
Ciências, Rio de Janeiro, pp 241–276
Bernsteen L, Gregory CR, Aronson LR, Lirtzman RA, Brummer DG (1999) Acute toxoplasmosis
following renal transplantation in three cats and a dog. J Am Vet Med Assoc 215:1123–1126
Bicca-Marques JC (1992) Drinking behavior in the black howler monkey (Alouatta caraya). Folia
Primatol (Basel) 58:107–111
Bicca-Marques JC, Silva VM, Gomes DF (2011) In: Reis NR, Perachi AL, Pedro WA, Lima IP
(eds) Mamíferos do Brasil, 2nd edn. Londrina, PR. pp. 107–150.
Neotropical Primates and Their Susceptibility to Toxoplasma gondii… 283
Blader IJ, Saeij JP (2009) Communication between Toxoplasma gondii and its host: impact on
parasite growth, development, immune evasion, and virulence. APMIS 117:458–476
Blanchard N, Gonzalez F, Schaeffer M, Joncker NT, Cheng T, Shastri AJ, Robey EA, Shastri N
(2008) Immunodominant, protective response to the parasite Toxoplasma gondii requires anti-
gen processing in the endoplasmic reticulum. Nat Immunol 9:937–944
Boinski S (1987) Mating patterns in squirrel-monkeys (Saimiri-Oerstedi) - implications for sea-
sonal sexual dimorphism. Behav Ecol Sociobiol 21:13–21
Boothroyd JC (2009) Toxoplasma gondii: 25 years and 25 major advances for the field. Int J
Parasitol 39:935–946
Bordignon MO, Setz EZF, Caselli CB (2008) Gênero Callicebus Thomas 1903. In: Reis NR,
Perachi AL, Andrade FR (eds) Primatas brasileiros. Technical Books, Londrina, pp 153–166
Borst GHA, Vanknapen F (1984) Acute acquired toxoplasmosis in primates in a zoo. J Zoo Wildl
Med 15:60–62
Bouer A, Werther K, Catao-Dias JL, Nunes AL (1999) Outbreak of toxoplasmosis in Lagothrix
lagotricha. Folia Primatol (Basel) 70:282–285
Bouer A, Werther K, Machado RZ, Nakaghi AC, Epiphanio S, Catao-Dias JL (2010) Detection of
anti-Toxoplasma gondii antibodies in experimentally and naturally infected non-human pri-
mates by Indirect Fluorescence Assay (IFA) and indirect ELISA. Rev Bras Parasitol Vet
19:26–31
Bradley PJ, Ward C, Cheng SJ, Alexander DL, Coller S, Coombs GH, Dunn JD, Ferguson DJ,
Sanderson SJ, Wastling JM, Boothroyd JC (2005) Proteomic analysis of rhoptry organelles
reveals many novel constituents for host-parasite interactions in Toxoplasma gondii. J Biol
Chem 280:34245–34258
Brown CCB, Barker DC (2007) Alimentary system. In: Maxie MG (ed) Pathology of domestic
animals. Elsevier, Philadelphia, PA, pp 1–296
Brown AD, Zunino GE (1990) Dietary variability in Cebus apella in extreme habitats—evidence
for adaptability. Folia Primatol 54:187–195
Buzoni-Gatel D, Werts C (2006) Toxoplasma gondii and subversion of the immune system. Trends
Parasitol 22:448–452
Cadavid AP, Canas L, Estrada JJ, Ramirez LE (1991) Prevalence of anti-Toxoplasma gondii anti-
bodies in Cebus spp in the Santa Fe Zoological Park of Medellin, Colombia. J Med Primatol
20: 259–261
Canale GR, Guidorizzi CE, Kierulff MC, Gatto CA (2009) First record of tool use by wild popula-
tions of the yellow-breasted capuchin monkey (Cebus xanthosternos) and new records for the
bearded capuchin (Cebus libidinosus). Am J Primatol 71:366–372
Carme B, Ajzenberg D, Demar M, Simon S, Darde ML, Maubert B, de Thoisy B (2009) Outbreaks
of toxoplasmosis in a captive breeding colony of squirrel monkeys. Vet Parasitol
163:132–135
Cedillo-Pelaez C, Rico-Torres CP, Salas-Garrido CG, Correa D (2011) Acute toxoplasmosis in
squirrel monkeys (Saimiri sciureus) in Mexico. Vet Parasitol 80:368–371
Correa D, Canedo-Solares I, Ortiz-Alegria LB, Caballero-Ortega H, Rico-Torres CP (2007)
Congenital and acquired toxoplasmosis: diversity and role of antibodies in different compart-
ments of the host. Parasite Immunol 29:651–660
Cunha RGT (2008) Gênero Aotus Illiger 1811. In: Reis NR, Perachi AL, Andrade FR (eds)
Primatas brasileiros. Technical Books, Londrina, pp 115–125
Cunningham AA, Buxton D, Thomson KM (1992) An epidemic of toxoplasmosis in a captive
colony of squirrel monkeys (Saimiri sciureus). J Comp Pathol 107:207–219
Däubener W, Hadding U (1997) Cellular immune reactions directed against Toxoplasma gondii
with special emphasis on the central nervous system. Med Microbiol Immunol
185:195–206
de Camps S, Dubey JP, Saville WJ (2008) Seroepidemiology of Toxoplasma gondii in zoo animals
in selected zoos in the midwestern United States. J Parasitol 94: 648–653
De Rodaniche E (1954) Spontaneous toxoplasmosis in the whiteface monkey, Cebus capucinus, in
Panama. Am J Trop Med Hyg 3:1023–1025
284 J.L. Catão-Dias et al.
de Thoisy B, Demar M, Aznar C, Carme B (2003) Ecologic correlates of Toxoplasma gondii expo-
sure in free-ranging neotropical mammals. J Wildl Dis 39: 456–459
Debierre-Grockiego F, Schwarz RT (2010) Immunological reactions in response to apicomplexan
glycosylphosphatidylinositols. Glycobiology 20:801–811
Debierre-Grockiego F, Campos MA, Azzouz N, Schmidt J, Bieker U, Resende AG, Santos Mansur
D, Weingart R, Schmidt RR, Golenbock DT, Gazzinelli RT, Schwarz RT (2007) Activation of
TLR2 and TLR4 by glycosylphosphatidylinositols derived from Toxoplasma gondii. J Immunol
179:1129–1137
Defler TR (2005) Primates of Colombia: conservation international. 550 p
Denkers EY, Butcher BA (2005) Sabotage and exploitation in macrophages parasitized by intracel-
lular protozoans. Trends Parasitol 21:35–41
Denkers EY, Gazzinelli RT (1998) Regulation and function of T-cell-mediated immunity during
Toxoplasma gondii infection. Clin Microbiol Rev 11:569–588
Denkers EY, Yap G, Scharton-Kersten T, Charest H, Butcher BA, Caspar P, Heiny S, Sher A (1997)
Perforin-mediated cytolysis plays a limited role in host resistance to Toxoplasma gondii. J
Immunol 159:1903–1908
Dietz HH, Henriksen P, Bille-Hansen V, Henriksen SA (1997) Toxoplasmosis in a colony of new
world monkeys. Vet Parasitol 68:299–304
Dubey JP (2009) History of the discovery of the life cycle of Toxoplasma gondii. Int J Parasitol
39:877–882
Dubey JP, Jones JL (2008) Toxoplasma gondii infection in humans and animals in the United
States. Int J Parasitol 38:1257–1278
Dubey JP, Su CL (2009) Population biology of Toxoplasma gondii: what’s out and where did they
come from. Mem Inst Oswaldo Cruz 104:190–195
Dubey JP, Lindsay DS, Speer CA (1998) Structures of Toxoplasma gondii tachyzoites, bradyzoites,
and sporozoites and biology and development of tissue cysts. Clin Microbiol Rev 11:267–299
Elmore SA, Jones JL, Conrad PA, Patton S, Lindsay DS, Dubey JP (2010) Toxoplasma gondii:
epidemiology, feline clinical aspects, and prevention. Trends Parasitol 26:190–196
Epiphanio S, Guimaraes MA, Fedullo DL, Correa SH, Catao-Dias JL (2000) Toxoplasmosis in
golden-headed lion tamarins (Leontopithecus chrysomelas) and emperor marmosets (Saguinus
imperator) in captivity. J Zoo Wildl Med 31:231–235
Epiphanio S, Sa LR, Teixeira RH, Catao-Dias JL (2001) Toxoplasmosis in a wild-caught black lion
tamarin (Leontopithecus chrysopygus). Vet Rec 149:627–628
Epiphanio S, Sinhorini IL, Catao-Dias JL (2003) Pathology of toxoplasmosis in captive new world
primates. J Comp Pathol 129:196–204
Feagle JG (1999) Primate adaptation and evolution. Academic, San Diego, CA
Fedigan LM (1990) Vertebrate predation in Cebus capucinus: meat eating in a neotropical monkey.
Folia Primatol (Basel) 54:196–205
Ferguson DJ, Hutchison WM, Dunachie JF, Siim JC (1974) Ultrastructural study of early stages of
asexual multiplication and microgametogony of Toxoplasma gondii in the small intestine of
the cat. Acta Pathol Microbiol Scand B Microbiol Immunol 82:167–181
Fernandez-Duque E (2006) Aotinae: social monogamy in the only nocturnal haplorhines. In:
Campbell CJ, Fuentes A, MacKinnon KC, Panger M, Bearder SK (eds) Primates in perspec-
tive. Oxford University Press, New York, NY, pp 139–154
Ferraroni JJ, Marzochi MC (1980) [Prevalence of Toxoplasma gondii infection in domestic and
wild animals, and human groups of the Amazonas region]. Mem Inst Oswaldo Cruz 75: 99–109
Ferreira RG, Resende BD, Mannu M, Ottoni EB, Izar P (2002) Bird predation and prey-transfer in
brown capuchin monkeys (Cebus apella). Neotrop Primates 10:84–89
Fleagle JG, Tejedor MF (2002) Early platyrrhines of southern South America. Cambridge Stud
Biol Evolut Anthropol 33:161–173
Fragaszy DM, Visalberghi E, Robinson JG (1990) Variability and adaptability in the genus Cebus.
Folia Primatol 54:114–118
Fragaszy DM, Visalberghi E, Fedigan LM (2004) The complete capuchin—the biology of the
genus Cebus. Cambridge University Press, Cambridge, UK
Neotropical Primates and Their Susceptibility to Toxoplasma gondii… 285
Freese CH, Oppehheimer JR (1981) The Capuchin monkeys, Genus Cebus. In: Coimbra-Filho AF,
Mittermeir RA (eds) Ecology and behavior of neotropical primates. Academia Brasileira de
Ciências, Rio de Janeiro, pp 331–390
Frenkel JK, Dubey JP, Miller NL (1969) Toxoplasma gondii: fecal forms separated from eggs of
the nematode Toxocara cati. Science 164:432–433
Frenkel JK, Dubey JP, Miller NL (1970) Toxoplasma gondii in cats: fecal stages identified as coc-
cidian oocysts. Science 167:893–896
Fujigaki S, Saito K, Takemura M, Maekawa N, Yamada Y, Wada H, Seishima M (2002)
L-tryptophan-L-kynurenine pathway metabolism accelerated by Toxoplasma gondii infection
is abolished in gamma interferon-gene-deficient mice: cross-regulation between inducible
nitric oxide synthase and indoleamine-2,3-dioxygenase. Infect Immun 70:779–786
Furuta T, Une Y, Omura M, Matsutani N, Nomura Y, Kikuchi T, Hattori S, Yoshikawa Y (2001)
Horizontal transmission of Toxoplasma gondii in squirrel monkeys (Saimiri sciureus). Exp
Anim 50:299–306
Galetti M (1990) Predation on squirrel (Sciurus aestuans) by Capuchin Monkey (Cebus apella).
Mammalia 54:152–154
Garcia JL, Svoboda WK, Chryssafidis AL, de Souza ML, Shiozawa MM, de Moraes AL, Teixeira
GM, Ludwig G, da Silva LR, Hilst C, Navarro IT (2005) Sero-epidemiological survey for
toxoplasmosis in wild New World monkeys (Cebus spp.; Alouatta caraya) at the Parana river
basin, Parana State, Brazil. Vet Parasitol 133:307–311
Gardiner CH, Fayer R, Dubey JP (1998) An atlas of protozoan parasites in animal tissues. United
States Department of Agriculture, Agriculture Handbook, Washington, DC
Gazzinelli R, Xu YH, Hieny S, Cheever A, Sher A (1992) Simultaneous depletion of Cd4+ and
Cd8+ Lymphocytes-T is required to reactivate chronic infection with Toxoplasma-gondii. J
Immunol 149:175–180
Gazzinelli RT, Eltoum I, Wynn TA, Sher A (1993a) Acute cerebral toxoplasmosis is induced by
in vivo neutralization of TNF-alpha and correlates with the down-regulated expression of
inducible nitric oxide synthase and other markers of macrophage activation. J Immunol 151:
3672–3681
Gazzinelli RT, Hieny S, Wynn TA, Wolf S, Sher A (1993b) Interleukin-12 is required for the
T-lymphocyte-independent induction of interferon-gamma by an intracellular parasite and
induces resistance in T-cell-deficient hosts. Proc Natl Acad Sci U S A 90:6115–6119
Gazzinelli RT, Hayashi S, Wysocka M, Carrera L, Kuhn R, Muller W, Roberge F, Trinchieri G,
Sher A (1994) Role of Il-12 in the initiation of cell-mediated-immunity by Toxoplasma-gondii
and Its regulation by Il-10 and nitric-oxide. J Eukaryot Microbiol 41:S9
Gilbert KA, Stouffer PC (1989) Use of a ground-water source by mantled howler monkeys
(Alouatta palliata). Biotropica 21:380
Giudice AM, Mudry MD (2000) Drinking behavior in the black howler monkey (Alouatta caraya).
Zoocriadores 3:11–19
Glander KE (1978) Drinking from arboreal water sources by mantled howling monkeys (Alouatta
palliata Gray). Folia Primatol 29:206–217
Groves CP (2001) Primate taxonomy. Smithsonian Institution Press, Washington, DC
Gustafsson K, Wattrang E, Fossum C, Heegaard PM, Lind P, Uggla A (1997) Toxoplasma gondii
infection in the mountain hare (Lepus timidus) and domestic rabbit (Oryctolagus cuniculus). II.
Early immune reactions. J Comp Pathol 117:361–369
Gyimesi ZS, Lappin MR, Dubey JP (2006) Application of assays for the diagnosis of toxoplasmo-
sis in a colony of woolly monkeys (Lagothrix lagotricha). J Zoo Wildl Med 37:276–280
Hegab SM, Al-Mutawa SA (2003) Immunopathogenesis of toxoplasmosis. Clin Exp Med
3:84–105
Hessler JR, Woodard JC, Tucek PC (1971) Lethal toxoplasmosis in a woolly monkey. J Am Vet
Med Assoc 159:1588–1594
Hill WC (1960) Primates. Comparative anatomy and taxonomy IV. Cebidae Part A. University
Press, Edinburgh, p xxii, 523
Hill RD, Gouffon JS, Saxton AM, Su C (2011) Differential gene expression in mice infected with
distinct Toxoplasma strains. Infect Immun 80:968–974
286 J.L. Catão-Dias et al.
Hutchison WM, Dunachie JF, Siim JC, Work K (1970) Coccidian-like nature of Toxoplasma gon-
dii. Br Med J 1:142–144
Ingberman B, Stone AI, Cheida CC (2008) Gênero Saimiri (Voigt 1831). In: Reis NR, Perachi AL,
Andrade FR (eds) Primatas Brasileiros. Technical Books, Londrina, pp 41–46
Innes EA (1997) Toxoplasmosis: comparative species susceptibility and host immune response.
Comp Immunol Microbiol Infect Dis 20:131–138
Innes EA (2010) A brief history and overview of Toxoplasma gondii. Zoonoses Public Health 57:1–7
Izawa K (1978) Frog eating behavior of wild black-capped capuchin (Cebus apella). Primates
19:633–642
Izawa K (1979) Foods and feeding behavior of wild black-capped capuchin (Cebus apella).
Primates 20:57–76
Jamieson SE, Peixoto-Rangel AL, Hargrave AC, de Roubaix LA, Mui EJ, Boulter NR, Miller EN,
Fuller SJ, Wiley JS, Castellucci L, Boyer K, Peixe RG, Kirisits MJ, Elias LD, Coyne JJ, Correa-
Oliveira R, Sautter M, Smith NC, Lees MP, Swisher CN, Heydemann P, Noble AG, Patel D,
Bardo D, Burrowes D, McLone D, Roizen N, Withers S, Bahia-Oliveira LMG, McLeod R,
Blackwell JM (2010) Evidence for associations between the purinergic receptor P2X(7)
(P2RX7) and toxoplasmosis. Genes Immun 11:374–383
Johnson LL, Sayles PC (1997) Interleukin-12, dendritic cells, and the initiation of host-protective
mechanisms against Toxoplasma gondii. J Exp Med 186:1799–1802
Juan-Salles C, Prats N, Marco AJ, Ramos-Vara JA, Borras D, Fernandez J (1998) Fatal acute toxo-
plasmosis in three golden lion tamarins (Leontopithecus rosalia). J Zoo Wildl Med 29:55–60
Kierulff MCM, Raboy BE, Procopio de Oliveira P, Miller K, Passos FC, Prado F (2002) Behavioral
ecology of lion tamarins. In: Kleiman DG, Rylands AB (eds) Lion tamarins: biology and con-
servation. Smithsonian Institution Press, Washington, pp 157–187
Kinzey WG (1992) Dietary adaptations in the Pitheciinae. Am J Phys Anthropol 88:499–514
Klevar S (2007) Tissue cyst forming coccidia; Toxoplasma gondii and Neospora caninum as a
cause of disease in farm animals. Acta Vet Scand 49:S1
Lang C, Gross U, Luder CG (2007) Subversion of innate and adaptive immune responses by
Toxoplasma gondii. Parasitol Res 100:191–203
Lees MP, Fuller SJ, McLeod R, Boulter NR, Miller CM, Zakrzewski AM, Mui EJ, Witola WH,
Coyne JJ, Hargrave AC, Jamieson SE, Blackwell JM, Wiley JS, Smith NC (2010) P2X7
receptor-mediated killing of an intracellular parasite, Toxoplasma gondii, by human and murine
macrophages. J Immunol 184:7040–7046
Leite TN, Maja Tde A, Ovando TM, Cantadori DT, Schimidt LR, Guercio AC, Cavalcanti A, Lopes
FM, Da Cunha IA, Navarro IT (2008) Occurrence of infection Leishmania spp. and Toxoplasma
gondii in monkeys (Cebus apella) from Campo Grande, MS. Rev Bras Parasitol Vet 17(Suppl
1):307–310
Luder CG, Stanway RR, Chaussepied M, Langsley G, Heussler VT (2009) Intracellular survival of
apicomplexan parasites and host cell modification. Int J Parasitol 39:163–173
Machado FS, Aliberti J (2009) Lipoxins as an immune-escape mechanism. Adv Exp Med Biol
666:78–87
Malmasi A, Mosallanejad B, Mohebali M, Sharifian Fard M, Taheri M (2009) Prevention of shed-
ding and re-shedding of Toxoplasma gondii oocysts in experimentally infected cats treated with
oral Clindamycin: a preliminary study. Zoonoses Public Health 56:102–104
Marshall LG (1988) Land mammals and the Great American Interchange. Am Sci 76:380–388
Mason NJ, Fiore J, Kobayashi T, Masek KS, Choi Y, Hunter CA (2004) TRAF6-dependent
mitogen-activated protein kinase activation differentially regulates the production of interleu-
kin-12 by macrophages in response to Toxoplasma gondii. Infect Immun 72:5662–5667
Mendes SL, Silva MP, Strier KB (2010) O Muriqui. Vitória, ES, Brasil, Instituto de Pesquisas da
Mata Atlântica-IPEMA. 95 p
Miller CM, Zakrzewski AM, Ikin RJ, Boulter NR, Katrib M, Lees MP, Fuller SJ, Wiley JS, Smith
NC (2011) Dysregulation of the inflammatory response to the parasite, Toxoplasma gondii, in
P2X(7) receptor-deficient mice. Int J Parasitol 41:301–308
Neotropical Primates and Their Susceptibility to Toxoplasma gondii… 287
Scanga CA, Aliberti J, Jankovic D, Tilloy F, Bennouna S, Denkers EY, Medzhitov R, Sher A
(2002) Cutting edge: MyD88 is required for resistance to Toxoplasma gondii infection and
regulates parasite-induced IL-12 production by dendritic cells. J Immunol 168:5997–6001
Seed JR (1996) Protozoa: pathogenesis and defenses. In: Baron S (ed) Medical microbiology.
University of Texas Medical Branch at Galveston, Galveston, TX
Sharma SD (1990) Immunology of toxoplasmosis. In: Wyler DJ (ed) Modern parasite biology:
cellular, immunological and molecular aspects. W.H. Freeman and Company, New York, NY,
pp 184–199
Sibley LD, Khan A, Ajioka JW, Rosenthal BM (2009) Genetic diversity of Toxoplasma gondii in
animals and humans. Philos Trans R Soc Lond B Biol Sci 364:2749–2761
Silva JCR (2007) Toxoplasmose. In: Cubas ZS, Silva JCR, Catão-Dias JL (eds) Tratado de animais
selvagens: medicina veterinária. Editora Roca, São Paulo, pp 768–784
Silva JS Jr (2001) Especiação nos macacos-prego e caiararas, gênero Cebus Erxleben, 1777
(Primates, Cebidae). Ph.D. Thesis, Universidade Federal do Rio de Janeiro, Rio de Janeiro
Simpson GG (1980) Splendid isolation: the curious history of South American mammals. Yale
University, New Haven, CT
Snowdon CT, Soini P (1988) The tamarins, genus Saguinus. In: Mittermeier RA, Rylands AB,
Coimbra-Filho AF, Fonseca GAB (eds) Ecology and behavior of neotropical primates. World
Wildlife Fund, Washington, DC, pp 223–298
Sogorb S F, Jamra LF, Guimaraes EC, Deane MP (1972) Toxoplasmose espontanea em animais
domesticos e silvestres, em Sao Paulo. Revista. Inst Med trop S Paulo 14: 314–320
Soini P (1988) The pygmy marmoset, genus Cebuella. In: Mittermeier RA, Rylands AB, Coimbra-
Filho AF, Fonseca GAB (eds) Ecology and behavior of neotropical primates. World Wildlife
Fund, Washington, DC, pp 79–129
Splendore A (1909) A new protozoan parasite of rabbit found in histological lesions similar to
human Kala-Azar. Rev Soc Sci S Paulo 3:109–112
Stuart M, Pendergast V, Rumfelt S, Pierberg S, Greenspan L, Glander K, Clarke M (1998) Parasites
of wild howlers (Alouatta spp.). Int J Primatol 19:493–512
Subauste CS (2009) CD40, autophagy and Toxoplasma gondii. Mem Inst Oswaldo Cruz
104:267–272
Suzuki Y, Orellana MA, Schreiber RD, Remington JS (1988) Interferon-gamma: the major media-
tor of resistance against Toxoplasma gondii. Science 240:516–518
Tait ED, Hunter CA (2009) Advances in understanding immunity to Toxoplasma gondii. Mem Inst
Oswaldo Cruz 104:201–210
Terborgh J (1983) Five New world primates. Princeton University Press, Princeton, NJ
Van Roosmalen MGM, Van Roosmalen T (2003) The description of a new marmoset genus,
Callibella (Callitrichinae, Primates), including its molecular phylogenetic status. Neotrop
Primates 11:1–10
Villarino AV, Stumhofer JS, Saris CJ, Kastelein RA, de Sauvage FJ, Hunter CA (2006) IL-27 limits
IL-2 production during Th1 differentiation. J Immunol 176:237–247
Vouldoukis I, Mazier D, Moynet D, Thiolat D, Malvy D, Mossalayi MD (2011) IgE mediates kill-
ing of intracellular Toxoplasma gondii by human macrophages through CD23-dependent,
interleukin-10 sensitive pathway. PLoS One 6:e18289
Webb SD (1976) Mammalian faunal dynamics of the Great American Interchange. Paleobiology
2:220–234
Werner H, Janitschke K, Kijhler H (1969) Uber Beobachtungen an Marmoset-Affen Saguinus
(Oedipomidas) Oedipus nach oraler und intraperitonealer infektion mit verschiedenen en zys-
tenbildenden Toxoplasma-Stgmmen unterschiedlicher Virulenz. I. Mitteilung: Klinische,
pathologisch-anatomische, histologische und parasitologische zentralblatt Bakteriologie
Befunde. Parasitenkunde fir Infektionskrankheiten und Hygiene 209:553–569
Wright PC (1981) The night monkeys, genus Aotus. In: Coimbra-Filho AF, Mittermeir RA (eds)
Ecology and behavior of neotropical primates. Academia Brasileira de Ciências, Rio de
Janeiro, pp 211–240
Neotropical Primates and Their Susceptibility to Toxoplasma gondii… 289
Yarovinsky F, Zhang DK, Andersen JF, Bannenberg GL, Serhan CN, Hayden MS, Hieny S,
Sutterwala FS, Flavell RA, Ghosh S, Sher A (2005) TLR11 activation of dendritic cells by a
protozoan profilin-like protein. Science 308:1626–1629
Zanon CMV, Reis NR, Filho HO (2008) Gênero Ateles E.Geoffroy 1806. In: Reis NR, Perachi AL,
Andrade FR (eds) Primatas brasileiros. Technical Books, Londrina, pp 169–173
The Evolution of SIV in Primates and
the Emergence of the Pathogen of AIDS
Introduction
Three decades have passed since the first reports of opportunistic diseases in
previously healthy individuals and the description of acquired immunodeficiency
syndrome (AIDS) as a new human disease (CDC 1981a, b, c). Shortly after, the
causative agent was identified as a T-cell tropic retrovirus (Barre-Sinoussi et al.
1983) that eventually came to be termed human immunodeficiency virus 1 (HIV-1).
HIV-1 is a retrovirus of the Lentivirus genus and the first primate lentivirus to be
discovered. A second human lentivirus, HIV-2, was later discovered in patients
suffering from AIDS in West Africa (Clavel et al. 1986).
Lentiviruses have since been discovered to naturally infect over 40 different
African primate species, termed simian immunodeficiency viruses (SIVs). Study of
SIV infection of the majority of these species is difficult, as many are endangered
and (rightly) protected in the wild, with limited or no captive populations available
for study. However, some important facts have been established. Firstly, it is now
clear that HIV-1 has originated from SIVcpz of the common chimpanzee (Pan trog-
lodytes) and HIV-2 from SIVsmm of the sooty mangabey (Cercocebus atys).
Secondly, some species of SIV-infected African primates are present in European
and US research centers, and thus, the natural history of their infection has been
studied in detail. In particular, the SIV infection of two African green monkey spe-
cies (Chlorocebus sabaeus and C. pygerythrus) and sooty mangabeys has been stud-
ied intensively. In these species, it is clear that the vast majority of individuals do
not progress to AIDS.
In contrast, Asian macaques are not infected with SIV in the wild, but can
develop AIDS when experimentally infected with SIV from other species,
J.F. Brinkworth and K. Pechenkina (eds.), Primates, Pathogens, and Evolution, 291
Developments in Primatology: Progress and Prospects,
DOI 10.1007/978-1-4614-7181-3_10, © Springer Science+Business Media New York 2013
292 E.J.D. Greenwood et al.
providing a model for disease in humans caused by HIV. Rhesus macaques (Macaca
mulatta) and pig-tailed macaques (Macaca nemestrina) are the two species princi-
pally used in these studies. These models were first established after macaques in
numerous American primate centers developed AIDS-like clinical signs and were
also found to be infected with T-cell tropic retroviruses, which collectively came to
be termed SIVmac (Benveniste et al. 1986; Daniel et al. 1984). Due to the similarity
between SIVmac and SIVsmm, it was hypothesized early on that SIVmac could
have its origins in SIVsmm infection of sooty mangabeys (Murphey-Corb et al.
1986), which has since been confirmed. It is likely that SIVsmm was unknowingly
transmitted from sooty mangabeys into macaques during invasive experiments for
the study of prion diseases in American primate centers and subsequently spread
within captive macaque populations (reviewed by Apetrei et al. 2006).
Comparison of the pathogenic infection of humans and macaques with HIV/SIV
with the nonpathogenic infection of sooty mangabeys and African green monkeys
has provided key insight into the pathways most important in the development of
AIDS in susceptible species and the host mechanisms that have evolved in African
primate species to avoid disease as a result of lentivirus infection.
In this chapter, we will first discuss the age and diversity of primate/human len-
tiviruses, the outcome of SIV transmission into humans, and the mechanisms pro-
posed to have promoted the pandemic spread of HIV-1. Next, we will compare the
natural history of pathogenic HIV infection of humans and SIV infection of Asian
macaques (the best available animal model of human HIV/AIDS) with the non-
pathogenic SIV infection of sooty mangabeys and African green monkeys. We will
discuss in depth the mechanisms that have been proposed to explain the dichoto-
mous outcome of lentivirus infection between these groups. We will then examine
the specific differences between HIV-1 and other primate lentiviruses, including
potential mechanisms for the high pathogenicity of HIV-1. Finally, we will review
what is known of the host–virus relationship in the species in which the HIV-1 lin-
eage evolved and suggest that examination of this relationship is of special impor-
tance to HIV-1 and SIV research.
SIVs observed in wild African primates are generally species-specific: multiple iso-
lates of virus from one primate species generally form monophyletic lineages in
phylogenetic trees. The degree to which SIVs differ within one species varies, but is
largely biased by the number of isolates sequenced and their geographical distribu-
tion (Bibollet-Ruche et al. 2004; Liegeois et al. 2012). Species with highly diver-
gent SIVs have also been observed, which are usually the result of cross-species
transmissions, sometimes followed by recombination between distant SIVs
(Aghokeng et al. 2007; Liegeois et al. 2012; Souquiere et al. 2001). High frequen-
cies of recombination are a characteristic of primate lentiviruses (Chen et al. 2006),
but such recombination events are most apparent when heterologous viruses
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 293
recombine. These so-called mosaic or chimeric viruses are described for African
green monkeys, mandrills (Mandrillus sphinx), and chimpanzees (Bailes et al.
2003; Jin et al. 1994; Takemura and Hayami 2004) and can be identified when com-
paring phylogenetic trees created using alignments from different parts of the
genome, as in Fig. 1.
A precise age of primate lentiviruses has been difficult to ascertain, but has been
estimated in several studies. Such dating methods normally require the use of a
“molecular clock,” in which sequences are compared, and a known or estimated rate
of genetic change is applied to estimate the time to the most recent common ances-
tor. Estimates resulting from molecular clock methods are dependant on how this
rate of change is estimated. While this is relatively simple for eukaryotic species, as
the rate of genetic change is both slow and well established for different species,
calibrating a molecular clock for retroviruses, which change extremely rapidly and
often recombine, is much more challenging.
Using only relatively modern SIV and HIV sequences of known dates (from
1975 to 2005) to calibrate a molecular clock resulted in a estimate that the primate
lentivirus lineage is only centuries old (Wertheim and Worobey 2009). However,
this dating was controversial as it was already suspected that using only modern
sequences to extrapolate the history of a possibly ancient lineage would lead to
erroneous estimates (Sharp et al. 2000).
However, a recent study of SIV infection of primates on the African island of
Bioko has allowed for a new calibration of the molecular clock estimate of the age
of this lineage (Worobey et al. 2010). The island has been separated from the
African mainland for a period of 10,000–12,000 years and accommodates a num-
ber of primate species. Individuals from four of these species were found to be
infected with SIV. Most importantly, the Bioko drill (Mandrillus leucophaeus
poensis) is infected with an SIV similar to that isolated from the mainland drill
(Mandrillus leucophaeus leucophaeus). The time to the most recent common
ancestor of SIVdrl from Bioko and SIVdrl from mainland Africa is therefore
known to be at least 10,000 years old, providing a new method for calibrating the
molecular clock. The resulting estimate is that SIVs have been present in African
primates for 76,000 years.
Finally, evidence exists that the primate lentivirus lineage is ancient. The
genomes of a number of lemur species of genera Cheirogaleus and Microcebus
contain sequences of a lentivirus that has at some point infected germ line cells and
become integrated into the genome—an endogenous lentivirus. After integration
into the genome, the sequences of endogenous retroviruses are expected to be sub-
ject to the same rate of mutation as other host genomic sequences. This rate is well
established for eukaryotic species and is much slower than the rate of mutation of
exogenous retroviruses. Endogenous retroviruses are therefore ideal for estimating
dates on ancient timescales. It seems that there were two independent integration
events, both estimated to have occurred around 4 million years ago (Gifford et al.
2008; Gilbert et al. 2009). Unless SIV was introduced to Madagascar independently
and prior to the introduction of SIV to the African mainland, this would indicate that
the African primate lentivirus lineage is at least equally as ancient.
294 E.J.D. Greenwood et al.
Fig. 1 Phylogenetic trees demonstrating the relationships between different SIVs and HIV-1 and
HIV-2. Trees were created from alignments using nucleic acid sequences from the (a) gp41 and
(b) protease genes. Scale bar indicates 0.9 substitutions per site. SIVagmSab is excluded from the
protease tree as a recombination event has occurred in the region used for this alignment
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 295
HIV/SIV are retroviruses of the Lentivirus genus. The genomes of all retroviruses
include the genes gag, pol, and env, the major structural and enzymatic proteins of
the virus. All known members of the Lentivirus genus also possess the regulatory
proteins Rev and Tat, including the oldest recognized lentivirus, “RELIK,” an
endogenous retrovirus estimated to have integrated into the genome of the European
rabbit (Oryctolagus cuniculus) over 12 million years ago (Katzourakis et al. 2007;
Keckesova et al. 2009). All extant lentiviruses found in primates, cats, sheep, and
cattle possess an additional gene, vif.
In contrast, four genes are unique to extant primate lentiviruses; nef and vpr are
found in all primate lentiviruses, while vpx and vpu are found in nonoverlapping
subsets of these viruses. Our understanding of the functions of vif, nef, vpr, vpx, and
vpu has increased dramatically during recent years, and it can be concluded that all
factors have evolved to play a role in counteracting host defense mechanisms called
restriction factors (Ayinde et al. 2010; Kirchhoff 2010).
Vpx, one of the two genes unique to primate lentiviruses, evolved in the Papionini
tribe of primates, and the distribution of this gene within primate SIVs is likely to
have been increased as a result of several recombination events between SIVs of
different primate species (Takemura and Hayami 2004). It is found in SIVsmm (of
sooty mangabeys), SIVrcm (red-capped mangabey, Cercocebus torquatus),
SIVmnd-2 (mandrill), and SIVdrl (drill). The other gene unique to primate lentivi-
ruses, vpu, was acquired by an SIV within a subset of guenon species (Bailes et al.
2003). Guenons (tribe Cercopithecini) are a species-rich group of primates; how-
ever, only four guenon species, which associate in the wild, carry SIVs with the vpu
gene. These are the mona monkey (Cercopithecus mona), greater spot-nosed mon-
key (C. nictitans), and mustached monkey (C. cephus), infected with SIVmon,
296 E.J.D. Greenwood et al.
There are four recognized HIV-1 groups: M, N, O, and P. As each of these HIV-1
groups is closest in sequence homology to different isolates of SIV in chimpanzees
or gorillas, it is clear that there have been at least four cross-species transmission
events into humans, with M and N group viruses having their source in chimpanzees
of the Pan troglodytes troglodytes subspecies and O and P group most closely
related to SIV of gorillas (Keele et al. 2006; Plantier et al. 2009; Van Heuverswyn
et al. 2006). To prevent confusion, it should also be noted that there is a separate
alphabetical nomenclature for describing different subtypes (also referred to as
clades) of HIV-1 group M viruses. Based on sequence homology, HIV-1 group M
viruses are assigned to subtypes from A to K or identified to have been generated
through recombination between viruses of previously identified subtypes.
Non-M HIV-1 viruses are mostly restricted to Cameroon. Group O infection
accounts for around 1 % of all HIV-1 infection in Cameroon (Vergne et al. 2003),
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 297
while N has been identified in less than 20 individuals (Vallari et al. 2010a) and
group P in only two cases so far (Plantier et al. 2009; Vallari et al. 2010b). Clinical
data regarding non-M infections is limited, but it is important to note that group O
and group N infections have been identified in patients with AIDS (Ayouba et al.
2000; Gurtler et al. 1994), while the first group P infection to be identified was asso-
ciated with depleted CD4+ T cells and a high viral load (Plantier et al. 2009). There
is therefore no evidence that HIV-1 non-M infections are less pathogenic than HIV-1
M infections, despite their more limited spread.
In addition to HIV-1, there is a second human immunodeficiency virus, HIV-2.
In contrast to HIV-1, HIV-2 infects globally only between 1 and 2 million individu-
als and is predominantly confined to West Africa (Campbell-Yesufu and Gandhi
2011). In common with HIV-1, HIV-2 consists of several groups, each the result of
a different cross-species transmission event. Eight groups, A–H, have been recog-
nized, with only groups A and B having spread more extensively within the human
population. The remaining groups are limited to infections of single individuals.
HIV-2 is the result of transmission of SIV from the sooty mangabey, and because of
this different source, HIV-2 has a different genetic structure to HIV-1. Viral genomes
of several primate lentiviruses, along with HIV-1 and HIV-2, are compared in Fig. 2.
There are critical differences between HIV-1 and HIV-2 infection of humans that are
discussed at the end of this chapter.
As with dating the age of the primate lentivirus lineage, dating the cross-species
transmission events leading to HIV has been difficult. However, the availability of
two HIV-1 group M sequences attained from archived samples from 1959 and 1960,
from Kinshasa (formally Leopoldville), Democratic Republic of Congo (DRC, for-
mally Belgian Congo and, later, Zaire), allows for greater confidence in dating the
origin of the HIV-1 group M pandemic. While the chimpanzees infected with
SIVcpz isolates most closely related to HIV-1 group M are found in Cameroon, it
seems that Kinshasa is a candidate for the epicenter of the HIV-1 pandemic. A study
conducted in 1997 indicated that there is unparalleled genetic diversity of HIV-1 M
in the DRC, with all subtypes represented, along with recombinant forms not repre-
sented outside of the DRC (Vidal et al. 2000). Notably, HIV-1 extracted from the
two samples from 1959 and 1960 show a high degree of genetic divergence—com-
parable to the genetic difference between two contemporary isolates of different
subtypes. This suggests that by 1960, HIV-1 had already circulated extensively in
humans in this region (Worobey et al. 2008).
Using sequences from these two early samples, along with later samples to cali-
brate a molecular clock, leads to an estimate that the cross-species transmission
event leading to HIV-1 group M most likely occurred near the start of the twentieth
century, between 1873 and 1924 (Worobey et al. 2008). Dating the cross-species
transmission events that resulted in the other HIV-1 and HIV-2 groups is more prob-
lematic due to the markedly fewer available sequences and the lack of sequences
from older achieved material. Estimates for the date of the four cross-species trans-
mission events leading to HIV-1 groups O and N and HIV-2 groups A and B are all
also within or near the beginning of the twentieth century (de Sousa et al. 2010;
Wertheim and Worobey 2009).
298 E.J.D. Greenwood et al.
Fig. 2 Annotated depiction of the genomes of HIV-1, HIV-2, and relevant SIVs. Diagrams are
based on annotated sequences, with the sequence used named in brackets. Numbers 1–3 at left
indicate reading frames
Firstly, Worobey and colleagues have suggested that the development of large
urban centers, not present in western Central Africa prior to 1900, facilitated the
spread of HIV-1 M. They assert that prior to 1910, there was not a single site with a
population greater than 10,000 people in western Central Africa (Worobey et al.
2008). Kinshasa underwent particularly rapid growth, from only a few thousand
people in 1905, to around 40,000 in 1940, and over 400,000 in 1961 (Chitnis et al.
2000).
However, this rapid urbanization, combined with the (often-forced) movement of
workers, led to disruption of social norms that may have been a greater contributing
factor than the size of the urban centers alone. De Sousa and colleagues carried out
extensive analysis of colonial medical articles and reports regarding Kinshasa from
the start of the twentieth century to the country’s independence (and general expul-
sion of Belgian authorities) in 1960 (Chin 2007; de Sousa et al. 2010). They note
that in 1929 there were a large number of commercial sex workers, possibly due to
the heavily male biased population, at 4:1 males to females. They also find that
sexually transmitted diseases, including syphilis and other genital ulcerative dis-
eases, were highly prevalent, with one survey in 1930–1932 finding that 5 % of the
female population had active genital ulcers (de Sousa et al. 2010). While the low
risk of HIV-1 transmission from heterosexual sex, estimated to be between 0.01 and
1.1 % per act (Boily et al. 2009), could be viewed as a barrier to early spread of
HIV-1 through purely sexual contact, the risk of transmission is considerably
increased when genital ulcers (and other non-ulcerative sexually transmitted dis-
eases) are present in either the HIV-positive or HIV-negative sexual partner, with
most studies finding a 2–5-fold increase in risk of infection, but some finding the
risk of transmission to be almost 20-fold higher (Boily et al. 2009; Fleming and
Wasserheit 1999).
The role of non-sterile injections throughout sub-Saharan Africa in the twentieth
century has also been proposed as a major factor allowing the early spread of HIV-1.
Marx and colleagues propose that initial “serial human passage” of the virus through
non-sterile injections allowed the adaptation to humans which they suggest would
be required for subsequent spread by sexual transmission (Marx et al. 2001). They
point to numerous mass injection campaigns in sub-Saharan Africa and evidence
for large numbers of injections involving the reuse of non-sterilized injection mate-
rials in the twentieth century, especially since the discovery of penicillin and its use
in Africa from the 1950s onwards.
The mechanisms reviewed above that are proposed to have facilitated the spread
of HIV-1 span the entirety of the twentieth century. Interestingly, in their 2008 anal-
ysis, Worobey et al. propose a model of the growth of HIV-1 that suggests relatively
slow growth prior to 1960, followed by much more rapid expansion after this date.
Although their analysis is likely to be subject to bias due to the limited availability
of samples available, this computational analysis has some support from the early
reports of AIDS in Africa, as it correlates with marked increases in the opportunistic
diseases that now define AIDS, such as Kaposi’s sarcoma and esophageal candidia-
sis, occurring in Kinshasa, Uganda, Zambia, and Rwanda from the late 1970s to the
early 1980s (Quinn et al. 1986). This would be consistent with aggressive spread of
300 E.J.D. Greenwood et al.
HIV-1 in the 1960s and early 1970s, due to the approximately decade-long incuba-
tion period between HIV-1 infection and the development of AIDS (see below).
Together, these lines of evidence tend not to favor extensive spread in urban cen-
ters in the early part of the twentieth century. May and colleagues have proposed an
elegant model of rural spread among “loosely interlinked villages” (May et al.
2001). One significant attraction of this model is that it predicts that HIV-1 could
persist at a very low level for decades within rural villages, with only limited spread
within each village compensated for by introduction to new villages, preventing the
virus from becoming extinct. In this model, the period of extremely slow spread is
then followed by much more rapid expansion of HIV, without requiring mecha-
nisms postulated above. Of course, this does not exclude a role for the other factors
reviewed, especially as urbanization, high levels of other sexually transmitted dis-
eases, and the reuse of non-sterilized injection materials could all have contributed
to the introduction of HIV-1 to urban centers and extensive spread within these
centers—which was almost certainly a requirement for HIV dissemination through-
out and beyond Africa.
While HIV-1 group M subtype B causes only a minority of HIV-1 infections
in Africa, it is the most prevalent subtype in a large number of countries outside
of Africa, including the USA and Europe. The disparity between the prevalence
of subtype B outside and within Africa could indicate that the pandemic beyond
Africa was the result of a single transmission from Africa. Gilbert and col-
leagues examined sequences from archived samples collected in 1982 and 1983
of Haitian nationals that had immigrated to the USA after 1975 and hospitalized
with AIDS prior to 1981 (Gilbert et al. 2007). They focused on these patients as
shortly after the recognition of AIDS as a novel syndrome firstly primarily in
homosexual men in the USA in 1981 (Gottlieb et al. 1981), and intravenous
drug users in 1982 (CDC 1982b), there were reports of Haitian nationals in the
USA suffering from AIDS (CDC 1982c). The majority of these Haitian patients
indicated they had practiced neither homosexual sex nor intravenous drug use,
suggesting that being a recent Haitian immigrant was an independent risk factor
for the development of AIDS.
In their analysis of sequences from these patients, in addition to other sequences
of HIV-1 isolates from Haiti, and sequences representing HIV-1 subtype B isolates
from multiple other countries, Gilbert and colleagues find that subtype B is the most
genetically diverse within Haiti, indicating an earlier introduction of HIV-1 subtype
B into this nation than any other, and that the HIV-1 epidemic in Haiti was the result
of a single introduction of the virus from Africa. They also show that with few
exceptions, the pandemic spread of HIV-1 subtype B beyond Haiti (including North
American, South American, European, and Asian countries) is the result of another
single founder event linked to Haiti, which they postulate to be most likely the intro-
duction of HIV-1 to the USA through immigration.
The authors estimate that the introduction of HIV-1 from Africa into Haiti
occurred between 1962 and 1970. Interestingly, it seems that there was a large
movement of Haitians to the DRC (then Zaire) in the early 1960s following two
events: firstly, political problems appearing in Haiti after Francois Duvalier
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 301
(‘Papa Doc’) took power in 1957 and, secondly, the vacancies for “executive”
positions (administrators, healthcare workers, and teachers) in Zaire following
the expulsion of the Belgian authorities in 1960. Many Haitians returned to Haiti
or to other nations in the late 1960s and early 1970s (Chin 2007; Molez 1998; Piot
et al. 1984), possibly indicating a potential mechanism by which HIV-1 was
introduced into Haiti. Gilbert et al. also estimate that the single founder event
leading to the spread of HIV-1 subtype B from Haiti occurred between 1966 and
1972 (Gilbert et al. 2007), which is constant with the earliest cases of AIDS in the
USA, retrospectively diagnosed as occurring in 1978 (CDC 1982a).
The natural history of the SIV infection of African green monkeys or sooty
mangabeys with SIVagm and SIVsmm, respectively, is very different. These ani-
mals generally do not develop disease despite peripheral blood viral loads similar to
those of humans and rhesus macaques. A recent study of sooty mangabeys has
demonstrated that SIV infection does result in an extremely slow depletion of CD4+
T cells (Taaffe et al. 2010), but unlike in humans and rhesus macaques, the magni-
tude of CD4+ T-cell loss does not correlate with viral load. CD4+ T-cell loss has not
been reported in African green monkeys. However, in both African green monkeys
and sooty mangabeys, measurements of CD4+ T cells are complicated by factors
relating to expression of CD4 (see below). For both of these species, only a single
case of an AIDS-like condition in an SIV-infected animal has been described (Ling
et al. 2004; Traina-Dorge et al. 1992). The mechanisms by which these species cir-
cumvent disease development are of great interest and are discussed below.
Entry of HIV-1 into the cell is mediated by the use of CD4 as a receptor, along with
a coreceptor. CD4 is expressed principally on CD4+ T cells, and it binds to MHC-II
on antigen-presenting cells. These T cells, upon activation, become “T helper cells,”
which direct the adaptive immune response primarily through cytokine production.
The coreceptor used by HIV-1 is principally the chemokine receptor C-C chemo-
kine receptor type 5 (CCR5), but in some cases a different chemokine receptor,
C-X-C chemokine receptor type 4 (CXCR4). Interestingly, humans lacking CCR5
expression through an inherited mutation are almost completely resistant to infec-
tion by both sexual and parenteral routes (Dean et al. 1996; Wilkinson et al. 1998).
Partly because of the restriction of CCR5 expression to memory T cells, memory
CD4+ T cells are preferentially infected compared to naïve CD4+ T cells. In humans
and rhesus macaques, quantitative PCR detection of the HIV-1 genome in different
CD4 T-cell subsets consistently finds that the majority of infected cells have a mem-
ory CD4+ cell phenotype, though infected naïve T cells are also found (Brenchley
et al. 2004a; Mattapallil et al. 2005; Ostrowski et al. 1999). Macrophages and den-
dritic cells, the primary antigen-presenting cells of the immune system, also express
CD4 and CCR5 and can be productively infected by HIV-1. An obvious potential
mechanism by which species may adapt to SIV infection would therefore be to
restrict the expression of the viral receptor and coreceptor.
Interesting observations have been made regarding CD4 expression in sooty
mangabeys and African green monkeys. Adult uninfected African green monkeys
have a much lower frequency of peripheral blood CD4+ T cells than other primates,
including humans, sooty mangabeys, and rhesus macaques (Beaumier et al. 2009).
They have a large peripheral blood T-cell population with a unique phenotype of
CD4-CD8αdim. Interestingly, it seems that while naïve T cells express normal levels
of CD4, activation and development of a subset of these into a memory phenotype
is associated with down-modulation of CD4 and upregulation of CD8α. Curiously,
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 303
these cells retain the functionality of T helper cells and are restricted to antigen
presented on MHC-II, despite the lack of CD4 expression (Beaumier et al. 2009).
As in humans and rhesus macaques, CD4+ T cells with a memory phenotype make
up the majority of SIVagm-infected CD4+ T cells. However, while the limited num-
ber of memory T cells that maintain CD4 expression has the highest level of infec-
tion, the CD4-CD8αdim population contains relatively low numbers of infected cells.
Thus, the loss of CD4 expression in differentiation from naïve to memory appar-
ently protects a subset of CD4+ memory T cells from infection.
Similarly, a large population of CD4-CD8- (“double-negative”) peripheral blood
T cells has been identified in both infected and uninfected sooty mangabeys (Milush
et al. 2011). These cells were first identified when it was discovered that some
strains of SIVsmm are in fact capable of causing very rapid and severe loss of CD4+
T cells in sooty mangabeys, to levels that would be associated with AIDS in humans
and macaques. Despite this, these animals maintain a competent immune system,
and this led to the discovery that the CD4–CD8- T cells seem to be capable of per-
forming functions normally associated with CD4+ T cells. In both species, the seg-
regation of functions normally carried out by CD4+ T cells to cells resistant to
infection means that regardless of the extent of CD4+ T-cell depletion, a minimum
level of immune function can be maintained.
Expression of the coreceptor CCR5 has also been examined in primate species,
in the context of SIV infection. One report has compared the frequency of CCR5+
CD4+ T cells in the blood and lymph nodes of uninfected individuals of multiple
primate species—comparing those species that are known to be SIV infected with
the wild (including African green monkeys and sooty macaques) with humans and
rhesus macaques. This report demonstrates that sooty mangabeys, African green
monkeys, and numerous other African species that are host to SIV have a much
lower frequency of CD4+ T cells that express CCR5 (Pandrea et al. 2007a). The low
frequency of CCR5 expression has therefore been assumed to be convergent evolu-
tion by SIV-exposed species to restrict target cell availability.
A recent study has further demonstrated that upon in vitro stimulation, naïve T
cells from sooty mangabeys show more restricted CCR5 upregulation compared to
naïve T cells from rhesus macaques. This failure to upregulate CCR5 is especially
pronounced in cells that take on a central memory phenotype after activation. This
may protect central memory T cells from infection; as in sooty mangabeys, this
population shows a much lower frequency of infection than the Tcm population of
rhesus macaques (Paiardini et al. 2011). However, it should be stressed that in this
comparison, different viral isolates were used to infect the two different species, and
the possibility of minor differences in viral tropism cannot be excluded.
Further examination of sooty mangabeys has demonstrated a relatively high fre-
quency of CCR5 deletion mutant alleles, with 8 % homozygous for CCR5 deletions
in a large captive population (Riddick et al. 2010). Interestingly, in contrast to HIV-
1-infected humans, these animals are not resistant to SIV infection. In the captive
population studied, prevalence of naturally acquired SIV infection was similar in
the CCR5 deletion homozygous animals as in heterozygous animals or those with
two functional CCR5 alleles (Riddick et al. 2010). Viruses isolated from both
304 E.J.D. Greenwood et al.
homozygous CCR5 deletion animals and those expressing CCR5 were able to use a
variety of additional coreceptors, including CXCR6, GPR15, and GPR1, in addition
to maintaining CCR5 utilization. Similarly, in red-capped mangabeys, two studies
have shown a high frequency of one CCR5 deletion (CCR5 Δ24), with approxi-
mately 60–70 % of animals being homozygous for this deletion (Beer et al. 2001;
Chen et al. 1998). Cells from red-capped mangabeys with the homozygous CCR5
deletion were completely resistant to infection by CCR5-using viruses from other
species. Interestingly, SIV of red-capped mangabeys has apparently adapted to this
selection pressure, as this virus uses principally CCR2b as a coreceptor and cannot
use CCR5 (Beer et al. 2001; Chen et al. 1998).
The clinical picture of HIV-1 infection of humans suggests that damage to the
immune system is a chronic, gradual process, resulting in the cumulative immuno-
logical collapse that allows the onset of opportunistic disease. In contrast, a number
of studies suggest that a great deal of irreversible damage occurs early in infection
in the secondary lymphoid organs and gut-associated immune system.
The gut-associated lymphoid system (GALT) can be divided into inductive sites
(gut-draining mesenteric lymph nodes and Peyer’s patches), at which T-cell
responses are generated through interaction with antigen-presenting cells and effec-
tor sites, primarily lymphocytes residing in the lamina propria (lamina propria lym-
phocytes, LPLs) and between the epithelial cells of the mucosal surface
(intraepithelial lymphocytes, IELs). There are a large number of CD4+ T cells in the
Peyer’s patches and the lamina propria, with expression of CCR5 at a much higher
frequency in these compartments than CD4+ T cells in the peripheral blood or other
secondary lymphoid tissues.
Several studies have demonstrated a dramatic depletion of CD4+ T cells in one
or more of these compartments from the very earliest stage of HIV-1 infection, with
little or no recovery of these cells in the chronic phase of infection (Brenchley et al.
2004b; Estes et al. 2008; Guadalupe et al. 2003; Mehandru et al. 2004). The result
of this T-cell loss has been the topic of much discussion in the field. Some authors
have stated that the gut mucosal immune system contains the majority of lympho-
cytes in the body (Guadalupe et al. 2003; Mehandru et al. 2004; Veazey et al. 1998),
with some further stating that the acute loss of CD4+ T cells in the gut therefore
“reflects the loss of most CD4+ T-cells in the body” (Brenchley and Paiardini 2011).
However, available literature on this subject (Ganusov and De Boer 2007) finds that
it is more likely that at most around 20 % of all lymphocytes are resident in the
human gut, with a similar percentage of the total CD4+ T cells found there.
Despite this, there does seem to be at least one important consequence of damage
to the mucosal immune system by HIV-1 infection, as it appears to result in the loss
of integrity of the mucosal barrier. This allows translocation of bacteria and
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 305
bacterial products and their systemic dissemination, which in turn causes the pro-
duction of pro-inflammatory cytokines by cells of the innate immune system. The
level of bacterial translocation, as measured by concentration of lipopolysaccharide
(LPS) in the peripheral blood, is significantly upregulated in chronically infected
HIV-1 patients when compared to uninfected controls in European and US cohorts
(Brenchley et al. 2006). LPS translocation also shows a significant correlation with
the level of peripheral blood activated CD8+ T cells, as measured by expression of
the markers CD38 and HLA-DR (Brenchley et al. 2006). This acute damage to the
gut mucosa therefore seems to be a major contributing factor to the chronic immune
activation that is thought to drive the progression from chronic HIV-1 infection to
AIDS (see below). The damage to the gut mucosal immune system in SIVmac-
infected rhesus macaques is similar to humans, with profound depletion of CD4+ T
cells occurring within weeks of infection at multiple sites of the GI tract and
increased levels of plasma LPS (Brenchley et al. 2006; Ling et al. 2007; Mattapallil
et al. 2005; Veazey et al. 1998).
The importance of LPS translocation in HIV-1 pathogenesis, supported by data
from HIV-1+ patients in the USA and Europe, is not entirely supported by two
studies carried out in sub-Saharan Africa. One study finds no difference in LPS
levels between patient samples taken from pre-infection to AIDS (Redd et al.
2009), while another finds a significant difference only between uninfected
patients and patients that have already progressed to AIDS—not earlier in the
disease process (Nowroozalizadeh et al. 2010). In this second study, there is how-
ever a nonsignificant trend for increased plasma LPS in HIV-1-positive patients
compared to negative control individuals and a significant inverse correlation
between peripheral blood CD4+ T-cell counts and plasma LPS. However, these
findings would be compatible with a theory that LPS translocation is allowed to
occur due to loss of CD4+ T cells by other mechanisms and becomes progres-
sively worse throughout the course of disease—rather than necessarily indicating
that LPS translocation is a driving factor in immune activation and loss of CD4+
T cells from the outset of infection.
However, the importance of loss of the integrity of the gut mucosal barrier in
pathogenic lentivirus infections is emphasized by the different outcome in African
nonhuman primates. Surprisingly, two studies have demonstrated that in nonpatho-
genic infection of African green monkeys and sooty mangabeys, there is also mas-
sive depletion of gut mucosal T cells, comparable with depletion seen in
SIVmac-infected rhesus macaques (Gordon et al. 2007; Pandrea et al. 2007b).
Recovery of CD4+ T cells in this compartment after acute infection is variable, with
some animals restoring numbers of cells to near pre-infection levels and some main-
taining very low levels throughout the chronic phase of infection. Despite this, the
mucosal barrier appears to remain intact, as SIV-infected African green monkeys
and sooty mangabeys do not show increased levels of plasma LPS (Brenchley et al.
2006; Gordon et al. 2007; Pandrea et al. 2007b).
The critical difference between the different outcomes may involve a specific
subset of CD4+ T cells, Th17 cells. These cells are postulated to be involved in
defense against bacterial pathogens through the production of the cytokines IL-17
306 E.J.D. Greenwood et al.
and IL-22 (Brenchley et al. 2008; Liang et al. 2006) and are found at higher fre-
quency in the gut mucosa compared with peripheral blood or lungs (Brenchley et al.
2008; Cecchinato et al. 2008). In HIV-1-infected humans and SIVmac-infected rhe-
sus macaques, in addition to the general destruction of gut CD4+ T cells, there is a
specific greater loss of Th17 cells—they are underrepresented in the residual CD4+
T cells (Brenchley et al. 2008; Cecchinato et al. 2008). In contrast, in the SIV infec-
tion of sooty mangabeys, total gut CD4+ T cells are depleted in SIV infection but
the percentage of Th17 cells within the remaining CD4+ cells is not altered
(Brenchley et al. 2008). Furthermore, in a study directly comparing SIVagm infec-
tion of pig-tailed macaques and African green monkeys, significant specific deple-
tion of Th17 only occurs in pig-tailed macaques, which progress to AIDS after
SIVagm infection (Favre et al. 2009). Thus, maintenance of the Th17 cell popula-
tion seems to play an important role in sustaining a competent mucosal barrier,
which in turn prevents systemic spread of bacteria from the gut and the subsequent
immune activation which follows in AIDS-susceptible species. How natural host
species have evolved to avoid this specific depletion of Th17 cells in the face of
massive general loss of CD4+ T cells in the gut has yet to be elucidated.
Studies using both flow cytometry to measure levels of activated T cells and
microarray analysis to measure gene expression in the peripheral blood and lym-
phoid tissues, throughout the course of infection, have demonstrated striking differ-
ences in immune activation in SIV-infected African green monkeys and sooty
mangabeys compared to disease susceptible HIV-1-infected humans and SIV-
infected rhesus and pig-tailed macaques. In the acute stage of infection, in all spe-
cies, there is a dramatic increase in the number of circulating activated T cells,
expression of genes that are induced by type I interferon, and soluble markers of
immune activation in the plasma (Harris et al. 2010; Kornfeld et al. 2005; Li et al.
2009; Meythaler et al. 2009; Stacey et al. 2009). In all species, the level of immune
activation is reduced in the transition from acute to chronic phase of infection.
However, the key difference between pathogenic and nonpathogenic infections is
the magnitude of this reduction in immune activation after the acute phase. In
humans and macaques, the reduction is only partial, and multiple markers of
immune activation remain highly elevated compared to pre-infection levels or unin-
fected controls.
In contrast, in sooty mangabeys and African green monkeys, the reduction in
immune activation after acute infection is more complete. In sooty mangabeys, a
small number of markers remain slightly (but significantly) upregulated when
chronically SIV-infected animals are compared to uninfected controls (Meythaler
et al. 2009; Silvestri et al. 2003). Interestingly, the extent of low-level chronic
immune activation does correlate inversely with CD4+ T-cell counts (Silvestri et al.
2003). Chronically infected African green monkeys are indistinguishable from
uninfected animals in the majority of markers, as immune activation is completely
suppressed (Kornfeld et al. 2005; Lederer et al. 2009; Lozano Reina et al. 2009).
Of particular interest is the strong induction of interferon-sensitive genes in the
lymph nodes of all species, which are then reduced to near-normal levels in sooty
mangabeys and African green monkeys, while remaining high in humans and rhe-
sus and pig-tailed macaques. Continuous immune activation has severe conse-
quences for the secondary lymphoid environment in pathogenic infections (see
below). In addition, the type I interferon response is likely connected to the other
elements of immune activation. Purified activated CD4+ T cells from HIV-1-
infected humans show much higher expression of interferon-sensitive genes than
activated cells from uninfected humans (Sedaghat et al. 2008), suggesting that type
I interferon plays a role in the excessive T-cell activation seen in HIV-1 infection. In
addition, the down-modulation of a type I interferon receptor on monocytes (which
is likely to be a measure of previous type I interferon exposure) correlates with
lower CD4+ T-cell counts, higher viral loads, and higher expression of CD38 on
CD8+ T cells (Hardy et al. 2009).
The mechanism by which sooty mangabeys and African green monkeys restrict
the interferon response and other elements of immune activation after the acute
stage of infection has been difficult to elucidate and may well be different for each
species. Down-modulation of the interferon response in the lymphoid tissue
occurs in natural host species despite similar levels of viral replication in second-
ary lymphoid tissues (Gueye et al. 2004). One study has shown that in vitro,
308 E.J.D. Greenwood et al.
Important changes occur in the T-cell zone, causing dramatic effects on the
whole body T-cell population. Firstly, infected humans and rhesus macaques show
significantly higher levels of collagen deposition in this area compared to unin-
fected controls (Diaz et al. 2010; Estes et al. 2007). As might be expected, the level
of collagen deposition correlates inversely with the number of naïve (and total)
CD4+ T cells resident in the lymph node, suggesting that collagen deposition
discourages habitation of this niche. In addition, the degree of collagen deposition
before the start of antiretroviral treatment inversely correlates with the increase in
peripheral blood T cells after 6 months of treatment (Schacker et al. 2002) or after
an even longer duration of treatment (Kumarasamy et al. 2009). Deposition of col-
lagen may therefore physically restrict the size of the niche available for CD4+ T
cells to occupy. In addition, a study in rhesus macaques has demonstrated that col-
lagen deposition restricts access of naïve T cells to the fibroblastic reticular cell
(FRC) network. This network provides the framework for migration of T cells
through the T-cell zone, and cells of this network also produce IL-7, a key stimulus
for naïve T-cell survival. Loss of access to this network in these animals leads to
apoptosis of naïve T cells in the node (Zeng et al. 2011).
The cause of this damage to the secondary lymphoid tissues has been
connected to two factors. First, in HIV-1-infected humans, there is abnormal
accumulation of CD4+ and CD8+ effector memory T cells in lymphoid tissues.
Effector memory cells are rare in the lymphoid tissue of uninfected humans, as
their normal role is to respond to antigen in non-lymphoid tissue. The extent of the
infiltration correlates with the area of collagen deposition in the tissue (Brenchley
et al. 2004b). These cells are presumably recruited to the lymph nodes as a result
of the pro-inflammatory environment induced by HIV-1, and so the presence of
such cells could either be a cause of damage or only a consequence of the inflam-
mation ongoing in the lymph node.
The second factor is the increased prevalence of TGFβ1-producing regulatory T
cells (Tregs) in the lymphatic tissue of infected humans and rhesus macaques. Tregs
are CD4+ T cells with an immunosuppressive function, including the production of
the cytokine TGFβ1. This cytokine has multiple functions, most prominently as an
immunosuppressant and in mediating wound repair. However, inappropriate expres-
sion is known to have a role in tissue fibrosis in other disease pathways (Branton and
Kopp 1999). In both humans and rhesus macaques, increased numbers of TGFβ1-
expressing cells and increased collagen deposition occur in the acute stage of infec-
tion and become exacerbated through the chronic stage of infection. Interestingly,
increased numbers of T cells producing the immunosuppressive factors IDO and
IL-10 are also found in infected rhesus macaques (Estes et al. 2006). It seems likely
that TGFβ1 is produced as part of a negative feedback response to the high levels of
immune activation, with the side effect of promoting damaging fibrosis in the
immune microenvironment.
The damaging effect of this immunosuppressive response is somewhat paradoxi-
cal, especially given that greater or similar levels of IL-10, IDO, and even TGFβ1
are found in gene expression analyses of SIV-infected African green monkeys and
sooty mangabeys when compared to rhesus and pig-tailed macaques (Bosinger
310 E.J.D. Greenwood et al.
et al. 2009; Jacquelin et al. 2009; Lederer et al. 2009). However, when using immu-
nohistology to analyze expression at the protein level, SIV-infected sooty mang-
abeys do not show increased numbers of TGFβ1-producing T cells in the lymph
node, and increased collagen deposition is not found in the T-cell zone. It seems
likely that the pathological aspect of TGFβ1 production in the lymph nodes of
humans and macaques is due to the failure, despite an immunosuppressive response,
to bring levels of immune activation down after acute infection. The ability men-
tioned above of sooty mangabeys and African green monkeys to resolve innate and
adaptive immune activation after chronic infection therefore likely saves the lymph
node environment from destruction.
The destruction of the lymph node environment in pathogenic infection leads to
loss of naïve T cells and the inability to generate de novo CD4+ and CD8+ T-cell
responses by preventing normal interactions between T cells and antigen-presenting
cells. Both naïve and memory T cells may become inappropriately activated due to the
high levels of interferon and are lost through subsequent apoptosis or direct HIV
infection. The combination of loss of memory T cells and the environment required to
generate new responses presumably drives the immune system to the critical point at
which AIDS occur (Heeney 1995). Existing memory cells to a specific antigen/patho-
gen are sufficiently depleted that there is no longer a meaningful memory response to
opportunistic infections, and the generation of new responses is sufficiently delayed
by the loss of proper lymph node environment and depletion of naïve T cells that an
immune response cannot be mounted in time to prevent illness and death. The key
pathways leading to disease in humans and macaques and how they are prevented in
African green monkeys and sooty mangabeys are outlined in Fig. 3.
As discussed, SIVsmm generally does not cause disease in sooty mangabey but has
the potential to cause an AIDS-like disease upon direct inoculation of Asian
macaques (McClure et al. 1989; Silvestri et al. 2005). The disease caused in rhesus
macaques is a commonly used animal model for AIDS in humans caused by HIV-1.
SIVsmm has also been transmitted into humans, causing the HIV-2 epidemic.
Numerous lines of evidence demonstrate that in humans, HIV-2 is not as pathogenic
as HIV-1. In surveys taking place in Guinea-Bissau, Holmgren and colleagues
found only a twofold increase in mortality in individuals infected with HIV-2, com-
parable with previous findings within the same geographical region (Holmgren
et al. 2007; Poulsen et al. 1997; Ricard et al. 1994). In contrast, HIV-1 is associated
with a 10–15-fold increase in mortality rate (reviewed in (Jaffar et al. 2004).
Longitudinal investigations with statistically meaningful cohorts indicate that about
80 % of HIV-2-infected individuals do not develop disease and maintain a normal
life expectancy (van der Loeff et al. 2010). One-third of all HIV-2-infected
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 311
Fig. 3 Pathways leading to the development of AIDS in progressive SIV/HIV infection compared
with the nonprogressive SIV infection of African primates
individuals have an undetectable plasma viral load (van der Loeff et al. 2010).
Nevertheless, progressive HIV-2 infections show clinical features that are indistin-
guishable from AIDS caused by HIV-1 (Martinez-Steele et al. 2007).
312 E.J.D. Greenwood et al.
Given the considerable difference in the outcome of HIV-1 and HIV-2 infection
of humans, is it appropriate to use the disease of rhesus macaques infected with
viruses of the SIVsm/SIVmac/HIV-2 lineage as a model for disease caused by
HIV-1 in humans? If it is the case that there are specific aspects of the HIV-1 lineage
that make this virus more pathogenic than HIV-2 in humans, then this model
becomes less attractive in understanding HIV-1 pathogenesis. We will now examine
the differences between these viruses to establish if this is the case.
Despite their different origins, HIV-1 and HIV-2 are relatively closely related
retroviruses with 60 % similarity at the amino acid level in the capsid and poly-
merase proteins and 30 % similarity in the envelope protein (Guyader et al. 1987).
However, the genomes of both viral lineages differ by two genes. While vpx is pres-
ent in the HIV-2/SIVsmm lineage (in addition to SIVrcm, SIVdrl, and SIVmnd2),
the accessory gene vpu, not vpx, is found in HIV-1, in addition to SIVcpz and the
SIVs of four species of the Cercopithecus genus (Barlow et al. 2003; Courgnaud
et al. 2002, 2003; Huet et al. 1990; Strebel et al. 1988).
Both vpu and vpx seem to have evolved to allow lentiviruses to escape host
restriction mechanisms. In macaques, the accessory viral gene vpx has been directly
shown to contribute to virulence. Animals infected with an SIVmac mutant lacking
the vpx gene show lower virus burdens, delayed declines in CD4 lymphocytes, and
either a lack of disease or delayed progression to disease (Gibbs et al. 1995; Hirsch
et al. 1998). The Vpx of the SIVsmm/SIVmac/HIV-2 viral lineage enables the virus
to efficiently replicate in primate macrophages by antagonizing a cellular restriction
factor termed SAM domain and HD domain 1 (SAMHD1) that is expressed in this
cell type (Laguette et al. 2011; Sharova et al. 2008). That vpx can provide a fitness
advantage to the virus was further confirmed in vivo when a wild-type and a vpx-
deleted SIVsmm isolate were inoculated into pig-tailed macaques, resulting in the
vpx mutant showing a strong competitive disadvantage in the early virus dissemina-
tion (Hirsch et al. 1998).
vpu, the accessory gene uniquely expressed in the SIVcpz/HIV-1 lineage, includ-
ing the small subset of guenon SIVs, also seems to posess various functions that
may impact pathogenicity. Vpu helps HIV-1 to escape host restriction and enhances
release of virus particles (Klimkait et al. 1990). Vpu contributes to HIV-1-induced
CD4 receptor downregulation, which enhances virus replication (Willey et al.
1992). In addition, it increases the release of progeny virions from infected cells and
cell-to-cell spread of viral particles by antagonizing tetherin, an interferon-induced
host restriction factor that directly cross-links virions on the host cell surface (Neil
et al. 2008). However, both CD4 and tetherin downregulation are not exclusive to
the lineage of HIV-1 and are facilitated by the nef or env gene in other lentiviruses
(Gupta et al. 2009; Le Tortorec and Neil 2009; Zhang et al. 2009). However, two
functions of vpu have recently been discovered that are thus far unique to HIV-1.
Firstly, Vpu inhibits the recycling of CD1d receptor from endosomal compartments,
strongly inhibiting the ability of infected DC to activate CD1d-restricted natural
killer T cells (NKT cells) (Moll et al. 2010). Secondly, Vpu downregulates the NKT
and B cell coactivator (NTB-A) at the surface of infected cells and as a result inter-
feres with the degranulation of NK cells that recognize the infected cells (Shah et al.
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 313
2010). Importantly, HIV-2-infected cells do not downregulate NTB-A and are killed
more efficiently by NK cells than HIV-1-infected cells (Shah et al. 2010).
Furthermore, vpu seems to have influenced the evolution of the other viral genes.
A mechanism encoded in the viral gene nef is believed to have evolved to restrict
T-cell activation, through down modulation of the T-cell receptor (also referred to as
CD3) in infected cells. However, in all viruses that carry vpu this function of nef is
reported to be lost (Schindler et al. 2006). Immunodeficiency viruses depend on
activated T cells for replication, but accelerated immune activation induces a range
of host defense mechanisms. It has been suggested that acquisition of the vpu gene
by this viral lineage has allowed the virus to trigger immune activation, increasing
viral replication, without the consequence of host-mediated restriction (Kirchhoff
2009). There are therefore reasons to believe that vpu plays an important role in
HIV-1-induced disease, and models of human AIDS should include this factor.
While pig-tailed macaques can become transiently infected with HIV-1 (Agy
et al. 1992), the infection does not persist. Other macaque species are resistant to
infection, as macaque-encoded restriction factors provide dominant-acting blocks
the establishment of infection (Stremlau et al. 2004). Using the macaque model to
investigate influence of the vpu gene in vivo therefore requires more complex
approaches. Our increased understanding of host restriction factors has led to novel
approaches in which HIV-1 isolates have been artificially equipped with viral gene
variants capable of antagonizing the macaque’s host restriction factors (Hatziioannou
et al. 2009). However, to date, these artificial HIV variants are not pathogenic. An
alternative approach has been to infect macaques with partial recombinants between
SIVs and HIV-1 isolates, called SHIVs. The first SHIVs were created by exchang-
ing the SIV envelope for an HIV-1 envelope sequence in order to test HIV-1 vaccine
candidates (Stremlau et al. 2004). Some SHIVs are equipped with the HIV-1 vpu
gene in addition to the HIV envelope, and a few studies indicate that the presence of
vpu can influence the pathogenic outcome in pig-tailed macaques (Hout et al. 2005;
Singh et al. 2001, 2003; Stephens et al. 2002). However, the mechanistic back-
ground remains unknown, especially as the various described functions of vpu are
yet not characterized for these Asian primate hosts. The lack of a suitable model for
HIV-1 has driven the search for alternative animal models such as the rodent model
in which mice were “humanized” with bone marrow/liver/thymus grafts from
human donors (Melkus et al. 2006). To date, these models are still in a premature
state and only further research will establish whether insights into HIV-1 virulence
can be gained with such models.
As stated previously, other than HIV-1, only SIVcpz of chimpanzees and SIVmon/
mus/gsn/den of four primate species of the guenon genus carry the vpu gene. Given
the difficulties in assessing the role of vpu in existing in vivo models described
314 E.J.D. Greenwood et al.
CD4+ cells in the periarteriolar lymphoid sheaths (PALS) of the spleens of these
animals was shown in the SIVcpz-infected animals. Particularly severe depletion
was shown in an animal that had become infected during the study, three years prior
to death, whose death was attributed to an “AIDS-like” disease by the group report-
ing this finding. A more recent article reports also identified a Pan troglodytes trog-
lodytes chimpanzee with AIDS-like clinical signs in a Cameroonian sanctuary
(Etienne et al. 2011). A pathogenic outcome of SIVcpz infection would lend further
support to the importance of vpu in HIV-1 infection.
From 1989 to 2005, seven SIVcpz-infected chimpanzees were housed in primate
centers in Europe and the USA (Heeney et al. 2006): one naturally infected sch-
weinfurthii animal, two experimentally infected schweinfurthii animals, and four
experimentally infected verus animals. Of these seven animals, four are still alive,
including Noah, a naturally infected animal that first tested positive in 1989, aged
around 2 years old, and remains asymptomatic after over 20 years of infection
(Greenwood et al., unpublished data). Three animals have died of cardiac condi-
tions, which are common in captive chimpanzees (Seiler et al. 2009).
Two of these chimpanzees have already been studied in some detail as part of a
cohort of HIV-1-infected chimpanzees. As with the HIV-1-infected chimpanzees,
these two SIVcpz-infected chimpanzees seem to lack the profound changes to the
immune system that are found in HIV-1-infected humans (Gougeon et al. 1997;
Rutjens et al. 2008, 2010). Given the apparent pathogenic outcome of SIVcpz infec-
tion of wild chimpanzees, it should be possible to identify, in these captive animals,
disease mechanisms that are shared with HIV-1-infected humans. Further study of
samples from these and other captive SIVcpz-infected animals is therefore war-
ranted, especially if this is possible using previously archived samples and/or non-
invasive methods.
As previously mentioned, SIVcpz is the result of recombination between SIVrcm
and SIVgsn/mon/mus. The importance of the vpu gene seems to be emphasized by
the fact that it is retained in the chimeric virus in favor of vpx. The small subset of
Cercopithecus species infected with these vpu-carrying viruses have not been
investigated for their outcome of infection. The only knowledge we have of these
viruses is their genomic sequences and their prevalence in the wild. Interestingly,
the few studies that investigated significant numbers of these Cercopithecus ani-
mals found an exceptionally low prevalence of the vpu-harboring SIVs in mus-
tached and greater spot-nosed guenon populations in Cameroon (2–4 %). This is in
stark contrast to the seroprevalence seen in primate populations that carry an SIV
lacking the vpu gene, which reported seroprevalence of 50–90 % (Aghokeng et al.
2006, 2009; Ellis et al. 2004). It therefore seems that the outcome of SIVgsn/
SIVmon infection is likely to differ from other SIV infection of natural hosts. It is
possible to postulate two possible mechanisms for this low prevalence. SIV infec-
tion in these species could be much more pathogenic than in other species, perhaps
due to the presence of vpu—which leads to infected animals being rapidly removed
from the population. Alternatively, these species could have evolved novel mecha-
nisms to prevent the spread of the virus within the population, though presumably
the evolutionary pressure for this to occur would again have to be the result of
316 E.J.D. Greenwood et al.
Conclusion
The African primate origins of HIV-1 and HIV-2 are now well established, and a
relatively thorough history of how HIV-1 has spread from western Central Africa
throughout the world has been proposed. However, the mechanisms by which HIV
causes AIDS in humans remain heavily debated. It should be noted that it is not
possible to review here the entirety of the vast and conflicting literature describing
the observed alterations to the immune system in HIV infection and the various
mechanisms proposed to cause CD4+ T-cell loss and AIDS in HIV in infected
humans.
Instead, we have highlighted the key differences found throughout the course of
infection between the pathogenic HIV/SIV infection of humans and macaques and
the nonpathogenic SIV infection of two well-studied African primate species and
used these differences to propose a relatively straightforward model by which lenti-
virus infection induces AIDS in humans and macaques but not in sooty mangabeys
and African green monkeys.
Finally, we have reviewed the possible limitation of current animal models that
specifically the HIV-1/SIVcpz viral lineage may have greater potential to cause
disease than other primate lentiviruses, which is not well addressed by current pri-
mate models. Specifically, we have noted the functions of the vpu gene. While some
of the functions of vpu are carried out by other genes in primate lentiviruses (such
as down-modulation of CD4 and tetherin), other functions are so far described
uniquely for this gene. It is also noted that the virus–host relationship in primate
species infected with a vpu-harboring virus seems to differ from that of other African
primates. While further examination of primate species infected with vpu-carrying
viruses—chimpanzees, gorillas, and members of the Cercopithecus genus—is
clearly complicated by ethical and practical considerations, noninvasive studies of
these species have already proven to be fruitful. Expansion of such noninvasive
studies may lead to new insights in the subject of lentivirus pathogenesis and host
adaptation not available in other animal models.
Acknowledgments We would like to thank Joel Wertheim for creating the HIV/SIV alignments
used to create Fig. 1.
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 317
References
Addo MM, Yu XG, Rathod A, Cohen D, Eldridge RL, Strick D, Johnston MN, Corcoran C, Wurcel
AG, Fitzpatrick CA et al (2003) Comprehensive epitope analysis of human immunodeficiency
virus type 1 (HIV-1)-specific T-cell responses directed against the entire expressed HIV-1 genome
demonstrate broadly directed responses, but no correlation to viral load. J Virol 77(3):2081–2092
Aghokeng AF, Liu W, Bibollet-Ruche F, Loul S, Mpoudi-Ngole E, Laurent C, Mwenda JM, Langat
DK, Chege GK, McClure HM et al (2006) Widely varying SIV prevalence rates in naturally
infected primate species from Cameroon. Virology 345(1):174–189
Aghokeng AF, Bailes E, Loul S, Courgnaud V, Mpoudi-Ngolle E, Sharp PM, Delaporte E, Peeters
M (2007) Full-length sequence analysis of SIVmus in wild populations of mustached monkeys
(Cercopithecus cephus) from Cameroon provides evidence for two co-circulating SIVmus lin-
eages. Virology 360(2):407–418
Aghokeng AF, Ayouba A, Mpoudi-Ngole E, Loul S, Liegeois F, Delaporte E, Peeters M (2009)
Extensive survey on the prevalence and genetic diversity of SIVs in primate bushmeat provides
insights into risks for potential new cross-species transmissions. Infect Genet Evol 10(3):386–396
Agy MB, Frumkin LR, Corey L, Coombs RW, Wolinsky SM, Koehler J, Morton WR, Katze MG
(1992) Infection of Macaca nemestrina by human immunodeficiency virus type-1. Science
257(5066):103–106
Aina O, Dadik J, Charurat M, Amangaman P, Gurumdi S, Mang E, Guyit R, Lar N, Datong P,
Daniyam C et al (2005) Reference values of CD4 T lymphocytes in human immunodeficiency
virus-negative adult Nigerians. Clin Diagn Lab Immunol 12(4):525–530
Alter HJ, Eichberg JW, Masur H, Saxinger WC, Gallo R, Macher AM, Lane HC, Fauci AS (1984)
Transmission of HTLV-III infection from human plasma to chimpanzees: an animal model for
AIDS. Science 226(4674):549–552
Apetrei C, Lerche NW, Pandrea I, Gormus B, Silvestri G, Kaur A, Robertson DL, Hardcastle J,
Lackner AA, Marx PA (2006) Kuru experiments triggered the emergence of pathogenic
SIVmac. AIDS 20(3):317–321
Ayinde D, Maudet C, Transy C, Margottin-Goguet F (2010) Limelight on two HIV/SIV accessory
proteins in macrophage infection: is Vpx overshadowing Vpr? Retrovirology 7:35
Ayouba A, Souquieres S, Njinku B, Martin PM, Muller-Trutwin MC, Roques P, Barre-Sinoussi F,
Mauclere P, Simon F, Nerrienet E (2000) HIV-1 group N among HIV-1-seropositive individuals
in Cameroon. AIDS 14(16):2623–2625
Bailes E, Gao F, Bibollet-Ruche F, Courgnaud V, Peeters M, Marx PA, Hahn BH, Sharp PM (2003)
Hybrid origin of SIV in chimpanzees. Science 300(5626):1713
Barre-Sinoussi F, Chermann JC, Rey F, Nugeyre MT, Chamaret S, Gruest J, Dauguet C et al (1983)
Isolation of a T-lymphotropic retrovirus from a patient at risk for acquired immune deficiency
syndrome (AIDS). Science 220(4599):868–871
Barlow KL, Ajao AO, Clewley JP (2003) Characterization of a novel simian immunodeficiency
virus (SIVmonNG1) genome sequence from a mona monkey (Cercopithecus mona). J Virol
77(12):6879–6888
Beaumier CM, Harris LD, Goldstein S, Klatt NR, Whitted S, McGinty J, Apetrei C, Pandrea I,
Hirsch VM, Brenchley JM (2009) CD4 down-regulation by memory CD4+ T cells in vivo ren-
ders African green monkeys resistant to progressive SIVagm infection. Nat Med 15(8):879–885
Beaumont T, van Nuenen A, Broersen S, Blattner WA, Lukashov VV, Schuitemaker H (2001)
Reversal of human immunodeficiency virus type 1 IIIB to a neutralization-resistant phenotype
in an accidentally infected laboratory worker with a progressive clinical course. J Virol
75(5):2246–2252
Beer BE, Foley BT, Kuiken CL, Tooze Z, Goeken RM, Brown CR, Hu J, St Claire M, Korber BT,
Hirsch VM (2001) Characterization of novel simian immunodeficiency viruses from red-
capped mangabeys from Nigeria (SIVrcmNG409 and -NG411). J Virol 75(24):12014–12027
Begtrup K, Melbye M, Biggar RJ, Goedert JJ, Knudsen K, Andersen PK (1997) Progression to
acquired immunodeficiency syndrome is influenced by CD4 T-lymphocyte count and time
since seroconversion. Am J Epidemiol 145(7):629–635
318 E.J.D. Greenwood et al.
Benveniste RE, Arthur LO, Tsai CC, Sowder R, Copeland TD, Henderson LE, Oroszlan S (1986)
Isolation of a lentivirus from a macaque with lymphoma: comparison with HTLV-III/LAV and
other lentiviruses. J Virol 60(2):483–490
Bibollet-Ruche F, Bailes E, Gao F, Pourrut X, Barlow KL, Clewley JP, Mwenda JM, Langat DK,
Chege GK, McClure HM et al (2004) New simian immunodeficiency virus infecting De
Brazza’s monkeys (Cercopithecus neglectus): evidence for a cercopithecus monkey virus
clade. J Virol 78(14):7748–7762
Boesch C, Boesch-Achermann H (2000) The chimpanzees of the Taï Forest: behavioural ecology
and evolution. Oxford University Press, Oxford, p 316, viii
Boily MC, Baggaley RF, Wang L, Masse B, White RG, Hayes RJ, Alary M (2009) Heterosexual
risk of HIV-1 infection per sexual act: systematic review and meta-analysis of observational
studies. Lancet Infect Dis 9(2):118–129
Bosinger SE, Li Q, Gordon SN, Klatt NR, Duan L, Xu L, Francella N, Sidahmed A, Smith AJ,
Cramer EM (2009) Global genomic analysis reveals rapid control of a robust innate response
in SIV-infected sooty mangabeys. J Clin Invest 119(12):3556–3572
Branton MH, Kopp JB (1999) TGF-beta and fibrosis. Microbes Infect 1(15):1349–1365
Brenchley JM, Paiardini M (2011) Immunodeficiency lentiviral infections in natural and nonnatu-
ral hosts. Blood 118:847–854
Brenchley JM, Hill BJ, Ambrozak DR, Price DA, Guenaga FJ, Casazza JP, Kuruppu J, Yazdani J,
Migueles SA, Connors M et al (2004a) T-cell subsets that harbor human immunodeficiency
virus (HIV) in vivo: implications for HIV pathogenesis. J Virol 78(3):1160–1168
Brenchley JM, Schacker TW, Ruff LE, Price DA, Taylor JH, Beilman GJ, Nguyen PL, Khoruts A,
Larson M, Haase AT et al (2004b) CD4+ T cell depletion during all stages of HIV disease
occurs predominantly in the gastrointestinal tract. J Exp Med 200(6):749–759
Brenchley JM, Price DA, Schacker TW, Asher TE, Silvestri G, Rao S, Kazzaz Z, Bornstein E,
Lambotte O, Altmann D et al (2006) Microbial translocation is a cause of systemic immune
activation in chronic HIV infection. Nat Med 12(12):1365–1371
Brenchley JM, Paiardini M, Knox KS, Asher AI, Cervasi B, Asher TE, Scheinberg P, Price DA,
Hage CA, Kholi LM et al (2008) Differential Th17 CD4 T-cell depletion in pathogenic and
nonpathogenic lentiviral infections. Blood 112(7):2826–2835
Brown CR, Czapiga M, Kabat J, Dang Q, Ourmanov I, Nishimura Y, Martin MA, Hirsch VM
(2007) Unique pathology in simian immunodeficiency virus-infected rapid progressor
macaques is consistent with a pathogenesis distinct from that of classical AIDS. J Virol
81(11):5594–5606
Campbell-Yesufu OT, Gandhi RT (2011) Update on human immunodeficiency virus (HIV)-2
infection. Clin Infect Dis 52(6):780–787
CDC (1981a) Follow-up on Kaposi’s Sarcoma and pneumocystis pneumonia. MMWR Morb
Mortal Wkly Rep 30:409–410
CDC (1981b) Kaposi’s sarcoma and pneumocycstis pneumonia among homosexual men—New
York City and California. MMWR Morb Mortal Wkly Rep 30:305–308
CDC (1981c) Pneumocycstis pneumonia—Los Angeles. MMWR Morb Mort Wkly Rep 30:1–3
CDC (1982a) Epidemiologic aspects of the current outbreak of Kaposi’s sarcoma and opportunis-
tic infections. N Engl J Med 306(4):248–252
CDC (1982b) Epidemiologic notes and reports update on Kaposi’s sarcoma and opportunistic
infections in previously healthy persons—United States. MMWR Morb Mortal Wkly Rep
31:300–301
CDC (1982c) Opportunistic infections and Kaposi’s sarcoma among Haitians in the United States.
MMWR Morb Mortal Wkly Rep 31:353–354
Cecchinato V, Trindade CJ, Laurence A, Heraud JM, Brenchley JM, Ferrari MG, Zaffiri L,
Tryniszewska E, Tsai WP, Vaccari M et al (2008) Altered balance between Th17 and Th1 cells
at mucosal sites predicts AIDS progression in simian immunodeficiency virus-infected
macaques. Mucosal Immunol 1(4):279–288
Chen Z, Kwon D, Jin Z, Monard S, Telfer P, Jones MS, Lu CY, Aguilar RF, Ho DD, Marx PA
(1998) Natural infection of a homozygous delta24 CCR5 red-capped mangabey with an R2b-
tropic simian immunodeficiency virus. J Exp Med 188(11):2057–2065
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 319
Estes JD, Li Q, Reynolds MR, Wietgrefe S, Duan L, Schacker T, Picker LJ, Watkins DI, Lifson JD,
Reilly C et al (2006) Premature induction of an immunosuppressive regulatory T cell response
during acute simian immunodeficiency virus infection. J Infect Dis 193(5):703–712
Estes JD, Wietgrefe S, Schacker T, Southern P, Beilman G, Reilly C, Milush JM, Lifson JD, Sodora
DL, Carlis JV (2007) Simian immunodeficiency virus-induced lymphatic tissue fibrosis is
mediated by transforming growth factor beta 1-positive regulatory T cells and begins in early
infection. J Infect Dis 195(4):551–561
Estes J, Baker JV, Brenchley JM, Khoruts A, Barthold JL, Bantle A, Reilly CS, Beilman GJ,
George ME, Douek DC et al (2008) Collagen deposition limits immune reconstitution in the
gut. J Infect Dis 198(4):456–464
Etienne L, Nerrienet E, LeBreton M, Bibila GT, Foupouapouognigni Y, Rousset D, Nana A, Djoko
CF, Tamoufe U, Aghokeng AF et al (2011) Characterization of a new simian immunodefi-
ciency virus strain in a naturally infected Pan troglodytes troglodytes chimpanzee with AIDS
related symptoms. Retrovirology 8:4
Fabre PH, Rodrigues A, Douzery EJ (2009) Patterns of macroevolution among Primates inferred
from a supermatrix of mitochondrial and nuclear DNA. Mol Phylogenet Evol 53(3):808–825
Favre D, Lederer S, Kanwar B, Ma ZM, Proll S, Kasakow Z, Mold J, Swainson L, Barbour JD,
Baskin CR et al (2009) Critical loss of the balance between Th17 and T regulatory cell popula-
tions in pathogenic SIV infection. PLoS Pathog 5(2):e1000295
Fleming DT, Wasserheit JN (1999) From epidemiological synergy to public health policy and
practice: the contribution of other sexually transmitted diseases to sexual transmission of HIV
infection. Sex Transm Infect 75(1):3–17
Ganusov VV, De Boer RJ (2007) Do most lymphocytes in humans really reside in the gut? Trends
Immunol 28(12):514–518
Gea-Banacloche JC, Migueles SA, Martino L, Shupert WL, McNeil AC, Sabbaghian MS, Ehler L,
Prussin C, Stevens R, Lambert L et al (2000) Maintenance of large numbers of virus-specific
CD8+ T cells in HIV-infected progressors and long-term nonprogressors. J Immunol
165(2):1082–1092
Gibbs JS, Lackner AA, Lang SM, Simon MA, Sehgal PK, Daniel MD, Desrosiers RC (1995)
Progression to AIDS in the absence of a gene for vpr or vpx. J Virol 69(4):2378–2383
Gifford RJ, Katzourakis A, Tristem M, Pybus OG, Winters M, Shafer RW (2008) A transitional
endogenous lentivirus from the genome of a basal primate and implications for lentivirus evo-
lution. Proc Natl Acad Sci USA 105(51):20362–20367
Gilbert MT, Rambaut A, Wlasiuk G, Spira TJ, Pitchenik AE, Worobey M (2007) The emergence of
HIV/AIDS in the Americas and beyond. Proc Natl Acad Sci USA 104(47):18566–18570
Gilbert C, Maxfield DG, Goodman SM, Feschotte C (2009) Parallel germline infiltration of a len-
tivirus in two Malagasy lemurs. PLoS Genet 5(3):e1000425
Gordon SN, Klatt NR, Bosinger SE, Brenchley JM, Milush JM, Engram JC, Dunham RM,
Paiardini M, Klucking S, Danesh A et al (2007) Severe depletion of mucosal CD4+ T cells in
AIDS-free simian immunodeficiency virus-infected sooty mangabeys. J Immunol
179(5):3026–3034
Gottlieb MS, Schroff R, Schanker HM, Weisman JD, Fan PT, Wolf RA, Saxon A (1981)
Pneumocystis carinii pneumonia and mucosal candidiasis in previously healthy homosexual
men: evidence of a new acquired cellular immunodeficiency. N Engl J Med 305(24):
1425–1431
Gougeon ML, Lecoeur H, Boudet F, Ledru E, Marzabal S, Boullier S, Roue R, Nagata S, Heeney
J (1997) Lack of chronic immune activation in HIV-infected chimpanzees correlates with the
resistance of T cells to Fas/Apo-1 (CD95)-induced apoptosis and preservation of a T helper 1
phenotype. J Immunol 158(6):2964–2976
Guadalupe M, Reay E, Sankaran S, Prindiville T, Flamm J, McNeil A, Dandekar S (2003) Severe
CD4+ T-cell depletion in gut lymphoid tissue during primary human immunodeficiency virus
type 1 infection and substantial delay in restoration following highly active antiretroviral ther-
apy. J Virol 77(21):11708–11717
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 321
Gueye A, Diop OM, Ploquin MJ, Kornfeld C, Faye A, Cumont MC, Hurtrel B, Barre-Sinoussi F,
Muller-Trutwin MC (2004) Viral load in tissues during the early and chronic phase of non-
pathogenic SIVagm infection. J Med Primatol 33(2):83–97
Gupta RK, Mlcochova P, Pelchen-Matthews A, Petit SJ, Mattiuzzo G, Pillay D, Takeuchi Y, Marsh
M, Towers GJ (2009) Simian immunodeficiency virus envelope glycoprotein counteracts teth-
erin/BST-2/CD317 by intracellular sequestration. Proc Natl Acad Sci USA 106(49):
20889–20894
Gurtler LG, Hauser PH, Eberle J, von Brunn A, Knapp S, Zekeng L, Tsague JM, Kaptue L (1994)
A new subtype of human immunodeficiency virus type 1 (MVP-5180) from Cameroon. J Virol
68(3):1581–1585
Guyader M, Emerman M, Sonigo P, Clavel F, Montagnier L, Alizon M (1987) Genome organiza-
tion and transactivation of the human immunodeficiency virus type 2. Nature 326(6114):
662–669
Hardy GA, Sieg SF, Rodriguez B, Jiang W, Asaad R, Lederman MM, Harding CV (2009)
Desensitization to type I interferon in HIV-1 infection correlates with markers of immune acti-
vation and disease progression. Blood 113(22):5497–5505
Harris LD, Tabb B, Sodora DL, Paiardini M, Klatt NR, Douek DC, Silvestri G, Muller-Trutwin M,
Vasile-Pandrea I, Apetrei C et al (2010) Down-regulation of robust acute type I interferon
responses distinguishes nonpathogenic simian immunodeficiency virus (SIV) infection of nat-
ural hosts from pathogenic SIV infection of rhesus macaques. J Virol 84(15):7886–7891
Hatziioannou T, Ambrose Z, Chung NP, Piatak M Jr, Yuan F, Trubey CM, Coalter V, Kiser R,
Schneider D, Smedley J et al (2009) A macaque model of HIV-1 infection. Proc Natl Acad Sci
USA 106(11):4425–4429
Hazenberg MD, Otto SA, Cohen Stuart JW, Verschuren MC, Borleffs JC, Boucher CA, Coutinho
RA, Lange JM, Rinke de Wit TF, Tsegaye A et al (2000) Increased cell division but not thymic
dysfunction rapidly affects the T-cell receptor excision circle content of the naive T cell popula-
tion in HIV-1 infection. Nat Med 6(9):1036–1042
Hazenberg MD, Otto SA, van Benthem BH, Roos MT, Coutinho RA, Lange JM, Hamann D, Prins
M, Miedema F (2003) Persistent immune activation in HIV-1 infection is associated with pro-
gression to AIDS. AIDS 17(13):1881–1888
Heeney JL (1995) AIDS: a disease of impaired Th-cell renewal? Immunol Today 16(11):
515–520
Heeney JL, Rutjens E, Verschoor EJ, Niphuis H, ten Haaft P, Rouse S, McClure H, Balla-
Jhagjhoorsingh S, Bogers W, Salas M et al (2006) Transmission of simian immunodeficiency
virus SIVcpz and the evolution of infection in the presence and absence of concurrent human
immunodeficiency virus type 1 infection in chimpanzees. J Virol 80(14):7208–7218
Hirsch VM, Dapolito G, Johnson PR, Elkins WR, London WT, Montali RJ, Goldstein S, Brown C
(1995) Induction of AIDS by simian immunodeficiency virus from an African green monkey:
species-specific variation in pathogenicity correlates with the extent of in vivo replication.
J Virol 69(2):955–967
Hirsch VM, Sharkey ME, Brown CR, Brichacek B, Goldstein S, Wakefield J, Byrum R, Elkins
WR, Hahn BH, Lifson JD et al (1998) Vpx is required for dissemination and pathogenesis of
SIV(SM) PBj: evidence of macrophage-dependent viral amplification. Nat Med 4(12):
1401–1408
Holmgren B, da Silva Z, Vastrup P, Larsen O, Andersson S, Ravn H, Aaby P (2007) Mortality
associated with HIV-1, HIV-2, and HTLV-I single and dual infections in a middle-aged and
older population in Guinea-Bissau. Retrovirology 4:85
Hout DR, Gomez ML, Pacyniak E, Gomez LM, Inbody SH, Mulcahy ER, Culley N, Pinson DM,
Powers MF, Wong SW et al (2005) Scrambling of the amino acids within the transmembrane
domain of Vpu results in a simian-human immunodeficiency virus (SHIVTM) that is less
pathogenic for pig-tailed macaques. Virology 339(1):56–69
Huet T, Cheynier R, Meyerhans A, Roelants G, Wain-Hobson S (1990) Genetic organization of a
chimpanzee lentivirus related to HIV-1. Nature 345(6273):356–359
322 E.J.D. Greenwood et al.
Jacquelin B, Mayau V, Targat B, Liovat AS, Kunkel D, Petitjean G, Dillies MA, Roques P, Butor
C, Silvestri G et al (2009) Nonpathogenic SIV infection of African green monkeys induces a
strong but rapidly controlled type I IFN response. J Clin Invest 119(12):3544–3555
Jaffar S, Grant AD, Whitworth J, Smith PG, Whittle H (2004) The natural history of HIV-1 and HIV-2
infections in adults in Africa: a literature review. Bull World Health Organ 82(6):462–469
Jin MJ, Hui H, Robertson DL, Muller MC, Barre-Sinoussi F, Hirsch VM, Allan JS, Shaw GM,
Sharp PM, Hahn BH (1994) Mosaic genome structure of simian immunodeficiency virus from
west African green monkeys. EMBO J 13(12):2935–2947
Juompan LY, Hutchinson K, Montefiori DC, Nidtha S, Villinger F, Novembre FJ (2008) Analysis
of the immune responses in chimpanzees infected with HIV type 1 isolates. AIDS Res Hum
Retroviruses 24(4):573–586
Katzourakis A, Tristem M, Pybus OG, Gifford RJ (2007) Discovery and analysis of the first endog-
enous lentivirus. Proc Natl Acad Sci USA 104(15):6261–6265
Kaufmann GR, Cunningham P, Kelleher AD, Zaunders J, Carr A, Vizzard J, Law M, Cooper DA
(1998) Patterns of viral dynamics during primary human immunodeficiency virus type 1 infec-
tion. The Sydney Primary HIV Infection Study Group. J Infect Dis 178(6):1812–1815
Keckesova Z, Ylinen LM, Towers GJ, Gifford RJ, Katzourakis A (2009) Identification of a RELIK
orthologue in the European hare (Lepus europaeus) reveals a minimum age of 12 million years
for the lagomorph lentiviruses. Virology 384(1):7–11
Keele BF, Van Heuverswyn F, Li Y, Bailes E, Takehisa J, Santiago ML, Bibollet-Ruche F, Chen Y,
Wain LV, Liegeois F et al (2006) Chimpanzee reservoirs of pandemic and nonpandemic HIV-1.
Science 313(5786):523–526
Keele BF, Jones JH, Terio KA, Estes JD, Rudicell RS, Wilson ML, Li Y, Learn GH, Beasley TM,
Schumacher-Stankey J et al (2009) Increased mortality and AIDS-like immunopathology in
wild chimpanzees infected with SIVcpz. Nature 460(7254):515–519
Kibaya RS, Bautista CT, Sawe FK, Shaffer DN, Sateren WB, Scott PT, Michael NL, Robb ML,
Birx DL, de Souza MS (2008) Reference ranges for the clinical laboratory derived from a rural
population in Kericho, Kenya. PLoS One 3(10):e3327
Kirchhoff F (2009) Is the high virulence of HIV-1 an unfortunate coincidence of primate lentiviral
evolution? Nat Rev Microbiol 7(6):467–476
Kirchhoff F (2010) Immune evasion and counteraction of restriction factors by HIV-1 and other
primate lentiviruses. Cell Host Microbe 8(1):55–67
Klimkait T, Strebel K, Hoggan MD, Martin MA, Orenstein JM (1990) The human immunodefi-
ciency virus type 1-specific protein vpu is required for efficient virus maturation and release. J
Virol 64(2):621–629
Kornfeld C, Ploquin MJ, Pandrea I, Faye A, Onanga R, Apetrei C, Poaty-Mavoungou V, Rouquet
P, Estaquier J, Mortara L et al (2005) Antiinflammatory profiles during primary SIV infection
in African green monkeys are associated with protection against AIDS. J Clin Invest
115(4):1082–1091
Kumarasamy N, Venkatesh KK, Devaleenol B, Poongulali S, Yephthomi T, Pradeep A, Saghayam
S, Flanigan T, Mayer KH, Solomon S (2009) Factors associated with mortality among HIV-
infected patients in the era of highly active antiretroviral therapy in southern India. Int J Infect
Dis 14(2):e127–131
Laguette N, Sobhian B, Casartelli N, Ringeard M, Chable-Bessia C, Segeral E, Yatim A et al (2011)
SAMHD1 is the dendritic- and myeloid-cell-specific HIV-1 restriction factor counteracted by
Vpx. Nature 474(7353):654–657
Le Tortorec A, Neil SJ (2009) Antagonism to and intracellular sequestration of human tetherin by
the human immunodeficiency virus type 2 envelope glycoprotein. J Virol 83(22):11966–11978
Lederer S, Favre D, Walters KA, Proll S, Kanwar B, Kasakow Z, Baskin CR, Palermo R, McCune JM,
Katze MG (2009) Transcriptional profiling in pathogenic and non-pathogenic SIV infections reveals
significant distinctions in kinetics and tissue compartmentalization. PLoS Pathog 5(2):e1000296
Leendertz SA, Locatelli S, Boesch C, Kucherer C, Formenty P, Liegeois F, Ayouba A, Peeters M,
Leendertz FH (2011) No evidence for transmission of SIVwrc from western red colobus
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 323
monkeys (Piliocolobus badius badius) to wild West African chimpanzees (Pan troglodytes
verus) despite high exposure through hunting. BMC Microbiol 11(1):24
Li Q, Smith AJ, Schacker TW, Carlis JV, Duan L, Reilly CS, Haase AT (2009) Microarray analysis
of lymphatic tissue reveals stage-specific, gene expression signatures in HIV-1 infection. J
Immunol 183(3):1975–1982
Liang SC, Tan XY, Luxenberg DP, Karim R, Dunussi-Joannopoulos K, Collins M, Fouser LA
(2006) Interleukin (IL)-22 and IL-17 are coexpressed by Th17 cells and cooperatively enhance
expression of antimicrobial peptides. J Exp Med 203(10):2271–2279
Liegeois F, Boue V, Mouacha F, Butel C, Mve-Ondo B, Pourrut X, Leroy E, Peeters M, Rouet F
(2012) New STLV-3 strains and a divergent SIVmus strain identified in non-human primate
bushmeat in Gabon. Retrovirology 9(1):28
Ling B, Apetrei C, Pandrea I, Veazey RS, Lackner AA, Gormus B, Marx PA (2004) Classic AIDS
in a sooty mangabey after an 18-year natural infection. J Virol 78(16):8902–8908
Ling B, Veazey RS, Hart M, Lackner AA, Kuroda M, Pahar B, Marx PA (2007) Early restoration
of mucosal CD4 memory CCR5 T cells in the gut of SIV-infected rhesus predicts long term
non-progression. AIDS 21(18):2377–2385
Locatelli S, Liegeois F, Lafay B, Roeder AD, Bruford MW, Formenty P, Noe R, Delaporte E,
Peeters M (2008) Prevalence and genetic diversity of simian immunodeficiency virus infection
in wild-living red colobus monkeys (Piliocolobus badius badius) from the Tai forest, Cote
d’Ivoire SIVwrc in wild-living western red colobus monkeys. Infect Genet Evol 8(1):1–14
Lozano Reina JM, Favre D, Kasakow Z, Mayau V, Nugeyre MT, Ka T, Faye A, Miller CJ, Scott-
Algara D, McCune JM et al (2009) Gag p27-specific B- and T-cell responses in Simian immu-
nodeficiency virus SIVagm-infected African green monkeys. J Virol 83(6):2770–2777
Mandl JN, Barry AP, Vanderford TH, Kozyr N, Chavan R, Klucking S, Barrat FJ, Coffman RL,
Staprans SI, Feinberg MB (2008) Divergent TLR7 and TLR9 signaling and type I interferon
production distinguish pathogenic and nonpathogenic AIDS virus infections. Nat Med
14(10):1077–1087
Martinez-Steele E, Awasana AA, Corrah T, Sabally S, van der Sande M, Jaye A, Togun T, Sarge-
Njie R, McConkey SJ, Whittle H et al (2007) Is HIV-2- induced AIDS different from HIV-1-
associated AIDS? Data from a West African clinic. AIDS 21(3):317–324
Marx PA, Alcabes PG, Drucker E (2001) Serial human passage of simian immunodeficiency virus
by unsterile injections and the emergence of epidemic human immunodeficiency virus in
Africa. Philos Trans R Soc Lond B Biol Sci 356(1410):911–920
Mattapallil JJ, Douek DC, Hill B, Nishimura Y, Martin M, Roederer M (2005) Massive infection
and loss of memory CD4+ T cells in multiple tissues during acute SIV infection. Nature
434(7037):1093–1097
May RM, Gupta S, McLean AR (2001) Infectious disease dynamics: What characterizes a success-
ful invader? Philos Trans R Soc Lond B Biol Sci 356(1410):901–910
McClure HM, Anderson DC, Fultz PN, Ansari AA, Lockwood E, Brodie A (1989) Spectrum of
disease in macaque monkeys chronically infected with SIV/SMM. Vet Immunol Immunopathol
21(1):13–24
Mehandru S, Poles MA, Tenner-Racz K, Horowitz A, Hurley A, Hogan C, Boden D, Racz P,
Markowitz M (2004) Primary HIV-1 infection is associated with preferential depletion of CD4+
T lymphocytes from effector sites in the gastrointestinal tract. J Exp Med 200(6):761–770
Melkus MW, Estes JD, Padgett-Thomas A, Gatlin J, Denton PW, Othieno FA, Wege AK, Haase AT,
Garcia JV (2006) Humanized mice mount specific adaptive and innate immune responses to
EBV and TSST-1. Nat Med 12(11):1316–1322
Mellors JW, Rinaldo CR Jr, Gupta P, White RM, Todd JA, Kingsley LA (1996) Prognosis in HIV-1
infection predicted by the quantity of virus in plasma. Science 272(5265):1167–1170
Meythaler M, Martinot A, Wang Z, Pryputniewicz S, Kasheta M, Ling B, Marx PA, O’Neil S, Kaur
A (2009) Differential CD4+ T-lymphocyte apoptosis and bystander T-cell activation in rhesus
macaques and sooty mangabeys during acute simian immunodeficiency virus infection. J Virol
83(2):572–583
324 E.J.D. Greenwood et al.
Milush JM, Mir KD, Sundaravaradan V, Gordon SN, Engram J, Cano CA, Reeves JD, Anton E, O’Neill
E, Butler E et al (2011) Lack of clinical AIDS in SIV-infected sooty mangabeys with significant
CD4+ T cell loss is associated with double-negative T cells. J Clin Invest 121(3):1102–1110
Molez JF (1998) The historical question of acquired immunodeficiency syndrome in the 1960s in
the Congo River basin area in relation to cryptococcal meningitis. Am J Trop Med Hyg
58(3):273–276
Moll M, Andersson SK, Smed-Sorensen A, Sandberg JK (2010) Inhibition of lipid antigen presen-
tation in dendritic cells by HIV-1 Vpu interference with CD1d recycling from endosomal com-
partments. Blood 116(11):1876–1884
Morgan D, Mahe C, Mayanja B, Okongo JM, Lubega R, Whitworth JA (2002) HIV-1 infection in
rural Africa: is there a difference in median time to AIDS and survival compared with that in
industrialized countries? AIDS 16(4):597–603
Muller MC, Saksena NK, Nerrienet E, Chappey C, Herve VM, Durand JP, Legal-Campodonico P,
Lang MC, Digoutte JP, Georges AJ et al (1993) Simian immunodeficiency viruses from central
and western Africa: evidence for a new species-specific lentivirus in tantalus monkeys. J Virol
67(3):1227–1235
Murphey-Corb M, Martin LN, Rangan SR, Baskin GB, Gormus BJ, Wolf RH, Andes WA, West M,
Montelaro RC (1986) Isolation of an HTLV-III-related retrovirus from macaques with simian
AIDS and its possible origin in asymptomatic mangabeys. Nature 321(6068):435–437
Neel C, Etienne L, Li Y, Takehisa J, Rudicell RS, Bass IN, Moudindo J, Mebenga A, Esteban A,
Van Heuverswyn F et al (2010) Molecular epidemiology of simian immunodeficiency virus
infection in wild-living gorillas. J Virol 84(3):1464–1476
Neil SJ, Zang T, Bieniasz PD (2008) Tetherin inhibits retrovirus release and is antagonized by
HIV-1 Vpu. Nature 451(7177):425–430
Nowroozalizadeh S, Mansson F, da Silva Z, Repits J, Dabo B, Pereira C, Biague A, Albert J,
Nielsen J, Aaby P et al (2010) Microbial translocation correlates with the severity of both
HIV-1 and HIV-2 infections. J Infect Dis 201(8):1150–1154
Ogg GS, Jin X, Bonhoeffer S, Dunbar PR, Nowak MA, Monard S, Segal JP, Cao Y, Rowland-Jones
SL, Cerundolo V et al (1998) Quantitation of HIV-1-specific cytotoxic T lymphocytes and
plasma load of viral RNA. Science 279(5359):2103–2106
Ostrowski MA, Chun TW, Justement SJ, Motola I, Spinelli MA, Adelsberger J, Ehler LA, Mizell
SB, Hallahan CW, Fauci AS (1999) Both memory and CD45RA+/CD62L+ naive CD4(+) T
cells are infected in human immunodeficiency virus type 1-infected individuals. J Virol
73(8):6430–6435
Paiardini M, Cervasi B, Reyes-Aviles E, Micci L, Ortiz AM, Chahroudi A, Vinton C, Gordon SN,
Bosinger SE, Francella N et al (2011) Low levels of SIV infection in sooty mangabey central
memory CD4(+) T cells are associated with limited CCR5 expression. Nat Med
17(7):830–836
Pandrea I, Apetrei C, Gordon S, Barbercheck J, Dufour J, Bohm R, Sumpter B, Roques P, Marx
PA, Hirsch VM et al (2007a) Paucity of CD4+CCR5+ T cells is a typical feature of natural SIV
hosts. Blood 109(3):1069–1076
Pandrea IV, Gautam R, Ribeiro RM, Brenchley JM, Butler IF, Pattison M, Rasmussen T, Marx PA,
Silvestri G, Lackner AA et al (2007b) Acute loss of intestinal CD4+ T cells is not predictive of
simian immunodeficiency virus virulence. J Immunol 179(5):3035–3046
Phair J, Munoz A, Detels R, Kaslow R, Rinaldo C, Saah A (1990) The risk of Pneumocystis carinii
pneumonia among men infected with human immunodeficiency virus type 1. Multicenter
AIDS Cohort Study Group. N Engl J Med 322(3):161–165
Phillips A, Lee CA, Elford J, Janossy G, Bofill M, Timms A, Kernoff PB (1989) Prediction of
progression to AIDS by analysis of CD4 lymphocyte counts in a haemophilic cohort. AIDS
3(11):737–741
Piot P, Quinn TC, Taelman H, Feinsod FM, Minlangu KB, Wobin O, Mbendi N, Mazebo P, Ndangi
K, Stevens W et al (1984) Acquired immunodeficiency syndrome in a heterosexual population
in Zaire. Lancet 2(8394):65–69
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 325
Plantier JC, Leoz M, Dickerson JE, De Oliveira F, Cordonnier F, Lemee V, Damond F, Robertson
DL, Simon F (2009) A new human immunodeficiency virus derived from gorillas. Nat Med
15(8):871–872
Poulsen AG, Aaby P, Larsen O, Jensen H, Naucler A, Lisse IM, Christiansen CB, Dias F, Melbye
M (1997) 9-year HIV-2-associated mortality in an urban community in Bissau, West Africa.
Lancet 349(9056):911–914
Quinn TC, Mann JM, Curran JW, Piot P (1986) AIDS in Africa: an epidemiologic paradigm.
Science 234(4779):955–963
Redd AD, Dabitao D, Bream JH, Charvat B, Laeyendecker O, Kiwanuka N, Lutalo T, Kigozi G,
Tobian AA, Gamiel J et al (2009) Microbial translocation, the innate cytokine response, and
HIV-1 disease progression in Africa. Proc Natl Acad Sci USA 106(16):6718–6723
Ricard D, Wilkins A, N'Gum PT, Hayes R, Morgan G, Da Silva AP, Whittle H (1994) The effects
of HIV-2 infection in a rural area of Guinea-Bissau. AIDS 8(7):977–982
Riddick NE, Hermann EA, Loftin LM, Elliott ST, Wey WC, Cervasi B, Taaffe J, Engram JC, Li B,
Else JG et al (2010) A novel CCR5 mutation common in sooty mangabeys reveals SIVsmm
infection of CCR5-null natural hosts and efficient alternative coreceptor use in vivo. PLoS
Pathog 6(8):e1001064
Rieder P, Joos B, von Wyl V, Kuster H, Grube C, Leemann C, Boni J, Yerly S, Klimkait T, Burgisser
P et al (2010) HIV-1 transmission after cessation of early antiretroviral therapy among men
having sex with men. AIDS 24(8):1177–1183
Rodriguez B, Sethi AK, Cheruvu VK, Mackay W, Bosch RJ, Kitahata M, Boswell SL, Mathews
WC, Bangsberg DR, Martin J et al (2006) Predictive value of plasma HIV RNA level on rate of
CD4 T-cell decline in untreated HIV infection. JAMA 296(12):1498–1506
Rutjens E, Vermeulen J, Verstrepen B, Hofman S, Prins JM, Srivastava I, Heeney JL, Koopman G
(2008) Chimpanzee CD4+ T cells are relatively insensitive to HIV-1 envelope-mediated inhibi-
tion of CD154 up-regulation. Eur J Immunol 38(4):1164–1172
Rutjens E, Mazza S, Biassoni R, Koopman G, Ugolotti E, Fogli M, Dubbes R, Costa P, Mingari
MC, Greenwood EJ et al (2010) CD8+ NK cells are predominant in chimpanzees, character-
ized by high NCR expression and cytokine production, and preserved in chronic HIV-1 infec-
tion. Eur J Immunol 40(5):1440–1450
Saez-Cirion A, Lacabaratz C, Lambotte O, Versmisse P, Urrutia A, Boufassa F, Barre-Sinoussi F,
Delfraissy JF, Sinet M, Pancino G et al (2007) HIV controllers exhibit potent CD8 T cell capac-
ity to suppress HIV infection ex vivo and peculiar cytotoxic T lymphocyte activation pheno-
type. Proc Natl Acad Sci USA 104(16):6776–6781
Schacker TW, Nguyen PL, Beilman GJ, Wolinsky S, Larson M, Reilly C, Haase AT (2002)
Collagen deposition in HIV-1 infected lymphatic tissues and T cell homeostasis. J Clin Invest
110(8):1133–1139
Schellens IM, Borghans JA, Jansen CA, De Cuyper IM, Geskus RB, van Baarle D, Miedema F
(2008) Abundance of early functional HIV-specific CD8+ T cells does not predict AIDS-free
survival time. PLoS One 3(7):e2745
Schindler M, Munch J, Kutsch O, Li H, Santiago ML, Bibollet-Ruche F, Muller-Trutwin MC,
Novembre FJ, Peeters M, Courgnaud V et al (2006) Nef-mediated suppression of T cell activa-
tion was lost in a lentiviral lineage that gave rise to HIV-1. Cell 125(6):1055–1067
Schmokel J, Sauter D, Schindler M, Leendertz FH, Bailes E, Dazza MC, Saragosti S, Bibollet-
Ruche F, Peeters M, Hahn BH et al (2010) The presence of a vpu gene and the lack of Nef-
mediated downmodulation of T cell receptor-CD3 are not always linked in primate lentiviruses.
J Virol 85(2):742–752
Sedaghat AR, German J, Teslovich TM, Cofrancesco J Jr, Jie CC, Talbot CC Jr, Siliciano RF
(2008) Chronic CD4+ T-cell activation and depletion in human immunodeficiency virus type 1
infection: type I interferon-mediated disruption of T-cell dynamics. J Virol 82(4):1870–1883
Seiler BM, Dick EJ Jr, Guardado-Mendoza R, VandeBerg JL, Williams JT, Mubiru JN, Hubbard
GB (2009) Spontaneous heart disease in the adult chimpanzee (Pan troglodytes). J Med
Primatol 38(1):51–58
326 E.J.D. Greenwood et al.
Shah AH, Sowrirajan B, Davis ZB, Ward JP, Campbell EM, Planelles V, Barker E (2010)
Degranulation of natural killer cells following interaction with HIV-1-infected cells is hindered
by downmodulation of NTB-A by Vpu. Cell Host Microbe 8(5):397–409
Sharova N, Wu Y, Zhu X, Stranska R, Kaushik R, Sharkey M, Stevenson M (2008) Primate lenti-
viral Vpx commandeers DDB1 to counteract a macrophage restriction. PLoS Pathog 4(5):
e1000057
Sharp PM, Hahn BH (2011) Origins of HIV and the AIDS pandemic. Cold Spring Harb Perspect
Med 1(1):a006841
Sharp PM, Bailes E, Gao F, Beer BE, Hirsch VM, Hahn BH (2000) Origins and evolution of AIDS
viruses: estimating the time-scale. Biochem Soc Trans 28(2):275–282
Sharp PM, Shaw GM, Hahn BH (2005) Simian immunodeficiency virus infection of chimpanzees.
J Virol 79(7):3891–3902
Sieg SF, Rodriguez B, Asaad R, Jiang W, Bazdar DA, Lederman MM (2005) Peripheral S-phase T
cells in HIV disease have a central memory phenotype and rarely have evidence of recent T cell
receptor engagement. J Infect Dis 192(1):62–70
Silvestri G, Sodora DL, Koup RA, Paiardini M, O'Neil SP, McClure HM, Staprans SI, Feinberg
MB (2003) Nonpathogenic SIV infection of sooty mangabeys is characterized by limited
bystander immunopathology despite chronic high-level viremia. Immunity 18(3):441–452
Silvestri G, Fedanov A, Germon S, Kozyr N, Kaiser WJ, Garber DA, McClure H, Feinberg MB,
Staprans SI (2005) Divergent host responses during primary simian immunodeficiency virus
SIVsm infection of natural sooty mangabey and nonnatural rhesus macaque hosts. J Virol
79(7):4043–4054
Singh DK, McCormick C, Pacyniak E, Lawrence K, Dalton SB, Pinson DM, Sun F, Berman NE,
Calvert M, Gunderson RS et al (2001) A simian human immunodeficiency virus with a non-
functional Vpu (deltavpuSHIV(KU-1bMC33)) isolated from a macaque with neuroAIDS has
selected for mutations in env and nef that contributed to its pathogenic phenotype. Virology
282(1):123–140
Singh DK, Griffin DM, Pacyniak E, Jackson M, Werle MJ, Wisdom B, Sun F, Hout DR, Pinson
DM, Gunderson RS et al (2003) The presence of the casein kinase II phosphorylation sites of
Vpu enhances the CD4(+) T cell loss caused by the simian-human immunodeficiency virus
SHIV(KU-lbMC33) in pig-tailed macaques. Virology 313(2):435–451
Souquiere S, Bibollet-Ruche F, Robertson DL, Makuwa M, Apetrei C, Onanga R, Kornfeld C,
Plantier JC, Gao F, Abernethy K et al (2001) Wild Mandrillus sphinx are carriers of two types
of lentivirus. J Virol 75(15):7086–7096
Stacey AR, Norris PJ, Qin L, Haygreen EA, Taylor E, Heitman J, Lebedeva M, DeCamp A, Li D,
Grove D et al (2009) Induction of a striking systemic cytokine cascade prior to peak viremia in
acute human immunodeficiency virus type 1 infection, in contrast to more modest and delayed
responses in acute hepatitis B and C virus infections. J Virol 83(8):3719–3733
Stephens EB, McCormick C, Pacyniak E, Griffin D, Pinson DM, Sun F, Nothnick W, Wong SW,
Gunderson R, Berman NE et al (2002) Deletion of the vpu sequences prior to the env in a
simian-human immunodeficiency virus results in enhanced Env precursor synthesis but is less
pathogenic for pig-tailed macaques. Virology 293(2):252–261
Strebel K, Klimkait T, Martin MA (1988) A novel gene of HIV-1, vpu, and its 16-kilodalton prod-
uct. Science 241(4870):1221–1223
Stremlau M, Owens CM, Perron MJ, Kiessling M, Autissier P, Sodroski J (2004) The cytoplasmic
body component TRIM5alpha restricts HIV-1 infection in Old World monkeys. Nature
427(6977):848–853
Switzer WM, Qari SH, Wolfe ND, Burke DS, Folks TM, Heneine W (2006) Ancient origin and
molecular features of the novel human T-lymphotropic virus type 3 revealed by complete
genome analysis. J Virol 80(15):7427–7438
Taaffe J, Chahroudi A, Engram J, Sumpter B, Meeker T, Ratcliffe S, Paiardini M, Else J, Silvestri
G (2010) A five-year longitudinal analysis of sooty mangabeys naturally infected with simian
immunodeficiency virus reveals a slow but progressive decline in CD4+ T-cell count whose
magnitude is not predicted by viral load or immune activation. J Virol 84(11):5476–5484
The Evolution of SIV in Primates and the Emergence of the Pathogen of AIDS 327
Takemura T, Hayami M (2004) Phylogenetic analysis of SIV derived from mandrill and drill. Front
Biosci 9:513–520
Traina-Dorge V, Blanchard J, Martin L, Murphey-Corb M (1992) Immunodeficiency and lympho-
proliferative disease in an African green monkey dually infected with SIV and STLV-I. AIDS
Res Hum Retroviruses 8(1):97–100
Tugume SB, Piwowar EM, Lutalo T, Mugyenyi PN, Grant RM, Mangeni FW, Pattishall K,
Katongole-Mbidde E (1995) Hematological reference ranges among healthy Ugandans. Clin
Diagn Lab Immunol 2(2):233–235
Vallari A, Bodelle P, Ngansop C, Makamche F, Ndembi N, Mbanya D, Kaptue L, Gurtler LG,
McArthur CP, Devare SG et al (2010a) Four new HIV-1 group N isolates from Cameroon:
prevalence continues to be low. AIDS Res Hum Retroviruses 26(1):109–115
Vallari A, Holzmayer V, Harris B, Yamaguchi J, Ngansop C, Makamche F, Mbanya D, Kaptue L,
Ndembi N, Gurtler L et al (2010b) Confirmation of putative HIV-1 group P in Cameroon.
J Virol 85(3):1403–1407
van der Loeff MF, Larke N, Kaye S, Berry N, Ariyoshi K, Alabi A, van Tienen C, Leligdowicz A,
Sarge-Njie R, da Silva Z et al (2010) Undetectable plasma viral load predicts normal survival
in HIV-2-infected people in a West African village. Retrovirology 7:46
Van Dooren S, Salemi M, Vandamme AM (2001) Dating the origin of the African human T-cell
lymphotropic virus type-i (HTLV-I) subtypes. Mol Biol Evol 18(4):661–671
Van Heuverswyn F, Li Y, Neel C, Bailes E, Keele BF, Liu W, Loul S, Butel C, Liegeois F, Bienvenue
Y et al (2006) Human immunodeficiency viruses: SIV infection in wild gorillas. Nature
444(7116):164
Veazey RS, DeMaria M, Chalifoux LV, Shvetz DE, Pauley DR, Knight HL, Rosenzweig M,
Johnson RP, Desrosiers RC, Lackner AA (1998) Gastrointestinal tract as a major site of CD4+
T cell depletion and viral replication in SIV infection. Science 280(5362):427–431
Vergne L, Bourgeois A, Mpoudi-Ngole E, Mougnutou R, Mbuagbaw J, Liegeois F, Laurent C,
Butel C, Zekeng L, Delaporte E et al (2003) Biological and genetic characteristics of HIV
infections in Cameroon reveals dual group M and O infections and a correlation between
SI-inducing phenotype of the predominant CRF02_AG variant and disease stage. Virology
310(2):254–266
Vidal N, Peeters M, Mulanga-Kabeya C, Nzilambi N, Robertson D, Ilunga W, Sema H, Tshimanga
K, Bongo B, Delaporte E (2000) Unprecedented degree of human immunodeficiency virus type
1 (HIV-1) group M genetic diversity in the Democratic Republic of Congo suggests that the
HIV-1 pandemic originated in Central Africa. J Virol 74(22):10498–10507
Wertheim JO, Worobey M (2009) Dating the age of the SIV lineages that gave rise to HIV-1 and
HIV-2. PLoS Comput Biol 5(5):e1000377
Wilkinson DA, Operskalski EA, Busch MP, Mosley JW, Koup RA (1998) A 32-bp deletion within
the CCR5 locus protects against transmission of parenterally acquired human immunodefi-
ciency virus but does not affect progression to AIDS-defining illness. J Infect Dis 178(4):
1163–1166
Willey RL, Maldarelli F, Martin MA, Strebel K (1992) Human immunodeficiency virus type 1 Vpu
protein induces rapid degradation of CD4. J Virol 66(12):7193–7200
Worobey M, Gemmel M, Teuwen DE, Haselkorn T, Kunstman K, Bunce M, Muyembe JJ, Kabongo
JM, Kalengayi RM, Van Marck E et al (2008) Direct evidence of extensive diversity of HIV-1
in Kinshasa by 1960. Nature 455(7213):661–664
Worobey M, Telfer P, Souquiere S, Hunter M, Coleman CA, Metzger MJ, Reed P, Makuwa M,
Hearn G, Honarvar S et al (2010) Island biogeography reveals the deep history of SIV. Science
329(5998):1487
Zeng M, Smith AJ, Wietgrefe SW, Southern PJ, Schacker TW, Reilly CS, Estes JD, Burton GF,
Silvestri G, Lifson JD et al (2011) Cumulative mechanisms of lymphoid tissue fibrosis and T
cell depletion in HIV-1 and SIV infections. J Clin Invest 121(3):998–1008
Zhang F, Wilson SJ, Landford WC, Virgen B, Gregory D, Johnson MC, Munch J, Kirchhoff F,
Bieniasz PD, Hatziioannou T (2009) Nef proteins from simian immunodeficiency viruses are
tetherin antagonists. Cell Host Microbe 6(1):54–67
Part III
Primates, Pathogens and Health
Microbial Exposures and Other Early
Childhood Influences on the Subsequent
Function of the Immune System
Introduction
J.F. Brinkworth and K. Pechenkina (eds.), Primates, Pathogens, and Evolution, 331
Developments in Primatology: Progress and Prospects,
DOI 10.1007/978-1-4614-7181-3_11, © Springer Science+Business Media New York 2013
332 G.A.W. Rook
siblings (especially older male siblings) in their family when they were 11 years old
(Strachan 1989). Then Matricardi and colleagues found that army recruits with evi-
dence of infections attributable to faecal–oral transmission were less likely to have
allergic manifestations (Matricardi et al. 1998). Such data were considered consis-
tent with a protective influence of postnatal infection that might be lost in the pres-
ence of modern hygiene (Matricardi et al. 1998; Strachan 1989; Strachan et al.
1996). A few years later it was pointed out that type 1 diabetes (T1D; caused by
autoimmune destruction of the insulin-secreting β-cells in the pancreas) is increas-
ing at the same rate and in the same countries (mostly rich and developed) as the
allergic disorders (Stene and Nafstad 2001). Similarly, a parallel rise in inflamma-
tory bowel diseases (IBD; Crohn’s disease and ulcerative colitis) had clearly started
at the beginning of the twentieth century, rising from rare and sporadic in 1900 to
400–500/100,000 by the 1990s in rich northern developed countries (the epidemio-
logical basis for these assertions is expanded below) (Elliott et al. 2005).
How can we make sense of all this? There are two obvious problems. First, aller-
gic disorders, autoimmune diseases and inflammatory bowel diseases involve dif-
ferent immune mechanisms. Allergic disorders are mediated largely by T helper 2
cells (Th2) secreting IL-4, IL-5 and IL-13, while the autoimmune disorders that are
increasing (multiple sclerosis (MS) and T1D) involve T helper 1 (Th1) and T helper
17 (Th17 cells) secreting IFN-γ, IL-17 and other mediators. The various forms of
IBD (Crohn’s disease, ulcerative colitis and celiac disease) involve a variety of
effector cell types. Is there an “umbrella” concept associated with hygiene or eco-
nomic development that can explain increases in pathologies that involve such
diverse effector cell types?
The second problem is the diversity of the epidemiology. What is the connection
between economic development, latitude, the farming environment, pet ownership
and dirty older siblings? Indeed, is there any possibility of finding a mechanism
with such broad explanatory power?
Since about 1998 it has been apparent that the first problem can be solved by the
view that the increase in the prevalence of these chronic inflammatory diseases is
due at least in part to defects in maturation of the immune system, leading notably
to defective immunoregulation (Rook and Stanford 1998). Thus, one cause of these
increases lies in a broad imbalance between immunoregulatory mechanisms and
effector mechanisms (whether Th1, Th17 or Th2). A failure of immunoregulatory
mechanisms can indeed lead to simultaneous increases in diverse types of pathol-
ogy. We know this because genetic defects of Foxp3, a transcription factor that
plays a crucial role in the development and function of regulatory lymphocytes,
leads to the X-linked autoimmunity–allergic dysregulation syndrome (XLAAD)
that includes aspects of allergy, autoimmunity and enteropathy (Wildin et al. 2002).
So although the recent increases in chronic inflammatory disorders are too rapid to
be due to genetic changes, there could be a lifestyle-associated factor that has a
broad detrimental effect on regulation of the immune system.
More recently the second problem (i.e. the factor that is common to economic
deprivation, farms, siblings, dogs and poor hygiene) has also been resolved. As will
be explained in detail below, we now know that the correct functioning of the
Microbial Exposures and Other Early Childhood Influences… 333
immune system is heavily dependent on the microbial input that it receives in utero,
in the neonatal period and also throughout life (Ege et al. 2008; Rook 2010).
Epidemiological and experimental studies have progressively identified organisms
that have appropriate immunoregulatory properties, are depleted from modern envi-
ronments and can downregulate chronic inflammatory disorders in experimental
models. Some of these are now entering clinical trials in humans. In the context of
this book, the interesting thing about the identified organisms is that not only do
they resolve the farm/dog/sibling dilemma, but they also take us back at least as far
as early hominins. The organisms identified are viruses, bacteria and helminths,
often harmless, that coevolved with humans but are depleted from the modern urban
environment. These can be divided into three overlapping groups: (1) organisms
that form part of the coevolved human microbiota that are altered by modern diets,
living conditions and antibiotics; (2) infections that will have been present in early
humans, usually harmless, transmitted by the faecal–oral route very early in life,
that have been depleted since the second epidemiological transition; and (3) harm-
less environmental organisms in mud, untreated water and fermenting vegetable
material that are eliminated by the modern city lifestyle. I refer to these organisms
collectively as “Old Friends”, to emphasise our long association with them and our
dependency on their presence (Rook 2010). Below, I discuss the evolutionary sig-
nificance and mechanisms of the immunoregulatory effects of these organisms.
As so often in medicine, a Darwinian approach has huge explanatory power. But
it is emphasised that this chapter is drawn from the medical literature and discusses
the effects on the immune system of recent human cultural and technological devel-
opments. I leave to others the interesting task of asking whether there were other
similar immunological turning points at earlier stages in hominoid evolution.
This chapter starts with a general account of the role of the immunoregulatory
“Old Friends”, in which immunology has been kept to a minimum. Those who want
more detail of the immunology and molecular mechanisms will find that in the later
section entitled “Mechanisms”.
dropped from the genome of one of them. Access to that gene is now “entrusted” to
the other species. This idea is at first surprising to most readers, but it is in fact rather
commonplace. For instance, most mammals can synthesise vitamin C, but large
primates and guinea pigs have lost the relevant pathways. In effect, man and guinea
pig are now in a state of evolved dependence on fruit and vegetables. Of course the
same is true for many other genes involved in the synthesis of vitamins and other
essential nutrients that we have to consume after other organisms have created them
for us (Resta 2009).
The role of this evolved dependence is most clearly seen in germ-free animals,
propagated by caesarean section under sterile conditions and maintained in a sterile
environment. The immune systems of germ-free animals fail to develop correctly
and are functionally distorted. There is lack of cellularity and, above all, lack of
effective immunoregulation (discussed in detail later). In 2005 Mazmanian and col-
leagues showed that a single polysaccharide from an intestinal commensal,
Bacteroides fragilis, could partly correct these developmental abnormalities
(Mazmanian et al. 2005). More recently they have shown, using three different
models of intestinal inflammation, that the same polysaccharide, given by mouth,
can turn on crucial anti-inflammatory, immunoregulatory pathways (Mazmanian
et al. 2008). In the discussion of the latter paper they state:
We propose that the mammalian genome does not encode for all functions required for
immunological development but rather that mammals depend on critical interactions with
their microbiome (the collective genomes of the microbiota) for health.
To put it even more simply, some genes needed for the development and regula-
tion of the mammalian immune system might have been “entrusted” to microorgan-
isms: a clear example of “evolved dependence”. It is obvious that these organisms
have to be those with which mammals have coevolved for a very long time and that
were always present. They cannot be organisms that merely cause sporadic infec-
tions or high death rates. The latter can modify the human genome by elimination
of susceptible genotypes, but they cannot be entrusted with the role of supplying
genes and functions that we need.
It may be useful to think about the state of evolved dependence seen in the mam-
malian immune system in the context of the Environment of Evolutionary
Adaptedness (EEA). The term EEA was first used in 1969 by John Bowlby, who
was concerned that those aspects of human behaviour that are genetically deter-
mined might be adapted to the hunter-gatherer existence rather than to modern city
life (Bowlby 1971 first published by Hogarth press in 1969). The basis for this was
the view that since the start of agriculture and pastoralism about 10,000 years ago,
much human adaptation to new environments has been cultural and technological
rather than genetic. For example, we have not adapted genetically to living in cold
Microbial Exposures and Other Early Childhood Influences… 335
places: we have learnt to make fur coats and electric heaters. Humans easily detect
problems within the physical environment and invent appropriate technological
adaptations. But humans are not equipped with a sense that tells us when the envi-
ronment is inappropriate for our immune systems. Only since the development of
modern biology have we begun to understand that we have an immune system and
that we need to think in a Darwinian and anthropological way about the nature and
timing of the inputs that it receives.
Epidemiological Transitions
It is possible to predict the identity of the organisms on which our immune systems
have evolved dependence (i.e. that were part of the immune system’s EEA) by con-
sidering the history of man’s changing microbial exposures. In 1971 Omran coined
the term “epidemiological transition” to describe the major watersheds in human
development that led to massive changes in mortality (discussed in Armelagos et al.
2005). Palaeolithic populations carried organisms inherited from primate ancestors
(“heirloom” species), including many viruses, as shown in Fig. 1 (Armelagos et al.
2005; Van Blerkom 2003). In addition they would have been exposed to zoonoses
that they picked up as they scavenged carrion (Armelagos et al. 2005; Despres et al.
1992; Hoberg 2006). Phylogenetic trees for these organisms confirm the extremely
ancient association with humans, probably for a million years or more (Despres
et al. 1992; Hoberg 2006). Finally, Palaeolithic populations will have consumed
several milligrams of harmless environmental saprophytes every day, since these are
ubiquitous in soil and untreated water (Delmont et al. 2011). We have called these
“pseudocommensals” because of their inevitable continuous presence until chlori-
nation and purification of water in the modern era. The organisms that have been
found to be important for the hygiene hypothesis belong within these three
categories.
About 10,000 years ago, the shift to agriculture and husbandry created the first
(Neolithic) epidemiological transition (Fig. 1) (Armelagos et al. 2005). This will
have had little effect on exposure to the “pseudocommensals” or to the heirloom
species. However, the more sedentary lifestyle increased faecal–oral transmission
and combined with husbandry caused prolonged contact with animals. The latter
led to adaptation to humans of a number of animal viruses shown in Fig. 1 (Van
Blerkom 2003). However, the viruses acquired during the Neolithic such as influ-
enza (B and C), smallpox, mumps and measles cannot have become endemic until
populations were large enough. This required communities of several hundreds of
thousands, which did not occur until the appearance of cities 2000–3000 years ago
(Armelagos et al. 2005). Since this represents only 100–150 generations from the
present, extremely strong selection pressure would have been required for evolved
dependence to appear, and this seems unlikely. Moreover, most humans did not live
in such large groups, and these viruses were, for example, absent from pre-Colum-
bian American populations. In any case, one would not expect “evolved
336 G.A.W. Rook
Fig. 1 Aspects of man’s microbiological history that are most relevant to the hygiene hypothesis.
Epidemiological data, laboratory and animal models and preliminary clinical trials investigating
the hygiene hypothesis implicate organisms that are thought to have accompanied mammalian and
human evolution. This relationship was long enough for the establishment of evolved dependence
on these organisms, that must be tolerated, and so have developed roles in the initiation of regula-
tory pathways. Organisms that evolved during the Neolithic are less likely to be relevant in this
context, and the first epidemiological transition did not reduce human contact with organisms
associated with animals, faeces and mud. On the other hand, the second epidemiological transition
has led to gene–environment misfit, as the “Old Friends” from the Palaeolithic were progressively
removed from the modern environment
dependence” on contact with organisms that were both dangerous and sporadic, and
as explained later, these “recently” acquired viruses tend to cause or exacerbate
chronic inflammatory disorders, rather than prevent them.
In short, there were dramatic changes to the human microbial environment after
the first epidemiological transition, but this did not result in loss of exposure to the
organisms that had coevolved with humans and are now implicated by epidemiology,
experimental models and clinical trials in the hygiene hypothesis, as detailed below.
Until the modern era more than 97 % of the population still lived in rural environ-
ments, close to mud, animals and faeces, which were the sources of these organisms.
The situation did not change until the mid-nineteenth century. Since then some
populations have undergone a second epidemiological transition in which concrete,
tarmac, diminished contact with animals, public health measures and, more recently,
antibiotics have resulted in diminished (or delayed) exposure to the “Old Friends”
that were present in earlier eras. The dramatic effects of the second epidemiological
Microbial Exposures and Other Early Childhood Influences… 337
prevalent in Finnish Karelians than in Russian Karelians despite the fact that the geno-
types associated with susceptibility are at the same frequency in the two populations
(Kondrashova et al. 2005). Meanwhile striking and repeatedly confirmed large global
variations in the incidence of multiple sclerosis (Rosati 2001) (an autoimmune inflam-
matory disorder where the immune system attacks the brain) are found to correlate
inversely with the prevalence of the helminth Trichuris trichiura, which is an excellent
surrogate marker of economic underdevelopment and of the likelihood of exposure to
other developing country infections (Fleming and Cook 2006).
Inflammatory bowel diseases (IBD) are at least partly due to inappropriate, unregu-
lated immune responses to bowel contents (Boden and Snapper 2008). The early
medical literature suggests that IBD had been extremely rare before the 1900s (Kirsner
1995). Then during the early twentieth century IBD was uncommon, sporadic and
usually seen in individuals from the upper classes, urban areas and northern latitudes
(Kirsner 1995). Later in the twentieth century, the incidence increased steadily, as
proven in repeat studies in individual countries (England, USA, Scotland, Wales,
Denmark, Sweden, Israel), using reproducible methods (reviewed and referenced in
Elliott et al. 2005). Meanwhile the greater incidence of IBD in hygienic urban settings
continues to be obvious, and the incidence increases (using standardised diagnostic
criteria) as hygiene and modernisation progress (Klement et al. 2008).
Gut Microbiota
Proof, using modern methods, that the microbiota of city-dwelling European (EU)
children differs dramatically from that of rural Africans was obtained recently (De
Filippo et al. 2010). The faecal microbiota of children from Burkina Faso (BF) had
Microbial Exposures and Other Early Childhood Influences… 339
Entero-mammary circulation
Ectoparasites
Fig. 2 Locations of some of the organisms that can be regarded as “Old Friends” and that play a
role in setting up immunoregulatory circuits. AOB ammonia-oxidising bacteria
dog strikingly increases the abundance and diversity of bacterial taxa in house dust
(Fujimura et al. 2010). The role of prenatal exposure is supported by the observation
that pregnant women (not sensitised to dog) who had a dog or cat in the home had
higher circulating regulatory T cell (Treg) levels compared with women who did not
have pets (Wegienka et al. 2009). Treg readily suppress allergic responses (dis-
cussed in the mechanisms section below).
Organisms highlighted in recent studies of the hygiene hypothesis that are transmit-
ted by the faecal–oral route include H. pylori, Salmonella, hepatitis A virus (HAV),
enteroviruses and Toxoplasma gondii (Fig. 1 and Table 1) (Matricardi et al. 2000;
Pelosi et al. 2005; Seiskari et al. 2007a; Umetsu et al. 2005).
A number of studies have demonstrated an association between infection with the
hepatitis A virus (HAV) and protection against the development of asthma (Matricardi
et al. 2000; Umetsu et al. 2005). This infection was probably universal and coevolved
with humans (Van Blerkom 2003). It was transmitted to the neonate by faecal–oral
contamination and harmless if transmitted in this way (it is of course dangerous in
the modern world when infection occurs in an adult). Its incidence rapidly declined
during the twentieth century. HAV directly infects lymphocytes via the receptor
TIM-1 (T-cell immunoglobulin and mucin domain containing 1). This is thought to
alter differential survival of T cell subsets and so bias the immune response away
from allergy-mediating Th2 and towards immunoregulation (Umetsu et al. 2005).
Apart from HAV, mentioned in the previous paragraph, most studies implicate
childhood viruses as triggers of asthma rather than as protective organisms.
The viruses most frequently implicated as inducers of allergic disorders are human
rhinovirus (HRV) (Jackson 2010) and respiratory syncytial virus (RSV) (reviewed
inYoo et al. 2007). A prospective study of 95,310 children from birth through early
childhood showed that the risk of developing asthma was greatest for children born
approximately 4 months before the winter virus peak (Wu et al. 2008). This finding
implied a role for early exposure to winter viruses in the triggering of later asthma
(Wu et al. 2008). Animal models provide further support. It is possible to provoke a
chronic inflammatory Th2-biased airway disease in mice using Sendai virus. This is
a mouse parainfluenza-type I virus that is similar to RSV. In this mouse model acute
lung disease appears 3 weeks after infection when IL-13-producing CD4+ Th2 cells
are recruited to the airways. Seven weeks after infection there is a chronic phase
associated with continuing production of IL-13 and Th2-mediated inflammation
(Holtzman et al. 2009).
Microbial Exposures and Other Early Childhood Influences… 341
Table 1 (continued)
Organism or location Disease or model or effect References
Gut microbiota
Segmented filamentous Th17 cells (mice) Gaboriau-Routhiau et al. (2009),
bacteria Ivanov et al. (2009), Wu et al.
(2010)
Clostridia species Treg in lamina propria Geuking et al. (2011)
(mice)
Bacillus fragilis IL-10 and Treg (mice) Round et al. (2011)
Faecalibacterium Crohn’s disease (human Sokol et al. (2008)
prausnitzii and animal studies)
Other microbiota
Skin microbiota; ammonia- Nitrite, nitric oxide Whitlock and Feelisch (2009)
oxidising bacteria
Lung microbiota Asthma Huang et al. (2011)
Oral and periodontal IBD Singhal et al. (2011)
microbiota
Gut organisms transported ? immunoregulation Donnet-Hughes et al. (2010)
to breast milk
Ectoparasites
Various Response to TLR agonists Friberg et al. (2010)
in vitro
Environmental saprophyte
Mycobacterium vaccae Allergy (mouse, dog) Ricklin-Gutzwiller et al. (2007),
Zuany-Amorim et al. (2002)
KEY: T1D type 1 diabetes, IBD inflammatory bowel disease, EAE experimental autoimmune
encephalomyelitis, TLR toll-like receptor
Helminths
The role of recent changes to the microbial microbiota of the skin, lung and breast
has received almost no attention. Before the invention of modern soaps and deter-
gents, the skin was probably colonised by ammonia-oxidising bacteria (AOB)
(Fig. 2). These are ubiquitous in soil, but they are exquisitely sensitive to alkyl-
benzene sulfonate detergents (Whitlock and Feelisch 2009). AOB can convert the
high concentrations of urea and ammonia found in human sweat into nitrite and
NO which are efficiently absorbed via the skin, so this source of nitrite might have
been biologically significant for immunoregulation in which NO is known to play
a major role (Whitlock and Feelisch 2009).
The lung is not sterile. It is estimated that there are about 2,000 bacterial genomes per
square centimetre of the surface of the bronchial tree, though there is a risk of con-
tamination from the bronchoscope used to gather the samples. Pathogenic
Proteobacteria, particularly Haemophilus spp., were more abundant in asthmatic air-
ways (Hilty et al. 2010). Interestingly bacterial concentrations and diversity were sig-
nificantly higher among asthmatic patients, as determined by assaying 16S ribosomal
RNA amplicons and the relative abundance of members of several bacterial families
correlated with bronchial hyperresponsiveness (Huang et al. 2011). Nothing is yet
known of differences between lung microbiota of hunter-gatherers and modern urban
man, but they must be striking (Fig. 2). This will be a difficult area to study because
truly sterile samples of material from healthy individuals probably cannot be obtained.
Breast milk is not sterile (Donnet-Hughes et al. 2010), and there appears to be an
“entero-mammary” circulation. Human peripheral blood mononuclear cells and
breast milk cells contain bacteria during lactation, due to increased translocation
from the gut to Peyer’s patches and thence into circulating dendritic cells. Bacteria
can also be seen in the glandular tissue of healthy breast and in mononuclear cells
in breast milk, which when “sterile” contains small numbers of cultivable organisms
Microbial Exposures and Other Early Childhood Influences… 345
(<103). The mononuclear cells seem to be partially matured dendritic cells that
might promote tolerogenic responses in the neonate (Donnet-Hughes et al. 2010). It
remains to be seen whether this is an important part of the colonisation and educa-
tion of the neonatal gut and immune system, but if it is, it must be severely disrupted
in many modern humans (Fig. 2).
Not only does human milk contain bacteria as described above, but it also con-
tains complex oligosaccharides that favour growth and establishment of selected
strains such as Bifidobacterium longum subsp. infantis in the infant gut (Zivkovic
et al. 2010). Thus, the decline in breastfeeding due to pressure to use formula milk
substitutes must be altering infant microbiota.
Environmental “Pseudocommensals”
Mud and untreated water contain milligrams of various saprophytic bacterial spe-
cies per litre. Therefore, milligram quantities were inevitably consumed by every-
one until modern chlorinated water supplies were developed. We have designated
these “pseudocommensals” because they were consumed regularly and inevitably
throughout mammalian evolution (Fig. 2). These too turn out to have immunoregu-
latory roles and are currently entering further clinical trials (Ricklin-Gutzwiller
et al. 2007; Zuany-Amorim et al. 2002). Some non-colonising lactobacillus strains
present in fermenting vegetable matter must have been encountered daily, and many
Lactobacillus strains are immunoregulatory (Smits et al. 2005) and are entering
clinical trials as probiotics (Sheil et al. 2007).
Ectoparasites
In a study of wild rodents, it emerged that various aspects of the innate immune
response were profoundly affected by the load of ectoparasites (i.e. fleas, lice, etc.)
(Friberg et al. 2010). It is perhaps not surprising that repeated “injections” of phar-
macologically active materials, many of them designed to modulate the host immune
response, can provoke lasting systemic effects. Such exposures must be greatly
diminished in modern society, but further work is needed to determine the impor-
tance of this factor (Friberg et al. 2010).
Malaria clearly needs to be studied in this context, especially in view of its impor-
tance in human evolution. Plasmodium vivax induces Treg population expansion in
humans (Bueno et al. 2010). Similarly, placental malaria, whether active at delivery
346 G.A.W. Rook
If the depletion of the organisms listed in the previous sections is contributing to the
increases in disorders of immunoregulation that have occurred since the second
epidemiological transition, it should be possible to demonstrate that administration
of these organisms can prevent and/or treat the same conditions in animal models
and in human clinical trials. A few examples have been cited in context above, and
a detailed analysis would be beyond the scope of this review, but this issue is
expanded briefly below.
Animal Models
Clinical Trials
Mechanisms
How and why do the Old Friends modulate our immune systems? Numerous mech-
anisms exist, but there are two related underlying evolutionary principles that prob-
ably apply to all the “Old Friends”. First, the encounters with a broad range of
microbial molecules that trigger the innate immune system via pattern recognition
receptors (PRR) such as Toll-like receptors (TLR), and also experience of diverse
microbial antigens, may be a necessary maturation stimulus for the immune system.
Second, the “Old Friends” persist as commensals (or “pseudocommensals”), carrier
states or chronic subclinical infections. Therefore, the host–“Old Friend” relation-
ship evolved so that rather than provoking needless damaging aggressive immune
responses, an anti-inflammatory equilibrium is established. This translates into the
priming of immunoregulation. Maturation and immunoregulation are considered
separately below, though they are clearly overlapping concepts.
It is suggested that the neonatal immune system defaults to a Th2 bias, in the absence
of stimuli that drive maturation (Pfefferle et al. 2010). Since this is the arm of the
348 G.A.W. Rook
Anti-inflammatory
mechanisms
Helminths
IL-10+ regulatory B cells
Fig. 3 Some mechanisms involved in the immunoregulatory properties of the “Old Friends” and
microbiota. A key pathway is the modification of DC so that they tend to drive Treg. Such DC also
process self-antigens, allergens, etc. and so drive crucial specific regulatory cell populations. The
intestinal helminths also exert indirect effects by modulating the microbiota. The individual path-
ways (some taken from experimental systems in the mouse, as illustrations) are referenced in the
main text: RA retinoic acid, DC dendritic cell, Aldh1a2 retinaldehyde dehydrogenase 2, GPR43
G-protein-coupled receptor 43, RegIIIγ regenerating islet-derived 3γ, CD103 an integrin and
marker of intestinal regulatory DC, IDO indoleamine 2,3, dioxygenase
immune response that drives allergic disorders, this point is of some interest. It now
seems that children exposed to high quantities and high diversity of microbial materi-
als are protected from subsequent allergic disorders (Ege et al. 2011). Interestingly
this exposure can take place before birth. Exposing pregnant women to barns and
farm animals during gestation reduced the risk of allergic disease in their offspring
(Ege et al. 2008; von Mutius and Radon 2008). Similarly, prenatal exposure to house-
hold pets reduces fetal immunoglobulin E production (Aichbhaumik et al. 2008).
Maturation following microbial exposures seems to push the system towards Th1,
with increased production of the Th1 signature cytokine, IFN-γ (Pfefferle et al. 2010).
Interestingly, this protective effect of prenatal exposure to microbial components
has been reproduced in a mouse allergy model and strongly suggests that epigenetic
factors are involved (Conrad et al. 2009). It is also possible that accelerated matura-
tion and Th1 bias induced by microbial exposure indirectly protects from allergy by
facilitating rapid removal of the viruses that are implicated in allergic disorders
(RSV and HRV discussed earlier) (Jackson 2010; Yoo et al. 2007).
These points all raise the interesting issue that it might take several generations
for the full consequences of loss of exposure to “Old Friends” to be manifested.
This greatly complicates the epidemiology.
However, maturation and Th1 bias is not the whole story and might in fact be less
important than priming of immunoregulation. For example, farm exposures during
pregnancy increased the number and function of cord blood Treg cells, in addition
to reducing the extent of Th2 cytokine secretion after in vitro stimulation (Schaub
et al. 2009). Immunoregulation is discussed in the next sections.
Microbial Exposures and Other Early Childhood Influences… 349
Initiation of Immunoregulation
The Old Friends persist as commensals, carrier states or chronic subclinical infec-
tions and so coevolved a role as inducers of immunoregulation. This avoids point-
less and damaging inflammatory attacks on organisms that are essential to our
health (such as gut microbiota) or impossible to remove (such as established hel-
minth infections). For instance, a futile effort to destroy Brugia malayi microfilariae
results in lymphatic blockage and elephantiasis (Babu et al. 2006). Thus, many
helminths induce expansion of Treg populations. For instance, the percentage of
circulating Treg was positively related to the level of infection with Schistosoma
haematobium in Zimbabwean children aged 8–13 years (Nausch et al. 2011). Some
mechanisms are summarised in Fig. 3.
A frequent mechanism is modulation of dendritic cells (DC) such that these drive
Treg rather than Th1, Th17 or Th2 effector cells (referenced in Rook 2009). These
modified DC can be regarded as DCreg. Then the constitutive presence of the “Old
Friends” causes continuous background activation of the DCreg and of Treg specific
for the Old Friends themselves, resulting in background bystander suppression of
inflammation. Meanwhile these DCreg inevitably sample self, gut contents and
allergens and so induce Treg specific for the illicit target antigens of the three groups
of chronic inflammatory disorder. Release of the anti-inflammatory cytokines,
IL-10 and TGF-β is often involved in the anti-inflammatory effects of these cells
(Zuany-Amorim et al. 2002).
that binds the TGF-βRΙΙ and causes FoxP3-negative T cells to become functional
Foxp3+ Treg (Grainger et al. 2010). But this helminth also modulates DC function in
an anti-inflammatory way (Hang et al. 2010) and may exert indirect immunomodula-
tory effects via induced changes in the bacterial microbiota (Walk et al. 2010).
This topic was extensively reviewed recently, and Fig. 3 lists some of the known
mechanisms (Ehlers and Rook 2011; Round and Mazmanian 2009). In the 1980s it
was revealed that defined alterations to the microbiota could reproducibly either
increase or decrease susceptibility to autoimmune arthritis (Kohashi et al. 1985).
Similar findings have been published recently using experimental autoimmune
encephalomyelitis (EAE), a mouse model of multiple sclerosis (Lee et al. 2010).
Modulation of the bowel microbiota could alter susceptibility to EAE by mecha-
nisms that involved the ability of intestinal DC to prime Th1, Th17 or Treg responses
(Lee et al. 2010). Some bacteria have also been found to enhance numbers and
activity of Treg (Atarashi et al. 2011) or even to secrete single molecules that lead
directly to expansion of Treg populations (Round et al. 2011). Short-chain fatty
acids (SCFA) produced by many gut bacteria also have an anti-inflammatory role in
the gut. SCFAs bind the G-protein-coupled receptor 43 (GPR43, also known as
FFAR2) and exert an anti-inflammatory effect that proved relevant in models of
colitis, arthritis and asthma (Maslowski et al. 2009). Other mechanisms include
induction of regulatory B cells (Correale et al. 2008) and regulatory macrophages
(Schnoeller et al. 2008), modulation of Treg/Th17 balance (Ivanov et al. 2008) and
indirect effects via epithelial cell products that cause DC to drive Foxp3+ cells with
gut-homing properties (Iliev et al. 2009), and induced secretion of REGIIIγ, a
C-type lectin with bactericidal effects on Gram-positive bacteria (Cash et al. 2006).
Interestingly there is evidence that very clean Specific Pathogen-Free (SPF)
mice (i.e. animals with a “normal” microbiota, in theory lacking pathogens) have
abnormally functioning Treg that can fail to secrete IL-10 and can switch function
to an aggressive cell type (Erdman et al. 2010). Humans in rich Western cities are
not SPF, but some modern babies must be getting close to the SPF state, with less
diverse commensals and microbiota and little exposure to pathogens.
Fig. 4 Interaction of genetics and loss of the “Old Friends”. The Old Friends had to be tolerated
and so coevolved roles as triggers of immunoregulatory pathways. In areas with very high loads of
these and other organisms, particularly helminths, compensatory genetic variants accumulated, to
partially restore inflammatory responses. In the absence of the Old Friends, not only is immuno-
regulation inadequately primed, but also these genetic variants cause excessive inflammation and
become risk factors for chronic inflammatory disorders. Thus, genetic variants that were advanta-
geous and did not cause disease in the past start to do so in the absence of the Old Friends (refer-
enced in main text). Several aspects of modern life are potentially interacting with the lack of “Old
Friends” at the level of immunoregulation. Obesity is associated with altered gut microbiota and
excessive release of proinflammatory cytokines. Lack of vitamin D exacerbates immunodysregula-
tion, as does the triggering of Th17 cells by dioxins (changes in the microbiota also impact on
Th17 development). Diesel particulates drive Th2 cells. Social stressors also drive inflammation,
and delayed exposure to childhood viruses may cause them trigger allergy and autoimmunity
late (for instance, at weaning) can be protective when given very early (Filippi and
von Herrath 2008; Harrison et al. 2008; Serreze et al. 2000). The mechanism again
involves immunoregulation. Such viruses, at least in a mouse model, can activate
invariant natural killer T cells that induce TGF-β-producing plasmacytoid DCs
(pDC) in the pancreatic lymph nodes (Diana et al. 2011). These regulatory pDC
then drive development of CD4 + CD25+ Tregs that synergise with upregulated pro-
grammed cell death-1 ligand 1 (PD-L1) to shut off the autoimmune response (Filippi
et al. 2009).
In parts of the world where there was a heavy load of organisms causing immuno-
regulation, there has been selection for single nucleotide polymorphisms (SNP) or
other variants that partially compensate for the immunoregulation. This is seen for
several proinflammatory cytokines (Fumagalli et al. 2009) and IgE (Barnes et al.
352 G.A.W. Rook
The modern human diet encourages obesity, and adipose tissue releases proinflam-
matory mediators that will exacerbate the effects of any immunoregulatory defects
(Collins and Bercik 2009). Moreover, obesity is associated with phylum-level
changes in the microbiota and reduced bacterial diversity (Turnbaugh et al. 2009),
which will have immunoregulatory consequences (Round and Mazmanian 2009).
Psychological stress also modulates gut microbiota and gut permeability, while both
obesity and stress result in greater release of proinflammatory cytokines (Collins
and Bercik 2009). This all leads to a vicious circle, because the tendency to develop
obesity is modulated by the nature of the microbiota. For example, low levels of
Bifidobacteria and high levels of Staphylococcus aureus in infant microbiota may
predict the development of obesity later in life (Kalliomaki et al. 2008).
Vitamin D
Humans need sunlight to drive formation of vitamin D3, which is rarely present at
adequate levels in the diet. Vitamin D is involved in driving regulatory cells such as
Treg (Xystrakis et al. 2006). In rich developed countries deficiency of vitamin D is
increasingly common, partly because of fears of melanoma. This deficiency is
implicated in the increases in chronic inflammatory disorders, some cancers and
allergic disorders (Brehm et al. 2010; Herr et al. 2011; Honeyman and Harrison
2009). Moreover, vitamin D may also protect from allergic disorders indirectly by
enhancing immunity to respiratory viruses (Sabetta et al. 2010), some of which are
implicated in the causation of allergic disorders (Jackson 2010; Yoo et al. 2007).
This area has become deeply controversial following a recent report from the
Institute of Medicine in the USA (Ross et al. 2011) that recommended much lower
levels of vitamin D than most workers in the field would want to see (Heaney and
Holick 2011). The recommended levels are certainly likely to be lower than those
that would have been found in scantily clad hunter-gatherer humans.
Pollution
and allergic sensitisation (Riedl and Diaz-Sanchez 2005). There must be many other
novel molecules in our environments that have significant though poorly docu-
mented proinflammatory effects.
The realisation that the hygiene or “Old Friends” hypothesis is largely a matter of
immunoregulation leads to the possibility that several other groups of disease might
also be increasing as a consequence of diminished exposure to organisms with
which we coevolved (Rook 2010). Two examples are particularly worthy of men-
tion. First, chronically raised levels of proinflammatory cytokines cause symptoms
of depression and are often found in depressed patients (Raison et al. 2010). Second,
chronic inflammation is oncogenic, and once tumours have developed, ongoing
inflammation releases growth factors and angiogenic factors that encourage tumour
growth. The epidemiology of some human cancers that are increasing in frequency
is similar to that of the allergic disorders and of T1D, and it is likely that failed
immunoregulation is playing a role (Rook and Dalgleish 2011). Thus, the conse-
quences of a lifestyle that no longer resembles the EEA of the primate or even that
of early hominins may be very broad and still not fully revealed.
Conclusions
References
Aichbhaumik N, Zoratti EM, Strickler R, Wegienka G, Ownby DR, Havstad S, Johnson CC (2008)
Prenatal exposure to household pets influences fetal immunoglobulin E production. Clin Exp
Allergy 38(11):1787–1794
Araujo MI, Lopes AA, Medeiros M, Cruz AA, Sousa-Atta L, Sole D, Carvalho EM (2000) Inverse
association between skin response to aeroallergens and Schistosoma mansoni infection. Int
Arch Allergy Immunol 123(2):145–148
Armelagos GJ, Brown PJ, Turner B (2005) Evolutionary, historical and political economic per-
spectives on health and disease. Soc Sci Med 61(4):755–765
Atarashi K, Tanoue T, Shima T, Imaoka A, Kuwahara T, Momose Y, Cheng G, Yamasaki S, Saito
T, Ohba Y et al (2011) Induction of colonic regulatory T cells by indigenous Clostridium spe-
cies. Science 331:337–341
Babu S, Blauvelt CP, Kumaraswami V, Nutman TB (2006) Regulatory networks induced by live
parasites impair both Th1 and Th2 pathways in patent lymphatic filariasis: implications for
parasite persistence. J Immunol 176(5):3248–3256
Barnes KC, Grant AV, Gao P (2005) A review of the genetic epidemiology of resistance to parasitic
disease and atopic asthma: common variants for common phenotypes? Curr Opin Allergy Clin
Immunol 5(5):379–385
Blackley CH (1873) Experimental researches on the causes and nature of Catarrhus Aestivus (Hay-
fever and Hay-asthma). Baillière Tindall and Cox, London
Blount D, Hooi D, Feary J, Venn A, Telford G, Brown A, Britton J, Pritchard D (2009)
Immunological profiles of subjects recruited for a randomized, placebo controlled clinical trial
of hookworm infection. Am J Trop Med Hyg 81:911–916
Boden EK, Snapper SB (2008) Regulatory T cells in inflammatory bowel disease. Curr Opin
Gastroenterol 24(6):733–741
Bowlby J (1971) Attachment and loss, Volume 1: Attachment. Penguin, Harmondsworth,
Middlesex, England, p 478 (first published by Hogarth press in 1969)
Brehm JM, Schuemann B, Fuhlbrigge AL, Hollis BW, Strunk RC, Zeiger RS, Weiss ST, Litonjua
AA (2010) Serum vitamin D levels and severe asthma exacerbations in the Childhood Asthma
Management Program study. J Allergy Clin Immunol 126(1):52–58, e55
Broadhurst MJ, Leung JM, Kashyap V, McCune JM, Mahadevan U, McKerrow JH, Loke P (2010)
IL-22+ CD4+ T cells are associated with therapeutic trichuris trichiura infection in an ulcer-
ative colitis patient. Sci Transl Med 2(60):60ra88
Bueno LL, Morais CG, Araujo FF, Gomes JA, Correa-Oliveira R, Soares IS, Lacerda MV, Fujiwara
RT, Braga EM (2010) Plasmodium vivax: induction of CD4 + CD25 + FoxP3+ regulatory T
356 G.A.W. Rook
cells during infection are directly associated with level of circulating parasites. PLoS One
5(3):e9623
Buning J, Homann N, von Smolinski D, Borcherding F, Noack F, Stolte M, Kohl M, Lehnert H,
Ludwig D (2008) Helminths as governors of inflammatory bowel disease. Gut
57(8):1182–1183
Cash HL, Whitham CV, Behrendt CL, Hooper LV (2006) Symbiotic bacteria direct expression of
an intestinal bactericidal lectin. Science 313(5790):1126–1130
Collins SM, Bercik P (2009) The relationship between intestinal microbiota and the central ner-
vous system in normal gastrointestinal function and disease. Gastroenterology
136(6):2003–2014
Conrad ML, Ferstl R, Teich R, Brand S, Blumer N, Yildirim AO, Patrascan CC, Hanuszkiewicz A,
Akira S, Wagner H et al (2009) Maternal TLR signaling is required for prenatal asthma protec-
tion by the nonpathogenic microbe Acinetobacter lwoffii F78. J Exp Med 206(13):2869–2877
Cooper PJ, Chico ME, Rodrigues LC, Ordonez M, Strachan D, Griffin GE, Nutman TB (2003)
Reduced risk of atopy among school-age children infected with geohelminth parasites in a rural
area of the tropics. J Allergy Clin Immunol 111(5):995–1000
Cooper PJ, Chico ME, Vaca MG, Moncayo AL, Bland JM, Mafla E, Sanchez F, Rodrigues LC,
Strachan DP, Griffin GE (2006) Effect of albendazole treatments on the prevalence of atopy in
children living in communities endemic for geohelminth parasites: a cluster-randomised trial.
Lancet 367(9522):1598–1603
Correale J, Farez M (2007) Association between parasite infection and immune responses in mul-
tiple sclerosis. Ann Neurol 61(2):97–108
Correale J, Farez MF (2011) The impact of parasite infections on the course of multiple sclerosis.
J Neuroimmunol 233(1–2):6–11
Correale J, Farez M, Razzitte G (2008) Helminth infections associated with multiple sclerosis
induce regulatory B cells. Ann Neurol 64(2):187–199
De Filippo C, Cavalieri D, Di Paola M, Ramazzotti M, Poullet JB, Massart S, Collini S, Pieraccini
G, Lionetti P (2010) Impact of diet in shaping gut microbiota revealed by a comparative study
in children from Europe and rural Africa. Proc Natl Acad Sci USA 107(33):14691–14696
de Mazancourt C, Loreau M, Dieckmann U (2005) Understanding mutualism when there is adap-
tation to the partner. J Ecol 93:305–314
Delmont TO, Robe P, Cecillon S, Clark IM, Constancias F, Simonet P, Hirsch PR, Vogel TM
(2011) Accessing the soil metagenome for studies of microbial diversity. Appl Environ
Microbiol 77(4):1315–1324
Despres L, Imbert-Establet D, Combes C, Bonhomme F (1992) Molecular evidence linking homi-
nid evolution to recent radiation of schistosomes (Platyhelminthes: Trematoda). Mol Phylogenet
Evol 1(4):295–304
Diana J, Vedran Brezar V, Beaudoin L, Dalod M, Mellor A, Tafuri A, von Herrath M, Boitard C,
Mallone R, Lehuen A (2011) Viral infection prevents diabetes by inducing regulatory T cells
through NKT cell–plasmacytoid dendritic cell interplay. J Exp Med 208:729–745
Donnet-Hughes A, Perez PF, Dore J, Leclerc M, Levenez F, Benyacoub J, Serrant P, Segura-
Roggero I, Schiffrin EJ (2010) Potential role of the intestinal microbiota of the mother in neo-
natal immune education. Proc Nutr Soc 69(3):407–415
dos Santos VM, Muller M, de Vos WM (2010) Systems biology of the gut: the interplay of food,
microbiota and host at the mucosal interface. Curr Opin Biotechnol 21:1–12
Eder W, Ege MJ, von Mutius E (2006) The asthma epidemic. N Engl J Med 355(21):2226–2235
Ege MJ, Herzum I, Buchele G, Krauss-Etschmann S, Lauener RP, Roponen M, Hyvarinen A,
Vuitton DA, Riedler J, Brunekreef B et al (2008) Prenatal exposure to a farm environment
modifies atopic sensitization at birth. J Allergy Clin Immunol 122(2):407–412, 412 e401–404
Ege MJ, Mayer M, Normand AC, Genuneit J, Cookson WO, Braun-Fahrlander C, Heederik D,
Piarroux R, von Mutius E (2011) Exposure to environmental microorganisms and childhood
asthma. N Engl J Med 364(8):701–709
Ehlers S, Rook GAW (2011) The role of bacterial and parasitic infections in chronic inflammatory
disorders and autoimmunity. In: Kaufmann SHE, Rouse BT, Sacks DL (eds) Immunology of
infectious diseases. American Society for Microbiology, Washington, DC, pp 521–535
Microbial Exposures and Other Early Childhood Influences… 357
Elliott DE, Crawford C, Lie J, Blum A, Metwali A, Qadir K, Weinstock JV (1999) Exposure to
helminthic parasites protects mice from intestinal inflammation. Gastroenterology 16:A706
Elliott DE, Li J, Blum A, Metwali A, Qadir K, Urban JF Jr, Weinstock JV (2003) Exposure to schisto-
some eggs protects mice from TNBS-induced colitis. Am J Physiol 284(3):G385–G391
Elliott DE, Summers RW, Weinstock JV (2005) Helminths and the modulation of mucosal inflam-
mation. Curr Opin Gastroenterol 21(1):51–58
Erdman SE, Rao VP, Olipitz W, Taylor CL, Jackson EA, Levkovich T, Lee CW, Horwitz BH, Fox
JG, Ge Z et al (2010) Unifying roles for regulatory T cells and inflammation in cancer. Int J
Cancer 126(7):1651–1665
Farias AS, Talaisys RL, Blanco YC, Lopes SC, Longhini AL, Pradella F, Santos LM, Costa FT
(2011) Regulatory T cell induction during Plasmodium chabaudi infection modifies the clinical
course of experimental autoimmune encephalomyelitis. PLoS One 6(3):e17849
Feary JR, Venn AJ, Mortimer K, Brown AP, Hooi D, Falcone FH, Pritchard DI, Britton JR (2010)
Experimental hookworm infection: a randomized placebo-controlled trial in asthma. Clin Exp
Allergy 40(2):299–306
Filippi CM, von Herrath MG (2008) Viral trigger for type 1 diabetes: pros and cons. Diabetes
57(11):2863–2871
Filippi CM, Estes EA, Oldham JE, von Herrath MG (2009) Immunoregulatory mechanisms trig-
gered by viral infections protect from type 1 diabetes in mice. J Clin Invest
119(6):1515–1523
Flanagan KL, Halliday A, Burl S, Landgraf K, Jagne YJ, Noho-Konteh F, Townend J, Miles DJ,
van der Sande M, Whittle H et al (2010) The effect of placental malaria infection on cord blood
and maternal immunoregulatory responses at birth. Eur J Immunol 40(4):1062–1072
Fleming JO, Cook TD (2006) Multiple sclerosis and the hygiene hypothesis. Neurology
67(11):2085–2086
Fleming J, Isaak A, Lee J, Luzzio C, Carrithers M, Cook T, Field A, Boland J, Fabry Z (2011)
Probiotic helminth administration in relapsing-remitting multiple sclerosis: a phase 1 study.
Mult Scler 17(6):743–754
Flohr C, Tuyen LN, Quinnell RJ, Lewis S, Minh TT, Campbell J, Simmons C, Telford G, Brown
A, Hien TT et al (2010) Reduced helminth burden increases allergen skin sensitization but not
clinical allergy: a randomized, double-blind, placebo-controlled trial in Vietnam. Clin Exp
Allergy 40(1):131–142
Fredericks CA, Drabant EM, Edge MD, Tillie JM, Hallmayer J, Ramel W, Kuo JR, Mackey S,
Gross JJ, Dhabhar FS (2010) Healthy young women with serotonin transporter SS polymor-
phism show a pro-inflammatory bias under resting and stress conditions. Brain Behav Immun
24:350–357
Friberg IM, Bradley JE, Jackson JA (2010) Macroparasites, innate immunity and immunoregula-
tion: developing natural models. Trends Parasitol 26:540–549
Fujimura KE, Johnson CC, Ownby DR, Cox MJ, Brodie EL, Havstad SL, Zoratti EM, Woodcroft
KJ, Bobbitt KR, Wegienka G et al (2010) Man’s best friend? The effect of pet ownership on
house dust microbial communities. J Allergy Clin Immunol 126(2):410–412, 412 e411–413
Fumagalli M, Pozzoli U, Cagliani R, Comi GP, Riva S, Clerici M, Bresolin N, Sironi M (2009)
Parasites represent a major selective force for interleukin genes and shape the genetic predis-
position to autoimmune conditions. J Exp Med 206(6):1395–1408
Gaboriau-Routhiau V, Rakotobe S, Lecuyer E, Mulder I, Lan A, Bridonneau C, Rochet V, Pisi A,
De Paepe M, Brandi G et al (2009) The key role of segmented filamentous bacteria in the coor-
dinated maturation of gut helper T cell responses. Immunity 31(4):677–689
Gale EA (2002) A missing link in the hygiene hypothesis? Diabetologia 45(4):588–594
Geuking MB, Cahenzli J, Lawson MA, Ng DC, Slack E, Hapfelmeier S, McCoy KD, Macpherson
AJ (2011) Intestinal bacterial colonization induces mutualistic regulatory T cell responses.
Immunity 34(5):794–806
Grainger JR, Smith KA, Hewitson JP, McSorley HJ, Harcus Y, Filbey KJ, Finney CAM, Greenwood
EJD, Knox DP, Wilson MS et al (2010) Helminth secretions induce de novo T cell Foxp3
expression and regulatory function through the TGF-beta pathway. J Exp Med
207(11):2331–2341
358 G.A.W. Rook
Hagel I, Lynch NR, Perez M, Di Prisco MC, Lopez R, Rojas E (1993) Modulation of the allergic
reactivity of slum children by helminthic infection. Parasite Immunol 15(6):311–315
Hang L, Setiawan T, Blum AM, Urban J, Stoyanoff K, Arihiro S, Reinecker HC, Weinstock JV
(2010) Heligmosomoides polygyrus infection can inhibit colitis through direct interaction with
innate immunity. J Immunol 185(6):3184–3189
Harrison LC, Honeyman MC, Morahan G, Wentworth JM, Elkassaby S, Colman PG, Fourlanos S
(2008) Type 1 diabetes: lessons for other autoimmune diseases? J Autoimmun 31(3):306–310
Heaney RP, Holick MF (2011) Why the IOM recommendations for vitamin D are deficient. J Bone
Miner Res 26(3):455–457
Herr C, Greulich T, Koczulla RA, Meyer S, Zakharkina T, Branscheidt M, Eschmann R, Bals R
(2011) The role of vitamin D in pulmonary disease: COPD, asthma, infection, and cancer.
Respir Res 12:31
Hilty M, Burke C, Pedro H, Cardenas P, Bush A, Bossley C, Davies J, Ervine A, Poulter L, Pachter
L et al (2010) Disordered microbial communities in asthmatic airways. PLoS One 5(1):e8578
Hoberg EP (2006) Phylogeny of Taenia: species definitions and origins of human parasites.
Parasitol Int 55(Suppl):S23–S30
Holtzman MJ, Byers DE, Benoit LA, Battaile JT, You Y, Agapov E, Park C, Grayson MH, Kim EY,
Patel AC (2009) Immune pathways for translating viral infection into chronic airway disease.
Adv Immunol 102:245–276
Honeyman MC, Harrison LC (2009) Alternative and additional mechanisms to the hygiene hypoth-
esis. In: Rook GAW (ed) The hygiene hypothesis and Darwinian medicine. Birkhäuser, Basel,
pp 279–298
Huang SL, Tsai PF, Yeh YF (2002) Negative association of Enterobius infestation with asthma and
rhinitis in primary school children in Taipei. Clin Exp Allergy 32(7):1029–1032
Huang YJ, Nelson CE, Brodie EL, Desantis TZ, Baek MS, Liu J, Woyke T, Allgaier M, Bristow J,
Wiener-Kronish JP et al (2011) Airway microbiota and bronchial hyperresponsiveness in
patients with suboptimally controlled asthma. J Allergy Clin Immunol 127(2):372–381,
e371–373
Iliev ID, Mileti E, Matteoli G, Chieppa M, Rescigno M (2009) Intestinal epithelial cells promote
colitis-protective regulatory T-cell differentiation through dendritic cell conditioning. Mucosal
Immunol 2(4):340–350
Ivanov II, Frutos Rde L, Manel N, Yoshinaga K, Rifkin DB, Sartor RB, Finlay BB, Littman DR
(2008) Specific microbiota direct the differentiation of IL-17-producing T-helper cells in the
mucosa of the small intestine. Cell Host Microbe 4(4):337–349
Ivanov II, Atarashi K, Manel N, Brodie EL, Shima T, Karaoz U, Wei D, Goldfarb KC, Santee CA,
Lynch SV et al (2009) Induction of intestinal Th17 cells by segmented filamentous bacteria.
Cell 139(3):485–498
Jackson DJ (2010) The role of rhinovirus infections in the development of early childhood asthma.
Curr Opin Allergy Clin Immunol 10(2):133–138
Jeon KW (1972) Development of cellular dependence on infective organisms: micrurgical studies
in amoebas. Science 176:1122–1123
Kalliomaki M, Collado MC, Salminen S, Isolauri E (2008) Early differences in fecal microbiota
composition in children may predict overweight. Am J Clin Nutr 87(3):534–538
Kirsner JB (1995) The historical basis of idiopathic inflammatory bowel diseases. Inflamm Bowel
Dis 1:2–26
Klement E, Lysy J, Hoshen M, Avitan M, Goldin E, Israeli E (2008) Childhood hygiene is associ-
ated with the risk for inflammatory bowel disease: a population-based study. Am J Gastroenterol
103(7):1775–1782
Kohashi O, Kohashi Y, Takahashi T, Ozawa A, Shigematsu N (1985) Reverse effect of gram-
positive bacteria vs. gram-negative bacteria on adjuvant-induced arthritis in germfree rats.
Microbiol Immunol 29:487–497
Koloski NA, Bret L, Radford-Smith G (2008) Hygiene hypothesis in inflammatory bowel disease:
a critical review of the literature. World J Gastroenterol 14(2):165–173
Microbial Exposures and Other Early Childhood Influences… 359
Peters M, Kauth M, Scherner O, Gehlhar K, Steffen I, Wentker P, von Mutius E, Holst O, Bufe A
(2010) Arabinogalactan isolated from cowshed dust extract protects mice from allergic airway
inflammation and sensitization. J Allergy Clin Immunol 126(3):648–656, e641–644
Pfefferle PI, Buchele G, Blumer N, Roponen M, Ege MJ, Krauss-Etschmann S, Genuneit J,
Hyvarinen A, Hirvonen MR, Lauener R et al (2010) Cord blood cytokines are modulated by
maternal farming activities and consumption of farm dairy products during pregnancy: the
PASTURE Study. J Allergy Clin Immunol 125(1):108–115 e101–103
Raison CL, Lowry CA, Rook GAW (2010) Inflammation, sanitation and consternation: loss of
contact with co-evolved, tolerogenic micro-organisms and the pathophysiology and treatment
of major depression. Arch Gen Psychiatry 67(12):1211–1224
Resta SC (2009) Effects of probiotics and commensals on intestinal epithelial physiology: implica-
tions for nutrient handling. J Physiol 587(Pt 17):4169–4174
Reyes A, Haynes M, Hanson N, Angly FE, Heath AC, Rohwer F, Gordon JI (2010) Viruses in the
faecal microbiota of monozygotic twins and their mothers. Nature 466(7304):334–338
Ricklin-Gutzwiller ME, Reist M, Peel JE, Seewald W, Brunet LR, Roosje PJ (2007) Intradermal
injection of heat-killed Mycobacterium vaccae in dogs with atopic dermatitis: a multicentre
pilot study. Vet Dermatol 18(2):87–93
Riedl M, Diaz-Sanchez D (2005) Biology of diesel exhaust effects on respiratory function.
J Allergy Clin Immunol 115(2):221–228, quiz 229
Riedler J, Braun-Fahrlander C, Eder W, Schreuer M, Waser M, Maisch S, Carr D, Schierl R,
Nowak D, von Mutius E (2001) Exposure to farming in early life and development of asthma
and allergy: a cross-sectional survey. Lancet 358(9288):1129–1133
Robinson CL, Baumann LM, Romero K, Combe JM, Gomez A, Gilman RH, Cabrera L, Gonzalvez
G, Hansel NN, Wise RA et al (2011) Effect of urbanisation on asthma, allergy and airways
inflammation in a developing country setting. Thorax 66(12):1051–1057
Rook GAW (2009) The broader implications of the hygiene hypothesis. Immunology 126:3–11
Rook GAW (2010) 99th Dahlem conference on infection, inflammation and chronic inflammatory
disorders: Darwinian medicine and the ‘hygiene’ or ‘old friends’ hypothesis. Clin Exp Immunol
160(1):70–79
Rook GAW, Dalgleish A (2011) Infection, immunoregulation and cancer. Immunol Rev
240:141–159
Rook GAW, Stanford JL (1998) Give us this day our daily germs. Immunol Today 19:113–116
Rosati G (2001) The prevalence of multiple sclerosis in the world: an update. Neurol Sci
22(2):117–139
Ross AC, Manson JE, Abrams SA, Aloia JF, Brannon PM, Clinton SK, Durazo-Arvizu RA,
Gallagher JC, Gallo RL, Jones G et al (2011) The 2011 report on dietary reference intakes for
calcium and vitamin D from the Institute of Medicine: what clinicians need to know. J Clin
Endocrinol Metab 96(1):53–58
Round JL, Mazmanian SK (2009) The gut microbiota shapes intestinal immune responses during
health and disease. Nat Rev Immunol 9(5):313–323
Round JL, Lee SM, Li J, Tran G, Jabri B, Chatila TA, Mazmanian SK (2011) The Toll-like receptor
2 pathway establishes colonization by a commensal of the human microbiota. Science
332(6032):974–977
Sabetta JR, DePetrillo P, Cipriani RJ, Smardin J, Burns LA, Landry ML (2010) Serum
25-hydroxyvitamin d and the incidence of acute viral respiratory tract infections in healthy
adults. PLoS One 5(6):e11088
Schaub B, Liu J, Hoppler S, Schleich I, Huehn J, Olek S, Wieczorek G, Illi S, von Mutius E (2009)
Maternal farm exposure modulates neonatal immune mechanisms through regulatory T cells.
J Allergy Clin Immunol 123(4):774–782.e775
Schnoeller C, Rausch S, Pillai S, Avagyan A, Wittig BM, Loddenkemper C, Hamann A, Hamelmann
E, Lucius R, Hartmann S (2008) A helminth immunomodulator reduces allergic and inflamma-
tory responses by induction of IL-10-producing macrophages. J Immunol 180(6):4265–4272
Scrivener S, Yemaneberhan H, Zebenigus M, Tilahun D, Girma S, Ali S, McElroy P, Custovic A,
Woodcock A, Pritchard D et al (2001) Independent effects of intestinal parasite infection and
domestic allergen exposure on risk of wheeze in Ethiopia: a nested case–control study. Lancet
358(9292):1493–1499
Microbial Exposures and Other Early Childhood Influences… 361
van Die I, Cummings RD (2010) Glycan gimmickry by parasitic helminths: a strategy for modulat-
ing the host immune response? Glycobiology 20(1):2–12
Veldhoen M, Hirota K, Westendorf AM, Buer J, Dumoutier L, Renauld JC, Stockinger B (2008)
The aryl hydrocarbon receptor links TH17-cell-mediated autoimmunity to environmental tox-
ins. Nature 453(7191):106–109
von Ehrenstein OS, von Mutius E, Illi S, Baumann L, Bohm O, von Kries R (2000) Reduced risk
of hay fever and asthma among children of farmers. Clin Exp Allergy 30:187–193
von Mutius E (2010) 99th Dahlem conference on infection, inflammation and chronic inflamma-
tory disorders: farm lifestyles and the hygiene hypothesis. Clin Exp Immunol 160(1):130–135
von Mutius E, Radon K (2008) Living on a farm: impact on asthma induction and clinical course.
Immunol Allergy Clin North Am 28(3):631–647
von Mutius E, Vercelli D (2010) Farm living: effects on childhood asthma and allergy. Nat Rev
Immunol 10(12):861–868
Walk ST, Blum AM, Ewing SA, Weinstock JV, Young VB (2010) Alteration of the murine gut
microbiota during infection with the parasitic helminth Heligmosomoides polygyrus. Inflamm
Bowel Dis 16(11):1841–1849
Wegienka G, Havstad S, Zoratti EM, Woodcroft KJ, Bobbitt KR, Ownby DR, Johnson CC (2009)
Regulatory T cells in prenatal blood samples: variability with pet exposure and sensitization.
J Reprod Immunol 81(1):74–81
Weinstock JV, Elliott DE (2009) Helminths and the IBD hygiene hypothesis. Inflamm Bowel Dis
15(1):128–133
Whitlock DR, Feelisch M (2009) Soil bacteria, nitrite, and the skin. In: Rook GAW (ed) The
hygiene hypothesis and Darwinian medicine. Basel, Birkhäuser, pp 103–116
Wildin RS, Smyk-Pearson S, Filipovich AH (2002) Clinical and molecular features of the immu-
nodysregulation, polyendocrinopathy, enteropathy, X linked (IPEX) syndrome. J Med Genet
39(8):537–545
Wu P, Dupont WD, Griffin MR, Carroll KN, Mitchel EF, Gebretsadik T, Hartert TV (2008)
Evidence of a causal role of winter virus infection during infancy in early childhood asthma.
Am J Respir Crit Care Med 178(11):1123–1129
Wu HJ, Ivanov II, Darce J, Hattori K, Shima T, Umesaki Y, Littman DR, Benoist C, Mathis D
(2010) Gut-residing segmented filamentous bacteria drive autoimmune arthritis via T helper 17
cells. Immunity 32(6):815–827
Xystrakis E, Kusumakar S, Boswell S, Peek E, Urry Z, Richards DF, Adikibi T, Pridgeon C,
Dallman M, Loke TK et al (2006) Reversing the defective induction of IL-10-secreting regula-
tory T cells in glucocorticoid-resistant asthma patients. J Clin Invest 116(1):146–155
Yan F, Polk DB (2010) Probiotics: progress toward novel therapies for intestinal diseases. Curr
Opin Gastroenterol 26(2):95–101
Yazdanbakhsh M, Wahyuni S (2005) The role of helminth infections in protection from atopic
disorders. Curr Opin Allergy Clin Immunol 5(5):386–391
Yoo J, Tcheurekdjian H, Lynch SV, Cabana M, Boushey HA (2007) Microbial manipulation of
immune function for asthma prevention: inferences from clinical trials. Proc Am Thorac Soc
4(3):277–282
Zaccone P, Fehervari Z, Jones FM, Sidobre S, Kronenberg M, Dunne DW, Cooke A (2003)
Schistosoma mansoni antigens modulate the activity of the innate immune response and pre-
vent onset of type 1 diabetes. Eur J Immunol 33(5):1439–1449
Zivkovic AM, German JB, Lebrilla CB, Mills DA (2010) Microbes and health Sackler colloquium:
human milk glycobiome and its impact on the infant gastrointestinal microbiota. Proc Natl
Acad Sci USA 108:4653–4658. doi:10.1073/pnas.1000083107, www.pnas.org/cgi/doi/
Zuany-Amorim C, Sawicka E, Manlius C, Le Moine A, Brunet LR, Kemeny DM, Bowen G, Rook
G, Walker C (2002) Suppression of airway eosinophilia by killed Mycobacterium vaccae-
induced allergen-specific regulatory T-cells. Nat Med 8:625–629
Make New Friends and Keep the Old?
Parasite Coinfection and Comorbidity
in Homo sapiens
Introduction
Across species, the fitness costs of parasitic infection have been a major force
shaping host adaptations to avoid infection (Hart 2009; Schmid-Hempel 2003;
Sheldon and Verhulst 1996), diminish the cost of infection (Minchella 1985; Råberg
et al. 2009), and even advertise resistance to infection to possible mates (Hamilton
and Zuk 1982; Moller 1990). These adaptations, in turn, have shaped selection on
parasite transmission and virulence, leading to coevolved host–parasite systems.
Host–parasite interactions are further shaped by local environments and proximate
host factors that influence transmission risk and infectious outcomes, including age,
sex, and nutritional and immune status (Anderson and May 1981; Anderson 1991;
Quinnell et al. 1995; Schad and Anderson 1985; Woolhouse 1992).
Host–parasite interactions are often observed and modeled as hosts interacting
with a single parasite species. Yet as is increasingly observed in animal populations,
including humans, coinfection with two or more species (alternately termed
“multiple-species” or “polyparasitic” infection) may be the rule in nature (Howard
et al. 2001; Booth et al. 1998; Bordes and Morand 2009; Pullan and Brooker 2008).
Coinfecting species may include any number of “typical parasites” (e.g., helminths,
flukes, tapeworms) and/or pathogens (e.g., bacteria, viruses, protozoa), each with
different associated exposure risks, infectious sites, reproductive strategies, virulence,
J.F. Brinkworth and K. Pechenkina (eds.), Primates, Pathogens, and Evolution, 363
Developments in Primatology: Progress and Prospects,
DOI 10.1007/978-1-4614-7181-3_12, © Springer Science+Business Media New York 2013
364 M. Martin et al.
and associated immune responses (Alizon 2008; May et al. 2009; Rigaud et al. 2010;
de Roode et al. 2005; Van Baalen and Sabelis 1995). Of particular interest is the role
of immune responses in mediating coinfection risk, as an immune response gener-
ated by one species may either increase or decrease a host’s susceptibility to infec-
tion with another (Christensen et. al. 1987; Cox 2001; Supali et al. 2010). At the
same time, infecting species may competitively inhibit establishment or replication
by other species (Lim and Heyneman 1972; Fredensborg and Poulin 2005). As such,
coinfection may increase, decrease, or have no effect on host fitness, depending on
the individual species involved (Fellous and Koella 2009).
For humans—whose habitats span from hot, humid jungles to dry deserts, frozen
tundras, and sterile office buildings—infection risk may be especially varied.
Differences in parasitic and pathogenic exposure appear to have been a major force
shaping genetic variation across human populations (Fumagalli et al. 2011). Given
its ubiquity in nature and among nonindustrialized populations (Howard et al.
2001), multiple-species infection was likely equally common and varied among
human and hominin ancestral populations. Along with more transient infections,
hominin ancestors would have harbored multiple symbiotic organisms common to
other mammals: commensal bacteria, pseudo-commensals, ectoparasites, and hel-
minths (Armelagos and Harper 2005; Rook 2008). As proposed by the “hygiene
hypothesis,” continuous exposure to these organisms during mammalian evolution
may have favored the evolution of immunoregulatory systems that required their
antigenic input to develop appropriately (Jackson et al. 2008). In modern industrial-
ized, hygienic environments, infection with many of these “old friends” (particu-
larly helminths) is exceedingly rare, and consequently, disorders of immunoregulation
(i.e., allergy, asthma, chronic inflammatory conditions) have become increasingly
common (Rook 2008).
However, these old friends are also not without costs. First, helminth-induced
immunoregulation, which downregulates proinflammatory responses, may decrease
resistance to other parasites and more virulent pathogens, resulting in increased
infection intensity or exacerbated immunopathology (Graham et al. 2005; Pullan
and Brooker 2008). Second, exposure to helminth coinfection may increase invest-
ment in immune function (Bordes and Morand 2009), which may divert energy
away from other fitness-enhancing allocations, such as growth and reproduction
(Sheldon and Verhulst 1996; Adamo 2001; Uller et al. 2006; Blackwell et al. 2010;
Muehlenbein et al. 2010). Given the risk of helminth coinfection in ancestral envi-
ronments, potentially divergent immune responses, and the costs of increased
investment in immune function, several questions arise. How does helminth coin-
fection risk and associated morbidity vary across environments and with different
interacting species? How costly are multiple-species infections involving helminths
in humans? What multiple-species infections would have been typical for ancestral
populations? Finally, how have recent environmental changes altered helminth
coinfection risk in modern populations, and what are the consequences for human
health?
In this chapter, we review known aspects of immune responses to helminths and
other parasitic and pathogenic threats and consider how coinfections involving
Make New Friends and Keep the Old? Parasite Coinfection and Comorbidity… 365
helminths may affect human immune function and health, both past and present.
We first review host, environmental, and parasitic characteristics that influence the
likelihood of helminth coinfection. We then examine helminth-protozoa coinfection
and helminth-associated morbidity among the Tsimane of lowland Bolivia. The
Tsimane are a subsistence-scale, forager-horticulturalist population afflicted with a
high burden of both parasites and pathogens. We examine coinfection involving
helminths and protozoa and interactions between helminths and the risk of other
infections and inflammatory conditions. Although many aspects of the Tsimane
environment are unlikely to match those of ancestral hunter-gatherer populations,
the Tsimane disease ecology is likely more representative of ancestral conditions
than that of a contemporary industrialized or transitioning population. Our intent is
to provide an example of the complex interactions between infecting species that
would likely have been present through much of human history.
Mammalian, and indeed all vertebrate, immune systems have evolved to counter
invasions from diverse parasites and pathogens. Responses may be optimized to
clear infection and/or minimize damage from infection, depending on the infec-
tious agent’s own evolved strategy to evade or exploit host immune pathways
(Allen and Maizels 2011). The immune system makes strong phenotypic com-
mitments in response to infection, which may lead to biased immune responses
that in turn influence coinfection outcomes (Bradley and Jackson 2008). Each
of the several different types of immune defense, therefore, has its own costs
and benefits, and organisms must allocate resources appropriately to invest in
defenses that are useful for local pathogens (Long and Nanthakumar 2004;
McDade 2005).
The vertebrate immune system is generally divided into two levels of response:
innate and adaptive immunity. Innate immunity is the first line of defense, found in
all plants and animals; it recognizes and responds to generic signals of invasion
(e.g., unchanging structures on bacteria cell walls) with nonspecific responses
including inflammation, induction of acute-phase proteins (e.g., C-reactive protein),
activation of the complement system (a cascade of proteins that assist antibodies
and phagocytic cells in pathogen clearance), and activation and recruitment of white
blood cells, or leukocytes, to target and clear infected host cells and extracellular
viruses, bacteria, and protozoa.
Adaptive immunity is found only in vertebrates and, compared to innate immu-
nity, is highly specific, highly flexible in its recognition capabilities, and capable of
antigen-specific memory. Importantly, helminth diversity—a proxy for coinfection
risk—may have selected for increased investment in adaptive immunity during
mammalian evolution (Bordes and Morand 2009). Adaptive immunity is activated
when particles from invading organisms (antigens) are engulfed and processed by
phagocytic cells of the innate immune system. These cells present the antigens to
effector cells of the adaptive system (T and B cells), which are then activated and
clonally expanded. Activated B cells release antibodies, which bind to antigen and
Make New Friends and Keep the Old? Parasite Coinfection and Comorbidity… 367
facilitate pathogen clearance. Activated B cells may also develop into memory B
cells, which are the basis of acquired immunity.
Activated T cells undergo further differentiation into various subgroups of T
cells that direct different immune responses. These subgroups include cytotoxic T
cells, helper T cells (TH), and regulatory T cells (Treg cells). Helper T cells are further
differentiated into TH1, TH2, and TH17 cells based on the cytokines they are associ-
ated with. In brief, TH1 cells and associated cytokines such as IFN-γ stimulate
inflammatory and cell-killing activity important in clearance of pathogens (e.g.,
protozoa, trypanosomes, bacteria, viruses). TH2-associated cytokines (primarily
IL-4, IL-5, IL-13) stimulate antibodies including immunoglobulin E (IgE), IgG1,
and (in humans) IgG4 production, as well as basophils, eosinophils, and mast cells,
which are important in mediating clearance and tissue repair associated with typical
parasites (e.g., helminths, flatworms) (Allen and Maizels 2011). TH1 and TH2
responses are directly antagonistic, with IL-4 inhibiting IFN-γ production and vice
versa (Maizels and Yazdanbakhsh 2003).
Clearance of coinfecting species that provoke similar immune responses may be
enhanced through cross-immunity (Lello and Hussell 2008; Supali et al. 2010),
which can also diminish the likelihood of future coinfection with other commonly
associated species (Karvonen et al. 2009). Conversely, immune responses directed
against one species may suppress responses against other species if the coinfecting
species invoke antagonistic immune responses. As is discussed below, increasing
evidence suggests that helminths may bias immune function in a manner that
increases susceptibility to viral and bacterial infections.
Helminths are a large category of parasite known to have significant effects on host
fitness. The term “helminth” refers collectively to wormlike parasites and encom-
passes two phyla of major human parasites: Platyhelminthes (flatworms), which
include tapeworms (e.g., Taenia spp. and Hymenolepis spp.) and flukes (e.g.,
Schistosoma mansoni and Schistosoma japonicum), and Nematoda (roundworms),
which include ascarids (e.g., Ascaris lumbricoides), filarial worms (e.g., Wuchereria
bancrofti), pinworm (e.g., Enterobius vermicularis), whipworm (Trichuris trichi-
ura), threadworm (Strongyloides stercoralis), and hookworm (referring to
Ancylostoma duodenale and Necator americanus, which are often undifferentiated
in microscopic identification). Helminths—which are long lived, grow to sexual
maturity in hosts, but do not replicate in hosts—evoke relatively gentle immune
responses in mammals, quite distinct from the strong inflammatory responses
evoked by transient microbial pathogens that present imminent threats to host fit-
ness (Jackson et al. 2008; Allen and Maizels 2011).
368 M. Martin et al.
Helminths shift T cell populations towards a TH2 immune response, with corre-
sponding decreases in TH1 and proinflammatory responses (Cooper et al. 2000;
Fallon and Mangan 2007; Fox et al. 2000; Hewitson et al. 2009; Maizels and
Yazdanbakhsh 2003; Yazdanbakhsh et al. 2002). TH2 cells activated by helminths in
mucosal tissues induce production of IL-13 and IL-4 cytokines that drive mucosal
and muscular responses to dislodge the parasites. In non-mucosal tissues, TH2-
induced pathways and innate immune cells such as eosinophils, basophils, and mast
cells help drive parasite killing. IgE secreted from B cells is important in protecting
hosts from extraintestinal and encysted stages of helminths and may facilitate
antibody-induced larval killing following concomitant or secondary infections
(Allen and Maizels 2011).
Many helminths (as well as commensal bacteria) also induce Treg activity in order
to enhance their own survival in the host (Maizels et al. 2009). Treg cells release
cytokines that suppress TH1 and TH2 responses in order to minimize immunopathol-
ogy and epithelial damage caused by immune activation (Rook 2008; Round and
Mazmanian 2009). Treg activity may also promote production of IgG4 over IgE and
reduce expulsion of worms from the host (Mingomataj et al. 2006). Enhanced and
spontaneous production of Treg cytokines has also been observed in children with
chronic A. lumbricoides or T. trichiura infection and A. lumbricoides/T. trichiura
coinfection, suggesting endemic exposure to multiple helminths promotes stronger
immunoregulation (Turner et al. 2008; Figueiredo et al. 2010).
As such, the prototypical TH2/Treg response induced by helminths may be better
characterized as a “tolerance” response that contains the extent of helminth infec-
tion while limiting damage to the host (Jackson et al. 2008). In this case, the inter-
ests of host and parasite may align. Chronic exposure to helminths, which present
low or intermediate threats to host fitness compared to pathogens, would favor a
continual tolerance response over successive, highly inflammatory responses that
would be energetically costly and highly immunopathogenic to hosts. A tolerance
response would also be favored by helminths, as the Treg response enhances their
own long-term survival in the host, while TH2 responses may limit establishment
and competition by secondary invaders, allowing established parasites to monopo-
lize host resources (Jackson et al. 2008).
The tolerance response suggests that mammals share a deep coevolutionary leg-
acy with helminths. The IgE antibody, which is integral to antihelminth responses,
is a derived innovation in the mammalian lineage (Jackson et al. 2008). However,
more recent coevolution is also apparent: in humans, helminths appear to have
played a major role in genetic population divergence since the appearance of ana-
tomically modern humans within the last 200,000 years (Fumagalli et al. 2010).
More recent evidence of helminth infection in human history comes from mummi-
fied remains dating to approximately 30,000 BP in the Old World (A. lumbricoides)
and 7,837 BP in the New World (E. vermicularis) (Gonçalves et al. 2003), as well
as historical writings from classical Egyptian and Greek physicians (Cox 2002).
Today, however, helminth exposure and associated immune phenotypes are var-
ied across human populations. Hygiene, medicine, and socioeconomic development
Make New Friends and Keep the Old? Parasite Coinfection and Comorbidity… 369
Fig. 1 Variation in human IgE by population. Values are geometric means for each published
population (Adapted from Blackwell et al. (2010), which contains the complete references for each
population)
diseases such as asthma do show elevated IgE levels (e.g., Bergmann et al. 1995;
Holford-Strevens et al. 1984; Lindberg and Arroyave 1986). Variations on the
“hygiene” and “old friends” hypotheses posit that a major factor in the rise of aller-
gic, autoimmune, and inflammatory disorders in industrialized populations today is
a mismatch between a human immune system that coevolved with “old friends”
(e.g., helminths and commensals) and a modern hygienic environment in which
these “old friends” are largely absent (Rook 2008). Without early and regular anti-
genic input from helminths, immune system development is altered, resulting in
“inappropriate” TH2 immune responses to harmless environmental antigens. When
induced by helminths, those same immune responses may depress allergic reactions
(Maizels 2005; Wilson and Maizels 2004), to the extent that prescribed low-dose
infections may be effective clinical treatments (Blount et al. 2009; Feary et al.
2010). The downregulation of inflammatory pathways induced by low-grade hel-
minth infections may also protect against diabetes (Maizels et al. 2009), obesity
(Wu et al. 2011), and immunopathology associated with opportunistic bacterial
infections (Anthony et al. 2008).
At the same time, chronic helminth infection can be costly to hosts. Presently, the
most common human helminth infections involve intestinal nematodes, especially
the soil-transmitted helminths (STH) A. lumbricoides, T. trichiura, and hookworm.
Adult STH reside and replicate in the host’s gut and pass eggs through host feces.
A. lumbricoides and T. trichiura eggs are ingested by hosts, whereas hookworm
eggs penetrate the skin, generally the bottom of the feet (Hotez et al. 2008; Jackson
et al. 2008). STH are widespread but are most prevalent in tropic and subtropic
regions (Chan et al. 1994; Silva et al. 2003). In many nonindustrialized nations,
STH infections are endemic and—despite increased efforts to improve sanitary con-
ditions, access to health care, and implement large-scale control programs—remain
a significant cause of morbidity, particularly among children (Bethony et al. 2006).
Complications ensuing from STH infections in children and adults (e.g., anemia,
growth faltering, reduced work output, and impaired cognitive ability), compounded
with poor nutrition and poverty, likely contribute to poor economic growth in these
areas (Guyatt 2000).
Coinfections involving multiple STH, or at least one STH and another parasite or
pathogen, are exceedingly common in nonindustrialized populations (Hotez et al.
2008). Multiple STH infections are often associated with increased infection inten-
sity, egg output, and morbidity (Booth et al. 1998; Brooker et al. 2000; Ellis et al.
2007; Pullan and Brooker 2008). It is also increasingly documented that helminth
infection and helminth-typical immune biasing may diminish immune responses to
vaccines, viruses, and bacteria, resulting in increased susceptibility to other infec-
tious diseases (e.g., HIV/AIDS (Bentwich et al. 1995); BCG, typhoid, measles, and
polio vaccines (Labeaud et al. 2009); tuberculosis (Lienhardt et al. 2002)). In sum,
helminths may interact with human immune function and other host factors in a
myriad of complex ways that have varying implications for human health. The risks
and effects of helminth coinfection therefore, while of clear clinical and epidemio-
logical significance, are also of relevance to researchers working across evolution-
ary and ecological fields.
Make New Friends and Keep the Old? Parasite Coinfection and Comorbidity… 371
Human immune pathways may have been selected to counter the disease ecologies
of our predecessors, which included constant exposure to multiple parasites and
pathogens. Unfortunately, much of our understanding of human immune function
and health has derived from populations living under evolutionarily novel condi-
tions in which common parasites and pathogens are largely absent. Wider surveying
of the patterns of helminth coinfection and comorbidity across a range of environ-
ments are needed to better understand the varying consequences of endemic hel-
minth exposure—and lack thereof—in both nonindustrialized and industrialized
populations.
In this section we present and review data on helminth coinfection and comor-
bidity patterns in an Amazonian, small-scale subsistence population, the Tsimane of
lowland Bolivia. We focus specifically on the risk of coinfection involving hel-
minths and Giardia lamblia (aka Giardia intestinalis, Giardia duodenalis), a com-
mon Tsimane intestinal protozoan. We then examine the links between helminth
infection and other medical diagnoses. This research provides an example of the
interactions that may occur when multiple species and conditions afflict a single
population (Table 1).
Since 2002, the Tsimane have been participants in the ongoing Tsimane Health and
Life History Project (THLHP). THLHP researchers have worked extensively in
Tsimane villages, collecting demographic, anthropological, and biomedical data
while also providing primary medical care. The data presented in this chapter was
collected from participants seen by a mobile team of THLHP physicians, who trav-
eled annually through Tsimane villages from 2007 to 2010.
Patients seen by THLHP physicians were given routine physical exams (patient
history, symptom investigation, blood pressure and temperature, height and weight).
Physicians administered vitamins, antibiotics, and antihelmintics as warranted, fol-
lowing on-site analysis of participant blood and fecal samples. Ethnographic and
epidemiological information on the Tsimane, methods for age estimation, subject
sampling, biomarker collection, and physician diagnostics have been described
elsewhere (Gurven et al. 2007, 2008, 2009).
Results of parasitic infection presented and reviewed in this chapter were
obtained through community sampling and patient diagnostics conducted from
2004 to 2010. Fecal samples collected by THLHP researchers were analyzed using
two methods. From 2004 to 2008, fecal samples were analyzed for the presence of
helminth eggs, larvae, and protozoa by direct identification on wet mounts.
Beginning in 2007, fecal samples were also preserved in 10 % formalin solution
following direct identification and later quantitatively analyzed using a modified
Percoll (Amersham Pharmacia) technique (Eberl et al. 2002). Methods of fecal
sample collection and parasite identification using both methods have been described
in greater detail elsewhere (Vasunilashorn et al. 2010; Blackwell et al. 2011). Data
presented here were aggregated from the two methods, with individuals coded as
either infected or not infected if helminths were detected by either method
(n = 3,628).
Half of all multiple-species infections observed involved infection with at least two
helminth species. As has been shown elsewhere (e.g., Howard et al. 2001), we found
that individual helminth infection increases the risk of coinfection with other hel-
minths. The strongest positive association we observed was between A. lumbricoi-
des and T. trichiura, with T. trichiura-infected subjects nearly six times as likely to
be coinfected with A. lumbricoides (Table 3). Each helminth infection was associ-
ated with higher odds of infection with another helminth. A. lumbricoides (OR = 1.40,
p < 0.001), T. trichiura (OR = 1.47, p = 0.05), and S. stercoralis (OR = 3.28, p < 0.001)
were all predictive of hookworm infection, while hookworm was associated with
A. lumbricoides (OR = 2.32, p < 0.001), and A. lumbricoides was associated with
T. trichiura (OR = 2.54, p < 0.001). Howard et al. (2001), in a survey of 60 interna-
tional studies of helminth coinfection, found that in ~70 % of cases, the risk of
coinfection with A. lumbricoides and T. trichiura was significantly higher than
would be expected by independent transmission. In the same study, increased risks
376 M. Martin et al.
Table 3 Odds ratios for infection with one parasite given infection with another
Dependent
Independent G. lamblia Hookworm A. lumbricoides T. trichiura S. stercoralis
G. lamblia 0.54*** 0.64* 0.89ns 0.51t
Hookworm 0.54*** 2.32*** 1.14ns 1.77ns
A. lumbricoides 0.72** 1.40*** 2.54** 0.97ns
T. trichiura 1.13ns 1.47* 2.03t 1.11ns
S. stercoralis 0.83ns 3.28*** 1.08ns 1.10ns
Age (decades) 1.04* 2.22*** 0.87* 1.05ns 1.06ns
Sex (male) 1.03ns 1.10ns 0.44** 0.77ns 0.91ns
Odds ratios were calculated in binomial logistic mixed models with all independent variables in
one model, controlling for age, sex, and repeat observations
Parameter significance: tp ≤ 0.10; *p ≤ 0.05; **p ≤ 0.01; ***p ≤ 0.001; nsnonsignificant
Table 4 Odds of infection based on receipt of antihelmintic or antiprotozoal drugs at the previous
medical visit
Any helminth Hookworm A. lumbricoides G. lamblia
(Intercept) 0.41*** 0.33*** 0.05*** 0.77**
Received antihelmintic 1.15 1.12 0.41*** 1.29*
Received antiprotozoal 1.46** 1.61*** 0.83 0.93
Age (years) 1.02*** 1.02*** 1.00 1.00
Sex (male) 1.06 1.13t 0.72* 1.05
Dist. to San Borja (10 km) 1.14*** 1.09*** 1.19*** 0.92***
Parameter values are odds ratios estimated in separate generalized logistic mixed model for each
parasite or pathogen.
Parameter significance: tp ≤ 0.10; *p ≤ 0.05; **p ≤ 0.01; ***p ≤ 0.001
Previously, we have also shown that early and chronic elevated IgE—characteristic
of endemic helminth exposure—is also associated with growth deficits (Blackwell
et al. 2010). Thus, endemic exposure to multiple helminths may be a significant
cause of morbidity in the Tsimane. Future research with the Tsimane will investi-
gate host factors that increase susceptibility to helminth coinfection and infection
intensity and will evaluate if certain helminth coinfections and higher infection
intensity are associated with increased morbidity.
In contrast to the higher odds of helminth coinfection, we have found that the risk
of helminth-giardia coinfection among the Tsimane is significantly less common
than would be predicted by independent transmission. G. lamblia infection was
associated with significantly lower odds of infection with both hookworm
(OR = 0.54, p < 0.001) and A. lumbricoides (OR = 0.64, p < 0.001), while hookworm
and A. lumbricoides were conversely associated with lower odds of G. lamblia
infection (OR = 0.54, p < 0.001; OR = 0.72, p < 0.001). To put the size of this effect
into perspective, 31 % of those infected with any helminth were also infected with
G. lamblia, compared to 45 % of those without helminth infection. Of those with G.
lamblia infection, 49 % were infected with at least one helminth, compared to 64 %
of those without G. lamblia.
Given the apparent antagonism between helminth and G. lamblia infection, we
examined how treatment for helminths affected later risk of infection with G. lam-
blia, and vice versa (Table 4). Receiving an antihelmintic had no effect on the odds
of being infected with any helminth (OR = 1.15, p = 0.28) or with hookworm
(OR = 1.12, p = 0.34) but did reduce the odds of A. lumbricoides infection (OR = 0.41,
p < 0.001). However, receipt of antihelmintics was also associated with increased
odds for G. lamblia infection (OR = 1.29, p = 0.03). Antiprotozoal agents had no
effect on the odds of G. lamblia infection 1 year later (OR = 0.93, p = 0.55) but were
378 M. Martin et al.
It has been suggested that the immunomodulatory properties of helminths may pro-
tect against allergies, autoimmune, and inflammatory disorders. However, helminth-
induced immune biasing may increase susceptibility to other infectious diseases. To
examine potential interactions between helminths, G. lamblia, and other medical
conditions, we grouped THLHP patient disease diagnoses into broad categories rep-
resenting the most common types of complaint. Excluding diagnoses of helminthia-
sis and giardiasis, these included gastrointestinal problems (43 % of 3,391 patient
examinations), muscle or back pain (34 %), upper respiratory illnesses (28 %), uri-
nary tract infections (13 % cases), fungal infections (8 %), arthritis (6 %), skin
infections (3 % cases), and traumatic burns or injuries (2 %). For analysis, we
divided the sample by age into children ≤ 16 years of age and adults over age 16
since many diagnoses were not equally prevalent in children and adults (Table 5).
Controlling for age, sex, and village location, helminth infections were associated
with greater odds of upper respiratory infection in children (OR = 1.33, p = 0.04).
Table 5 Association between current helminth and giardia infection and likelihood of medical diagnosis during medical visit
Odds ratios
Giardia Dist. to town
Sample Medical diagnosis (cases) Cases (%) Helminth infected infected Age (years) Sex (male) (per 10 km)
Children ≤ 16 years Gastrointestinal problems 454 (42 %) 1.09 0.88 0.97t 1.01 0.91**
n = 894 Fungal infections 70 (6 %) 1.17 0.88 1.00 1.23 1.06
obs = 1,086 Upper respiratory infections 474 (44 %) 1.33* 0.99 0.93*** 0.73* 1.11***
Urinary tract infections 19 (2 %) 0.58 0.75 1.24 0.46 0.62
Skin infections 57 (5 %) 0.84 0.37 0.79t 0.46 1.13
Trauma 17 (2 %) 0.54 0.68 1.13 0.51 0.65
Muscle or back pain 10 (1 %) 0.50 0.28 1.54 1.45 1.33
Adults > 16 Gastrointestinal problems 989 (43 %) 1.31** 1.01 1.00 1.02 0.97
n = 1,439 Fungal infections 186 (8 %) 1.00 0.49* 0.99 1.09 1.02
obs = 2,305 Upper respiratory infections 474 (21 %) 1.09 0.84 0.99*** 0.97 0.99
Urinary tract infections 423 (18 %) 1.03 0.96 1.00 0.35*** 0.88***
Skin infections 35 (2 %) 0.20 0.53 0.99 1.01 1.07
Trauma 55 (2 %) 1.75 1.07 0.95 1.35 0.81
Arthritisa 198 (9 %) 0.68* 1.13 1.06*** 0.48*** 1.13**
Muscle or back pain 1,143 (50 %) 0.72*** 1.32** 1.00 2.01*** 1.07***
Parameter values are odds ratios estimated in separate generalized logistic mixed model for each medical diagnosis
Make New Friends and Keep the Old? Parasite Coinfection and Comorbidity…
There are several limitations of our analysis that limit wider extrapolation of our
results and interpretations. First, infectious status among the Tsimane is diagnosed
on the basis of a single fecal sample only, which may underestimate the true preva-
lence of both single- and multiple-species infections in this population. Second,
THLHP community sampling does not permit us to conclusively discriminate
between acute, chronic, or resolving infections, which may influence our interpreta-
tions of parasite associations in hosts. We also as yet do not have data on the pos-
sible risk of coinfection involving helminths and more virulent diseases known to
afflict the Tsimane, including leishmaniasis, dengue fever, tuberculosis, leptospiro-
sis, and other common viral and bacterial infections. While we have shown that
endemic helminth infection in the Tsimane may be protective against a common
intestinal protozoan, helminth-induced TH2/Treg biasing may increase susceptibility
to coinfection with these more virulent pathogens.
Finally, although we have shown that the prevalence of helminth coinfection
among the Tsimane is lower than that of several African populations, Tsimane coin-
fection rates may have increased in recent decades and may continue to increase in
coming years. The Tsimane have only been permanently settled in their current
territory since the mid-twentieth century (Gurven et al. 2007; Huanca 2006).
Increased sedentism, population growth and density, interactions with domesticates,
fecal contamination of local water sources, and emerging social threats (e.g., pros-
titution and alcohol abuse) have likely increased the rate of exposure to existing and
Make New Friends and Keep the Old? Parasite Coinfection and Comorbidity… 381
Conclusions
At the same time, helminth-giardia coinfection in the Tsimane occurs less fre-
quently than would be predicted by independent transmission. The lower risk of
helminth-giardia coinfection may be mediated by direct competition or controlled
by strong TH2 mechanisms invoked by chronic helminth challenge. Future research
is needed to elucidate the mechanisms of helminth-giardia antagonism, particularly
given the clinical implications of increased infection risk following antihelmintic or
antiprotozoal administration. The results reviewed here also suggest that helminth
infection and elevated IgE in adulthood is associated with lower incidence of
inflammatory-associated morbidity (Blackwell et al. 2010; Vasunilashorn et al.
2010). These findings are all consistent with the proposal that some inflammation-
linked “diseases of modernity,” such as obesity and heart disease, may be due to the
absence of “old friends” with which our immune systems coevolved.
We intend this discussion of the costs, benefits, and altered risks of coinfection
associated with helminths to illustrate the importance of considering multiple-
species infections and the role of infectious communities in affecting the evolution
of immune responses in hosts, the virulence of pathogens, and implications for
health and treatment. In sum, we have presented evidence of a complex trade-off in
the risks and benefits of helminth infection in humans. These trade-offs are likely to
vary by life stage, environment, and host factors influencing coinfection and
morbidity risk, which are constantly changing across the human landscape. These
factors may influence differences in patterns of coinfection prevalence and associ-
ated morbidity observed today and must be considered in models of ancestral
parasite and pathogen coinfection risk. Identifying and understanding both the
ancestral and emerging risks of coinfection pose an important challenge for research-
ers working across varied human populations, which will be best met by an inte-
grated cultural, epidemiological, ecological, and evolutionary approach.
References
Abdul-Wahid A, Faubert G (2008) Characterization of the local immune response to cyst antigens
during the acute and elimination phases of primary murine giardiasis. Int J Parasitol
38:691–703
Adamo S (2001) Changes in lifetime immunocompetence in male and female Gryllus texensis
(formerly G. integer): trade-offs between immunity and reproduction. Anim Behav
62:417–425
Alizon S (2008) Decreased overall virulence in coinfected hosts leads to the persistence of virulent
parasites. Am Nat 172:E67–E79
Allen JE, Maizels RM (2011) Diversity and dialogue in immunity to helminths. Nat Rev Immunol
11:375–388
Anderson RM (1991) Populations and infectious diseases: ecology or epidemiology? J Anim Ecol
60:1–50
Anderson R, May R (1979) Population biology of infectious diseases: Part I. Nature 280:
361–367
Anderson RM, May RM (1981) The population dynamics of micro-parasites and their invertebrate
hosts. Phil Trans R Soc B 291:451–524
Make New Friends and Keep the Old? Parasite Coinfection and Comorbidity… 383
Anthony RM, Rutitzky LI, Urban JFU Jr, Stadecker MJ, Gause WC (2008) Protective immune
mechanisms in helminth infection. Nat Rev Immunol 7:975–987
Armelagos GJ, Harper KN (2005) Genomics at the origins of agriculture: Part II. Evol Anthropol
14:109–121
Bentwich Z, Kalinkovich A, Weisman Z (1995) Immune activation is a dominant factor in the
pathogenesis of African AIDS. Immunol Today 16:187–191
Bergmann RL, Schulz J, Gunther S, Dudenhausen JW, Bergmann KE, Bauer CP, Dorsch W,
Schmidt E, Luck W, Lau S (1995) Determinants of cord-blood IgE concentrations in 6401
German neonates. Allergy 50:65–71
Bethony J, Brooker S, Albonico M, Geiger SM, Loukas A, Diemert D, Hotez PJ (2006)
Soil-transmitted helminth infections: ascariasis, trichuriasis, and hookworm. Lancet
367:1521–1532
Blackwell AD, Snodgrass JJ, Madimenos FC, Sugiyama LS (2010) Life history, immune function,
and intestinal helminths: trade-offs among immunoglobulin E, C-reactive protein, and growth
in an Amazonian population. Am J Hum Biol 22:836–848
Blackwell AD, Gurven MD, Sugiyama LS, Madimenos FC, Liebert MA, Martin MA, Kaplan HS,
Snodgrass JJ (2011) Evidence for a peak shift in a humoral response to helminths: age profiles
of IgE in the Shuar of Ecuador, the Tsimane of Bolivia, and the U.S. NHANES. PLoS Negl
Trop Dis 5:12
Blount D, Hooi D, Feary J, Venn A, Telford G, Brown A, Britton J, Pritchard D (2009) Immunologic
profiles of persons recruited for a randomized, placebo-controlled clinical trial of hookworm
infection. Am J Trop Med Hyg 81:911–916
Booth M, Bundy DA, Albonico M, Chwaya HM, Alawi KS, Savioli L (1998) Associations among
multiple geohelminth species infections in schoolchildren from Pemba Island. Parasitology
116:85–93
Bordes F, Morand S (2009) Coevolution between multiple helminth infestations and basal immune
investment in mammals: cumulative effects of polyparasitism? Parasitol Res 106:33–37
Bradley JE, Jackson JA (2008) Measuring immune system variation to help understand host-
pathogen community dynamics. Parasitology 135:807–823
Brooker S, Miguel EA, Moulin S, Luoba AI, Bundy DA, Kremer M (2000) Epidemiology of single
and multiple species of helminth infections among school children in Busia District, Kenya.
East Afr Med J 77:157–161
Brooker S, Bethony J, Hotez P (2004) Human hookworm infection in the 21st century. Adv
Parasitol 58:197–288
Buckley CE, Larrick JW, Kaplan JE (1985) Population differences in cutaneous methacholine
reactivity and circulating IgE concentrations. J Allergy Clin Immunol 76:847
Chan MS, Medley GF, Jamison D, Bundy DA (1994) The evaluation of potential global morbidity
attributable to intestinal nematode infections. Parasitology 109:373–387
Christensen NO, Nansen P, Fagbemi BO, Monrad J (1987) Heterologous antagonistic and syner-
gistic interactions between helminths and between helminths and protozoans in concurrent
experimental infection of mammalian hosts. Parasitol Res 73:387–410
Chunge RN, Nagelkerke N, Karumba PN, Kaleli N, Wamwea M, Mutiso N, Andala EO, Gachoya
J, Kiarie R, Kinoti SN (1992) Longitudinal study of young children in Kenya: intestinal para-
sitic infection with special reference to Giardia lamblia, its prevalence, incidence and duration,
and its association with diarrhoea and with other parasites. Acta Trop 50:39–49
Cooper PJ, Chico ME, Sandoval C, Espinel I, Guevara A, Kennedy MW, Urban JF Jr, Griffin GE,
Nutman TB (2000) Human infection with Ascaris lumbricoides is associated with a polarized
cytokine response. J Infect Dis 182:1207–1213
Cooper PJ, Alexander N, Moncayo A-L, Benitez SM, Chico ME, Vaca MG, Griffin GE (2008)
Environmental determinants of total IgE among school children living in the rural tropics:
importance of geohelminth infections and effect of anthelmintic treatment. BMC Immunol
9:33
Cox FE (2001) Concomitant infections, parasites and immune responses. Parasitology 122:S23–S38
Cox FEG (2002) History of human parasitology. Clin Microbiol Rev 15:595–612
384 M. Martin et al.
de Roode JC, Helinski MEH, Anwar MA, Read AF (2005) Dynamics of multiple infection and
within-host competition in genetically diverse malaria infections. Am Nat 166:531–542
Eberl M, Hagan P, Ljubojevic S, Thomas AW, Wilson RA (2002) A novel and sensitive method to
monitor helminth infections by faecal sampling. Acta Trop 83:183–187
Ellis MK, Raso G, Li Y-S, Rong Z, Chen H-G, McManus DP (2007) Familial aggregation of
human susceptibility to co- and multiple helminth infections in a population from the Poyang
Lake region, China. Int J Parasitol 37:1153–1161
Esrey SA, Potash JB, Roberts L, Shiff C (1991) Effects of improved water supply and sanitation
on ascariasis, diarrhoea, dracunculiasis, hookworm infection, schistosomiasis, and trachoma.
Bull World Health Organ 69:609–621
Fallon PG, Mangan NE (2007) Suppression of TH2-type allergic reactions by helminth infection.
Nat Rev Immunol 7:220–230
Feary JR, Venn AJ, Mortimer K, Brown AP, Hooi D, Falcone FH, Pritchard DI, Britton JR (2010)
Experimental hookworm infection: a randomized placebo-controlled trial in asthma. Clin Exp
Allergy 40:299–306
Fellous S, Koella JC (2009) Infectious dose affects the outcome of the within-host competition
between parasites. Am Nat 173:E177–E184
Figueiredo CA, Barreto ML, Rodrigues LC, Cooper PJ, Silva NB, Amorim LD, Alcantara-Neves
NM (2010) Chronic intestinal helminth infections are associated with immune hyporespon-
siveness and induction of a regulatory network. Infect Immun 78:3160–3167
Fleming FM, Brooker S, Geiger SM, Caldas IR, Correa-oliveira R, Hotez PJ, Bethony JM (2006)
Synergistic associations between hookworm and other helminth species in a rural community
in Brazil. Trop Med Int Health 11:56–64
Foster Z, Byron E, Reyes-García V, Huanca T, Vadez V, Apaza L, Pérez E, Tanner S, Gutierrez Y,
Sandstrom B, Yakhedts A, Osborn C, Gody RA, Leonard WR (2005) Physical growth and
nutritional status of Tsimane’ Amerindian children of lowland Bolivia. Am J Phys Anthropol
126:343–351
Fox JG, Beck P, Dangler CA, Whary MT, Wang TC, Shi HN, Nagler-Anderson C (2000) Concurrent
enteric helminth infection modulates inflammation and gastric immune responses and reduces
helicobacter-induced gastric atrophy. Nat Med 6:536–542
Fredensborg BL, Poulin R (2005) Larval helminths in intermediate hosts: does competition early
in life determine the fitness of adult parasites? Int J Parasitol 35:1061–1070
Fumagalli M, Pozzoli U, Cagliani R, Comi GP, Bresolin N, Clerici M, Sironi M (2010) The land-
scape of human genes involved in the immune response to parasitic worms. BMC Evol Biol
10:264
Fumagalli M, Sironi M, Pozzoli U, Ferrer-Admettla A, Pattini L, Nielsen R (2011) Signatures of
environmental genetic adaptation pinpoint pathogens as the main selective pressure through
human evolution. PLoS Genet 7:e1002355
Gonçalves MLC, Araújo A, Ferreira LF (2003) Human intestinal parasites in the past: new findings
and a review. Mem Inst Oswaldo Cruz 98(Suppl I):103–118
Graham AL, Lamb TJ, Read AF, Allen JE (2005) Malaria-filaria coinfection in mice makes malar-
ial disease more severe unless filarial infection achieves patency. J Infect Dis 191:410–421
Grant AV, Araujo MI, Ponte EV, Oliveira RR, Cruz AA, Barnes KC, Beaty TH (2008) High herita-
bility but uncertain mode of inheritance for total serum IgE level and Schistosoma mansoni
infection intensity in a schistosomiasis-endemic Brazilian population. J Infect Dis
198:1227–1236
Gurven M, Kaplan H, Supa AZ (2007) Mortality experience of Tsimane Amerindians of Bolivia:
regional variation and temporal trends. Am J Hum Biol 19:376–398
Gurven M, Kaplan H, Winking J, Finch C, Crimmins EM (2008) Aging and inflammation in two
epidemiological worlds. J Gerontol A Biol Sci Med Sci 63:196–199
Gurven M, Kaplan H, Winking J, Rodriguez DE, Vasunilashorn S, Kim JK, Finch C, Crimmins E
(2009) Inflammation and infection do not promote arterial aging and cardiovascular disease
risk factors among lean horticulturalists. PLoS One 4:e6590
Guyatt H (2000) Do intestinal nematodes affect productivity in adulthood? Parasitol Today
16:153–158
Make New Friends and Keep the Old? Parasite Coinfection and Comorbidity… 385
Hagel I, Cabrera M, Sánchez P, Rodríguez P, Lattouf JJ (2006) Role of the low affinity IgE receptor
(CD23) on the IgE response against Ascaris lumbricoides in Warao Amerindian children from
Venezuela. Invest Clin 47:241–251
Hagel I, Cabrera M, Puccio F, Santaella C, Buvat E, Infante B, Zabala M, Cordero R, Di Prisco MC
(2011) Co-infection with Ascaris lumbricoides modulates protective immune responses against
Giardia duodenalis in school Venezuelan rural children. Acta Trop 117:189–195
Hall A, Holland C (2000) Geographical variation in Ascaris lumbricoides fecundity and its impli-
cations for helminth control. Parasitology 16:540–544
Hamilton WD, Zuk M (1982) Heritable true fitness and bright birds: a role for parasites ? Science
218:384–387
Hart BL (2009) Adaptations to parasites: an ethological approach. J Parasitol 78:256–265
Haswell-Elkins MR, Elkins DB, Anderson RM (1987) Evidence for predisposition in humans to
infection with Ascaris, hookworm, Enterobius and Trichuris in a South Indian fishing commu-
nity. Parasitology 95:323–337
Hewitson JP, Grainger JR, Maizels RM (2009) Helminth immunoregulation: the role of parasite
secreted proteins in modulating host immunity. Mol Biochem Parasitol 167:1–11
Holford-Strevens V, Warren P, Wong C, Manfreda J (1984) Serum total immunoglobulin E levels
in Canadian adults. J Allergy Clin Immunol 73:516–522
Holland CV (2009) Predisposition to ascariasis: patterns, mechanisms and implications.
Parasitology 136:1537–1547
Hotez PJ, Brindley PJ, Bethony JM, King CH, Pearce EJ, Jacobson J (2008) Helminth infections:
the great neglected tropical diseases. J Clin Invest 118:1311–1321
Howard SC, Donnelly CA, Chan MS (2001) Methods for estimation of associations between mul-
tiple species parasite infections. Parasitology 122:233–251
Howard SC, Donnelly CA, Kabatereine NB, Ratard RC, Brooker S (2002) Spatial and intensity-
dependent variations in associations between multiple species helminth infections. Acta Trop
83:141–149
Huanca T (2006) Oral tradition, landscape, and identity in tropical forest. Wa-Gui, La Paz
Iancovici Kidon M, Stein M, Geller-Bernstein C, Weisman Z, Steinberg S, Greenberg Z, Handzel
ZT, Bentwich Z (2005) Serum immunoglobulin E levels in Israeli-Ethiopian children: environ-
ment and genetics. Isr Med Assoc J 7:799–802
Jackson JA, Frieber IM, Little S, Bradley JE (2008) Review series on helminths, immune modula-
tion and the hygiene hypothesis: immunity against helminths and immunological phenomena
in modern human populations: coevolutionary legacies? Immunology 126:18–27
Jiménez JC, Pinon A, Dive D, Capron M, Dei-Cas E, Convit J (2009) Antibody response in chil-
dren infected with Giardia intestinalis before and after treatment with Secnidazole. Am J Trop
Med Hyg 80:11–15
Kalyoncu AF, Stålenheim G (1992) Serum IgE levels and allergic spectra in immigrants to Sweden.
Allergy 47:277–280
Kaplan JE, Larrick JW, Yost JA (1980) Hyperimmunoglobulinemia E in the Waorani, an isolated
Amerindian population. Am J Trop Med Hyg 29:1012–1017
Karvonen A, Seppälä O, Tellervo VE (2009) Host immunization shapes interspecific associations
in trematode parasites. J Anim Ecol 78:945–952
Kron MA, Ammunariz M, Pandey J, Guzman JR (2000) Hyperimmunoglobulinemia E in the
absence of atopy and filarial infection: the Huaorani of Ecuador. Allergy Asthma Proc
21:335–341
Labeaud AD, Malhotra I, King MJ, King CL, King CH (2009) Do antenatal parasite infections
devalue childhood vaccination? PLoS Negl Trop Dis 3:1–6
Lafferty KD, Kuris AM (2002) Trophic strategies, animal diversity and body size. Trends Ecol
Evol 17:507–513
Lafferty KD, Sammond DT, Kuris AM (1994) Analysis of larval trematode communities. Ecology
75:2275–2285
Lello J, Hussell T (2008) Functional group/guild modelling of inter-specific pathogen interactions:
a potential tool for predicting the consequences of co-infection. Parasitology 135:825–839
386 M. Martin et al.
Lienhardt C, Azzurri A, Amedei A, Fielding K, Sillah J, Sow OY, Bah B, Benagiano M, Diallo A,
Manetti R, Maneh K, Gustafson P, Bennett S, D’Elios MM, McAdam K, Del Prete G (2002)
Active tuberculosis in Africa is associated with reduced Th1 and increased Th2 activity in vivo.
Eur J Immunol 32:1605–1613
Lim HK, Heyneman D (1972) Intramolluscan inter-trematode antagonism: a review of factors
influencing the host-parasite system and its possible role in biological control. Adv Parasitol
10:191–268
Lindberg R, Arroyave C (1986) Levels of IgE in serum from normal children and allergic children
as measured by an enzyme immunoassay. J Allergy Clin Immunol 78:614–618
Long KZ, Nanthakumar N (2004) Energetic and nutritional regulation of the adaptive immune
response and trade-offs in ecological immunology. Am J Hum Biol 16:499–507
Lynch NR, Lopez R, Isturiz G, Tenias-Salazar E (1983) Allergic reactivity and helminthic infec-
tion in Amerindians of the Amazon basin. Int Arch Allergy Appl Immunol 72:369–372
Maizels RM (2005) Infections and allergy - helminths, hygiene and host immune regulation. Curr
Opin Immunol 17:656–661
Maizels RM, Yazdanbakhsh M (2003) Immune regulation by helminth parasites: cellular and
molecular mechanisms. Nat Rev Immunol 3:733–744
Maizels RM, Pearce EJ, Artis D, Yazdanbakhsh M, Wynn TA (2009) Regulation of pathogenesis
and immunity in helminth infections. J Exp Med 206:2059–2066
Matowicka-Karna J, Dymicka-Piekarska V, Kemona H (2009) IFN-gamma, IL-5, IL-6 and IgE in
patients infected with Giardia intestinalis. Folia Histochem Cytobiol 47:93–97
May RM, Anderson RM (1979) Population biology of infectious diseases: Part II. Nature
280:455–461
May RM, Nowak MA, Nowak A (2009) Coinfection and the evolution of parasite virulence. Proc
Biol Sci 261:209–215
McDade TW (2005) Life history, maintenance, and the early origins of immune function. Am J
Hum Biol 17:81–94
Minchella DJ (1985) Host life-history variation in response to parasitism. Parasitology
90:205–216
Mingomataj EC, Xhixha F, Gjata E (2006) Helminths can protect themselves against rejection
inhibiting hostile respiratory allergy symptoms. Allergy 61:400–406
Moller AP (1990) Parasites and sexual selection: current status of the Hamilton and Zuk hypothe-
sis. J Evol Biol 3:319–328
Muehlenbein MP, Hirschtick JL, Bonner JZ, Swartz AM (2010) Toward quantifying the usage
costs of human immunity: altered metabolic rates and hormone levels during acute immune
activation in men. Am J Hum Biol 22:546–556
Mupfasoni D, Karibushi B, Koukounari A, Ruberanziza E, Kaberkuka T, Kramer MH, Mukabayire
O, Kabera M, Nizeyimana V, Deville MA, Ruxin J, Webster JP, Fenwick A (2009) Polyparasite
helminth infections and their association to anemia and undernutrition in Northern Rwanda.
PLoS Negl Trop Dis 3:e517
Ortega YR, Adam R (1997) Giardia: overview and update. Clin Infect Dis 25:545–549
Perlmann H, Helmby H, Hagstedt M, Carlson J, Larsson PH, Troye-Blomberg M, Perlmann P
(1994) IgE elevation and IgE anti-malarial antibodies in Plasmodium falciparum malaria: asso-
ciation of high IgE levels with cerebral malaria. Clin Exp Immunol 97:284–292
Perlmann P, Perlmann H, ElGhazali G, Blomberg MT (1999) IgE and tumor necrosis factor in
malaria infection. Immunol Lett 65:29–33
Pullan R, Brooker S (2008) The health impact of polyparasitism in humans: are we under-
estimating the burden of parasitic diseases? Parasitology 135:783–794
Quinnell RJ, Grafen A, Woolhouse MEJ (1995) Changes in parasite aggregation with age: a dis-
crete infection model. Parasitology 111:635–644
Råberg L, Graham AL, Read AF (2009) Decomposing health: tolerance and resistance to parasites
in animals. Philos Trans R Soc Lond B Biol Sci 364:37–49
Raso G, Luginbühl A, Adjoua CA, Tian-Bi NT, Silué KD, Matthys B, Vounatsou P, Wang Y, Dumas
ME, Holmes E, Singer BH, Tanner M, N’Goran E, Utzinger J (2004) Multiple parasite infec-
Make New Friends and Keep the Old? Parasite Coinfection and Comorbidity… 387
tions and their relationship to self-reported morbidity in a community of rural Côte d’Ivoire. Int
J Epidemiol 33:1092–1102
Rigaud T, Perrot-Minnot M-J, Brown MJF (2010) Parasite and host assemblages: embracing the
reality will improve our knowledge of parasite transmission and virulence. Proc Biol Sci
277:3693–3702
Rook GAW (2008) Review series on helminths, immune modulation and the hygiene hypothesis:
the broader implications of the hygiene hypothesis. Immunology 126:3–11
Round JL, Mazmanian SK (2009) The gut microbiota shapes intestinal immune responses during
health and disease. Nat Rev Immunol 9:313–324
Rousham EK (1994) An increase in Giardia duodenalis infection among children receiving peri-
odic antihelmintic treatment in Bangladesh. J Trop Pediatr 40:329–333
Sayasone S, Mak TK, Vanmany M, Rasphone O, Vounatsou P, Utzinger J, Akkhavong K, Odermatt
P (2011) Helminth and intestinal protozoa infections, multiparasitism and risk factors in
Champasack Province, Lao People’s Democratic Republic. PLoS Negl Trop Dis 5:e1037
Schad GA, Anderson RM (1985) Predisposition to hookworm infection in humans. Science
228:1537–1540
Schmid-Hempel P (2003) Variation in immune defence as a question of evolutionary ecology. Proc
Biol Sci 270:357–366
Sheldon BC, Verhulst S (1996) Ecological immunology: costly parasite defenses and trade-offs in
evolutionary ecology. Trends Ecol Evol 11:317–321
Silva NRD, Brooker S, Hotez PJ, Montresor A, Engels D, Savioli L (2003) Soil-transmitted hel-
minth infections: updating the global picture. Trends Parasitol 19:547–551
Supali T, Verweij JJ, Wiria AE, Djuardi Y, Hamid F, Kaisar MMM, Wammes LJ, van Lieshout L,
Luty AJF, Sartono E, Yazdanbakhsh M (2010) Polyparasitism and its impact on the immune
system. Int J Parasitol 40:1171–1176
Turner JD, Jackson JA, Faulkner H, Behnke J, Else KJ, Kamgno J, Boussinesq M, Bradley JE
(2008) Intensity of intestinal infection with multiple worm species is related to regulatory cyto-
kine output and immune hyporesponsiveness. J Infect Dis 197:1204–1212
Uller T, Isaksson C, Olsson M (2006) Immune challenge reduces reproductive output and growth
in a lizard. Funct Ecol 20:873–879
Van Baalen M, Sabelis MW (1995) The dynamics of multiple infection and the evolution of viru-
lence. Am Nat 146:881–910
Vasunilashorn S, Crimmins EM, Kim JK, Winking J, Gurven M, Kaplan H, Finch CE (2010)
Blood lipids, infection, and inflammatory markers in the Tsimane of Bolivia. Am J Hum Biol
22:731–740
Walson JL, John-Stewart G (2007) Treatment of helminth co-infection in individuals with HIV-1:
a systematic review of the literature. PLoS Negl Trop Dis 1:e102
Weidinger S, Gieger C, Rodriguez E, Baurecht H, Mempel M, Klopp N, Gohlke H, Wagenpfeil S,
Ollert M, Ring J, Behrendt H, Heinrich J, Novak N, Bieber T, Kramer U, Berdel D, von Berg
A, Bauer CP, Herbath O, Koletzko S, Prokisch H, Mehta D, Meitinger T, Depner M, von
Mutuis E, Liang L, Moffatt M, Cookson W, Kabesch M, Wichmann HE, Illig T (2008)
Genome-wide scan on total serum IgE levels identifies FCER1A as novel susceptibility locus.
PLoS Genet 4:e1000166
Wilson MS, Maizels RM (2004) Regulation of allergy and autoimmunity in helminth infection.
Clin Rev Allergy Immunol 26:35–50
Wolfe M (1992) Giardiasis. Clin Microbiol Rev 5:93–100
Woolhouse ME (1992) A theoretical framework for the immunoepidemiology of helminth infec-
tion. Parasite Immunol 14:563–578
Wu D, Molofsky AB, Liang H-E, Ricardo-Gonzalez RR, Jouihan HA, Bando JK, Chawla A,
Locksley RM (2011) Eosinophils sustain adipose alternatively activated macrophages associ-
ated with glucose homeostasis. Science 332:243–247
Yazdanbakhsh M, Kremsner PG, van Ree R (2002) Allergy, parasites, and the hygiene hypothesis.
Science 296:490–494
Primates, Pathogens, and Evolution:
A Context for Understanding Emerging
Disease
Introduction
1
Here we use the term “pathogen” broadly, to include both microparasites (viruses, bacteria, and
fungi) and macroparasites (such as worms).
K.N. Harper (*)
Department of Environmental Health Sciences, Columbia University, Black Building, Rm
1618, 650 W 168th Street, New York, NY 10032, USA
e-mail: [email protected]
M.K. Zuckerman
Department of Anthropology and Middle Eastern Cultures, Mississippi State University,
Mississippi State, MS 39762, USA
e-mail: [email protected]
B.L. Turner
Department of Anthropology, Georgia State University, Atlanta, GA 30303, USA
e-mail: [email protected]
G.J. Armelagos
Department of Anthropology, Emory University, Atlanta, GA 30322, USA
e-mail: [email protected]
J.F. Brinkworth and K. Pechenkina (eds.), Primates, Pathogens, and Evolution, 389
Developments in Primatology: Progress and Prospects,
DOI 10.1007/978-1-4614-7181-3_13, © Springer Science+Business Media New York 2013
390 K.N. Harper et al.
wild great ape populations, exacerbating existing threats posed by habitat loss,
human encroachment, and hunting (Boesch and Boesch-Achermann 2000; Leendertz
et al. 2004; Wolfe et al. 1998). Whether occurring as epidemics, small outbreaks, or
single deaths, infectious diseases can negatively affect the viability of small or iso-
lated NHP populations, due to the characteristically low reproductive rate of many
NHP species, especially great apes (Boesch and Boesch-Achermann 2000; Ferber
2000; Goodall 1970, 1983; Nishida 1990; Wallis 2000; Wolfe et al. 1998).
Understanding how pathogens move between human and NHP populations and
sometimes become established in a novel host requires knowledge of a pathogen’s
evolutionary history in its natural host, as well as the potential for transmission in
both its natural and novel hosts. However, establishing these evolutionary and epi-
demiological relationships has been complicated by methodological missteps and
conceptual pitfalls. Molecular methods allow us to construct a better picture of the
past disease-scape of primates, and the past few years have witnessed a veritable
explosion of information regarding major pathogen transmission events between
humans and NHPs. These data have helped to elucidate pathogens’ histories, though
much remains to be done. In this chapter, we discuss some of the theoretical issues
and methodological advances involved in reconstructing the evolutionary relation-
ships between pathogenic organisms, humans, and NHPs. Using examples of major
human diseases and their causative agents, specifically malaria (Plasmodium spp.)
and HIV (human immunodeficiency virus), we discuss the implications for under-
standing emerging infectious diseases in both humans and our closest relatives.
First, we explore the differentiation of the human disease-scape, attempting to
reconstruct which pathogens diverged with their human hosts over evolutionary
time. Next, we outline which human pathogens are believed to have resulted from
cross-species transmission from NHPs. Finally, we discuss examples of cross-
species parasite transmission from humans to NHPs and consider their ramifications
upon conservation biology.
Theoretical Considerations
Studies of human disease ecology suggest that infectious diseases affecting human
populations can generally be grouped into two broad categories: those with a long-
standing, millennial relationship with humans and the latter those that have been
Primates, Pathogens, and Evolution: A Context for Understanding Emerging Disease 391
more recently acquired. Sprent (1969a, b) was the first to make this distinction,
identifying two distinct classes of microbes that afflicted hunter-gatherers in the
Paleolithic: “heirloom species” and “souvenir species.” Heirloom species are patho-
gens that originated in our anthropoid ancestors and continued to infect hominins
and eventually modern humans. In contrast to heirloom species, souvenir species
are newer evolutionary acquisitions that are typically “picked up” via exposure to
zoonotic reservoirs or vectors (Kilks 1990). These diseases are generally zoonoses,
i.e., infectious diseases that can be transmitted from animals to humans. Zoonoses
can be contracted through a number of routes: from sympatric reservoir host species
through cross transfer, through insect or animal bites, or through the preparation and
consumption of contaminated animal flesh. It appears that many of these souvenir
species are capable of temporary host switches without significant levels of adapta-
tion, as long as the novel host is sufficiently similar to the natural host in terms of
resources available to the pathogen (Kellog 1896, in Brooks and Ferrao 2005, p.
1292). Thus, these microbes can remain specialized in their natural hosts while
infecting multiple hosts, including humans, across a wider ecological niche. In
some cases, however, a souvenir species enters the human population and stays
there permanently, adapting to its novel host and becoming established; indeed, as
will be discussed, this category of souvenir species includes some of the major
pathogens that affect human societies.
Given the complexity of pathogen host specificity and adaptation, it should
come as no surprise that reconstructing the history of human pathogens is a diffi-
cult endeavor. Often, in studying the origins of human pathogens, it is necessary to
gather samples from our close NHP relatives to assess the presence or absence of
a pathogen in a given species as well as to determine its genetic relationship to
human variants. As the examples ahead will show, this process frequently repre-
sents a limiting factor in our search for disease origins due to the difficulty inher-
ent in collecting biological samples from wild NHPs, many of which live in remote
areas of the world as well as being endangered and enjoying special protections. In
order to illustrate the effect that host switching and limited sampling of NHPs can
have on our understanding of a pathogen’s history in humans, we present the cau-
tionary tale of malignant malaria. By beginning our discussion of human pathogen
origins with this example, we hope to demonstrate the careful attention to sam-
pling necessary when considering the evolutionary history of even intensively
studied pathogens such as Plasmodium falciparum, never mind the myriad less-
studied ones discussed in this volume (e.g., Treponema, Toxoplasma, Pediculus,
and Pthirus lice).
illness (Snow et al. 2005) and nearly one million deaths annually (WHO 2010).
The pathogen has exerted tremendous selective pressure upon the human genome in
recent history (Hamblin et al. 2002; Sabeti et al. 2002; Tishkoff et al. 2001), but
fundamental questions persist about the Plasmodium’s history in humans.
Early molecular studies indicated that the protozoan’s closest relative was
P. reichenowi (Escalante and Ayala 1994), a parasite species found in chimpanzees.
This species was, for many years, represented by only a single strain isolated from
a chimp captured in the Democratic Republic of Congo several decades ago (Collins
et al. 1986). For this reason, many researchers believed that humans and chimpan-
zees both harbored their own distinct malaria strains; humans were infected by
P. falciparum, while chimps carried P. reichenowi. This was consistent with a sce-
nario in which each host species possessed its own heirloom Plasmodium species.
More intensive sampling, enabled by advances in noninvasive techniques, has
recently generated a novel twist on the origins of this pathogen by suggesting that
P. falciparum should instead be considered a souvenir species in humans. By exam-
ining blood as well as noninvasive samples collected from many wild chimpanzees,
researchers identified many new Plasmodium isolates among chimpanzee popula-
tions (Prugnolle et al. 2010; Rich et al. 2009). Sequence analysis showed that the
global genetic diversity of human P. falciparum was very low relative to the diver-
sity of chimpanzee P. reichenowi strains, indicating that the human species had
diverged more recently than the chimpanzee species. This finding was inconsistent
with P. falciparum being an heirloom pathogen in humans. Moreover, a more com-
prehensive study of fecal samples from nearly 3,000 wild great apes, including
chimpanzees, gorillas, and bonobos from throughout Central Africa, prompted
consideration of an alternative scenario regarding P. falciparum’s original host
(Liu et al. 2010). This study demonstrated that the gorilla branch of the Plasmodium
phylogeny included the strains most closely related genetically to human falci-
parum strains. Thus, it would appear that all circulating P. falciparum strains might
be the result of a single, very successful cross-species transmission event from
gorillas to humans.
Given the propensity of Plasmodium to switch primate hosts (Garamszegi 2009)
and the numerous twists and turns in the history of malaria thus far, it has been
noted that only further in-depth sampling of wild animals will confirm that no still-
closer relative is going undetected and that the current story is the correct one
(Prugnolle et al. 2011). Even with regard to a well-studied disease such as malaria,
our understanding may well change with future developments, which hinge on sam-
ple availability. As Wolfe et al. (2007) note, while resolving the debate surrounding
the origin of malaria will not necessarily assist with global eradication of the dis-
ease, it may contribute to our broader understanding of the dynamics of disease
emergence. With the potential importance of such knowledge in mind, as well as the
complexity involved in determining a pathogen’s history, we now turn to a discus-
sion of which human infections appear to be due to heirloom pathogens and which
appear to be caused by souvenir species instead.
Primates, Pathogens, and Evolution: A Context for Understanding Emerging Disease 393
During the Paleolithic, small human population sizes and, to a lesser extent, low
population density would have limited the diversity of possible heirloom species
affecting human and hominin populations by preventing sustained transmission of
many viruses and bacteria (Dunn et al. 2010, p. 2590). However, some species of
parasites and bacteria appear to have thrived in this setting. Characteristically, heir-
loom pathogens are organisms that are able to persist in small, dispersed popula-
tions; generate incomplete or short-lived immunity; or have a chronic course,
enabling prolonged transmission to new hosts. Several potential heirloom species
have been identified, including macroparasites such as head and body lice (Pediculus
humanus) (Reed et al. 2004) and pinworms (Enterobius vermicularis) (Hugot et al.
1999) as well as Staphylococci (Cockburn 1967, 1971; Sprent 1962, 1969a). Other
possible heirloom pathogens include the causative agents of yaws (Treponema pal-
lidum subsp. pertenue) (Harper et al. 2008) and typhoid (Salmonella typhi)
(Roumagnac et al. 2006).
The identification of extant pathogen species that belong to heirloom lineages
has been controversial, however, as the example of P. falciparum, above, illustrates.
Novel genetic data have assisted in this process, helping to clarify which species
most likely belong in the heirloom category by providing information on pathogen
divergence times and host histories. In addition, studies of host specificity in patho-
gens that infect wild NHP populations suggest that some major classes of parasites,
such as helminths, are more likely to be heirlooms due to their species specificity
than are protozoa and viruses, which are typically able to infect a wider swathe of
hosts (Pedersen et al. 2005).
Unique patterns of human behavior have no doubt guided the adaptation of heir-
loom pathogens. Weiss and Wrangham (1999), for example, note that while chim-
panzees and humans share 95–99 % of their DNA (Olson and Varki 2003), we only
share approximately 50 % of our pathogens. More specifically, pathogens and para-
sites affecting wild NHP species primarily include helminths, viruses, and protozoa,
while pathogens affecting humans are dominated by fungi and bacteria (Pedersen
et al. 2005) (Fig. 1). The reasons for these differences are poorly understood but
likely reflect divergent characteristics of host ecology and behavior.
The evolution of human herpesviruses may offer one example of an heirloom
pathogen molded over time by uniquely human behaviors. Phylogenetic studies
indicate that all eight members of the human herpesvirus family (Herpesviridae)
likely derive from an ancestral viral genome which infected the last common
ancestor of hominins and great apes (Gentry et al. 1988). Herpes simplex virus
(HSV) spreads from sites of initial infection in skin or mucosal surfaces to neuronal
cell bodies in order to establish latent infection, forming a long-term relationship
with its host. This long latency would have allowed HSV to persist in small, low-
density Paleolithic populations. There are two types of herpes simplex: HSV-1 pri-
marily produces oral herpes infection, while HSV-2 primarily produces genital
infection. HSV-2 appears to be the only type of herpesvirus among primates that is
394 K.N. Harper et al.
Fig. 1 Comparison of the taxonomic distribution of parasites in (a) free-living nonhuman pri-
mates and (b) humans. This data was based on a survey of 369 nonhuman primates and 1,415
humans (Figure reproduced from Pedersen et al. (2005))
transmitted primarily via sexual contact, and some have suggested that it may owe
its existence to uniquely human sexual practices. The divergence of HSV-1 and
HSV-2 dates back approximately 8–10 million years (McGeouch et al. 1995) and
presumably reflects the development of species-specific tropisms for the epithelium
of the oropharynx and the urogenital tract, respectively. For HSV-1 and HSV-2 to
take on these distinct tropisms, oral and genital sites had to become microbiologi-
cally isolated from each other, while oral–oral and genital–genital contact between
the hosts had to be maintained. McGeouch et al. (1995) have suggested that the
evolution of continual sexual attractiveness of hominin females throughout the
entire menstrual cycle, with an expected attendant increase in the frequency of sex-
ual intercourse, and the adoption of close face-to-face mating among hominins,
which may have facilitated the practice of kissing, provided the necessary condi-
tions for the evolutionary divergence of HSV-1 and HSV-2. This hypothesis, of
course, awaits rigorous testing, and more generally, the issue of how the behaviors
of particular primate species provide niches conducive to sexual transmission
remains a subject for further study.
Even absent the unique behavioral practices that characterize humans, however,
millions of years of evolution in different hosts ensures divergence among the
microbes that inhabit the bodies of humans and NHPs. For instance, the gut micro-
bial communities of the great apes, including humans, have evolved independently
with their hosts, diverging over the years in a manner consistent with host specia-
tion patterns (Ochman et al. 2010). Common shared ancestors are hypothesized to
exist between gut microflora species of humans and chimpanzees (Ushida et al.
2010) and also strepsirrhines (Bo et al. 2010). This suggests that some of the sym-
biotic microbes in human and NHP guts could be heirlooms. Understanding the
intersecting roles of physiological similarity, behavioral divergence, environmen-
tal context, and microbial colonization is therefore crucial when reconstructing
the evolutionary histories of heirloom microbes, both beneficial and pathogenic, in
humans and NHPs.
Primates, Pathogens, and Evolution: A Context for Understanding Emerging Disease 395
Similar to the example of malaria discussed above, investigation into the NHP
origins of HIV/SIV has demonstrated that continued research on even the most
well-studied disease agents can yield important and surprising insights into their
origins and evolution. HIV represents the best-known example of an NHP pathogen
transmitted to and then sustained within humans, and it remains one of the most
serious pandemics in history. The UNAIDS Report on the Global AIDS Epidemic
2010 estimates that 33.3 million people are infected with HIV worldwide, among
them 2.6 million children. It is estimated that some 25 million people worldwide
have died from HIV-/AIDS-related diseases (UNAIDS 2010). The majority of indi-
viduals with HIV live in sub-Saharan Africa (UNAIDS 2010). Adding to the dis-
ease burden in sub-Saharan Africa and other regions with high rates of HIV is the
fact that this infection disproportionately affects young adults (Patton et al. 2009),
leading to high morbidity and mortality among those individuals who would other-
wise be among the most economically active in their societies. A result of this pattern
of infection is a demographic crisis in which young children and older adults carry
the burden of looking after themselves, each other, and those who are infected. The
fact that a disproportionate burden of infection is found in countries with high rates
of political and economic inequality compounds this crisis even further (Fox 2012).
There are two types of HIV: HIV-1 and HIV-2. Scholars have understood for
some time that both types evolved from the simian immunodeficiency viruses (SIVs),
with HIV-1 deriving from the SIV variant of chimps (SIVcpz) and HIV-2 from the
SIV of sooty mangabeys (SIVsmm) (Gao et al. 1999; Hahn et al. 2000; Hirsch et al.
1995; Peeters et al. 2002; Weiss and Wrangham 1999). SIV infection is quite com-
mon in NHPs over 40 species-specific SIV variants have been documented in
396 K.N. Harper et al.
have noted that recombination between SIVs and circulating HIVs may pose a threat
to human health if it results in novel strains with an increased ability to exploit
our species.
In conclusion, though the sequence of events that allowed for permanent
establishment and spread of HIV-1 and HIV-2 in humans remains uncertain
(Pepin 2011), molecular and epidemiological studies are providing a richer context
for understanding these souvenir pathogens. As discussed, it has been known for
some time that at least two host switches, one from chimpanzees and one from sooty
mangabeys, resulted in HIV-1 and HIV-2, respectively. Recent research, though, has
demonstrated that a third host switch from gorillas may have resulted in one rare
HIV subtype (HIV-1 group O) (Van Heuverswyn et al. 2006). Additionally, molecular
epidemiological studies have demonstrated that cross-species exposure to SIV is
ongoing (Djoko et al. 2012) and may at times even result in low-level transmission
among humans (Plantier et al. 2009). Thus, the potential for novel forms of HIV to
take root in humans appears to be present, though most SIV transmission events
seem to fizzle out quickly.
Malaria and HIV, both discussed extensively in this chapter, are examples of patho-
gens that originally came from NHPs but found a permanent home in humans.
There are multiple reasons why a given NHP host species may or may not become
an established source of infection for humans. Phylogenetic relatedness and physi-
cal and environmental proximity are two primary—and tightly interwoven—factors
which are likely to affect this dynamic. In an assessment of the animal origins of 25
major human infectious human diseases, Wolfe et al. (2007) partially attributed the
finding that the majority of tropical infectious diseases arose in the Old World rather
than the New World to the greater genetic distance separating New World monkeys
and humans. It represents roughly twice the genetic distance between Old World
monkeys and humans and many times that between humans and Old World apes. At
the same time, physical proximity and shared habitats are likely to play a substantial
role. For example, Wolfe et al. (2007) partially attributed the high number of souve-
nir infections arising in the Old World to the greater evolutionary time available for
transfers between primates and humans there (c. 7–8 million years) as compared to
the New World (c. < 14,000 years).
On a regional level, some populations have maintained fairly frequent exposure
to NHPs. In Africa, South America, and Asia, in particular, communities living in
close proximity to NHPs often become involved in activities associated with a high
risk of exposure to NHP pathogens. Bushmeat handling and consumption provides
one of the most effective means for the spread of pathogens from NHPs to humans
398 K.N. Harper et al.
(Fig. 2). Almost 100 % of villagers in rural forested areas of Cameroon have
reported eating NHPs, with over 70 % involved in hunting and 30 % active in butch-
ering (Wolfe et al. 2004a). Both activities involve repeated contact with potentially
infective body fluids and tissues. They also generate opportunities for pathogen
transmission to other individuals and communities linked by the bushmeat trade. A
market survey of two cities in Equatorial Guinea recorded 4,222 primate carcasses
on sale over 424 days (Fa et al. 1995), illustrating the great importance of NHP
hunting to local economies.
Primates, Pathogens, and Evolution: A Context for Understanding Emerging Disease 399
The prevalence of human infections derived from NHPs suggests that these
pathogens are able to exploit between-species interactions effectively. For example,
Wolfe et al. (2005) found that bushmeat hunters in rural Cameroon are infected with
a wide variety of human T-lymphotropic viruses (HTLVs), which are linked to leu-
kemia, lymphoma, and HTLV-associated myelopathy, as well as multiple simian
T-lymphotropic virus (STLV)-1-like viruses. The high diversity of these viruses in
Cameroon indicates ongoing cross-species transmission from NHPs to humans.
Similarly, at least three independent examples of NHP-to-human transmission of
simian foamy virus (SFV), which to date has not been linked to any signs of disease
in humans, have been confirmed in hunters (Wolfe et al. 2004b). Each event derived
from distinct NHP lineages—De Brazza’s guenons, mandrills, and gorillas—indi-
cating that many species can potentially infect humans. Bites and scratches by
NHPs, in particular, appear to be a very efficient means of transmitting viruses.
In one study, over 35 % of hunters reporting such wounds were seropositive for SFV
infection, and almost 2 % of serum samples belonging to adults from Cameroon
were found to be seropositive as well, underscoring the high prevalence of cross-
species transmission opportunities (Calattini et al. 2007).
While the factors leading to introduction of NHP pathogens into human popula-
tions are increasingly well characterized, the conditions that facilitate their estab-
lishment in our species are poorly understood. For example, though studies of
Cameroonian hunters suggest that exposure to STLVs via NHP blood contributes to
a greater diversity of HTLVs than is found in other populations (Wolfe et al. 2005),
only HTLV-1 and HTLV-2 appear to have established themselves worldwide.
Similarly, reports of the presence of dual HIV-1 and SFV infections in a commercial
sex worker from the Democratic Republic of Congo and in a blood donor from
Cameroon suggest the potential for SFV to be transmitted from human to human via
sex or blood, as is HIV (Switzer et al. 2008). Such human-to-human transmission
events appear to occur relatively rarely in the case of SFV, however. At present, our
ability to predict which souvenir pathogens will “take off,” flourish, and become
established within human populations is poor. Nonetheless, continued attention to
the movement of pathogens across the NHP-human interface may contribute to our
knowledge of this process.
It is worrisome that the trend of increased contact between NHPs and humans
seems to be intensifying. Population pressure, expanded ecotourism and conserva-
tion programs, and ever more powerful technological advances are enabling humans
to encroach further and further into NHP habitats (Auzel and Hardin 2001). For
instance, the villages surrounding logging concessions in Equatorial Africa have
grown rapidly, often having increased from a few hundred individuals to several
thousand. Political instability and forced migration, from Liberia into Sierra Leone,
for example, have also played a role in the dramatic redistribution of human popula-
tions into areas of increased contact with NHPs (Hodges and Heistermann 2003).
This new proximity, paired with the desirability of NHPs as prey, has increased
opportunities for cross-species transmission events. In addition, people in these
areas can utilize new and improved roads to transport bushmeat from remote vil-
lages to major cities, greatly increasing the number of humans who come into
400 K.N. Harper et al.
contact with NHP carcasses. In truth, contact with wild NHPs now spans continents,
reaching consumers who have never set foot in wild NHP habitat (Ellicott 2011);
examination of bushmeat samples confiscated at US airports has revealed NHP tis-
sue infected with SFV and herpesviruses (Smith et al. 2012). It is expected that
these trends will intensify in the future, in the absence of decisive community-level
and governmental actions to regulate ecotourism and constrain environmental
destruction and the bushmeat trade.
Naturally, the increasing proximity between humans and NHPs also leads to greater
opportunity for human pathogens to infect both captive and wild NHPs. In general,
as the level of interaction between humans and NHPs increases, so does the risk of
transmission of diseases such as measles and tuberculosis (Wolfe et al. 1998). Not
surprisingly, there are many well-documented examples of captive NHPs becoming
infected with human pathogens, with “immunologically naïve” great apes proving
especially susceptible. For instance, there are frequent reports of tuberculosis infec-
tions of human origin among captive NHPs (Montali et al. 2001). Poliovirus can
also infect chimpanzees and gorillas, as well as more distantly related anthropoids,
like Colobus monkeys (Brack 1987; Suleman et al. 1984). As such, accidental expo-
sure to infected laboratory workers has led to poliovirus infections of chimpanzees
and gorillas since the 1940s (Ruch 1959). In another example, Arcobacter butzleri,
which is a member of the same bacterial family as Campylobacter
(Campylobacteraceae) and is associated with chronic diarrhea in humans, was
implicated in a spate of cases of chronic diarrhea among captive primate popula-
tions at a research center (Andersen et al. 1993).
There is accumulating evidence for similar episodes of disease transmission in
the wild (Adams et al. 1999; Homsy 1999; Wallis 2000; Wolfe et al. 1998; Woodford
et al. 2002). There are a number of “likely” instances of human-to-NHP transmis-
sion. In perhaps the most infamous instance, in 1966, six chimpanzees at Gombe
Stream National Park in Tanzania died from a polio-like virus, and six others were
paralyzed for life, shortly after a polio epidemic swept through neighboring human
settlements (Wallis 2000). Unfortunately, as no biological samples were collected
from the animals, it was impossible to verify if the epidemic was due to a poliovirus
introduced by local human populations or researchers (Wolfe et al. 1998). Confirmed
examples of human to wild NHP transmission are relatively rare, due to the diffi-
culty inherent in collecting samples but include Cryptosporidium infections in
mountain gorillas (Nizeyi et al. 2002) as well as the cases discussed below.
Respiratory diseases, in particular, have been recognized as a major source of
morbidity and mortality among free-living NHPs. This class of diseases is widely
regarded as the most important cause of morbidity and mortality among wild great
Primates, Pathogens, and Evolution: A Context for Understanding Emerging Disease 401
Conclusion
The surveillance of humans living in close proximity to NHPs has revealed tantaliz-
ing clues about the process of disease transmission from NHPs to humans. Perhaps
intensive study of circulating strains of HIV/SIV, HTLV/STLV, and SFV will
increase our knowledge of the features which characterize cross-species transmis-
sion events that result in subsequent sustained transmission between humans. In
addition, the explosion of knowledge surrounding pathogens of NHP origin, such as
P. falciparum and HIV, in the last few years underscores the need to delve into the
disease-scape of closely related NHPs to better understand our own infections. For
instance, it was not until 2010 that researchers demonstrated that wild chimps in the
Taï Forest appeared to be naturally infected with five different Plasmodium species
(Kaiser et al. 2010). Casting a wider net for pathogens may lead to similar advances
in our understanding of the other pathogens that make up the NHP disease-scape.
It is likely that the rapid development of sophisticated, noninvasive means of
NHP sampling will provide insight into the existence and/or prevalence of various
pathogens. These noninvasive methods include approaches similar to those devel-
oped in primatology to study endocrinology (e.g., Deschner et al. 2003; Hodges and
Heistermann 2003), characterize NHP genetics (e.g., Boesch et al. 2006; Bradley
Primates, Pathogens, and Evolution: A Context for Understanding Emerging Disease 403
et al. 2004; Vigilant et al. 2001), and perform urine assessment (e.g., Knott 1996;
Krief et al. 2005). For instance, fecal samples have already been used to assay SIV
and malaria in NHPs. In addition, a recent retrospective study of the epidemics in
the Taï Forest suggests that it is possible to monitor infections caused by paramyxo-
viruses, such as hMPV and respiratory syncytial virus, using fecal samples (Köndgen
et al. 2010). A pioneering study using aDNA techniques to study curated early-
twentieth-century NHP skeletal material even suggests that we can explore the his-
tory of viruses such as STLV using museum specimens (Calvignac et al. 2008).
Such noninvasive approaches may also help address the role of human pathogens
in NHP demographic declines. There are many unresolved questions. Do human
pathogens associated with epidemics in NHPs tend to rapidly infect small groups
before burning out? Or are some human-derived pathogens circulating continuously
and even adapting to their new hosts? In some cases, whether or not the pathogens
sweeping through NHP populations derive from humans is not clear. For example,
S. pneumoniae was recently found to be responsible for clusters of sudden death in
the chimpanzees of the Taï Forest (Chi et al. 2007). However, comparison to
sequences obtained from people living nearby suggested that the pathogen might
not be of human origin. Chimpanzees in different areas, but in frequent contact with
one another, were also found to harbor distinct S. pneumoniae clones. If not from
humans, from what reservoir did this pathogen arise? Similarly, adenoviruses were
isolated from two chimpanzees with signs of acute respiratory disease in Mahale
Mountains National Park (Tong et al. 2010). Sequence analysis identified two dis-
tinct viruses, but their origin was unclear; they could have been acquired from
humans, acquired from another species, or circulating among chimpanzees for some
time. Further investigation may shed light on the processes underlying such
outbreaks.
Continued study of the relationship between humans, pathogens, and NHPs con-
fers several substantial benefits. First, knowledge in this area is fundamental when
assessing the impact of pathogen exchange between species (Wolfe et al. 2007),
including research on the footprint of natural selection imposed by various infec-
tious diseases upon the human genome (e.g., Hamblin et al. 2002; Sabeti et al. 2002;
Tishkoff et al. 2001). Second, understanding more about how NHP pathogens are
introduced into human populations and then spread has practical implications.
According to Wolfe et al. (2007), benefits include a better understanding of disease
emergence and the potential for novel laboratory models helpful in studying public
health threats. Applications might include the development of indicators useful in
monitoring pathogen transmission between NHPs and high-risk individuals, such as
hunters and wildlife veterinarians; predicting which NHP pathogens might repre-
sent a future threat; and detecting and even controlling local human outbreaks
before they become epidemics (Wolfe et al. 2007). Third, some scholars have argued
that wild NHPs can serve as “sentinel species” for predicting disease outbreaks
among humans (Leendertz et al. 2006; Rouquet et al. 2005).
Finally, studying the relationship between pathogens, NHPs, and humans has
important conservation implications. Köndgen et al. (2008) and others have argued
that the close proximity between NHPs and humans, which is critical to both
404 K.N. Harper et al.
References
Adams H, Sleeman J, New J (1999) A medical survey of tourists visiting Kibale National Park,
Uganda, to determine the potential risk of disease transmission to chimpanzees (Pan troglo-
dytes) from ecotourism. In: Baer C (ed) Proceedings of the American Association of Zoo
Veterinarians. American Association of Zoo Veterinarians, Media, Philadelphia, PA
Andersen K, Kiehlbauch JA, Anderson DC, McClure HM, Wachsmuth IK (1993) Arcobacter
(Campylobacter) butzleri-associated diarrheal illness in a nonhuman primate population. Infect
Immun 61(5):2220–2223
Apetrei C, Metzger M, Richardson D, Ling B, Telfer P, Reed P, Robertson D, Marx P (2005)
Detection and partial characterization of simian immunodeficiency virus SIVsm strains from
bush meat samples from rural Sierra Leone. J Virol 79(4):2631–2636
Auzel P, Hardin R (2001) Colonial history, concessionary politics, and collaborative management
of Equatorial African rain forests. In: Bakarr M, Fonseca GD, Mittermeier R, Rylands A,
Painemilla K (eds) Hunting and Bushmeat utilization in the African Rain Forest: Perspectives
Toward a Blueprint for Conservation Action. Conservation International, Washington
Bibollet-Ruche F, Galat-Luong A, Cuny G, Sarni-Manchado P, Galat G, Durand J-P, Pourrut X,
Veas F (1996) Simian immunodeficiency virus infection in a patas monkey (Erythrocebus
patas): evidence for cross-species transmission from African green monkeys (Cercopithecus
aethiops sabaeus) in the wild. J Gen Virol 77(773–781)
Bo X, Zun-xi H, Xiao-yan W, Run-chi G, Xiang-hua T, Yue-lin M, Yun-Juan Y, Hui S, Li-da Z
(2010) Phylogenetic analysis of the fecal flora of the wild pygmy loris. Am J Primatol
72:699–706
Boesch C, Boesch-Achermann H (2000) The chimpanzees of the Taï Forest: behavioural ecology
and evolution. Oxford University Press, Oxford/New York
Boesch C, Kohou G, Nene H, Vigilant L (2006) Male competition and paternity in wild chimpan-
zees of Taï Forest. Am J Phys Anthropol 130:103–115
Boivin G, De Serres G, Cote S, Gilca R, Abed Y, Rochette L, Bergeron M, Dery P (2003) Human
metapneumovirus infections in hospitalized children. Emerg Infect Dis 9:634–640
Brack M (1987) Agents transmissible from simians to man. Springer, Berlin
Bradley B, Doran-Sheehy D, Lukas D, Boesch C, Vigilant L (2004) Dispersed male networks in
western Gorillas. Curr Biol 14:510–513
Brooks D, Ferrao A (2005) The historical biogeography of co-evolution: emerging infectious
diseases are evolutionary accidents waiting to happen. J Biogeogr 32:1291–1299
Calattini S, Betsem E, Froment A, Mauclère P, Tortevoye P, Schmitt C, Njouom R, Saib A, Gessain
A (2007) Simian foamy virus transmission from apes to humans, rural Cameroon. Emerg Infect
Dis 13(9):1314–1320
406 K.N. Harper et al.
Calvignac S, Terme J, Hensley S, Jalinot P, Greenwood A, Hänni C (2008) Ancient DNA identifi-
cation of early 20th century simian T-cell leukemia virus type 1. Mol Biol Evol
25(6):1093–1098
Chen Z, Luckay A, Sodora DL, Telfer P, Reed P, Gettie A, Kanu JM, Sadek RF, Yee J, Ho D et al
(1997) Human immunodeficiency virus type 2 (HIV-2) seroprevalence and characterization of
a distinct HIV-2 genetic subtype from the natural range of simian immunodeficiency virus-
infected sooty mangabeys. J Virol 71(5):3953–3960
Chi F, Leider M, Leendertz F, Bergmann C, Boesch C, Schenk S, Pauli G, Ellerbrok H, Hakenbeck
R (2007) New Streptococcus pneumoniae clones in deceased wild chimpanzees. J Bacteriol
189(16):6085–6088
Cockburn T (1967) Infections of the order primates. In: Cockburn T (ed) Infectious diseases: their
evolution and eradication. CC Thomas, Springfield, IL
Cockburn T (1971) Infectious disease in ancient populations. Curr Anthropol 12(1):45–62
Collins W, Skinner J, Pappaioanou M, Broderson J, Mehaffey P (1986) The sporogonic cycle of
Plasmodium reichenowi. J Parasitol 72(2):292–298
Deschner T, Heistermann M, Hodges K, Boesch C (2003) Timing and probability of ovulation in
relation to sex skin swelling in wild West African chimpanzees, Pan troglodytes verus. Anim
Behav 66:551–560
Djoko C, Wolfe N, Aghokeng A, Lebreton M, Liegeois F, Tamoufe U, Schneider B, Ortiz N,
Mbacham W, Carr J et al (2012) Failure to detect simian immunodeficiency virus infection in
a large Cameroonian cohort with high non-human primate exposure. Ecohealth 9(1):17–23
Dunn R, Davies T, Harris N, Gavin M (2010) Global drivers of human pathogen richness and
prevalence. Proc R Soc B 277:2587–2595
Ellicott C (2011 March 11) Meat from chimpanzees ‘is on sale in Britain’ in lucrative black mar-
ket. Daily Mail, London
Escalante A, Ayala F (1994) Phylogeny of the malarial genus Plasmodium, derived from rRNA
gene sequences. Proc Natl Acad Sci USA 91:11373–11377
Fa J, Juste J, Val JD, Castroviejo J (1995) Impact of market hunting on mammal species in equato-
rial Guinea. Conserv Biol 9(5):1107–1115
Ferber D (2000) Primatology. Human diseases threaten great apes. Science 289:1277–1278
Fox A (2012) The HIV-poverty thesis re-examines: poverty, wealth or inequality as a social deter-
minant of HIV infection in Sub-Saharan Africa? J Biosoc Sci 25:1–22
Gao F, Bailes E, Robertson D, Chen Y, Rodenburg C, Michael S, Cummins L, Arthur L, Peeters M,
Shaw GM et al (1999) Origin of HIV-1 in the chimpanzee Pan troglodytes troglodytes. Nature
397:436–441
Garamszegi L (2009) Patterns of co-speciation and host switching in primate malaria parasites.
Malar J 8:110–124
Gentry G, Lowe M, Alford G, Nevias R (1988) Sequence analysis of herpesviral enzymes suggest
an ancient origin for human sexual behavior. Proc Natl Acad Sci 85:2658–2661
Goldberg T, Gillespie T, Rwego I, Wheeler E, Estoff E, Chapman C (2007) Patterns of gastrointes-
tinal bacterial exchange between chimpanzees and humans involved in research and tourism in
western Uganda. Biol Conserv 135:511–517
Goodall J (1970) In the shadow of man. Collins, London
Goodall J (1983) Population dynamics during a 15-year period in one community of free-living
chimpanzees in the Gombe National Park, Tanzania. Z Tierpsychol 61:1–60
Goodall J (1986) The chimpanzees of Gombe: patterns of behaviour. Harvard University Press,
Cambridge, MA
Hahn B, Shaw G, Cock KD, Sharp P (2000) AIDS as a zoonosis: scientific and public health impli-
cations. Science 287(5453):607–614
Hamblin M, Thompson E, Rienzo AD (2002) Complex signatures of natural selection at the Duffy
blood group locus. Am J Hum Genet 70(2):369–383
Hanamura S, Kiyono M, Lukasik-Braum M, Mlengeya T, Fujimoto M, Nakamura M, Nishida T
(2007) Chimpanzee deaths at Mahale caused by a flu-like disease. Primates 49:77–80
Primates, Pathogens, and Evolution: A Context for Understanding Emerging Disease 407
J.F. Brinkworth and K. Pechenkina (eds.), Primates, Pathogens, and Evolution, 411
Developments in Primatology: Progress and Prospects,
DOI 10.1007/978-1-4614-7181-3, © Springer Science+Business Media New York 2013
412 Index
D Enteroviruses, 340–342
DAMPS. See Danger-associated molecular Environment of evolutionary adaptedness
patterns (DAMPS) (EEA), 333–335, 352, 354
Danger-associated molecular patterns Eosinophils, 35, 367, 368
(DAMPS), 19 granule content, 38, 39
DARC. See Duffy antigen receptor for and heterophils, 38
chemokines (DARC) origins, 35
Daubentonia (aye-aye), 23 phenotypes, 38
Dendritic cells (DCs), 26–29, 36, 44, 141, 263, role in immunity, 35
266, 344, 345, 349, 350 tissue remodeling, 38, 39
description, 31 Epidemiological transitions, 333, 335–337,
immunodeficiency viruses, 33 346, 352, 354–355
inter-primate differences, 26, 32 Epidemiology, 180, 268–274, 336, 341–342,
markers for primates, 32 354, 401
morphology and function, 31–32 leishmaniasis, 260
myeloid, 32 malaria, 118, 120
plasmacytoid, 32, 100, 307–308, 351 toxoplasma, NWP (see New World
TLR expression, 33 primates (NWPs))
Dengue Treponema, 189–214
inter-primate differences, 93 Episodic selection, 71, 73
Diabetes, type 1 (T1D), 126, 332, 337, 341, Erythrocebus patas (Patas monkey), 170, 171,
342, 354 197, 201, 396
Disease susceptibility, 3–6, 79, 94, 101, and Treponema sp., 197, 201
143, 307 Erythrocyte, 81, 118–130, 144
Duffy antigen receptor for chemokines enzymes and deficiency
(DARC), 80, 136 G6PD, 137–138
alleles, 134 PK, 138–139
antigens, 133–134 genetic variation and malaria
diversity in humans, 133–135 enzymes and deficiency, 137–139
haplotypes, 134 hemoglobin variants (see Hemoglobin)
Hepatocystis, 81 membrane proteins and surface
Plasmodium, 81, 133–134, 137 antigens, 131–137
Plasmodium vivax, 134–135 membrane proteins and surface antigens
ABO, 135–137
DARC, 133–135
E elliptocytosis and ovalocytosis,
Early humans, 333, 342 131–132
and chimpanzees, 168 glycophorin A and B receptors,
gorillas, 168 132–133
and lice, 168 Escherichia coli, 92, 95, 136, 401
and pathogens, 165 Eulemur (Brown lemur), 23
Ebola, 389, 395 Euoticus (needled-clawed bushbaby), 23
Ecology, NWPs. See New World primates
(NWPs)
Ecotourism, 7, 399–401, 403–404 F
Ectoparasite hypothesis, 169, 172. See also Faecalibacterium prausnitzi, 339, 342
Lice Fasciola hepatica, 341
Eczema, 45, 337, 343 Firmicutes
EEA. See Environment of evolutionary and Crohn’s disease, 339
adaptedness (EEA) Flukes, 363, 367
E3L, 4 Fluorescent treponemal antibody absorption
Enterobacteriaceae, 338–339 (FTA-ABS), 196, 204, 209
Enterobius vermicularis (pinworm), 341, 343, Fossil apes
344, 367, 393 Nakalipithecus nakayami, 174
Index 415
Helicobacter pylori, 136, 340, 341 Hepatitis C virus (HCV), 93, 102
Heligmosomoides polygyrus, 341, 349 interprimate differences, 93
Helminths, 34, 37, 38, 44, 333, 337, 338, TLR7 polymorphism, 102
341, 343–344, 348–351, 354, Herpes simplex virus (HSV), 102, 223,
363–382, 393 393, 394
adults, 370, 378–380, 382 HSV emergence, 393–394
affect, 44, 365, 377, 382, 393 HSV emergence, 393–394
chronic infection, 343, 349, 354, 364, 368, γ-Herpesviruses, 91, 221–242
370, 377, 380–382 cancer, 222–224
cost of infection to host, 370 Epstein-Barr virus (EBV), 221, 223–225
definition, 367 herpesvirus saimiri (HVS), 221, 222, 224
diminish immune responses, 370 and IL-2, 236
and G. lamblia infection, 371, interprimate differences in
373–378, 380 manifestation, 236
human IgE, 44, 367–370 human, 221, 223–225, 230, 234,
and IgE, 37, 44, 368–369, 377, 380 236–238, 240
immune phenotypes, 368–369 Herpesvirus saimiri (HVS), 221, 222, 224,
immunological role, 343 226–228, 231, 233–239, 241, 242
immunomodulatory effects, 350, 380 Kaposi sarcoma (see Kaposi’s sarcoma-
and immunoregulation, 333, 348–351, 364, associated herpesvirus (KSHV))
368, 380 Kaposi’s sarcoma-associated herpesvirus
induced immune responses in humans, 367 (KSAV), 221, 223–233
induce Treg activity, 368, 378 Callitrix jacchus (marmoset) model,
and inflammatory disease, 44, 343–344, 225
378, 380 control of lymphocyte activation,
morbidity, 364, 365, 370, 371 226–228
multispecies infections in humans, and HIV, 224
364–366, 375 interference with BCR signaling, 232
and neutrophils, 34 and multicentric Castleman’s disease
prototypical TH2/Treg response, 368 (MCD), 223–225, 228
shift T cell populations, 368 and primary effusion lymphoma (PEL),
soil transmitted helminth (STH) 223–225, 232
infections, 370 subordination of apoptosis, 231
T cell responses, 368 life cycle, 221–222, 240
TH2 immune response, 367, 368, 370 Lymphocryptovirus, 221, 224
tolerance response, 368 lytic replication, 221, 222, 224, 230, 231,
treatment, 163, 165, 337, 341, 343, 349, 234, 240, 241
370, 377, 378, 382 miRNAs (see microRNAs (miRNAs))
Tsimane of Bolivia, 371–381 molecular piracy, 222–223, 226–233, 242
Hemoglobin, 118, 120–131 mouse herpesvirus 68 (mHV68), 221, 224
geographic distribution of variants, phylogenetic tree, 221, 223
121–122 phylogeny, 223
geographic distribution of variants, respective host and malignancies, 221,
121–122 222, 224
HBB alleles and malaria, 120–131, 146 Rhadinovirus, 221, 224, 226, 236
HBC, malaria resistance, 120, 122, 123, rhesus rhadinovirus (RRV), 221, 222, 224,
128, 129, 131 233–236
HbC, malaria resistance, 122, 123, 128, 129 prevalence, 234
HBE, malaria resistance, 120, 122, 123, similarity to KSHV, 234, 235
128–131 STP, 226–231, 236–239, 342
HbE, malaria resistance, 122, 123, Tip, 226–230, 237–239, 242
128–129, 131 viral signaling molecules, 239
HBS, malaria resistance, 120–123, 128 virus-induced tumorigenesis, 222, 239
HbS, malaria resistance, 121–123, 129 Herpesvirus saimiri (HVS), 221, 222, 224,
Hepatitis A virus (HAV), 340, 341, 350 226–231, 233–239, 241, 242
Index 417
Herpesvirus saimiri U RNAs (HSUR), 241 HSUR. See Herpesvirus saimiri U RNAs
HERV-W, 5 (HSUR)
HIV. See Human immunodeficiency virus HSV. See Herpes simplex virus (HSV)
(HIV) HTLVs. See Human T-lymphotropic viruses
HIV-1. See Human immunodeficiency virus-1 (HTLVs)
(HIV-1) Human disease-scape, 213, 390–400
HIV-2, 67, 100, 291, 294, 297, 298, 310–313, heirloom pathogens and human Paleolithic
316, 395–397 ecology, 390, 393–394
HLA. See Human leukocyte antigen (HLA) HIV, 395–397
hMPV. See Human metapneumovirus (hMPV) malaria, 390–392
Holobiont, 4 pathogens to humans
Hominins bushmeat handling and consumption,
Ardipithecus kaddaba, 166 397–398
Ardipithecus ramidus, 166 ecotourism and constrain
Australopithecus afarensis, 166, 175, environmental destruction, 400
177–180 genetic distance, 397
Australopithecus africanus, 166 HTLVs, 399
Australopithecus anamensis, 166, 175 SFV, 399–400
Australopithecus bahrelghazali, 166, 175 STLV, 399
Homo erectus, 167 transmission, individuals and
Homo habilis, 172 communities, 398
Homo neanderthalensis, 167 souvenir species and changing human
Kenyanthropus platyops, 166, 175 ecology, 395
Orrorin tugenensis, 166 Human immunodeficiency virus-1 (HIV-1),
Paranthropus, 175 8, 73, 80, 102, 106, 169, 291,
Sahelanthropus tchadensis, 166 292, 294, 296–302, 304–316,
Homo sapiens 395–397, 399
Dengue virus, 93 in Africa, 291, 297–300, 305, 316, 396
gram-negative bacterial sepsis, 92, 94, 98 animal models
hepatitis C virus (HCV), 93 HIV-2, 310–313
historical exposure to pathogens, 106 rhesus macaques, 310, 312
HIV, 92, 291–292, 296–315, 395–397 SHIVs, 313
Kaposi’s sarcoma-associated herpesvirus T cells, 312–314
(KSHV), 224, 225 vpu and vpx, 312
lice and loss of hair, 169, 171 annotated depiction, genomes, 298
monkey B virus, 94 CCR5, 75, 106, 302, 314
Neisseria gonorrhoeae, 92, 99 CD8+ T cells, 305–307
Plasmodium falciparum, 92 CD4+ T cells and disease progression,
Schistosoma mansoni, 92, 99 301, 306
TLR polymorphisms, 102 chimpanzees, 80, 169, 291, 296, 297,
Toxoplasma gondii, 92, 262, 267–268 313–316, 396, 397
Hookworm, 343, 347, 367, 370, 372–378 cross-species transmission, 296–298, 397
and antagonism, 377–378 discovery of AIDS and HIVs, 291
and Ascaris lumbricoides, 367, 370, emergence, 8, 291, 292, 294, 296–302,
373–378 304–316
infections, 343, 370, 373–378 groups, 292, 296–298, 300, 301, 314, 315,
odds ratios, infection, 376 396, 397
Tsimane intestinal parasites, 371–373 humanized mice, 313
Tsimane single-and multiple-species immune activation, 306–308
infections, 373, 374 infection, 80, 106, 291, 292, 296, 297,
Host–parasite interactions, 260, 363 299–302, 304–308, 311–316, 395,
Host-pathogen coevolution, 2–3, 105, 148 396, 399
hRSV. See Human respiratory syncytial virus LPS, 305
(hRSV) lymph node environment, 308–310
418 Index
V Z
Varecia, (ruffed lemur), 23 Zoonotic transmissions, 168–169, 391