Pneumocystis: from a doubtful unique entity to a group of highly diversified fungal species

16
MINIREVIEW Pneumocystis: from a doubtful unique entity to a group of highly diversi¢ed fungal species Magali Chab ´ e 1,2 ,C´ ecile-Marie Aliouat-Denis 1,2 , Laurence Delhaes 1,3 , El Moukhtar Aliouat 1,2 , Eric Viscogliosi 1 & Eduardo Dei-Cas 1,2,3 1 Biology and Diversity of Emergent Eukaryotic Pathogens (BDEEP)–Center for Infection and Immunity of Lille, Pasteur Institute of Lille, Inserm U1019, CNRS UMR 8204, University Lille-Nord-de-France, Lille, France; 2 Department of Parasitology–Mycology, Faculty of Biological and Pharmaceutical Sciences, University of Lille-Nord-de-France, Lille, France; and 3 Department of Parasitology–Mycology, Faculty of Medicine, University of Lille-Nord- de-France, Biology-Pathology Centre, University Hospital Center, Lille, France Correspondence: Magali Chab ´ e, Department of Parasitology–Mycology, Faculty of Biological and Pharmaceutical Sciences, University of Lille-Nord-de-France, 3, rue du Professeur Laguesse BP83, 59006 Lille, France. Tel.: 133 3 2096 4010; fax: 133 3 2087 7276; e-mail: magali.chabe@univ- lille2.fr Received 11 May 2010; revised 10 October 2010; accepted 12 October 2010. Final version published online 26 November 2010. DOI:10.1111/j.1567-1364.2010.00698.x Editor: Teun Boekhout Keywords Pneumocystis spp.; phylogenetic species concept; life cycle; epidemiology. Abstract At the end of the 20th century the unique taxonomically enigmatic entity called Pneumocystis carinii was identified as a heterogeneous group of microscopic Fungi, constituted of multiple stenoxenic biological entities largely spread across ecosystems, closely adapted to, and coevolving in parallel with, mammal species. The discoveries and reasoning that led to the current conceptions about the taxonomy of Pneumocystis at the species level are examined here. The present review also focuses on the biological, morphological and phylogenetical features of Pneumocystis jirovecii, Pneumocystis oryctolagi, Pneumocystis murina, P. carinii and Pneumocystis wakefieldiae, the five Pneumocystis species described until now, mainly on the basis of the phylogenetic species concept. Interestingly, Pneumocystis organisms exhibit a successful adaptation enabling them to dwell and replicate in the lungs of both immunocompromised and healthy mammals, which can act as infection reservoirs. The role of healthy carriers in aerial disease transmission is nowadays recognized as a major contribution to Pneumocystis circulation, and Pneumocystis infection of nonimmunosuppressed hosts has emerged as a public health issue. More studies need to be undertaken both on the clinical consequences of the presence of Pneumocystis in healthy carriers and on the intricate Pneumocystis life cycle to better define its epidemio- logy, to adapt existing therapies to each clinical context and to discover new drug targets. Introduction The present review examines the discoveries and reasoning that led to the current conceptions about the taxonomy of Pneumocystis at the species level. Considered for a long time an enigmatic pulmonary unicellular parasite of unknown biological significance in man and other mammals, it is now well established that the ascomycetous Pneumocystis genus contains numerous host species-specific species widely spread across ecosystems. Interestingly, advances in this field have had a beneficial impact on the understanding of the natural history of Pneumocystis infection. On the basis of the identi- fication of several species in the genus and the demonstration that each Pneumocystis species can only infect its own specific host (Aliouat et al., 1993b, 1994; Furuta et al., 1993; Gigliotti et al., 1993; Atzori et al., 1999; Durand-Joly et al., 2002), new patterns of Pneumocystis pneumonia (PcP) epidemiology are emerging. These new concepts, associated with the fact that Pneumocystis organisms are actively transmitted by the air- borne route (Walzer et al., 1977; Hughes, 1982; Soulez et al., 1991; Dumoulin et al., 2000; Chab´ e et al., 2004), and perhaps also vertically (Cer´ e et al., 1997; Demanche et al., 2003; Sanchez et al., 2007; Montes-Cano et al., 2009), provide new insights to identify reservoirs and infection sources, and therefore to conceive rational, efficient and cost-effective prevention measures against PcP (Vargas et al., 2000; Miller et al., 2001; Rabodonirina, 2001; Durand-Joly et al., 2003; Calder ´ on et al., 2004). FEMS Yeast Res 11 (2011) 2–17 c 2010 Federation of European Microbiological Societies Published by Blackwell Publishing Ltd. All rights reserved YEAST RESEARCH

Transcript of Pneumocystis: from a doubtful unique entity to a group of highly diversified fungal species

M I N I R E V I E W

Pneumocystis: fromadoubtful unique entity toagroupof highlydiversi¢ed fungal speciesMagali Chabe1,2, Cecile-Marie Aliouat-Denis1,2, Laurence Delhaes1,3, El Moukhtar Aliouat1,2,Eric Viscogliosi1 & Eduardo Dei-Cas1,2,3

1Biology and Diversity of Emergent Eukaryotic Pathogens (BDEEP)–Center for Infection and Immunity of Lille, Pasteur Institute of Lille, Inserm U1019,

CNRS UMR 8204, University Lille-Nord-de-France, Lille, France; 2Department of Parasitology–Mycology, Faculty of Biological and Pharmaceutical

Sciences, University of Lille-Nord-de-France, Lille, France; and 3Department of Parasitology–Mycology, Faculty of Medicine, University of Lille-Nord-

de-France, Biology-Pathology Centre, University Hospital Center, Lille, France

Correspondence: Magali Chabe,

Department of Parasitology–Mycology,

Faculty of Biological and Pharmaceutical

Sciences, University of Lille-Nord-de-France,

3, rue du Professeur Laguesse BP83, 59006

Lille, France. Tel.: 133 3 2096 4010; fax: 133

3 2087 7276; e-mail: magali.chabe@univ-

lille2.fr

Received 11 May 2010; revised 10 October

2010; accepted 12 October 2010.

Final version published online 26 November

2010.

DOI:10.1111/j.1567-1364.2010.00698.x

Editor: Teun Boekhout

Keywords

Pneumocystis spp.; phylogenetic species

concept; life cycle; epidemiology.

Abstract

At the end of the 20th century the unique taxonomically enigmatic entity called

Pneumocystis carinii was identified as a heterogeneous group of microscopic Fungi,

constituted of multiple stenoxenic biological entities largely spread across

ecosystems, closely adapted to, and coevolving in parallel with, mammal species.

The discoveries and reasoning that led to the current conceptions about

the taxonomy of Pneumocystis at the species level are examined here. The present

review also focuses on the biological, morphological and phylogenetical features

of Pneumocystis jirovecii, Pneumocystis oryctolagi, Pneumocystis murina, P. carinii

and Pneumocystis wakefieldiae, the five Pneumocystis species described until

now, mainly on the basis of the phylogenetic species concept. Interestingly,

Pneumocystis organisms exhibit a successful adaptation enabling them to

dwell and replicate in the lungs of both immunocompromised and healthy

mammals, which can act as infection reservoirs. The role of healthy carriers in

aerial disease transmission is nowadays recognized as a major contribution to

Pneumocystis circulation, and Pneumocystis infection of nonimmunosuppressed

hosts has emerged as a public health issue. More studies need to be undertaken

both on the clinical consequences of the presence of Pneumocystis in healthy

carriers and on the intricate Pneumocystis life cycle to better define its epidemio-

logy, to adapt existing therapies to each clinical context and to discover new

drug targets.

Introduction

The present review examines the discoveries and reasoning

that led to the current conceptions about the taxonomy of

Pneumocystis at the species level. Considered for a long time

an enigmatic pulmonary unicellular parasite of unknown

biological significance in man and other mammals, it is now

well established that the ascomycetous Pneumocystis genus

contains numerous host species-specific species widely spread

across ecosystems. Interestingly, advances in this field have

had a beneficial impact on the understanding of the natural

history of Pneumocystis infection. On the basis of the identi-

fication of several species in the genus and the demonstration

that each Pneumocystis species can only infect its own specific

host (Aliouat et al., 1993b, 1994; Furuta et al., 1993; Gigliotti

et al., 1993; Atzori et al., 1999; Durand-Joly et al., 2002), new

patterns of Pneumocystis pneumonia (PcP) epidemiology are

emerging. These new concepts, associated with the fact that

Pneumocystis organisms are actively transmitted by the air-

borne route (Walzer et al., 1977; Hughes, 1982; Soulez et al.,

1991; Dumoulin et al., 2000; Chabe et al., 2004), and perhaps

also vertically (Cere et al., 1997; Demanche et al., 2003;

Sanchez et al., 2007; Montes-Cano et al., 2009), provide new

insights to identify reservoirs and infection sources, and

therefore to conceive rational, efficient and cost-effective

prevention measures against PcP (Vargas et al., 2000; Miller

et al., 2001; Rabodonirina, 2001; Durand-Joly et al., 2003;

Calderon et al., 2004).

FEMS Yeast Res 11 (2011) 2–17c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

YEA

ST R

ESEA

RC

H

The Pneumocystis genus: from Protoctiststo Fungi

Carlos Chagas (1879–1934) first discovered Pneumocystis

cystic forms in 1909 after he had been charged with the task

of malaria control during the construction of the central

railroad in Minas Gerais state, Brazil (Delaporte, 1999).

Chagas observed such forms in the lungs of guinea pigs

inoculated with the blood of two children with trypanoso-

miasis. He misidentified them as a new schizogonic state of

Trypanosoma cruzi and proposed a new trypanosome genus

named Schizotrypanum (Chagas, 1909). Chagas observed

the same cystic forms in the lungs of a human case of acute

American trypanosomiasis (Chagas, 1911; Delaporte, 1999).

In 1910, at the Pasteur Institute of Sao Paulo, Antonio

Carini (1872–1950) found similar cysts in the lungs of

Rattus norvegicus infected by Trypanosoma lewisi (Carini,

1910). He sent the slides to Alphonse Laveran, whose two

fellows, husband and wife Delanoe, observed similar pul-

monary cysts in trypanosome-free Parisian rats (R. norvegi-

cus). After inoculating Trypanosoma-free rats with the cysts

and checking that they did not lead to blood-circulating

trypanosomes, Delanoe & Delanoe concluded that pulmon-

ary cystic forms reported by Chagas and Carini indeed

constituted a new microorganism unrelated to trypano-

somes (Delanoe & Delanoe, 1912; Calderon-Sandubete

et al., 2002). Pneumocystis carinii was named and described

as a new biological entity (Delanoe & Delanoe, 1912). The

choice of the Latin name stems from the tropism of the

organism for the lungs (Pneumo-), its round shape

(-cystis) and the name of Dr Antonio Carini, who provided

the tissue samples (Delanoe & Delanoe, 1912). Chagas then

retracted his first description and ascertained the French

couple’s position by supportive data (Chagas, 1913).

At this time, P. carinii was hypothetically linked with

coccidian protozoa (Aragao, 1913; Chagas, 1913). The

controversy about the taxonomic classification of Pneumo-

cystis, especially its fungal vs. protozoan nature, began and

continued to gain momentum until the end of the 1980s.

However, since 1988 many pieces of evidence have been

brought together that indicate that Pneumocystis belongs in

the group of Fungi. For example, phylogenetic analyses of

Pneumocystis based on rRNA gene (Edman et al., 1988,

1989a; Stringer et al., 1989) as well as mitochondrial gene

sequences (Pixley et al., 1991; Sinclair et al., 1991; Wakefield

et al., 1992) showed significant homology with fungal DNA

sequences. Recently, analyses of P. carinii expressed sequence

tags (EST) and cDNA sequences in the framework of the

Pneumocystis genome project (PGP) initiative (http://pgp.

cchmc.org) confirmed the overwhelming homology of

Pneumocystis to Fungi (Cushion et al., 2007). Furthermore,

Pneumocystis genes encoding thymidylate synthase (TS) and

dihydrofolate reductase (DHFR) were found to be located

on different chromosomes, indicating that TS and DHFR

enzymatic functions reside in two distinct polypeptide

chains, unlike all protozoa studied to date (Krungkrai et al.,

1990; Anderson, 2005; Sienkiewicz et al., 2008). Finally, a

gene encoding a protein that showed marked structural

similarity with the elongation factor 3 (EF-3) of Saccha-

romyces cerevisiae and Candida albicans, was characterized

in Pneumocystis (Ypma-Wong et al., 1992). Whereas EF-1

and EF-2 are conserved among Fungi and higher eukaryotes,

EF-3 is unique to Fungi.

Today, the assignment of the Pneumocystis genus to the

group of Fungi needs no further proof, but the identity of

the closest extant relative to the Pneumocystis genus has been

the subject of debate. In fact, the Pneumocystis genus was

placed in the fungal phylum Ascomycota, subphylum Taph-

rinomycotina (Eriksson & Winka, 1997), Order Pneumocys-

tidales (Eriksson, 1994), Class Pneumocystidomycetes (sensu

Eriksson & Winka, 1997), Family Pneumocystidaceae (Eriks-

son, 1994), Genus Pneumocystis (Delanoe & Delanoe, 1912).

The Taphrinomycotina (formerly Archiascomycota) taxon

was initially created based on the rRNA gene phylogeny and

comprises highly diverse members such as the fission yeast

Schizosaccharomyces pombe, the plant pathogen Taphrina

deformans, the anamorphic yeast-like Saitoella complicata,

Protomyces sp. and Neolecta irregularis, the only member

bearing a fruiting body structure (Nishida & Sugiyama,

1993, 1994). The question whether Taphrinomycotina is a

monophyletic (James et al., 2006; Liu et al., 2006; Spatafora

et al., 2006; Sugiyama et al., 2006) or a paraphyletic

(Eriksson, 1999; Baldauf et al., 2000; Landvik et al., 2001;

Diezmann et al., 2004) group has long been debated. Indeed,

the inferred phylogenetic analyses testing this issue differ in

the sampling of genes and fungal species as well as in the

chosen method of comparison. However, a recent phyloge-

nomic analysis, combining both nuclear and mitochondrial

gene sequences, revealed the monophyly of Taphrinomyco-

tina, which was placed as a sister group of Saccharomycoti-

na1Pezizomycotina (Liu et al., 2009). This confirmed the

assignment of the Pneumocystis genus to the group of

Taphrinomycotina, with S. pombe as the closest extant

relative species (Liu et al., 2009).

Pneumocystis organisms: what are they?

Little notice was paid to Pneumocystis organisms until they

were linked to the epidemics of interstitial plasma cell

pneumonia that plagued malnourished infants and small

children living in European orphanages as a consequence of

World War II (see Calderon-Sandubete et al., 2002 for

review). As a primary cause of mortality among AIDS

patients in the 1980s, Pneumocystis and pneumocystosis

have became a focus of research attention.

FEMS Yeast Res 11 (2011) 2–17 c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

3Pneumocystis species

Lack of efficient culture systems to propagate Pneumocys-

tis organisms in vitro has hindered the understanding of the

basic biology of Pneumocystis. In addition, as culture meth-

ods to isolate current Pneumocystis jirovecii from clinical

samples are unavailable, most Pneumocystis cell biology

studies have been developed using parasites of animal

origin. Most data about Pneumocystis cell structure resulted

from research on Pneumocystis oryctolagi in the 1990s,

thanks to significant improvements in transmission electron

microscopy (TEM)-fixation methods (Goheen et al., 1989;

Palluault et al., 1992a, b), ultrastructural cytochemistry

(Yoshikawa & Yoshida, 1987; Yoshikawa et al., 1988; Pal-

luault et al., 1990, 1992a, b; Kwon et al., 1998), and compu-

ter-aided 3D reconstructions (Palluault et al., 1991a, b).

These studies revealed specific ultrastructural features of

P. oryctolagi (Dei-Cas et al., 2006), but most morphological

data on this species can be extended to other Pneumocystis

species (Dei-Cas et al., 2004) and are briefly summarized

below.

All known Pneumocystis life-cycle stages are found in the

lungs of infected hosts, although parasites may spread

exceptionally to other organs. The usually accepted life cycle

(Fig. 1) of Pneumocystis species involves an ameboid, thin-

walled, mononuclear vegetative or trophic form which

becomes a thick-walled cystic stage (ascus), in which a

multiple nuclear division leads to the formation of eight

ascospores. These forms are able to leave the cyst, presum-

ably by a pore-like zone located at the thickest part of the

cyst cell wall (Itatani, 1994), to attach specifically to type-I

epithelial alveolar cells and to evolve towards the cystic stage.

The transition from trophic forms to mature cyst or ascus

occurs in three consecutive sporocytic stages (early, inter-

mediate and late sporocytes) (Yoshida, 1989; Dei-Cas et al.,

2004).

Ascospores, which result from invaginations of the spor-

ocyte cell membrane, present a single mitochondrion, a

well-developed rough or smooth endoplasmic reticulum

and an electron-dense one-layered cell wall with a clearly

visible outer membrane (Barton & Campbell, 1967; Vavra &

Kucera, 1970; Haque et al., 1987; Palluault et al., 1992a, b).

Pneumocystis trophic forms and early sporocytes also have a

thin, monolayered, electron-dense cell wall with an outer

membrane (Palluault et al., 1992a). Mononuclear trophic

forms are irregular in size (4–8 mm long) and shape and

present numerous filopodia, filiform structures markedly

abundant and tree-like in mouse-derived Pneumocystis

(Dei-Cas et al., 1991). A single mitochondrion with budding

zones occupies a considerable part of the cell volume

(Palluault et al., 1991a, b). Early sporocytes are usually

rounded, mononuclear cells with a cell wall similar to that

Cyst

Late sporocyte

n

Intermediatesporocyte

n

Earlysporocyte

R!

2n

n + n

Trophicforms

Conjugation• ste2• ste3• mapk• ste11/ste20

Synaptonemalcomplexe

Meiosis:• mei2• ran1

Mitosis:• cdc2• cdc13• cdc25• cdc42

Spindle polebody

Fig. 1. A hypothetical life-cycle of Pneumocystis species. Parasites are represented as observed in the lung using TEM. Pleomorphic, thin-walled

mononuclear trophic forms are shown attached to type-I epithelial alveolar cells (at the top). Trophic form (2n) evolves into early sporocyte in which a

synaptonemal complex is indicated. Meiotic nuclear division (R!) leads to thick-walled sporocytic and cystic stages, in which multiple nuclear divisions in

intermediate and late sporocytes lead to the formation of eight haploid spores or ascospores (n). These forms are able to leave the cyst, to attach

specifically to type-I epithelial alveolar cells and, likely, to develop conjugation as illustrated at the top (left) (n1n), where spindle pole bodies are clearly

visible. Main molecular factors involved in conjugation, meiosis or mitosis identified in Pneumocystis carinii have been listed in the figure (see

homologies, functions and references in Table 1). Conjugation and synaptonemal complexes have been drawn according to Yoshida et al. (1984) and

Itatani (1996), respectively.

FEMS Yeast Res 11 (2011) 2–17c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

4 M. Chabe et al.

of trophic forms. Transition from early to intermediate

sporocyte is denoted (1) by an increase of the cell wall

thickness (from 40 to 100 nm), resulting mainly from the

synthesis of a b-(1,3) glucan-rich, electron-lucent middle

layer, and (2) by the occurrence of multiple nuclear divi-

sions. In the late sporocytes, the resulting eight nuclei are

visible and a plasma membrane invagination progressively

demarcates the ascospores. The mature cyst or ascus is about

4–6mm in diameter, round and thick-walled. This life stage

also presents an outer membrane (De Stefano et al., 1990;

Palluault et al., 1992a). It contains a maximum of eight

spores. Mature cysts were detected in the bronchial lumen,

suggesting that they could gain the external environment to

be transmitted to other hosts by the airborne route (Dei-

Cas, 2000). However, the infectious forms of Pneumocystis

species have not been identified (Chabe et al., 2009).

In the parasite cell wall there is a highly polymorphic

116–120-kDa mannose-rich glycoprotein (major surface

glycoprotein) that constitutes the major surface antigen of

Pneumocystis (Kovacs et al., 1993). Cholesterol constitutes

about 78% of the total sterols in Pneumocystis (Kaneshiro,

1998) and is largely present in the parasite plasma mem-

brane. Pneumocystis organisms, like rust fungi, do not

contain ergosterol (Kaneshiro, 1998).

Pneumocystis fungi: an ascomycetous lifecycle

Pneumocystis species are ascomycetous pathogens appar-

ently able to sexually reproduce (mating, meiotic division,

production of ascospores) within their hosts, a property

more frequent in pathogenic fungi of plants than animals

(Dei-Cas & Vernes, 1986; Sexton & Howlett, 2006).

On the basis of detailed ultrastructural investigations

(Matsumoto & Yoshida, 1984; Yoshida, 1989; Peters et al.,

2001; Dei-Cas et al., 2004), cell sorting (Aliouat-Denis et al.,

2009; Martinez et al., 2009) and other approaches (Yamada

et al., 1986; Wyder et al., 1994, 1998; Cornillot et al., 2002),

crucial steps in the Pneumocystis life cycle are being clarified

(Fig. 1). At least in the species P. carinii, meiosis occurs in

early sporocytes, as revealed by the detection of synaptone-

mal complexes in the nucleus of these life-cycle stages

(Matsumoto & Yoshida, 1984; Peters et al., 2001). The

nucleus of the ascospore should be haploid, and a process

of nuclear fusion may therefore occur (Itatani, 1996; Smu-

lian et al., 2001) to produce diplophasic forms able to

resume the process of cyst generation. Although haploidy

has been predicted for most stages based on karyotype and

quantitative fluorescence analyses (Yamada et al., 1986;

Wyder et al., 1994, 1998; Cushion, 1998; Stringer & Cushion,

1998), more recent data indicate that haploidy and diploidy

coexist in Pneumocystis populations (Cornillot et al., 2002;

Dei-Cas et al., 2004; Aliouat-Denis et al., 2009). Thus,

replication of Pneumocystis organisms results at least

from the generation of haploid ascospores in the cyst, and

in vitro observations have suggested that trophic forms

could result exclusively from cyst development (Aliouat

et al., 1999).

In Fungi, the mating process is usually initiated after

mutual secretion of pheromones by fungal cells of opposite

mating types. Pheromone secretion is also stimulated by

environmental stress such as nutrition deprivation (Li et al.,

2007). Pheromones recognize a heterotrimeric G-coupled

transmembrane receptor located at the cell surface of the

opposite mating type. In turn, a mitogen-activated protein

kinase (Mapkp) signal transduction cascade is activated (Li

et al., 2007). Once activated, Mapkp controls many cell

effectors that halt the mitotic cell cycle, initiate transcription

of genes involved in mating, and eventually allow the fusion

of both cells (Harigaya & Yamamoto, 2007).

Many genetic factors involved in mating, meiosis and

mitosis regulation (Table 1, Fig. 1) have been identified in

P. carinii, thanks, at least in part, to the PGP (http://pgp.

cchmc.org/) (Cushion, 2004; Cushion et al., 2007). Also,

using degenerate PCR, library screening, heterologous

expression, yeast complementation, computer modeling,

immunoprecipitation and biochemical characterization,

key Pneumocystis genes or proteins potentially involved in

mating, meiosis or mitosis have been identified (Table 1, Fig.

1). The topic was reviewed recently by Aliouat-Denis et al.

(2009). Thomas et al. (1998b) identified a gene encoding a

Mapkp in P. carinii that is homologous to other fungal

Mapkps. Thus, heterologous expression of P. carinii mapk

was shown to restore pheromone signaling in S. cerevisiae

fus3/kss1 double mutants (Vohra et al., 2003a). In addition,

P. carinii Mapkp was reported to phosphorylate the P. carinii

homologue of S. pombe ste11 (ste12 in S. cerevisiae), a gene

which encodes a transcriptional factor needed for the

pheromone-induced expression of genes required for mat-

ing (Vohra et al., 2003b).

Once conjugation has occurred, activation of ste11 indir-

ectly turns on mei2, which plays pivotal roles in both the

induction and the progression of meiosis (Harigaya &

Yamamoto, 2007). Pneumocystis carinii mei2, a homologue

of S. pombe mei2, has recently been identified using the PGP

database (Burgess et al., 2008). The same group also

identified another kinase gene, P. carinii ran1, as an S. pombe

ran1 homologue. Ran1p is known to directly phosphorylate

and inhibit the activity of Mei2p. Both P. carinii genes

exhibited functional activity in meiotic control when ex-

pressed in S. pombe.

Pneumocystis carinii ste3, an a-factor pheromone receptor

homologue, was identified from an EST database that was

created as part of the PGP (Smulian et al., 2001). This

G-protein-coupled receptor was later reported to be exclu-

sively expressed in a subpopulation of trophic forms (Vohra

FEMS Yeast Res 11 (2011) 2–17 c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

5Pneumocystis species

et al., 2004), a finding that is consistent with the expression

pattern of pheromone receptors in other fungi. So far, no

ligand has been identified for this receptor. Genes encoding

functional elements of the pheromone response signal

transduction cascade, such as ste12 and ste20 orthologues,

were found to be clustered around the Pneumocystis ste3

gene (Smulian et al., 2001). Pneumocystis carinii ste20 was

shown to be expressed following adherence of the fungus to

extracellular matrix components or lung epithelial cells

(Kottom et al., 2003). Heterologous expression of P. carinii

ste20 conferred pseudohyphal growth and also revealed the

gene to be functional in mating signaling pathways in S.

cerevisiae mutant strains (Kottom et al., 2003). Additional

orthologues of fungal genes associated with either the

mating/sexual mode of replication or stress/nutritional

deprivation were identified in the P. carinii EST database,

pointing to these conditions as triggers of Pneumocystis

mating (Cushion et al., 2007). A b-1,3-endoglucanase gene,

potentially involved in the regulation of cell wall biosynthe-

sis, was recently characterized in P. carinii (Villegas et al.,

2010). Lastly, many cell division-cycle genes have been

identified and characterized in Pneumocystis species (Tho-

mas et al., 1998a; Kaiser et al., 1999; Kottom et al., 2000;

Gustafson et al., 2001; Chabe et al., 2004; Arcenas et al.,

2006; Cushion et al., 2007; Krajicek et al., 2010).

Pneumocystis genus: from one tomultiple species

A notion, which lasted for almost a century after the

description of Pneumocystis in rats, guinea pigs, humans

and other mammals, was that Pneumocystis pneumonia

(PcP) has a zoonotic pattern. This misleading belief was

consistent with the erroneous concept of a unique Pneumo-

cystis species (i.e. P. carinii Delanoe & Delanoe, 1912).

Frenkel (1976) was the first to consider rat- and human-

derived Pneumocystis as distinct species on the basis of

apparent host species specificity, antigenic and ultrastruc-

tural differences. He proposed to name the human Pneumo-

cystis organisms ‘P. jiroveci’ (Frenkel, 1976). But it was only

in the early 1990s that the idea really emerged of a

Pneumocystis genus containing numerous highly divergent

genuine taxonomic entities. Indeed, molecular techniques

have revealed the existence of great genomic diversity

among isolates of Pneumocystis from different mammalian

species (see Aliouat-Denis et al., 2008 for review). These

Table 1. Pneumocystis life cycle: main Pneumocystis genes/proteins associated with mating, meiosis or mitosis

Gene Protein Homologue to� Function References

ste2w – Saccharomyces cerevisiae a-mating factor pheromone

receptor

Smulian et al. (2001); Cushion (2004)

ste3 Ste3p Coprinopsis cinerea a-mating factor pheromone

receptor

Smulian et al. (2001); Vohra et al. (2004)

mapk Mapkp Fusarium solani,

Schizosaccharomyces pombe SPK1,

Saccharomyces cerevisiae FUS3

Mitogen-activated protein kinase

(fungal differentiation and

proliferation)

Thomas et al. (1998b); Vohra et al.

(2003a)

ste11/ste20z Ste11p/Ste20p Schizosaccharomyces pombe

Ste11p, Cryptococcus neoformans

Ste20p

Ste11/Ste20p are involved in

pheromone response signal

transduction cascade

Smulian et al. (2001); Vohra et al.

(2003a); Kottom et al. (2003)

mei2 Mei2p Schizosaccharomyces pombe Induction and progression of

meiosis

Burgess et al. (2008)

ran1 Ran1p Pneumocystis carinii Temperature-dependent meiosis

inhibition activity

Burgess et al. (2008); Burgess et al.

(2009)

cdc2 Cdc2p Schizosaccharomyces pombe Cell-division-cycle serine–threonine

kinase (mitosis inductor)

Thomas et al. (1998a)‰

cdc13 Cdc13p Pneumocystis carinii Cdc13p binds to Cdc2 (needed for

Cdc2p kinase activity)

Kottom et al. (2000)

con7w – Magnaporthe grisea Likely involved in fungal

sporogenesis

Cushion et al. (2007)

cdc25 Cdc25p Schizosaccharomyces pombe Cell-division-cycle mitotic inducer

phosphatase

Gustafson et al. (2001)

cdc42 Cdc42p Schizophyllum commune Interacting with Ste20. Regulatory

GTPase function.

Krajicek et al. (2010)

�Only the greatest or most significant fungal homologies are mentioned.wGenes that were shown to be transcribed in Pneumocystis carinii.zAlso homologue to Saccharomyces cerevisiae Ste12p.‰cdc2 has also been described in Pneumocystis murina (Chabe et al., 2004) and in Pneumocystis jirovecii (Kaiser et al., 1999; Arcenas et al., 2006).

FEMS Yeast Res 11 (2011) 2–17c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

6 M. Chabe et al.

observations were strengthened by the demonstration of a

narrow Pneumocystis host specificity revealed by cross-

infection experiments (Aliouat et al., 1993b, 1994; Gigliotti

et al., 1993).

In 1994, at the 3rd International Workshops on Oppor-

tunistic Protists (IWOP-3, Cleveland), the use of a tempor-

ary trinomial nomenclature (i.e. ‘formae speciales’) to name

Pneumocystis organisms from different mammals, was

unanimously accepted (Anonymous, 1994). Indeed, the

nomenclature of special forms, which is governed by the

requirements of the International Code of Botanical

Nomenclature (ICBN), is used to distinguish, within a

fungal parasite species, the special forms (formae speciales)

that are exclusively characterized by their adaptation to

different hosts (ICBN, chapter 1, article no. 4, note no. 3).

For instance, the human Pneumocystis strain was named

P. carinii f.sp. hominis, the mouse strain P. carinii f.sp. muris,

the rabbit strain P. carinii f.sp. oryctolagi, etc. Since 1994,

sequence comparisons, phylogenetic analyses, population

genetics and even ultrastructural studies revealed that diver-

gence existing among the best-characterized Pneumocystis

special forms reached the level of divergence found among

fungal species (Dei-Cas et al., 1994, 2006; Keely et al., 1994,

2003, 2004; Banerji et al., 1995; Mazars et al., 1995, 1997;

Wakefield, 1998; Frenkel, 1999; Stringer et al., 2001; Cushion

et al., 2004; Redhead et al., 2006). In accordance with this,

Frenkel (1999) described again human-derived Pneumocystis

at the species level, naming it P. jirovecii, but this time in the

framework of ICBN. He took advantage of this work to

characterize rat-derived Pneumocystis according to ICBN

rules, retaining the name P. carinii (Frenkel, 1999) proposed

by Delanoe & Delanoe (1912). Afterwards, a discussion

among the Pneumocystis investigators at the IWOP-7 (Cin-

cinnati, 2001) ended with a general agreement to pursue the

process of describing and naming the Pneumocystis species

(Cushion & Beck, 2001). To date, three other Pneumocystis

species have been formerly described: Pneumocystis wake-

fieldiae Cushion et al. (2004), in R. norvegicus, Pneumocystis

murina Keely et al. (2004), in laboratory mouse (Mus

musculus), and P. oryctolagi Dei-Cas et al. (2006), in Old

World rabbits (Oryctolagus cuniculus) (Table 2).

Which species concept for Pneumocystis?

Fungal taxonomy issues are often complex, and Pneumocys-

tis taxonomy does not depart from this rule. The reasons for

this complexity include the very nature and odd character-

istics of some fungi, such as the inability to grow in vitro, but

also the heterogeneity of the species concepts. Indeed, at

least 20 defined species concepts exist, leading to many ways

of conceiving a species (Mayden, 1997; Hey, 2001). Fungal

taxonomy issues, especially criteria and strategies to describe

new species, have been comprehensively clarified (Taylor

et al., 2000). Taylor highlighted the distinction made by

Mayden between theoretical species concepts, such as the

evolutionary species concept (ESC), and operational ones,

such as the morphological species concept (MSC), the

biological species concept (BSC), and the phylogenetic

species concept (PSC) (Mayden, 1997; Taylor et al., 2000).

MSC, BSC and PSC specify criteria to recognize species and

are thus in common use, whereas the ESC cannot be directly

applied (Mayden, 1997). Therefore, Taylor proposed the

term ‘species recognition’ for the operational species con-

cepts, i.e. morphological species recognition (MSR), biolo-

gical species recognition (BSR) and phylogenetic species

recognition (PSR).

MSR is the dominant method and has permitted the

diagnosis of nearly 70 000 fungal species, which serve as a

base for future comparisons (Taylor et al., 2000). At first

sight, MSR is of little use for describing new species in the

Pneumocystis genus because Pneumocystis species show little

difference at the light microscopic level. However, host

species-related divergence, at least among some Pneumocys-

tis species, was found using TEM (Dei-Cas et al., 1994, 2004,

2006; Nielsen et al., 1998). Indeed, ultrastructural differ-

ences allowed rabbit-derived P. oryctolagi to be distinguished

from rodent-derived Pneumocystis species (Dei-Cas et al.,

1994; Nielsen et al., 1998). Pneumocystis murina organisms

show thinner and more numerous filopodia than P. orycto-

lagi (Dei-Cas et al., 2006) and the density of membrane-

limited cytoplasmic granules in Pneumocystis is different

among rats, mice and rabbits (Nielsen et al., 1998). In

addition to these morphological differences, phenotypic

divergence based on other features was reported to exist

among Pneumocystis species from diverse mammals. These

phenotypic differences, such as cystic-to-trophic form ratio,

location in the pulmonary alveolus, in vivo doubling time

and pathology, are detailed in Table 2. Furthermore, cross-

infection experiments, performed by several groups, in

SCID mice or Nude rats with inocula of parasites isolated

from rats, mice, monkeys, ferrets, rabbits or immuno-

suppressed patients, revealed that Pneumocystis-free labora-

tory animals were refractory to infection elicited by Pneu-

mocystis organisms derived from a heterologous host species

(Aliouat et al., 1993b, 1994; Furuta et al., 1993; Gigliotti

et al., 1993; Atzori et al., 1999; Durand-Joly et al., 2002).

Thus, Pneumocystis species show an outstanding selective

infectivity, demonstrating narrow host-species specificity.

Further differences were reported in the Pneumocystis in

vitro behavior (Aliouat et al., 1993a). For instance, rat-

derived Pneumocystis seems to have a higher capacity for

attaching to target cells in vitro than mouse-derived patho-

gens. Moreover, in vitro attachment of rat Pneumocystis

seems to be more sensitive to pentamidine or cytochalasin-

B than attachment of mouse-derived organisms (Aliouat

et al., 1993a).

FEMS Yeast Res 11 (2011) 2–17 c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

7Pneumocystis species

Table 2. Genotypic and phenotypic signatures differentiating Pneumocystis species (modified from Aliouat-Denis et al., 2008)

Features Pneumocystis carinii

Pneumocystis

wakefieldiae Pneumocystis murina

Pneumocystis

oryctolagi Pneumocystis jirovecii

Specific host Rat (Rattus norvegicus) Rat (Rattus norvegicus) Mouse (Mus musculus)

Rabbit (Oryctolagus

cuniculus) Man (Homo sapiens)Genes�

mtLSUrRNA U20169 (Cushion,

1998)

U20173 (Cushion,

1998)

AF257179 (S. Ruan

et al., unpublished

data)

S42915 (Wakefield

et al., 1992)

S42926 (Wakefield

et al., 1992)

mtSSUrRNA Hunter & Wakefield

(1996)

Hunter & Wakefield

(1996)

Hunter & Wakefield

(1996)

Hunter & Wakefield

(1996)

nucrRNA M86760 (Edman et al.,

1988)

L27658 (Ortiz-Rivera

et al., 1995; Cushion

et al., 1993)

AY532651 (Keely

et al., 2004)

DQ010098 (L. Li et al.,

unpublished data)

Liu et al. (1992)

ITS Ortiz-Rivera et al.

(1995)

L27658 (Ortiz-Rivera

et al., 1995)

AY532651 (Keely

et al., 2004)

DQ010098 (L. Li et al.,

unpublished data)

Ortiz-Rivera et al.

(1995); Lu et al. (1994)

TS M25415 (Edman et al.,

1989b)

Keely et al. (1994) Mazars et al. (1995) Mazars et al. (1995) Mazars et al. (1995)

HSP70 U80967 (Stedman

et al., 1998)

U80969 (Stedman

et al., 1998)

AY353254 (C.R.

Icenhour et al.,

unpublished data)

DQ435616 (Dei-Cas

et al., 2006)

U80970 (Stedman

et al., 1998)

Electrophoretic

karyotypes (number

and size range of

bands)

12–16 bands

(308–680 kb) Cushion

et al. (2004)

14 bands

(308–660 kb) Cushion

et al. (2004)

17 bands

(309–634 kb) Keely

et al. (2004)

14 bands

(300–700 kb) Cho

et al. (1999)

13 bands

(370–810 kb) Stringer

& Cushion (1998)

Size of

Pneumocystis

genome

8.2 Mb (Keely et al.,

2004)

7.7 Mb (Keely et al.,

2004)

8.2 Mb (Keely et al.,

2004)

7.7 Mb (Stringer &

Cushion, 1998)

Main phenotypic signatures (from Dei-Cas et al., 2006 unless otherwise indicated)

Organisms in lung

dry smears (TBO or

Giemsa stains)

Closely clustered Closely clustered Clustered Detached from each

other

Closely clustered

Cystic-to-trophic

form ratio

0.02–0.05 ? 0.02–0.05 0.10–0.15 ?

Size of cystic forms 4.03–4.42mm

(Laakkonen & Sukura,

1997)

5–8mm (Cushion

et al., 2004)

5–8mm (Keely et al.,

2004)

4–6 mm (Dei-Cas et al.,

2006)

3.5–5mm (Frenkel,

1976)

Filopodia thickness

and richness

Smaller than those of

rabbit

? Thinner and more

numerous than those

of rabbit, human and

macaque

Thicker and less

abundant than those

of mice

Thicker and less

abundant than those

of mice

Location Filling alveolar lumen Filling alveolar lumen Filling alveolar lumen Lining alveolar

epithelium

Filling alveolar lumen

(AIDS) or lining

alveolar epithelium

(epidemic or infantile

PcP)

In vivo doubling

time

4.5 days (in corticoid-

treated rats) (Aliouat

et al., 1999)

? 10.5 days (SCID mice)

(Aliouat et al., 1999)

1.7 days (untreated

rabbits) (Aliouat et al.,

1999)

?

Intra-alveolar

eosinophilic

honeycomb material

Present ? Present Rare Present

Fibrosis Frequent ? Frequent Rare Frequent

�Gene sequences available from at least two different Pneumocystis species. Accession numbers of the Pneumocystis DNA sequences and/or references

are given when they are available.

mtLSUrRNA, large subunit of mitochondrial rRNA; mtSSUrRNA, small subunit of mitochondrial rRNA; nucrRNA: rRNA from nuclear genome; ITS,

internal transcribed spacer from nuclear rRNA gene locus; TS, thymidylate synthase; HSP70, heat shock protein 70; ?, unknown.

FEMS Yeast Res 11 (2011) 2–17c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

8 M. Chabe et al.

Concepts of biological species (‘groups of actually or

potentially interbreeding natural populations, which are

reproductively isolated from other such groups’; Mayr,

1963) play a prominent role in describing new fungal species

(Taylor et al., 2000). In Fungi, BSR has been used to identify

groups of mating compatible individuals, thus equated

with species. To some degree, BSC could be applied to

identify new Pneumocystis species. Indeed, sexual forms of

Pneumocystis have not been identified and sustained in vitro

cultures are not yet feasible, thus preventing laboratory

mating experiments from being set up. To answer the

question of whether gene flow occurs between Pneumocystis

subpopulations, population genetic approaches were uti-

lized (Mazars et al., 1997). Actually, multilocus enzyme

electrophoresis (MEE) was used to analyze five Pneumocystis

enzyme systems: malate dehydrogenase (MDH), glucose

phosphate isomerase (GPI), leucine aminopeptidase

(LAP), malic enzyme (ME) and 6-phosphogluconate

dehydrogenase (6PGDH) in Pneumocystis isolates from

rabbits, mice and rats (22 weaning rabbits, 30 corticoste-

roid-treated rats and 17 corticosteroid-treated mice) (Ma-

zars et al., 1997). Linkage disequilibrium analysis of allele

segregation and recombination of genotypes occurring

at these five loci clearly showed that Pneumocystis strains

from different host species constitute genetically distinct

populations, isolated from each other for a very long time

and therefore representing truly diverse lineages (Mazars

et al., 1997).

Today, the PSC (evolutionary lineages with a unique

combination of DNA orthologue sequences) is of growing

importance in fungal taxonomy (Taylor et al., 2000). For

Taylor, PSR comes closer than MSR and BSR to recognizing

species consistent with the ESC because, once progeny

evolutionary species have formed from an ancestor, changes

in gene sequences occur and can be recognized before

changes have occurred in mating behavior or morphology

(Taylor et al., 2000). Moreover, the genealogical concor-

dance phylogenetic species recognition method (Taylor

et al., 2000), which assesses the relationships among several

phylogenetic trees constructed on the basis of gene

sequences, increases the reliability of PSC-based species deter-

mination, making it possible to detect the occurrence of

genetic exchange (i.e. a concordance of gene trees is indica-

tive of genetic isolation) (Taylor et al., 2000). Beyond doubt,

PSC is the main species concept for use in the description of

new Pneumocystis species. Thanks to the abundance of

Pneumocystis gene sequence data, the genealogical concor-

dance condition, required to describe new fungal species

(Taylor et al., 2000), has been fulfilled by all Pneumocystis

species described until now, and all have emerged as mono-

phyletic clades on the basis of several genes (Fig. 2; Dei-Cas

et al., 2006). In agreement with this, the topological con-

cordance among seven Pneumocystis gene trees suggests that

the Pneumocystis organisms found in different host species

do not mate (Keely & Stringer, 2005). Genetic variation at

rRNA-encoding and DHFR loci between Pneumocystis spe-

cies were also used to estimate the times of Pneumocystis

speciation (Keely et al., 2003, 2004). According to these

estimations, P. carinii and P. jirovecii diverged early from

each other, about 100 million years ago (Keely et al., 2003).

Although this kind of estimation has to be viewed with

caution, the chronology matches well the divergence time of

primates from rodents (Nei et al., 2001). The same approach

applied to Pneumocystis from rodent subfamily Murinae

(P. carinii and P. murina) suggested they diverged from each

other 30–40 million years ago (Keely et al., 2004), which is

Fig. 2. Maximum-likelihood phylogeny of 12

Pneumocystis taxa, including the five described

Pneumocystis species, inferred from

mitochondrial large subunit ribosomal RNA

(mtLSUrRNA) and mitochondrial small-subunit

rRNA (mtSSUrRNA) concatenated gene

sequences, carried out using MRBAYES v3.1.2

(Ronquist & Huelsenbeck, 2003). Bayesian

posterior probabilities, calculated using a

Markov chain Monte Carlo sampling approach

(Green, 1995) implemented in MRBAYES v3.1.2,

are given as percentages near the individual

nodes. Nodes with values of o 50% are not

shown. Scale bar: 0.1 substitutions (corrected)

per base pair (from Aliouat-Denis et al., 2008).

FEMS Yeast Res 11 (2011) 2–17 c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

9Pneumocystis species

consistent with the evolutionary time frame of their hosts

(Nei et al., 2001).

To date, the five accepted Pneumocystis species (P. carinii

Frenkel, 1999; P. jirovecii Frenkel, 1999; P. wakefieldiae

Cushion et al., 2004; P. murina Keely et al., 2004; P. oryctolagi

Dei-Cas et al., 2006) have all been described in the frame-

work of ICBN, on the basis of the morphological and/or

biological (Mayr, 1963) and phylogenetic (Taylor et al.,

2000) species concepts (Dei-Cas et al., 2006) examined

above.

Practical and theoretical implications ofnew notions on Pneumocystis taxonomy

The discovery of the genetic diversity of Pneumocystis genus

was likely the most important achievement in Pneumocystis

research for the last 20 years. Research on this topic revealed

that the genus contains numerous species with diverse

biological properties, each one being stenoxenous, i.e.

strictly infectious to one mammal species. Relatively large

molecular studies have been developed to explore the

presence and the circulation of Pneumocystis species in

ecosystems. Thus, Pneumocystis isolates from domestic,

synanthropic or wild mammals could be identified on the

basis of new species-specific DNA sequences that hitherto

confirmed the narrow host range of Pneumocystis species

(see Aliouat-Denis et al., 2008 for review). The results of

these researches have radically changed our conceptions

about Pneumocystis development, infection sources and

reservoir.

It was considered until recently that Pneumocystis organ-

isms were able to multiply exclusively in deeply immunode-

pressed hosts. Recent research has drastically changed

this view. Sensitive molecular methods allowed the detection

of Pneumocystis organisms in respiratory samples from

healthy or hospitalized subjects without severe immuno-

depression (Calderon et al., 1996, 2004; Dei-Cas, 2000;

Peterson & Cushion, 2005). In addition, experiments

focused on the behavior of Pneumocystis organisms in

immunocompetent hosts, demonstrated that Pneumocystis

cells are able to multiply in their lungs (Chabe et al., 2004).

It was also long believed that infection could be con-

tracted by the airborne route from animal or hypothetical

environmental sources, where the unique species of the

genus (P. carinii) able to cause PcP in both humans and

animals developed saprophytic growth. We know now

that the Pneumocystis genus includes numerous species,

each one restricted to one mammal species. Therefore,

animals cannot constitute a source of Pneumocystis infection

in humans (Gigliotti et al., 1993; Durand-Joly et al., 2002).

The idea of a zoonotic pattern to Pneumocystis infection,

although still retained by some authors (Stewart et al., 2005;

Youn, 2009), is no longer valid, as only human beings

could represent a potential infection source for other human

beings.

Which humans play the role of Pneumocystis infection

source? It might be considered that only hosts with PcP

transmit the infection. However, it was shown that immu-

nocompetent hosts with subclinical Pneumocystis infection

can transmit the infection by the airborne route to both

susceptible and healthy hosts (Chabe et al., 2004). Several

pieces of evidence suggest that the real impact of Pneumo-

cystis infection in humans or other mammals is beyond PcP.

For instance, seroconversion revealed Pneumocystis primary

infection in 4 90% of healthy children worldwide (re-

viewed in Aliouat-Denis et al., 2008), a condition that could

be associated with benign respiratory symptoms (Stagno

et al., 1980; Totet et al., 2004; Larsen et al., 2007). Immuno-

competent adults are often found to be Pneumocystis carriers

( = subjects without PcP but harboring very low Pneumocys-

tis rates), and Pneumocystis carriage was shown to be more

frequent in subjects with chronic respiratory conditions, e.g.

chronic obstructive pulmonary disease (COPD) (Calderon

et al., 1996, 2007), which is a major cause of disability and

the fourth leading cause of death in the world (Tan & Ng,

2008). Pneumocystis infection of immunocompetent hosts is

emerging as a relevant issue for human as well as animal

health (Rajagopalan-Levasseur et al., 1998; Cavallini-

Sanches et al., 2007) and, interestingly, typical PcP seems to

be a rare event in the natural history of Pneumocystis

infection.

What about the environment as a potential Pneumocystis

infection source? It cannot be formally excluded that (1)

Pneumocystis growth could be supported by unknown

inanimate, abiotic or vegetal substrates, or (2) some Pneu-

mocystis life-cycle stages could retain their infectious power

for some time after being released into the environment

from infected hosts. Hypotheses (1) and (2) were strength-

ened by the detection of Pneumocystis DNA in air- and

water-filtrate samples (Wakefield, 1996; Casanova-Cardiel &

Leibowitz, 1997). Also, Pneumocystis DNA was successfully

amplified from the air at conventional animal facilities

where Pneumocystis-infected laboratory animals were

housed (Olsson et al., 1996; Latouche et al., 1997) and from

hospital wards housing PcP patients (Bartlett et al., 1994;

Olsson et al., 1998). Against hypothesis (1) is that no

evidence of Pneumocystis environmental growth has been

reported so far. However, some Pneumocystis forms could

retain their viability after being in the air, as P. jirovecii

mRNA could be amplified from hospital air samples

(Latouche et al., 2001; Maher et al., 2001). In agreement with

this, there is some evidence of the existence of a Pneumocystis

‘dormant’ life-cycle stage (Kaneshiro & Maiorano, 1996; Chin

et al., 1999). Alternatively, the fact that Pneumocystis DNA or

even mRNA can be detected in the environment of Pneumo-

cystis-infected animals or humans (Bartlett et al., 1997;

FEMS Yeast Res 11 (2011) 2–17c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

10 M. Chabe et al.

Olsson et al., 1998; Sing et al., 1999; Latouche et al., 2001;

Maher et al., 2001) suggests that Pneumocystis cells are

released with the exhaled air from infected hosts. This

observation supports the existence of an active host-to-host

transmission rather than the existence of hypothetical Pneu-

mocystis life stages able either to grow out of the hosts or to

remain viable in the environment for extended periods.

As Pneumocystis organisms are able to replicate in the

lungs of immunocompetent hosts, healthy members of

mammal populations, as well as severely infected susceptible

hosts, could play a major role as a Pneumocystis reservoir

(Dei-Cas, 2000; Chabe et al., 2004; Cushion et al., 2007;

Aliouat-Denis et al., 2008). The following features of Pneu-

mocystis organisms support this hypothesis: (1) a long

genetic isolation and coevolution with their host species

entailing Pneumocystis speciation and narrow host-species

specificity (Banerji et al., 1995; Mazars et al., 1995, 1997;

Demanche et al., 2001; Guillot et al., 2001; Hugot et al.,

2003; Keely et al., 2003, 2004; Dei-Cas et al., 2006; Aliouat-

Denis et al., 2008); (2) a highly compatible relationship with

mammalian host physiology (Dei-Cas et al., 1991, 2004;

Settnes & Nielsen, 1991; Aliouat et al., 1993a; Beck et al.,

1998; Cushion et al., 2007); and (3) a low pathogenic power

in healthy hosts (Chabe et al., 2004). In addition, Pneumo-

cystis colonized hosts are usually much more numerous than

hosts with PcP (Aliouat-Denis et al., 2008).

On the whole, molecular taxonomy research has played a

crucial role in the shift from an old conceptual Pneumocystis

infection framework to a new one. Main issues involve

taxonomy, host range and terminology. In the old concep-

tual framework, Pneumocystis organisms were considered

undefined or even enigmatic protists belonging to a unique

euryxenic taxonomic entity (P. carinii) transmissible by the

airborne route between mammals of different species. In this

old framework, terminology used to name Pneumocystis life-

cycle stages was taken from protistology (trophozoites,

precysts, cysts, intracystic bodies). In the new conceptual

framework, the Pneumocystis genus is a highly diversified

group of parasitic microfungi that contains numerous

stenoxenic species closely adapted to, and coevolving with,

mammal species. Accordingly, in this new framework, the

terminology used to name Pneumocystis life-cycle stages was

taken from mycology (trophic forms, sporocytes, ascus,

ascospores).

Conclusion

Hitherto, the identification of Pneumocystis species

(P. jirovecii excepted) was performed on parasites harvested

from immunosuppressed laboratory rodents or nonimmuno-

compromised laboratory, domestic, meat or wild Old World

rabbits. We need to explore Pneumocystis species diversity in

natural ecosystems to avoid bias derived from conventional

breeding. In fact, besides the potential impact of breeding

conditions on the Pneumocystis populations (e.g. housing in

close host-to-host proximity and isolation from the natural

environment), it could also be speculated that laboratory

animals come from a limited number of standard colonies

infected originally by a small number of Pneumocystis

strains. Furthermore, the characterization of Pneumocystis

species living in the lungs of wild animals remains a crucial

way forward to understand how these pathogens dissemi-

nate across ecosystems. Thanks to long adaptation of

Pneumocystis to its mammalian host, the topology of

Pneumocystis phylogenetic trees could shed light on the

mammalian host evolution history, helping to solve taxo-

nomic uncertainties. Although MSR and BSR have signifi-

cantly contributed to new species descriptions in the

Pneumocystis genus, the PSR seems to be a promising tool

to describe new Pneumocystis species, which are usually

represented in wild mammals by low parasite rates (Mazars

et al., 1997; Laakkonen, 1998; Palmer et al., 2000; Aliouat-

Denis et al., 2008; Chabe et al., 2010).

As a result of advancing knowledge about the natural

history of Pneumocystis infection, the perception of its

clinical impact on public health is evolving. The role of

healthy carriers in airborne disease transmission is nowa-

days recognized as a major contribution to Pneumocystis

circulation (Chabe et al., 2004). Recent data indicate that the

transplacental route is another human-to-human mode of

transmission (Montes-Cano et al., 2009). Finally, the low

parasite burden usually found in Pneumocystis carriers could

be the cause of benign respiratory symptoms in healthy

small children (Stagno et al., 1980; Totet et al., 2004; Larsen

et al., 2007) or worsening of the symptoms in COPD

patients (Calderon et al., 2007). More studies need to be

undertaken both on the clinical consequences of the pre-

sence of Pneumocystis in chronic pulmonary contexts and on

the intricate Pneumocystis life cycle to better define preven-

tion measures, adapt existing therapies to each clinical

context and discover new drug targets.

Acknowledgements

This work was supported by the ANR-ERA-NET ‘Pneumo-

cystis’ PathoGenoMics Program (ANR-06-PATHO-009-01),

by ANR ‘Biodiversite’ Program (CERoPath network, ANR-

07-BDIV-012) and by the French Ministry of Research

(EA3609-University Lille-Nord-de-France & Lille Pasteur

Institute).

References

Aliouat EM, Dei-Cas E, Ouaissi A, Palluault F, Soulez B & Camus

D (1993a) In vitro attachment of Pneumocystis carinii from

mouse and rat origin. Biol Cell 77: 209–217.

FEMS Yeast Res 11 (2011) 2–17 c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

11Pneumocystis species

Aliouat EM, Mazars E, Dei-Cas E, Cesbron J & Camus D (1993b)

Intranasal inoculation of mouse, rat or rabbit-derived

Pneumocystis in SCID mice. J Protozool Res 3: 94–98.

Aliouat EM, Mazars E, Dei-Cas E, Delcourt P, Billaut P & Camus

D (1994) Pneumocystis cross infection experiments using SCID

mice and Nude rats as recipient host, showed strong host-

species specificity. J Eukaryot Microbiol 41: S71.

Aliouat EM, Dujardin L, Martinez A, Duriez T, Ricard I & Dei-

Cas E (1999) Pneumocystis carinii growth kinetics in culture

systems and in hosts: involvement of each life cycle parasite

stage. J Eukaryot Microbiol 46: S116–S117.

Aliouat-Denis CM, Chabe M, Demanche C, Aliouat EM,

Viscogliosi E, Guillot J, Delhaes L & Dei-Cas E (2008)

Pneumocystis species, co-evolution and pathogenic power.

Infect Genet Evol 8: 708–726.

Aliouat-Denis CM, Martinez A, Aliouat EM, Pottier M, Gantois

N & Dei-Cas E (2009) The Pneumocystis life cycle. Mem I

Oswaldo Cruz 104: 419–426.

Anderson AC (2005) Targeting DHFR in parasitic protozoa. Drug

Discov Today 10: 121–128.

Anonymous (1994) Revised nomenclature for Pneumocystis

carinii. The Pneumocystis Workshop. J Eukaryot Microbiol 41:

S121–S122.

Aragao HB (1913) Nota sobre as schizogonias e gametogonidas

dos trypanosomas. Bras Med 11: 271.

Arcenas RC, Uhl JR, Buckwalter SP, Limper AH, Crino D, Roberts

GD & Wengenack NL (2006) A real-time polymerase chain

reaction assay for detection of Pneumocystis from

bronchoalveolar lavage fluid. Diagn Microbiol Infect Dis 54:

169–175.

Atzori C, Agostoni F, Angeli E, Mainini A, Micheli V & Cargnel A

(1999) P. carinii host specificity: attempt of cross infections

with human derived strains in rats. J Eukaryot Microbiol 46:

S112.

Baldauf SL, Roger AJ, Wenk-Siefert I & Doolittle WF (2000) A

kingdom-level phylogeny of eukaryotes based on combined

protein data. Science 290: 972–977.

Banerji S, Lugli EB, Miller RF & Wakefield AE (1995) Analysis of

genetic diversity at the arom locus in isolates of Pneumocystis

carinii. J Eukaryot Microbiol 42: 675–679.

Bartlett MS, Lee CH, Lu JJ, Bauer NL, Bettz JF, McLaughlin GL &

Smith JW (1994) Pneumocystis carinii detected in air.

J Eukaryot Microbiol 41: S75.

Bartlett MS, Vermund SH, Jacobs R et al. (1997) Detection of

Pneumocystis carinii DNA in air samples: likely environmental

risk to susceptible persons. J Clin Microbiol 35: 2511–2513.

Barton EG & Campbell WG (1967) Further observations on the

ultrastructure of Pneumocystis. Arch Pathol 83: 527–534.

Beck JM, Preston AM, Wagner JG, Wilcoxen SE, Hossler P,

Meshnick SR & Paine R III (1998) Interaction of rat

Pneumocystis carinii and rat alveolar epithelial cells in vitro. Am

J Physiol 275: L118–L125.

Burgess JW, Kottom TJ & Limper AH (2008) Pneumocystis carinii

exhibits a conserved meiotic control pathway. Infect Immun

76: 417–425.

Burgess JW, Kottom TJ, Villegas LR, Lamont JD, Baden EM,

Ramirez-Alvarado M & Limper AH (2009) The Pneumocystis

meiotic PCRan1p kinase exhibits unique temperature-

regulated activity. Am J Resp Cell Mol 41: 714–721.

Calderon EJ, Regordan C, Medrano FJ, Ollero M & Varela JM

(1996) Pneumocystis carinii infection in patients with chronic

bronchial disease. Lancet 347: 977.

Calderon EJ, Varela JM, Medrano FJ et al. (2004) Epidemiology of

Pneumocystis carinii pneumonia in southern Spain. Clin

Microbiol Infect 10: 673–676.

Calderon EJ, Rivero L, Respaldiza N, Morilla R, Montes-Cano

MA, Friaza V, Munoz-Lobato F, Varela JM, Medrano FJ & De

La Horra C (2007) Systemic inflammation in patients with

chronic obstructive pulmonary disease who are colonized with

Pneumocystis jirovecii. Clin Infect Dis 45: 17–19.

Calderon-Sandubete EJ, Varela-Aguilar JM, Medrano-Ortega FJ,

Nieto-Guerrer V, Respaldiza-Salas N, de la Horra-Padilla C &

Dei-Cas E (2002) Historical perspective on Pneumocystis

carinii infection. Protist 153: 303–310.

Carini A (1910) Formas de eschizogonia do Trypanosoma lewisi.

Bol Soc de Med e Cir de Sao Paulo 18: 204.

Casanova-Cardiel L & Leibowitz MJ (1997) Presence of

Pneumocystis carinii DNA in pond water. J Eukaryot Microbiol

44: S28.

Cavallini-Sanches EM, Pescador C, Rozza D, Spanamberg A,

Borba MR, Ravazzolo AP, Driemeier D, Guillot J & Ferreiro L

(2007) Detection of Pneumocystis spp. in lung samples from

pigs in Brazil. Med Mycol 45: 395–399.

Cere N, Drouet-Viard F, Dei-Cas E, Chanteloup N & Coudert P

(1997) In utero transmission of Pneumocystis carinii sp. f.

oryctolagi. Parasite 4: 325–330.

Chabe M, Dei-Cas E, Creusy C, Fleurisse L, Respaldiza N, Camus

D & Durand-Joly I (2004) Immunocompetent hosts as a

reservoir of Pneumocystis organisms: histological and RT-PCR

data demonstrate active replication. Eur J Clin Microbiol 23:

89–97.

Chabe M, Nevez G, Totet A, Frealle E, Delhaes L, Aliouat EM &

Dei-Cas E (2009) Transmission de Pneumocystis. J Mycol Med

19: 276–284.

Chabe M, Herbreteau V, Hugot JP, Bouzard N, Deruyter L,

Morand S & Dei-Cas E (2010) Pneumocystis carinii and

Pneumocystis wakefieldiae in wild Rattus norvegicus trapped in

Thailand. J Eukaryot Microbiol 57: 213–217.

Chagas C (1909) Nova tripanozomiazaea humana. Mem I

Oswaldo Cruz 1: 159–218.

Chagas C (1911) Nova entidade morbida do homen: regumo

general de estudos etiologicos e clinicos. Mem I Oswaldo Cruz

3: 219–275.

Chagas C (1913) Revisao do ciclo evolutivo do Trypanosoma

cruzi. Braz Med 11: 225.

Chin K, Luttrell TD, Roe JD, Shadzi S, Wyder MA & Kaneshiro ES

(1999) Putative Pneumocystis dormant forms outside the

mammalian host, and long-term culture derived from them:

initial characterizations. J Eukaryot Microbiol 46: S95–S99.

FEMS Yeast Res 11 (2011) 2–17c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

12 M. Chabe et al.

Cho SR, Park YG, Moon HN, Lee SH & Hong ST (1999)

Karyotypes of Pneumocystis carinii derived from several

mammals. Korean J Parasitol 37: 271–275.

Cornillot E, Keller B, Cushion MT, Metenier G & Vivares CP

(2002) Fine analysis of the Pneumocystis carinii f. sp. carinii

genome by two-dimensional pulsed-field gel electrophoresis.

Gene 293: 87–95.

Cushion M, Keely S & Stringer J (2004) Molecular and

phenotypic description of Pneumocystis wakefieldiae sp.nov., a

new species in rats. Mycologia 96: 429–438.

Cushion MT (1998) Genetic heterogeneity of rat-derived

Pneumocystis. FEMS Immunol Med Mic 22: 51–58.

Cushion MT (2004) Pneumocystis: unraveling the cloak of

obscurity. Trends Microbiol 12: 243–249.

Cushion MT & Beck JM (2001) Summary of Pneumocystis

research presented at the 7th International Workshop on

Opportunistic Protists. J Eukaryot Microbiol 48: S101–S105.

Cushion MT, Zhang J, Kaselis M, Giuntoli D, Stringer SL &

Stringer JR (1993) Evidence for two genetic variants of

Pneumocystis carinii coinfecting laboratory rats. J Clin

Microbiol 31: 1217–1223.

Cushion MT, Smulian AG, Slaven BE, Sesterhenn T, Arnold J,

Staben C, Porollo A, Adamczak R & Meller J (2007)

Transcriptome of Pneumocystis carinii during fulminate

infection: carbohydrate metabolism and the concept of a

compatible parasite. PLoS ONE 2: e423.

Dei-Cas E (2000) Pneumocystis infections: the iceberg? Med Mycol

1: 23–32.

Dei-Cas E & Vernes A (1986) Parasitic adaptation of pathogenic

fungi to the mammalian hosts. CRC Crit Rev Microbiol 13:

173–218.

Dei-Cas E, Jackson H, Palluault F, Aliouat EM, Hancock V, Soulez

B & Camus D (1991) Ultrastructural observations on the

attachment of Pneumocystis carinii in vitro. J Protozool 38:

S205–S207.

Dei-Cas E, Mazars E, Odberg-Ferragut C et al. (1994)

Ultrastructural, genomic, isoenzymatic and biological features

make it possible to distinguish rabbit Pneumocystis from other

mammal Pneumocystis strains. J Eukaryot Microbiol 41: S84.

Dei-Cas E, Aliouat EM & Cailliez JC (2004) Cellular structure.

Pneumocystis Pneumonia, 3rd edn (Walzer PD & Cushion MT,

eds), pp. 61–94. Marcel Dekker, New York.

Dei-Cas E, Chabe M, Moukhlis R et al. (2006) Pneumocystis

oryctolagi sp. nov., an uncultured fungus causing pneumonia

in rabbits at weaning: review of current knowledge, and

description of a new taxon on genotypic, phylogenetic and

phenotypic bases. FEMS Microbiol Rev 80: 853–871.

Delanoe P & Delanoe E (1912) Sur les rapports des kystes de

carinii du poumon des rats avec le Trypanosoma lewisi. CR

Acad Sci 155: 658–660.

Delaporte F (1999) La Maladie de Chagas. Histoire d’un Fleau

Continental. Payot & Rivages, Paris.

Demanche C, Berthelemy M, Petit T, Polack B, Wakefield AE,

Dei-Cas E & Guillot J (2001) Phylogeny of Pneumocystis carinii

from 18 primate species confirms host specificity and suggests

coevolution. J Clin Microbiol 39: 2126–2133.

Demanche C, Petit T, Moisson P, Ollivet F, Rigoulet J, Chermette

R, Dei-Cas E, Wakefield AE & Guillot J (2003) Assessment of

Pneumocystis species carriage in captive primates. Vet Rec 28:

811–813.

De Stefano JA, Cushion MT, Sleught R & Walzer PD (1990)

Analysis of Pneumocystis carinii cyst wall. I. Evidence for an

outer surface membrane. J Protozool 37: 428–435.

Diezmann S, Cox CJ, Sconian G, Vilgalys RJ & Mitchell TG

(2004) Phylogeny and evolution of medical species of Candida

and related taxa: a multigenic analysis. J Clin Microbiol 42:

5624–5635.

Dumoulin A, Mazars E, Seguy N, Gargallo-Viola D, Vargas S,

Cailliez JC, Aliouat EM, Wakefield AE & Dei-Cas E (2000)

Transmission of Pneumocystis carinii disease from

immunocompetent contacts of infected hosts to susceptible

hosts. Eur J Clin Microbiol 19: 671–678.

Durand-Joly I, Aliouat EM, Recourt C, Guyot K, Francois N,

Wauquier M, Camus D & Dei-Cas E (2002) Pneumocystis

carinii f. sp. hominis is not infectious for SCID mice. J Clin

Microbiol 40: 1862–1865.

Durand-Joly I, Soula F, Chabe M, Dalle JH, Lafitte JJ, Senechal M,

Pinon A, Camus D & Dei-Cas E (2003) Long-term

colonization with Pneumocystis jirovecii in hospital staffs: a

challenge to prevent nosocomial pneumocystosis. J Eukaryot

Microbiol 50: S614–S615.

Edman JC, Kovacs JA, Masur H, Santi DV, Elwood HJ & Sogin

ML (1988) Ribosomal RNA sequence shows Pneumocystis

carinii to be a member of the fungi. Nature 334: 519–522.

Edman JC, Kovacs JA, Masur H, Santi DV, Elwood HJ & Sogin

ML (1989a) Ribosomal RNA sequence shows Pneumocystis

carinii to be a member of the fungi. J Protozool 36: S18–S20.

Edman U, Edman JC, Lundgren B & Santi DV (1989b) Isolation

and expression of the Pneumocystis carinii thymidylate

synthase gene. P Natl Acad Sci USA 86: 6503–6507.

Eriksson OE (1994) Pneumocystis carinii, a parasite in lungs of

mammals, referred to a new family and order

(Pneumocystidaceae, Pneumocystidales, Ascomycota). Systema

Ascomycetum 13: 165–180.

Eriksson OE (1999) Outline of Ascomycota – 1999. Myconet 3:

1–88.

Eriksson OE & Winka K (1997) Supraordinal taxa of the

Ascomycota. Myconet 1: 1–16.

Frenkel JK (1976) Pneumocystis jiroveci n. sp. from man:

morphology, physiology, and immunology in relation to

pathology. Natl Cancer I Monogr 43: 13–30.

Frenkel JK (1999) Pneumocystis pneumonia, an

immunodeficiency-dependent disease (IDD): a critical

historical overview. J Eukaryot Microbiol 46: S89–S92.

Furuta T, Fujita M, Mukai R, Sakakibara I, Sata T, Miki K,

Hayami M, Kojima S & Yoshikawa Y (1993) Severe pulmonary

pneumocystosis in simian acquired immunodeficiency

syndrome induced by simian immunodeficiency virus: its

characterization by the polymerase-chain-reaction method

FEMS Yeast Res 11 (2011) 2–17 c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

13Pneumocystis species

and failure of experimental transmission to immunodeficient

animals. Parasitol Res 79: 624–628.

Gigliotti F, Harmsen AG, Haidaris CG & Haidaris PJ (1993)

Pneumocystis carinii is not universally transmissible between

mammalian species. Infect Immun 61: 2886–2890.

Goheen M, Blumershine R, Hull M, Bartlett M & Smith J (1989)

Enhancement of the fine structure of Pneumocystis carinii

using potassium ferrocyanide and tannic acid. Proceedings of

the 47th Annual Meeting of the Electron Microscopy Society of

America (Bailey GW, ed), pp. 984–985. San Francisco.

Green PJ (1995) Reversible jump Markov chain Monte Carlo

computation and Bayesian model determination. Biometrika

82: 711–732.

Guillot J, Demanche C, Hugot JP, Berthelemy M, Wakefield AE,

Dei-Cas E & Chermette R (2001) Parallel phylogenies of

Pneumocystis species and their mammalian hosts. J Eukaryot

Microbiol 47: S113–S115.

Gustafson MP, Thomas CF Jr, Rusnak F, Limper AH & Leof EB

(2001) Differential regulation of growth and checkpoint

control mediated by a Cdc25 mitotic phosphatase from

Pneumocystis carinii. J Biol Chem 276: 835–843.

Haque AU, Plattner SB, Cook RT & Hart MN (1987)

Pneumocystis carinii: taxonomy as viewed by electron

microscopy. Am J Clin Pathol 87: 504–510.

Harigaya Y & Yamamoto M (2007) Molecular mechanisms

underlying the mitosis–meiosis decision. Chromosome Res 15:

523–537.

Hey J (2001) The mind of the species problem. Trends Ecol Evol

16: 326–329.

Hughes WT (1982) Natural mode of acquisition for de novo

infection with Pneumocystis carinii. J Infect Dis 145: 842–848.

Hugot JP, Demanche C, Barriel V, Dei-Cas E & Guillot J (2003)

Phylogenetic systematics and evolution of primate-derived

Pneumocystis based on mitochondrial or nuclear DNA

sequence comparison. Syst Biol 52: 735–744.

Hunter JA & Wakefield AE (1996) Genetic divergence at the

mitochondrial small subunit ribosomal RNA gene among

isolates of Pneumocystis carinii from five mammalian host

species. J Eukaryot Microbiol 43: S24–S25.

Itatani CA (1994) Ultrastructural demonstration of a pore in the

cyst wall of Pneumocystis carinii. J Parasitol 80: 644–648.

Itatani CA (1996) Ultrastructural morphology of intermediate

forms and forms suggestive of conjugation in the life cycle of

Pneumocystis carinii. J Parasitol 82: 163–171.

James TY, Kauff F, Schoch CL et al. (2006) Reconstructing the

early evolution of Fungi using a six-gene phylogeny. Nature

443: 818–822.

Kaiser K, Rabodonirina M, Mayencon M & Picot S (1999) Cdc2

gene of Pneumocystis carinii hominis and its expression during

culture. J Eukaryot Microbiol 46: S130.

Kaneshiro ES (1998) The lipids of Pneumocystis carinii. Clin

Microbiol Rev 11: 27–41.

Kaneshiro ES & Maiorano JN (1996) Survival and infectivity of

Pneumocystis carinii outside the mammalian host. J Eukaryot

Microbiol 43: S35.

Keely S, Pai HJ, Baughman R, Sidman C, Sunkin C, Stringer JR &

Stringer SL (1994) Pneumocystis species inferred from analysis

of multiple genes. J Eukaryot Microbiol 41: S94.

Keely SP & Stringer JR (2005) Nomenclature and genetic

variation of Pneumocystis. Pneumocystis Pneumonia, 3rd edn

(Walzer PD & Cushion MT, eds), pp. 39–59. Marcel Dekker,

New York.

Keely SP, Fischer JM & Stringer JR (2003) Evolution and

speciation of Pneumocystis. J Eukaryot Microbiol 50:

S624–S626.

Keely SP, Fischer JM, Cushion MT & Stringer JR (2004)

Phylogenetic identification of Pneumocystis murina sp. nov., a

new species in laboratory mice. Microbiology 150: 1153–1165.

Kottom TJ, Thomas CF Jr, Mubarak KK, Leof EB & Limper AH

(2000) Pneumocystis carinii uses a functional cdc13 B-type

cyclin complex during its life cycle. Am J Resp Cell Mol 22:

722–731.

Kottom TJ, Kohler JR, Thomas CF, Fink GR & Limper AH (2003)

Lung epithelial cells and extracellular matrix components

induce expression of Pneumocystis carinii STE20, a gene

complementing the mating and pseudohyphal growth defects

of STE20 mutant yeast. Infect Immun 71: 6463–6471.

Kovacs JA, Powell F, Edman JC, Lundgren B, Martinez A, Drew B

& Angus CW (1993) Multiple genes encode the major surface

glycoprotein of Pneumocystis carinii. J Biol Chem 268:

6034–6040.

Krajicek BJ, Kottom TJ, Villegas L & Limper AH (2010)

Characterization of the PcCdc42 small G protein from

Pneumocystis carinii, which interacts with the PcSte20 life cycle

regulatory kinase. Am J Physiol Lung C 298: L252–L260.

Krungkrai J, Webster HK & Yuthavong Y (1990) Folate and

cobalamin metabolism in Plasmodium falciparum. Parasitol

Today 6: 388–391.

Kwon KY, Kim SP & Limper AH (1998) Recognition of

Pneumocystis carinii antigen on its surface by

immunohistochemistry and immunoelectron microscopy.

J Korean Med Sci 13: 131–137.

Laakkonen J (1998) Pneumocystis carinii in wildlife. Int J Parasitol

28: 241–252.

Laakkonen J & Sukura A (1997) Pneumocystis carinii of the

common shrew, Sorex araneus, shows a discrete phenotype.

J Eukaryot Microbiol 44: 117–121.

Landvik S, Eriksson OE & Berbee ML (2001) Neolecta – a fungal

dinosaur? Evidence from b-tubulin amino acid sequences.

Mycologia 93: 1151–1163.

Larsen HH, von Linstow ML, Lundgren B, Hogh B, Westh H &

Lundgren JD (2007) Primary Pneumocystis infection in infants

hospitalized with acute respiratory tract infection. Emerg Infect

Dis 13: 66–72.

Latouche S, Olsson M, Polack B, Brun-Pascaud M, Bernard C &

Roux P (1997) Detection of Pneumocystis carinii f. sp. in air

samples collected in animal rooms. J Eukaryot Microbiol 44:

46–47.

Latouche S, Totet A, Lacube P, Bolognini J, Nevez G & Roux P

(2001) Development of an RT-PCR on the heat shock protein

FEMS Yeast Res 11 (2011) 2–17c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

14 M. Chabe et al.

70 gene for viability detection of Pneumocystis carinii f. sp.

hominis in patients with pneumocystosis and in air sample.

J Eukaryot Microbiol 48: S176–S177.

Li L, Wright SJ, Krystofova S, Park G & Borkovich KA (2007)

Heterotrimeric G protein signaling in filamentous fungi. Annu

Rev Microbiol 61: 423–452.

Liu Y, Rocourt M, Pan S, Liu C & Leibowitz MJ (1992) Sequence

and variability of the 5.8S and 26S rRNA genes of Pneumocystis

carinii. Nucleic Acids Res 20: 3763–3772.

Liu Y, Leigh JW, Brinkmann H, Cushion MT, Rodriguez-Ezpeleta

N, Philippe H & Lang BF (2009) Phylogenomic analyses

support the monophyly of Taphrinomycotina, including

Schizosaccharomyces fission yeasts. Mol Biol Evol 26: 27–34.

Liu YJ, Hodson MC & Hall BD (2006) Loss of flagellum

happened only once in the fungal lineage: phylogenetic

struture of kingdom Fungi inferred from RNA polymerase II

subunit genes. BMC Evol Biol 6: 74.

Lu JJ, Bartlett MS, Shaw MM, Queener SF, Smith JW, Ortiz-

Rivera M, Leibowitz MJ & Lee CH (1994) Typing of

Pneumocystis carinii strains that infect humans based on

nucleotide sequence variations of internal transcribed spacers

of rRNA genes. J Clin Microbiol 32: 2904–2912.

Maher NH, Vermund SH, Welsh DA, Dillon HK, Awooda A &

Unnasch TR (2001) Development and characterization of a

molecular viability assay for Pneumocystis carinii f sp hominis.

J Infect Dis 183: 1825–1827.

Martinez A, Aliouat EM, Pottier M, Gantois M, Pincon C,

Standaert-Vitse A, Dei-Cas E & Aliouat CM (2009) High-

speed cell sorting of infectious trophic and cystic forms of

Pneumocystis carinii. J Eukaryot Microbiol 56: 446–453.

Matsumoto Y & Yoshida Y (1984) Sporogony in Pneumocystis

carinii: synaptonemal complexes and meiotic nuclear divisions

observed in precysts. J Protozool 31: 420–428.

Mayden RL (1997) A hierarchy of species concepts: the

denouement in the saga of the species problem. Species: The

Units of Biodiversity (Claridge MF, Dawah HA & Wilson MR,

eds), pp. 381–424. Chapman & Hall, London.

Mayr E (1963) Animal Species and Evolution. Harvard University

Press, Cambridge, MA.

Mazars E, Odberg-Ferragut C, Dei-Cas E, Fourmaux MN, Aliouat

EM, Brun-Pascaud M, Mougeot G & Camus D (1995)

Polymorphism of Pneumocystis carinii from different host

species. J Eukaryot Microbiol 42: 26–32.

Mazars E, Guyot K, Durand I, Dei-Cas E, Boucher S, Ben

Abderrazak S, Banuls AL, Tibayrenc M & Camus D (1997)

Isoenzyme diversity in Pneumocystis carinii from rats, mice

and rabbits. J Infect Dis 175: 655–660.

Miller RF, Ambrose HE & Wakefield AE (2001) Pneumocystis

carinii f. sp. hominis DNA in immunocompetent health care

workers in contact with patients with P. carinii pneumonia.

J Clin Microbiol 39: 3877–3882.

Montes-Cano MA, Chabe M, Fontillon-Alberdi M, de-Lahorra C,

Respaldiza N, Medrano FJ, Varela JM, Dei-Cas E & Calderon

EJ (2009) Vertical transmission of Pneumocystis jirovecii in

humans. Emerg Infect Dis 15: 125–127.

Nei M, Xu P & Glazko G (2001) Estimation of divergence times

from multiprotein sequences for a few mammalian species and

several distantly related organisms. P Natl Acad Sci USA 98:

2497–2502.

Nielsen MH, Settnes OP, Aliouat EM, Cailliez JC & Dei-Cas E

(1998) Different ultrastructural morphology of Pneumocystis

carinii derived from mice, rats, and rabbits. APMIS 106:

771–779.

Nishida H & Sugiyama J (1993) Phylogenetic relationships

among Taphrina, Saitoella, and other higher fungi. Mol Biol

Evol 10: 431–436.

Nishida H & Sugiyama J (1994) Archiascomycetes: detection of a

major new lineage within the Ascomycota. Mycoscience 35:

361–366.

Olsson M, Sukura A, Lindberg LA & Linder E (1996) Detection of

Pneumocystis carinii DNA by filtration of air. Scand J Infect Dis

28: 279–282.

Olsson M, Lidman C, Latouche S, Bjorkman A, Roux P, Linder E

& Wahlgren M (1998) Identification of Pneumocystis carinii f.

sp. hominis gene sequences in filtered air in hospital

environments. J Clin Microbiol 36: 1737–1740.

Ortiz-Rivera M, Liu Y, Felder R & Leibowitz MJ (1995)

Comparison of coding and spacer region sequences of

chromosomal rRNA-coding genes of two sequevars of

Pneumocystis carinii. J Eukaryot Microbiol 42: 44–49.

Palluault F, Dei-Cas E, Slomianny C, Soulez B & Camus D (1990)

Golgi complex and lysosomes in rabbit-derived Pneumocystis

carinii. Biol Cell 70: 73–82.

Palluault F, Pietrzyck B, Dei-Cas E, Slomianny C, Soulez B &

Camus D (1991a) Three-dimensional reconstruction of

rabbit-derived Pneumocystis carinii. I. Trophozoite. J Protozool

38: 402–407.

Palluault F, Pietrzyck B, Dei-Cas E, Slomianny C, Soulez B &

Camus D (1991b) Three-dimensional reconstruction of

rabbit-derived Pneumocystis carinii. II. Intermediate precyst.

J Protozool 38: 407–411.

Palluault F, Slomianny C, Soulez B, Dei-Cas E & Camus D

(1992a) High osmotic pressure enables fine ultrastructural and

cytochemical studies on Pneumocystis carinii. I: epon

embedding. Parasitol Res 78: 437–444.

Palluault F, Soulez B, Slomianny C, Dei-Cas E, Cesbron JY &

Camus D (1992b) High osmotic pressure for Pneumocystis

carinii London Resin White embedding enables fine

immunocytochemistry studies: I. Golgi complex and cell-wall

synthesis. Parasitol Res 78: 482–488.

Palmer RJ, Settnes OP, Lodal J & Wakefield AE (2000) Population

structure of rat-derived Pneumocystis carinii in Danish wild

rats. Appl Environ Microb 66: 4954–4961.

Peters SE, English K, Rana A, Akter S, Malik S, Warburton NC &

Duckett JG (2001) Synaptonemal complexes in the pre-cyst of

Pneumocystis carinii. J Eukaryot Microbiol 48: S134.

Peterson JC & Cushion MT (2005) Pneumocystis: not just

pneumonia. Curr Opin Microbiol 8: 393–398.

FEMS Yeast Res 11 (2011) 2–17 c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

15Pneumocystis species

Pixley FJ, Wakefield AE, Banerji S & Hopkin JM (1991)

Mitochondrial gene sequences show fungal homology for

Pneumocystis carinii. Mol Microbiol 5: 1347–1351.

Rabodonirina M (2001) La pneumocystose est-elle une maladie

contagieuse? Lettre l’Infectiol 16: 110–112.

Rajagopalan-Levasseur P, Allaert A, Dridba M, Odberg-Ferragut

C, Jouault T, Creusy C, Camus D & Dei-Cas E (1998) Response

to Pneumocystis infection in an immunocompetent host.

FEMS Immunol Med Mic 22: 107–121.

Redhead SA, Cushion MT, Frenkel JK & Stringer JR (2006)

Pneumocystis and Trypanosoma cruzi: nomenclature and

typifications. J Eukaryot Microbiol 53: 2–11.

Ronquist F & Huelsenbeck JP (2003) MrBayes 3: Bayesian

phylogenetic inference under mixed models. Bioinformatics 19:

1572–1574.

Sanchez CA, Chabe M, Aliouat el M, Durand-Joly I, Gantois N,

Conseil V, Lopez C, Duriez T, Dei-Cas E & Vargas SL (2007)

Exploring transplacental transmission of Pneumocystis

oryctolagi in first-time pregnant and multiparous rabbit does.

Med Mycol 45: 701–707.

Settnes OP & Nielsen MJ (1991) Host–parasite relationship in

Pneumocystis carinii infection: activation of the plasmalemmal

vesicular system in type I alveolar epithelial cells. J Protozool

38: S174–S176.

Sexton AC & Howlett BJ (2006) Parallels in fungal pathogenesis

on plant and animal hosts. Eukaryot Cell 5: 1941–1949.

Sienkiewicz N, Jaroslawski S, Wyllie S & Fairlamb AH (2008)

Chemical and genetic validation of dihydrofolate reductase-

thymidylate synthase as a drug target in African trypanosomes.

Mol Microbiol 69: 520–533.

Sinclair K, Wakefield AE, Banerji S & Hopkin JM (1991)

Pneumocystis carinii organisms derived from rat and human

hosts are genetically distinct. Mol Biochem Parasitol 45:

183–184.

Sing A, Wonhas C, Bader L, Luther M & Heesemann J (1999)

Detection of Pneumocystis carinii DNA in the air filter of a

ventilated patient with AIDS. Clin Infect Dis 29: 952–953.

Smulian AG, Sesterhenn T, Tanaka R & Cushion MT (2001) The

ste3 pheromone receptor gene of Pneumocystis carinii is

surrounded by a cluster of signal transduction genes. Genetics

157: 991–1002.

Soulez B, Palluault F, Cesbron JY, Dei-Cas E, Capron A & Camus

D (1991) Introduction of Pneumocystis carinii in a colony of

SCID mice. J Protozool 38: S123–S125.

Spatafora JW, Sung GH, Johnson D et al. (2006) A five-gene

phylogeny of Pezizomycotina. Mycologia 98: 1018–1028.

Stagno S, Pifer LL, Hughes WT, Brasfield DM & Tiller RE (1980)

Pneumocystis carinii pneumonitis in young

immunocompetent infants. Pediatrics 66: 56–62.

Stedman TT, Butler DR & Buck GA (1998) The HSP70 gene

family in Pneumocystis carinii: molecular and phylogenetic

characterization of cytoplasmic members. J Eukaryot Microbiol

45: 589–599.

Stewart WC, Pollock KG, Browning LM, Young D, Smith-Palmer

A & Reilly WJ (2005) Survey of zoonoses recorded in Scotland

between 1993 and 2002. Vet Rec 157: 697–702.

Stringer JR & Cushion MT (1998) The genome of Pneumocystis

carinii. FEMS Immunol Med Mic 22: 15–26.

Stringer JR, Cushion MT & Wakefield AE (2001) New

nomenclature for the genus Pneumocystis. J Eukaryot Microbiol

48: 184–189.

Stringer SL, Stringer JR, Blase MA, Walzer PD & Cushion MT

(1989) Pneumocystis carinii: sequence from ribosomal RNA

implies a close relationship with fungi. Exp Parasitol 68:

450–461.

Sugiyama J, Hosaka K & Suh SO (2006) Early diverging

Ascomycota: phylogenetic divergence and related evolutionary

enigmas. Mycologia 298: 996–1005.

Tan WC & Ng TP (2008) COPD in Asia: where east meets west.

Chest 133: 517–527.

Taylor JW, Jacobson DJ, Kroken S, Kasuga T, Geiser DM, Hibbett

DS & Fisher MC (2000) Phylogenetic species recognition and

species concepts in fungi. Fungal Genet Biol 31: 21–32.

Thomas CF, Anders RA, Gustafson MP, Leof EB & Limper AH

(1998a) Pneumocystis carinii contains a functional cell-

division-cycle Cdc2 homologue. Am J Resp Cell Mol 18:

297–306.

Thomas CF Jr, Kottom TJ, Leof EB & Limper AH (1998b)

Characterization of a mitogen-activated protein kinase from

Pneumocystis carinii. Am J Physiol 275: L193–L199.

Totet A, Latouche S, Lacube P, Pautard JC, Jounieaux V, Raccurt

C, Roux P & Nevez G (2004) Pneumocystis jirovecii

dihydropteroate synthase genotypes in immunocompetent

infants and immunosuppressed adults, Amiens, France. Emerg

Infect Dis 10: 667–673.

Vargas SL, Ponce CA, Gigliotti F, Ulloa AV, Prieto S, Munoz MP &

Hughes WT (2000) Transmission of Pneumocystis carinii DNA

from a patient with P. carinii pneumonia to

immunocompetent contact health care workers. J Clin

Microbiol 38: 1536–1538.

Vavra J & Kucera K (1970) Pneumocystis carinii delanoe, its

ultrastructure and ultrastructural affinities. J Protozool 17:

463–483.

Villegas LR, Kottom TJ & Limper AH (2010) Characterization of

PCEng2 a {beta}-1,3-endoglucanase homologue in

Pneumocystis carinii with activity in cell wall regulation. Am J

Resp Cell Mol 43: 192–200.

Vohra PK, Puri V, Kottom TJ, Limper AH & Thomas CF Jr

(2003a) Pneumocystis carinii STE11, an HMG-box protein, is

phosphorylated by the mitogen activated protein kinase PCM.

Gene 312: 173–179.

Vohra PK, Puri V & Thomas CF Jr (2003b) Complementation

and characterization of the Pneumocystis carinii MAPK, PCM.

FEBS Lett 551: 139–146.

Vohra PK, Park JG, Sanyal B & Thomas CF Jr (2004) Expression

analysis of PCSTE3, a putative pheromone receptor from the

lung pathogenic fungus Pneumocystis carinii. Biochem Bioph

Res Co 319: 193–199.

FEMS Yeast Res 11 (2011) 2–17c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

16 M. Chabe et al.

Wakefield AE (1996) DNA sequences identical to Pneumocystis

carinii f. sp. carinii and Pneumocystis carinii f. sp. hominis in

samples of air spora. J Clin Microbiol 34: 1754–1759.

Wakefield AE (1998) Genetic heterogeneity in Pneumocystis

carinii: an introduction. FEMS Immunol Med Mic 22: 97–101.

Wakefield AE, Peters SE, Banerji S, Bridge PD, Hall GS,

Hawksworth DL, Guiver LA, Allen AG & Hopkin JM (1992)

Pneumocystis carinii shows DNA homology with the

Ustomycetous red yeast fungi. Mol Microbiol 6: 1903–1911.

Walzer PD, Schnelle V, Armstrong D & Rosen PP (1977) Nude

mouse: a new experimental model for Pneumocystis carinii

infection. Science 197: 177–179.

Wyder MA, Rasch EM & Kaneshiro ES (1994) Assessment of

Pneumocystis carinii DNA content. J Eukaryot Microbiol 41:

S120.

Wyder MA, Rasch EM & Kaneshiro ES (1998) Quantitation of

absolute Pneumocystis carinii nuclear DNA content. Trophic

and cystic forms isolated from infected rat lungs are haploid

organisms. J Eukaryot Microbiol 45: 233–239.

Yamada M, Matsumoto Y, Hamada S, Fujita S & Yoshida Y (1986)

Demonstration and determination of DNA in Pneumocystis

carinii by fluorescence microscopy with 40,6-diamidino-2-

phenylindole (DAPI). Zbl Bakt Mik Hyg A 262: 240–246.

Yoshida Y (1989) Ultrastructural studies of Pneumocystis carinii. J

Protozool 36: 53–60.

Yoshida Y, Matsumoto Y, Yamada M, Okabayashi K, Yoshikawa H

& Nakazawa M (1984) Pneumocystis carinii: electron

microscopic investigation on the interaction of

trophozoite and alveolar lining cell. Zbl Bakt Mik Hyg A 256:

390–399.

Yoshikawa H & Yoshida Y (1987) Localization of silver deposits

on Pneumocystis carinii treated with Gomori’s methenamine

silver nitrate stain. Zbl Bakt Mik Hyg A 264: 363–372.

Yoshikawa H, Morioka H & Yoshida Y (1988) Ultrastructural

detection of carbohydrates in the pellicle of Pneumocystis

carinii. Parasitol Res 74: 537–543.

Youn H (2009) Review of zoonotic parasites in medical and

veterinary fields in the Republic of Korea. Korean J Parasitol 47:

S133–S141.

Ypma-Wong MF, Fonzi WA & Sypherd PS (1992) Fungus-specific

translation elongation factor 3 gene present in Pneumocystis

carinii. Infect Immun 60: 4140–4145.

FEMS Yeast Res 11 (2011) 2–17 c� 2010 Federation of European Microbiological SocietiesPublished by Blackwell Publishing Ltd. All rights reserved

17Pneumocystis species