ArticlePDF AvailableLiterature Review

Opportunistic yeast pathogens: Reservoirs, virulence mechanisms, and therapeutic strategies

Authors:

Abstract and Figures

Life-threatening invasive fungal infections are becoming increasingly common, at least in part due to the prevalence of medical interventions resulting in immunosuppression. Opportunistic fungal pathogens of humans exploit hosts that are immunocompromised, whether by immunosuppression or genetic predisposition, with infections originating from either commensal or environmental sources. Fungal pathogens are armed with an arsenal of traits that promote pathogenesis, including the ability to survive host physiological conditions and to switch between different morphological states. Despite the profound impact of fungal pathogens on human health worldwide, diagnostic strategies remain crude and treatment options are limited, with resistance to antifungal drugs on the rise. This review will focus on the global burden of fungal infections, the reservoirs of these pathogens, the traits of opportunistic yeast that lead to pathogenesis, host genetic susceptibilities, and the challenges that must be overcome to combat antifungal drug resistance and improve clinical outcome.
Content may be subject to copyright.
REVIEW
Opportunistic yeast pathogens: reservoirs, virulence mechanisms,
and therapeutic strategies
Elizabeth J. Polvi Xinliu Li Teresa R. O’Meara
Michelle D. Leach Leah E. Cowen
Received: 24 December 2014 / Revised: 6 February 2015 / Accepted: 11 February 2015
ÓSpringer Basel 2015
Abstract Life-threatening invasive fungal infections are
becoming increasingly common, at least in part due to the
prevalence of medical interventions resulting in immuno-
suppression. Opportunistic fungal pathogens of humans
exploit hosts that are immunocompromised, whether by
immunosuppression or genetic predisposition, with infec-
tions originating from either commensal or environmental
sources. Fungal pathogens are armed with an arsenal of
traits that promote pathogenesis, including the ability to
survive host physiological conditions and to switch be-
tween different morphological states. Despite the profound
impact of fungal pathogens on human health worldwide,
diagnostic strategies remain crude and treatment options
are limited, with resistance to antifungal drugs on the rise.
This review will focus on the global burden of fungal in-
fections, the reservoirs of these pathogens, the traits of
opportunistic yeast that lead to pathogenesis, host genetic
susceptibilities, and the challenges that must be overcome
to combat antifungal drug resistance and improve clinical
outcome.
Keywords Opportunistic Fungi Yeast Pathogen
Candida Cryptococcus Histoplasma Pneumocystis
Abbreviations
ABPA Allergic bronchopulmonary aspergillosis
APECED Autoimmune polyendocrinopathy–
candidiasis–ectodermal dystrophy
cAMP Cyclic AMP
CDC Centers for Disease Control and Prevention
CGD Chronic granulomatous disease
CLR C-type lectin receptors
CMC Chronic mucocutaneous candidiasis
CNS Central nervous system
COPD Chronic obstructive pulmonary disease
CRP C-reactive protein
ECM Extracellular matrix
GUT Gastrointestinally induced transition
GXM Glucuronoxylomannan
GXMGal Glucuronoxylomannogalactan
HAART Highly active antiretroviral therapies
HIES Hyper-IgE syndrome
HSP Heat shock protein
IFN-cInterferon-c
IL Interleukin
IRIS Immune reconstitution inflammatory
syndrome
MPO Myeloperoxidase
MR Mannose receptors
NF-jB Nuclear factor-jB
NO Nitric oxide
PAMP Pathogen-associated molecular pattern
PCP Pneumocystis pneumonia
PKA Protein kinase A
PKC Protein kinase C
PRR Pattern recognition receptors
SAP Secreted aspartyl proteinase
TGF-bTransforming growth factor-b
E. J. Polvi and X. Li contributed equally to this work.
E. J. Polvi X. Li T. R. O’Meara M. D. Leach
L. E. Cowen (&)
Department of Molecular Genetics, University of Toronto,
1 King’s College Circle, Medical Sciences Building, Room
4368, Toronto, ON M5S 1A8, Canada
e-mail: leah.cowen@utoronto.ca
M. D. Leach
Aberdeen Fungal Group, Institute of Medical Sciences, School
of Medical Sciences, University of Aberdeen, Foresterhill,
Aberdeen, UK
Cell. Mol. Life Sci.
DOI 10.1007/s00018-015-1860-z Cellular and Molecular Life Sciences
123
TLR Toll-like receptors
TNF-aTumor necrosis factor-a
Treg cell Regulatory T cell
VVC Vulvovaginal candidiasis
Introduction
The emergence of fungi occurred approximately 1.6 billion
years ago [1,2]. With an estimated 1.5 million species
occupying a range of environments [3], fungal species are
extraordinarily diverse. Fungi are a major cause of disease
in insects, amphibians, plants and even other fungi, with
incidences of infection reaching unprecedented levels in
recent years. Strikingly, millions of acres of the world’s
forests have become victim to fungal infections. The blue
stain fungus Grosmannia clavigera has devastated North
American pine forests [4], and elm trees have been in
significant decline since the twentieth century due to Dutch
elm blight. Fungi also pose a major threat to food supplies,
being responsible for the destruction of wheat and barley
crops by wheat leaf rust and rice by rice blast fungus.
Additionally, fungi pose a significant threat to animals,
crippling bat and amphibian species worldwide [5].
Of the 1.5 million fungal species, around 300 are re-
ported to be pathogenic in humans, and only a minority of
these are common human pathogens [6]. Many fungi are
commensal, forming part of our natural microbiota. Indeed,
a recent study by Findley et al. [7] illustrated the diversity
of fungal species on three different foot sites, and there is a
growing appreciation that fungi have an important role in
defining commensal microbial communities [8]. This
commensal role allows fungi to infect humans in multiple
ways. Ranging from cutaneous infections affecting 29
million North Americans each year [9], to superficial skin
infections, to more than 2 million invasive fungal infec-
tions per year worldwide [10], commensal fungi have
important roles in disease, being capable of switching to
opportunistic pathogens. It is the opportunistic nature of
fungal pathogens that triggered a stark rise in fungal in-
fections in the late twentieth century, primarily in hosts
with impaired immunity due to medical interventions such
as chemotherapy for cancer, organ transplantation, or in-
fection with HIV [11]. Overall, fungi are the cause of
billions of infections worldwide, killing over 1.5 million
people annually [10,12]. Species of Aspergillus,Candida,
Cryptococcus and Pneumocystis are at the forefront of
fungal infections, accounting for approximately 90 % of
human mortality cases [10]. Poor diagnostic tools and an-
tifungal drug resistance accounts for a 50 % or higher
mortality rate in patients with systemic fungal infections
[11].
It has been postulated that global warming is in part
responsible for the drastic increase in fungal infections,
leading to the devastating loss of forests and crops
worldwide. Remarkably, the common denominator of all
human pathogenic fungi is the ability to grow at host
temperatures [13,14]. The steady rise in temperature upon
climate change will selectively enable adaptation of fungi,
broadening the array of species able to survive at host
temperature [15]. Indeed, clinical isolates of the generally
benign yeast Saccharomyces cerevisiae that exhibit en-
hanced capacity to grow at higher temperature (41 °C)
compared to laboratory strains, were able to survive in
mice [16]. The common fungal pathogens Cryptococcus
neoformans,Histoplasma capsulatum and Aspergillus fu-
migatus are found in environments as diverse as pigeon
excreta and soil, yet each retains the capacity to grow at
37 °C. Many of the genes required for growth at high
temperatures in these pathogens are necessary for virulence
and some are required for survival [1719]. Consequently,
high temperature growth is essential for pathogenesis. This
review will focus on the reservoirs and mechanisms of
pathogenic yeast infections, the challenges faced upon in-
fection and the hurdles that must be overcome to combat
antifungal drug resistance.
Global burden of fungal infections
Superficial and mucosal fungal infections are extremely
common, but life-threatening systemic fungal infections
are generally limited to immunocompromised individuals
[20,21]. The population of vulnerable individuals expe-
riencing some form of altered immune function due to the
HIV/AIDS epidemic, hospitalization, chronic illnesses,
antibiotic-mediated microbiota alteration, or che-
motherapies continues to increase [10,22,23], and has led
to billions of cases of invasive fungal infections
worldwide.
Cryptococcus
Cryptococcus neoformans is the leading cause of deaths
due to fungal infections, with a global burden of nearly 1
million cases annually, and more than 620,000 deaths
worldwide [23]. Cryptococcus is ubiquitous and globally
distributed, with more than 70 % of children older than 5
demonstrating serum reactivity against cryptococcal pro-
teins [24]. A major risk factor for cryptococcal infections is
immunosuppression due to HIV/AIDS. In resource-limited
areas with high rates of HIV/AIDS, the mortality rate ap-
proaches 70 % [23], and cryptococcal meningitis is
considered an AIDS-defining illness [25]. Left untreated,
cryptococcal meningitis is uniformly fatal.
E. J. Polvi et al.
123
In Western Europe and North America, the overall
mortality rate of cryptococcal infections is 10 %, with
approximately 700 deaths per year [23]. However, nearly
20 % of cryptococcosis cases in the US are in non-HIV
patients [26], and the mortality rate of these infections can
be greater than 30 % [27]. Many of these infections are in
patients with conditions associated with immunosuppres-
sion or with solid organ transplants, although there are
some incidences of underlying genetic susceptibility to
fungal infections [2628].
Another source of cryptococcosis in immunocompetent
individuals is infections due to Cryptococcus gattii, which
has caused an outbreak in the Pacific Northwest over the
last 15 years [2931]. The majority of these cases have
been in immunocompetent individuals, and many cases
were fatal despite antifungal therapies [31]. Although the
origin of the outbreak was Vancouver Island, cases of C.
gattii infections have been documented along the pacific
coast and in Florida [32,33].
An additional complication in cryptococcal treatment is
the occurrence of immune reconstitution inflammatory
syndrome (IRIS), where restoration of immune function
after starting highly active antiretroviral therapies
(HAART) results in enhanced and destructive immune
responses to subclinical or cleared microbial infections
[34]. Recent studies have reported that between 8 and 50 %
of AIDS patients develop cryptococcal IRIS after treatment
with HAART, and the mortality rate ranges between 30 and
60 % [35,36].
Candida
Candida species are normal members of the human mu-
cosal microbiota; studies have estimated between 25 and
40 % of people are colonized with Candida albicans at any
given point [37,38]. However, C. albicans also causes
more than 400,000 deaths per year due to invasive can-
didiasis [39]. Disseminated Candida infections have an
attributable mortality rate approaching 40 %, even with
antifungal therapies [39].
Patients with HIV/AIDS or other forms of impaired
immunity, especially neutropenia, are vulnerable to dis-
seminated candidiasis [25,39,40]. Additionally, Candida
species are the fourth most common cause of hospital-ac-
quired bloodstream infections, and the incidence of
Candida infections is increasing [41]. Candida albicans is
the most common causal agent of nosocomial infection
among the Candida species, followed by C. glabrata,C.
parapsilosis, and C. tropicalis [41]. Many of these infec-
tions may be due to breakdown of mucosal surfaces during
hospitalization or treatment, which then allows the super-
ficial Candida colonization access to the bloodstream [42,
43]. Finally, C. albicans can form biofilms on medically
implanted devices and catheters, creating a reservoir for
bloodstream infections [44,45].
Recently, mortality rates due to C. albicans infections in
high-risk patients, such as stem cell transplant recipients or
those undergoing therapy for leukemia, have decreased due
to prophylactic fluconazole therapies [39,42]. However,
resistance can occur [46,47], and the Centers for Disease
Control and Prevention (CDC) has ranked fluconazole-re-
sistant Candida as a serious threat [48].
Pneumocystis
Pneumocystis is an almost ubiquitous colonizer of human
lungs, with approximately 80 % of children demonstrating
serum antibodies to the organism [4951]. However, it can
also cause pneumocystis pneumonia (PCP) in immuno-
compromised hosts. In the early years of the AIDS
epidemic, PCP was the most common clinical manifesta-
tion of decreased immune function [52]. Before HAART,
approximately 65 % of AIDS patients in the United States
exhibited PCP [52,53]. Recent studies in developing na-
tions have suggested that up to 40 % of HIV-infected
patients exhibit PCP [52,54]. An extrapolation of the
current epidemiological reports suggests that there are
more than 400,000 cases per year [10,54].
Pneumocystis pneumonia remains one of the leading
causes of morbidity and mortality worldwide [50]. In de-
veloped nations with readily available HAART, the
mortality rate has dropped to approximately 10 % [55,56].
However, in a recent study of pulmonary disease in AIDS
patients in Uganda, the number of cases of PCP was lower
than for tuberculosis or cryptococcal pneumonia, but the
mortality was higher, with 75 % of the patients dying
within 2 months of admission [57]. Additionally, Pneu-
mocystis infections may also contribute to impaired lung
function and increased mortality of patients with chronic
obstructive pulmonary disease (COPD) [50,58].
Aspergillus
Aspergillus is a ubiquitous filamentous fungus, and people
are faced with continuous exposure to Aspergillus spores.
As an opportunistic pathogen, Aspergillus mainly causes
disease in patients with impaired immune function. Most
infections are acquired exogenously instead of through
reactivation of latent infections [59]. The most vulnerable
population includes patients with neutropenia, organ or
bone marrow transplant, or those undergoing immunosup-
pressive therapies [42,59,60]. Invasive aspergillosis is
uniformly fatal if untreated, and even with current treat-
ment options, mortality remains at 50 % [42,61]. Current
reports suggest that there are more than 200,000 cases of
invasive aspergillosis per year [10].
Opportunistic yeast pathogens
123
Aspergillus can also cause allergic bronchopulmonary
aspergillosis (ABPA), and recent estimates suggest that the
global burden of ABPA is 4,837,000 patients [62,63].
Patients with lung disease, such as COPD, cystic fibrosis,
or emphysema are especially susceptible to ABPA [59,60,
63]. Infection with Aspergillus can increase the mortality of
these underlying diseases, contributing to the 100,000
deaths per year [10].
Histoplasma
Histoplasma capsulatum is endemic to the Midwest United
States and Central America, though it has a global distri-
bution [64]. Histoplasma is the most prevalent cause of
systemic mycosis in Central America, and it infects several
hundred thousand individuals annually [64,65]. Recent
studies estimate nearly 80 % of the adult population in the
endemic areas show previous infections by Histoplasma
[64]. Most of these infections occur in immunocompetent
hosts, but the majority does not display symptoms. How-
ever, individuals who are exposed to large numbers of
spores can suffer from acute pulmonary histoplasmosis.
Additionally, the disease is more severe in patients with
underlying lung disease [64,65]. Disseminated histoplas-
mosis usually only occurs in individuals with compromised
immune systems [64,66]. Mortality due to disseminated
histoplasmosis remains at 30 %, but it can be reduced with
HAART [65].
Paracoccidioides
Paracoccidioides is an endemic dimorphic fungal patho-
gen in South America, and it is the most common cause
of invasive mycosis in this region [65]. Nearly 10 million
people are infected with this organism, but in most cases,
the host immune system can prevent the conidia from
causing disease [65]. However, individuals with compro-
mised immune systems or those with decreased lung
function, such as smokers, are vulnerable to expansion of
the yeast from the lungs into disseminated disease [22,
65]. Interestingly, there is not a strong association of
paracoccidiomycosis with HIV [67]. In some cases (be-
tween 5 and 13 %), patients experience relapse of
paracoccidiomycosis due to reactivation of quiescent yeast
cells [68].
Penicillium
Penicillium marneffei is an endemic fungal pathogen in
Southeast Asia, causing disease primarily in immunocom-
promised patients, although infections in immunocompetent
hosts also occur [69,70]. In a study of AIDS patients in
northern Thailand, it was the third most common cause of
infection, after Mycobacterium tuberculosis and C. neofor-
mans [71]. As an opportunistic pathogen, the prevalence of
P. marneffei infections is increasing with the expansion of
the HIV epidemic in Southeast Asia [72].
Penicillium marneffei infections are more common in
the rainy season and are positively correlated with in-
creased humidity [73,74]. This suggests that disease occurs
from primary infections shortly after exposure to the fun-
gus [73]. These disseminated infections are uniformly fatal
without antifungal treatments, but even with rapid admin-
istration of therapeutics, the mortality rate remains
approximately 20 % [71,74]. In children with immune
deficiencies, the mortality rate is higher, at 55 % [70].
Coccidioides
Coccidiomycosis, also known as valley fever, is endemic to
the Southwestern United States and Central America, and it
causes thousands of hospitalizations and approximately
*400 deaths per year in the United States [75]. Coccid-
iomycosis is caused by Coccidioides immitis and
Coccidioides posadasii, which occupy different geographic
regions [76]. Although most infections are self-limiting
lung infections, Coccidioides can also cause disseminated
disease. Risk factors include compromised immune sys-
tems, but agricultural or construction work, outdoor
activity, African or Asian ethnicity, increased age, or
pregnancy also increase vulnerability to Coccidioides.
Additionally, uncontrolled diabetes increases the risk of
disseminated infections [77]. In patients with impaired T
cell function and disseminated infection, mortality rates
were 50 % [76].
Blastomyces
Blastomycosis is caused by infection with Blastomyces
dermatitidis, a dimorphic fungal pathogen that is endemic
in the Midwestern United States. Most infections are due to
spore inhalation during outdoor activities, and ap-
proximately 30–50 % of infections are asymptomatic [78].
Of the symptomatic infections, most can be treated with
antifungal drugs. However, mortality ranges between 4 and
22 % [78]. In a recent outbreak in Wisconsin, 70 % of the
cases were hospitalized and 5 % died [79]. Only 25 % of
the infected patients had any form of compromised im-
mune system. However, there appears to be genetic risk
factors for blastomycosis as the incidence rate is 12 times
higher for Asians [79].
Rhizopus and Mucor
Mucormycosis is due to infections by basal fungi, most
often by species of Rhizopus and Mucor [80]. Vulnerable
E. J. Polvi et al.
123
populations include patients with hematopoietic stem cell
transplants or other hematological malignancies, those with
diabetes, and those with compromised immune systems
[80,81]. The association between mucormycosis and dia-
betes is particularly strong; a recent study showed that
70 % of mucormycosis patients had diabetic ketoacidosis
[81]. The mortality rate can be as high as 90 % for dis-
seminated infections [82,83]. Cutaneous mucormycosis
can also occur after natural disasters, when environmental
spores are disturbed and patients have open wounds [84].
The case-fatality rate for cutaneous mucormycosis can be
as high as 80 % [84,85].
Sporothrix
Sporotrichosis is usually caused by traumatic inoculation
by material contaminated with Sporothrix schenckii,
although there are some cases of inhalational sporotrichosis
[86]. Although the infections are restricted to the site of
infection, there are cases of disseminated sporotrichosis,
mostly in immunocompromised individuals [67]. A recent
study of cases in HIV-infected patients in Brazil found that
the incidence of sporotrichosis was increasing to the point
where there the incidence was comparable to histoplas-
mosis and cryptococcosis [67]. Of the patients with both
HIV and sporotrichosis, almost 40 % were hospitalized for
disseminated infection [67].
Trichosporon
Like Candida,Trichosporon can exist as a member of the
normal human microbiota [87]. However, patients with
malignant hematological disorders are vulnerable to in-
fection with Trichosporon [87]. Trichosporon infections
are the second most common cause of disseminated fungal
infections in these patients, although the overall rate is low
[88]. The mortality rate of patients with hematological
disorders and Trichosporon remains at 11 %, even with
antifungal therapies [88].
Reservoirs and nature of infection
Here we discuss the reservoirs and nature of infection,
focusing on the predominant opportunistic yeast and yeast-
like pathogens, Cryptococcus,Candida,Histoplasma, and
Pneumocystis.
Cryptococcus
Cryptococcus species are found ubiquitously in nature, but
only C. neoformans and C. gattii are considered serious
human pathogens [89,90]. The source of human infections
are thought to be exclusively environmental, as there is no
evidence of human to human transmission other than
through contaminated medical devices [91]. Furthermore,
concordance in phenotypic and genotypic characterization
between environmental and clinical isolates within the
same geographical region suggests they belong to the same
fungal population [89,92,93].
Environmental isolates of these yeasts have been re-
ported all around the globe, recovered from soil, dust,
avian excreta, trees and other plants, domestic and wild
animals, as well as marine mammals [94]. Cryptococcus
neoformans is most often found in excreta from pigeons
and other birds. In southern Africa, isolates are commonly
found in the decayed hollows of the mopane tree [95].
Cryptococcus gattii is mostly isolated from tropical and
subtropical regions in association with eucalyptus trees, as
well as from soil, trees, and animals in the Pacific North-
west [94,96]. Plant materials have also been shown to
promote fertility and virulence of C. gattii [97].
Cryptococcal infection occurs upon inhalation of air-
borne fungal cells, generally spores or desiccated yeast [98].
Once inside the host lung, the yeast particles will encounter
either alveolar macrophages or dendritic cells and trigger an
immune response [99]. This can lead to clearance of the
infection, or result in a localized asymptomatic latent in-
fection that is contained within a granuloma, where the
yeast is enveloped by immune cells [99,100]. This latency
period could last for years before reactivation of dormant
infection occurs and disease symptoms develop [89,101].
In the case of C. neoformans, reactivation occurs when host
immunity is compromised, such as in patients with HIV/
AIDS [89]. Cryptococcus gattii infections appear to be new
events as they occur in both immunocompromised and
immunocompetent individuals [102].
Active infection in the lung leads to pneumonia-like
illness, with common symptoms including cough, fever,
and dyspnea, among others [103,104]. Fungal cells can
subsequently disseminate from the lung through vascular or
lymphatic systems to cause systemic infections, with the
central nervous system (CNS) being the preferred desti-
nation [91,100]. Cryptococcus can cross the blood–brain
barrier either directly by transcytosis through the en-
dothelial cells or with the aid of host monocytes [105107].
Infection of the CNS and brain parenchyma is life threat-
ening and results in cryptococcal meningitis and
meningoencephalitis [104,108]. Neurological symptoms
include headaches, visual and hearing impairment, sei-
zures, and mental status change [108,109].
Candida
Although a large number of Candida species have been
documented, only a few are known to cause disease in
Opportunistic yeast pathogens
123
humans [11,110]. Over 90 % of all Candida infections are
caused by five species: C. albicans,Candida glabrata,
Candida parapsilosis,Candida tropicalis, and Candida
krusei. Among these, C. albicans is by far the most
prevalent, responsible for 90–100 % of superficial mucosal
infections and 40–70 % of disseminated infections [11,
110].
Unlike most other pathogenic fungi, Candida species are
a natural commensal of the human microbiome [111], found
rarely in the soil and external environments, suggesting
adaptation to a parasitic lifestyle [112]. The majority of
Candida infections are from endogenous sources, derived
from commensal populations acquired prior to disease de-
velopment [11,113]. Exogenous sources of infection are
also common, especially in healthcare settings where
transmission can occur from healthcare workers, other pa-
tients, and contaminated medical devices [11,114].
Normally, Candida species exist harmlessly on our skin
and mucosal surfaces as part of the commensal microbiota,
colonizing the skin, oral cavity, gastrointestinal tract, and
reproductive tract [111,115]. However, disruption of the
normal microbial flora or compromising the immune sys-
tem may enable this fungus to overgrow, resulting in
symptomatic infections. Due to the long evolutionary his-
tory with the human host, C. albicans is well adapted to
survive and proliferate in the host environment. The switch
from commensalism to pathogenesis involves significant
changes in the fungus, including regulation of key viru-
lence genes that allows it to quickly sense and adjust to
different sites of infection in the body. Candida species
must overcome different stresses, such as changes in tem-
perature, pH, osmolarity, oxygen and nutrient availability
[14,112,115118].
Inflammations of the genitourinary tract are common
clinical manifestations of Candida infections, which include
vulvovaginal candidiasis (VVC) in women, balanitis and
balanoposthitis in men, and candiduria in both sexes [119].
VVC, commonly referred to as thrush or yeast infection,
typically presents as isolated occurrences of mild to mod-
erate infection in otherwise healthy women, and are
generally easily controlled with a single-dose therapy [110].
Episodes can occur either sporadically or due to predispo-
sition resulting from a number of risk factors, including
antibiotic use, pregnancy, diabetes, and immunosuppression
[21,120,121]. A subgroup of patients will experience severe
or recurrent VVC, in which case more rigorous and long-
term antifungal therapies are required [110]. It is estimated
that the majority of women develop at least one episode of
VVC in their lifetime [122]. Candida balanitis occurs at a
lower frequency than VVC, though the exact incidence is
unclear due to the lack of population studies [119]. The
disease is generally sexually transmitted and has been as-
sociated with diabetes and antibiotic use [123]. Candiduria,
the presence of Candida species in the urinary tract, is
commonly diagnosed in hospitalized patients, especially
those with a urinary catheter [124126]. Most cases of can-
diduria are asymptomatic or self-limiting, though it can
cause increased vulnerability to bloodstream infections in
high-risk patients, as well as increased mortality rates and
hospital costs [119,127].
Candida infections can also develop in the mouth or
throat, resulting in oropharyngeal candidiasis. Clinical
manifestations are typically characterized by white or red
lesions on the surfaces of the tongue and oropharynx,
which result in pain and burning sensations, alteration in
taste, and tissue damage [128,129]. Infection is generally
associated with immunosuppression, antibiotic use, en-
docrine alterations, and denture use [128,130]. It is
estimated that over 90 % of HIV-infected patients will
develop oropharyngeal candidiasis at some point during the
progression of their illness [131].
Invasive candidiasis occurs when Candida species break
the mucosal barrier, penetrating into deeper tissue and
gaining access to the bloodstream [132]. Dissemination via
the bloodstream allows the fungus to invade almost all
body sites and organs, resulting in lethal systemic disease
[110]. Major risk factors include immunosuppression, in-
vasive medical procedures, and extended stay in the
intensive care unit [11,133,134]. Early clinical symptoms
of invasive candidiasis are non-specific and resemble other
nosocomial infections, which impede accurate diagnosis
and delays treatment, contributing to increased mortality
[132,135,136].
Pneumocystis
Pneumocystis species are ubiquitous in nature, infecting a
wide variety of mammalian hosts [11,137]. While rodents
were originally thought to be a natural reservoir, it is now
known that Pneumocystis is not zoonotic. Current pheno-
typic and genetic evidence indicates that each mammalian
species that contracts the disease has its own species of
Pneumocystis [137139]. Pneumocystis jirovecii is re-
sponsible for human infections. However, study of the
organism has been hindered by the inability to culture
Pneumocystis in vitro. Much of what we now know about
P. jirovecii is based on direct clinical evidence or ex-
trapolated from related animal models [50,140].
PCR-based strategies have been used to type and track
the spread of P. jirovecii [141,142]. Children and im-
munocompromised individuals are identified as the most
important sources of P. jirovecii infection [142145].
While some environmental sources have been identified,
transmission is thought to mostly occur through inhalation
of airborne pathogens that spread from human to human
[146,147].
E. J. Polvi et al.
123
Pneumocystis has strong tropism for the lung, where it
exists as an alveolar pathogen without invading the host
[140]. Infections are typically asymptomatic or subclinical
in the healthy host, but can develop into pneumocystic
pneumonia in immunocompromised individuals, especially
those with HIV infections [50]. Common clinical symp-
toms include progressive dyspnea, nonproductive cough,
low-grade fever, and eventual respiratory failure with dis-
ease progression [140].
Histoplasma
Histoplasma capsulatum exists as a mold in soil environ-
ments and is often associated with bird and bat droppings,
which enhance proliferation of the organism by accelerat-
ing sporulation [148]. Once contaminated, soil can yield
the organism for years after the birds are gone [64]. In-
fections occur through inhalation of the aerosolized H.
capsulatum from disturbed soils. Air currents can also
carry the fungal spore for miles, resulting in infections far
away from contaminated sites [148].
Most infections by H. capsulatum are asymptomatic in
healthy individuals [148]. In children who were exposed
for the first time and in individuals who experienced heavy
exposure to the organism, pulmonary histoplasmosis may
develop [64]. Common symptoms include fever, malaise,
cough, and chest pain. Pulmonary infections can be com-
plicated with infection of the mediastinal lymph nodes,
resulting in Granulomatous mediastinitis.
In immunocompromised patients, such as those infected
with HIV, disseminated histoplasmosis may occur [64,
149,150]. Initial clinical presentation often includes fever,
anorexia, and weight loss. Subsequent severe disease
manifestation includes sepsis, as well as failure of respi-
ratory, renal, and multi-organ systems.
Mechanisms of pathogenesis
Pathogenicity is the ability of a microbe to cause damage to
the host [151]. The ability of an opportunistic pathogen to
cause disease requires the expression of virulence factors.
Here, we outline some of the traits that enable pathogenesis
of the major opportunistic yeast pathogens of humans, C.
albicans,C. neoformans, and H. capsulatum.
High temperature growth
Fungi must be capable of surviving in human host condi-
tions. One trait absolutely required for pathogenicity in
humans is the ability to grow at human body temperature,
37 °C[14]. The most prevalent fungal pathogen, C. albi-
cans, is a natural commensal of mammals and is thus
intrinsically competent for growth at 37 °C. Despite this
thermal adaptation, the heat shock response has been re-
tained in C. albicans and influences virulence [152],
presumably reflecting the importance of febrile tem-
peratures during systemic infection [153]. In C. albicans,
the heat shock response is governed by the transcription
factor Hsf1, and the molecular chaperones Hsp70 and
Hsp90 [14]. Mutant C. albicans strains containing inactive
Hsf1 are thermosensitive and display reduced virulence in
a murine model of systemic candidiasis [154]. Further-
more, the membrane dynamics of C. albicans are crucial
for sensing changes in temperature, with depletion of
OLE1, encoding a fatty acid desaturase, impairing activa-
tion of Hsf1 [155]. Temperature is a central cue that
enables the morphogenetic transition of C. albicans from
yeast to filamentous growth through complex cellular cir-
cuitry [156].
Although the saprobe C. neoformans is usually found in
environmental locales, the ability to grow at human
physiological temperatures allows it to cause disease.
Several signaling pathways have been identified as essen-
tial for C. neoformans growth at elevated temperature,
including Ras1 and Cdc42 signaling [157], the unfolded
protein response [158], the histone acetyltransferase Gcn5
[159], the cell wall integrity pathway [160], and calcineurin
signaling [19]. In addition, increased expression of C.
neoformans Hsp90 and other heat-shock proteins is ob-
served upon infection of a murine lung [161], suggesting a
role for Hsp90 in adapting to host conditions. Mutants that
are unable to grow at 37 °C are unable to cause disease in
murine models of cryptococcosis.
Like C. neoformans, the dimorphic fungus H. capsula-
tum can be found in the soil. However, H. capsulatum
forms mycelia in the environment and changes to the yeast
form upon inhalation by a mammalian host. The increase in
temperature to 37 °C is key for this transition, and mem-
brane dynamics may be involved in sensing this cue [162].
For example, addition of saturated fatty acids with a con-
current increase in temperature amplifies the H. capsulatum
heat shock response [163].
Adaptation to pH
Opportunistic fungal pathogens must not only adapt to
variations in temperature, but also pH. In the host, the pH
can range from acidic (in the vaginal and gastrointestinal
tracts and the immune cell phagolysosome) to basic (such
as in the blood or saliva) [164]. Therefore, pH is a powerful
signal for entry into the host bloodstream or other mi-
croenvironments. Additionally, environmental pH can lead
to alterations in nutrient bioavailability.
In C. albicans, alkaline conditions result in cleavage and
activation of the transcription factor Rim101, which leads
Opportunistic yeast pathogens
123
to the upregulation of genes required for growth in alkaline
conditions, including genes involved in iron acquisition
and morphogenesis [165]. Additionally, Rim101 regulates
expression of genes with roles in adhesion and modulating
the fungal cell wall [166]. PHR1 and PHR2 are pH-re-
sponsive genes regulated by Rim101, with PHR2 being
required for virulence at acidic infection sites (such as the
vaginal tract), whereas PHR1 is required at alkaline sites of
infection (such as the bloodstream) [167]. Mutants lacking
Rim101 have reduced virulence in a mouse model of sys-
temic candidiasis [168].
Cryptococcus neoformans also uses the Rim101 tran-
scription factor to adapt to varying environmental cues.
Similar to C. albicans,CnRim101 is required for growth in
neutral/alkaline conditions, and also regulates expression
of genes required for iron and metal homeostasis [169].
However, in C. neoformans, Rim101 is also activated by
the cyclic AMP (cAMP)/protein kinase A (PKA) pathway
[169]. Rim101 is required to modulate many important C.
neoformans virulence traits, such as attachment of the
polysaccharide capsule [169], the morphological change to
titan cells [170], and modulation of the host immune sys-
tem [171,172]. Upon phagocytosis by macrophages, C.
neoformans enters into an acidic phagolysosome; as a
facultative intracellular pathogen, C. neoformans is well
adapted for growth in acidic conditions [173].
Nutrient limitation
Key to survival is growth. Organisms are often nutrient
starved within the host and must adapt to changes in
available host nutrients. One mechanism by which C. al-
bicans can respond to limited glucose is by secreting
ammonia and autoalkalizing its surroundings to induce
filamentation [174]. For C. neoformans, phagocytosis by
macrophages induces a starvation stress response that in-
cludes upregulation of amino acid and sugar transporters,
gluconeogenesis and fatty acid metabolism [175].
As iron levels are tightly controlled within the host,
microbes must have mechanisms to acquire this essential
cofactor [176]. Although C. albicans and C. neoformans do
not appear to produce their own siderophores, which are
high-affinity iron-binding proteins, they are capable of
transporting xenosiderophores produced by other organ-
isms [177,178]. Additionally, both C. neoformans and C.
albicans are able to use iron from hemoglobin and ferritin
as iron sources [179,180]. Cryptococcus neoformans uti-
lizes a ferric reductase system involving the iron permease
Cft1 and ferroxidase Cfo1 to acquire iron from host
transferrin [181,182]. Mutants lacking these enzymes
show a reduced virulence in a mouse model of crypto-
coccosis. Iron levels are sensed by a master transcription
factor, Cir1, which directs the expression of genes for iron
acquisition. Cir1 is also involved in regulating growth at
physiological temperature, capsule formation, and melanin
production [183], highlighting the central role of iron
regulation in C. neoformans virulence.
Candida albicans is able to acquire iron from he-
moglobin by first binding hemoglobin to a receptor on the
cell wall and subsequently endocytosing this complex
[184]. The heme oxygenase Hmx1 then releases ferrous
iron [185], carbon monoxide, and biliverdin [186]. As
carbon monoxide has immunosupressive properties [187],
its production by Hmx1 may decrease the ability of the host
immune system to clear a C. albicans infection. The hmx1
mutant is unable to cause disease as this mutant is unable to
grow in iron-limited conditions and has a defect in carbon
monoxide production [186,188]. Additionally, the tran-
scriptional activator Sef1 induces iron-uptake genes and
enables virulence, whereas the transcriptional repressor
Sfu1 diminishes expression of iron-uptake genes and en-
ables gastrointestinal commensalism [116]. These genetic
programs may be fundamental to enabling C. albicans to
survive iron depletion in the bloodstream while minimizing
iron toxicity in the gut.
Histoplasma capsulatum is able to both produce side-
rophores and utilize reductive iron assimilation to obtain
the iron necessary for growth [189]. Histoplasma capsu-
latum produces hydroxyamate-type siderophores, and these
siderophores are required for proliferation within macro-
phages and infection in mice [189,190]. Production of
siderophores is negatively regulated by the GATA-type
transcription factor SRE1 [191]. Histoplasma capsulatum
can also take up ferrioxamine B, a xenosiderophore, via
ferric reductases [192].
Changes in cellular shape and size
Host physiological temperature induces a filament-to-yeast
transition in dimorphic fungi, such as H. capsulatum
(Fig. 1). This switch from mycelia to yeast growth at host
temperatures is a requirement for H. capsulatum virulence,
and the histidine kinase Drk1 is a key regulator of this
transition [193]. DRK1-silenced strains show a drastic re-
duction in virulence in a mouse model of histoplasmosis
[193]. In addition, Drk1 regulates expression of two im-
portant H. capsulatum virulence genes, CBP1 and AGS1
[193]. The DNA-binding protein Ryp1 is also essential for
yeast growth, and is related to the C. albicans master
transcriptional regulator of phenotypic switching, Wor1.
Ryp1 is a master regulator of morphogenesis, as it is re-
quired for the expression of yeast-phase-specific virulence
genes and repression of mycelial-specific genes [194].
Recently, a temperature-responsive regulatory circuit
composed of four Ryp proteins was identified that controls
both the transition from filamentous to yeast forms and the
E. J. Polvi et al.
123
expression of virulence genes such as YPS3 and CBP1
[195].
Strikingly, host physiological temperature induces the
opposite morphological transition in the commensal fun-
gus, C. albicans, from yeast to filament compared to the
filament-to-yeast transition initiated in dimorphic fungi
(Fig. 1). Candida albicans is able to undergo various
morphological transitions that facilitate its ability to colo-
nize and infect different niches within the human host. The
most striking of these morphological transitions is that
from yeast to filamentous forms. Typically, chains or
branches of elongated connected cells are referred to as
pseudohyphae [196], whereas cells containing defined
septa between cells and no constrictions in their cell walls
are known as hyphae [197]. Various different environ-
mental cues induce the morphological transition from yeast
to filaments. Many of these cues require a concurrent in-
crease in temperature to 37 °C, and many of the cues
mimic host physiological conditions. For example, serum
is a potent inducer of filamentation at 37 °C, with key
effectors being bacterial peptidoglycans and glucose [198
200]. At host physiological temperature, deprivation of
nutrients such as carbon or nitrogen can induce a
filamentous program, as can alkaline conditions (pH [6.5)
or elevated CO
2
[201]. The requirement of 37 °C for
filamentation to occur in many cues is due at least in part to
the Hsp90-mediated repression of filamentation, which is
relieved at elevated temperature due to global problems in
protein folding that can overwhelm the functional capacity
of Hsp90 [202]. Many cues induce filamentation in C. al-
bicans via the cAMP–PKA signaling pathway, which
activates transcription factors such as Efg1 [203] and Flo8
[204]. Filamentation is also governed by repressors such as
Nrg1, whose downregulation and degradation is required
for both hyphal initiation and maintenance [205].
The ability of C. albicans to transition between yeast
and filamentous states is key for its dissemination and
virulence in different niches within the host. Invasion of
epithelial cells by C. albicans can occur by either endo-
cytosis or by active penetration by hyphal cells [206]. Cells
locked in the yeast form are reduced in virulence [207,
208], and defective in adhering to and invading oral ep-
ithelial cells [209]. Similarly, cells unable to revert to the
yeast form also have attenuated virulence and show a de-
crease in kidney fungal burden in a murine model of
systemic candidiasis [210]. It is thought that yeast cells are
Fig. 1 Yeast virulence traits. Top,C. albicans can exist as yeast,
filaments, or as a biofilm. Wild-type yeast cells were grown in rich
medium at 30 °C for 24 h. Filaments were grown in RPMI at 37 °C
for 24 h. The biofilm is fluconazole treated and was formed in a rat
central venous catheter. Scale bar is 20 lm. Middle,C. neoformans
cells stained with India ink to highlight the capsule. Arrow indicates
the size of a titan cell as compared to a normal cell. Scale bar 10 lm.
Bottom,H. capsulatum can exist as either yeast or as mycelia. Scale
bar is 5 lm
Opportunistic yeast pathogens
123
required for dissemination throughout the body, while
filamentous cells are essential for penetrating tissues,
establishing infection and causing mortality [211]. A recent
study has shown that a clinical isolate locked in the yeast
form is more fit in a murine model of commensal infection
[212]. Furthermore, another recent study demonstrated that
cells locked in the yeast form are better able to colonize the
gastrointestinal tract in a mouse model, while constitu-
tively filamentous cells showed a decrease in colonization
[213]. Taken together, this demonstrates that the poly-
morphic nature of C. albicans is a key virulence trait that
allows it to adapt to the diverse conditions encountered
within the host.
While both yeast and filamentous morphologies con-
tribute to the pathogenesis of C. albicans, filamentous cells
do possess specific virulence characteristics that increase
their pathogenicity. For example, the sustained polarized
growth exhibited by a filamentous cell provides a physical
force that allows the fungal cell to penetrate host tissues,
allowing for further invasion of tissues and organs [214].
Hyphal cells also express Als3, a hypha-specific protein
found on the fungal cell surface [215]. Als3 is an adhesin
that mediates binding to host cells, and induces endocytosis
by binding to the host cell receptors E-cadherin and
N-cadherin [215]. Hwp1 is another hyphal-specific fungal
cell surface protein required to mediate attachment to ep-
ithelial cells [216]. In addition, the expression of genes
encoding secreted aspartyl proteinases (SAPs) is coordi-
nated with morphological state with SAP4 and SAP6 being
expressed by hyphal cells [217,218]. Additionally, con-
nections exist between C. albicans filamentation and drug
resistance. Certain proteins such as Hsp90 [219], O-man-
nosyltransferases [220], and the transcription factor Ndt80
[221] have roles in both morphogenesis and drug
resistance.
In addition to the striking morphological differences
between yeast and filamentous forms, C. albicans can
also undergo distinct phenotypic switches. One prominent
example is the switch between white and opaque cells.
The more common white cells have a round or oval
shape, while opaque cells are oblong and dimpled. This
transition is governed by the regulator Wor1, with ex-
pression of WOR1 leading to opaque cell formation
[222]. Opaque cells are mating competent, with progeny
showing recombination between chromosomes and alter-
ations in ploidy [223,224]. White and opaque cells show
distinct preferences for different host environments and
niches. Since opaque cells are not stable at 37 °C, they
preferentially colonize surface environments such as the
skin [225]. Furthermore, while white cells require the
elevated temperature of 37 °C for filamentation, opaque
cells filament preferentially at 25 °C[226]. Finally,
white-opaque switching may aid in the evasion of host
defenses, where white phase cells are preferentially rec-
ognized by neutrophils in certain environmental
conditions [227,228]. A novel phenotypic switch was
recently discovered upon passage of wild-type white cells
through the mouse gastrointestinal tract [117]. Induction
of WOR1 in this environment resulted in morphologically
and functionally distinct GUT (gastrointestinally induced
transition) cells, which are optimized for the digestive
tract. A third phenotypic variant, the ‘gray’ phenotype
has recently been identified [229]. This variant shows
elevated levels of SAP expression and the propensity to
colonize cutaneous tissue, similar to opaque cells. The
master regulators Wor1 and Efg1 may act coordinately to
regulate switching between white, gray and opaque cel-
lular states [229].
Distinct morphological transitions are also important
for virulence in C. neoformans (Fig. 1). While a typical
C. neoformans cell is approximately 5–10 lM, specific
conditions can induce the formation of titan cells, which
can be as large as 100 lM[230]. Distinct features of titan
cells include a thicker cell wall and denser capsule [231],
as well as a tetraploid or octaploid DNA content [230].
Although the exact mechanism for titan cell formation is
unclear, in part due to the difficulty in culturing titan cells
in vitro [232], several signaling pathways have been im-
plicated. The G protein-coupled receptors Ste3a (a
pheromone receptor) and Gpr5 are required for titan cell
production in vivo [170,230] as is signaling through the
cAMP/PKA pathway [170,231]. The Rim101 transcrip-
tion factor downstream of PKA signaling is also required
for titan cell production in vivo [170]. Titan cells play a
central role in C. neoformans pathogenicity, in that they
interact with the host immune system. The large size of
titan cells may prevent their phagocytosis [231,233]; titan
cells may also protect typical-sized cells from being
phagocytosed [233], and promote dissemination from the
lungs [234].
Biofilms
Biofilms are complex three-dimensional surface-associat-
ed communities of yeast and hyphal cells within an
extracellular matrix (ECM), and can be found on medical
devices such as catheters and artificial joints [235], or on
mucosal surfaces [236], contributing to virulence (Fig. 1).
The first step in biofilm formation is adherence to either
an abiotic or biotic surface. In C. albicans, cell wall ad-
hesins such as Eap1 and Als1, and other cell wall proteins
have roles in attachment of the biofilm to the surface,
which then stimulates changes in gene expression [237].
Adherence is followed by proliferation of yeast cells. This
basal layer of yeast cells may contribute to anchoring the
biofilm to the surface [238]. Next, yeast cells undergo the
E. J. Polvi et al.
123
morphogenetic transition to filamentous growth. The for-
mation of hyphae is important, as strains defective in
filamentation have a vastly reduced ability to form bio-
films, as with mutants lacking a major transcriptional
regulator of morphogenesis, Efg1 [238]. Accumulation of
the ECM is part of the biofilm maturation process. The
ECM is composed of mainly protein and carbohydrates,
including mannans and glucans; however, lipids and nu-
cleic acids are also found in the ECM [239]. Finally,
yeast cells are dispersed from the biofilm. These cells
have elevated pathogenicity, and an increased ability to
adhere [237].
Biofilms are notoriously difficult to treat with anti-
fungals, and have especially high levels of resistance to
azoles and polyenes [237]. Resistance to azoles is de-
pendent on the molecular chaperone Hsp90 [45], glucan
modification enzymes that are required for delivery and
organization of the ECM [240], and b-1,3 glucan in the
ECM that may sequester azoles [241]. Increased expres-
sion of efflux pumps early in biofilm development, and
the presence of persister cells also contribute to antifungal
resistance [242].
Secreted factors
Secretion of proteinases such as SAPs is crucial for both
pathogenesis and nutrient acquisition. As mentioned
above, C. albicans SAP4 and SAP6 expression is coor-
dinated with hyphal formation [217,218], while
expression of SAP1 is regulated by the white-to-opaque
phenotypic switch [243]. During infection, the SAP pro-
teins have roles in adherence to host cells [244] and
degradation of host proteins such as mucin [245], perhaps
allowing penetration and colonization of tissues. SAP
proteins may also degrade host antimicrobial factors such
as lactoferrin, complement and the proteinase inhibitor
cystatin A [244]. Secreted phospholipases cleave ester
linkages in glycophospholipids, and can have roles in host
cell penetration and invasion [246]. Mutants lacking the
phospholipase Plb1 have reduced virulence in a murine
model of Candida infection [247].
Histoplasma also secretes several factors important for
pathogenesis. The calcium-binding protein Cbp1 is se-
creted from yeast-phase H. capsulatum, and is required for
growth in calcium-depleted conditions and for virulence in
a murine model of pulmonary histoplasmosis [248].
Resolution of the structure by NRM revealed that Cbp1
may be a lipid-binding protein, and possibly interacts with
glycolipids in the host [249]. Yps3 is both a secreted factor
as well as a cell wall protein in H. capsulatum, and while
its exact function is yet to be elucidated, silencing of YPS3
results in reduced organ colonization in a murine model of
infection [250].
Evasion of the host immune system
Capsule and cell wall
The capsule of C. neoformans is an important virulence
determinant, playing key roles in surviving environ-
mental conditions and modulating the host immune
response (Fig. 1). The capsule is primarily composed of
the polysaccharides glucuronoxylomannan (GXM) and
glucuronoxylomannogalactan (GXMGal) [251]. Mono-
mers of simple sugars are modified and polymerize
within the cell before being secreted and attached to the
cell wall surface. An important modification of capsule
monomers is xylosylation, such that deletion of the
xylosyltransferase gene CXT1 results in an altered cap-
sule structure and reduced virulence [252].
O-acetylation contributes to the antigenicity of the
capsule by affecting binding of antibodies and activation
of the complement cascade [253]. Hyaluronic acid,
synthesized by Cps1, is also added to the capsule and
may play a role in facilitating passage of C. neoformans
across the blood–brain barrier [254]. Finally, antigenic
variation of capsule monomers can lead to differential
binding of antibodies [255].
Just as conditions that mimic physiologically relevant
host conditions induce filamentation in C. albicans,
similar cues also induce capsule formation in C.
neoformans. Upon infection of a host, the capsule in-
creases in size with the thickness being determined by
the location of infection. Infection of the lungs results in
a thicker capsule than does infection of the brain [256],
demonstrating that regulation of capsule synthesis is
dependent on the environmental niche. Nutrient-poor
conditions such as low glucose and low nitrogen can
contribute to capsule induction by signaling through the
cAMP–PKA pathway [257259]. Similarly, as iron
levels are tightly controlled within the host, conditions
of low iron are a potent inducer of capsule formation
[260]. The iron-sensing transcription factor Cir1 is one
of the main regulators of capsule induction in response
to iron poor conditions [183]. Induction of capsule for-
mation in low iron conditions is also signaled through
the cAMP–PKA pathway [251]. Additionally, the cAMP
pathway is used to signal capsule formation in the
presence of high CO
2
levels [251,261]. Physiological
pH is another cue that is sensed by C. neoformans to
regulate capsule formation, and the conserved Rim101
pH-sensing pathway integrates signals from the cAMP–
PKA pathway to regulate attachment of the capsule to
the fungal cell wall [169]. However, it appears that al-
kaline pH alone is insufficient to induce an enlarged
capsule, and instead this condition must be combined
with others, such as nutrient deprivation or the presence
Opportunistic yeast pathogens
123
of serum [262]. Finally, stress response pathways can
havearepressiveeffectoncapsule induction, such as
the Hog1 and protein kinase C (PKC) pathways that
respond to osmotic stress [251].
The capsule is essential for C. neoformans interactions
with the host immune system. The polysaccharides GXM
and GXMGal have immunosuppressive properties such as
modulation of macrophage, neutrophil and dendritic cell
activities, and also inhibition of pro-inflammatory cytokine
production [263]. GXMGal reduces B cell activity and
inhibits activation of T cells. The capsule also has an-
tiphagocytic properties, inhibiting phagocytosis in vitro
due to the large size of the capsule, and possibly by hiding
pathogen-associated molecular patterns (PAMPs) on the
cell surface [264].
To avoid interaction with host phagocytes that would
initiate the production of reactive oxygen species and the
secretion of pro-inflammatory cytokines, H. capsulatum
antigenic cell surface bglucans are hidden under a layer of
a-(1,3)-glucan. Silencing of AGS1, encoding the a-glucan
synthase, severely attenuates virulence of H. capsulatum in
mouse lungs [265]. Loss of a-(1,3)-glucan allows for
recognition of the bglucans by the host receptor dectin-1
[266]. Additionally, heat shock protein 60 (HSP60) on the
cell surface binds to the CR3 receptor on macrophages,
followed by phagocytosis [267]. However, interactions
with CR3 do not typically result in a strong immune re-
sponse unless other costimulatory signals are present [268,
269], allowing H. capsulatum to grow and survive within
macrophages.
Escape from macrophages
Fungi have evolved multiple mechanisms to avoid being
killed upon phagocytosis by immune cells such as mac-
rophages. One mechanism is to escape the phagocyte,
either by inducing macrophage lysis or by non-lytic
exocytosis. Non-lytic expulsion of fungal cells was first
described in C. neoformans, where live yeasts were able
to escape macrophages without killing the host cell [270
272]. This prevents release of pro-inflammatory cytokines
and potentially aids in transversal of cryptococcal cells
across the blood–brain barrier [105]. This phenomenon of
non-lytic expulsion has also been observed in C. albicans,
although at lower frequency than for C. neoformans
[273]. Candida albicans can also escape from macro-
phages via filamentation. These filaments can induce
pyroptosis, a programmed pro-inflammatory macrophage
death [274,275]. This process is hypothesized to be de-
pendent on the transition to hyphae, which stimulate host
pyroptotic caspases and result in macrophage death [274,
275].
Nitric oxide detoxification
Nitric oxide (NO) is a nitrogen radical and antimicrobial
effector produced by the host immune system in response
to fungal infections. Candida albicans responds to this
stress by increasing the expression of an NO dioxygenase
flavohemoglobin, YHB1 [276,277], which metabolizes
NO. Candida albicans cells lacking YHB1 have increased
sensitivity to NO and reduced virulence in a murine model
of systemic candidiasis [276,277].
Cryptococcus neoformans also employs an enzymatic
defense against NO produced by the host. The flavohe-
moglobin denitrosylase Fhb1 consumes NO, and cells in
which FHB1 is disrupted have reduced survival in mac-
rophages and decreased virulence in a murine inhalation
model [278]. A global analysis of the C. neoformans re-
sponse to nitrosative stress revealed that the glutathione
reductase Glr1 is upregulated in response to NO [279].
Deletion of GLR1 renders C. neoformans sensitive to ni-
trosative stress, and avirulent in a mouse inhalation model
[279].
Although the presence of a flavohemoglobin in H.
capsulatum is not apparent, a shotgun genomic microarray
did identify NOR1, a P450 nitric oxide reductase homo-
logue, whose expression is increased by nitrosative stress
[280]. Overexpression of this gene decreased the sensitivity
of H. capsulatum to reactive nitrogen species [280].
Melanin
In C. neoformans, melanins are pigmented molecules with
antioxidant properties, synthesized by the laccase enzymes
Lac1 and Lac2 [264]. Melanization occurs during infection
[281], and overexpression of melanin results in decreased
recognition by the host and modulation of host cell immune
responses [282]. Melanized cells are also less susceptible to
microbicidal peptides and recognition by macrophages, as
well as being less susceptible to antifugal drugs in vitro
[283]. Together, this demonstrates an important role for
melanin in the virulence of C. neoformans.
Urea metabolism
In addition to being used as a nitrogen source by many
fungi, C. albicans and C. neoformans also use urea meta-
bolism to modulate the host immune system and alter
dissemination [284]. Urease is considered a cryptococcal
virulence factor as urease expression is linked with in-
creased invasion across the blood–brain barrier [285],
increased fungal burden, and altered host immune re-
sponses [286288]. Urease-positive C. neoformans induce
a strong non-protective Th2 response as demonstrated by
E. J. Polvi et al.
123
an increased accumulation of eosinophils in the lung, in-
creased serum IgE and higher levels of Th2 cytokines
[288], demonstrating a role for urease in influencing the
host immune response.
While C. albicans does not contain a urease enzyme, it
does contain a urea amidolyase, encoded by DUR1,2 [289].
Deletion of DUR1,2 results in cells that are less virulent
than wild-type cells in a murine model of disseminated
candidiasis and results in decreased kidney fungal burden
[290]. Additionally, deletion of DUR1,2 causes a decreased
inflammatory response in the kidneys, reduced neutrophil
infiltration and altered host cytokine and chemokine pro-
duction, suggesting a role for urea metabolism in
modulating host immune responses [290]. A dur1,2 mutant
also shows impaired escape from host macrophages as it is
unable to form filaments in the presence of urea or argi-
nine, a urea-precursor [291]. This may be coupled to the
defect of the dur1,2 strain in auto-alkalinization of the
environment [174].
Host immune responses
The interactions between opportunistic fungal infections
and the host immune system are complex and only be-
ginning to be understood. Outcomes of infections are
dependent on interactions between the host immune re-
sponse and the intrinsic virulence of the pathogen.
Opportunistic fungi such as C. albicans constantly in-
teract with the host immune system, and perturbations to
this balance can result in a switch from commensalism
to pathogenesis, as discussed above. Alterations to the
innate or adaptive host immune responses, either due to
immunosuppression or genetic predisposition, can result
in severe and often deadly fungal infections. The inter-
play between fungi and the host immune system has
been the subject of several recent reviews [8,292294].
Here we summarize this interaction, and describe host
factors that can increase susceptibility to fungal
infections.
Host response to fungal pathogens
Innate
Physical barriers such as the skin and mucosal epithelial
cells at sites continuously in contact with fungal pathogens
are the first line of defense against fungal pathogens. The
next barrier is the innate immune response, which is crucial
for preventing invasive and systemic infections. The innate
immune response is required for recognition of fungal
PAMPs. The fungal cell wall is composed of many
PAMPs, including both a- and b-glucans, chitin, chitosan,
and mannans [295]. These PAMPs can be recognized by
host pattern recognition receptors (PRRs) on phagocytic
cells such as macrophages, dendritic cells and neutrophils
[292,293]. Specific PAMPs on different fungal species can
be recognized by specific PRRs, including complement
components, toll-like receptors (TLRs), C-type lectin re-
ceptors (CLRs), and mannose receptors (MR), among
others [296299]. These interactions will shape the re-
sulting immune response, including phagocytosis of the
pathogen, production of inflammatory cytokines, and acti-
vation of T cells.
Upon phagocytosis, fungal cells can be killed by ef-
fector molecules such as proteases, defensins, and cationic
peptides as well as by the production of reactive oxygen
and nitrogen species [294]. Neutrophils are highly effective
at containing fungal infections, including inhibition of the
C. albicans morphological transition [300]. Neutropenic
patients, including those with acute leukemia, are at high
risk of developing invasive Candida and Aspergillus in-
fections, and are often given prophylactic antifungal
therapies as a preventative measure [301]. However, in-
creased neutrophils in mouse kidneys are associated with
pathogenesis of C. albicans at late time points [302].
Adaptive
The adaptive immune system involves stimulation of T
cells by antigen-presenting cells [292,303]. These antigen-
presenting cells can generate specific cytokine profiles that
can bias the immune system towards Th1-, Th2-, Treg-, or
Th17-responses [303]. The balance between these re-
sponses is crucial for the ability to clear the pathogen
without causing autoimmune damage.
The adaptive immune response plays a pivotal role in
preventing fungal infections, as highlighted by the inci-
dence of fungal disease in AIDS patients [10,304]. HIV
infection depletes the pool of CD4
?
T cells, consequently
causing a loss of antifungal immunity. This results in im-
paired production of interferon-c(IFN-c) and tumor
necrosis factor-a(TNF-a)[304]. Additionally, the
population of memory B cells is depleted [305], and
macrophage and dendritic cell function is impaired upon
HIV infection [306], contributing to increased suscepti-
bility to fungal infections.
Th1 responses include the production of IFN-c, which is
key for controlling many disseminated fungal infections
[303]. Cryptococcus neoformans strains that induce Th1-
biased responses are generally cleared from the lung and do
not cause disease [307,308]. The decrease in Th1-type
cell-mediated immune responses in AIDS patients may
explain their increased susceptibility to cryptococcal in-
fections compared to patients with other types of immune
deficiencies [309].
Opportunistic yeast pathogens
123
Th17 responses appear to be critical for host defense
against fungal infections, especially Candida [310]. In HIV
patients, susceptibility to mucosal candidiasis can be at-
tributed to a loss of mucosal Th17 cells, which is
accompanied by a decrease in the integrity of epithelial
cells and alterations to the intestinal environment [311].
Decreases in the Th17 response also leads to increased
severity of Pneumocystis infections in mice [312].
Regulatory T (Treg) cells continually survey the host for
signs of mucosal infections and aid in preventing reinfec-
tion. However, Treg cells are able to suppress inflammation
during systemic candidiasis by producing interleukin 4 (IL-
4), IL-10, and transforming growth factor-b(TGF-b),
which inhibit inflammatory Th1 and Th17 responses
making them detrimental for clearing the infection [313].
Depletion of Treg cells results in decreased C. albicans
growth in a murine model of candidiasis [313], and a de-
crease in Treg cells is also associated with an increase in
the Th17 responses that are beneficial in clearing Histo-
plasma infections [314].
In contrast, Th2 responses to fungal infections are often
deleterious [292]. During C. neoformans infections, the
fungi are able to induce a strong Th2-biased immune re-
sponse, with increased levels of IL-4 and IL-5, thus
favoring persistence of the infection [309]. Mice with in-
creased IL-4 levels also demonstrated increased
Histoplasma infections [315]. In addition to favoring per-
sistence of the infection, Th2 responses are also linked with
allergic bronchopulmonary aspergillosis and increased
susceptibility to invasive aspergillosis [316]. In contrast,
Th2 responses may be protective against Pneumocystis
pneumonia, with affected HIV individuals showing low
levels of Th2 cytokines [317].
Immune reconstitution inflammatory syndrome
The recovered immune system in AIDS patients treated
with antiretrovirals can overreact to PAMPs exposed dur-
ing fungal infections, even if the infections have been
cleared by antifungal therapies. This overactive inflam-
mation is known as immune reconstitution inflammatory
syndrome (IRIS), and it can cause significant mortality due
to damage by the host immune system. IRIS is a major
consideration in the treatment of AIDS patients with
cryptococcosis. Recent studies have noted that between 8
and 50 % of AIDS patients develop cryptococcal IRIS,
even when no live cryptococcal cells are recovered from
the patient [35,318]. IRIS can also occur in transplant
patients after reduction in the immunosuppresive therapies
used to prevent organ rejection [319]. The inflammatory
immune response in these IRIS patients consists of in-
creased IL-6 and C-reactive protein (CRP) levels.
Additionally, many patients who go on to develop IRIS
have decreased TNF-aand other Th1 cytokines prior to
antiretroviral therapy [320]. The current standard of care
for IRIS patients includes administration of steroids to
decrease systemic inflammatory responses and prevent
further damage to host tissues [36,90].
Genetic susceptibility to fungal infections
Chronic mucocutaneous candidiasis
Recurrent Candida infections of cutaneous or mucosal
surfaces in the absence of immunosuppressive conditions
(such as HIV), is described as chronic mucocutaneous
candidiasis (CMC). IL-17 signaling and activation of T
cells are essential for defending against mucocutaneous C.
albicans infections and genetic mutations that perturb this
signaling pathway can lead to CMC [321].
Numerous single nucleotide polymorphisms have been
associated with increased susceptibility to CMC (see [8,
292,321]). For example, loss of function of the CLR
dectin-1 can lead to increased onychomycosis and muco-
cutaneous candidiasis [322,323]. This is the result of
defective recognition of bglucan in the C. albicans cell
wall and impaired cytokine production (including IL-1b
[323] and IL-17 [322]). Neutrophil function remains
unaffected, and thus systemic candidiasis is not associated
with dectin-1 mutations [323]. The caspase recruitment
domain-containing protein 9 (CARD9) is a signaling pro-
tein downstream of many CLRs, including dectin-1 [321]
and activates the nuclear factor-jB (NF-jB) pathway
[324]. Autosomal recessive mutations in CARD9 can result
in decreased development of Th17 cells and is associated
with CMC [324]. Autosomal dominant gain-of-function
mutations such as those in STAT1 (signal transducer and
activator of transcription 1) can also lead to inhibition of
Th17 development, and increased susceptibility to CMC
[325,326]. Additionally, autosomal recessive mutations in
the IL-17 receptor IL-17RA, and deficiencies in the cy-
tokine IL-17F can result in CMC [327]. Mice with deficient
IL-17A production show an impaired ability to clear a
cutaneous C. albicans infection [328], suggesting that both
IL-17A and IL-17F are required for the defense against
chronic mucocutaneous candidiasis.
Additional syndromes can lead to CMC. One of these is
hyper-IgE syndrome (HIES), in which patients experience
recurrent pulmonary infections, eczema, staphylococcal
infections, and CMC, among other symptoms. Mutations in
STAT3 have been identified as a cause of HIES [329], in
which an inability of CD4
?
T lymphocytes to differentiate
into Th17 cells results in impaired IL-17 production [330].
Mutations in DOCK8 (dedicator of cytokinesis 8) have been
identified in the autosomal recessive form of this syndrome,
in which there is defective T cell activation and Th17 cell
E. J. Polvi et al.
123
differentiation [331]. Finally, the rare autoimmune
polyendocrinopathy–candidiasis–ectodermal dystrophy
(APECED) syndrome is characterized by development of
CMC in almost all patients, and is caused by mutations in
the autoimmune regulator AIRE [321,332]. The strong
prevalence of CMC in these patients may be due to the
production of neutralizing autoantibodies against Th17-
produced cytokines, specifically IL-17A, IL-17F and IL-22
[333].
Invasive fungal infections
While mucocutaneous Candida infections are primarily
caused by defects in the adaptive immune system, defi-
ciencies in the innate immune response, such as those
caused by primary immunodefiencies can lead to invasive
yeast infections. A few such cases are outlined below.
Chronic granulomatous disease (CGD) is the result of
mutations in genes encoding components of the NADPH
oxidase complex, impairing the ability of phagocytes to
produce reactive oxygen species [334]. While the primary
fungal infection associated with CGD is invasive
aspergillosis, [335] infections by Candida and Tri-
chosporon species have also been noted [321].
Myeloperoxidase (MPO) deficiency is a disorder in which
the production of antimicrobial hypochlorous acid by
phagocytic cells is impaired, and mice deficient for either
NADPH or MPO show increased mortality in response to a
high dose of C. albicans [336]. While the majority of MPO
patients do not suffer from fungal infections, those who do
acquire infections usually also have diabetes [337]. Patients
presenting with monocytopenia are more susceptible to
histoplasmosis and cryptococcal infections [338], and
cryptococcosis has also been observed in patients with
hyper-IgE or hyper-IgM syndromes and idiopathic CD4
lymphocytopenia [321]. Finally, patients with severe
combined immunodeficiency commonly acquire Pneumo-
cystis pneumonia, as do those with hyper-IgM syndrome
[321].
Management of fungal infections
Diagnosis
The first step in achieving the proper management of in-
fectious disease is the effective diagnosis and species
identification of the pathogen. Unfortunately, shortcomings
in current diagnostic techniques remain one of the greatest
challenges in the field. In most cases, diagnosis still relies
on traditional culture methods and histopathology [10,87,
339,340]. Culturing the organism and subsequent identi-
fication is a slow process that takes several days, leading to
considerable delays in treatment. Such delays have direct
and significant impact on disease outcome and patient
mortality, especially for severe invasive diseases [136,
340]. A number of novel molecular diagnostic approaches
have been developed to improve the current system, in-
cluding PCR-based assays and antigen detection systems
[339,341]. However, these are not regular practice in the
clinic, and still need to be standardized and tested in large
patient cohorts before they can be incorporated into clinical
guidelines [87,90,110,339].
Antifungal drugs
Due to the close evolutionary relatedness between humans
and fungi, the number of targets that can be selectively
exploited for drug development is limited [10,342]. There
are also very few drugs currently being developed, pri-
marily because antifungals are not predicted to generate a
large enough financial return for pharmaceutical companies
[10]. Polyenes, azoles, and echinocandins represent the
three most common classes of antifungals currently used in
the clinic, each with their own advantages and limitations.
Overall, host toxicity, cross reactivity with other drugs, and
development of drug resistance pose great challenges to
current antifungals.
Polyenes
Polyenes are broad-spectrum natural product antifungals
discovered in the early 1950s from the bacterial genus
Streptomyces [342,343]. Polyenes bind to ergosterol in the
fungal membrane, generating aqueous pores that result in
leakage of fungal cell content and eventually cell death
[344,345]. Amphotericin B, one of the most successful
polyene derivatives, was a gold standard in treating serious
fungal infections. However, amphotericin B is known to
cause severe systemic toxicity and nephrotoxicity [342,
346]. Various drug delivery systems have been developed
to improve its safety profile, including the popular lipid
formulation [347,348]. While its clinical use decreased
with the development of azoles in the late 1980s, it is still
widely deployed to treat life-threatening disseminated and
invasive mycoses [90,110]. Fortunately, despite its long
history of clinical use, resistance to polyenes remains a rare
occurrence [349,350].
Azoles
Azoles are synthetic compounds that were first introduced
as antifungals in the late 1980s and early 1990s [344].
Their low toxicity led to extensive use in the clinic. Azoles
disrupt the biosynthesis of ergosterol by inhibiting the
cytochrome P-450-dependent enzyme lanosterol
Opportunistic yeast pathogens
123
demethylase (also referred to as 14a-sterol demethylase)
[351]. Azoles are chemically classified as either imidazoles
if they have two nitrogen atoms in the azole ring, or tria-
zoles if they have three [343,351]. Imidazoles are typically
limited to topical treatment of superficial infections, while
triazoles are used more broadly in both superficial and
systemic infections due to their superior pharmacokinetic
and safety profile. The most commonly used azoles in the
clinic include fluconazole, itraconazole, voriconazole, and
posaconazole.
Azoles remain the drug of choice as initial therapy for
most fungal infections and are often recommended as
prophylaxis for high-risk patients [90,110,352]. Wide-
spread use of azoles has led to increasing reports of azole
resistance in the clinic, which is associated with greater
treatment difficulties and patient mortality [353]. There has
also been an increased incidence of infections caused by
intrinsically azole-resistant fungal species, including C.
glabrata and C. krusei, creating major challenges for future
treatments [87].
Echinocandins
Echinocandins first entered the market in 2001, represent-
ing the newest class of antifungals to reach the clinic [354].
Echinocandins are large semi-synthetic lipopeptides that
inhibit the cell wall enzyme complex b-1,3-D-glucan syn-
thase [355], thereby decreasing the concentration of b-
(1,3)-glucan in the fungal cell wall and causing the sub-
sequent loss of cell wall integrity. Echinocandins are
generally well tolerated with little to no side effects [356].
Echinocandins are active against Candida and Asper-
gillus species, including azole-resistant Candida isolates,
but show no in vivo activity against Cryptococcus [90,110,
342]. Currently available echinocandins include caspo-
fungin, anidulafungin, and micafungin [354]. All three are
not orally bioavailable and need to be administered intra-
venously. Echinocandin resistance has been reported in the
clinic, and the incidence appears to be on the rise [357].
Vaccines
An ideal solution to fungal management is the prevention
of disease through vaccination. However, there is currently
no clinically available vaccine for any fungal pathogen.
Over the years, a number of vaccines have been in devel-
opment against different fungi, with promising results in
animal models [10,358]. However, only a few have been
translated to human clinical trials, with the major barrier
being the lack of funding [359]. Efficacy trials are also
difficult to conduct as high-risk patients often routinely
receive antifungal prophylaxis [10]. Nonetheless, two
subunit vaccines against Candida have recently
demonstrated success in Phase I clinical trials, one of
which is currently being tested in a Phase II trial in the
United States [360]. This vaccine represents a promising
advance and may stimulate further development of much
needed immunotherapeutics and vaccines [358,361,362].
Concluding remarks
A significant advance in our understanding of invasive
fungal infections has been made in recent years, motivated
at least in part by the increasing incidence of these infec-
tions. Given the growing number of transplant patients,
those in receipt of immunosuppressive therapy and the
global HIV pandemic, this comes as no surprise. Host
niches are complex, dynamic and distinct to each indi-
vidual, as are the pathogens’ response strategies for
coexisting with or evading the host immune response.
Taking steps towards understanding the underlying im-
munopathogenesis of fungal infections permits early and
accurate diagnosis and treatment. We must continue to
seek out novel diagnostics that will allow for rapid treat-
ment of fungal infections with the appropriate antifungal.
With the rise in antifungal drug resistance now a real
threat, researchers must strive to discover novel antifungals
and antifungal drug combinations, as well as effective
immunotherapeutic strategies that combat resistance, im-
prove patient outcome, shorten hospital stays and ease the
economic burden.
Acknowledgments We thank the J. Andrew Alspaugh and Chad
Rappleye labs for images and Cowen lab members for helpful dis-
cussions. EJP is supported by a Canadian Institutes of Health
Research (CIHR) Frederick Banting and Charles Best CGS Doctoral
Award, XL by a University of Toronto Fellowship, MDL by a Sir
Henry Wellcome Postdoctoral Fellowship (Wellcome Trust 096072),
and LEC by a Ministry of Research and Innovation Early Researcher
Award, Canada Research Chair in Microbial Genomics and Infectious
Disease, Natural Sciences and Engineering Research Council Dis-
covery Grant #355965, and by Canadian Institutes of Health Research
Grants MOP-86452 and MOP-119520.
References
1. Wang DY, Kumar S, Hedges SB (1999) Divergence time esti-
mates for the early history of animal phyla and the origin of
plants, animals and fungi. Proc Biol Sci 266:163–171
2. Butterfield NJ (2005) Probable proterozoic fungi. Paleobiology
31:165–182
3. Hawksworth DL (2001) The magnitude of fungal diversity: the
1.5 million species estimate revisited. Mycol Res 105(12):1422–
1432
4. DiGuistini S, Wang Y, Liao NY, Taylor G, Tanguay P, Feau N,
Henrissat B, Chan SK, Hesse-Orce U, Alamouti SM, Tsui CK,
Docking RT, Levasseur A, Haridas S, Robertson G, Birol I, Holt
RA, Marra MA, Hamelin RC, Hirst M, Jones SJ, Bohlmann J,
Breuil C (2011) Genome and transcriptome analyses of the
E. J. Polvi et al.
123
mountain pine beetle-fungal symbiont Grosmannia clavigera,a
lodgepole pine pathogen. Proc Natl Acad Sci USA 108:2504–
2509
5. Fisher MC, Henk DA, Briggs CJ, Brownstein JS, Madoff LC,
McCraw SL, Gurr SJ (2012) Emerging fungal threats to animal,
plant and ecosystem health. Nature 484:186–194
6. Taylor LH, Latham SM, Woolhouse ME (2001) Risk factors for
human disease emergence. Philos Trans R Soc Lond B Biol Sci
356:983–989
7. Findley K, Oh J, Yang J, Conlan S, Deming C, Meyer JA,
Schoenfeld D, Nomicos E, Park M, Program NIHISCCS, Kong
HH, Segre JA (2013) Topographic diversity of fungal and
bacterial communities in human skin. Nature 498:367–370
8. Underhill DM, Iliev ID (2014) The mycobiota: interactions
between commensal fungi and the host immune system. Nat Rev
Immunol 14:405–416
9. Bickers DR, Lim HW, Margolis D, Weinstock MA, Goodman
C, Faulkner E, Gould C, Gemmen E, Dall T, American
Academy of Dermatology A, Society for Investigative D (2006)
The burden of skin diseases: 2004 a joint project of the Amer-
ican Academy of Dermatology Association and the Society for
Investigative Dermatology. J Am Acad Dermatol 55:490–500
10. Brown GD, Denning DW, Gow NA, Levitz SM, Netea MG,
White TC (2012) Hidden killers: human fungal infections. Sci
Transl Med 4:165rv13
11. Pfaller MA, Diekema DJ (2010) Epidemiology of invasive
mycoses in North America. Crit Rev Microbiol 36:1–53
12. Brown GD, Denning DW, Levitz SM (2012) Tackling human
fungal infections. Science 336:647
13. Bergman A, Casadevall A (2010) Mammalian endothermy op-
timally restricts fungi and metabolic costs. MBio 1(5):e00212-
10
14. Leach MD, Cowen LE (2013) Surviving the heat of the moment:
a fungal pathogens perspective. PLoS Pathog 9:e1003163
15. Garcia-Solache MA, Casadevall A (2010) Global warming will
bring new fungal diseases for mammals. MBio 1(1):e00061-10
16. McCusker JH, Clemons KV, Stevens DA, Davis RW (1994)
Saccharomyces cerevisiae virulence phenotype as determined
with CD-1 mice is associated with the ability to grow at 42 °C
and form pseudohyphae. Infect Immun 62:5447–5455
17. Bhabhra R, Miley MD, Mylonakis E, Boettner D, Fortwendel J,
Panepinto JC, Postow M, Rhodes JC, Askew DS (2004) Dis-
ruption of the Aspergillus fumigatus gene encoding nucleolar
protein CgrA impairs thermotolerant growth and reduces viru-
lence. Infect Immun 72:4731–4740
18. Lamoth F, Juvvadi PR, Fortwendel JR, Steinbach WJ (2012)
Heat shock protein 90 is required for conidiation and cell wall
integrity in Aspergillus fumigatus. Eukaryot Cell 11:1324–1332
19. Odom A, Muir S, Lim E, Toffaletti DL, Perfect J, Heitman J
(1997) Calcineurin is required for virulence of Cryptococcus
neoformans. EMBO J 16:2576–2589
20. Havlickova B, Czaika VA, Friedrich M (2008) Epidemiological
trends in skin mycoses worldwide. Mycoses 51:2–15
21. Sobel JD (2007) Vulvovaginal candidosis. Lancet 369:1961–
1971
22. Marques SA, Robles AM, Tortorano AM, Tuculet MA, Negroni
R, Mendes RP (2000) Mycoses associated with AIDS in the
Third World. Med Mycol 38:269–279
23. Park BJ, Wannemuehler KA, Marston BJ, Govender N, Pappas
PG, Chiller TM (2009) Estimation of the current global burden
of cryptococcal meningitis among persons living with HIV/
AIDS. AIDS 23(4):525–530
24. Goldman DL, Khine H, Abadi J, Lindenberg DJ, Pirofski LA,
Niang R, Casadevall A (2001) Serologic evidence for Crypto-
coccus neoformans infection in early childhood. Pediatrics
107(5):E66
25. Schneider E, Whitmore S, Glynn KM, Dominguez K, Mitsch A,
Centers for Disease Control and Prevention (CDC) (2008)
McKenna MT (2008) Revised surveillance case definitions for
HIV infection among adults, adolescents, and children aged
\18 months and for HIV infection and aids among children
aged 18 months to \13 years—United States. MMWR Recomm
Rep 57(10):1–12
26. Pyrgos V, Seitz AE, Steiner CA, Prevots DR, Williamson PR
(2013) Epidemiology of Cryptococcal meningitis in the US:
1997–2009. PLoS ONE 8(2):e56269
27. Bratton EW, El Husseini N, Chastain CA, Lee MS, Poole C,
Stu
¨rmer T, Juliano JJ, Weber DJ, Perfect JR (2012) Comparison
and temporal trends of three groups with cryptococcosis: HIV-
infected, solid organ transplant, and HIV-negative/non-trans-
plant. PLoS ONE 7(8):e43582
28. Singh N, Dromer F, Perfect JR, Lortholary O (2008) Immuno-
compromised hosts: cryptococcosis in solid organ transplant
recipients: current state of the science. Clin Infect Dis
47(10):1321–1327
29. Bartlett KH, Cheng P-Y, Duncan C, Galanis E, Hoang L, Kidd
S, Lee M-K, Lester S, MacDougall L, Mak S, Morshed M,
Taylor M, Kronstad JW (2012) A decade of experience: Cryp-
tococcus gattii in British Columbia. Mycopathologia 173(5–6):
311–319
30. Byrnes EJ, Li W, Lewit Y, Ma H, Voelz K, Ren P, Carter DA,
Chaturvedi V, Bildfell RJ, May RC, Heitman J (2010) Emer-
gence and pathogenicity of highly virulent Cryptococcus gattii
genotypes in the northwest United States. PLoS Pathog
6(4):e1000850
31. Galanis E, Macdougall L, Kidd S, Morshed M, British Colombia
Cryptococcus gattii Working Group (2010) Epidemiology of
Cryptococcus gattii, British Columbia, Canada, 1999–2007.
Emerging Infect Dis 16(2):251–257
32. Byrnes I, Edmond J, Bildfell RJ, Frank SA, Mitchell TG, Marr
KA, Heitman J (2009) Molecular evidence that the range of the
Vancouver Island outbreak of Cryptococcus gattii infection has
expanded into the pacific northwest in the United States. J Infect
Dis 199(7):1081–1086
33. Kunadharaju R, Choe U, Harris JR, Lockhart SR, Greene JN
(2013) Cryptococcus gattii, Florida, USA, 2011. Emerg Infect
19(3):519–521
34. Perfect JR, Bicanic T (2014) Cryptococcosis diagnosis and
treatment: what do we know now. Fungal Genet Biol. doi:10.
1016/j.fgb.2014.10.003
35. Boulware DR, Bonham SC, Meya DB, Wiesner DL, Park GS,
Kambugu A, Janoff EN, Bohjanen PR (2010) Paucity of initial
cerebrospinal fluid inflammation in cryptococcal meningitis is
associated with subsequent immune reconstitution inflammatory
syndrome. J Infect Dis 202(6):962–970
36. Shelburne SA 3rd, Darcourt J, White AC Jr, Greenberg SB,
Hamill RJ, Atmar RL, Visnegarwala F (2005) The role of im-
mune reconstitution inflammatory syndrome in AIDS-related
Cryptococcus neoformans disease in the era of highly active
antiretroviral therapy. Clin Infect Dis 40:1049–1052
37. Fetter A, Partisani M, Koenig H, Kremer M, Lang J-M (1993)
Asymptomatic oral Candida albicans carriage in HIV-infection:
frequency and predisposing factors. J Oral Pathol Med
22(2):57–59
38. Martins MD, Lozano-Chiu M, Rex JH (1998) Declining rates of
oropharyngeal candidiasis and carriage of Candida albicans
associated with trends toward reduced rates of carriage of flu-
conazole-resistant C. albicans in human immunodeficiency
virus-infected patients. Clin Infect Dis 27(5):1291–1294
39. Horn DL, Neofytos D, Anaissie EJ, Fishman JA, Steinbach WJ,
Olyaei AJ, Marr KA, Pfaller MA, Chang C-H, Webster KM
(2009) Epidemiology and outcomes of candidemia in 2019
Opportunistic yeast pathogens
123
patients: data from the prospective antifungal therapy alliance
registry. Clin Infect Dis 48(12):1695–1703
40. Koh AY, Kohler JR, Coggshall KT, Van Rooijen N, Pier GB
(2008) Mucosal damage and neutropenia are required for Can-
dida albicans dissemination. PLoS Pathog 4(2):e35
41. Wisplinghoff H, Bischoff T, Tallent SM, Seifert H, Wenzel RP,
Edmond MB (2004) Nosocomial bloodstream infections in US
hospitals: analysis of 24,179 cases from a prospective nation-
wide surveillance study. Clin Infect Dis 39(3):309–317
42. Freifeld AG, Bow EJ, Sepkowitz KA, Boeckh MJ, Ito JI, Mullen
CA, Raad II, Rolston KV, Young JAH, Wingard JR (2011)
Clinical practice guideline for the use of antimicrobial agents in
neutropenic patients with cancer: 2010 update by the Infectious
Diseases Society of America. Clin Infect Dis 52(4):e56–e93
43. Nucci M, Anaissie E (2001) Revisiting the source of can-
didemia: skin or gut? Clin Infect Dis 33(12):1959–1967
44. Nett J, Andes DR (2006) Candida albicans biofilm develop-
ment, modeling a host–pathogen interaction. Curr Opin
Microbiol 9(4):340–345
45. Robbins N, Uppuluri P, Nett J, Rajendran R, Ramage G, Lopez-
Ribot JL, Andes D, Cowen LE (2011) Hsp90 governs dispersion
and drug resistance of fungal biofilms. PLoS Pathog 7:e1002257
46. Hill JA, Ammar R, Torti D, Nislow C, Cowen LE (2013) Ge-
netic and genomic architecture of the evolution of resistance to
antifungal drug combinations. PLoS Genet 9(4):e1003390
47. Mansfield BE, Oltean HN, Oliver BG, Hoot SJ, Leyde SE,
Hedstrom L, White TC (2010) Azole drugs are imported by
facilitated diffusion in Candida albicans and other pathogenic
fungi. PLoS Pathog 6(9):e1001126
48. Centres for Disease Control and Prevention (2013) Antibiotic
resistance threats in the United States, 2013. U.S. Deparment of
Health and Human Services. Centres for Disease Control and
Prevention, Atlanta
49. Djawe K, Huang L, Daly KR, Levin L, Koch J, Schwartzman A,
Fong S, Roth B, Subramanian A, Grieco K, Jarlsberg L, Walzer
PD (2010) Serum antibody levels to the Pneumocystis jirovecii
major surface glycoprotein in the diagnosis of P. jirovecii
pneumonia in HIV
?
patients. PLoS ONE 5(12):e14259
50. Morris A, Norris KA (2012) Colonization by Pneumocystis
jirovecii and its role in disease. Clin Microbiol Rev 25:297–317
51. Pifer LL, Hughes WT, Stagno S, Woods D (1978) Pneumocystis
carinii infection: evidence for high prevalence in normal and
immunosuppressed children. Pediatrics 61(1):35–41
52. Morris A, Lundgren JD, Masur H, Walzer PD, Hanson DL,
Frederick T, Huang L, Beard CB, Kaplan JE (2004) Current
epidemiology of Pneumocystis pneumonia. Emerging Infect Dis
10(10):1713–1720
53. Masur H, Michelis MA, Greene JB, Onorato I, Stouwe RA,
Holzman RS, Wormser G, Brettman L, Lange M, Murray HW,
Cunningham-Rundles S (1981) An outbreak of community-
acquired Pneumocystis carinii pneumonia: initial manifestation
of cellular immune dysfunction. N Engl J Med 305(24):1431–
1438
54. Fisk DT, Meshnick S, Kazanjian PH (2003) Pneumocystis car-
inii pneumonia in patients in the developing world who have
acquired immunodeficiency syndrome. Clin Infect Dis 36(1):
70–78
55. Huang L, Cattamanchi A, Davis JL, Boon Sd, Kovacs J,
Meshnick S, Miller RF, Walzer PD, Worodria W, Masur H,
International HIV-associated Opportunistic Pneumonias (IHOP)
Study; Lung HIV Study (2011) HIV-associated Pneumocystis
pneumonia. Prom Am Thorac Soc 8(3):294–300
56. Kelley CF, Checkley W, Mannino DM, Franco-Paredes C, Del
Rio C, Holguin F (2009) Trends in hospitalizations for AIDS-
associated Pneumocystis jirovecii Pneumonia in the United
States (1986 to 2005). Chest 136(1):190–197
57. Kyeyune R, den Boon S, Cattamanchi A, Davis JL, Worodria
W, Yoo SD, Huang L (2010) Causes of early mortality in HIV-
infected TB suspects in an East African referral hospital. J Ac-
quir Immune Defic Syndr 55(4):446–450
58. Sivam S, Sciurba FC, Lucht LA, Zhang Y, Duncan SR, Norris
KA, Morris A (2011) Distribution of Pneumocystis jirovecii in
lungs from colonized COPD patients. Diagn Microbiol Infect
Dis 71(1):24–28
59. Wald A, Leisenring W, van Burik JA, Bowden RA (1997)
Epidemiology of Aspergillus infections in a large cohort of
patients undergoing bone marrow transplantation. J Infect Dis
175:1459–1466
60. Marr KA, Carter RA, Crippa F, Wald A, Corey L (2002) Epi-
demiology and outcome of mould infections in hematopoietic
stem cell transplant recipients. Clin Infect Dis 34:909–917
61. Baddley JW, Andes DR, Marr KA, Kauffman CA, Kontoyiannis
DP, Ito JI, Schuster MG, Brizendine KD, Patterson TF, Lyon
GM, Boeckh M, Oster RA, Chiller T, Pappas PG (2013) Anti-
fungal therapy and length of hospitalization in transplant
patients with invasive aspergillosis. Med Mycol 51(2):128–135
62. Agarwal R, Chakrabarti A (2013) Allergic bronchopulmonary
aspergillosis in asthma: epidemiological, clinical and therapeutic
issues. Future Microbiol 8(11):1463–1474
63. Denning DW, Pleuvry A, Cole DC (2013) Global burden of
allergic bronchopulmonary aspergillosis with asthma and its
complication chronic pulmonary aspergillosis in adults. Med
Mycol 51(4):361–370
64. Kauffman CA (2007) Histoplasmosis: a clinical and laboratory
update. Clin Microbiol Rev 20(1):115–132
65. Colombo AL, Tobo
´n A, Restrepo A, Queiroz-Telles F, Nucci M
(2011) Epidemiology of endemic systemic fungal infections in
Latin America. Med Mycol 49(8):785–798
66. Goodwin RAJ, Shapiro JL, Thurman GH, Thurman SS, des Prez
RM (1980) Disseminated Histoplasmosis: clinical and patho-
logic correlations. Medicine (Baltimore) 59(1):1–33
67. Freitas DFS, Valle ACFD, da Silva MBT, Campos DP, Lyra
MR, de Souza RV, Veloso VG, Zancope
´-Oliveira RM, Bastos
FI, Galhardo MCG (2014) Sporotrichosis: an emerging ne-
glected opportunistic infection in HIV-infected patients in Rio
de Janeiro, Brazil. PLoS Negl Trop Dis 8(8):e3110
68. Sylvestre TF, Silva LRF, Cavalcante RDS, Moris DV, Venturini
J, Vicentini AP, Carvalho LRD, Mendes RP (2014) Prevalence
and serological diagnosis of relapse in Paracoccidioidomycosis
patients. PLoS Negl Trop Dis 8(5):e2834
69. Duong TA (1996) Infection due to Penicillium marneffei,an
emerging pathogen: review of 155 reported cases. Clin Infect
Dis 23(1):125–130
70. Lee PPW, Chan K-W, Lee T-L, Ho MH-K, Chen X-Y, Li C-H,
Chu K-M, Zeng H-S, Lau Y-L (2012) Penicilliosis in children
without HIV infection–are they immunodeficient? Clin Infect
Dis 54(2):e8–e19
71. Supparatpinyo K, Chiewchanvit S, Hirunsri P, Uthammachai C,
Nelson KE, Sirisanthana T (1992) Penicillium marneffei infec-
tion in patients infected with human immunodeficiency virus.
Clin Infect Dis 14(4):871–874
72. Vanittanakom N, Cooper CR, Fisher MC, Sirisanthana T (2006)
Penicillium marneffei infection and recent advances in the epi-
demiology and molecular biology aspects. Clin Microbiol Rev
19(1):95–110
73. Bulterys PL, Le T, Quang VM, Nelson KE, Lloyd-Smith JO
(2013) Environmental predictors and incubation period of
AIDS-associated Penicillium marneffei infection in Ho Chi
Minh City, Vietnam. Clin Infect Dis 56(9):1273–1279
74. Le T, Wolbers M, Chi NH, Quang VM, Chinh NT, Lan NPH,
Lam PS, Kozal MJ, Shikuma CM, Day JN, Farrar J (2011)
Epidemiology, seasonality, and predictors of outcome of AIDS-
E. J. Polvi et al.
123
associated Penicillium marneffei infection in Ho Chi Minh City,
Vietnam. Clin Infect Dis 52(7):945–952
75. Flaherman VJ, Hector R, Rutherford GW (2007) Estimating
severe coccidioidomycosis in California. Emerg Infect Dis
13(7):1087–1090
76. Brown J, Benedict K, Park B, Thompson G 3rd (2013) Coc-
cidioidomycosis: epidemiology. Clin Epidemiol 5:185–197
77. Santelli AC, Blair JE, Roust LR (2006) Coccidioidomycosis in
patients with diabetes mellitus. Am J Med 119(11):964–969
78. Khuu D, Shafir S, Bristow B, Sorvillo F (2014) Blastomycosis
mortality rates, United States, 1990–2010. Emerg Infect Dis
20(11):1789–1794
79. Roy M, Benedict K, Deak E, Kirby MA, McNiel JT, Sickler CJ,
Eckardt E, Marx RK, Heffernan RT, Meece JK, Klein BS,
Archer JR, Theurer J, Davis JP, Park BJ (2013) A large com-
munity outbreak of blastomycosis in Wisconsin with geographic
and ethnic clustering. Clin Infect Dis 57(5):655–662
80. Mendoza L, Vilela R, Voelz K, Ibrahim AS, Voigt K, Lee SC
(2014) Human fungal pathogens of Mucorales and Entomoph-
thorales. Cold Spring Harb Perspect Med. doi:10.1101/
cshperspect.a019562
81. Chakrabarti A, Das A, Mandal J, Shivaprakash MR, George VK,
Tarai B, Rao P, Panda N, Verma SC, Sakhuja V (2006) The
rising trend of invasive zygomycosis in patients with uncon-
trolled diabetes mellitus. Med Mycol 44(4):335–342
82. Lanternier F, Sun H-Y, Ribaud P, Singh N, Kontoyiannis DP,
Lortholary O (2012) Mucormycosis in organ and stem cell
transplant recipients. Clin Infect Dis 54(11):1629–1636
83. Roden MM, Zaoutis TE, Buchanan WL, Knudsen TA, Sarkisova
TA, Schaufele RL, Sein M, Sein T, Chiou CC, Chu JH, Kon-
toyiannis DP, Walsh TJ (2005) Epidemiology and outcome of
zygomycosis: a review of 929 reported cases. Clin Infect Dis
41(5):634–653
84. Green JP, Karras DJ (2012) Update on emerging infections:
news from the Centers for Disease Control and Prevention.
Notes from the field: fatal fungal soft-tissue infections after a
tornado–Joplin, Missouri, 2011. Ann Emerg Med 59(1):53–55
85. Bonifaz A, Va
´zquez-Gonza
´lez D, Tirado-Sa
´nchez A, Ponce-
Olivera RM (2013) Cutaneous zygomycosis. Clin Dermatol
30(4):413–419
86. Barros MBDL, de Almeida Paes R, Schubach AO (2011)
Sporothrix schenckii and Sporotrichosis. Clin Microbiol Rev
24(4):633–654
87. Miceli MH, Diaz JA, Lee SA (2011) Emerging opportunistic
yeast infections. Lancet Infect Dis 11:142–151
88. Girmenia C, Pagano L, Martino B, D’Antonio D, Fanci R,
Specchia G, Melillo L, Buelli M, Pizzarelli G, Venditti M,
Martino P, GIMEMA Infection Program (2005) Invasive in-
fections caused by Trichosporon species and Geotrichum
capitatum in patients with hematological malignancies: a ret-
rospective multicenter study from Italy and review of the
literature. J Clin Microbiol 43(4):1818–1828
89. Lin X, Heitman J (2006) The biology of the Cryptococcus
neoformans species complex. Annu Rev Microbiol 60:69–105
90. Perfect JR, Dismukes WE, Dromer F, Goldman DL, Graybill JR,
Hamill RJ, Harrison TS, Larsen RA, Lortholary O, Nguyen MH,
Pappas PG, Powderly WG, Singh N, Sobel JD, Sorrell TC
(2010) Clinical practice guidelines for the management of
cryptococcal disease: 2010 update by the infectious diseases
society of America. Clin Infect Dis 50:291–322
91. Idnurm A, Bahn YS, Nielsen K, Lin X, Fraser JA, Heitman J
(2005) Deciphering the model pathogenic fungus Cryptococcus
neoformans. Nat Rev Microbiol 3:753–764
92. Sorrell TC, Chen SC, Ruma P, Meyer W, Pfeiffer TJ, Ellis DH,
Brownlee AG (1996) Concordance of clinical and environ-
mental isolates of Cryptococcus neoformans var. gattii by
random amplification of polymorphic DNA analysis and PCR
fingerprinting. J Clin Microbiol 34:1253–1260
93. Yamamoto Y, Kohno S, Koga H, Kakeya H, Tomono K, Kaku
M, Yamazaki T, Arisawa M, Hara K (1995) Random amplified
polymorphic DNA analysis of clinically and environmentally
isolated Cryptococcus neoformans in Nagasaki. J Clin Microbiol
33:3328–3332
94. Cogliati M (2013) Global molecular epidemiology of Crypto-
coccus neoformans and Cryptococcus gattii: an atlas of the
molecular types. Scientifica (Cairo) 2013:675213
95. Litvintseva AP, Carbone I, Rossouw J, Thakur R, Govender NP,
Mitchell TG (2011) Evidence that the human pathogenic fungus
Cryptococcus neoformans var. grubii may have evolved in
Africa. PLoS ONE 6:e19688
96. Kidd SE, Chow Y, Mak S, Bach PJ, Chen H, Hingston AO,
Kronstad JW, Bartlett KH (2007) Characterization of environ-
mental sources of the human and animal pathogen Cryptococcus
gattii in British Columbia, Canada, and the Pacific Northwest of
the United States. Appl Environ Microbiol 73:1433–1443
97. Chen SC, Meyer W, Sorrell TC (2014) Cryptococcus gattii in-
fections. Clin Microbiol Rev 27:980–1024
98. Velagapudi R, Hsueh YP, Geunes-Boyer S, Wright JR, Heitman
J (2009) Spores as infectious propagules of Cryptococcus
neoformans. Infect Immun 77:4345–4355
99. Coelho C, Bocca AL, Casadevall A (2014) The intracellular life
of Cryptococcus neoformans. Annu Rev Pathol 9:219–238
100. Sabiiti W, May RC (2012) Mechanisms of infection by the
human fungal pathogen Cryptococcus neoformans. Future Mi-
crobiol 7:1297–1313
101. Garcia-Hermoso D, Janbon G, Dromer F (1999) Epi-
demiological evidence for dormant Cryptococcus neoformans
infection. J Clin Microbiol 37:3204–3209
102. Kronstad JW, Attarian R, Cadieux B, Choi J, D’Souza CA,
Griffiths EJ, Geddes JM, Hu G, Jung WH, Kretschmer M, Saikia
S, Wang J (2011) Expanding fungal pathogenesis: Cryptococcus
breaks out of the opportunistic box. Nat Rev Microbiol 9:193–203
103. Nadrous HF, Antonios VS, Terrell CL, Ryu JH (2003) Pul-
monary cryptococcosis in nonimmunocompromised patients.
Chest 124:2143–2147
104. Voelz K, May RC (2010) Cryptococcal interactions with the
host immune system. Eukaryot Cell 9:835–846
105. Charlier C, Nielsen K, Daou S, Brigitte M, Chretien F, Dromer F
(2009) Evidence of a role for monocytes in dissemination and
brain invasion by Cryptococcus neoformans. Infect Immun
77:120–127
106. Chen SH, Stins MF, Huang SH, Chen YH, Kwon-Chung KJ,
Chang Y, Kim KS, Suzuki K, Jong AY (2003) Cryptococcus
neoformans induces alterations in the cytoskeleton of human
brain microvascular endothelial cells. J Med Microbiol
52:961–970
107. Chang YC, Stins MF, McCaffery MJ, Miller GF, Pare DR, Dam
T, Paul-Satyaseela M, Kim KS, Kwon-Chung KJ (2004) Cryp-
tococcal yeast cells invade the central nervous system via
transcellular penetration of the blood-brain barrier. Infect Im-
mun 72:4985–4995
108. Bicanic T, Harrison TS (2004) Cryptococcal meningitis. Br Med
Bull 72:99–118
109. Chen SC, Slavin MA, Heath CH, Playford EG, Byth K, Marriott
D, Kidd SE, Bak N, Currie B, Hajkowicz K, Korman TM,
McBride WJ, Meyer W, Murray R, Sorrell TC, Australia, New
Zealand Mycoses Interest Group-Cryptococcus S (2012) Clin-
ical manifestations of Cryptococcus gattii infection:
determinants of neurological sequelae and death. Clin Infect Dis
55:789–798
110. Pappas PG, Kauffman CA, Andes D, Benjamin DK Jr, Calandra
TF, Edwards JE Jr, Filler SG, Fisher JF, Kullberg BJ, Ostrosky-
Opportunistic yeast pathogens
123
Zeichner L, Reboli AC, Rex JH, Walsh TJ, Sobel JD, Infectious
Diseases Society of A (2009) Clinical practice guidelines for the
management of candidiasis: 2009 update by the Infectious
Diseases Society of America. Clin Infect Dis 48:503–535
111. Odds FC (1988) Candida and candidosis, 2nd edn. Baillie
`re
Tindall, London
112. Hube B (2004) From commensal to pathogen: stage- and tissue-
specific gene expression of Candida albicans. Curr Opin Mi-
crobiol 7:336–341
113. Taylor BN, Harrer T, Pscheidl E, Schweizer A, Rollinghoff M,
Schroppel K (2003) Surveillance of nosocomial transmission of
Candida albicans in an intensive care unit by DNA finger-
printing. J Hosp Infect 55:283–289
114. Asmundsdottir LR, Erlendsdottir H, Haraldsson G, Guo H, Xu J,
Gottfredsson M (2008) Molecular epidemiology of candidemia:
evidence of clusters of smoldering nosocomial infections. Clin
Infect Dis 47:e17–e24
115. Gow NA, Hube B (2012) Importance of the Candida albicans
cell wall during commensalism and infection. Curr Opin Mi-
crobiol 15:406–412
116. Chen C, Pande K, French SD, Tuch BB, Noble SM (2011) An
iron homeostasis regulatory circuit with reciprocal roles in
Candida albicans commensalism and pathogenesis. Cell Host
Microbe 10:118–135
117. Pande K, Chen C, Noble SM (2013) Passage through the
mammalian gut triggers a phenotypic switch that promotes
Candida albicans commensalism. Nat Genet 45:1088–1091
118. Perez JC, Kumamoto CA, Johnson AD (2013) Candida albicans
commensalism and pathogenicity are intertwined traits directed
by a tightly knit transcriptional regulatory circuit. PLoS Biol
11:e1001510
119. Achkar JM, Fries BC (2010) Candida infections of the
genitourinary tract. Clin Microbiol Rev 23:253–273
120. de Leon EM, Jacober SJ, Sobel JD, Foxman B (2002) Preva-
lence and risk factors for vaginal Candida colonization in
women with type 1 and type 2 diabetes. BMC Infect Dis 2:1
121. Spinillo A, Capuzzo E, Acciano S, De Santolo A, Zara F (1999)
Effect of antibiotic use on the prevalence of symptomatic vul-
vovaginal candidiasis. Am J Obstet Gynecol 180:14–17
122. Sobel JD, Faro S, Force RW, Foxman B, Ledger WJ, Nyirjesy
PR, Reed BD, Summers PR (1998) Vulvovaginal candidiasis:
epidemiologic, diagnostic, and therapeutic considerations. Am J
Obstet Gynecol 178:203–211
123. Edwards S (1996) Balanitis and balanoposthitis: a review.
Genitourin Med 72:155–159
124. Kauffman CA, Vazquez JA, Sobel JD, Gallis HA, McKinsey
DS, Karchmer AW, Sugar AM, Sharkey PK, Wise GJ, Mangi R,
Mosher A, Lee JY, Dismukes WE (2000) Prospective multi-
center surveillance study of funguria in hospitalized patients.
The National Institute for Allergy and Infectious Diseases
(NIAID) Mycoses Study Group. Clin Infect Dis 30:14–18
125. Bougnoux ME, Kac G, Aegerter P, D’Enfert C, Fagon JY,
CandiRea Study G (2008) Candidemia and candiduria in
critically ill patients admitted to intensive care units in France:
incidence, molecular diversity, management and outcome. In-
tensive Care Med 34:292–299
126. Lundstrom T, Sobel J (2001) Nosocomial candiduria: a review.
Clin Infect Dis 32:1602–1607
127. Alvarez-Lerma F, Nolla-Salas J, Leon C, Palomar M, Jorda R,
Carrasco N, Bobillo F, Group ES (2003) Candiduria in critically
ill patients admitted to intensive care medical units. Intensive
Care Med 29:1069–1076
128. Coronado-Castellote L, Jimenez-Soriano Y (2013) Clinical and
microbiological diagnosis of oral candidiasis. J Clin Exp Dent
5:e279–e286
129. Samaranayake LP, Keung Leung W, Jin L (2009) Oral mucosal
fungal infections. Periodontol 49:39–59
130. Pankhurst CL (2013) Candidiasis (oropharyngeal). Clin Evid
(Online) 2013:1304
131. de Repentigny L, Lewandowski D, Jolicoeur P (2004) Im-
munopathogenesis of oropharyngeal candidiasis in human
immunodeficiency virus infection. Clin Microbiol Rev
17:729–759
132. Eggimann P, Garbino J, Pittet D (2003) Epidemiology of Can-
dida species infections in critically ill non-immunosuppressed
patients. Lancet Infect Dis 3:685–702
133. Pappas PG, Alexander BD, Andes DR, Hadley S, Kauffman CA,
Freifeld A, Anaissie EJ, Brumble LM, Herwaldt L, Ito J, Kon-
toyiannis DP, Lyon GM, Marr KA, Morrison VA, Park BJ,
Patterson TF, Perl TM, Oster RA, Schuster MG, Walker R,
Walsh TJ, Wannemuehler KA, Chiller TM (2010) Invasive
fungal infections among organ transplant recipients: results of
the transplant-associated infection surveillance network
(TRANSNET). Clin Infect Dis 50:1101–1111
134. Eggimann P, Bille J, Marchetti O (2011) Diagnosis of invasive
candidiasis in the ICU. Ann Intensive Care 1:37
135. Ellepola AN, Morrison CJ (2005) Laboratory diagnosis of in-
vasive candidiasis. J Microbiol 43:65–84
136. Morrell M, Fraser VJ, Kollef MH (2005) Delaying the empiric
treatment of Candida bloodstream infection until positive blood
culture results are obtained: a potential risk factor for hospital
mortality. Antimicrob Agents Chemother 49:3640–3645
137. Gigliotti F, Wright TW (2012) Pneumocystis: where does it
live? PLoS Pathog 8:e1003025
138. Durand-Joly I, El Aliouat M, 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
139. Sinclair K, Wakefield AE, Banerji S, Hopkin JM (1991) Pneu-
mocystis carinii organisms derived from rat and human hosts are
genetically distinct. Mol Biochem Parasitol 45:183–184
140. Thomas CF Jr, Limper AH (2004) Pneumocystis pneumonia.
N Engl J Med 350:2487–2498
141. Beard CB, Roux P, Nevez G, Hauser PM, Kovacs JA, Unnasch
TR, Lundgren B (2004) Strain typing methods and molecular
epidemiology of Pneumocystis pneumonia. Emerg Infect Dis
10:1729–1735
142. Peterson JC, Cushion MT (2005) Pneumocystis: not just pneu-
monia. Curr Opin Microbiol 8:393–398
143. Wakefield AE, Lindley AR, Ambrose HE, Denis CM, Miller RF
(2003) Limited asymptomatic carriage of Pneumocystis jiroveci
in human immunodeficiency virus-infected patients. J Infect Dis
187:901–908
144. 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
145. Morris A, Wei K, Afshar K, Huang L (2008) Epidemiology and
clinical significance of pneumocystis colonization. J Infect Dis
197:10–17
146. Choukri F, Menotti J, Sarfati C, Lucet JC, Nevez G, Garin YJ,
Derouin F, Totet A (2010) Quantification and spread of Pneu-
mocystis jirovecii in the surrounding air of patients with
Pneumocystis pneumonia. Clin Infect Dis 51:259–265
147. 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
148. Wheat LJ, Freifeld AG, Kleiman MB, Baddley JW, McKinsey
DS, Loyd JE, Kauffman CA, Infectious Diseases Society of
America (2007) Clinical practice guidelines for the management
E. J. Polvi et al.
123
of patients with histoplasmosis: 2007 update by the Infectious
Diseases Society of America. Clin Infect Dis 45:807–825
149. Adenis A, Nacher M, Hanf M, Vantilcke V, Boukhari R, Blachet
D, Demar M, Aznar C, Carme B, Couppie P (2014) HIV-asso-
ciated histoplasmosis early mortality and incidence trends: from
neglect to priority. PLoS Negl Trop Dis 8:e3100
150. Adenis AA, Aznar C, Couppie P (2014) Histoplasmosis in HIV-
infected patients: a review of new developments and remaining
gaps. Curr Trop Med Rep 1:119–128
151. Casadevall A, Pirofski LA (1999) Host-pathogen interactions:
redefining the basic concepts of virulence and pathogenicity.
Infect Immun 67:3703–3713
152. Nicholls S, Leach MD, Priest CL, Brown AJ (2009) Role of the
heat shock transcription factor, Hsf1, in a major fungal pathogen
that is obligately associated with warm-blooded animals. Mol
Microbiol 74:844–861
153. Leach MD, Tyc KM, Brown AJ, Klipp E (2012) Modelling the
regulation of thermal adaptation in Candida albicans, a major
fungal pathogen of humans. PLoS ONE 7:e32467
154. Nicholls S, MacCallum DM, Kaffarnik FA, Selway L, Peck SC,
Brown AJ (2011) Activation of the heat shock transcription
factor Hsf1 is essential for the full virulence of the fungal
pathogen Candida albicans. Fungal Genet Biol 48:297–305
155. Leach MD, Cowen LE (2014) Membrane fluidity and temperature
sensing are coupled via circuitry comprised of Ole1, Rsp5, and
Hsf1 in Candida albicans. Eukaryot Cell 13:1077–1084
156. O’Meara T, Cowen LE (2014) Hsp90-dependent regulatory
circuitry controlling temperature-dependent fungal development
and virulence. Cell Microbiol 16(4):473–481
157. Nichols CB, Perfect ZH, Alspaugh JA (2007) A Ras1-Cdc24
signal transduction pathway mediates thermotolerance in the
fungal pathogen Cryptococcus neoformans. Mol Microbiol
63:1118–1130
158. Cheon SA, Jung KW, Chen YL, Heitman J, Bahn YS, Kang HA
(2011) Unique evolution of the UPR pathway with a novel bZIP
transcription factor, Hxl1, for controlling pathogenicity of
Cryptococcus neoformans. PLoS Pathog 7:e1002177
159. O’Meara TR, Hay C, Price MS, Giles S, Alspaugh JA (2010)
Cryptococcus neoformans histone acetyltransferase Gcn5 reg-
ulates fungal adaptation to the host. Eukaryot Cell 9:1193–1202
160. Gerik KJ, Bhimireddy SR, Ryerse JS, Specht CA, Lodge JK
(2008) PKC1 is essential for protection against both oxidative
and nitrosative stresses, cell integrity, and normal manifestation
of virulence factors in the pathogenic fungus Cryptococcus
neoformans. Eukaryot Cell 7:1685–1698
161. Hu G, Cheng PY, Sham A, Perfect JR, Kronstad JW (2008)
Metabolic adaptation in Cryptococcus neoformans during early
murine pulmonary infection. Mol Microbiol 69:1456–1475
162. Leach M, Cowen L (2014) To sense or die: mechanisms of
temperature sensing in fungal pathogens. Curr Fungal Infect Rep
8:185–191
163. Carratu L, Franceschelli S, Pardini CL, Kobayashi GS, Horvath
I, Vigh L, Maresca B (1996) Membrane lipid perturbation
modifies the set point of the temperature of heat shock response
in yeast. Proc Natl Acad Sci USA 93:3870–3875
164. Selvig K, Alspaugh JA (2011) pH response pathways in fungi:
adapting to host-derived and environmental signals. Mycobi-
ology 39:249–256
165. Bensen ES, Martin SJ, Li M, Berman J, Davis DA (2004)
Transcriptional profiling in Candida albicans reveals new
adaptive responses to extracellular pH and functions for
Rim101p. Mol Microbiol 54:1335–1351
166. Nobile CJ, Solis N, Myers CL, Fay AJ, Deneault JS, Nantel A,
Mitchell AP, Filler SG (2008) Candida albicans transcription
factor Rim101 mediates pathogenic interactions through cell
wall functions. Cell Microbiol 10:2180–2196
167. De Bernardis F, Muhlschlegel FA, Cassone A, Fonzi WA (1998)
The pH of the host niche controls gene expression in and viru-
lence of Candida albicans. Infect Immun 66:3317–3325
168. Davis D, Edwards JE Jr, Mitchell AP, Ibrahim AS (2000)
Candida albicans RIM101 pH response pathway is required for
host-pathogen interactions. Infect Immun 68:5953–5959
169. O’Meara TR, Norton D, Price MS, Hay C, Clements MF, Ni-
chols CB, Alspaugh JA (2010) Interaction of Cryptococcus
neoformans Rim101 and protein kinase A regulates capsule.
PLoS Pathog 6:e1000776
170. Okagaki LH, Wang Y, Ballou ER, O’Meara TR, Bahn YS,
Alspaugh JA, Xue C, Nielsen K (2011) Cryptococcal titan cell
formation is regulated by G-protein signaling in response to
multiple stimuli. Eukaryot Cell 10:1306–1316
171. O’Meara TR, Holmer SM, Selvig K, Dietrich F, Alspaugh JA
(2013) Cryptococcus neoformans Rim101 is associated with cell
wall remodeling and evasion of the host immune responses.
MBio 4(1):e00522-12
172. O’Meara TR, Xu W, Selvig KM, O’Meara MJ, Mitchell AP,
Alspaugh JA (2014) The Cryptococcus neoformans Rim101
transcription factor directly regulates genes required for adap-
tation to the host. Mol Cell Biol 34:673–684
173. Feldmesser M, Kress Y, Novikoff P, Casadevall A (2000)
Cryptococcus neoformans is a facultative intracellular pathogen
in murine pulmonary infection. Infect Immun 68:4225–4237
174. Vylkova S, Carman AJ, Danhof HA, Collette JR, Zhou H,
Lorenz MC (2011) The fungal pathogen Candida albicans au-
toinduces hyphal morphogenesis by raising extracellular pH.
MBio 2:e00055-11
175. Derengowski Lda S, Paes HC, Albuquerque P, Tavares AH,
Fernandes L, Silva-Pereira I, Casadevall A (2013) The tran-
scriptional response of Cryptococcus neoformans to ingestion by
Acanthamoeba castellanii and macrophages provides insights
into the evolutionary adaptation to the mammalian host.
Eukaryot Cell 12:761–774
176. Weinberg ED (2009) Iron availability and infection. Biochim
Biophys Acta 1790:600–605
177. Tangen KL, Jung WH, Sham AP, Lian T, Kronstad JW (2007)
The iron- and cAMP-regulated gene SIT1 influences ferriox-
amine B utilization, melanization and cell wall structure in
Cryptococcus neoformans. Microbiology 153:29–41
178. Heymann P, Gerads M, Schaller M, Dromer F, Winkelmann G,
Ernst JF (2002) The siderophore iron transporter of Candida
albicans (Sit1p/Arn1p) mediates uptake of ferrichrome-type
siderophores and is required for epithelial invasion. Infect Im-
mun 70:5246–5255
179. Almeida RS, Wilson D, Hube B (2009) Candida albicans iron
acquisition within the host. FEMS Yeast Res 9:1000–1012
180. Jung WH, Saikia S, Hu G, Wang J, Fung CK, D’Souza C, White
R, Kronstad JW (2010) HapX positively and negatively reg-
ulates the transcriptional response to iron deprivation in
Cryptococcus neoformans. PLoS Pathog 6:e1001209
181. Jung WH, Sham A, Lian T, Singh A, Kosman DJ, Kronstad JW
(2008) Iron source preference and regulation of iron uptake in
Cryptococcus neoformans. PLoS Pathog 4:e45
182. Jung WH, Hu G, Kuo W, Kronstad JW (2009) Role of fer-
roxidases in iron uptake and virulence of Cryptococcus
neoformans. Eukaryot Cell 8:1511–1520
183. Jung WH, Sham A, White R, Kronstad JW (2006) Iron regula-
tion of the major virulence factors in the AIDS-associated
pathogen Cryptococcus neoformans. PLoS Biol 4:e410
184. Weissman Z, Shemer R, Conibear E, Kornitzer D (2008) An
endocytic mechanism for haemoglobin-iron acquisition in
Candida albicans. Mol Microbiol 69:201–217
185. Santos R, Buisson N, Knight S, Dancis A, Camadro JM, Lesu-
isse E (2003) Haemin uptake and use as an iron source by
Opportunistic yeast pathogens
123
Candida albicans: role of CaHMX1-encoded haem oxygenase.
Microbiology 149:579–588
186. Navarathna DH, Roberts DD (2010) Candida albicans heme
oxygenase and its product CO contribute to pathogenesis of
candidemia and alter systemic chemokine and cytokine ex-
pression. Free Radic Biol Med 49:1561–1573
187. Otterbein LE, Bach FH, Alam J, Soares M, Tao LuH, Wysk M,
Davis RJ, Flavell RA, Choi AM (2000) Carbon monoxide has
anti-inflammatory effects involving the mitogen-activated pro-
tein kinase pathway. Nat Med 6:422–428
188. Pendrak ML, Chao MP, Yan SS, Roberts DD (2004) Heme
oxygenase in Candida albicans is regulated by hemoglobin and
is necessary for metabolism of exogenous heme and hemoglobin
to alpha-biliverdin. J Biol Chem 279:3426–3433
189. Hwang LH, Mayfield JA, Rine J, Sil A (2008) Histoplasma
requires SID1, a member of an iron-regulated siderophore gene
cluster, for host colonization. PLoS Pathog 4:e1000044
190. Hilty J, George Smulian A, Newman SL (2011) Histoplasma
capsulatum utilizes siderophores for intracellular iron acquisi-
tion in macrophages. Med Mycol 49:633–642
191. Chao LY, Marletta MA, Rine J (2008) Sre1, an iron-modulated
GATA DNA-binding protein of iron-uptake genes in the fungal
pathogen Histoplasma capsulatum. Biochemistry 47:7274–7283
192. Timmerman MM, Woods JP (2001) Potential role for extracel-
lular glutathione-dependent ferric reductase in utilization of
environmental and host ferric compounds by Histoplasma cap-
sulatum. Infect Immun 69:7671–7678
193. Nemecek JC, Wuthrich M, Klein BS (2006) Global control of
dimorphism and virulence in fungi. Science 312:583–588
194. Nguyen VQ, Sil A (2008) Temperature-induced switch to the
pathogenic yeast form of Histoplasma capsulatum requires
Ryp1, a conserved transcriptional regulator. Proc Natl Acad Sci
USA 105:4880–4885
195. Beyhan S, Gutierrez M, Voorhies M, Sil A (2013) A tem-
perature-responsive network links cell shape and virulence traits
in a primary fungal pathogen. PLoS Biol 11:e1001614
196. Warenda AJ, Konopka JB (2002) Septin function in Candida
albicans morphogenesis. Mol Biol Cell 13:2732–2746
197. Sudbery P, Gow N, Berman J (2004) The distinct morphogenic
states of Candida albicans. Trends Microbiol 12:317–324
198. Xu XL, Lee RT, Fang HM, Wang YM, Li R, Zou H, Zhu Y,
Wang Y (2008) Bacterial peptidoglycan triggers Candida albi-
cans hyphal growth by directly activating the adenylyl cyclase
Cyr1p. Cell Host Microbe 4:28–39
199. Brown V, Sexton JA, Johnston M (2006) A glucose sensor in
Candida albicans. Eukaryot Cell 5:1726–1737
200. Hudson DA, Sciascia QL, Sanders RJ, Norris GE, Edwards PJ,
Sullivan PA, Farley PC (2004) Identification of the dialysable
serum inducer of germ-tube formation in Candida albicans.
Microbiology 150:3041–3049
201. Shapiro RS, Robbins N, Cowen LE (2011) Regulatory circuitry
governing fungal development, drug resistance, and disease.
Microbiol Mol Biol Rev 75:213–267
202. Shapiro RS, Uppuluri P, Zaas AK, Collins C, Senn H, Perfect
JR, Heitman J, Cowen LE (2009) Hsp90 orchestrates tem-
perature-dependent Candida albicans morphogenesis via Ras1-
PKA signaling. Curr Biol 19:621–629
203. Bockmuhl DP, Ernst JF (2001) A potential phosphorylation site for
an A-type kinase in the Efg1 regulator protein contributes to hy-
phal morphogenesis of Candida albicans. Genetics 157:1523–1530
204. Cao F, Lane S, Raniga PP, Lu Y, Zhou Z, Ramon K, Chen J, Liu
H (2006) The Flo8 transcription factor is essential for hyphal
development and virulence in Candida albicans. Mol Biol Cell
17:295–307
205. Lu Y, Su C, Liu H (2014) Candida albicans hyphal initiation
and elongation. Trends Microbiol 22:707–714
206. Dalle F, Wachtler B, L’Ollivier C, Holland G, Bannert N,
Wilson D, Labruere C, Bonnin A, Hube B (2010) Cellular in-
teractions of Candida albicans with human oral epithelial cells
and enterocytes. Cell Microbiol 12:248–271
207. Rocha CR, Schroppel K, Harcus D, Marcil A, Dignard D, Taylor
BN, Thomas DY, Whiteway M, Leberer E (2001) Signaling
through adenylyl cyclase is essential for hyphal growth and
virulence in the pathogenic fungus Candida albicans. Mol Biol
Cell 12:3631–3643
208. Leberer E, Harcus D, Dignard D, Johnson L, Ushinsky S,
Thomas DY, Schroppel K (2001) Ras links cellular morpho-
genesis to virulence by regulation of the MAP kinase and cAMP
signalling pathways in the pathogenic fungus Candida albicans.
Mol Microbiol 42:673–687
209. Park H, Myers CL, Sheppard DC, Phan QT, Sanchez AA, Ed-
wards JE, Filler SG (2005) Role of the fungal Ras-protein kinase
A pathway in governing epithelial cell interactions during
oropharyngeal candidiasis. Cell Microbiol 7:499–510
210. Murad AM, Leng P, Straffon M, Wishart J, Macaskill S,
MacCallum D, Schnell N, Talibi D, Marechal D, Tekaia F,
d’Enfert C, Gaillardin C, Odds FC, Brown AJ (2001) NRG1
represses yeast-hypha morphogenesis and hypha-specific gene
expression in Candida albicans. EMBO J 20:4742–4752
211. Saville SP, Lazzell AL, Monteagudo C, Lopez-Ribot JL (2003)
Engineered control of cell morphology in vivo reveals distinct
roles for yeast and filamentous forms of Candida albicans
during infection. Eukaryot Cell 2:1053–1060
212. Hirakawa MP, Martinez DA, Sakthikumar S, Anderson MZ,
Berlin A, Gujja S, Zeng Q, Zisson E, Wang JM, Greenberg JM,
Berman J, Bennett RJ, Cuomo CA (2014) Genetic and pheno-
typic intra-species variation in Candida albicans. Genome Res.
doi:10.1101/gr.174623.114
213. Vautier S, Drummond RA, Chen K, Murray GI, Kadosh D,
Brown AJ, Gow NA, MacCallum DM, Kolls JK, Brown GD
(2014) Candida albicans colonization and dissemination from
the murine gastrointestinal tract: the influence of morphology
and Th17 immunity. Cell Microbiol. doi:10.1111/cmi.12388
214. Brand A (2012) Hyphal growth in human fungal pathogens and
its role in virulence. Int J Microbiol. doi:10.1155/2012/517529
215. Liu Y, Filler SG (2011) Candida albicans Als3, a multifunc-
tional adhesin and invasin. Eukaryot Cell 10:168–173
216. Staab JF, Bradway SD, Fidel PL, Sundstrom P (1999) Adhesive
and mammalian transglutaminase substrate properties of Can-
dida albicans Hwp1. Science 283:1535–1538
217. Hube B, Monod M, Schofield DA, Brown AJ, Gow NA (1994)
Expression of seven members of the gene family encoding se-
cretory aspartyl proteinases in Candida albicans. Mol Microbiol
14:87–99
218. White TC, Agabian N (1995) Candida albicans secreted as-
partyl proteinases: isoenzyme pattern is determined by cell type,
and levels are determined by environmental factors. J Bacteriol
177:5215–5221
219. Cowen LE (2013) The fungal Achilles’ heel: targeting Hsp90 to
cripple fungal pathogens. Curr Opin Microbiol 16(4):377–384
220. Prill SK, Klinkert B, Timpel C, Gale CA, Schroppel K, Ernst JF
(2005) PMT family of Candida albicans: five protein manno-
syltransferase isoforms affect growth, morphogenesis and
antifungal resistance. Mol Microbiol 55:546–560
221. Sellam A, Askew C, Epp E, Tebbji F, Mullick A, Whiteway M,
Nantel A (2010) Role of transcription factor CaNdt80p in cell
separation, hyphal growth, and virulence in Candida albicans.
Eukaryot Cell 9:634–644
222. Huang G, Wang H, Chou S, Nie X, Chen J, Liu H (2006) Bis-
table expression of WOR1, a master regulator of white-opaque
switching in Candida albicans. Proc Natl Acad Sci USA
103:12813–12818
E. J. Polvi et al.
123
223. Miller MG, Johnson AD (2002) White-opaque switching in
Candida albicans is controlled by mating-type locus home-
odomain proteins and allows efficient mating. Cell 110:293–302
224. Forche A, Alby K, Schaefer D, Johnson AD, Berman J, Bennett
RJ (2008) The parasexual cycle in Candida albicans provides an
alternative pathway to meiosis for the formation of recombinant
strains. PLoS Biol 6:e110
225. Kvaal C, Lachke SA, Srikantha T, Daniels K, McCoy J, Soll DR
(1999) Misexpression of the opaque-phase-specific gene PEP1
(SAP1) in the white phase of Candida albicans confers increased
virulence in a mouse model of cutaneous infection. Infect Im-
mun 67:6652–6662
226. Si H, Hernday AD, Hirakawa MP, Johnson AD, Bennett RJ
(2013) Candida albicans white and opaque cells undergo dis-
tinct programs of filamentous growth. PLoS Pathog 9:e1003210
227. Sasse C, Hasenberg M, Weyler M, Gunzer M, Morschhauser J
(2013) White-opaque switching of Candida albicans allows
immune evasion in an environment-dependent fashion. Eukaryot
Cell 12:50–58
228. Kolotila MP, Diamond RD (1990) Effects of neutrophils and
in vitro oxidants on survival and phenotypic switching of Can-
dida albicans WO-1. Infect Immun 58:1174–1179
229. Tao L, Du H, Guan G, Dai Y, Nobile CJ, Liang W, Cao C,
Zhang Q, Zhong J, Huang G (2014) Discovery of a ‘‘white-gray-
opaque’ tristable phenotypic switching system in Candida al-
bicans: roles of non-genetic diversity in host adaptation. PLoS
Biol 12:e1001830
230. Okagaki LH, Strain AK, Nielsen JN, Charlier C, Baltes NJ,
Chretien F, Heitman J, Dromer F, Nielsen K (2010) Crypto-
coccal cell morphology affects host cell interactions and
pathogenicity. PLoS Pathog 6:e1000953
231. Zaragoza O, Garcia-Rodas R, Nosanchuk JD, Cuenca-Estrella
M, Rodriguez-Tudela JL, Casadevall A (2010) Fungal cell gi-
gantism during mammalian infection. PLoS Pathog 6:e1000945
232. Zaragoza O, Nielsen K (2013) Titan cells in Cryptococcus
neoformans: cells with a giant impact. Curr Opin Microbiol
16:409–413
233. Okagaki LH, Nielsen K (2012) Titan cells confer protection
from phagocytosis in Cryptococcus neoformans infections.
Eukaryot Cell 11:820–826
234. Crabtree JN, Okagaki LH, Wiesner DL, Strain AK, Nielsen JN,
Nielsen K (2012) Titan cell production enhances the virulence
of Cryptococcus neoformans. Infect Immun 80:3776–3785
235. Kojic EM, Darouiche RO (2004) Candida infections of medical
devices. Clin Microbiol Rev 17:255–267
236. Dongari-Bagtzoglou A, Kashleva H, Dwivedi P, Diaz P, Vasi-
lakos J (2009) Characterization of mucosal Candida albicans
biofilms. PLoS ONE 4:e7967
237. Finkel JS, Mitchell AP (2011) Genetic control of Candida al-
bicans biofilm development. Nat Rev Microbiol 9:109–118
238. Douglas LJ (2003) Candida biofilms and their role in infection.
Trends Microbiol 11:30–36
239. Zarnowski R, Westler WM, Lacmbouh GA, Marita JM, Bothe
JR, Bernhardt J, Lounes-Hadj Sahraoui A, Fontaine J, Sanchez
H, Hatfield RD, Ntambi JM, Nett JE, Mitchell AP, Andes DR
(2014) Novel entries in a fungal biofilm matrix encyclopedia.
MBio 5:e01333-14
240. Taff HT, Nett JE, Zarnowski R, Ross KM, Sanchez H, Cain MT,
Hamaker J, Mitchell AP, Andes DR (2012) A Candida biofilm-
induced pathway for matrix glucan delivery: implications for
drug resistance. PLoS Pathog 8(8):e1002848
241. Nett J, Lincoln L, Marchillo K, Massey R, Holoyda K, Hoff B,
VanHandel M, Andes D (2007) Putative role of b-1,3 glucans in
Candida albicans biofilm resistance. Antimicrob Agents Che-
mother 51:510–520
242. Taff HT, Mitchell KF, Edward JA, Andes DR (2013) Mechan-
isms of Candida biofilm drug resistance. Future Microbiol
8:1325–1337
243. Morrow B, Srikantha T, Soll DR (1992) Transcription of the
gene for a pepsinogen, PEP1, is regulated by white-opaque
switching in Candida albicans. Mol Cell Biol 12:2997–3005
244. Naglik JR, Challacombe SJ, Hube B (2003) Candida albicans
secreted aspartyl proteinases in virulence and pathogenesis.
Microbiol Mol Biol Rev 67:400–428
245. de Repentigny L, Aumont F, Bernard K, Belhumeur P (2000)
Characterization of binding of Candida albicans to small in-
testinal mucin and its role in adherence to mucosal epithelial
cells. Infect Immun 68:3172–3179
246. Schaller M, Borelli C, Korting HC, Hube B (2005) Hydrolytic
enzymes as virulence factors of Candida albicans. Mycoses
48:365–377
247. Leidich SD, Ibrahim AS, Fu Y, Koul A, Jessup C, Vitullo J,
Fonzi W, Mirbod F, Nakashima S, Nozawa Y, Ghannoum MA
(1998) Cloning and disruption of caPLB1, a phospholipase B
gene involved in the pathogenicity of Candida albicans. J Biol
Chem 273:26078–26086
248. Sebghati TS, Engle JT, Goldman WE (2000) Intracellular
parasitism by Histoplasma capsulatum: fungal virulence and
calcium dependence. Science 290:1368–1372
249. Beck MR, Dekoster GT, Cistola DP, Goldman WE (2009) NMR
structure of a fungal virulence factor reveals structural ho-
mology with mammalian saposin B. Mol Microbiol 72:344–353
250. Bohse ML, Woods JP (2007) RNA interference-mediated si-
lencing of the YPS3 gene of Histoplasma capsulatum reveals
virulence defects. Infect Immun 75:2811–2817
251. O’Meara TR, Alspaugh JA (2012) The Cryptococcus neofor-
mans capsule: a sword and a shield. Clin Microbiol Rev
25:387–408
252. Klutts JS, Doering TL (2008) Cryptococcal xylosyltransferase 1
(Cxt1p) from Cryptococcus neoformans plays a direct role in the
synthesis of capsule polysaccharides. J Biol Chem 283:14327–
14334
253. Kozel TR, Levitz SM, Dromer F, Gates MA, Thorkildson P,
Janbon G (2003) Antigenic and biological characteristics of
mutant strains of Cryptococcus neoformans lacking capsular
O acetylation or xylosyl side chains. Infect Immun 71:2868–
2875
254. Jong AY, Wu C-H, Chen H-M, Luo F, Kwon-Chung KJ, Chang
YC, LaMunyon CW, Plaas A, Huang S-H (2007) Identification
and characterization of CPS1 as a hyaluronic acid synthase
contributing to the pathogenesis of Cryptococcus neoformans
infection. Eukaryot Cell 6(8):1486–1496
255. McFadden DC, Fries BC, Wang F, Casadevall A (2007) Capsule
structural heterogeneity and antigenic variation in Cryptococcus
neoformans. Eukaryot Cell 6:1464–1473
256. Rivera J, Feldmesser M, Cammer M, Casadevall A (1998) Or-
gan-dependent variation of capsule thickness in Cryptococcus
neoformans during experimental murine infection. Infect Immun
66:5027–5030
257. Palmer DA, Thompson JK, Li L, Prat A, Wang P (2006) Gib2, a
novel Gb-like/RACK1 homolog, functions as a Gbsubunit in
cAMP signaling and is essential in Cryptococcus neoformans.
J Biol Chem 281:32596–32605
258. Xue C, Hsueh YP, Chen L, Heitman J (2008) The RGS protein
Crg2 regulates both pheromone and cAMP signalling in Cryp-
tococcus neoformans. Mol Microbiol 70:379–395
259. Xue C, Bahn YS, Cox GM, Heitman J (2006) G protein-coupled
receptor Gpr4 senses amino acids and activates the cAMP-PKA
pathway in Cryptococcus neoformans. Mol Biol Cell
17:667–679
Opportunistic yeast pathogens
123
260. Vartivarian SE, Anaissie EJ, Cowart RE, Sprigg HA, Tingler
MJ, Jacobson ES (1993) Regulation of cryptococcal capsular
polysaccharide by iron. J Infect Dis 167:186–190
261. Klengel T, Liang W-J, Chaloupka J, Ruoff C, Schro
¨ppel K,
Naglik JR, Eckert SE, Mogensen EG, Haynes K, Tuite MF,
Levin LR, Buck J, Muhlschlegel FA (2005) Fungal adenylyl
cyclase integrates CO
2
sensing with cAMP signaling and viru-
lence. Curr Biol 15:2021–2026
262. Zaragoza O, Casadevall A (2004) Experimental modulation of
capsule size in Cryptococcus neoformans. Biol Proced Online
6:10–15
263. Vecchiarelli A, Pericolini E, Gabrielli E, Kenno S, Perito S,
Cenci E, Monari C (2013) Elucidating the immunological
function of the Cryptococcus neoformans capsule. Future Mi-
crobiol 8:1107–1116
264. Coelho C, Bocca AL, Casadevall A (2014) The tools for viru-
lence of Cryptococcus neoformans. Adv Appl Microbiol
87:1–41
265. Rappleye CA, Engle JT, Goldman WE (2004) RNA interference
in Histoplasma capsulatum demonstrates a role for a-(1,3)-
glucan in virulence. Mol Microbiol 53:153–165
266. Rappleye CA, Eissenberg LG, Goldman WE (2007) Histoplas-
ma capsulatum a-(1,3)-glucan blocks innate immune
recognition by the b-glucan receptor. Proc Natl Acad Sci USA
104:1366–1370
267. Long KH, Gomez FJ, Morris RE, Newman SL (2003) Identifi-
cation of heat shock protein 60 as the ligand on Histoplasma
capsulatum that mediates binding to CD18 receptors on human
macrophages. J Immunol 170:487–494
268. Ehlers MR (2000) CR3: a general purpose adhesion-recognition
receptor essential for innate immunity. Microbes Infect
2:289–294
269. Mihu MR, Nosanchuk JD (2012) Histoplasma virulence and
host responses. Int J Microbiol. doi:10.1155/2012/268123
270. Ma H, Croudace JE, Lammas DA, May RC (2006) Expulsion of
live pathogenic yeast by macrophages. Curr Biol 16:2156–2160
271. Nicola AM, Robertson EJ, Albuquerque P, Derengowski Lda S,
Casadevall A (2011) Nonlytic exocytosis of Cryptococcus
neoformans from macrophages occurs in vivo and is influenced
by phagosomal pH. MBio 2:e00167-11
272. Alvarez M, Casadevall A (2006) Phagosome extrusion and host-
cell survival after Cryptococcus neoformans phagocytosis by
macrophages. Curr Biol 16:2161–2165
273. Bain JM, Lewis LE, Okai B, Quinn J, Gow NA, Erwig LP
(2012) Non-lytic expulsion/exocytosis of Candida albicans
from macrophages. Fungal Genet Biol 49:677–678
274. Uwamahoro N, Verma-Gaur J, Shen HH, Qu Y, Lewis R, Lu J,
Bambery K, Masters SL, Vince JE, Naderer T, Traven A (2014)
The pathogen Candida albicans hijacks pyroptosis for escape
from macrophages. MBio 5:e00003–e00014
275. Wellington M, Koselny K, Sutterwala FS, Krysan DJ (2014)
Candida albicans triggers NLRP3-mediated pyroptosis in mac-
rophages. Eukaryot Cell 13:329–340
276. Hromatka BS, Noble SM, Johnson AD (2005) Transcriptional
response of Candida albicans to nitric oxide and the role of the
YHB1 gene in nitrosative stress and virulence. Mol Biol Cell
16:4814–4826
277. Ullmann BD, Myers H, Chiranand W, Lazzell AL, Zhao Q,
Vega LA, Lopez-Ribot JL, Gardner PR, Gustin MC (2004) In-
ducible defense mechanism against nitric oxide in Candida
albicans. Eukaryot Cell 3:715–723
278. de Jesus-Berrios M, Liu L, Nussbaum JC, Cox GM, Stamler JS,
Heitman J (2003) Enzymes that counteract nitrosative stress
promote fungal virulence. Curr Biol 13:1963–1968
279. Missall TA, Pusateri ME, Donlin MJ, Chambers KT, Corbett JA,
Lodge JK (2006) Posttranslational, translational, and
transcriptional responses to nitric oxide stress in Cryptococcus
neoformans: implications for virulence. Eukaryot Cell
5:518–529
280. Nittler MP, Hocking-Murray D, Foo CK, Sil A (2005) Identi-
fication of Histoplasma capsulatum transcripts induced in
response to reactive nitrogen species. Mol Biol Cell
16:4792–4813
281. Nosanchuk JD, Valadon P, Feldmesser M, Casadevall A (1999)
Melanization of Cryptococcus neoformans in murine infection.
Mol Cell Biol 19:745–750
282. Huffnagle GB, Chen GH, Curtis JL, McDonald RA, Strieter
RM, Toews GB (1995) Down-regulation of the afferent phase of
T cell-mediated pulmonary inflammation and immunity by a
high melanin-producing strain of Cryptococcus neoformans.
J Immunol 155:3507–3516
283. Casadevall A, Rosas AL, Nosanchuk JD (2000) Melanin and
virulence in Cryptococcus neoformans. Curr Opin Microbiol
3:354–358
284. Rutherford JC (2014) The emerging role of urease as a general
microbial virulence factor. PLoS Pathog 10:e1004062
285. Olszewski MA, Noverr MC, Chen GH, Toews GB, Cox GM,
Perfect JR, Huffnagle GB (2004) Urease expression by Cryp-
tococcus neoformans promotes microvascular sequestration,
thereby enhancing central nervous system invasion. Am J Pathol
164:1761–1771
286. Cox GM, Mukherjee J, Cole GT, Casadevall A, Perfect JR
(2000) Urease as a virulence factor in experimental cryptococ-
cosis. Infect Immun 68:443–448
287. Singh A, Panting RJ, Varma A, Saijo T, Waldron KJ, Jong A,
Ngamskulrungroj P, Chang YC, Rutherford JC, Kwon-Chung
KJ (2013) Factors required for activation of urease as a viru-
lence determinant in Cryptococcus neoformans. MBio
4:e00220-13
288. Osterholzer JJ, Surana R, Milam JE, Montano GT, Chen GH,
Sonstein J, Curtis JL, Huffnagle GB, Toews GB, Olszewski MA
(2009) Cryptococcal urease promotes the accumulation of im-
mature dendritic cells and a non-protective T2 immune response
within the lung. Am J Pathol 174:932–943
289. Navarathna DH, Das A, Morschhauser J, Nickerson KW,
Roberts DD (2011) Dur3 is the major urea transporter in Can-
dida albicans and is co-regulated with the urea amidolyase
Dur1,2. Microbiology 157:270–279
290. Navarathna DH, Lionakis MS, Lizak MJ, Munasinghe J, Nick-
erson KW, Roberts DD (2012) Urea amidolyase (DUR1,2)
contributes to virulence and kidney pathogenesis of Candida
albicans. PLoS ONE 7:e48475
291. Ghosh S, Navarathna DH, Roberts DD, Cooper JT, Atkin AL,
Petro TM, Nickerson KW (2009) Arginine-induced germ tube
formation in Candida albicans is essential for escape from
murine macrophage line RAW 264.7. Infect Immun
77:1596–1605
292. Romani L (2011) Immunity to fungal infections. Nat Rev Im-
munol 11:275–288
293. Netea MG, Brown GD, Kullberg BJ, Gow NA (2008) An inte-
grated model of the recognition of Candida albicans by the
innate immune system. Nat Rev Microbiol 6:67–78
294. Romani L (2004) Immunity to fungal infections. Nat Rev Im-
munol 4:1–23
295. Latge JP (2010) Tasting the fungal cell wall. Cell Microbiol
12:863–872
296. Lionakis MS, Netea MG (2013) Candida and host determinants
of susceptibility to invasive candidiasis. PLoS Pathog
9:e1003079
297. Gow N, Netea M, Munro CA, Ferwerda G, Bates S, Mora-
Montes H, Walker LA, Jansen T, Jacobs L, Tsoni V, Brown G,
Odds FC, Van der Meer JW, Brown A, Kullberg BJ (2007)
E. J. Polvi et al.
123
Immune recognition of Candida albicans b-glucan by dectin 1.
J Infect Dis 196(10):1565–1571
298. Netea MG, Gow NA, Munro CA, Bates S, Collins C, Ferwerda
G, Hobson RP, Bertram G, Hughes HB, Jansen T, Jacobs L,
Buurman ET, Gijzen K, Williams DL, Torensma R, McKinnon
A, MacCallum DM, Odds FC, Van der Meer JW, Brown AJ,
Kullberg BJ (2006) Immune sensing of Candida albicans re-
quires cooperative recognition of mannans and glucans by lectin
and toll-like receptors. J Clin Invest 116:1642–1650
299. Dan JM, Wang JP, Lee CK, Levitz SM (2008) Cooperative
stimulation of dendritic cells by Cryptococcus neoformans
mannoproteins and CpG oligodeoxynucleotides. PLoS ONE
3(4):e2046
300. Brown GD (2011) Innate antifungal immunity: the key role of
phagocytes. Annu Rev Immunol 29:1–21
301. Bhatt VR, Viola GM, Ferrajoli A (2011) Invasive fungal in-
fections in acute leukemia. Ther Adv Hematol 2:231–247
302. Lionakis MS, Fischer BG, Lim JK, Swamydas M, Wan W,
Richard Lee CC, Cohen JI, Scheinberg P, Gao JL, Murphy PM
(2012) Chemokine receptor Ccr1 drives neutrophil-mediated
kidney immunopathology and mortality in invasive candidiasis.
PLoS Pathog 8:e1002865
303. van de Veerdonk FL, Netea MG (2010) T-cell subsets and an-
tifungal host defenses. Curr Fungal Infect Rep 4:238–243
304. Armstrong-James D, Meintjes G, Brown GD (2014) A neglected
epidemic: fungal infections in HIV/AIDS. Trends Microbiol
22:120–127
305. Pensieroso S, Galli L, Nozza S, Ruffin N, Castagna A, Tambussi
G, Hejdeman B, Misciagna D, Riva A, Malnati M, Chiodi F,
Scarlatti G (2013) B-cell subset alterations and correlated fac-
tors in HIV-1 infection. AIDS 27:1209–1217
306. Collini P, Noursadeghi M, Sabroe I, Miller RF, Dockrell DH
(2010) Monocyte and macrophage dysfunction as a cause of
HIV-1 induced dysfunction of innate immunity. Curr Mol Med
10:727–740
307. Olszewski MA, Huffnagle GB, McDonald RA, Lindell DM,
Moore BB, Cook DN, Toews GB (2000) The role of macro-
phage inflammatory protein-1 a/CCL3 in regulation of T cell-
mediated immunity to Cryptococcus neoformans infection.
J Immunol 165:6429–6436
308. Wormley FL Jr, Perfect JR, Steele C, Cox GM (2007) Protection
against cryptococcosis by using a murine gamma interferon-
producing Cryptococcus neoformans strain. Infect Immun
75:1453–1462
309. Hole CR, Wormley FL Jr (2012) Vaccine and im-
munotherapeutic approaches for the prevention of
cryptococcosis: lessons learned from animal models. Front Mi-
crobiol 3:291
310. Conti HR, Shen F, Nayyar N, Stocum E, Sun JN, Lindemann
MJ, Ho AW, Hai JH, Yu JJ, Jung JW, Filler SG, Masso-Welch
P, Edgerton M, Gaffen SL (2009) Th17 cells and IL-17 receptor
signaling are essential for mucosal host defense against oral
candidiasis. J Exp Med 206:299–311
311. Bixler SL, Mattapallil JJ (2013) Loss and dysregulation of Th17
cells during HIV infection. Clin Dev Immunol 2013:852418
312. Rudner XL, Happel KI, Young EA, Shellito JE (2007) Inter-
leukin-23 (IL-23)-IL-17 cytokine axis in murine Pneumocystis
carinii infection. Infect Immun 75:3055–3061
313. Netea MG, Sutmuller R, Hermann C, Van der Graaf CA, Van
der Meer JW, van Krieken JH, Hartung T, Adema G, Kullberg
BJ (2004) Toll-like receptor 2 suppresses immunity against
Candida albicans through induction of IL-10 and regulatory T
cells. J Immunol 172:3712–3718
314. Kroetz DN, Deepe GS Jr (2010) CCR5 dictates the equilibrium
of proinflammatory IL-17
?
and regulatory Foxp3
?
T cells in
fungal infection. J Immunol 184:5224–5231
315. Gildea LA, Gibbons R, Finkelman FD, Deepe GS Jr (2003)
Overexpression of interleukin-4 in lungs of mice impairs
elimination of Histoplasma capsulatum. Infect Immun
71:3787–3793
316. Cenci E, Mencacci A, Del Sero G, Bacci A, Montagnoli C,
d’Ostiani CF, Mosci P, Bachmann M, Bistoni F, Kopf M, Ro-
mani L (1999) Interleukin-4 causes susceptibility to invasive
pulmonary aspergillosis through suppression of protective type I
responses. J Infect Dis 180:1957–1968
317. Myers RC, Dunaway CW, Nelson MP, Trevor JL, Morris A,
Steele C (2013) STAT4-dependent and -independent Th2 re-
sponses correlate with protective immunity against lung
infection with Pneumocystis murina. J Immunol 190:6287–6294
318. Bicanic T, Meintjes G, Rebe K, Williams A, Loyse A, Wood R,
Hayes M, Jaffar S, Harrison TS (2009) Immune reconstitution
inflammatory syndrome in HIV-associated cryptococcal
meningitis: a prospective study. J Acquir Immune Defic Syndr
51(2):130–134
319. Perfect JR (2012) The impact of the host on fungal infections.
Am J Med 125:S39–S51
320. Boulware DR, Meya DB, Bergemann TL, Wiesner DL, Rhein J,
Musubire A, Lee SJ, Kambugu A, Janoff EN, Bohjanen PR
(2010) Clinical features and serum biomarkers in HIV immune
reconstitution inflammatory syndrome after cryptococcal
meningitis: a prospective cohort study. PLoS Med 7:e1000384
321. Lionakis MS (2012) Genetic susceptibility to fungal infections
in humans. Curr Fungal Infect Rep 6:11–22
322. Ferwerda B, Ferwerda G, Plantinga TS, Willment JA, van Spriel
AB, Venselaar H, Elbers CC, Johnson MD, Cambi A, Huysamen
C, Jacobs L, Jansen T, Verheijen K, Masthoff L, Morre SA,
Vriend G, Williams DL, Perfect JR, Joosten LA, Wijmenga C,
van der Meer JW, Adema GJ, Kullberg BJ, Brown GD, Netea
MG (2009) Human dectin-1 deficiency and mucocutaneous
fungal infections. N Engl J Med 361:1760–1767
323. Plantinga TS, van der Velden WJ, Ferwerda B, van Spriel AB,
Adema G, Feuth T, Donnelly JP, Brown GD, Kullberg BJ,
Blijlevens NM, Netea MG (2009) Early stop polymorphism in
human DECTIN-1 is associated with increased Candida
colonization in hematopoietic stem cell transplant recipients.
Clin Infect Dis 49:724–732
324. Glocker EO, Hennigs A, Nabavi M, Schaffer AA, Woellner C,
Salzer U, Pfeifer D, Veelken H, Warnatz K, Tahami F, Jamal S,
Manguiat A, Rezaei N, Amirzargar AA, Plebani A, Hanness-
chlager N, Gross O, Ruland J, Grimbacher B (2009) A
homozygous CARD9 mutation in a family with susceptibility to
fungal infections. N Engl J Med 361:1727–1735
325. Liu L, Okada S, Kong XF, Kreins AY, Cypowyj S, Abhyankar
A, Toubiana J, Itan Y, Audry M, Nitschke P, Masson C, Toth B,
Flatot J, Migaud M, Chrabieh M, Kochetkov T, Bolze A, Bor-
ghesi A, Toulon A, Hiller J, Eyerich S, Eyerich K, Gulacsy V,
Chernyshova L, Chernyshov V, Bondarenko A, Grimaldo RM,
Blancas-Galicia L, Beas IM, Roesler J, Magdorf K, Engelhard
D, Thumerelle C, Burgel PR, Hoernes M, Drexel B, Seger R,
Kusuma T, Jansson AF, Sawalle-Belohradsky J, Belohradsky B,
Jouanguy E, Bustamante J, Bue M, Karin N, Wildbaum G,
Bodemer C, Lortholary O, Fischer A, Blanche S, Al-Muhsen S,
Reichenbach J, Kobayashi M, Rosales FE, Lozano CT, Kilic SS,
Oleastro M, Etzioni A, Traidl-Hoffmann C, Renner ED, Abel L,
Picard C, Marodi L, Boisson-Dupuis S, Puel A, Casanova JL
(2011) Gain-of-function human STAT1 mutations impair IL-17
immunity and underlie chronic mucocutaneous candidiasis.
J Exp Med 208:1635–1648
326. van de Veerdonk FL, Plantinga TS, Hoischen A, Smeekens SP,
Joosten LA, Gilissen C, Arts P, Rosentul DC, Carmichael AJ,
Smits-van der Graaf CA, Kullberg BJ, van der Meer JW, Lilic
D, Veltman JA, Netea MG (2011) STAT1 mutations in
Opportunistic yeast pathogens
123
autosomal dominant chronic mucocutaneous candidiasis. N Engl
J Med 365:54–61
327. Puel A, Cypowyj S, Bustamante J, Wright JF, Liu L, Lim HK,
Migaud M, Israel L, Chrabieh M, Audry M, Gumbleton M,
Toulon A, Bodemer C, El-Baghdadi J, Whitters M, Paradis T,
Brooks J, Collins M, Wolfman NM, Al-Muhsen S, Galicchio M,
Abel L, Picard C, Casanova JL (2011) Chronic mucocutaneous
candidiasis in humans with inborn errors of interleukin-17 im-
munity. Science 332:65–68
328. Kagami S, Rizzo HL, Kurtz SE, Miller LS, Blauvelt A (2010)
IL-23 and IL-17A, but not IL-12 and IL-22, are required for
optimal skin host defense against Candida albicans. J Immunol
185:5453–5462
329. Minegishi Y, Saito M, Tsuchiya S, Tsuge I, Takada H, Hara T,
Kawamura N, Ariga T, Pasic S, Stojkovic O, Metin A, Kara-
suyama H (2007) Dominant-negative mutations in the DNA-
binding domain of STAT3 cause hyper-IgE syndrome. Nature
448:1058–1062
330. Milner JD, Brenchley JM, Laurence A, Freeman AF, Hill BJ,
Elias KM, Kanno Y, Spalding C, Elloumi HZ, Paulson ML,
Davis J, Hsu A, Asher AI, O’Shea J, Holland SM, Paul WE,
Douek DC (2008) Impaired T(H)17 cell differentiation in sub-
jects with autosomal dominant hyper-IgE syndrome. Nature
452:773–776
331. Engelhardt KR, McGhee S, Winkler S, Sassi A, Woellner C,
Lopez-Herrera G, Chen A, Kim HS, Lloret MG, Schulze I, Ehl
S, Thiel J, Pfeifer D, Veelken H, Niehues T, Siepermann K,
Weinspach S, Reisli I, Keles S, Genel F, Kutukculer N, Cam-
cioglu Y, Somer A, Karakoc-Aydiner E, Barlan I, Gennery A,
Metin A, Degerliyurt A, Pietrogrande MC, Yeganeh M, Baz Z,
Al-Tamemi S, Klein C, Puck JM, Holland SM, McCabe ER,
Grimbacher B, Chatila TA (2009) Large deletions and point
mutations involving the dedicator of cytokinesis 8 (DOCK8) in
the autosomal-recessive form of hyper-IgE syndrome. J Allergy
Clin Immunol 124:1289–1302
332. Mathis D, Benoist C (2009) Aire. Annu Rev Immunol
27:287–312
333. Kisand K, Boe Wolff AS, Podkrajsek KT, Tserel L, Link M,
Kisand KV, Ersvaer E, Perheentupa J, Erichsen MM, Bratanic
N, Meloni A, Cetani F, Perniola R, Ergun-Longmire B, Ma-
claren N, Krohn KJ, Pura M, Schalke B, Strobel P, Leite MI,
Battelino T, Husebye ES, Peterson P, Willcox N, Meager A
(2010) Chronic mucocutaneous candidiasis in APECED or
thymoma patients correlates with autoimmunity to Th17-asso-
ciated cytokines. J Exp Med 207:299–308
334. Segal BH, Leto TL, Gallin JI, Malech HL, Holland SM (2000)
Genetic, biochemical, and clinical features of chronic granulo-
matous disease. Medicine (Baltimore) 79:170–200
335. Winkelstein JA, Marino MC, Johnston RB Jr, Boyle J, Curnutte
J, Gallin JI, Malech HL, Holland SM, Ochs H, Quie P, Buckley
RH, Foster CB, Chanock SJ, Dickler H (2000) Chronic
granulomatous disease. Report on a national registry of 368
patients. Medicine (Baltimore) 79:155–169
336. Aratani Y, Kura F, Watanabe H, Akagawa H, Takano Y, Suzuki
K, Dinauer MC, Maeda N, Koyama H (2002) Critical role of
myeloperoxidase and nicotinamide adenine dinucleotide phos-
phate-oxidase in high-burden systemic infection of mice with
Candida albicans. J Infect Dis 185:1833–1837
337. Cech P, Papathanassiou A, Boreux G, Roth P, Miescher PA
(1979) Hereditary myeloperoxidase deficiency. Blood
53:403–411
338. Vinh DC, Patel SY, Uzel G, Anderson VL, Freeman AF, Olivier
KN, Spalding C, Hughes S, Pittaluga S, Raffeld M, Sorbara LR,
Elloumi HZ, Kuhns DB, Turner ML, Cowen EW, Fink D, Long-
Priel D, Hsu AP, Ding L, Paulson ML, Whitney AR, Sampaio
EP, Frucht DM, DeLeo FR, Holland SM (2010) Autosomal
dominant and sporadic monocytopenia with susceptibility to
mycobacteria, fungi, papillomaviruses, and myelodysplasia.
Blood 115:1519–1529
339. Arvanitis M, Anagnostou T, Fuchs BB, Caliendo AM, Mylon-
akis E (2014) Molecular and nonmolecular diagnostic methods
for invasive fungal infections. Clin Microbiol Rev 27:490–526
340. Richardson MD, Warnock DW (2012) Fungal infection diag-
nosis and management, 4th edn. Wiley-Blackwell, Chichester,
West Sussex, UK
341. Nguyen MH, Wissel MC, Shields RK, Salomoni MA, Hao B,
Press EG, Shields RM, Cheng S, Mitsani D, Vadnerkar A, Sil-
veira FP, Kleiboeker SB, Clancy CJ (2012) Performance of
Candida real-time polymerase chain reaction, b-D-glucan assay,
and blood cultures in the diagnosis of invasive candidiasis. Clin
Infect Dis 54:1240–1248
342. Denning DW, Hope WW (2010) Therapy for fungal diseases:
opportunities and priorities. Trends Microbiol 18:195–204
343. Gomez-Lopez A, Zaragoza O, Rodriguez-Tudela JL, Cuenca-
Estrella M (2008) Pharmacotherapy of yeast infections. Expert
Opin Pharmacother 9:2801–2816
344. Ghannoum MA, Rice LB (1999) Antifungal agents: mode of
action, mechanisms of resistance, and correlation of these
mechanisms with bacterial resistance. Clin Microbiol Rev
12:501–517
345. Masia Canuto M, Gutierrez Rodero F (2002) Antifungal drug
resistance to azoles and polyenes. Lancet Infect Dis 2:550–563
346. Deray G (2002) Amphotericin B nephrotoxicity. J Antimicrob
Chemother 49(Suppl 1):37–41
347. Dupont B (2002) Overview of the lipid formulations of am-
photericin B. J Antimicrob Chemother 49(Suppl 1):31–36
348. Wong-Beringer A, Jacobs RA, Guglielmo BJ (1998) Lipid for-
mulations of amphotericin B: clinical efficacy and toxicities.
Clin Infect Dis 27:603–618
349. Vincent BM, Lancaster AK, Scherz-Shouval R, Whitesell L,
Lindquist S (2013) Fitness trade-offs restrict the evolution of
resistance to amphotericin B. PLoS Biol 11:e1001692
350. Ellis D (2002) Amphotericin B: spectrum and resistance. J An-
timicrob Chemother 49(Suppl 1):7–10
351. Sheehan DJ, Hitchcock CA, Sibley CM (1999) Current and
emerging azole antifungal agents. Clin Microbiol Rev 12:40–79
352. Walsh TJ, Anaissie EJ, Denning DW, Herbrecht R, Kon-
toyiannis DP, Marr KA, Morrison VA, Segal BH, Steinbach WJ,
Stevens DA, van Burik JA, Wingard JR, Patterson TF, Infec-
tious Diseases Society of America (2008) Treatment of
aspergillosis: clinical practice guidelines of the Infectious Dis-
eases Society of America. Clin Infect Dis 46:327–360
353. Pfaller MA (2012) Antifungal drug resistance: mechanisms,
epidemiology, and consequences for treatment. Am J Med
125:S3–S13
354. Pound MW, Townsend ML, Drew RH (2010) Echinocandin
pharmacodynamics: review and clinical implications. J Antimi-
crob Chemother 65:1108–1118
355. Denning DW (2003) Echinocandin antifungal drugs. Lancet
362:1142–1151
356. Holt SL, Drew RH (2011) Echinocandins: addressing out-
standing questions surrounding treatment of invasive fungal
infections. Am J Health Syst Pharm 68:1207–1220
357. Arendrup MC, Perlin DS (2014) Echinocandin resistance: an
emerging clinical problem? Curr Opin Infect Dis 27:484–492
358. Nanjappa SG, Klein BS (2014) Vaccine immunity against fun-
gal infections. Curr Opin Immunol 28:27–33
359. Spellberg B (2011) Vaccines for invasive fungal infections.
F1000 Med Rep 3:13
360. NovaDigm Therapeutics, Inc (2013) Safety, tolerability, im-
munogenicity and efficacy of NDV-3A vaccine in preventing
recurrent vulvovaginal candidiasis. Identifier: NCT01926028
E. J. Polvi et al.
123
361. Schmidt CS, White CJ, Ibrahim AS, Filler SG, Fu Y, Yeaman
MR, Edwards JE Jr, Hennessey JP Jr (2012) NDV-3, a recom-
binant alum-adjuvanted vaccine for Candida and
Staphylococcus aureus, is safe and immunogenic in healthy
adults. Vaccine 30:7594–7600
362. De Bernardis F, Amacker M, Arancia S, Sandini S, Gremion C,
Zurbriggen R, Moser C, Cassone A (2012) A virosomal vaccine
against candidal vaginitis: immunogenicity, efficacy and safety
profile in animal models. Vaccine 30:4490–4498
Opportunistic yeast pathogens
123
... Additionally, iatrogenic immune suppression of various types, grades, and timings as well as genetic host patterns might influence the risk as well as the clinical characteristics of opportunistic infections (Polvi et al., 2015). In the context of epidemiological studies in a specific patient group with a congenital or acquired condition (such as HIV illness, antineoplastic treatment, transplant, etc.), Opportunistic infections frequently make the news of immune system dysfunction. ...
Article
Full-text available
The receptive patient and the bacterium serve as the two defining criteria for opportunistic bacteria. Theoretically, no saprophytic or typical commensal microbes can infiltrate a healthy receptive person. Only specific "undesirable" commensal species, "such as Vargues' Specific Pathogenic Bacteria", can infect this person. Several species from the typical commensal flora, or opportunistic bacteria in the wide meaning of the word, “may infiltrate an otherwise healthy host if their immune defences temporarily deteriorate”. Even species that were previously thought to be non-virulent may assault an immunosuppressed patient with a significant and protracted immune system depression, “including various saprophytic and commensal microbes”. Several bacteria that are typically found in water, food, and the air have recently become opportunistic pathogens in both people and animals. The issue is made more difficult by the introduction of many antibiotic-resistant strains of these opportunistic pathogens, which make hospital-acquired infections in susceptible hosts challenging to treat in the setting of illness.
... The types of virulence exhibited by fungi can be classified in two ways, including primary infections and opportunistic infections [20,50,71]. Primary infections occur in an immunologically normal host (healthy host) and usually result from the inhalation of fungal spores, which can lead to pneumonia as the primary symptom. ...
Article
Full-text available
Human fungal diseases are infections caused by any fungus that invades human tissues, causing superficial, subcutaneous, or systemic diseases. Fungal infections that enter various human tissues and organs pose a significant threat to millions of individuals with weakened immune systems globally. Over recent decades, the reported cases of invasive fungal infections have increased substantially and research progress in this field has also been rapidly boosted. This review provides a comprehensive list of human fungal pathogens extracted from over 850 recent case reports, and a summary of the relevant disease conditions and their origins. Details of 281 human fungal pathogens belonging to 12 classes and 104 genera in the divisions ascomycota, basidiomycota, entomoph-thoromycota, and mucoromycota are listed. Among these, Aspergillus stands out as the genus with the greatest potential of infecting humans, comprising 16 species known to infect humans. Additionally, three other genera, Curvularia, Exophiala, and Trichophyton, are recognized as significant genera, each comprising 10 or more known human pathogenic species. A phylogenetic analysis based on partial sequences of the 28S nrRNA gene (LSU) of human fungal pathogens was performed to show their phylogenetic relationships and clarify their taxonomies. In addition, this review summarizes the recent advancements in fungal disease diagnosis and therapeutics.
... F ungi pose a major threat to human health, infecting over a billion people and claiming 1.5 million lives worldwide annually (1). Candida albicans is a member of the normal human mucosal microbiota, where, as an opportunistic pathogen, it can cause diseases ranging from superficial infections to life-threatening systemic candidiasis in immunocompromised individuals, with mortality rates upward of 40% (1,2). C. albicans virulence is enabled in part by its ability to transition between yeast and filamentous growth states. ...
Article
Full-text available
For the human fungal pathogen Candida albicans , metabolic flexibility and the ability to transition between yeast and filamentous growth states are key virulence traits that enable disease in the host. These traits are particularly important during the interaction of C. albicans with macrophages, where the fungus must utilize multiple alternative carbon sources to survive after being phagocytosed, and filamentation is coupled to fungal escape and immune cell death. Here, we employed functional genomic screening of conditional-expression mutants covering >50% of the C. albicans genome to identify genes selectively required for filamentation inside macrophages. Through manual and machine learning-based image analyses, we uncovered a role for the mitochondrial ribosome, respiration, and the SNF1 AMP-activated kinase complex in governing filamentous growth within the phagosome, suggesting that C. albicans relies on respiration to evade the antifungal activities of macrophages. We demonstrate that downregulating the expression of these genes reduces ATP levels and impedes filamentation as well as growth under monoculture conditions in medium lacking glucose. In co-culture with physiological glucose concentration, downregulation of genes involved in mitochondrial function and respiration prevented C. albicans from expanding within the phagosome, escaping, and inducing immune cell death. Together, our work provides new insights into the impact of metabolism on the interaction between C. albicans and macrophages, highlighting respiration and the SNF1 AMP-activated kinase as key effectors of C. albicans metabolic flexibility and filamentation within phagocytes. IMPORTANCE Candida albicans is a leading human fungal pathogen that often causes life-threatening infections in immunocompromised individuals. The ability of C. albicans to transition between yeast and filamentous forms is key to its virulence, and this occurs in response to many host-relevant cues, including engulfment by host macrophages. While previous efforts identified C. albicans genes required for filamentation in other conditions, the genes important for this morphological transition upon internalization by macrophages remained largely enigmatic. Here, we employed a functional genomic approach to identify genes that enable C. albicans filamentation within macrophages and uncovered a role for the mitochondrial ribosome, respiration, and the SNF1 AMP-activated kinase complex. Additionally, we showed that glucose uptake and glycolysis by macrophages support C. albicans filamentation. This work provides insights into the metabolic dueling that occurs during the interaction of C. albicans with macrophages and identifies vulnerabilities in C. albicans that could serve as promising therapeutic targets.
... Candida albicans is a common colonizer of humans, with between 20% and 80% of the world's population estimated to be asymptomatically colonized at any given time [1,2], although this depends on many factors, including host health status and diet [3][4][5]. Colonization occurs at multiple body sites including the mouth, skin, GI tract, and vaginal tract [6,7]. ...
Article
Full-text available
Candida albicans is a frequent colonizer of human mucosal surfaces as well as an opportunistic pathogen. C. albicans is remarkably versatile in its ability to colonize diverse host sites with differences in oxygen and nutrient availability, pH, immune responses, and resident microbes, among other cues. It is unclear how the genetic background of a commensal colonizing population can influence the shift to pathogenicity. Therefore, we examined 910 commensal isolates from 35 healthy donors to identify host niche-specific adaptations. We demonstrate that healthy people are reservoirs for genotypically and phenotypically diverse C. albicans strains. Using limited diversity exploitation, we identified a single nucleotide change in the uncharacterized ZMS1 transcription factor that was sufficient to drive hyper invasion into agar. We found that SC5314 was significantly different from the majority of both commensal and bloodstream isolates in its ability to induce host cell death. However, our commensal strains retained the capacity to cause disease in the Galleria model of systemic infection, including outcompeting the SC5314 reference strain during systemic competition assays. This study provides a global view of commensal strain variation and within-host strain diversity of C. albicans and suggests that selection for commensalism in humans does not result in a fitness cost for invasive disease.
Article
Full-text available
In exploring the evolutionary trajectories of both pathogenesis and karyotype dynamics in fungi, we conducted a large-scale comparative genomic analysis spanning the Cryptococcus genus, encompassing both global human fungal pathogens and nonpathogenic species, and related species from the sister genus Kwoniella. Chromosome-level genome assemblies were generated for multiple species, covering virtually all known diversity within these genera. Although Cryptococcus and Kwoniella have comparable genome sizes (about 19.2 and 22.9 Mb) and similar gene content, hinting at preadaptive pathogenic potential, our analysis found evidence of gene gain (via horizontal gene transfer) and gene loss in pathogenic Cryptococcus species, which might represent evolutionary signatures of pathogenic development. Genome analysis also revealed a significant variation in chromosome number and structure between the 2 genera. By combining synteny analysis and experimental centromere validation, we found that most Cryptococcus species have 14 chromosomes, whereas most Kwoniella species have fewer (11, 8, 5, or even as few as 3). Reduced chromosome number in Kwoniella is associated with formation of giant chromosomes (up to 18 Mb) through repeated chromosome fusion events, each marked by a pericentric inversion and centromere loss. While similar chromosome inversion–fusion patterns were observed in all Kwoniella species with fewer than 14 chromosomes, no such pattern was detected in Cryptococcus. Instead, Cryptococcus species with less than 14 chromosomes showed reductions primarily through rearrangements associated with the loss of repeat-rich centromeres. Additionally, Cryptococcus genomes exhibited frequent interchromosomal translocations, including intercentromeric recombination facilitated by transposons shared between centromeres. Overall, our findings advance our understanding of genetic changes possibly associated with pathogenicity in Cryptococcus and provide a foundation to elucidate mechanisms of centromere loss and chromosome fusion driving distinct karyotypes in closely related fungal species, including prominent global human pathogens.
Chapter
Skin diseases have been treated with conventional topical and oral formulations to date. Both topical and oral routes encounter several challenges, e.g., malabsorption, high metabolism, toxicity, and drug-drug interactions. Traditional topical dosage forms, such as lotions, creams, and gels, have been unsuccessful to demonstrate controlled release of drug. Fungal infections impart an immeasurable impact on human health as well as on the global economy. Even then, it remains an ignored issue by public health authorities, although most of the deaths due to fungal infections are preventable. Worldwide, there are estimated to be huge rate of various fungal infection cases, i.e., aspergillosis (2.5 lacs), pulmonary aspergillosis (>30 lacs), candidiasis (7 lacs), Pneumocystis jirovecii pneumonia, (5 lacs), and more each year. Nanomedicines emerged as a suitable therapeutic option for treating fungal infections and hence decreasing the number of patients infected with these fungal infections. Nano-sized drugs can easily enter the skin micropores to achieve the desired effect. This chapter covers various drug delivery systems for antifungals, i.e., lipid-based drug carriers, liposomes, polymeric nanoparticles, carbon nanotubes, dendrimers, quantum dots, etc. All of these delivery systems have advantages as well as disadvantages. The main advantages are improved antifungal properties, e.g., bioavailability, reduced toxicity, and site-specific drug delivery, which expedite innovative therapeutic systems, while the main disadvantage is high production costs of these nanocarrier drug delivery systems. Thus, more research is needed to develop novel therapeutic alternatives that contribute to the advancement of the field.
Article
Full-text available
Formation of hyphae is a key virulence trait of the fungal pathogen Candida albicans. Hypha morphogenesis depends upon the cyclin Hgc1, which acts together with cyclin-dependent protein kinase Cdc28 to phosphorylate effectors that drive polarized growth. Hgc1 has also been implicated in gene regulation through its effects on two transcription factors, Efg1 and Ume6. Here we report RNA-seq analysis of two pairs of hgc1Δ/Δ mutants and their respective wild-type strains, which lie in two different genetic backgrounds. We find that hgc1Δ/Δ mutations alter expression of 271 genes in both genetic backgrounds, and 266 of those genes respond consistently with regard to up- or down-regulation. Consistency is similar to what has been observed with efg1Δ/Δ mutations and greater than observed with nrg1Δ/Δ mutations in these two backgrounds. The gene expression response includes genes under Efg1 control, as expected from prior studies. Hgc1-responsive genes also include ergosterol biosynthetic genes and bud neck-related genes, which may reflect interactions between Hgc1 and additional transcription factors as well as effects of Hgc1 on cellular length-to-width ratios.
Article
Fungal infections cause severe and life-threatening complications especially in immunocompromised individuals. Antifungals targeting cellular machinery and cell membranes including azoles are used in clinical practice to manage topical to systemic fungal infections. However, continuous exposure to clinically used antifungal agents in managing the fungal infections results in the development of multi-drug resistance via adapting different kinds of intrinsic and extrinsic mechanisms. The unique chemical composition of fungal membranes presents attractive targets for antifungal drug discovery as it is difficult for fungal cells to modify the membrane targets for emergence of drug resistance. Here, we discussed available antifungal drugs with their detailed mechanism of action and described different antifungal resistance mechanisms. We further emphasized structure-activity relationship studies of membrane-targeting antifungal agents, and classified membrane-targeting antifungal agents on the basis of their core scaffold with detailed pharmacological properties. This review aims to pique the interest of potential researchers who could explore this interesting and intricate fungal realm.
Article
Samples of ambient air collected with three different types of spore traps in a rural location were examined for the presence of Pneumocystis carinii by screening for P. carinii-specific DNA sequences by DNA amplification. Eleven spore trap samples were analyzed by nested PCR, using oligonucleotide primers designed for the gene encoding the mitochondrial large subunit rRNA of P. carinii f. sp. carinii and P. carinii f. sp. hominis. The samples were collected over a 3-year period during the months of May to September, with a range of sampling times from 9 to 240 h. One air sample from an animal facility housing P. carinii-infected rats was also examined. P. carinii-specific amplification products were obtained from samples from each of the spore traps. The amplification products from eight air samples were cloned and sequenced. The majority of the recombinants from each of these samples had sequences identical to those of P. carinii f. sp. carinii and P. carinii f. sp. hominis, and a number of clones had single-base differences. These data suggest that sequences identical to those of P. carinii f. sp. carinii and P. carinii f. sp. hominis can be detected in samples of air collected in a rural location and that P. carinii may be a component of the air spora of rural Oxfordshire.
Article
The functional properties of granulocytes in a diabetic patient deficient in myeloperoxidase (MPO) were compared with those of granulocytes in healthy subjects. The granulocytes of this patient had normal phagocytic activity. The microbicidal activity of the granulocytes was partially diminished with regard to Staphylococcus aureus and was almost nil with regard to Candida albicans. Fungicidal activity of normal granulocytes was shown to be impaired during the in vitro artificial hyperglycemic condition. The relationship among diabetes mellitus, MPO deficiency, and serious C. albicans infection was examined. Genetic investigation was carried out in 28 members of the proband's family. In close relatives of the patient, MPO values were found to be diminshed to a greater or lesser degree, thus suggesting variable expressivity of the heterozygote state of MPO deficiency.
Article
The fungal pathogen Cryptococcus neoformans has a predilection for the central nervous system (CNS), resulting in devastating meningoencephalitis. At present, it is unclear how C. neoformans traverses the blood-brain barrier (BBB) and causes CNS infection. The present study has examined and characterized the interaction of C. neoformans with human brain microvascular endothelial cells (HBMEC), which constitute the BBB. Adhesion of and transcytosis of HBMEC by C. neoformans was inoculum- and time-dependent and occurred with both encapsulated and acapsulated strains. C. neoformans induced marked morphological changes in HBMEC, for example membrane ruffling, irregular nuclear morphology and swelling of the mitochondria and the ER. These findings suggest that C. neoformans induced actin cytoskeletal reorganization of the host cells. In addition, it was observed that the dephosphorylated form of cofilin was increased during cryptococcal adherence to HBMEC, concomitant with the actin rearrangement. Cryptococcal binding to HBMEC was increased in the presence of Y27632, a Rho kinase (ROCK)-specific inhibitor. Since ROCK activates LIM kinase (LIMK), which phosphorylates cofilin (inactive form), this suggests the involvement of the ROCK→LIMK→cofilin pathway. In contrast, the phosphatase inhibitor sodium orthovanadate decreased adherence of Cryptococcus to HBMEC, concomitant with the increase of phosphorylation of cofilin. Furthermore, the tight junction marker protein occludin became Triton-extractable, indicating alteration of tight junctions in brain endothelial cells. This is the first demonstration that C. neoformans is able to adhere to and transcytose across the HBMEC monolayer and alter the cytoskeleton morphology in HBMEC. Further characterization of the interactions between C. neoformans and HBMEC should help the development of novel strategies to prevent cryptococcal meningitis and its associated morbidity.
Article
The topic of immunity to fungal infections is of interest to a wide range of disciplines, from microbiology to immunology. It is of particular interest in terms of therapy of HIV-infected individuals, and patients with cancer or individuals who have received transplants. Understanding the nature and function of the immune response to fungi is an exciting challenge that might set the stage for new approaches to the treatment of fungal diseases, from immunotherapy to vaccines. The past decade has witnessed the development of a wide range of new approaches to elucidate events that occur at the host-fungus interface.