ArticlePDF AvailableLiterature Review

Resistance of Candida spp. to antifungal drugs in the ICU: Where are we now?

Authors:

Abstract and Figures

Current increases in antifungal drug resistance in Candida spp. and clinical treatment failures are of concern, as invasive candidiasis is a significant cause of mortality in intensive care units (ICUs). This trend reflects the large and expanding use of newer broad-spectrum antifungal agents, such as triazoles and echinocandins. In this review, we firstly present an overview of the mechanisms of action of the drugs and of resistance in pathogenic yeasts, subsequently focusing on recent changes in the epidemiology of antifungal resistance in ICU. Then, we emphasize the clinical impacts of these current trends. The emergence of clinical treatment failures due to resistant isolates is described. We also consider the clinical usefulness of recent advances in the interpretation of antifungal susceptibility testing and in molecular detection of the mutations underlying acquired resistance. We pay particular attention to practical issues relating to ICU patient management, taking into account the growing threat of antifungal drug resistance.
Content may be subject to copyright.
Danie
`
le Maubon
Ce
´
cile Garnaud
Thierry Calandra
Dominique Sanglard
Muriel Cornet
Resistance of
Candida
spp. to antifungal drugs
in the ICU: where are we now?
Received: 15 May 2014
Accepted: 10 July 2014
! Springer-Verlag Berlin Heidelberg and
ESICM 2014
Take-home message: The emergence of
resistance is a warning signal triggering
improvements in antifungal drug use,
particularly in patients for whom the
potential benefit of treatment is unproven.
Practical proposals to detect and prevent the
risk of clinical failure are (i) accurate
assessments of prior antifungal exposure,
(ii) close clinical monitoring of patients
treated with antifungal drugs, (iii) routine
surveillance of in vitro susceptibility testing
and (iv) development of feasible methods
for rapid detection of mutations.
Electronic supplementary material The
online version of this article (doi:
10.1007/s00134-014-3404-7) contains sup-
plementary material, which is available to
authorized users.
D. Maubon (
)
) ! C. Garnaud ! M. Cornet
Parasitologie-Mycologie, Institut de
Biologie et de Pathologie, CHU de
Grenoble, Grenoble, France
e-mail: dmaubon@chu-grenoble.fr
Tel.: ?33 4 76 76 54 90
D. Maubon ! C. Garnaud ! M. Cornet
Laboratoire TIMC-TheREx, UMR 5525
CNRS-UJF, Universite
´
Grenoble Alpes,
Grenoble, France
T. Calandra
Infectious Diseases Service, Department of
Medicine, Centre Hospitalier Universitaire
Vaudois and University of Lausanne,
Lausanne, Switzerland
D. Sanglard
Institute of Microbiology, Centre
Hospitalier Universitaire Vaudois and
University of Lausanne, Lausanne,
Switzerland
Abstract Current increases in anti-
fungal drug resistance in Candida
spp. and clinical treatment failures are
of concern, as invasive candidiasis is
a significant cause of mortality in
intensive care units (ICUs). This
trend reflects the large and expanding
use of newer broad-spectrum anti-
fungal agents, such as triazoles and
echinocandins. In this review, we
firstly present an overview of the
mechanisms of action of the drugs
and of resistance in pathogenic
yeasts, subsequently focusing on
recent changes in the epidemiology of
antifungal resistance in ICU. Then,
we emphasize the clinical impacts of
these current trends. The emergence
of clinical treatment failures due to
resistant isolates is described. We
also consider the clinical usefulness
of recent advances in the interpreta-
tion of antifungal susceptibility
testing and in molecular detection of
the mutations underlying acquired
resistance. We pay particular atten-
tion to practical issues relating to ICU
patient management, taking into
account the growing threat of anti-
fungal drug resistance.
Keywords Antifungal resistance !
Resistance mechanisms !
Candida resistance !
Intensive care unit !
Clinical resistance !
Microbial resistance
Introduction
Invasive candidiasis is a major threat to intensive care
unit (ICU) patients, causing significant mortality. An
early initiation of antifungal therapy is crucial to improve
the prognosis [1, 2]. However, the performance of current
diagnostic tools for confirming the diagnosis remains
limited. Increasing numbers of ICU patients without
documented candidiasis are therefore receiving prophy-
lactic or empirical antifungal treatments in an attempt to
decrease Candida-related mortality. This strategy has
been encouraged by the introduction of new, better-tol-
erated antifungal drugs, such as triazoles and
echinocandins, leading to stronger selective pressure [3,
4]. Antifungal drug resistance was considered less prob-
lematic in Candida spp. than in other pathogens, but
Intensive Care Med
DOI 10.1007/s00134-014-3404-7
REVIEW
recent increases in resistance to both echinocandins and
azoles have led to clinical failures [5, 6]. This is a matter
of concern because of the limited number of drug classes
targeting different fungal components and because the
number of patients at risk receiving treatment is contin-
ually growing, thus further increasing antifungal drug
pressure.
In this review, we firstly summarize the basis of the
mechanisms of action and resistance concentrating on
recent advances that improve our understanding of
antifungal drugs. Then, we describe the current changes
in the epidemiology of Candida spp. resistance. We
enlighten their consequences for responses to antifungal
treatments and for the optimal choice for empiric, pre-
emptive and targeted strategies in ICU patients. The
clinical relevance of the new developments for labora-
tory antifungal drug testing and for the detection of
resistance-associated mutations is discussed with spe-
cific attention paid to practical approaches, to assess the
risk of clinical treatment failure and to improve its
prevention.
Targets and mechanisms of action of systemic
antifungal drugs
Fungi are more closely related to humans than other
pathogens, such as bacteria, limiting the number of
available antifungal targets. Despite the introduction of a
novel drug class exploiting a new target (echinocandins)
and new azole drugs with a broader spectrum of activity
(voriconazole, posaconazole), the antifungal arsenal still
remains restricted.
Antifungal agents acting on the cell wall and/
or plasma membrane
Echinocandins
Caspofungin, micafungin and anidulafungin block cell
wall synthesis by inhibiting (1,3)-b-
D-glucan synthase,
which catalyses the first step in the elongation of (1,3)-b-
D-glucans, a major cell wall component together with
chitin and mannoproteins. Echinocandins inhibit the cat-
alytic subunit (Fksp) encoded by two or three FKS genes,
depending on the fungal species (Fig. 1)[7].
Azoles
Triazoles—fluconazole, itraconazole, voriconazole and
posaconazole—are the azoles most commonly used to
treat invasive fungal infections. Isavuconazole (ISA;
BAL4815), is a novel triazole currently in global phase 3
clinical trials for treatment of invasive fungal infections.
It showed good activity against Candida spp. with
reduced susceptibility to currently used azoles (personal
communications: Smart JI, P983, ECCMID, Berlin, 2013
and Maertens J., O230, ECCMID, Barcelona, 2014).
Triazoles block the synthesis of the main sterol of fungal
membranes, ergosterol, by targeting the lanosterol-14a-
demethylase, also called Erg11p or Cyp51p (Fig. 1). This
blockade has three major effects: (a) ergosterol depletion
and changes in membrane permeability, (b) changes in the
activity of membrane-bound proteins, some of which are
involved in cell wall synthesis and (c) synthesis of toxic
sterols as a result of Erg3p activity and accumulation of
14a-methylated sterols (Fig. 1)[8]. Azoles have long
been considered to act solely on the cell membrane, but
there is growing evidence to suggest that they also act on
the cell wall structure. Studies have demonstrated com-
pensatory responses similar to those observed with cell
wall-disrupting agents [9, 10].
Polyenes
Amphotericin B (AMB) and its lipid and liposomal
derivatives bind ergosterol, causing pore formation and
ion leakage, with fungicidal effects (Fig. 1). It has been
suggested that pore formation is not required for the
fungicidal effect, which is dependent only on ergosterol
binding [11]. In addition, a recent study revealed that
AMB is able to aggregate and to act like a ‘sponge’’, thus
extracting this key component from cell membranes [12].
Cholesterol is the major sterol of the mammalian mem-
branes. Ergosterol and cholesterol have different
structures, but drug specificity is not absolute and AMB
has also been shown to bind cholesterol [13]. New for-
mulations involving liposome encapsulation (L-AMB),
AMB colloidal dispersion (ABCD) and AMB lipid com-
plex (ABLC) have increased drug specificity and
delivery, greatly reducing toxicity without decreasing
efficacy [14]. However, ABCD caused a similar number
of infusion-related reactions to AMB, and is no longer
available.
Antifungal agents acting on nucleic acids and protein
synthesis
Flucytosine is a pyrimidine analogue that is converted
to 5-fluorouracil, which inhibits both RNA and DNA
synthesis. Cytosine permease (Fcy2p), cytosine deami-
nase (Fcy1p), and uracil phosphoribosyl transferase
(Fur1p) activities are required for antifungal activity
(Fig. 1).
Antifungal drug resistance in Candida spp.
Tolerance and resistance due to cellular stress
responses
An increase in cell wall chitin content has been shown to
occur in response to the exposure to echinocandin and
azoles in C. albicans [10, 15]. The blockade of this cell
wall compensatory mechanism with calcineurin or protein
kinase C (PKC) inhibitors restores the fungicidal activity
of both azoles and echinocandins consistent with the
hypothesis that chitin accumulation plays a role in toler-
ance to these drugs [16, 17]. Furthermore, a high chitin
content has been associated with resistance to echino-
candins in mice and in ‘paradoxical growth’, defined as
the ability to develop in vitro at high, but not intermediate
concentrations of a drug [18]. The clinical impact of this
paradoxical growth in vitro remains unclear, as it is also
related to lower virulence [19].
Molecular mechanisms of antifungal drug resistance
in Candida spp.
The molecular mechanisms of antifungal drug resistance
are presented in Fig. 2 and Table S1 (electronic supple-
mentary material).
Echinocandins
Molecular resistance to echinocandins is mediated
principally by mutations in FKS genes: FKS1 in Can-
dida spp., and FKS1 and FKS2 in C. glabrata. These
mutations are located in two‘hotspot’regions,HS1
and HS2 and are mostly S645F/P/Y and S629P in FKS1
of C. albicans and C. glabrata respectively, and S663F/P
in C. glabrata FKS2 [7, 20]. These mutations confer
cross-resistance to all three echinocandins, by modify-
ing the catalytic and kinetic properties of the target
enzyme.
Azoles
Decreased susceptibility or resistance to azoles in
Candida spp. is mediated by various mechanisms,
which may operate simultaneously in a given isolate,
following sequential acquisition under drug pressure
[21, 22]. Drug efflux is a major mechanism, mediated
by mutations of genes encoding regulators of trans-
porters of the ATP-binding cassette (ABC) superfamily
or the major facilitator superfamily (MFS) [22, 23].
ABC transporter overexpression is associated with
cross-resistance to diverse azoles, whereas MFS trans-
porter overexpression is limited to resistance to fewer
Fig. 1 Targets and mechanisms of action of systemic antifungal
drugs. Sites and modes of action of the current classes of systemic
antifungal drugs used to treat invasive candidiasis. a Echinocandins
target cell wall synthesis, inhibiting (1,3)-b-
D-glucan synthesis,
which occurs on the inner side of the plasma membrane. b Azoles
target the ergosterol biosynthesis pathway in the endoplasmic
reticulum. They block 14a-demethylase (also called Erg11p or
Cyp51p), resulting in ergosterol depletion in the membrane and
activation of the Erg3p alternative pathway, leading to the synthesis
of toxic sterols. c Polyenes bind to cell membrane ergosterol
creating pores and aggregate, to act as a ‘sponge’’, thus resulting in
ion depletion. d Flucytosine acts in the nucleus, where its toxic
metabolites inhibit nucleic acid synthesis
azoles (fluconazole, voriconazole) (Table S1). A second
major mechanism is overproduction of the target
enzyme Erg11p [22]. Amino acid substitutions in
Erg11p may also decrease the affinity of the drugs for
this enzyme [24]. Finally, ERG3 mutations are associ-
ated with cross-resistance to azoles through a metabolic
bypass leading to the synthesis of fecosterol which is
able to replace ergosterol (Fig. 2)[25].
Other major genetic alterations may decrease azole
susceptibility. Aneuploidy, through chromosomal dupli-
cation or loss of heterozygosity, increases the copy
number of genes involved in azole resistance in C. albi-
cans and C. glabrata [22, 23, 26]. Respiratory and
mitochondrial deficiencies may also contribute to azole
resistance in these species [27].
Polyenes
Polyene resistance has been little described and the exact
mechanisms involved remain unclear, partly because of
the small number of clinical isolates displaying altered
susceptibility in vitro. Resistance is associated with
changes in membrane sterol composition due to mutations
in the genes of the ergosterol biosynthesis pathway:
ERG2, ERG3, ERG5, ERG6 and ERG11 [28].
Flucytosine
Two main mechanisms of flucytosine resistance are
known: (1) decreased uptake of the drug due to mutations
Fig. 2 Molecular mechanisms of echinocandin and azole resistance
in Candida spp. a Regular b-1,3-glucan synthesis on the inner side
of the fungal membrane. b Typical echinocandin activity. These
compounds block cell wall synthesis by inhibiting the Fksp subunit
of the b-1,3-glucan synthase. c Echinocandin resistance due to FKS
mutations. The target enzyme is less sensitive to echinocandins,
allowing the production of b-1,3-glucans. d Typical ergosterol
synthesis at the endosplamic reticulum and uptake of azole
antifungal drugs into the cytosol of the fungal cell. e Typical azole
activity. These molecules inhibit the lanosterol-14a-demethylase
(Erg11p), leading to (1) membrane ergosterol depletion and (2) the
production of toxic sterols via Erg3p. f Azole resistance due to (1)
the overproduction of transporters, increasing azole efflux, (2)
alteration of the target enzyme by mutations of ERG11, (3) Erg11p
overproduction, (4) mutations of ERG3 preventing the azole-
mediated production of toxic sterols which are substituted by the
non-toxic fecosterol
of the FCY2 gene encoding the cytosine permease, and (2)
impaired metabolism of the drug or its active metabolite
(5-FU) due to mutations of FCY1 or FUR1. Such muta-
tions have been described in clinical isolates of C.
albicans and C. lusitaniae [29, 30].
Antifungal drug resistance in Candida spp. biofilms
In ICUs, candidiasis may be favoured by biofilms for-
mation, mostly on catheters but also on other implanted
medical devices. Only a few antifungal drugs (L-AMB
and echinocandins) have some efficacy against yeasts
embedded in such complex structures [31]. Echinocandins
are active against biofilms, but are more effective against
biofilms containing C. albicans or C. glabrata than
against biofilms of C. tropicalis or C. parapsilosis [32,
33]. Conversely, yeast cells in biofilms are up to 1,000
times more resistant to azoles than their planktonic
counterparts [34].
The resistance of biofilms combines both planktonic
and biofilm-specific resistance mechanisms. Efflux pump
upregulation is involved in the early stages of biofilm
development, whereas the greater resistance of mature
biofilms is due to the presence of an extracellular matrix
(ECM) and persister cells, changes to the sterol compo-
sition of the membrane and the activation of stress-
induced pathways [3537]. The ECM plays a key role, by
sequestering antifungal agents and preventing their
interaction with the target. This action is mediated at least
by (1,3)-b-
D-glucan polymers [38]. Stress responses,
mediated by the PKC, calcineurin and heat shock pro-
tein 90 (HSP90) pathways, also control ECM production
[35, 39]. Extracellular DNA also affects biofilm resistance
and the treatment of C. albicans biofilms with DNase
potentiates the antifungal activity of echinocandins and
polyenes, but not fluconazole [40]. Persister cells have
been described in Candida spp. biofilms, particularly
those formed by C. krusei and C. albicans. These cells are
phenotypic variants able to survive in the presence of
antifungal agents. They can again proliferate when drug
pressure is released and may cause relapses often
described in clinical situations [35, 37].
Laboratory detection of antifungal resistance
Antifungal drug susceptibility testing assays
Methods for in vitro susceptibility testing are available
from the Clinical and Laboratory Standards Institute
(CLSI) and the European Committee on Antimicrobial
Susceptibility Testing (EUCAST) [41, 42]. Other com-
mercially available standardised tests, such as Etest
"
(bioMe
´
rieux), Sensititre YeastOne
"
(TREK diagnostic
systems), ATB fungus 2
"
(bioMe
´
rieux) and Vitek-2
"
(bioMe
´
rieux), are more appropriate for routine clinical
use [4345]. These tests determine the minimum inhibi-
tory concentration (MIC) or directly classify isolates as
susceptible (S), intermediate (I) or resistant (R), corre-
sponding to a high probability of treatment success (S), an
uncertain effect of treatment (I) or a high probability of
treatment failure (R). This classification is based on the
clinical breakpoints (CBPs) established for MIC inter-
pretation. Previous CBPs were not species-specific and
were too high to distinguish between C. glabrata isolates
susceptible and resistant to azoles and to detect emerging
resistance in C. albicans, C. tropicalis and C. parapsilo-
sis. In addition, clinical resistance to echinocandins due to
FKS mutations was increasingly being reported in patients
infected with ‘S’ strains, defined with a former CBP
of 2 lg/ml or less. Up to 45 % of FKS mutants have been
incorrectly considered as susceptible [7, 46, 47].
The revised CBPs in current use are species-specific
and were established on the basis of five parameters: dose
regimens; MIC distributions from multiple laboratories;
epidemiologic cut-off values defined with respect to the
higher MIC of wild-type isolates; pharmacokinetic/phar-
macodynamic parameters and clinical outcome [20]. The
key changes concern the susceptibility of C. glabrata and
C. parapsilosis to fluconazole and echinocandins,
respectively. The ‘S’ category was abolished by CLSI
and EUCAST for C. glabrata and fluconazole, consider-
ing all isolates to be intermediate or resistant. EUCAST
was even more severe in its approach, recommending, as
for C. krusei, that C. glabrata should not be tested with
fluconazole and that fluconazole should not be used for
C. glabrata infections [48]. The same removal of the ‘S’
category was recommended, albeit only by EUCAST, for
C. parapsilosis and echinocandins [20]. One other major
difference between CLSI and EUCAST is that this latter
does not determine CBPs for caspofungin. Indeed, CLSI
and EUCAST agree that there is a lack of interlaboratory
reproducibility in MIC values for caspofungin. Until this
problem, not seen with other echinocandins, is resolved,
neither CLSI nor EUCAST recommends caspofungin
resistance testing [4952]. EUCAST specifies that some
mutations decrease susceptibility to anidulafungin and
caspofungin but not micafungin, and thus recommends
the use of anidulafungin as a marker for echinocandin
resistance [20] (see Table 1 for simplified CLSI and
EUCAST updated CBPs). Both CLSI and EUCAST also
determined epidemiological cut-off values which are
more sensitive than CBP to detect non-wild-type isolates
exhibiting potential resistance mutations and
mechanisms.
AMB testing remains particularly challenging and
microbiological resistance is rarely detected [53]. Etest
"
(bioMe
´
rieux) was found to be superior to both CLSI and
EUCAST reference methods for identifying resistant and
intermediately susceptible C. glabrata isolates [54]. Thus,
AMB resistance is mainly identified through clinical
failure.
While current MIC testing protocols are adapted for
planktonic cells, these protocols are still not implemented
in biofilms. Since biofilms can be detected in infected
tissues, this is clearly another limitation in the interpre-
tation of susceptibility tests for predicting patient
outcome.
The performance of direct antifungal drug suscepti-
bility testing, through the use of Etest
"
(bioMe
´
rieux) on
positive blood samples, has been evaluated. Agreement
between direct and standard methods was high and false-
positive results for resistance to fluconazole and vorico-
nazole were obtained for 7 % of isolates, with false-
negative results obtained for 0.6 % of blood samples. No
errors were detected for caspofungin, but the method was
not reliable for AMB [55]. The new CBPs are species-
specific, so this approach requires a rapid identification
tool. Direct antifungal drug susceptibility testing should
therefore be re-evaluated, according to the current stan-
dards, for both categorisation and identification.
Molecular detection of mutations conferring
antifungal drug resistance
Molecular methods have been developed for the charac-
terisation of resistance-causing mutations. Culture-based
susceptibility assays take at least 24 h, but molecular
tools can assess resistance more rapidly and with greater
sensitivity. Both azole and echinocandin resistance
mutations are accurately detected with next-generation
sequencing platforms, allele-specific real-time probes,
melt curve analysis or microarrays, or microsphere-based
technologies, such as Luminex Mag Pix (Austin, TX)
[5660] (C. Garnaud, personal communication). More-
over, as in Aspergillus fumigatus azole resistance, the
direct detection of mutations in clinical samples may
make it possible to detect mutations earlier by eliminating
the time required for culture [61].
Update on the epidemiology of Candida spp.
antifungal resistance
When focusing on species distribution and antifungal
resistance, recent epidemiological studies, including the
SENTRY cohort, did not show major differences between
ICU and non-ICU patients [62, 63]. In both ICU and non-
ICU, the five main species (i.e. C. albicans, C. parapsi-
losis, C. glabrata, C. tropicalis and C. krusei) are
responsible for more than 90 % of invasive fungal
infections [6265]. C. albicans still stands in first place,
even if, since the early 2000s, a shift towards non-albi-
cans species was clearly noticed [66, 67]. The fluconazole
drug pressure may explain this trend but other factors,
mainly underlying conditions or antibacterial therapy,
have been suggested [68]. A main difference in species
distribution is related to geographical location. In
southern countries (Italy, Spain, South America)
C. parapsilosis ranks second while in northern countries
C. glabrata is the most frequent species after C. albicans
[69]. These site specificities highlight the importance of
local data on Candida epidemiology, specific to each
health-care centre.
Table 1 EUCAST and CLSI antifungal breakpoints for the main Candida species
Antifungal agent MIC breakpoint (mg/l)
C. albicans C. glabrata C. krusei C. parapsilosis C. tropicalis
BS [R BS [R BS [R BS [R BS [R
Amphotericin B
EUCAST 1 1 1 1 1 1 1 1 1 1
CLSI ND ND ND ND ND ND ND ND ND ND
Fluconazole
EUCAST 2 4 0.002 32 2 4 2 4
CLSI 2 4 0.002 32 2 4 2 4
Voriconazole
EUCAST 0.12 0.12 IE IE IE IE 0.12 0.12 0.12 0.12
CLSI 0.12 0.5 0.5 1 0.12 0.5 0.12 0.5
Anidulafungin
EUCAST 0.03 0.03 0.06 0.06 0.06 0.06 0.002 4 0.06 0.06
CLSI 0.25 0.5 0.12 0.25 0.25 0.5 2 4 0.25 0.5
Micafungin
EUCAST 0.016 0.016 0.03 0.03 IE IE 0.002 2 IE IE
CLSI 0.25 0.5 0.06 0.12 0.25 0.5 2 4 0.25 0.5
Adapted from Arendrup et al. [20] drug resistance updates (doi:10.1016/j.drup.2014.01.001) with permission. For complete data, see
Arendrup et al. [20]
ND not done, IE insufficient evidence
Indeed, some non-albicans species show intrinsic
resistance. For example C. glabrata and C. krusei are less
susceptible to azoles than other species (Table 2;[7072])
and C. parapsilosis is less susceptible to echinocandins
owing to naturally occurring polymorphisms of the FKS
genes [73, 74]. Breakthrough infections with these species
may therefore occur during azoles or echinocandin
treatment [75, 76].
Another major and increasing threat is the risk of
becoming infected with a strain which has acquired a
resistant phenotype. Acquired resistance is thought to be
rare in Candida spp., or at least less frequent than intrinsic
resistance. Fortunately, yeasts, unlike bacteria, do not
display the horizontal transmission of resistance genes
[77]. Moreover, cross-contamination between patients
and health-care workers has been described mostly for C.
albicans and C. parapsilosis but remains rare [69, 78, 79].
Acquired resistance thus results principally from the
selection of mutants subjected to drug pressure in
patients.
Acquired resistance to echinocandins is increasingly
reported for most of the clinically important Candida
spp. It remains uncommon in C. albicans (\1 %), C.
tropicalis (\5 %) and C. krusei (\7 %), but is now
becoming frequent in C. glabrata (8–15 %) [5, 63, 80,
81]. One recent study showed that the frequency of
echinocandin resistance in C. glabrata increased from
4.9 to 12.3 % between 2001 and 2010 [5]. It has been
shown that 7 days of exposure to echinocandin is suf-
ficient to induce FKS mutations in C. glabrata [5, 6].
The haploid trait of this species may partly explain the
higher level of expression of molecular resistance
exhibited by C. glabrata. FKS mutations have been
described in almost all the clinically important Candida
species:
C. albicans, C. glabrata, C. tropicalis [82, 83],
C. krusei [84] and C. kefyr [85] and breakthrough
infections are also increasingly reported [7, 75, 76, 86
90]. A recent study focusing on C. glabrata candidemia
described 18 % of FKS mutation, with prior echino-
candin exposure as the only independent risk factor for
the development of these mutations [91] confirming the
results obtained previously by Alexander et al. [5].
Interestingly, the nature and/or the number of FKS
mutations in C. glabrata and C. albicans influences the
level of resistance in vivo [91, 92]. Even if the micro-
biological resistance to echinocandins is still uncommon,
the growing incidence of FKS mutations is worrying and
needs to be very closely monitored. FKS resistance
mutations also need to be more deeply studied.
Azoles and especially fluconazole are widely pre-
scribed for ICU patients. Acquired fluconazole resistance
is frequent in C. glabrata (from 4 to 16 %), which
increasingly displays cross-resistance to voriconazole. So
far, multidrug-resistant phenotype against azole and ech-
inocandins has only been described for C. glabrata and is
a matter of serious concern [5, 63, 66, 80, 81, 93, 94].
Fluconazole resistance remains uncommon in C. albicans
(\5 %), but is more prevalent in C. parapsilosis (4–10 %)
and C. tropicalis (4–9 %) [63, 64, 81]. However, the
recent China-SCAN study reported higher rates of fluco-
nazole resistance in C. albicans (9.6 %) and C.
parapsilosis (19.3 %) which may reflect geographical
differences [93]. Again, most studies report that a previ-
ous history of azole pre-exposure increases the risk of
in vitro azole resistance (from 2 to 58 % in a 2013 study
by Montagna et al. [65]).
Resistance to AMB remains rare despite its use in
monotherapy for years. This may be due to its inherently
fungicidal effect, limiting the selection of mutants.
However, resistant isolates of C. glabrata and C. krusei
Table 2 Spectrum of activity of the antifungal agents used to treat invasive candidiasis
Candida spp. Polyenes Azoles Echinocandins Flucytosine
AMB formulations FLU ITRA VOR POSA CAS MIC ANI
C. albicans ?? ?? ?? ?? ?? ?? ?? ?? ??
C. glabrata ?
a
?/-?/- ? ? ?? ?? ?? ??
C. parapsilosis ?? ?? ?? ?? ?? ? ? ? ??
C. tropicalis ?? ?? ?? ?? ?? ?? ?? ?? ??
C. krusei ? 2 ?/- ? ? ?? ?? ?? 2
C. rugosa ?
a
? ? ?? ?? ? ? ? ??
C. guilliermondii ?? ?? ?? ?? ?? ? ? ? ??
C. lusitaniae ?? ?? ?? ?? ?? ?? ?? ?? ??
C. inconspicua ?? - ?/- ? ? ?? ?? ?? NS
C. norvegensis ?? - ?/-?/-?/- ?? ?? ?? NS
Adapted with permission from Denning DW, Hope WW (2010)
Trends Microbiol (doi:10.1016/j.tim.2010.02.004). In vitro inherent
activity: ?? good activity, ? mild activity, ?/- slight activity,
- no activity; NS not specified
AMB amphotericin B, FLU fluconazole, ITRA itraconazole, VOR
voriconazole, POSA posaconazole, CAS caspofungin, MIC mica-
fungin, ANI anidulafungin
a
This slight decrease in susceptibility to AMB is more pronounced
in North America than in Europe
are increasingly being reported and this new entity also
needs to be closely monitored [54, 9598].
Even if it remains uncommon, Candida spp. drug
resistance is clearly becoming an ‘every-day’ concern in
the mycology laboratory. Determining initial but also
subsequent MICs is necessary to assess microbial resis-
tance emergence.
Clinical impact of antifungal resistance
Clinical resistance
The failure of antifungal therapy or clinical resistance is
defined as a steady-to-worse infectious syndrome with no
improvement of attributable symptoms during the evalu-
ation, death being the ‘ultimate’ failure. However, it
remains difficult to assess whether the patient dies with or
of fungal infection. These criteria classify clinical out-
come in trials, but can also be applied for bedside
management [99]. Breakthrough infections are considered
as clinical resistance and are microbiologically docu-
mented. They have frequently, but not exclusively, been
described in cases of echinocandin exposure [5, 7, 75, 76,
8690, 100, 101].
Despite its recent spread, microbiological resistance is
not the major factor underlying clinical resistance.
Indeed, underlying diseases, immunosuppression, com-
plicated abdominal surgery, extreme age and renal failure,
all frequently encountered in ICU patients, are known to
be predictors of mortality in cases of invasive candidiasis
[102, 103]. Clinical failure may also occur when the
effective concentration of the chosen drug is not reached
at the infected site. This situation frequently occurs for
biofilms on prosthetic devices or catheters, abscesses,
chorioretinitis or endophthalmitis, or other sanctuary foci.
As a result of their multiple comorbidities and manage-
ment strategies, ICU patients may display higher
pharmacokinetic (PK) instability than other patients.
Thus, regular, complete investigations of deep infections
and assessments of the PK/pharmacodynamic properties
of antifungal drugs are essential for correct appraisal of
the clinical response. For example, given their poor
penetration into the eye, echinocandins are not recom-
mended in cases of suspected ocular secondary
dissemination of Candida, whereas echinocandins or
L-AMB are the drugs of choice when central catheters
cannot be removed [104]. Given the lower frequency of
primary resistance than initially thought and its beneficial
penetration properties, 5-fluorocytosine may be adminis-
tered in combination with other drugs to treat invasive
candidiasis at deep secondary sites [105]. A recent review
has provided a comprehensive analysis of the tissue
penetration properties of current antifungal agents [106].
Clinical relevance of in vitro antifungal drug
susceptibility testing and of the molecular detection
of mutations
Crude MICs are not sufficient to predict clinical outcome.
Candidemia due to C. tropicalis, for which the MICs of all
antifungal agents are very low, has been associated with a
higher mortality than for other species [102]. By contrast,
C. parapsilosis isolates have high MICs for echinocandins,
although treatment failure remains rare [104, 107]. These
discrepancies between laboratory tests on antifungal drugs
and clinical outcome have been extensively reported and
are due to several factors, including a species-dependent
virulence traits and patient-dependent conditions. The use
of the revised CBPs, partly taking these factors into
account, may improve the clinical predictive value of
in vitro susceptibility tests. Thus, close monitoring of MICs
together with accurate interpretation based on revised
CBPs is always warranted to ensure appropriate specific
treatment (Fig. 3). Antifungal susceptibility testing (AST)
on Candida strains isolated from deep sites is recom-
mended by the European Society of Clinical Microbiology
and Infectious Diseases (ESCMID) [104]. Reference
methods are preferred but commercial techniques can be
used if verification has been made that the endpoint for each
species mirrors those of reference methods [104].
There is growing evidence that the detection of muta-
tions, and especially FKS mutations, could be used as a
predictive marker of clinical failure. In one recent study,
FKS mutations were found in 7.9 % of 313 C. glabrata
isolates from blood samples, and up to 80 % of patients
infected with strains with both FKS mutations and high
MICs for caspofungin experienced clinical failure or
recurrent infection [5]. Another study identified C. glab-
rata FKS mutation as the only independent risk factor
associated with clinical failure and showed that the
detection of FKS mutations was superior to MIC for pre-
dicting treatment response [108]. The same group
subsequently showed that the Etest
"
method (bioMe
´
rieux,
France) and a MIC greater than 0.25 ll/ml for caspofungin
provided 100 % sensitivity and 94 % specificity for the
identification of FKS mutant isolates. Prior echinocandin
exposure and MIC values greater than 0.25, 0.06 and
0.03 ll/ml for caspofungin, anidulafungin and micafun-
gin, respectively, were found to be predictive of clinical
failure in 91, 89 and 78 % of patients with treatment
failure, respectively [6]. These findings have led to valu-
able, easy-to-use algorithms for predicting the outcome of
echinocandin treatment from MIC levels and prior echi-
nocandin exposure status. Not all mycology laboratories
are equipped to detect FKS mutations and the Etest
"
method performs well for the detection of non-wild-type
strains [109, 110]; this bedside strategy can therefore be
used to identify patients at risk of treatment failure, for
whom other antifungal treatments should be prescribed.
Similar predictive markers have been suggested for
azole resistance. Exposure to fluconazole in the last
30 days has been shown to have a significant impact on
species distribution and MIC [4]. Algorithms have also
been developed for assessment of the growing risk of C.
glabrata infections. Cohen et al. identified six indepen-
dent risk factors for C. glabrata fungemia in ICU patients:
age greater than 60 years, recent abdominal surgery, less
than 7 days between ICU admission and first positive
blood culture, recent use of cephalosporins, solid tumour
and absence of diabetes mellitus [111].
Integration of clinical and microbiological data (as
proposed in Fig. 3) is thus crucial to improve the pre-
diction of treatment response. Previous exposure to
fluconazole and echinocandin should be accurately mon-
itored, although the exact period to be considered remains
to be defined. Patients receiving prophylactic, empirical
or targeted antifungal therapy should be carefully moni-
tored for breakthrough infections. Local epidemiological
investigations and MIC determinations for Candida spp.
isolates are also crucial and should be interpreted with the
most appropriate, revised CBPs. Molecular tools are also
required for the rapid detection of mutant strains.
Impact of antifungal drug resistance on patient
management
Epidemiological changes have a direct impact on clinical
management, leading to the updating of international
expert committee recommendations [104, 112116].
These recommendations propose consensual attitudes to
the management of invasive candidiasis, but divergence
remains on several crucial, contentious points [107, 117,
118], which may be confusing for clinicians treating
patients.
All experts agree that patients with Candida-positive
blood cultures should be treated with systemic antifungal
drugs, but ESCMID cites echinocandins as the only initial
treatment with the highest levels of strength of recom-
mendation (A) and quality of evidence (I) [104], whereas
the European Conference on Infections in Leukaemia
(ECIL) and the Infectious Diseases Society of America
(IDSA) consider fluconazole at the AI level of recom-
mendation as a suitable alternative for patients with less
severe or stable infection not previously exposed to azoles
[10, 115, 116]. Indeed, in a recent study including 216
patients with Candida-induced septic shock, no difference
Fig. 3 Bedside strategy for circumventing antifungal drug resis-
tance in 2014. ATF antifungal drug, FCZ fluconazole, CAS
caspofungin, PK/PD pharmacokinetics and pharmacodynamics,
TDM therapeutic drug monitoring, MIC minimal inhibitory
concentration, CBP clinical breakpoint
in mortality was observed between patients treated with
fluconazole or with echinocandins [119]. In patients at
risk of C. glabrata candidemia, echinocandins should be
preferred. Voriconazole is not usually used as first-line
therapy but it offers an alternate option for intrinsically
less susceptible species (C. krusei or C. glabrata).
Because acquired mutations can lead to cross-resistance
to both fluconazole and voriconazole, a strain resistant to
fluconazole should not be treated with voriconazole
unless its susceptibility profile has been confirmed, or the
mutation genetically characterized.
All expert panels strongly recommend the removal of
central venous catheters ‘whenever possible’, but ESC-
MID guidelines suggest that replacement is not formally
required in patients treated with echinocandins or L-AMB
[104]. Catheter exchange via a guide wire entails a risk of
contaminating the new device with Candida and should
be restricted to patients with limited venous access [107,
116, 120]. Given the specific link between C. parapsilosis
and catheter infections and the low susceptibility of this
species to echinocandins, catheter removal is appropriate
in patients with invasive C. parapsilosis candidiasis. In
stabilised patients infected with a fluconazole-susceptible
isolate, with negative blood cultures, step-down therapy
onto oral fluconazole is recommended, over a period of
3–10 days, depending on the guidelines considered.
Conclusions
Although drug resistance is rapidly spreading in Candida
spp., antifungal treatments are still generally successful:
up to 80 % of C. albicans infections are cleared with
echinocandins. Treatment success rates are also generally
satisfactory for fluconazole. However, the emergence of
antifungal resistance must be considered at the patient
level in order to improve patient management. In ICUs,
intrinsic resistance of C. glabrata and C. krusei to
fluconazole can be detected and handled rapidly through
correct species identification, detailed assessment of
antifungal drug exposure and Candida spp. colonisation
history. The emergence of acquired resistance during or
after treatment is more worrying: it mostly involves C.
glabrata and the echinocandins and leads to breakthrough
infections or treatment failures. This highlights the need
for (a) accurate assessments of prior antifungal exposure,
(b) close monitoring of patients on antifungal drugs,
(c) the routine surveillance of in vitro susceptibility test-
ing and (d) the development of feasible methods for rapid
detection of mutations. The emergence of resistance
should also be considered at the community level as a
warning sign triggering improvements in antifungal drug
use, particularly in patients for whom the potential benefit
of treatment is unproven. Closer monitoring of antifungal
drug use is thus required.
Acknowledgments We are grateful to Audrey Le Goue
¨
llec for her
assistance in preparing the figures.
Conflicts of interest D. Maubon, C. Garnaud and M. Cornet
received a research grant from Pfizer in 2013. T. Calandra: board
membership: Pfizer; Consultancy: Pfizer, MSD; Speakers bureaus:
BioMe
´
rieux, Pfizer; Development & educational presentations:
MSD, Gilead Sciences (money to institution); Travel & meeting
expenses: Astellas, Pfizer.
References
1. Leo
´
n C, Ostrosky-Zeichner L,
Schuster M (2014) What’s new in the
clinical and diagnostic management of
invasive candidiasis in critically ill
patients. Intensive Care Med. doi:
10.1007/s00134-014-3281-0
2. Guery BP, Arendrup MC, Auzinger G
et al (2009) Management of invasive
candidiasis and candidemia in adult
non-neutropenic intensive care unit
patients: part II. Treatment Intensive
Care Med 35:206–214. doi:
10.1007/s00134-008-1339-6
3. Fournier P, Schwebel C, Maubon D
et al (2011) Antifungal use influences
Candida species distribution and
susceptibility in the intensive care
unit. J Antimicrob Chemother
66:2880–2886
4. Lortholary O, Desnos-Ollivier M,
Sitbon K et al (2011) Recent exposure
to caspofungin or fluconazole
influences the epidemiology of
candidemia: a prospective multicenter
study involving 2,441 patients.
Antimicrob Agents Chemother
55:532–538. doi:
10.1128/AAC.01128-10
5. Alexander BD, Johnson MD, Pfeiffer
CD et al (2013) Increasing
echinocandin resistance in Candida
glabrata: clinical failure correlates
with presence of FKS mutations and
elevated minimum inhibitory
concentrations. Clin Infect Dis
56:1724–1732. doi:10.1093/cid/cit136
6. Shields RK, Nguyen MH, Press EG
et al (2013) Caspofungin MICs
correlate with treatment outcomes
among patients with Candida glabrata
invasive candidiasis and prior
echinocandin exposure. Antimicrob
Agents Chemother 57:3528–3535. doi:
10.1128/AAC.00136-13
7. Perlin DS (2011) Current perspectives
on echinocandin class drugs. Future
Microbiol 6:441–457. doi:
10.2217/fmb.11.19
8. Odds FC, Brown AJP, Gow NAR
(2003) Antifungal agents: mechanisms
of action. Trends Microbiol
11:272–279
9. Sorgo AG, Heilmann CJ, Dekker HL
et al (2011) Effects of fluconazole on
the secretome, the wall proteome, and
wall integrity of the clinical fungus
Candida albicans. Eukaryot Cell
10:1071–1081. doi:
10.1128/EC.05011-11
10. Pfaller M, Riley J (1992) Effects of
fluconazole on the sterol and
carbohydrate composition of four
species of Candida. Eur J Clin
Microbiol Infect Dis 11:152–156
11. Gray KC, Palacios DS, Dailey I et al
(2012) Amphotericin primarily kills
yeast by simply binding ergosterol.
Proc Natl Acad Sci U S A
109:2234–2239. doi:
10.1073/pnas.1117280109
12. Anderson TM, Clay MC, Cioffi AG
et al (2014) Amphotericin forms an
extramembranous and fungicidal
sterol sponge. Nat Chem Biol
10:400–406. doi:
10.1038/nchembio.1496
13. Kotler-Brajtburg J, Price HD, Medoff
G et al (1974) Molecular basis for the
selective toxicity of amphotericin B
for yeast and filipin for animal cells.
Antimicrob Agents Chemother
5:377–382
14. Hamill RJ (2013) Amphotericin B
formulations: a comparative review of
efficacy and toxicity. Drugs
73:919–934. doi:
10.1007/s40265-013-0069-4
15. Walker LA, Gow NAR, Munro CA
(2010) Fungal echinocandin
resistance. Fungal Genet Biol
47:117–126. doi:
10.1016/j.fgb.2009.09.003
16. Sanglard D, Ischer F, Marchetti O et al
(2003) Calcineurin A of Candida
albicans: involvement in antifungal
tolerance, cell morphogenesis and
virulence. Mol Microbiol 48:959–976
17. LaFayette SL, Collins C, Zaas AK
et al (2010) PKC signaling regulates
drug resistance of the fungal pathogen
Candida albicans via circuitry
comprised of Mkc1, calcineurin, and
Hsp90. PLoS Pathog 6:e1001069. doi:
10.1371/journal.ppat.1001069
18. Lee KK, Maccallum DM, Jacobsen
MD et al (2012) Elevated cell wall
chitin in Candida albicans confers
echinocandin resistance in vivo.
Antimicrob Agents Chemother
56:208–217. doi:
10.1128/AAC.00683-11
19. Rueda C, Cuenca-Estrella M,
Zaragoza O (2014) Paradoxical
growth of Candida albicans in the
presence of caspofungin is associated
with multiple cell wall rearrangements
and decreased virulence. Antimicrob
Agents Chemother 58:1071–1083. doi:
10.1128/AAC.00946-13
20. Arendrup MC, Cuenca-Estrella M,
Lass-Flo
¨
rl C, Hope WW (2014)
Breakpoints for antifungal agents: an
update from EUCAST focussing on
echinocandins against Candida spp
and triazoles against Aspergillus spp.
Drug Resist Updat 16(6):81–95. doi:
10.1016/j.drup.2014.01.001
21. Coste A, Selmecki A, Forche A et al
(2007) Genotypic evolution of azole
resistance mechanisms in sequential
Candida albicans isolates. Eukaryot
Cell 6:1889–1904
22. Sanglard D, Coste A, Ferrari S (2009)
Antifungal drug resistance
mechanisms in fungal pathogens from
the perspective of transcriptional gene
regulation. FEMS Yeast Res
9:1029–1050. doi:
10.1111/j.1567-1364.2009.00578.x
23. Coste A, Turner V, Ischer F et al
(2006) A mutation in Tac1p, a
transcription factor regulating CDR1
and CDR2, is coupled with loss of
heterozygosity at chromosome 5 to
mediate antifungal resistance in
Candida albicans. Genetics
172:2139–2156. doi:
10.1534/genetics.105.054767
24. Morio F, Loge C, Besse B et al (2010)
Screening for amino acid substitutions
in the Candida albicans Erg11 protein
of azole-susceptible and azole-
resistant clinical isolates: new
substitutions and a review of the
literature. Diagn Microbiol Infect Dis
66:373–384. doi:
10.1016/j.diagmicrobio.2009.11.006
25. Morio F, Pagniez F, Lacroix C et al
(2012) Amino acid substitutions in the
Candida albicans sterol D5,6-
desaturase (Erg3p) confer azole
resistance: characterization of two
novel mutants with impaired
virulence. J Antimicrob Chemother
67:2131–2138. doi:
10.1093/jac/dks186
26. Pola
´
kova
´
S, Blume C, Za
´
rate JA et al
(2009) Formation of new
chromosomes as a virulence
mechanism in yeast Candida glabrata.
Proc Natl Acad Sci U S A
106:2688–2693. doi:
10.1073/pnas.0809793106
27. Ferrari S, Sanguinetti M, De Bernardis
F et al (2011) Loss of mitochondrial
functions associated with azole
resistance in Candida glabrata results
in enhanced virulence in mice.
Antimicrob Agents Chemother
55:1852–1860. doi:
10.1128/AAC.01271-10
28. Vincent BM, Lancaster AK, Scherz-
Shouval R et al (2013) Fitness trade-
offs restrict the evolution of resistance
to amphotericin B. PLoS Biol
11:e1001692. doi:
10.1371/journal.pbio.1001692
29. Spampinato C, Leonardi D (2013)
Candida infections, causes, targets,
and resistance mechanisms: traditional
and alternative antifungal agents. Bio
Med Res Int 2013:204237. doi:
10.1155/2013/204237
30. Florent M, Noe
¨
l T, Ruprich-Robert G
et al (2009) Nonsense and missense
mutations in FCY2 and FCY1 genes
are responsible for flucytosine
resistance and flucytosine-fluconazole
cross-resistance in clinical isolates of
Candida lusitaniae. Antimicrob
Agents Chemother 53:2982–2990. doi:
10.1128/AAC.00880-08
31. Delattin N, Cammue BPA, Thevissen
K (2014) Reactive oxygen species-
inducing antifungal agents and their
activity against fungal biofilms. Future
Med Chem 6:77–90. doi:
10.4155/fmc.13.189
32. Kuhn DM, George T, Chandra J et al
(2002) Antifungal susceptibility of
Candida biofilms: unique efficacy of
amphotericin B lipid formulations and
echinocandins. Antimicrob Agents
Chemother 46:1773–1780
33. Choi HW, Shin JH, Jung SI et al
(2007) Species-specific differences in
the susceptibilities of biofilms formed
by Candida bloodstream isolates to
echinocandin antifungals. Antimicrob
Agents Chemother 51:1520–1523. doi:
10.1128/AAC.01141-06
34. Lamfon H, Porter SR, McCullough M,
Pratten J (2004) Susceptibility of
Candida albicans biofilms grown in a
constant depth film fermentor to
chlorhexidine, fluconazole and
miconazole: a longitudinal study.
J Antimicrob Chemother 53:383–385.
doi:10.1093/jac/dkh071
35. Taff HT, Mitchell KF, Edward JA,
Andes DR (2013) Mechanisms of
Candida biofilm drug resistance.
Future Microbiol 8:1325–1337. doi:
10.2217/fmb.13.101
36. Tobudic S, Kratzer C, Lassnigg A,
Presterl E (2012) Antifungal
susceptibility of Candida albicans in
biofilms. Mycoses 55:199–204. doi:
10.1111/j.1439-0507.2011.02076.x
37. Mathe
´
L, Van Dijck P (2013) Recent
insights into Candida albicans biofilm
resistance mechanisms. Curr Genet.
doi:10.1007/s00294-013-0400-3
38. Mitchell KF, Taff HT, Cuevas MA
et al (2013) Role of matrix b-1,3
glucan in antifungal resistance of non-
albicans Candida biofilms.
Antimicrob Agents Chemother
57:1918–1920. doi:
10.1128/AAC.02378-12
39. Robbins N, Uppuluri P, Nett J et al
(2011) Hsp90 governs dispersion and
drug resistance of fungal biofilms.
PLoS Pathog 7:e1002257. doi:
10.1371/journal.ppat.1002257
40. Martins M, Henriques M, Lopez-Ribot
JL, Oliveira R (2012) Addition of
DNase improves the in vitro activity of
antifungal drugs against Candida
albicans biofilms. Mycoses 55:80–85.
doi:
10.1111/j.1439-0507.2011.02047.x
41. EUCAST (2008) EUCAST definitive
document EDef 7.1: method for the
determination of broth dilution MICs
of antifungal agents for fermentative
yeasts. Clin Microbiol Infect
14:398–405
42. CLSI Clinical and Laboratory
Standards Institute (2008) Reference
method for broth dilution antifungal
susceptibility testing of yeasts. Third
informational supplement. CLSI
document M27-S3. Clinical and
Laboratory Standards Institute, Wayne
43. Chryssanthou E, Cuenca-Estrella M
(2002) Comparison of the antifungal
susceptibility testing subcommittee
of the European Committee on
Antibiotic Susceptibility Testing
proposed standard and the E-test with
the NCCLS broth microdilution
method for voriconazole and
caspofungin susceptibility testing of
yeast species. J Clin Microbiol
40:3841–3844
44. Cuenca-Estrella M, Gomez-Lopez A,
Alastruey-Izquierdo A et al (2010)
Comparison of the Vitek 2 antifungal
susceptibility system with the clinical
and laboratory standards institute
(CLSI) and European Committee on
Antimicrobial Susceptibility Testing
(EUCAST) broth microdilution
reference methods and with the
sensititre yeastone and Etest
techniques for in vitro detection of
antifungal resistance in yeast isolates.
J Clin Microbiol 48:1782–1786. doi:
10.1128/JCM.02316-09
45. Lombardi G, Farina C, Andreoni S
et al (2004) Comparative evaluation of
Sensititre YeastOne vs. the NCCLS
M27A protocol and E-test for
antifungal susceptibility testing of
yeasts. Mycoses 47:397–401. doi:
10.1111/j.1439-0507.2004.01013.x
46. Pfaller MA, Diekema DJ, Andes D
et al (2011) Clinical breakpoints for
the echinocandins and Candida
revisited: integration of molecular,
clinical, and microbiological data to
arrive at species-specific interpretive
criteria. Drug Resist Updat
14:164–176. doi:
10.1016/j.drup.2011.01.004
47. Arendrup MC, Garcia-Effron G,
Buzina W et al (2009) Breakthrough
Aspergillus fumigatus and Candida
albicans double infection during
caspofungin treatment: laboratory
characteristics and implication for
susceptibility testing. Antimicrob
Agents Chemother 53:1185–1193. doi:
10.1128/AAC.01292-08
48. Pfaller MA, Andes D, Diekema DJ
et al (2010) Wild-type MIC
distributions, epidemiological cutoff
values and species-specific clinical
breakpoints for fluconazole and
Candida: time for harmonization of
CLSI and EUCAST broth
microdilution methods. Drug Resist
Updat 13:180–195. doi:
10.1016/j.drup.2010.09.002
49. Espinel-Ingroff A, Arendrup MC,
Pfaller MA et al (2013) Interlaboratory
variability of caspofungin MICs for
Candida spp. Using CLSI and
EUCAST methods: should the clinical
laboratory be testing this agent?
Antimicrob Agents Chemother
57:5836–5842. doi:
10.1128/AAC.01519-13
50. Pfaller MA, Messer SA, Diekema DJ
et al (2014) Use of micafungin as a
surrogate marker to predict
susceptibility and resistance to
caspofungin among 3,764 clinical
isolates of Candida by use of CLSI
methods and interpretive criteria.
J Clin Microbiol 52:108–114. doi:
10.1128/JCM.02481-13
51. Eschenauer GA, Nguyen MH, Shoham
S et al (2014) Real-world experience
with echinocandin MICs against
Candida species in a multicenter study
of hospitals that routinely perform
susceptibility testing of bloodstream
isolates. Antimicrob Agents
Chemother 58:1897–1906. doi:
10.1128/AAC.02163-13
52. Shields RK, Nguyen MH, Press EG
et al (2013) Anidulafungin and
micafungin minimum inhibitory
concentration breakpoints are superior
to caspofungin for identifying FKS
mutant Candida glabrata and
echinocandin resistance. Antimicrob
Agents Chemother. doi:
10.1128/AAC.01451-13
53. Park BJ, Arthington-Skaggs BA,
Hajjeh RA et al (2006) Evaluation of
amphotericin B interpretive
breakpoints for Candida bloodstream
isolates by correlation with therapeutic
outcome. Antimicrob Agents
Chemother 50:1287–1292. doi:
10.1128/AAC.50.4.1287-1292.2006
54. Krogh-Madsen M, Arendrup MC,
Heslet L, Knudsen JD (2006)
Amphotericin B and caspofungin
resistance in Candida glabrata isolates
recovered from a critically ill patient.
Clin Infect Dis 42:938–944. doi:
10.1086/500939
55. Guinea J, Recio S, Escribano P et al
(2010) Rapid antifungal susceptibility
determination for yeast isolates by use
of Etest performed directly on blood
samples from patients with fungemia.
J Clin Microbiol 48:2205–2212. doi:
10.1128/JCM.02321-09
56. Wiederhold NP, Grabinski JL, Garcia-
Effron G et al (2008) Pyrosequencing
to detect mutations in FKS1 that
confer reduced echinocandin
susceptibility in Candida albicans.
Antimicrob Agents Chemother
52:4145–4148. doi:
10.1128/AAC.00959-08
57. Park S, Perlin DS (2005) Establishing
surrogate markers for fluconazole
resistance in Candida albicans.
Microb Drug Resist 11:232–238. doi:
10.1089/mdr.2005.11.232
58. Loeffler J, Hagmeyer L, Hebart H et al
(2000) Rapid detection of point
mutations by fluorescence resonance
energy transfer and probe melting
curves in Candida species. Clin Chem
46:631–635
59. Wang H, Kong F, Sorrell TC et al
(2009) Rapid detection of ERG11
gene mutations in clinical Candida
albicans isolates with reduced
susceptibility to fluconazole by rolling
circle amplification and DNA
sequencing. BMC Microbiol 9:167.
doi:10.1186/1471-2180-9-167
60. Pham CD, Bolden CB, Kuykendall RJ,
Lockhart SR (2013) Development of a
Luminex-based multiplex assay for
detection of mutations conferring
resistance to echinocandins in
Candida glabrata. J Clin Microbiol.
doi:10.1128/JCM.03378-13
61. Zhao Y, Stensvold CR, Perlin DS,
Arendrup MC (2013) Azole resistance
in Aspergillus fumigatus from
bronchoalveolar lavage fluid samples
of patients with chronic diseases.
J Antimicrob Chemother
68:1497–1504. doi:10.1093/jac/dkt071
62. Bassetti M, Merelli M, Righi E et al
(2013) Epidemiology, species
distribution, antifungal susceptibility,
and outcome of candidemia across five
sites in Italy and Spain. J Clin
Microbiol 51:4167–4172. doi:
10.1128/JCM.01998-13
63. Pfaller MA, Messer SA, Moet GJ et al
(2011) Candida bloodstream
infections: comparison of species
distribution and resistance to
echinocandin and azole antifungal
agents in intensive care unit (ICU) and
non-ICU settings in the SENTRY
antimicrobial surveillance program
(2008–2009). Int J Antimicrob Agents
38:65–69. doi:
10.1016/j.ijantimicag.2011.02.016
64. Puig-Asensio M, Pema
´
n J, Zaragoza R
et al (2014) Impact of therapeutic
strategies on the prognosis of
candidemia in the ICU. Crit Care Med
42:1423–1432. doi:
10.1097/CCM.0000000000000221
65. Montagna MT, Caggiano G, Lovero G
et al (2013) Epidemiology of invasive
fungal infections in the intensive care
unit: results of a multicenter Italian
survey (AURORA Project). Infection
41:645–653. doi:
10.1007/s15010-013-0432-0
66. Diekema D, Arbefeville S, Boyken L
et al (2012) The changing
epidemiology of healthcare-associated
candidemia over three decades. Diagn
Microbiol Infect Dis 73:45–48. doi:
10.1016/j.diagmicrobio.2012.02.001
67. Chow JK, Golan Y, Ruthazer R et al
(2008) Factors associated with
candidemia caused by non-albicans
Candida species versus Candida
albicans in the intensive care unit.
Clin Infect Dis 46:1206–1213. doi:
10.1086/529435
68. Kanafani ZA, Perfect JR (2008)
Resistance to antifungal agents:
mechanisms and clinical impact. Clin
Infect Dis 46:120–128. doi:
10.1086/524071
69. Guinea J (2014) Global trends in the
distribution of Candida species
causing candidemia. Clin Microbiol
Infect. doi:10.1111/1469-0691.12539
70. Pfaller MA, Diekema DJ (2010)
Epidemiology of invasive mycoses in
North America. Crit Rev Microbiol
36:1–53
71. Guitard J, Angoulvant A, Letscher-
Bru V et al (2013) Invasive infections
due to Candida norvegensis and
Candida inconspicua: report of 12
cases and review of the literature. Med
Mycol. doi:
10.3109/13693786.2013.807444
72. Pfaller MA, Diekema DJ, Colombo
AL et al (2006) Candida rugosa, an
emerging fungal pathogen with
resistance to azoles: geographic and
temporal trends from the ARTEMIS
DISK antifungal surveillance program.
J Clin Microbiol 44:3578–3582. doi:
10.1128/JCM.00863-06
73. Garcia-Effron G, Katiyar SK, Park S
et al (2008) A naturally occurring
proline-to-alanine amino acid change
in FKS1p in Candida parapsilosis,
Candida orthopsilosis, and Candida
metapsilosis accounts for reduced
echinocandin susceptibility.
Antimicrob Agents Chemother
52:2305–2312. doi:
10.1128/AAC.00262-08
74. Canto
´
n E, Pema
´
n J, Sastre M et al
(2006) Killing kinetics of caspofungin,
micafungin, and amphotericin B
against Candida guilliermondii.
Antimicrob Agents Chemother
50:2829–2832. doi:
10.1128/AAC.00524-06
75. Kabbara N, Lacroix C, Peffault de
Latour R et al (2008) Breakthrough C.
parapsilosis and C. guilliermondii
blood stream infections in allogeneic
hematopoietic stem cell transplant
recipients receiving long-term
caspofungin therapy. Haematologica
93:639–640. doi:
10.3324/haematol.11149
76. Pfeiffer CD, Garcia-Effron G, Zaas
AK et al (2010) Breakthrough invasive
candidiasis in patients on micafungin.
J Clin Microbiol 48:2373–2380
77. Anderson JB (2005) Evolution of
antifungal-drug resistance:
mechanisms and pathogen fitness. Nat
Rev Microbiol 3:547–556. doi:
10.1038/nrmicro1179
78. Marco F, Lockhart SR, Pfaller MA
et al (1999) Elucidating the origins of
nosocomial infections with Candida
albicans by DNA fingerprinting with
the complex probe Ca3. J Clin
Microbiol 37:2817–2828
79. Ste
´
phan F, Bah MS, Desterke C et al
(2002) Molecular diversity and routes
of colonization of Candida albicans in
a surgical intensive care unit, as
studied using microsatellite markers.
Clin Infect Dis 35:1477–1483. doi:
10.1086/344648
80. Pfaller MA, Castanheira M, Lockhart
SR et al (2012) Frequency of
decreased susceptibility and resistance
to echinocandins among fluconazole-
resistant bloodstream isolates of
Candida glabrata. J Clin Microbiol
50:1199–1203. doi:
10.1128/JCM.06112-11
81. Cleveland AA, Farley MM, Harrison
LH et al (2012) Changes in incidence
and antifungal drug resistance in
candidemia: results from population-
based laboratory surveillance in
Atlanta and Baltimore, 2008–2011.
Clin Infect Dis 55:1352–1361. doi:
10.1093/cid/cis697
82. Garcia-Effron G, Chua DJ, Tomada JR
et al (2010) Novel FKS mutations
associated with echinocandin
resistance in Candida species.
Antimicrob Agents Chemother
54:2225–2227. doi:
10.1128/AAC.00998-09
83. Garcia-Effron G, Kontoyiannis DP,
Lewis RE, Perlin DS (2008)
Caspofungin-resistant Candida
tropicalis strains causing breakthrough
Fungemia in patients at high risk for
hematologic malignancies. Antimicrob
Agents Chemother 52:4181–4183. doi:
10.1128/AAC.00802-08
84. Hakki M, Staab JF, Marr KA (2006)
Emergence of a Candida krusei Isolate
with reduced susceptibility to
Caspofungin during therapy.
Antimicrob Agents Chemother
50:2522–2524. doi:
10.1128/AAC.00148-06
85. Fekkar A, Meyer I, Brossas JY et al
(2013) Rapid emergence of
echinocandin resistance during
Candida kefyr fungemia treatment
with caspofungin. Antimicrob Agents
Chemother 57:2380–2382. doi:
10.1128/AAC.02037-12
86. Pfaller MA (2012) Antifungal drug
resistance: mechanisms,
epidemiology, and consequences for
treatment. Am J Med 125:S3–S13.
doi:10.1016/j.amjmed.2011.11.001
87. Bizerra FC, Jimenez-Ortigosa C,
Souza ACR et al (2014) Breakthrough
candidemia due to multidrug resistant
C. glabrata during prophylaxis with
low dose of micafungin. Antimicrob
Agents Chemother 58:2438–2440. doi:
10.1128/AAC.02189-13
88. Chan TSY, Gill H, Hwang Y–Y et al
(2014) Breakthrough invasive fungal
diseases during echinocandin
treatment in high-risk hospitalized
hematologic patients. Ann Hematol
93:493–498. doi:
10.1007/s00277-013-1882-2
89. Sun H-Y, Singh N (2010)
Characterisation of breakthrough
invasive mycoses in echinocandin
recipients: an evidence-based review.
Int J Antimicrob Agents 35:211–218.
doi:10.1016/j.ijantimicag.2009.09.020
90. Myoken Y, Kyo T, Sugata T et al
(2006) Breakthrough fungemia caused
by fluconazole-resistant Candida
albicans with decreased susceptibility
to voriconazole in patients with
hematologic malignancies.
Haematologica 91:287–288
91. Beyda ND, John J, Kilic A et al (2014)
FKS mutant Candida glabrata; risk
factors and outcomes in patients with
candidemia. Clin Infect Dis. doi:
10.1093/cid/ciu407
92. Lackner M, Tscherner M, Schaller M
et al (2014) Positions and numbers of
FKS mutations in Candida albicans
selectively influence in vitro and
in vivo susceptibilities to echinocandin
treatment. Antimicrob Agents
Chemother 58:3626–3635. doi:
10.1128/AAC.00123-14
93. Liu W, Tan J, Sun J et al (2014)
Invasive candidiasis in intensive care
units in China: in vitro antifungal
susceptibility in the China-SCAN
study. J Antimicrob Chemother
69:162–167. doi:10.1093/jac/dkt330
94. Chapeland-Leclerc F, Hennequin C,
Papon N et al (2010) Acquisition of
flucytosine, azole, and caspofungin
resistance in Candida glabrata
bloodstream isolates serially obtained
from a hematopoietic stem cell
transplant recipient. Antimicrob
Agents Chemother 54:1360–1362. doi:
10.1128/AAC.01138-09
95. Pfaller MA, Diekema DJ (2007)
Epidemiology of invasive candidiasis:
a persistent public health problem.
Clin Microbiol Rev 20:133–163. doi:
10.1128/CMR.00029-06
96. Yang Y-L, Li S-Y, Cheng H–H et al
(2005) The trend of susceptibilities to
amphotericin B and fluconazole of
Candida species from 1999 to 2002 in
Taiwan. BMC Infect Dis 5:99. doi:
10.1186/1471-2334-5-99
97. Kontoyiannis DP, Lewis RE (2002)
Antifungal drug resistance of
pathogenic fungi. Lancet
359:1135–1144. doi:
10.1016/S0140-6736(02)08162-X
98. Silva S, Negri M, Henriques M et al
(2012) Candida glabrata, Candida
parapsilosis and Candida tropicalis:
biology, epidemiology, pathogenicity
and antifungal resistance. FEMS
Microbiol Rev 36:288–305. doi:
10.1111/j.1574-6976.2011.00278.x
99. Segal BH, Herbrecht R, Stevens DA
et al (2008) Defining responses to
therapy and study outcomes in clinical
trials of invasive fungal diseases:
Mycoses Study Group and European
Organization for Research and
Treatment of Cancer consensus
criteria. Clin Infect Dis 47:674–683.
doi:10.1086/590566
100. Walsh TJ, Finberg RW, Arndt C et al
(1999) Liposomal amphotericin B for
empirical therapy in patients with
persistent fever and neutropenia.
N Engl J Med 340:764–771. doi:
10.1056/NEJM199903113401004
101. Moen MD, Lyseng-Williamson KA,
Scott LJ (2009) Liposomal
amphotericin B. Drugs 69:361–392.
doi:
10.2165/00003495-200969030-00010
102. Andes DR, Safdar N, Baddley JW et al
(2012) Impact of treatment strategy on
outcomes in patients with candidemia
and other forms of invasive
candidiasis: a patient-level
quantitative review of randomized
trials. Clin Infect Dis 54:1110–1122.
doi:10.1093/cid/cis021
103. Puig-Asensio M, Pema
´
n J, Zaragoza R
et al (2014) Impact of therapeutic
strategies on the prognosis of
candidemia in the ICU. Crit Care Med.
doi:10.1097/CCM.0000000000000221
104. Cornely OA, Bassetti M, Calandra T
et al (2012) ESCMID guideline for the
diagnosis and management of Candida
diseases 2012: non-neutropenic adult
patients. Clin Microbiol Infect
18(Suppl 7):19–37. doi:
10.1111/1469-0691.12039
105. Pfaller MA, Messer SA, Boyken L
et al (2002) In vitro activities of
5-fluorocytosine against 8,803 clinical
isolates of Candida spp.: global
assessment of primary resistance using
national committee for clinical
laboratory standards susceptibility
testing methods. Antimicrob Agents
Chemother 46:3518–3521
106. Felton T, Troke PF, Hope WW (2014)
Tissue penetration of antifungal
agents. Clin Microbiol Rev 27:68–88.
doi:10.1128/CMR.00046-13
107. Glo
¨
ckner A, Cornely OA (2013)
Practical considerations on current
guidelines for the management of non-
neutropenic adult patients with
candidaemia: practical considerations
on current guidelines. Mycoses
56:11–20. doi:
10.1111/j.1439-0507.2012.02208.x
108. Shields RK, Nguyen MH, Press EG
et al (2012) The presence of an FKS
mutation rather than MIC is an
independent risk factor for failure of
echinocandin therapy among patients
with invasive candidiasis due to
Candida glabrata. Antimicrob Agents
Chemother 56:4862–4869. doi:
10.1128/AAC.00027-12
109. Baixench M-T, Aoun N, Desnos-
Ollivier M et al (2007) Acquired
resistance to echinocandins in
Candida albicans: case report and
review. J Antimicrob Chemother
59:1076–1083. doi:
10.1093/jac/dkm095
110. Bourgeois N, Laurens C, Bertout S
et al (2014) Assessment of
caspofungin susceptibility of Candida
glabrata by the Etest
"
, CLSI, and
EUCAST methods, and detection of
FKS1 and FKS2 mutations. Eur J Clin
Microbiol Infect Dis. doi:
10.1007/s10096-014-2069-z
111. Cohen Y, Karoubi P, Adrie C et al
(2010) Early prediction of Candida
glabrata fungemia in nonneutropenic
critically ill patients. Crit Care Med
38:826–830. doi:
10.1097/CCM.0b013e3181cc4734
112. Ruhnke M, Rickerts V, Cornely OA
et al (2011) Diagnosis and therapy of
Candida infections: joint
recommendations of the German
Speaking Mycological Society and the
Paul-Ehrlich-Society for
Chemotherapy. Mycoses 54:279–310.
doi:
10.1111/j.1439-0507.2011.02040.x
113. Colombo AL, Guimara
˜
es T, Camargo
LFA et al (2013) Brazilian guidelines
for the management of candidiasis—a
joint meeting report of three medical
societies: Sociedade Brasileira de
Infectologia, Sociedade Paulista de
Infectologia and Sociedade Brasileira
de Medicina Tropical. Braz J Infect
Dis 17:283–312. doi:
10.1016/j.bjid.2013.02.001
114. Bow EJ, Evans G, Fuller J et al (2010)
Canadian clinical practice guidelines
for invasive candidiasis in adults. Can
J Infect Dis Med Microbiol 21:e122–
e150
115. Castagna L, Bramanti S, Sarina B et al
(2012) ECIL 3-2009 update guidelines
for antifungal management. Bone
Marrow Transpl 47:866
116. Pappas PG, Kauffman CA, Andes D
et al (2009) Candida-clinical practice
guidelines for the management of
candidiasis: 2009 update by the
Infectious Diseases Society of
America. Clin Infect Dis 48:503–535.
doi:10.1086/596757
117. Leroux S, Ullmann AJ (2013)
Management and diagnostic
guidelines for fungal diseases in
infectious diseases and clinical
microbiology: critical appraisal. Clin
Microbiol Infect 19:1115–1121. doi:
10.1111/1469-0691.12426
118. Deshpande A, Gaur S, Bal AM (2013)
Candidaemia in the non-neutropenic
patient: a critique of the guidelines. Int
J Antimicrob Agents 42:294–300. doi:
10.1016/j.ijantimicag.2013.06.005
119. Bassetti M, Righi E, Ansaldi F et al
(2014) A multicenter study of septic
shock due to candidemia: outcomes
and predictors of mortality. Intensive
Care Med 40:839–845. doi:
10.1007/s00134-014-3310-z
120. Mermel LA, Allon M, Bouza E et al
(2009) Clinical practice guidelines for
the diagnosis and management of
intravascular catheter-related
infection: 2009 update by the
Infectious Diseases Society of
America. Clin Infect Dis 49:1–45. doi:
10.1086/599376
... Owing to the high burden of HIV/AIDS, PLHIV in Africa have an increased risk of OPC caused by a wide variety of Candida pathogens. A shift in the etiology associated with OPC towards antifungal-resistant NAC in Africa could explain the increase in cumulative prevalence, which increased from 32% in 2014 to 69% in 2019 and then decreased to 48% in 2022 (26, 39,40). ...
... The predominant NAC species were C. glabrata (29.6%), C. tropicalis (27.7%), C. krusei (17.0%), C. parapsilosis (8.1%) and C. dubliniensis (5.2%). Our results support observations that have been reported in several other studies regarding the epidemiological shift of OPC etiology toward NAC species, and this has accounted for their emergence as a signi cant Candida pathogen (26,39,40). For instance, the distribution of NAC species in our study agreed with studies conducted in Indonesia, Iran and India that identi ed C. glabrata (15-19%), C. krusei (4.6-15%) and C. tropicalis (4.6-10%) as the most prevalent NAC species (44)(45)(46). ...
Preprint
Full-text available
Background: The incidence of oropharyngeal candidiasis among people living with human immunodeficiency virus in Africa is on the rise. Oropharyngeal candidiasis is mainly caused by C. albicans; however, a shift in the etiology towards non-Candida albicans species is increasing. In addition, there are variations in the epidemiological distribution of Candida species causing oropharyngeal candidiasis among people living with human immunodeficiency virus in Africa. Objective: This review aimed to determine the prevalence of oropharyngeal candidiasis and the distribution of Candida species among people living with human immunodeficiency virus in Africa. Materials and Methods: This systematic review protocol was registered in the base PROSPERO database prior to its conduct (CRD42021254473). The Preferred Reporting Items for Systematic Reviews and Meta-Analyses Protocol guidelines (PRISMA-P) were followed for this study. The PubMed, Scopus and EMBASE databases were searched to identify published studies published between 1st January 2000 and 8th October 2022. The eligible studies were included in the meta-analysis and analyzed using a random effects model. The risk of bias of the included studies was assessed using the Joanna Briggs Institute quality assessment tool for prevalence studies. Results: The database search yielded 370 titles from PubMed (n=192), EMBASE (n=162) and SCOPUS (n=16). Fourteen studies with a total of 3,863 participants were included in the meta-analysis. The pooled prevalence of oropharyngeal candidiasis was 49.0% (95% CI: 37% - 62%). A total of 2,688 Candida isolates were reported; approximately 76.6% (n=2,060) were C. albicans, and 21.7% (n=582) were non-C. albicans. Among the non-Candida albicans species, C. glabrata was the most common isolate (29.6%), followed by C. tropicalis (27.7%), C. krusei (17.0%), C. parapsilosis (8.1%) and C. dubliniensis (5.2%). Out of 14 studies, 7 (50.0%) had a low risk of bias, 5 (35.7%) had a moderate risk of bias, and 2 (14.3%) had a high risk of bias. Conclusion: Almost half of people living with HIV in Africa have oropharyngeal candidiasis, and C. albicans remains the most frequent cause of oropharyngeal candidiasis.
... A 8 (MIC 80 = 0.7 µg/mL) showed the best inhibitory activity against C. albicans 5122. For the C. albicans 5172, A 16 (MIC 80 = 5.8 µg/mL) displayed the best inhibitory activity against it. ...
... According to the above data, the TAI values of twenty-one compounds (A 2 , A 4 , A 5 , A 6 , A 9 , A 10 , A 11 , A 13 , A 16 , B 1 , B 5 , B 7 , B 10 , B 11 , B 14 , C 1 , C 3 , C 4 , C 7 , D 1 , and D 5 ) with MIC 80 ≤ 64 µg/mL were calculated according to the formula TAI = ∑ n 1 1/ √ MIC 80 , which reflected the level of antifungal activity of the compounds, and the larger the index, the better the antifungal activity of the compound [35]. ...
Article
Full-text available
Fifty-two kinds of N′-phenylhydrazides were successfully designed and synthesized. Their antifungal activity in vitro against five strains of C. albicans (Candida albicans) was evaluated. All prepared compounds showed varying degrees of antifungal activity against C. albicans and their MIC80 (the concentration of tested compounds when their inhibition rate was at 80%), TAI (total activity index), and TSI (total susceptibility index) were calculated. The inhibitory activities of 27/52 compounds against fluconazole-resistant fungi C. albicans 4395 and 5272 were much better than those of fluconazole. The MIC80 values of 14/52 compounds against fluconazole-resistant fungus C. albicans 5122 were less than 4 μg/mL, so it was the most sensitive fungus (TSIB = 12.0). A11 showed the best inhibitory activity against C. albicans SC5314, 4395, and 5272 (MIC80 = 1.9, 4.0, and 3.7 μg/mL). The antifungal activities of B14 and D5 against four strains of fluconazole-resistant fungi were better than those of fluconazole. The TAI values of A11 (2.71), B14 (2.13), and D5 (2.25) are the highest. Further exploration of antifungal mechanisms revealed that the fungus treated with compound A11 produced free radicals and reactive oxygen species, and their mycelium morphology was damaged. In conclusion, the N′-phenylhydrazide scaffold showed potential in the development of antifungal lead compounds. Among them, A11, B14, and D5 demonstrated particularly promising antifungal activity and held potential as novel antifungal agents.
... Lifestyle of rural patients is different from that of urban patients, and some believe that rural patients are more at the risk of diabetes and foot ulcers. The lower level of hygiene, more contact with the soil, and the possibility of insect bites and damage during agricultural activities in rural areas are risk factors affecting the incidence of wounds in patients [51,52]. Results of the present study also showed that although the frequency of rural patients with DFUs was not significantly different from that of urban patients, the incidence of CI was 2.13 times higher in rural patients, compared to urban patients. ...
Article
Full-text available
Background and Purpose: In diabetic foot ulcers, if fungal agents, such as Candida species penetrate the cutaneous or depth of the ulcer, it can increase the wound severity and make it more difficult to heal. Materials and Methods: A cross-sectional study was performed on 100 diabetic patients with a foot ulcer from December 2019 to November 2020 in northern Iran. Patient data and wound grades were recorded in a questionnaire. Candida infection was confirmed by direct microscopic examination and culture. To identify the causative agent, polymerase chain reaction-restriction fragment length polymorphism using MspI enzyme and the partial amplification of hyphal wall proteins (HWP1) gene were performed. Results: Mean age of the participants was 62.1 ± 10.8 years old, and 95% of them had type 2 diabetes. Moreover, more than 83% of them had diabetes for a duration of 10 years. In addition, 59% of the patients were male, and 66% > of them had poor education levels. Besides, 99% of them were married, and 52% were rural. Furthermore, 95% of the participants had neuropathic symptoms and 88% used antibiotics. The HbA1C level was > 9% in 69% of them, and the mean ulcer grade of the patients was 2.6±1.05. Candida infection was detected in 13% of the deep tissue and 7% of the tissue surrounding the wound. The predominant Candida isolate was C. parapsilosis (71.5%) and C. albicans (14.3%). Infections caused by filamentous fungi were not detected. There was a statistically significant relationship between Candida infection and gender, rural lifestyle, HbA1C, and ulcer grade. Conclusion: Mycological evaluations of diabetic foot ulcers are often ignored. The present study revealed that C. parapsilosis is the most common causative agent of deep-seated foot ulcer infection in these patients and may require specific treatment. Therefore, more attention of physicians to Candida infections, early diagnosis, and prompt treatment can help accelerate wound healing and prevent amputation.
... Além disso, estudos tem demonstrado a resistência da Candida spp. aos antifúngicos devido ao seu uso frequente e duradouro (MAUBON et al., 2014;BAILLY et al., 2016). Nesse sentido, T. vulgaris L. (tomilho), uma planta aromática e medicinal, tem sido empregada. ...
... Although the treatment of candidiasis with antifungal drugs is highly effective, their use is associated with adverse reactions, such as changes in taste, gastrointestinal and allergic symptoms (Bakhshi et al., 2012). In addition, studies have shown the resistance of Candida spp. to antifungal drugs due to their frequent and long-lasting use (Maubon et al., 2014;Bailly et al., 2016), which is common among denture wearers (Bailly et al., 2016;Lewis & Williams, 2017;Gad & Fouda, 2020). ...
Article
Full-text available
The objective of this study was to provide a review on Candida albicans, focusing on its main virulence factors, pathogenesis, and methods of diagnosis and control of infections caused by this microorganism, known as candidiasis. Its main virulence factors are adhesion, polymorphism and dimorphism, which aid in tissue invasion, phenotypic variability, tolerance to toxins, and the presence of enzymes such as proteases and phospholipases. These factors confer the fungus with the ability to colonize, establish itself, and consequently, cause infections. C. albicans can proliferate on the skin and mucous membranes of the oropharyngeal cavity, gastrointestinal tract, and vaginal tract. The result of this colonization is the formation of white plaques or nodules with erythematous borders in the infected area. Additionally, it may cause pain and burning or be asymptomatic. The diagnosis of candidiasis is based on the symptoms presented by the host. Cultures, histopathological examinations, blood cultures, and serum beta-glucan tests can also be used. The treatment of candidiasis is carried out with antifungals such as nystatin, clotrimazole, fluconazole, itraconazole, and amphotericin B. However, research on medicinal plant products has been conducted to provide an integrative and complementary approach to controlling this pathogen. Thymus vulgaris L. is a good example of this. It is a plant with various phytochemicals and recognized biological activities, including antifungal effects. Thus, this study demonstrated some morphological and pathological characteristics of C. albicans. It was also possible to understand how candidiasis manifests, and how it can be diagnosed and treated both conventionally and integratively.
Article
Full-text available
Introduction Candida albicans (C. albicans) can form biofilms; a critical virulence factor that provides effective protection from commercial antifungals and contributes to public health issues. The development of new antifungal therapies, particularly those targeting biofilms, is imperative. Thus, this study was conducted to investigate the antifungal and antibiofilm effects of Lactobacillus salivarius (L. salivarius), zinc nanoparticles (ZnNPs) and nanocomposites (ZnNCs) on C. albicans isolates from Nile tilapia, fish wash water and human fish sellers in Sharkia Governorate, Egypt. Methods A cross-sectional study collected 300 samples from tilapia, fish wash water, and fish sellers (100 each). Probiotic L. salivarius was immobilized with ZnNPs to synthesize ZnNCs. The study assessed the antifungal and antibiofilm activities of ZnNPs, L. salivarius, and ZnNCs compared to amphotericin (AMB). Results Candida spp. were detected in 38 samples, which included C. albicans (42.1%), C. glabrata (26.3%), C. krusei (21.1%), and C. parapsilosis (10.5%). A total of 62.5% of the isolates were resistant to at least one antifungal agent, with the highest resistance to nystatin (62.5%). However, 75% of the isolates were highly susceptible to AMB. All C. albicans isolates exhibited biofilm-forming capabilities, with 4 (25%) isolates showing strong biofilm formation. At least one virulence-associated gene (RAS1, HWP1, ALS3, or SAP4) was identified among the C. albicans isolates. Probiotics L. salivarius, ZnNPs, and ZnNCs displayed antibiofilm and antifungal effects against C. albicans, with ZnNCs showing significantly higher inhibitory activity. ZnNCs, with a minimum inhibitory concentration (MIC) of 10 µg/mL, completely reduced C. albicans biofilm gene expression. Additionally, scanning electron microscopy images of C. albicans biofilms treated with ZnNCs revealed asymmetric, wrinkled surfaces, cell deformations, and reduced cell numbers. Conclusion This study identified virulent, resistant C. albicans isolates with strong biofilm-forming abilities in tilapia, water, and humans, that pose significant risks to public health and food safety.
Thesis
Full-text available
Abstract This study aimed to investigate the effect of some physiological characteristics on three isolates of pathogenic Candida albicans, and a diagnostic study was conducted for candidiasis isolated from the mouth ,Urinary tract, vagina and skin of patients and healthy Iraqis isolated from 100 clinical samples collected from 50 males and 50 females from patients and healthy people from some Baghdad hospitals. The results showed that 56 isolates were positive for Candida in the mouth, skin, urinary tract and vagina. The highest injuries were recorded in the mouth of the patients, with a percentage of 30.36% and the lowest injuries in the vagina with a rate of 3.57%, while the highest injuries were in the healthy patients with a rate of 35.71%, and the lowest percentage of injuries in the vagina was 1.79%, and no urinary tract infections were recorded. The results of the positive isolates from the mouth revealed that there were 11 positive isolates in the female patients, and the age group was 51-60 years, the highest groups were infected with Candida spp. with 4 infections, and no infections appeared in the 1-10 years. As for healthy women, the number of infections reached 6, and the highest was in the age group 40-31 years, with 3 cases, while no infection appeared in the age groups between 11- 20 years and 60-51years. The isolated infections from the mouth of male patients amounted to 6 infections, and the highest percentage of infection appeared for the age group 31-40 years which numbered 3, while no infection was recorded in the age groups 1-30 years, 14 isolates were obtained from healthy men, and the highest infection was recorded for age groups ranging from 11- 20, 21-30 and 31-40 years as there were 4 injuries for each, and no injuries were recorded for the age group between 41-60 years. Isolated results from the vagina of three women showed that all of them were infected with candidiasis, the two patients numbered in the age groups 30-40 and 41-50 years were infected with candidiasis, as well as the appearance of infection in the only sample taken from healthy people in the age group 30-40 years , among 24 samples. The results of the isolated skin showed 7 injuries for women and 5 for men, there were two injuries in women patients in the age group 51-60 years and one infection in the age group 61-70 years and no injuries were recorded in the rest of the age groups, As for healthy women, 4 injuries were recorded in the age group 1-10, and no infection appeared in the other age groups of the study. While the sick men found one infection in the age group 51-60 years , and no infection was recorded in the other groups. As for healthy men, two infections appeared in the age group 11-20 years and one infection in the age groups 1-10 years and 30-21 years, and no infection appeared in the rest of the age groups. Eight samples were taken from the urine of sick and healthy women, and one positive isolate appeared in sick women in the age group 40-50 years and no infection appeared in healthy women, among the 8 samples collected from sick and healthy men, an infection appeared in the 61-70 years age group, and two positive samples were in the 71-80 years, while no infection appeared in the healthy men. When the samples were cultured on Sabouraud Dextrase Agar medium, the colonies of the Candida spp. appeared in a creamy white color, shiny, smooth to the touch, oval or spherical, and sticky. During the microscopic examination of Candida colonies after adding lactophenol blue dye, the results were oval or spherical cells with blue-green borders as a result of The Gram stain collects on the positive wall due to the presence of a peptidoglycan layer that has the ability to retain this pigment. Candida spp. were diagnosed based on color by growth on Chrom Agar medium, Candida albicans appeared in light green color, C. tropecalis was blue, C. glabrata was pale pink, and C. dublineiensis was dark green. Candida albicans showed growth At a temperature of 45 °C and the formation of the germ tube and no superficial growth was recorded, C. dublienensis showed its ability to form the germ tube. Morphology and microscopic diagnosis of three species of Candida albicans isolated from patients in a hospital Yarmouk in Baghdad, Tikrit Hospital in Salah Al-Din and Al-Kut Hospital in Wasit, with one isolation from each hospital using the Vitek2 Compact System. Wizard Genomic DNA was used, and the results showed DNA amplification of yeast isolates using special equipment for this purpose, agarose gel was used at a concentration of 1.5% to electrophoresis and then detected by using ethidium bromide dye. The three isolates contained one single bundle of extracted DNA. The same site was taken on the support plate in the agarose gel, as the PCR products were shown using the fungal primer pair (ITS1, ITS4). Produce different molecular sizes ranging from 510 -870 bp and the Candida albicans had a molecular size of 535 bp. The physiological study was carried out in two ways, the first is to show the effect of some physiological factors such as temperature, pH concentration and CO2 ratio, the second is to study the effect of some compounds phosphate, biotin and phenolic alcohol on the three isolates of Candida albicans. The effect of temperature recorded different growth on isolates, as the best rate of growth was in temperature 37°C. Whereas isolate K8 recorded the highest growth rate of 75*107 cfu/ml, while the lowest growth rate was at a temperature of 30°C for isolate T5 reaching 1.37*107 cfu/ml and no growth was recorded for the three isolates at 45°C. As for the effect of pH, the results showed that the best growth rate was at pH 6.2, as the highest growth rate in isolate K8 was 74.6 * 107 cfu/ml at pH 6.2, while the lowest growth rate recorded by isolate K8 reached 0. 5 * 107 cfu/ml at pH 9, while no growth occurred for isolates T5 and k8 at pH 4.The results showed that the high percentage of CO2 had a negative effect on the growth rate, as the best rate for the growth of all isolates was at a concentration of 0.03% of carbon dioxide, the highest growth rate in isolate K8 reached 76.7 * 107 cfu/ml, while the lowest growth rate for isolates K8 The isolates were at a 15% concentration of K8 23*107 cfu /ml. The results of adding the phosphate compound showed a difference in the growth rates of the three isolates. K8 recorded the highest level of colony growth reaching 74.8*107 cfu/ml and isolate T5 was the least developed recording 47.5*107 cfu/ml. When adding 10 µm of phosphate compound to the medium, isolate T5 recorded the highest growth rate, reaching 60*107 cfu/ml and the lowest growth rate appeared in isolate K8 which amounted to 3.2*107cfu/ml, while isolate T5 showed the highest growth rate which was 90*107 cfu/ml, The lowest growth rate was in isolate K8, which recorded 7.3*107 cfu/ml. The addition of phenol had a negative effect on the growth of Candida albicans, two different concentrations of phenol were used, which are 10 and 50 µm, the result was no growth of colonies in the three isolates. The results of the study showed that the addition of biotin to the culture media improves the growth of colonies well, as isolate T5 recorded the highest growth rate when adding biotin by 10 µm, reaching 123.1*107 fuc/ml, the least growth was isolate B17 which amounted to 31.4*107 fuc/ml, as well as isolate T5 recorded the highest. The growth rate when adding 50 µm reached 110*107 fuc/ml, while isolate K8 was the least growing reaching 40.8.*107 fuc/ml. The sensitivity and effect of the antifungal agents Amphotericin, Fluconazole, Miconazole, and Nystatin were examined, the results showed that the antifungal Fluconazole was the most inhibiting antifungal of the three isolates, with inhibition diameters ranging between 2.5 -4mm, the least inhibiting antifungal was Nystatin with inhibition diameters ranging from 1-1.2 mm. All results showed that there were significant differences under the probability level of P≤0.05.
Article
Full-text available
The fungal component of the microbiota, the mycobiota, has been neglected for a long time due to its poor richness compared to bacteria. Limitations in fungal detection and taxonomic identification arise from using metagenomic approaches, often borrowed from bacteriome analyses. However, the relatively recent discoveries of the ability of fungi to modulate the host immune response and their involvement in human diseases have made mycobiota a fundamental component of the microbial communities inhabiting the human host, deserving some consideration in host–microbe interaction studies and in metagenomics. Here, we reviewed recent data on the identification of yeasts of the Ascomycota phylum across human body districts, focusing on the most representative genera, that is, Saccharomyces and Candida . Then, we explored the key factors involved in shaping the human mycobiota across the lifespan, ranging from host genetics to environment, diet, and lifestyle habits. Finally, we discussed the strengths and weaknesses of culture‐dependent and independent methods for mycobiota characterization. Overall, there is still room for some improvements, especially regarding fungal‐specific methodological approaches and bioinformatics challenges, which are still critical steps in mycobiota analysis, and to advance our knowledge on the role of the gut mycobiota in human health and disease. This article is categorized under: Immune System Diseases > Genetics/Genomics/Epigenetics Immune System Diseases > Environmental Factors Infectious Diseases > Environmental Factors
Article
Full-text available
Infections caused by Candida species increased, as well as species resistant to the most conventional antimicrobials. Thus, the search for new natural sources of antimicrobials without adverse effects also increased. In this case, Astronium urundeuva Engl., a medicinal plant with several pharmacological activities, including antifungal activity, emerges as a source of antifungal compounds. Here, we evaluated the antifungal activity of A. urundeuva and its derivatives free or loaded into a nanostructured lipid system and combination studies against Candida species. A. urundeuva component's purification and identification were performed through chromatographic and spectroscopic techniques, respectively. The major components identified were gallic acid (1), methyl gallate (2), ethyl gallate (3), and 1,2,3,4,6-penta-O-galoyl-β-D-glucose (PGG) (4). A. urundeuva extract, its aqueous fraction, and PGG (4) exhibited the same MIC values between 0.2 and 62.5 µg/mL against Candida species being classified as significant activity, while the amphotericin B MIC was 0.1–1.2 µg/mL. Combination assay demonstrated that may occur synergism (0.37) between gallic acid (1) x methyl gallate (2). Thus, the results demonstrated as PGG (4) much as the gallic acid (1) x methyl gallate (2) combined might be the main responsible for the antifungal activity against Candida species, mainly Candida glabrata ATCC 2001 and that the tested nanostructured system maintained the antifungal activity. Through the antifungal activity of the extract and purified compounds, the extract may be employed in the development of antifungal herbal medicine, because its purification process is simple and greener than pure compounds.
Article
Full-text available
By CLSI interpretive criteria, anidulafungin and micafungin MICs determined by various methods were sensitive (60 to 70%) and highly specific (94 to 100%) for identifying FKS mutations among 120 Candida glabrata isolates. Anidulafungin and micafungin breakpoints were more specific than CLSI's caspofungin breakpoint in identifying FKS mutant strains and patients with invasive candidiasis who were likely to fail echinocandin treatment (P ≤ 0.0001 for both). Echinocandin MICs were most useful clinically when interpreted in the context of prior echinocandin exposure.
Article
Full-text available
Background: Echinocandins are recommended for Candia glabrata candidemia. Mutations in the FKS1 and FKS2 genes are associated with echinocandin resistance. Few studies have assessed risk factors for FKS mutant isolates and outcomes in patients receiving micafungin treatment. Methods: Patients with C. glabrata bloodstream infection admitted to a large, tertiary care hospital between 2009 and 2012 were included in this study. For each isolate, FKS1 and FKS2 genes were sequenced to identify mutations. Risk factors for FKS mutations and treatment outcomes in patients receiving an echinocandin were assessed using multivariate logistic regression. Results: Seventy-two patients were included in the study of which 13 (18%) had an FKS mutant isolate. The only significant predictor for FKS mutations was prior echinocandin exposure (odds ratio [OR], 19.9; 95% confidence interval [CI], 4.7-84.7; P ≤ .01). Treatment failure occurred in 17 (30%) of 57 patients who received an echinocandin and was more common in patients with FKS mutants (6 of 10; 60%) compared with non-FKS mutants (11 of 47; 23%). Underlying gastrointestinal disorder (OR, 4.7; 95% CI, 1.1-20.9; P = .04) and prior echinocandin exposure (OR, 8.3; 95% CI, 1.7-40.4; P ≤ .01) were independent predictors of echinocandin treatment failure. Treatment response and echinocandin minimum inhibitory concentrations varied among specific FKS mutations. Conclusions: FKS mutations were identified in 18% of 72 patients with C. glabrata candidemia. Common risk factors for FKS mutant isolates included previous echinocandin exposure, which also influenced response rates.
Article
Full-text available
Candida is the most common cause of severe yeast infections worldwide, especially in critically ill patients. In this setting, septic shock attributable to Candida is characterized by high mortality rates. The aim of this multicenter study was to investigate the determinants of outcome in critically ill patients with septic shock due to candidemia. This was a retrospective study in which patients with septic shock attributable to Candida who were treated during the 3-year study period at one or more of the five participating teaching hospitals in Italy and Spain were eligible for enrolment. Patient characteristics, infection-related variables, and therapy-related features were reviewed. Multiple logistic regression analysis was performed to identify the risk factors significantly associated with 30-day mortality. A total of 216 patients (mean age 63.4 ± 18.5 years; 58.3 % males) were included in the study. Of these, 163 (75 %) were admitted to the intensive care unit. Overall 30-day mortality was 54 %. Significantly higher Acute Physiology and Chronic Health Evaluation (APACHE) II scores, dysfunctional organs, and inadequate antifungal therapy were compared in nonsurvivors and survivors. No differences in survivors versus nonsurvivors were found in terms of the time from positive blood culture to initiation of adequate antifungal therapy. Multivariate logistic regression identified inadequate source control, inadequate antifungal therapy, and 1-point increments in the APACHE II score as independent variables associated with a higher 30-day mortality rate.
Article
Full-text available
The objectives of this study were to determine species distribution and in vitro antifungal susceptibility of Candida isolates identified in the multicentre China-SCAN study of invasive Candida infection (ICI) in intensive care units (ICUs) across China. Candida isolates from patients in the China-SCAN study with documented ICI were evaluated by a central laboratory. Species were identified using chromogenic culture media or the API 20C AUX kit. Susceptibility to fluconazole, voriconazole, itraconazole, caspofungin and amphotericin B was determined using the CLSI broth microdilution method (M27-A3) and updated clinical breakpoints or epidemiological cut-off values. A total of 389 isolates from 244 patients were analysed. Species identified most frequently were Candida albicans (40.1%), Candida parapsilosis (21.3%), Candida tropicalis (17.2%) and Candida glabrata (12.9%). Rarer species such as Lodderomyces elongisporus and Candida ernobii were also identified. Fluconazole susceptibility was evident in 85.9% (134/156) of C. albicans, 62.7% (42/67) of C. tropicalis and 48.2% (40/83) of C. parapsilosis isolates. Susceptibility to voriconazole was ≥90% among all species. All isolates were susceptible to amphotericin B and caspofungin except C. glabrata [86.0% (43/50) susceptible to caspofungin]. Cross-resistance between fluconazole and voriconazole was observed for C. parapsilosis and C. glabrata. Although C. albicans was the predominant single species, non-albicans species constituted >50% of isolates. Fluconazole susceptibility was lower in most non-albicans species, indicating that fluconazole resistance should be closely monitored. Susceptibility to voriconazole, amphotericin B and caspofungin is encouraging. Differences between these data and those from other regions emphasize the importance of assessing regional variations.
Article
Full-text available
Candidemia is the fourth most common kind of microbial bloodstream infection, with Candida albicans being the most common causative species. Echinocandins are employed as the first-line treatment for invasive candidiasis until the fungal species is determined and confirmed by clinical diagnosis. Echinocandins block the FKS glucan synthases responsible for embedding β-(1,3)-d-glucan in the cell wall. The increasing use of these drugs has led to the emergence of antifungal resistance, and elevated MICs have been associated with single-residue substitutions in specific hot spot regions of FKS1 and FKS2. Here, we show for the first time the caspofungin-mediated in vivo selection of a double mutation within one allele of the FKS1 hot spot 1 in a clinical isolate. We created a set of isogenic mutants and used a hematogenous murine model to evaluate the in vivo outcomes of echinocandin treatment. Heterozygous and homozygous double mutations significantly enhance the in vivo resistance of C. albicans compared with the resistance seen with heterozygous single mutations. The various FKS1 hot spot mutations differ in the degree of their MIC increase, substance-dependent in vivo response, and impact on virulence. Our results demonstrate that echinocandin EUCAST breakpoint definitions correlate with the in vivo response when a standard dosing regimen is used but cannot predict the in vivo response after a dose escalation. Moreover, patients colonized by a C. albicans strain with multiple mutations in FKS1 have a higher risk for therapeutic failure.
Article
Invasive Candida infections are important causes of morbidity and mortality in immunocompromized and hospitalized patients. This article provides the joint recommendations of the German-speaking Mycological Society (Deutschsprachige Mykologische Gesellschaft; DMyKG) and the Paul-Ehrlich-Society for Chemotherapy Section antifungal therapy (PEG-SAC) for diagnosis and treatment of invasive and superficial Candida infections. The recommendations are based on published results of clinical trials, case-series and expert opinion using the evidence criteria set forth by the Infectious Diseases Society of America (IDSA). Key recommendations are summarized here: The cornerstone of diagnosis remains the detection of the organism by culture with identification of the isolate at the species level; in vitro susceptibility testing is mandatory for invasive isolates. Options for initial therapy of candidemia and other invasive Candida infections in non-granulocytopenic patients include fluconazole or one of the three approved echinocandin compounds; liposomal amphotericin B and voriconazole are secondary alternatives because of their less favorable pharmacological properties. In granulocytopenic patients, an echinocandin or liposomal amphotericin B is recommended as initial therapy based on the fungicidal mode of action. Indwelling central venous catheters serve as a main source of infection independent of the pathogenesis of candidemia in the individual patients and should be removed whenever feasible. Preexisting immunosuppressive treatment, particularly by glucocorticosteroids, ought to be discontinued, if feasible, or reduced. The duration of treatment for uncomplicated candidemia is 14 days following the first negative blood culture and resolution of all associated symptoms and findings. Ophthalmoscopy is recommended prior to the discontinuation of antifungal chemotherapy to rule out endophthalmitis or chorioretinitis. Beyond these key recommendations, this article provides detailed recommendations for specific disease entities, for antifungal treatment in pediatric patients as well as a comprehensive discussion of epidemiology, clinical presentation and emerging diagnostic options of invasive and superficial Candida infections.
Article
Clin Microbiol Infect 2012; 18 (Suppl. 7): 19–37 This part of the EFISG guidelines focuses on non-neutropenic adult patients. Only a few of the numerous recommendations can be summarized in the abstract. Prophylactic usage of fluconazole is supported in patients with recent abdominal surgery and recurrent gastrointestinal perforations or anastomotic leakages. Candida isolation from respiratory secretions alone should never prompt treatment. For the targeted initial treatment of candidaemia, echinocandins are strongly recommended while liposomal amphotericin B and voriconazole are supported with moderate, and fluconazole with marginal strength. Treatment duration for candidaemia should be a minimum of 14 days after the end of candidaemia, which can be determined by one blood culture per day until negativity. Switching to oral treatment after 10 days of intravenous therapy has been safe in stable patients with susceptible Candida species. In candidaemia, removal of indwelling catheters is strongly recommended. If catheters cannot be removed, lipid-based amphotericin B or echinocandins should be preferred over azoles. Transoesophageal echocardiography and fundoscopy should be performed to detect organ involvement. Native valve endocarditis requires surgery within a week, while in prosthetic valve endocarditis, earlier surgery may be beneficial. The antifungal regimen of choice is liposomal amphotericin B +/− flucytosine. In ocular candidiasis, liposomal amphotericin B +/− flucytosine is recommended when the susceptibility of the isolate is unknown, and in susceptible isolates, fluconazole and voriconazole are alternatives. Amphotericin B deoxycholate is not recommended for any indication due to severe side effects.
Article
To determine the epidemiology of Candida bloodstream infections, variables influencing mortality, and antifungal resistance rates in ICUs in Spain. Prospective, observational, multicenter population-based study. Medical and surgical ICUs in 29 hospitals distributed throughout five metropolitan areas of Spain. Adult patients (≥ 18 yr) with an episode of Candida bloodstream infection during admission to any surveillance area ICU from May 2010 to April 2011. Candida isolates were sent to a reference laboratory for species identification by DNA sequencing and susceptibility testing using the methods and breakpoint criteria promulgated by the European Committee on Antimicrobial Susceptibility Testing. Prognostic factors associated with early (0-7 d) and late (8-30 d) mortality were analyzed using logistic regression modeling. We detected 773 cases of candidemia, 752 of which were included in the overall cohort. Among these, 168 (22.3%) occurred in adult ICU patients. The rank order of Candida isolates was as follows: Candida albicans (52%), Candida parapsilosis (23.7%), Candida glabrata (12.7%), Candida tropicalis (5.8%), Candida krusei (4%), and others (1.8%). Overall susceptibility to fluconazole was 79.2%. Cumulative mortality at 7 and 30 days after the first episode of candidemia was 16.5% and 47%, respectively. Multivariate analysis showed that early appropriate antifungal treatment and catheter removal (odds ratio, 0.27; 95% CI, 0.08-0.91), Acute Physiology and Chronic Health Evaluation II score (odds ratio, 1.11; 95% CI, 1.04-1.19), and abdominal source (odds ratio, 8.15; 95% CI, 1.75-37.93) were independently associated with early mortality. Determinants of late mortality were age (odds ratio, 1.04; 95% CI, 1.01-1.07), intubation (odds ratio, 7.24; 95% CI, 2.24-23.40), renal replacement therapy (odds ratio, 6.12; 95% CI, 2.24-16.73), and primary source (odds ratio, 2.51; 95% CI, 1.06-5.95). Candidemia in ICU patients is caused by non-albicans species in 48% of cases, C. parapsilosis being the most common among these. Overall mortality remains high and mainly related with host factors. Prompt adequate antifungal treatment and catheter removal could be critical to decrease early mortality.