The ATP Binding Cassette Transporter Gene CgCDR1 from C.glabrata.pdf

(2231 KB) Pobierz
ac119902753p
A NTIMICROBIAL A GENTS AND C HEMOTHERAPY ,
0066-4804/99/$04.00
Nov. 1999, p. 2753–2765
Vol. 43, No. 11
1
0
Copyright © 1999, American Society for Microbiology. All Rights Reserved.
The ATP Binding Cassette Transporter Gene CgCDR1 from
Candida glabrata Is Involved in the Resistance of Clinical
Isolates to Azole Antifungal Agents
DOMINIQUE SANGLARD, 1 * FRAN ¸ OISE ISCHER, 1
DAVID CALABRESE, 1
PAUL A. MAJCHERCZYK, 2
AND JACQUES BILLE 1
Institut de Microbiologie 1
and Division of Infectious Diseases, 2 Centre Hospitalier Universitaire Vaudois (CHUV),
1011 Lausanne, Switzerland
Received 17 June 1999/Returned for modification 13 August 1999/Accepted 7 September 1999
The resistance mechanisms to azole antifungal agents were investigated in this study with two pairs of
Candida glabrata clinical isolates recovered from two separate AIDS patients. The two pairs each contained a
fluconazole-susceptible isolate and a fluconazole-resistant isolate, the latter with cross-resistance to itracon-
azole and ketoconazole. Since the accumulation of fluconazole and of another unrelated substance, rhodamine
6G, was reduced in the azole-resistant isolates, enhanced drug efflux was considered as a possible resistance
mechanism. The expression of multidrug efflux transporter genes was therefore examined in the azole-
susceptible and azole-resistant yeast isolates. For this purpose, C. glabrata genes conferring resistance to azole
antifungals were cloned in a Saccharomyces cerevisiae strain in which the ATP binding cassette (ABC) trans-
porter gene PDR5 was deleted. Three different genes were recovered, and among them, only C. glabrata CDR1
( CgCDR1 ), a gene similar to the Candida albicans ABC transporter CDR genes, was upregulated by a factor of
5 to 8 in the azole-resistant isolates. A correlation between upregulation of this gene and azole resistance was
thus established. The deletion of CgCDR1 in an azole-resistant C. glabrata clinical isolate rendered the resulting
mutant (DSY1041) susceptible to azole derivatives as the azole-susceptible clinical parent, thus providing
genetic evidence that a specific mechanism was involved in the azole resistance of a clinical isolate. When
CgCDR1 obtained from an azole-susceptible isolate was reintroduced with the help of a centromeric vector in
DSY1041, azole resistance was restored and thus suggested that a trans -acting mutation(s) could be made
responsible for the increased expression of this ABC transporter gene in the azole-resistant strain. This study
demonstrates for the first time the determinant role of an ABC transporter gene in the acquisition of resistance
to azole antifungals by C. glabrata clinical isolates.
Patients with advanced human immunodeficiency virus in-
fection develop opportunistic infections due to the decrease in
their immunity. Oropharyngeal candidiasis (OPC) caused by
Candida albicans is a very common opportunistic infection in
these patients and is treated mainly with azole antifungal
agents, particularly with fluconazole. Treatment failures have
been observed following the repeated use of this agent in
relapses of OPC (21, 37, 45, 55). Different laboratories have
reported that C. albicans isolates sampled sequentially during
fluconazole treatment showed decreased susceptibility to flu-
conazole compared to that of the isolates sampled at the time
of the first episode of infection (6, 30, 41, 53). Clinical resistance
to fluconazole has been correlated with in vitro resistance of
the yeasts recovered from patients undergoing antifungal ther-
apy (54). This phenomenon has been also documented in other
yeast species, including Candida glabrata (4, 16), Candida tropi-
calis (28), and Candida krusei (57, 58), and in Cryptococcus
neoformans (27).
The increasing number of azole-resistant isolates recovered
in many institutions during the past decade has motivated
studies with the aim of understanding their mechanisms of
resistance at the molecular level. Until now, C. albicans isolates
have provided a major source for the discovery of mechanisms
of azole resistance. Recent findings have shown that increased
azole efflux is an important mechanism of resistance in yeast
clinical isolates. In azole-resistant C. albicans isolates, in-
creased azole efflux has been correlated with the upregulation
of multidrug efflux transporter genes from two distinct families,
the ATP binding cassette (ABC) transporters ( CDR1 and
CDR2 ) and the major facilitators ( C. albicans MDR1 ) (3, 27,
30, 50, 53, 60). Deletion of CDR1 in C. albicans leads to azole
hypersusceptibility and increased fluconazole accumulation
(51). Decrease in azole affinity of the target enzyme of these
antifungals, i.e., the cytochrome P450 lanosterol demethylase
(called CYP51A1, or ERG11), has also been explained at the
molecular level. Mutations in the genes encoding CYP51A1
( CYP51A1 ) have been detected in azole-resistant yeasts. These
mutations resulted, in some cases, in amino acid substitutions
with the probable effect of altering the binding properties of
azoles and thus contributing to a decrease in azole susceptibil-
ity in clinical yeast isolates (49). Another mechanism of azole
resistance originates from alterations in the ergosterol biosyn-
thesis pathway, often resulting in the absence of ergosterol.
This feature renders cells affected by this mechanism cross-
resistant to amphotericin B. A few C. albicans clinical isolates
possess this property and have been found to accumulate 14
a
-
methylergosta-8,24(28)-dien-3
b
,6
a
-diol, which is indicative of
5,6 sterol desaturase (24, 35). The above-men-
tioned resistance mechanisms can combine with each other in
C. albicans and complicate the analysis of such isolates, since it
is difficult to establish the role of an individual mechanism in
the decrease in azole susceptibility (48). Dissection of resis-
tance mechanisms by the use of genetics in C. albicans clinical
D
* Corresponding author. Mailing address: Institute of Microbiology,
University Hospital Lausanne (CHUV), Rue de Bugnon 44, CH-1011
Lausanne, Switzerland. Phone: 0041 21 3144083. Fax: 0041 21 3144060.
E-mail: Dominique.Sanglard@chuv.hospvd.ch.
2753
a defect in
690426656.003.png
 
2754
SANGLARD ET AL.
A NTIMICROB .A GENTS C HEMOTHER .
TABLE 1. Yeast strains used in this study
Strain
Genotype
Reference
S. cerevisiae
YKKB-13
MAT a ura3-52 lys2-801 (Am) ade2-101 (Oc) trp1-
D
63 his3-
D
200 leu2-
D
1
D
pdr5::TRP1
8
C. glabrata
ATCC 90030
ATCC reference strain
DSY528
Clinical isolate, azole susceptible from patient 1
This study
DSY530
Clinical isolate, azole resistant from patient 1
This study
DSY562
Clinical isolate, azole susceptible from patient 2
This study
DSY565
Clinical isolate, azole resistant from patient 2
This study
DSY1029
ura3 derivative of DSY565
This study
DSY1041
D
cgcdr1::hisG-URA3-hisG , derived from DSY1029
This study
DSY1056
D
cgcdr1::hisG , derived from DSY1041
This study
DSY671
ura3
61
DSY1033
D
cgcdr1::hisG-URA3-hisG , derived from DSY671
This study
DSY1067
D
cgcdr1::hisG , derived from DSY1033
This study
DSY1717
DSY1041 transformed with pDS670
This study
DSY1718
DSY1033 transformed with pDS670
This study
isolates resistant to azole antifungal agents has not been re-
ported yet. This lack of important information arises from the
difficulties of developing reliable genetic systems for this dip-
loid organism.
Historically, C. albicans accounted for 70 to 80% of organ-
isms isolated in patients infected by fungal species. However,
recent data report a population shift toward non- C. albicans
yeast species, such as C. glabrata , C. tropicalis ,or C. krusei (15).
Among the non- C. albicans species, C. glabrata has emerged as
an important nosocomial pathogen. Berrouane et al. (7) re-
ported that among Candida species, the proportion of C. gla-
brata infections in the Iowa University Hospitals from 1988 to
1994 increased significantly, while it remained unchanged for
other yeast species and even decreased for C. albicans . Other
investigators have noted similar increases in the frequency of
infections caused by C. glabrata , mostly in conjunction with the
use of azoles (34, 39, 40). We also observed that C. glabrata was
often recovered from cultures originating from AIDS patients
with OPC. C. glabrata is known to be less susceptible to flu-
conazole than most of the C. albicans fluconazole-susceptible
isolates. Rex et al. (43) reported that the minimal inhibitory
concentration inhibiting 50% of the yeast population investi-
gated (MIC 50 ) of fluconazole for 31 C. glabrata isolates was 16
m
MATERIALS AND METHODS
Strains and media. The yeast strains used in this study are listed in Table 1.
They were grown at 30°C on yeast extract-peptone-dextrose (YEPD) complex
medium containing 2% glucose, 1% Bacto peptone (Difco Laboratories, Detroit,
Mich.), and 0.5% yeast extract (Difco). YEPD agar plates contained 2% agar
(Difco) as a supplement. Yeast nitrogen base (YNB [Difco]) with 2% glucose
and 2% agar (Difco) with appropriate amino acids and bases was used as a
selective medium after transformation of S. cerevisiae YKKB-13 and C. glabrata .
Agar plates containing 50
m
(19) was used as a host for plasmid
constructions and propagation and was grown on standard media.
Fluconazole and rhodamine 6G accumulation. Fluconazole accumulation test-
ing was performed in duplicate with 3 H-labelled fluconazole (Amersham Life
Science, Little Chalfont, Buckinghamshire, United Kingdom) as described pre-
viously (53), but with a single incubation time of 20 min. Rhodamine 6G (Sigma,
Fluka Chemie AG, Buchs, Switzerland) accumulation testing was performed
with yeast cells grown to the logarithmic phase in 14 ml of sterile polystyrene
tubes with 2 ml of YEPD at 30°C under constant agitation. Rhodamine 6G
labelling of cells was performed in 1 ml of YEPD with 10 7
a
cells and containing
M rhodamine 6G. The mixture was incubated for 30 min at 30°C, after which
it was stopped by placing the tubes on ice. These conditions have been optimized
for minimal incubation time and maximal rhodamine 6G accumulation (data not
shown). The reaction mixture was then diluted 40-fold in cold sterile 0.1 M
phosphate-buffered saline (PBS) at pH 7.0 and then directly subjected to flow
cytometry in a FACScan fluorescence-activated cell sorter (FACS) (Becton
Dickinson, San Jose, Calif.). Fluorescence was measured at an excitation wave-
length of 488 nm and an emission wavelength of 515 nm (F1 detector). The
sheath fluid was Isotone II. Data were acquired for 1,500 cells with the FACScan
Lysis II software. Rhodamine 6G efflux was determined with 10 7 cells previously
loaded by incubation with 10
m
g/ml for 129 C. albicans
isolates. In several patients who responded poorly to flucon-
azole therapy, we noticed that C. glabrata isolates could persist
and that their susceptibility to azoles was decreased. Since
mechanisms of resistance to azoles have been less intensively
investigated in non- C. albicans species such as C. glabrata ,we
addressed here the molecular basis of resistance in two pairs of
isolates taken from two different AIDS patients with docu-
mented azole antifungal treatment failure. We first isolated C.
glabrata azole resistance genes by complementation of hyper-
susceptibility of a Saccharomyces cerevisiae ABC transporter
mutant. From the three different azole resistance genes iso-
lated, only CgCDR1 , which resembles the C. albicans ABC
transporter CDR genes, was upregulated in the azole-resistant
C. glabrata isolates from these two patients. By introducing a
genetic marker in an azole-resistant clinical isolate, not only
could evidence for the participation of this gene in azole re-
sistance be obtained, but the nature of the mutation or muta-
tions implicated in the upregulation of CgCDR1 could be pre-
dicted.
m
M rhodamine 6G at 30°C in YEPD. Cells were
washed three times with YEPD medium at 4°C to remove excess rhodamine 6G,
and efflux was started by incubation at 30°C in the same medium. The decrease
in fluorescence of loaded cells was then recorded at regular time intervals.
Drug susceptibility tests. Tests of susceptibility to azole antifungals were
performed by broth microdilution assay according to the National Committee for
Clinical Laboratory Standards (NCCLS) protocol M27-A (33) with RPMI-1640
medium (Difco) and incubation at 35°C for 48 h. Endpoint readings were re-
corded with a microplate reader (Bio-Rad, Hercules, Calif.), and the azole
concentration yielding at least 50% growth inhibition compared to the growth in
drug-free medium was defined as the MIC. Amphotericin B susceptibility was
measured according to growth in Antibiotic Medium 3 broth (Difco) as recom-
mended previously (42).
Susceptibility to different compounds of the C. glabrata isolates and of S.
cerevisiae strains containing C. glabrata drug resistance genes was also tested
qualitatively by spotting serial dilutions of yeast cultures onto complex YEPD
medium agar plates. This provides an easy visualization of growth differences
between different yeast strains. Since S. cerevisiae does not grow well in the
RPMI medium described in the NCCLS protocol M27-A, the use of the quali-
tative plate assay for drug susceptibility is more adequate. The following drugs
were solubilized in dimethyl sulfoxide: ketoconazole and itraconazole (Janssen
Pharmaceuticals, Beerse, Belgium), 4-nitroquinoline- N -oxide (Sigma), and
benomyl (Riedel-de-Ha¨n, Seelze, Germany). Fluconazole (Pfizer UK, Sand-
m
g of 5-fluoroorotic acid (5-FOA) per ml were made
for the introduction of the ura3 genetic marker in YNB selective medium with 50
m
g of uridine per ml. Escherichia coli DH5
10
g/ml, while the MIC 50 was 0.25
690426656.004.png
V OL . 43, 1999
AZOLE RESISTANCE IN C. GLABRATA
2755
TABLE 2. Antifungal drug susceptibility of C. glabrata isolates taken from two AIDS patients with OPC
MIC (
g/ml)
Time elapsed between
samplings (days)
Isolate
Fluconazole
Ketoconazole
Itraconazole
Amphotericin B
DSY528
4
0.062
0.125
0.5
DSY530
32
0.5
1
0.5
47
DSY562
4
0.031
0.125
0.5
DSY565
128
2
4
0.5
50
wich, United Kingdom), cycloheximide, fluphenazine, and crystal violet (Sigma)
were dissolved in water. Each plate contained 15 ml of agar. The drugs were
diluted in the corresponding solvents to achieve the concentrations used in
YEPD plates. Preliminary tests were performed to optimize drug concentrations
in YEPD plates so that growth differences between the different S. cerevisiae and
C. glabrata strains used in this study could be observed. To perform the suscep-
tibility tests, yeasts were grown overnight at 30°C with constant shaking in YEPD
liquid medium. The cultures were diluted to 2
80°C.
Typing of C. glabrata clinical isolates. Two different methods were utilized for
the typing analysis of the C. glabrata clinical isolates. The first method was based
on restriction fragment length polymorphism (RFLP) with Hin fI as restriction
enzyme as described by others (5, 11). The second method was based on the use
of repetitive probes Cg6 and Cg12 as described previously (29). In this method,
5
2
10 7 cells per ml in 0.9% NaCl.
Five microliters of this suspension and of serial dilutions of each yeast culture
was spotted onto each type of plate and incubated for 48 h at 30°C.
Ergosterol biosynthesis inhibition. Ergosterol biosynthesis inhibition assays
were performed with cellular extracts from C. glabrata strains. Cellular extracts
were from cells grown in 50 ml of YEPD medium and were obtained by me-
chanical disruption with glass beads (0.3 to 0.5 mm in diameter) in phosphate
buffer (0.1 M sodium phosphate [pH 7.5]). The extracts were centrifuged at
10,000
3
gof C. glabrata genomic DNA, prepared as described above, was cut with
Eco RI. The digested DNA was electrophoresed in a 0.7% agarose gel. Transfer
was made by vacuum blotting on GeneScreen Plus membranes (DuPont NEN,
Boston, Mass.) after depurination of the DNA with HCl. The Cg6 and Cg12
repetitive probes were prepared from
m
l
phage DNA by liquid lysate with the
kit (Qiagen, Chatsworth, Calif.). The probes were labelled with
[ 32 P]dATP by nick translation with a nick translation system (Gibco BRL, Life
Technologies, Inc., Rockville, Md.). After prehybridization for 30 min, mem-
branes were hybridized overnight at 65°C in 5
l
g for 10 min at 4°C. The assays were performed with increasing
fluconazole concentrations by the method described by Vanden Bossche et al.
(56). Each assay mixture contained 5 mg of cellular proteins from the individual
strain which was measured by the Bradford assay (Bio-Rad, Hercules, Calif.).
Isolation of DNA and RNA. The small-scale isolation of DNA and RNA from
C. glabrata was performed from cultures grown to the logarithmic growth phase
in YEPD medium at 30°C under constant shaking. One milliliter of each culture
was centrifuged in Eppendorf tubes at 4°C. After a washing step with TE (10 mM
Tris-HCl [pH 7.5], 1 mM EDTA), yeast DNA was extracted by adding 0.3 g of
glass beads (0.3 to 0.5 mm in diameter), 200
SSPE is 0.15 M NaCl
with 10 mM NaH 2 PO 4 and 1 mM EDTA) containing 5% dextran sulfate and
0.3% SDS. The membranes were washed four times for 30 min each at 48°C in
2
3
SSPE (1
3
SSPE containing 0.2% SDS. Fuji RX film was used for visualization of
hybridization patterns.
Yeast transformations. S. cerevisiae and C. glabrata were transformed by the
lithium acetate procedure as described by Sanglard et al. (51).
Cloning of C. glabrata azole resistance genes. The C. glabrata gene library was
constructed in the plasmid YEp24 (GenBank accession no. L09156). C. glabrata
genomic DNA from strain ATCC 90030 was partially digested with Sau 3A to
obtain fragment lengths ranging from 6 to 9 kb. The fragments were purified by
agarose gel electrophoresis and ligated to YEp24 previously digested with
Bam HI. The gene library was amplified in E. coli DH5
3
l of a breaking buffer (2% Triton
X-100, 1% sodium dodecyl sulfate [SDS], 10 mM Tris-HCl [pH 8.0], 1 mM
EDTA, 100 mM NaCl), and 200
m
l of phenol-chloroform-isoamyl alcohol (24:
24:1). After 1 min of vortexing in a Mini-Beadbeater (Biospec Products, Inc.,
Bartlesville, Okla.), the tubes were centrifuged at maximum speed for 10 min in
a microcentrifuge, and the supernatant was reextracted with chloroform-isoamyl
alcohol (24:1). Nucleic acids were then precipitated with 20
m
. C. glabrata genes
conferring resistance to azole antifungals were cloned by complementation of
hypersusceptibility to fluconazole of the S. cerevisiae strain YKKB-13. First,
about 80,000 Ura 1 clones were selected after transformation of YKKB-13 with
the C. glabrata genomic library. The Ura 1 clones were then pooled in separate
aliquots, part of which were spotted onto YEPD medium containing 5 and 10
a
m
l of 3 M sodium
acetate (pH 5.0) and 400
m
l of ethanol at
2
20°C, washed with 70% ethanol, and
l of TE. For the extraction of RNA, the yeast cell pellet was
mixed with 0.3 g of glass beads, 300
m
m
g
l of RNA extraction buffer (0.1 M Tris-HCl
[pH 7.5], 0.1 M LiCl, 10 mM EDTA, 0.5% SDS) and 300
m
80°C.
Plasmid rescue from S. cerevisiae. Episomal plasmids from the parent vector
YEp24 were rescued from S. cerevisiae transformants by electroporation in E.
coli . Yeast cells were grown in selective YNB medium to late log phase, and total
DNA was extracted as outlined above. One microliter of the DNA suspension
(50
2
l of phenol-chloro-
form-isoamyl alcohol (24:24:1). After 1 min of vortexing in the Mini-Beadbeater,
the aqueous phase was reextracted with phenol-chloroform-isoamyl alcohol, and
RNA was precipitated with 600
m
m
l of ethanol at
2
20°C for 1 h. The RNA pellet
was resuspended in 50
m
l of diethyl pyrocarbonate-treated H 2 O, and the con-
, and ampicillin-resistant clones
from each transformant were analyzed by restriction enzyme analysis.
Construction of C. glabrata vectors. To enable the replication of YEp24-
derived plasmids in C. glabrata and particularly of pNB126, the CgCEN and
CgARS sequences contained in pCgACU-5 were subcloned in these vectors.
pCgACU-5 (a kind gift of K. Kitada) was generated by inserting CgURA3 into
the Xho I site of pCgAC-5 (25). The presence of the CgCEN and CgARS se-
quences in plasmids contributes to their stable replication in C. glabrata in one
copy per cell. The CgCEN and CgARS sequences were recovered from
pCgACU-5 by digestion with Sal I and Xho I and subcloned into the single Sal I site
m
l) was electroporated into E. coli DH5
a
TABLE 3. Fluconazole accumulation and inhibition of ergosterol
biosynthesis by fluconazole in C. glabrata isolates
Isolate
Amt of
[ 3 H]fluconazole/2
3
IC 50 of fluconazole for ergosterol
biosynthesis (nM)
10 7
cells (cpm)
DSY528
300
6
20
55
6
10
DSY530
95
6
20
65
6
8
FIG. 1. Typing of C. glabrata clinical isolates used in this study. (A) Restric-
tion enzyme analysis of C. glabrata genomic DNA digested with Hin fI. Restric-
tion fragments were separated by 1% agarose gel electrophoresis and stained
with ethidium bromide. (B) Profiles of band patterns revealed by hybridization
with the repetitive element probes Cg6 and Cg12 as described by Lockhart et al.
(29). Molecular size standards were depicted on each photograph.
DSY562
313
6
22
40
6
12
DSY565
81
6
5
50
6
8
DSY1041
457
6
7
ND a
a ND, not determined.
m
centration was measured spectrophotometrically at A 260 and A 280 . RNA was
stored at
Qiagen
3
resuspended in 50
of fluconazole per ml, while the remaining cells were kept frozen at
690426656.005.png
2756
SANGLARD ET AL.
A NTIMICROB .A GENTS C HEMOTHER .
DNA sequencing. Sequence data were obtained with DNA fragments sub-
cloned from pNB124, pNB125, and pNB126 into pBluescript (Stratagene GmbH,
Z¨rich, Switzerland). Sequence data were generated on both DNA strands by
using reverse or universal primers and customized primers by automated se-
quencing in a Li-Cor 4200 sequencer (Li-Cor, Inc., Lincoln, Nebr.).
FIG. 2. Rhodamine 6G accumulation in C. glabrata clinical isolates. Cells
were labelled with rhodamine 6G and analyzed by FACS as indicated in Mate-
rials and Methods. FACS histograms are given for the unlabelled control
(DSY562) and for each yeast strain labelled with rhodamine 6G. The mean
fluorescence values for DSY562 and DSY565 were 691 and 113, respectively, and
those for DSY528 and DSY530 were 566 and 110, respectively. Accumulation
experiments with rhodamine 6G were repeated five times with these strains.
DSY565 and DSY530 mean fluorescence reached 16%
6
1.8% and 12%
6
3%
RESULTS
Origin and azole susceptibility of C. glabrata isolates. For
the study of mechanisms of resistance to azoles, C. glabrata
clinical isolates were selected retrospectively from a collection
of yeasts recovered from AIDS patients with OPC. The C.
glabrata strains were chosen from two different patients with
recurrent OPC and were first selected on the basis of de-
creased susceptibility to azole antifungals.
Patient 1 had his first OPC episode in October 1990 and was
diagnosed with AIDS. C. albicans was isolated initially from
the patient’s oral cavity in December 1993 and was subse-
quently found in recurrent episodes of OPC. In April 1994, the
first C. glabrata strain (DSY528) was isolated. At that time,
patient 1 had received a cumulative dose of 17.3 g of flucon-
azole, and his CD4 count in blood was 4 cells per mm 3 . After
treatment with 400 mg of fluconazole for 7 consecutive days,
OPC was still persistent 47 days later. DSY530 was isolated at
this time and showed a reduced susceptibility to fluconazole
compared to DSY528 (Table 2). Patient 2 had his first clinically
documented episode of OPC in November 1991 and was di-
agnosed with AIDS. C. albicans was first isolated from the oral
cavity in February 1993. In May 1995, C. glabrata DSY562 was
isolated for the first time together with C. albicans . The patient
had received a cumulative dose of 4.1 g of fluconazole, and his
CD4 count in blood was 34 cells per mm 3 . After two courses of
treatment with 200 mg of fluconazole for 7 consecutive days
each, OPC was still persistent 50 days later, and the C. glabrata
strain DSY565 was isolated, which showed reduced suscepti-
bility to fluconazole compared to that of DSY562 (Table 2). In
both patients, C. glabrata isolates were isolated in mixed cul-
ture with C. albicans . The less susceptible C. glabrata isolate of
each patient will be designated in this study as azole resistant,
without reference to the clinical MIC breakpoints proposed by
Rex et al. (44).
The C. glabrata strains from these two patients were typed by
two different methods. The results shown in Fig. 1 demonstrate
that identical banding patterns for the two C. glabrata strains
from a given patient could be obtained by these methods and
thus suggest that the azole-susceptible and azole-resistant iso-
lates from these two patients were related to each other. This
implies that no strain replacement occurred during azole ther-
apy in these patients, but rather that azole resistance devel-
oped from the original azole-susceptible isolates. The stability
of the resistance phenotype did not change after more than 50
consecutive passages (over 500 generations) in drug-free me-
dium. Thus, azole resistance in the azole-resistant isolates
could be due to genome alterations rather than to transient
adaptation to azole antifungals.
Fluconazole and rhodamine 6G accumulations in C. glabrata
isolates. We performed several experiments in order to deter-
mine the mechanisms of azole resistance in the azole-resistant
isolates. First, no changes in amphotericin B susceptibility were
observed in the four clinical strains used in this study (Table 2).
Therefore, since some alterations in the ergosterol biosynthetic
pathway are coupled with amphotericin B resistance (17, 35), it
is likely that no alterations in this pathway were occurring in
these strains. In agreement with this hypothesis was the detec-
tion of this sterol in ergosterol biosynthesis inhibition assays.
Second, fluconazole concentrations at which the ergosterol
biosynthesis was inhibited in cellular extracts by 50% (IC 50 ) did
of the values obtained with DSY562 and DSY528, respectively.
of pNB126 to generate pDS670. pDS670 contains CgCDR1 and could transform
successfully ura3 mutants from both S. cerevisiae and C. glabrata .
Disruption of CgCDR1. For the disruption of CgCDR1 in C. glabrata , a 4.9-kb
fragment from pNB126 was cloned into the same sites of pMTL21 (10), yielding
pDS458. An 800-bp fragment was removed from pDS458 by digestion with Bgl II,
thus creating a deletion in CgCDR1. The 3.7-kb Bam HI- Bgl II fragment from
pNKY51, which contained a hisG-URA3-hisG cassette with the S. cerevisiae
URA3 gene, was inserted in the Bgl II-treated pDS458, thus yielding pDS460.
Digestion of pDS460 with Pvu II and Xho I, the latter cutting in the polylinker site
of pMTL21, generated a fragment of 5.3 kb which was used to transform
DSY1029 by lithium acetate.
Northern and Southern blots. DNA was separated by conventional 1% aga-
rose gel electrophoresis in TAE buffer (40 mM Tris-acetate [pH 7.5], 1 mM
EDTA). For RNA electrophoresis, RNA samples were resuspended in a loading
buffer (50% formamide, 100 mM morpholinepropanesulfonic acid [MOPS] [pH
7.0], 6.4% formaldehyde, 5% glycerol, 5% of a water solution saturated with
bromophenol blue), denatured at 85°C for 5 min, and separated by 1% agarose
gel electrophoresis. The agarose was melted in a buffer containing 0.1 M MOPS,
0.6 M formaldehyde, and 10
m
3
SSC (1
3
SSC
gof
salmon sperm DNA per ml. 32 P-DNA-labelled probes were generated by random
priming (14) and added to the hybridization solution overnight. Washing steps
were performed at high stringency identical to those recommended by the sup-
plier (DuPont NEN, Boston, Mass.). Stripping of probes off Northern blots for
sequential hybridizations was achieved by boiling the membranes for 10 min in
TE buffer with 0.1% SDS.
The DNA probes used in Southern and Northern blots were as follows:
CgCDR1 , 1.8-kb Xba I- Bam HI fragment from pNB126; CgYAP1 , 2.3-kb Pst I- Nru I
fragment from pNB125; and CgMDR1 , 1.2-kb Eco RV fragment from pNB124.
The CgERG11 and CgURA3 probes were generated from fragments amplified by
PCR from genomic DNA. The primers for CgURA3 (GenBank accession no.
L13661) were 5
m
9
CTCGAGAACCAATTGCATCA 3
9
and 5
9
CTAGCTTCCTA
and amplified a 900-bp fragment. The primers for CgERG11
(GenBank accession no. L40389) were 5
9
9
ATGTCCACTGAAAACACTTCT
TTG 3
9
and 5
9
CTAGTACTTTTGTTCTGGATGTCT 3
9
and amplified the
1.6-kb CgERG11 open reading frame (ORF).
Quantifications of Northern blot bands were performed by scanning the hy-
bridized membranes in an Instant Imager (Packard Instrument Company, Me-
riden, Conn.). Signals were integrated by the software supplied by the manufac-
turer and normalized to the corresponding values of an internal standard. In the
case of C. glabrata , the internal standard was the CgURA3 probe.
g of ethidium bromide per ml. The electrophoresis
buffer was 0.1 M MOPS (pH 7.0). After completion of electrophoresis, RNA was
visualized under UV light, and the position of the rRNA was determined. Both
Northern and Southern blots were performed by vacuum blotting onto Gene-
Screen Plus membranes (DuPont NEN, Boston, Mass.). Membranes containing
RNA were baked under vacuum for2hat80°C. Membranes were prehybridized
at 42°C with a buffer consisting of 50% formamide, 1% SDS, 4
is 0.15 M NaCl plus 0.015 M sodium citrate), 10% dextran sulfate, and 100
TTGGATATG 3
690426656.001.png
V OL . 43, 1999
AZOLE RESISTANCE IN C. GLABRATA
2757
pdr5 mutant YKKB-13. The plasmid backbone is YEp24. Sau 3A-digested C. glabrata DNA was introduced at the Bam HI site of
YEp24 in the construction of the gene library. The transcription directions of the ORF of each azole resistance gene deduced from nucleotide sequencing are shown
by arrows. Ba, Bam HI; Bg, Bgl II; E, Eco RI; Ev, Eco RV; Xba, Xba I; [Ba/Sa], Bam HI site of YEp24 destroyed by insertion of a genomic Sau 3A site. (B) Drug resistance
profiles of S. cerevisiae YKKB-13 transformed with plasmids pNB124, pNB125, and pNB126. Yeast strains were spotted in serial dilutions onto YEPD medium
containing the drug at the indicated concentration. Plates were incubated at 30°C for 48 h.
D
not vary significantly between azole-susceptible and azole-re-
sistant isolates (Table 3). This led us to deduce that an alter-
ation in the affinity of the C. glabrata CYP51A1 proteins to
azoles was not the cause of azole resistance in these isolates.
We therefore tested the accumulation of two different sub-
stances in the azole-susceptible and azole-resistant isolates. As
shown in Table 3, the azole-resistant isolates were accumulat-
ing less fluconazole than their azole-susceptible parents. The
accumulation of fluconazole was reduced by factors of 3.8 and
3.1 in DSY565 and DSY530 compared to those in DSY562 and
DSY528, respectively. Failure in drug accumulation was not
restricted to a single compound, since intracellular levels of
rhodamine 6G were also affected in the azole-resistant isolates
(Fig. 2). The decreases in rhodamine 6G accumulation were
6.1- and 5.1-fold in DSY565 and DSY530 compared to those in
DSY562 and DSY528, respectively. Failure in accumulating
several unrelated drugs in the azole-resistant isolates was rem-
iniscent of the effect of multidrug efflux transporters observed
in C. albicans , and therefore we exploited the possibility that
multidrug efflux transporter genes were upregulated in the
azole-resistant C. glabrata isolates.
Cloning of C. glabrata azole resistance genes. Given the
observations mentioned above, we first attempted to isolate
multidrug efflux transporter genes from C. glabrata , since at the
time these studies were initiated, no C. glabrata gene encoding
a transporter was available. We used a strategy consisting of
isolation of C. glabrata genes which could confer azole resis-
tance to an S. cerevisiae multidrug efflux transporter mutant.
Since the absence of ABC transporter Pdr5p in S. cerevisiae
yields hypersusceptibility to azole antifungals, this genetic
background was suitable to isolate genes involved in azole
resistance in C. glabrata . From a total of about 80,000 Ura 1 S.
cerevisiae transformants, more than 200 individual yeasts able
to grow on media containing 5 and 10
m
m
g of fluconazole per ml, 0.1 and 1
m
gof
itraconazole per ml, 0.2 and 2
m
g of ketoconazole per ml, 0.25
m
g of cycloheximide per ml, 10
m
g of fluphenazine per ml, 50
m
g of benomyl per ml, 0.5
m
g of crystal violet per ml, and 0.5
g of nitroquinoline- N -oxide per ml. Among the 200 flucon-
azole-resistant clones, only 20 were specifically resistant to
fluconazole among the azole drugs tested while remaining re-
sistant to cycloheximide, benomyl, and nitroquinoline- N -oxide.
The other yeast transformants were able to grow on medium
containing different azole antifungals and other drugs, such as
cycloheximide and fluphenazine. Plasmids rescued from these
isolates were extracted and separated by gel electrophoresis.
FIG. 3. Cloning of azole resistance genes in C. glabrata . (A) Restriction maps of C. glabrata insert DNA from plasmids pNB124, pNB125, and pNB126 conferring
azole resistance in the S. cerevisiae
g of fluconazole per ml
were selected. Each fluconazole-resistant clone was then tested
for different drug resistance profiles by using agar plates con-
taining 10 and 25
m
690426656.002.png
Zgłoś jeśli naruszono regulamin