ArticlePDF Available

Contribution of Galactofuranose to the Virulence of the Opportunistic Pathogen Aspergillus fumigatus

Authors:

Abstract and Figures

The filamentous fungus Aspergillus fumigatus is responsible for a lethal disease called invasive aspergillosis that affects immunocompromised patients. This disease, like other human fungal diseases, is generally treated by compounds targeting the primary fungal cell membrane sterol. Recently, glucan synthesis inhibitors were added to the limited antifungal arsenal and encouraged the search for novel targets in cell wall biosynthesis. Although galactomannan is a major component of the A. fumigatus cell wall and extracellular matrix, the biosynthesis and role of galactomannan are currently unknown. By a targeted gene deletion approach, we demonstrate that UDP-galactopyranose mutase, a key enzyme of galactofuranose metabolism, controls the biosynthesis of galactomannan and galactofuranose containing glycoconjugates. The glfA deletion mutant generated in this study is devoid of galactofuranose and displays attenuated virulence in a low-dose mouse model of invasive aspergillosis that likely reflects the impaired growth of the mutant at mammalian body temperature. Furthermore, the absence of galactofuranose results in a thinner cell wall that correlates with an increased susceptibility to several antifungal agents. The UDP-galactopyranose mutase thus appears to be an appealing adjunct therapeutic target in combination with other drugs against A. fumigatus. Its absence from mammalian cells indeed offers a considerable advantage to achieve therapeutic selectivity.
Content may be subject to copyright.
EUKARYOTIC CELL, Aug. 2008, p. 1268–1277 Vol. 7, No. 8
1535-9778/08/$08.000 doi:10.1128/EC.00109-08
Copyright © 2008, American Society for Microbiology. All Rights Reserved.
Contribution of Galactofuranose to the Virulence of the Opportunistic
Pathogen Aspergillus fumigatus
Philipp S. Schmalhorst,
1
Sven Krappmann,
2
Wouter Vervecken,
3
Manfred Rohde,
4
Meike Mu¨ller,
5
Gerhard H. Braus,
2
Roland Contreras,
3
Armin Braun,
5
Hans Bakker,
1
and Franc¸oise H. Routier
1
*
Department of Cellular Chemistry, Hannover Medical School, Hannover, Germany
1
; Department of Molecular Microbiology and
Genetics, Georg August University, Go¨ttingen, Germany
2
; Department of Molecular Biology, Ghent University, and
Department for Molecular Biomedical Research, VIB, Ghent, Belgium
3
; Department of Microbial Pathogenicity,
Helmholtz Centre for Infection Research, Braunschweig, Germany
4
; and Department of
Immunology, Allergology and Immunotoxicology, Fraunhofer Institute of Toxicology
and Experimental Medicine, Hannover, Germany
5
Received 27 March 2008/Accepted 9 June 2008
The filamentous fungus Aspergillus fumigatus is responsible for a lethal disease called invasive aspergillosis
that affects immunocompromised patients. This disease, like other human fungal diseases, is generally treated
by compounds targeting the primary fungal cell membrane sterol. Recently, glucan synthesis inhibitors were
added to the limited antifungal arsenal and encouraged the search for novel targets in cell wall biosynthesis.
Although galactomannan is a major component of the A. fumigatus cell wall and extracellular matrix, the
biosynthesis and role of galactomannan are currently unknown. By a targeted gene deletion approach, we
demonstrate that UDP-galactopyranose mutase, a key enzyme of galactofuranose metabolism, controls the
biosynthesis of galactomannan and galactofuranose containing glycoconjugates. The glfA deletion mutant
generated in this study is devoid of galactofuranose and displays attenuated virulence in a low-dose mouse
model of invasive aspergillosis that likely reflects the impaired growth of the mutant at mammalian body
temperature. Furthermore, the absence of galactofuranose results in a thinner cell wall that correlates with an
increased susceptibility to several antifungal agents. The UDP-galactopyranose mutase thus appears to be an
appealing adjunct therapeutic target in combination with other drugs against A. fumigatus. Its absence from
mammalian cells indeed offers a considerable advantage to achieve therapeutic selectivity.
The filamentous fungus Aspergillus fumigatus is the primary
cause of invasive aspergillosis, an often fatal condition affecting
people with a weakened immune system. Along with the im-
munocompromised population, the incidence of invasive as-
pergillosis is constantly growing, but therapy remains problem-
atic. The sterol binding polyene amphotericin B and the
ergosterol biosynthesis inhibitor itraconazole have long been
the drugs of choice for treatment of this infection, but because
of their higher efficacy and lower toxicity, new triazoles, such as
voriconazole or posaconazole, are supplanting these drugs (28,
33). Additionally, a novel class of antifungal agents called the
echinocandins provides further options for treatment. These
compounds inhibit the synthesis of 1,3-glucan, a major cell
wall component, with resultant osmotic instability and lysis
(12). Their minimal toxicity and synergistic activity with vori-
conazole and amphotericin B make them particularly attractive
for combination therapy, although clinical validation is still
awaited (33, 35). Despite these advances in therapy, invasive
aspergillosis is often associated with significant morbidity and
mortality, emphasizing the need for novel therapeutic strate-
gies based on the fundamental knowledge of A. fumigatus
pathogenesis.
The development of echinocandins illustrates the viability of
targeting enzymes involved in cell wall biosynthesis and en-
courages the development of chitin synthesis inhibitors. Like
glucan and chitin, galactomannan is an abundant component
of the A. fumigatus cell wall (4). This polysaccharide, composed
of a linear mannan core branched with short 1,5-linked ga-
lactofuranose (Galf) chains (22), is bound covalently to the cell
wall 1,3-glucan, anchored to the lipid membrane by a glyco-
sylphosphatidylinositol, or released in the environment during
tissue invasion or growth in culture (3, 9, 14). Besides being an
abundant component of the extracellular matrix, secreted ga-
lactomannans are used for serological diagnostic of invasive
aspergillosis (1). The monosaccharide Galf has also been
found in the N- and O-glycans of some glycoproteins as well as
the glycosphingolipids of A. fumigatus (23, 29, 41, 47) and thus
represents an important constituent of the cell wall of this
fungus. Galf is otherwise infrequent in natural compounds but
prevalent in pathogens. Moreover, since Galf is absent from
higher eukaryotes and involved in the survival or virulence of
various bacteria, the enzymes involved in the biosynthesis of
Galf are considered attractive drug targets (32, 34).
Our understanding of Galf metabolism in eukaryotes is lim-
ited. Galf is most likely incorporated into cell surface compo-
nents by specific galactofuranosyltransferases that use UDP-
Galf as a donor. The work of Trejo and colleagues in the early
1970s already suggested the existence of an enzyme converting
UDP-galactopyranose into UDP-galactofuranose involved in
* Corresponding author. Mailing address: Department of Cellular
Chemistry (OE 4330), Hannover Medical School, 30625 Hannover,
Germany. Phone: 49 (511) 532-9807. Fax: 49 (511) 532-3956. E-mail:
Routier.Francoise@mh-hannover.de.
† Present address: Research Center for Infectious Diseases, Wu¨rz-
burg, Germany.
Published ahead of print on 13 June 2008.
1268
the biosynthesis of the fungal cell wall (48). This enzyme,
named UDP-galactopyranose mutase (UGM) and encoded by
the glf gene, was described first for bacteria (17, 30, 50) and
lately for several eukaryotic pathogens, including A. fumigatus
(2, 5). UGM is to date the only characterized enzyme involved
in the biosynthesis of galactofuranose-containing molecules in
eukaryotes, whereas several galactofuranosyltransferases have
been described for bacteria (15, 19, 27, 51). The identification
of this enzyme, highly conserved among lower eukaryotes and
present in many fungi, enables studies of the biological role of
galactofuranose in these organisms. The present report high-
lights the role of galactofuranose in A. fumigatus growth and
virulence.
MATERIALS AND METHODS
Strains, media, and growth conditions. A. fumigatus clinical isolate D141 (38)
was used as the wild-type (wt) strain in this study. All strains were grown at 37°C
on Aspergillus minimal medium (AMM) containing 1%
D-glucose as the carbon
source and 70 mM NaNO
3
as the nitrogen source (36), unless otherwise stated.
Phleomycin or 5-fluoro-2-deoxyuridine (FUDR) was added at 30 g/ml or 100
M, respectively, for selection purposes.
Generation of A. fumigatus mutant strains. The 5 and 3 flanking regions (1.5
and 2 kb, respectively) of the A. fumigatus glfA coding sequence were amplified
from genomic DNA by PCR with primers PS12/PS1 and PS3/PS4 (Table 1),
respectively, and cloned into the pBluescript II SK() vector (Stratagene) by use
of the restriction sites SacII/NotI and EcoRV/ClaI. A SpeI/NotI fragment re-
leased from pSK269 containing the phleo/tk blaster (18) was then inserted
between the two fragments to obtain the disruption plasmid pglfA. For recon-
stitution of the glfA gene locus, the plasmid pglfA* was constructed as follows.
The phleo/tk blaster of pglfA was first replaced with the original A. fumigatus
glfA gene by homologous recombination in Escherichia coli strain YZ2000 (Gene
Bridges, Leimen, Germany). A single point mutation was introduced by site-
directed mutagenesis. Briefly, nonmethylated plasmid DNA was generated from
a methylated parent plasmid by Phusion DNA polymerase (NEB) using com-
plementary primers that both carried the desired mutation (PS23s/PS23r [Table
1]). Prior to transformation, the parental, methylated DNA strand was specifi-
cally cleaved by DpnI to selectively obtain transformants that harbored the
mutated plasmid. Thus, codon 130 of the glfA coding sequence (GenBank ac-
cession number AJ871145) was changed from CTT to CTC, which generated a
new XhoI restriction site. Since gene reconstitution by homologous recombina-
tion could not be obtained with this construct, 5 and 3 flanking regions were
extended to 5 kb by replacement with recloned PCR fragments (primer pairs
PS28/PS1 and PS3/PS31) to obtain the final pglfA* construct.
The pglfA and pglfA* plasmids were linearized (KpnI/SacII) before poly-
ethylene glycol-mediated fusion of protoplasts, as described previously (37).
Transformants were grown on AMM plates containing 1.2 M sorbitol as the
osmotic stabilizer under appropriate selection conditions and singled out twice
before further analysis. Accurate gene deletion and reconstitution were con-
firmed by Southern hybridization. Southern probes were amplified from genomic
DNA by using primer pairs PS66A/PS67A, PS68A/PS69A, and PS20/PS21. All
primer sequences are provided in Table 1.
Western blot analysis. Cell wall glycoproteins and soluble polysaccharides
were extracted from 30 mg ground A. fumigatus mycelium by incubation in 1 ml
sample buffer (15% glycerol, 100 mM Tris-HCl, pH 6.8, 1.5% sodium dodecyl
sulfate, 0.25% -mercaptoethanol, 0.025% bromophenol blue) for 12 min at
95°C. A portion (20 l) of the supernatant was separated on a 10% sodium
dodecyl sulfate-polyacrylamide gel and transferred to nitrocellulose membranes.
The monoclonal antibody (MAb) EB-A2 (42) conjugated to horseradish perox-
idase (HRP) from a Platelia Aspergillus test (Bio-Rad, Hercules, CA) or HRP-
coupled lectin concanavalin A (ConA) (Sigma-Aldrich) was used in a 1:50 dilu-
tion or at 0.2 g/ml, respectively. HRP activity was visualized by an enhanced
chemiluminescence system (Pierce, Rockford, IL).
N-glycan analysis. N-glycans of secreted glycoproteins in the supernatant of an
A. fumigatus liquid culture were analyzed after peptide N-glycosidase F (PNGase
F)-mediated release and 8-amino-1,3,6-pyrenetrisulfonic acid (APTS) labeling
by capillary electrophoresis, as recently described (20). Separation was carried
out on a four-capillary electrophoresis DNA sequencer (3130 genetic analyzer;
Applied Biosystems, Foster City, CA). Oligomaltose and bovine RNase B N-
glycans (Prozyme, San Leandro, CA) served as reference oligosaccharides.
Purification and analysis of GIPCs. Mycelia (0.5 g) ground in liquid nitrogen
with a mortar and pestle were disrupted by sonication in 6 ml of CHCl
3
-methanol
(MeOH), 1:1. After addition of 3 ml CHCl
3
(to obtain a CHCl
3
/MeOH ratio of
2:1), glycosylinositolphosphoceramides (GIPCs) were extracted at room temper-
ature for at least 15 min on a rotating shaker. MeOH (3 ml) was then added to
lower the density, and the mixture was centrifuged for 10 min at 2,000 g to
remove insoluble material. Chloroform and H
2
O were then added to the super
-
natant to obtain a biphasic system with an 8:4:3 ratio of CHCl
3
/MeOH/H
2
O.
After centrifugation for 10 min at 2,000 g, GIPCs contained in the upper phase
were collected and applied to a C
18
SepPak cartridge (Waters, Eschborn, Ger
-
many) preequilibrated with 5 ml CHCl
3
/MeOH/H
2
O at a ratio of 3:48:47. After
washing of the column with 20 ml CHCl
3
/MeOH/H
2
O at a ratio of 3:48:47,
glycolipids were eluted with 5 ml MeOH and dried under a stream of nitrogen.
High-performance thin-layer chromatography and immunostaining with the
MAb MEST-1 were carried out as previously described (47).
Growth assay. For radial growth measurement, a 10-l drop containing 10,000
A. fumigatus conidia was placed in the center of an agar plate containing either
minimal (AMM) or complete (potato dextrose agar; Becton Dickinson Difco,
Heidelberg, Germany) medium. Plates were incubated at various temperatures,
and colony diameters were measured twice daily.
Antifungal susceptibility testing. The reference broth microdilution test was
applied for A. fumigatus antifungal susceptibility testing (21). Each antifungal
stock was diluted in 200 l double-strength RPMI 2%G (RPMI 1640 liquid
medium buffered with 165 mM 4-morpholinepropanesulfonic acid [MOPS] to
TABLE 1. DNA oligonucleotides used in this study
Oligonucleotide Sequence (533)
a
Description (restriction site)
PS1 ATAAGCGGCCGCAAGCTGGGAACGCGATTCAA 5 Flanking region pglfA reverse (NotI)
PS12 TATACCGCGGCTGCCAAGCTATCAGTTTCC 5 Flanking region pglfA sense (SacII)
PS3 ATCCGGTGCTCAGGTATTCGCCA 3 Flanking region pglfA sense (EcoRV)
PS4 ATCCATCGATCATATCCTATGCGGTCTCAG 3 Flanking region pglfA reverse (ClaI)
PS66A TTACGCATTCCCAGCAGTTG Southern blot probe 1 sense
PS67A TGCGCTGTGATGAATGGTGT Southern blot probe 1 reverse
PS68A TCCACAATACGTCCCCTACA Southern blot probe 2 sense
PS69A GTATGAACCCTCTCCCAATG Southern blot probe 2 reverse
PS20 AAGGTCGTTGCGTCAGTCCA Southern blot probe 3 sense
PS21 TCGATGTGTCTGTCCTCC Southern blot probe 3 reverse
PS23s ATGCCGCTCTCGAGGCTCGT Site-directed mutagenesis glfA* sense (XhoI)
PS23r CACGAGCCTCGAGAGCGGCA Site-directed mutagenesis glfA* reverse (XhoI)
PS28 ATATGCGGCCGCAAACAGGAGCGAAGTAGT 5 Flanking region pglfA* sense (NotI)
PS31 ATATCCCGGGAGTTTGGTGCTGTGGTAGGT 3 Flanking region pglfA* reverse (XmaI)
PS78 CGTGTCTATCGTACCTTGTTGCTT 18S rRNA gene fragment sense
PS79 AACTCAGACTGCATACTTTCAGAACAG 18S rRNA gene fragment reverse
Probe FAM-CCCGCCGAAGACCCCAACATG-TAMRA qPCR hybridization probe
a
Restriction sites are underlined. TAMRA, 6-carboxytetramethylrhodamine.
VOL. 7, 2008 A. FUMIGATUS UDP-GALACTOPYRANOSE MUTASE 1269
pH 7.0 and supplemented with 2% glucose) to obtain the highest concentration
to be tested. Nine serial 1:2 dilutions in double-strength RPMI 2%G were made,
and to each dilution a volume of 100 lofanA. fumigatus spore solution
(2.5 10
5
/ml in water) was added. Microtiter plates were incubated at 35°C, and
fungal growth in each well was read out visually after 3 days and compared to
growth in control wells that contained no antifungal.
Field emission scanning electron microscopy. For morphological studies and
measurements of the cell wall thickness, A. fumigatus wt and glfA mutant
mycelia were fixed in 5% formaldehyde and 2% glutaraldehyde in cacodylate
buffer (0.1 M cacodylate, 0.01 M CaCl
2
, 0.01 M MgCl
2
, 0.09 M sucrose, pH 6.9)
for1honice. Samples were washed several times with cacodylate buffer and
subsequently with TE buffer (20 mM Tris-HCl, 1 mM EDTA, pH 6.9) before
dehydration in a graded series of acetone (10, 30, 50, 70, 90, and 100%) on ice
for 15 min per step. Samples at the 100% acetone step were allowed to reach
room temperature before another change in 100% acetone. Samples were then
subjected to critical-point drying with liquid CO
2
(CPD 30; Balzers Union,
Liechtenstein). Dried samples were then mounted onto conductive carbon ad-
hesive tabs on an aluminum stub and sputter coated with a thin gold film (SCD
40; Balzers Union, Liechtenstein). For cell wall thickness measurements, myce-
lium was fractured by pressing another conductive carbon adhesive tab-covered
stub onto the sample and separating both stubs immediately thereafter. Frac-
tured hyphae were also made conductive by sputter coating with a gold film
before examination with a field emission scanning electron microscope (Zeiss
DSM 982 Gemini) using an Everhart Thornley SE detector and an in-lens SE
detector at a 50:50 ratio at an acceleration voltage of 5 kV and at calibrated
magnifications.
Mouse infection model. A low-dose mouse infection model of invasive as-
pergillosis for BALB/c mice which had been established previously (25) was
essentially used. The immunosuppressive state was established by intraperitoneal
injections of 100 mg cyclophosphamide (Endoxan; Baxter Chemicals)/kg of body
weight on days 4, 1, 0, 2, 5, 8, and 11 and a single subcutaneous dose (200
mg/kg) of a cortisone acetate suspension (Sigma) on day 1. Groups of 20 mice
were infected intranasally with 20,000 conidia of the wt strain, the glfA strain,
or the reconstituted glfA* strain on day 0. The control group received phosphate-
buffered saline (PBS) only. Survival was monitored for 13 days after infection,
and moribund animals were sacrificed. Coincidence of severely reduced mobility,
low body temperature, and breathing problems was defined as the moribundity
criterion. Statistical analysis of survival data was carried out using a log rank test
implemented in Prism 4 (GraphPad Software, San Diego, CA). For quantifica-
tion experiments, groups of three to five animals were killed 2, 4, and 6 days after
infection and lungs were removed for further analysis.
Lung histology. Female BALB/c mice were immunosuppressed and infected as
described above. The animals were killed after 5 days and their lungs removed
and fixed in 4% PBS-buffered paraformaldehyde overnight. Tissue samples were
dehydrated through a series of graded alcohols, cleared with xylene, and embed-
ded in paraffin. Tissue sections (5 m) were stained either with hematoxylin/
eosin or by the periodic acid-Schiff method for visualization of fungal cell walls.
Photomicrographs were taken with an Axiovert 200M microscope (Zeiss, Ger-
many) at 10 and 20 magnifications.
Preparation of genomic DNA from mouse lungs. Tissue homogenization was
essentially as described previously (7). Immediately after removal, mouse lungs
were transferred to a 2-ml screw cap containing 1.4-mm ceramic beads (lysing
matrix D; Qbiogene, Irvine, CA) and 20% glycerol-PBS. Tissue was disrupted
using a FastPrep FP120 instrument (Qbiogene) three times for 30 s each at speed
5, with intermediate cooling on ice. The disrupted tissues were further homog-
enized with approximately 250 mg acid-washed glass beads (0.45 to 0.5 mm;
Sigma-Aldrich) by vortexing three times for 30 s each, with intermediate cooling
on ice. A DNeasy blood and tissue kit (Qiagen, Hilden, Germany) was used to
extract genomic DNA from an equivalent of 8% of the starting tissue material of
this homogenate. DNA was finally recovered in 200 l elution buffer.
qPCR. Quantitative PCR (qPCR) was carried out essentially as described
previously (7). Primers for amplification of an 18S rRNA gene (GenBank ac-
cession number AB008401) fragment specific for A. fumigatus and a hybridiza-
tion probe labeled with 6-carboxyfluorescein (FAM) (5 end) and 6-carboxytet-
ramethylrhodamine (3 end) were designed using Primer Express software,
version 3.0 (Applied Biosystems) (Table 1). qPCR reactions were performed
with a 7500 Fast real-time PCR system instrument (Applied Biosystems) loaded
with MicroAmp optical 96-well plates sealed with an optical adhesive cover
(Applied Biosystems). Each qPCR reaction mixture (20 l) contained 5 l
sample DNA, 250 nM dual-labeled hybridization probe, 500 nM primers, 250
g/ml bovine serum albumin, and TaqMan Fast universal PCR master mix
(Applied Biosystems). The latter contains hot-start DNA polymerase, de-
oxynucleoside triphosphates, and the fluorescent dye carboxyrhodamine (ROX)
as a passive reference. Real-time PCR data were acquired using Sequence
Detection software, v1.3.1. The FAM/ROX fluorescence ratio was recorded at
every cycle, and a threshold cycle (C
T
) value was assigned to each reaction
product, defining the cycle number at which the FAM/ROX signal surpassed an
automatically defined threshold. C
T
values were corrected for differences in
yield of genomic DNA by normalization to the DNA concentration of a control
sample by using the formula C
T,norm
C
T,measured
log
2
(DNA
sample
/DNA
control
)
(7). Translation of sample C
T,norm
values into rRNA gene copy numbers was
done as follows. C
T
values of serial 1:10 dilutions containing 300 to 300,000
molecules (n) (calculated from M
w
and DNA concentration determined by
measurement of optical density at 260 nm) of a plasmid bearing the cloned A.
fumigatus 18S rRNA gene were plotted against n to generate a calibration curve
which was then used to assign an rRNA gene copy number to a given sample
C
T,norm
value. Conidial equivalents were calculated from gene copy numbers by
means of uninfected tissue samples that were spiked with defined numbers of
conidia before tissue homogenization (7). Samples, controls, and standards were
analyzed in triplicate.
RESULTS
Deletion and reconstitution of the glfA gene in A. fumigatus.
To begin investigating the role of Galf in A. fumigatus biology,
we deleted the gene encoding UGM (GenBank accession no.
AJ871145) and named it glfA, following the recommendations
for gene naming for Aspergillus. To do this, we generated a
deletion plasmid containing the regions flanking the glfA cod-
ing sequence separated by the bifunctional selection cassette
phleo/tk, which confers both resistance to phleomycin and sen-
sitivity to FUDR (18). This construct was used to transform
protoplasts of A. fumigatus clinical strain D141, which served as
the wt, and phleomycin-resistant transformants were analyzed
by Southern blotting using several digoxigenin-labeled probes
(Fig. 1). One of the clones that had undergone the desired
gene replacement (Fig. 1) was selected for further analysis and
named glfA.
The selected disruptant was further subjected to protoplast
transformation with a large DNA fragment encompassing the
glfA coding sequence which contained a single translationally
silent nucleotide exchange that generated an XhoI restriction
site. Gene replacement in the transformants resulted in the
reconstitution of the glfA locus (Fig. 1) as detected by FUDR
resistance and proven by Southern blot analysis for a selected
clone named glfA* (Fig. 1B). The silent mutation introduced in
the reconstituted strain allowed differentiation between the wt
and the glfA* mutant, as demonstrated in Fig. 1B (top), and
thus enabled us to rule out contamination by the wt strain. The
reconstitution of the glfA locus ensures that any phenotype
observed for the glfA strain can be reverted and hence be
securely attributed to the loss of the glfA gene.
Galf is absent from the A. fumigatus glfA mutant. To con-
firm that deletion of glfA indeed altered the expression of
Galf-containing glycoconjugates, aqueous mycelial extracts
were tested for reactivity to the Galf-specific MAb EB-A2. This
antibody recognizes preferably 1,5-linked Galf residues that
are present in all forms of galactomannan (cell wall bound,
membrane bound, and secreted) (42) as well as in some O-
glycans (23). Moreover, a second binding epitope, Galf(1,2)
Man, which is part of galactofuranosylated N-glycans, has been
postulated (29). Thus, EB-A2 can be used to detect galacto-
mannan and galactofuranosylated glycoproteins simulta-
neously. Western blot analysis of wt and glfA* total mycelial
extracts labeled with HRP-conjugated EB-A2 revealed a smear
migrating around 40 to 80 kDa, in accordance with previous
1270 SCHMALHORST ET AL. EUKARYOT.CELL
findings (42). In contrast, the glfA mycelial extract was not
stained at all, indicating the absence of Galf in the galacto-
mannan and glycoproteins of this mutant (Fig. 2A, left). ConA
used as the loading control bound slightly better to the glfA
extract than to those of the wt and the glfA* mutant (Fig. 2A,
right). The lack of Galf in the glfA mutant might increase the
accessibility of the mannan for ConA and thus could explain
this finding.
Similarly, the absence of Galf in glfA glycolipids was shown
by the absence of reactivity to the MAb MEST-1. This anti-
body, which recognizes 1,3- and 1,6-linked Galf residues
(43), labeled several A. fumigatus GIPCs after separation by
high-performance thin-layer chromatography, as previously
shown (47), but did not label glycosphingolipids extracted from
the glfA mutant (Fig. 2B, left). The upper bands observed in
Fig. 2B (left panel) might be attributed to GIPCs containing
one or two Galf and two or three mannose residues, as recently
described (41, 47). In addition, Simenel et al. (41) reported an
unusual GIPC containing a Galf residue substituted at position
6 by a choline phosphate. The lower band present in the wt
chromatogram could correspond to a similar GIPC. Staining of
glycolipids by orcinol was used as the loading control (Fig. 2B,
right). The simpler glfA chromatogram is compatible with the
absence of Galf-containing GIPCs. The uppermost band ob-
served in the chromatogram most probably corresponds to
Man(1,3)Man(1,2)Ins-P-Cer, while the band just beneath it
could be attributed to Man(1,2)Man(1,3)Man(1,2)Ins-P-Cer
(47). The chromatograms obtained from the reconstituted glfA*
mutant and the wt were indistinguishable (data not shown).
Additionally, N-glycans enzymatically released from A. fu-
FIG. 1. (A) Schematic representation of the chromosomal glfA locus in wt, glfA, and reconstituted glfA* strains. The thick black bars show
flanking regions used for homologous recombination. The positions of probes (probes 1 to 3) used for Southern blot analysis, along with the
respective restriction fragments (sizes in kb), are indicated. ble/tk, phleomycin resistance/thymidine kinase fusion gene; P, promoter; T, terminator.
(B) Southern blots of genomic DNA digested with the indicated restriction enzymes and hybridized to different digoxigenin-labeled probes.
FIG. 2. (A) Western blots of A. fumigatus mycelial extracts containing glycoproteins and cell wall polysaccharides stained with HRP conjugates
of either Galf-specific MAb EB-A2 (left) or -mannose binding lectin ConA (right). (B) A. fumigatus GIPCs separated by high-performance
thin-layer chromatography and stained with Galf-specific MAb MEST-1 (left) or orcinol (right). White bars indicate the origin.
V
OL. 7, 2008 A. FUMIGATUS UDP-GALACTOPYRANOSE MUTASE 1271
1272
migatus secreted proteins were analyzed by capillary electro-
phoresis after fluorescent labeling (8, 20). The profiles ob-
tained are presented in Fig. 3A (panels 1 and 2). The peaks
labeled 1, 2, 3, 4, and 5 present in both electropherograms
comigrated with reference oligosaccharides M5 to M9 (Fig.
3A, panel 9, and 3B). Moreover, digestion of these N-glycans
by Trichoderma reesei 1,2-mannosidase indicates that peaks 2,
3, 4, and 5 arise from substitution of oligosaccharide 1 with one
to four mannose residues linked in 1,2 (Fig. 3A, panels 3 and
4). The profile obtained with wt N-glycans (Fig. 3A, panel 1)
presents four additional peaks (labeled 1a, 2a, 3a, and 4a) that
were absent from glfA N-glycans. The retention times of these
peaks suggest that they arise from substitution of oligosaccha-
rides 1 to 4 with a single Galf residue. The presence of a
terminal nonreducing Galf residue in A. fumigatus N-glycans
has been reported previously (9) and was demonstrated by
hydrofluoric acid treatment of the N-glycans after T. reesei
1,2-mannosidase digestion (Fig. 3A, panels 5 and 6). This
mild acid treatment, known to release Galf, entirely converted
oligosaccharide 1a into oligosaccharide 1 (Fig. 3A, panels 3
and 5). In contrast, hydrofluoric acid treatment did not change
the profile of glfA N-glycans digested with 1,2-mannosidase
(Fig. 3A, panels 4 and 6).
Interestingly, the comparison of wt and glfA N-glycans di-
gested with T. reesei 1,2-mannosidase or jack bean mannosi-
dase helps with positioning of the Galf residue. 1,2-Manno-
sidase treatment converted the oligosaccharides 2a, 3a, and 4a
into 1a while the oligosaccharides 2, 3, and 4 generated 1 (Fig.
3A, compare panels 1 and 2 with panels 3 and 4). This indicates
that the Galf residue does not protect any mannose residues
from the exomannosidase digestion and thus does not substi-
tute an 1,2-linked mannose (Fig. 3A, panels 3 and 4). More-
over, jack bean mannosidase digestion of wt N-glycans resulted
in a major peak (peak 7), attributed to GalfMan
3
GlcNAc
2
from its retention time, in addition to Man
1
GlcNAc
2
(peak 6),
expected from digestion of high-mannose type N-glycans (Fig.
3A, panels 7 and 8). These experiments do not allow for the
determination of the detailed N-glycan structures but suggest
that they resemble the N-glycans of A. niger -glucosidase and
-galactosidase (44, 45). More importantly, these experiments
demonstrate the absence of Galf in the glfA N-glycans.
Loss of Galf alters morphology and growth of A. fumigatus.
The glfA strain exhibited a marked growth defect on solid
minimal media or complete media compared to the wt. This
effect could be observed for a wide range of temperatures (Fig.
4) and was statistically different in all cases (P 0.001, t test,
n 3). The most severe effect was found at 42°C, with a 75%
reduction in radial growth (Fig. 4B). In parallel, glfA conidia-
tion was diminished by 90% at 37°C and was almost absent at
42°C. In contrast, the onsets and rates of germination of wt,
glfA, and glfA* conidia were similar. In minimal media at
37°C, the conidia of all strains started forming germ tubes at
3.2 h and reached 100% germination within 8 to 9 h (data not
shown).
Scanning electron micrographs of intact mycelium, conidio-
phores, and conidia of glfA did not reveal any obvious mor-
phological differences when compared to wt. However, the
observation of fractured mycelium revealed a marked reduc-
tion of the glfA cell wall thickness (Fig. 5). Measurements
indicated that the cell wall thickness of wt A. fumigatus varies
from 85 to 315 nm, which is in good agreement with earlier
findings (39). In contrast, glfA cell wall thickness ranged from
FIG. 3. (A) Electropherograms of fluorescently labeled N-glycans enzymatically released from secreted A. fumigatus glycoproteins. Oligosac-
charides from the wt and the glfA mutant were untreated (panels 1 and 2), digested with T. reesei 1,2-exomannosidase with or without
hydrofluoric acid (HF) treatment (panels 3 to 6), or digested with jack bean -mannosidase (panels 7 and 8). Bovine RNase B N-glycans served
as the reference (panel 9). (B) Structures of bovine RNase B reference N-glycans. (C) Major N-glycans found on A. niger -galactosidase and
-glucosidase (44, 45). Black squares, N-acetylglucosamine; gray circles, mannose; white pentagon, galactofuranose.
FIG. 4. (A) Colony morphology of A. fumigatus on minimal agar after 2 days. Bar, 1 cm. (B) Absolute and relative (rel.) (compared to the wt)
growth rates derived from three independent experiments. P values derived from a t test indicate statistical significance (***, P 0.001; ns, not
significant).
V
OL. 7, 2008 A. FUMIGATUS UDP-GALACTOPYRANOSE MUTASE 1273
85 to 150 nm. The mean values (standard deviations) of cell
wall thickness obtained from 25 measurements were 227.5 nm
(15.98 nm) and 109.7 nm (11.3 nm) for wt and glfA hy-
phae, respectively. The cell wall of glfA was thus approxi-
mately half the thickness of the wt cell wall.
glfA is more susceptible to drugs. The structural cell wall
defect caused by the Galf deficiency was accompanied by an
increased susceptibility to several antifungal agents (Table 2).
MICs determined by a broth microdilution test were slightly
reduced for amphotericin B and caspofungin in the glfA mu-
tant. A more pronounced increase in susceptibility was seen for
voriconazole (0.04 mg/liter for glfA, compared to 0.3 mg/liter
for the wt) and nikkomycin Z (63 to 125 mg/liter for glfA and
500 mg/ml for the wt), suggesting an increased permeability of
the cell wall caused by the loss of Galf. In contrast, the sensi-
tivity toward oxidative stress remained unchanged, as indicated
by equal MICs for H
2
O
2
in both wt and glfA strains.
glfA displays attenuated virulence in a murine model of
invasive aspergillosis. The influence of the glfA deletion on the
pathogenicity of A. fumigatus was assessed in a low-dose mouse
infection model of invasive aspergillosis (25). Cyclophospha-
mide was used to induce neutropenia in female BALB/c mice,
and a single dose of cortisone acetate was administered before
intranasal infection with 20,000 A. fumigatus conidia. Neutro-
penia was maintained throughout the observation period of 13
days, and survival was recorded daily (Fig. 6A). Ninety percent
of the animals infected with the wt did not survive beyond day
7 after infection, whereas half of the mice infected with glfA
were still alive on day 13. A log rank test of wt and glfA
survival data confirmed that the observed difference was sta-
tistically significant (P 0.0004). The attenuation in virulence
could clearly be attributed to the absence of glfA, since animals
infected with the reconstituted glfA* strain showed a survival
pattern nearly identical to that of the wt (no significant differ-
ence by the log rank test, P 0.559). A histological examina-
tion of lung tissue from mice infected with the wt, glfA, and
glfA* strains 5 days after inoculation showed evident fungal
growth surrounding bronchioles and tissue penetration (Fig.
7). For each strain, inflammatory cells were rarely observed at
the sites of infection.
To correlate the delay in the onset and progression of mor-
tality with a growth defect, the fungal burden in lungs of in-
fected mice was determined by qPCR (Fig. 6B). Mice were
treated and infected as described above. After 2, 4, and 6 days,
animals were sacrificed and their lungs taken. DNA was iso-
lated from homogenized lung tissue and fungal content deter-
mined by amplification of a part of A. fumigatus ribosomal
DNA. As shown in Fig. 6B, growth of glfA was restricted in
vivo compared to that of the wt, which was in agreement with
the slower growth observed in vitro.
DISCUSSION
The essential role of the 1,3-glucan in cell wall organization
and growth of several pathogenic fungi has been the basis for
the development of the echinocandins (11). Likewise, inhibi-
tors of chitin biosynthesis are currently being explored as new
antifungal drugs since chitin is an important structural element
of the fungal cell wall (6). In contrast, although galactomannan
is a major component of the cell wall and the extracellular
matrix, the role of galactomannan had not yet been investi-
gated, since the enzymes involved in its biosynthesis are un-
known. Recently, we and others characterized the UGM of
various pathogenic eukaryotes, including A. fumigatus (2, 5). In
prokaryotes, like in the protozoan Leishmania, this enzyme is
the only route to the formation of UDP-Galf, the donor sub-
strate of galactofuranosyltransferases, and thus controls the
biosynthesis of all Galf-containing molecules. Likewise, A. fu-
migatus UGM was found to be essential for the biosynthesis of
galactomannan as well as some glycosphingolipids and glyco-
proteins. Like in other organisms (16, 32), deletion of the glfA
gene resulted in the complete absence of Galf, as shown for
instance by the absence of reactivity to the antibody EB-A2.
Besides demonstrating the lack of Galf in the glfA mutant,
our analyses provide useful structural information for A. fu-
migatus N-glycans. Treatment of wt secreted proteins with
PNGase F released galactofuranosylated high-mannose type
N-glycans. The size of the oligosaccharides and the presence of
a single Galf residue are in agreement with previous studies of
filamentous fungi (26, 29). Moreover, analysis of these oligo-
FIG. 5. Field emission scanning electron micrographs of cross-fractured mycelial walls of A. fumigatus wt and glfA hyphae. Panels 1 and 3
display the highest measurement of cell wall thickness. Panels 2 and 4 present two measurements illustrating the 50% reduction of glfA cell wall
thickness.
TABLE 2. MICs of various antifungal agents against A. fumigatus
mutants, obtained from a broth microdilution assay
Genotype
MIC (mg/liter)
a
of:
AmB Vor Cas NiZ H
2
O
2
wt 3.9 0.3 62.5 500 218
glfA 2.0 0.04 31.3 62.5–125 218
glfA* 3.9 0.3 62.5 500 218
a
Values for amphotericin B (AmB) and caspofungin (Cas) are MIC
90
values,
values for voriconazole (Vor) and nikkomycin Z (NiZ) are MIC
50
values, and
values for H
2
O
2
are MIC
100
values.
1274 SCHMALHORST ET AL. EUKARYOT.CELL
saccharides after digestion by jack bean or T. reesei 1,2-man-
nosidase helps with positioning of the Galf residue. These data
and the comparison with high-mannose standards suggest that
the N-glycans from A. fumigatus secreted proteins resemble
those of A. niger -
D-galactosidase and -D-glucosidase (44, 45,
49). These N-glycans might have arisen simply from trimming
of the Glc
3
Man
9
GlcNAc
2
precursor and substitution by a Galf
residue. Aspergillus spp. indeed contain several 1,2-mannosi-
dase genes, and trimming of high-mannose glycans has been
shown previously (13, 52). Interestingly, Galf addition has been
suggested to act as a stop signal for mannose addition, in
analogy to the role proposed for the 1,3-linked terminal man-
nose in Saccharomyces cerevisiae (29, 49). However, preventing
the addition of galactofuranose does not result in an increased
size of the oligosaccharides. On the contrary, Man
5
GlcNAc
2
is
the main oligosaccharide found in the glfA mutant, while
GalfMan
6
GlcNAc
2
is predominant in the wt.
Although glfA deletion has been shown to be lethal in My-
cobacterium smegmatis (32), the in vitro viability of the A.
fumigatus glfA mutant is unsurprising, since Galf occupies a
nonreducing terminal position in the molecules of this fungus.
Hence, the absence of Galf does not perturb the basic organi-
zation of the cell wall, as would the absence of the underlying
structures. Nevertheless, it resulted in marked alterations of
the cell surface and a notably thinner cell wall, as revealed by
electron microscopy. The basis of this drastic change is unclear
and difficult to attribute to a particular cell wall component,
since glycosylphosphatidylinositol/cell wall bound galactoman-
FIG. 6. (A) Survival of immunosuppressed mice infected intranasally with A. fumigatus wt (solid line), glfA (dotted line), or glfA* (dashed line)
strains and of uninfected mice (dotted and dashed line). Each group consisted of 20 animals. (B) qPCR determination of A. fumigatus burden
(measured as conidial equivalents [eq.] [see Materials and Methods]) in lung tissue from immunosuppressed mice infected with wt (solid line) or
glfA (dotted line) strains. Each data point represents the mean value obtained from three to five animals. Error bars indicate standard errors of
the means.
FIG. 7. Periodic acid-Schiff-stained lung sections of mice infected with wt, glfA,orglfA* A. fumigatus strains. Fungal colonies appear
purple/red. Infected sites are typically surrounded by areas of necrotic tissue but show no or hardly any infiltrating leukocytes. Bar, 100 m.
VOL. 7, 2008 A. FUMIGATUS UDP-GALACTOPYRANOSE MUTASE 1275
nan, N-glycans, O-glycans, and GIPCs are affected by UDP-
Galf deficiency. In other fungi, the loss of terminal sugar res-
idues has sometimes been associated with reduced cell wall
strength. For instance, a Schizosaccharomyces pombe mutant
deficient in cell wall galactosylation displays morphological
changes, attenuated growth, and a 25 to 35% reduction in cell
wall thickness (46).
The structural changes originating from the glfA deletion are
associated with slower growth, indicating that Galf plays an
important role in A. fumigatus morphogenesis. The tempera-
ture-sensitive growth defect at a higher temperature displayed
by the glfA mutant is reminiscent of that observed for the
AfPmt1 mutant, a mutant characterized by reduced O glyco-
sylation (53). Interestingly, an influence of Galf deficiency on
the growth rate was also observed for glfA mutants of As-
pergillus nidulans (F. H. Routier, unpublished data) and As-
pergillus niger (10). Conversely, glfA deletion had no effect on
the in vitro growth of Leishmania parasites (16), highlighting
that the role of Galf cannot be translated to every Galf-con-
taining organism.
The ability to thrive at 37°C is a characteristic of human
pathogens that has been shown to correlate with virulence
potential in the case of A. fumigatus (31). Consequently, mu-
tations that affect the growth of fungi at mammalian body
temperature are commonly associated with attenuated viru-
lence (40). In this study, we observed slower growth of the A.
fumigatus glfA mutant in vitro but also in vivo by using qPCR.
In agreement with this observation, the mutant was clearly
attenuated in virulence, showing a delay in both the onset and
the progression of mortality when tested in a low-dose mouse
infection model of invasive aspergillosis. An altered immune
response caused by the different cell wall structure of the glfA
mutant may also contribute to the attenuation in virulence.
However, no differences in adherence and uptake of wt and
glfA conidia by murine bone marrow-derived dendritic cells
or in production of tumor necrosis factor alpha or interleu-
kin-10 by infected murine bone-marrow derived macrophages
were observed (K. Kotz, F. Ebel, and F. H. Routier, unpub-
lished data).
The value of echinocandins in invasive aspergillosis treatment
resides in their synergistic effects with azoles and amphotericin
B. Similarly, chitin synthesis inhibitors demonstrate synergy
with echinocandins and azoles (24). These synergistic effects
that offer new options for combination antifungal therapy are
most likely due to greater cell wall permeability. We did note
an increase in susceptibility of the glfA mutant to several
antifungal agents, notably to voriconazole. However, in the
liquid culture conditions classically used for antifungal suscep-
tibility testing, the fungus is not surrounded by an extracellular
matrix. This extracellular matrix that delays the penetration of
drug is rich in galactomannan (3) and is probably altered in the
glfA mutant, as suggested by the compact appearance of
colonies on agar plates. In vivo, a greater increase in suscep-
tibility of the glfA mutant to drugs would therefore be ex-
pected. Besides the attenuated virulence, this suggests that
inhibitors of UGM might be useful in antifungal therapy. The
absence of Galf biosynthesis in mammals would represent a
considerable advantage for the development of antifungal
drugs with selective toxicity.
ACKNOWLEDGMENTS
We thank Florian La¨nger, Frank Ebel, and Jakob Engel for their
help with histopathology, cytokine analysis, and glycolipid analysis,
respectively. Monika Berger, Verena Grosse, Sabine Schild, Olaf
Macke, and Brigitte Philippens are thanked for excellent technical
assistance, and Anita Straus and Helio Takahashi are thanked for the
generous gift of the MEST-1 MAb. We are indebted to Rita Gerardy-
Schahn for her constant support.
REFERENCES
1. Aquino, V. R., L. Z. Goldani, and A. C. Pasqualotto. 2007. Update on the
contribution of galactomannan for the diagnosis of invasive aspergillosis.
Mycopathologia 163:191–202.
2. Bakker, H., B. Kleczka, R. Gerardy-Schahn, and F. H. Routier. 2005. Iden-
tification and partial characterization of two eukaryotic UDP-galactopyra-
nose mutases. Biol. Chem. 386:657–661.
3. Beauvais, A., C. Schmidt, S. Guadagnini, P. Roux, E. Perret, C. Henry, S.
Paris, A. Mallet, M. C. Pre´vost, and J. P. Latge´. 2007. An extracellular matrix
glues together the aerial-grown hyphae of Aspergillus fumigatus. Cell. Micro-
biol. 9:1588–1600.
4. Bernard, M., and J. P. Latge´. 2001. Aspergillus fumigatus cell wall: compo-
sition and biosynthesis. Med. Mycol. 39(Suppl 1):9–17.
5. Beverley, S. M., K. L. Owens, M. Showalter, C. L. Griffith, T. L. Doering,
V. C. Jones, and M. R. McNeil. 2005. Eukaryotic UDP-galactopyranose
mutase (GLF gene) in microbial and metazoal pathogens. Eukaryot. Cell
4:1147–1154.
6. Borgia, P. T., and C. L. Dodge. 1992. Characterization of Aspergillus nidulans
mutants deficient in cell wall chitin or glucan. J. Bacteriol. 174:377–383.
7. Bowman, J. C., G. K. Abruzzo, J. W. Anderson, A. M. Flattery, C. J. Gill,
V. B. Pikounis, D. M. Schmatz, P. A. Liberator, and C. M. Douglas. 2001.
Quantitative PCR assay to measure Aspergillus fumigatus burden in a murine
model of disseminated aspergillosis: demonstration of efficacy of caspofun-
gin acetate. Antimicrob. Agents Chemother. 45:3474–3481.
8. Callewaert, N., S. Geysens, F. Molemans, and R. Contreras. 2001. Ultrasen-
sitive profiling and sequencing of N-linked oligosaccharides using standard
DNA-sequencing equipment. Glycobiology 11:275–281.
9. Costachel, C., B. Coddeville, J. P. Latge´, and T. Fontaine. 2005. Glyco-
sylphosphatidylinositol-anchored fungal polysaccharide in Aspergillus fu-
migatus. J. Biol. Chem. 280:39835–39842.
10. Damveld, R. A., A. Franken, M. Arentshorst, P. J. Punt, F. M. Klis, C. A. van
den Hondel, and A. F. Ram. 2008. A novel screening method for cell wall
mutants in Aspergillus niger identifies UDP-galactopyranose mutase as an
important protein in fungal cell wall biosynthesis. Genetics 178:873–881.
11. Denning, D. W. 2002. Echinocandins: a new class of antifungal. J. Antimi-
crob. Chemother. 49:889–891.
12. Denning, D. W. 2003. Echinocandin antifungal drugs. Lancet 362:1142–1151.
13. Eades, C. J., and W. E. Hintz. 2000. Characterization of the class I alpha-
mannosidase gene family in the filamentous fungus Aspergillus nidulans.
Gene 255:25–34.
14. Fontaine, T., C. Simenel, G. Dubreucq, O. Adam, M. Delepierre, J. Lemoine,
C. E. Vorgias, M. Diaquin, and J. P. Latge´. 2000. Molecular organization of
the alkali-insoluble fraction of Aspergillus fumigatus cell wall. J. Biol. Chem.
275:27594–27607.
15. Guan, S., A. J. Clarke, and C. Whitfield. 2001. Functional analysis of the
galactosyltransferases required for biosynthesis of
D-galactan I, a component
of the lipopolysaccharide O1 antigen of Klebsiella pneumoniae. J. Bacteriol.
183:3318–3327.
16. Kleczka, B., A. C. Lamerz, G. van Zandbergen, A. Wenzel, R. Gerardy-
Schahn, M. Wiese, and F. H. Routier. 2007. Targeted gene deletion of
Leishmania major UDP-galactopyranose mutase leads to attenuated viru-
lence. J. Biol. Chem. 282:10498–10505.
17. Ko¨plin, R., J. R. Brisson, and C. Whitfield. 1997. UDP-galactofuranose
precursor required for formation of the lipopolysaccharide O antigen of
Klebsiella pneumoniae serotype O1 is synthesized by the product of the
rfbD
KPO1
gene. J. Biol. Chem. 272:4121–4128.
18. Krappmann, S., O. Bayram, and G. H. Braus. 2005. Deletion and allelic
exchange of the Aspergillus fumigatus veA locus via a novel recyclable marker
module. Eukaryot. Cell 4:1298–1307.
19. Kremer, L., L. G. Dover, C. Morehouse, P. Hitchin, M. Everett, H. R. Morris,
A. Dell, P. J. Brennan, M. R. McNeil, C. Flaherty, K. Duncan, and G. S.
Besra. 2001. Galactan biosynthesis in Mycobacterium tuberculosis. Identifi-
cation of a bifunctional UDP-galactofuranosyltransferase. J. Biol. Chem.
276:26430–26440.
20. Laroy, W., R. Contreras, and N. Callewaert. 2006. Glycome mapping on
DNA sequencing equipment. Nat. Protoc. 1:397–405.
21. Lass-Flo¨rl, C., M. Cuenca-Estrella, D. W. Denning, and J. L. Rodriguez-
Tudela. 2006. Antifungal susceptibility testing in Aspergillus spp. according to
EUCAST methodology. Med. Mycol. 44:319–325.
22. Latge´, J. P., H. Kobayashi, J. P. Debeaupuis, M. Diaquin, J. Sarfati, J. M.
Wieruszeski, E. Parra, J. P. Bouchara, and B. Fournet. 1994. Chemical and
1276 SCHMALHORST ET AL. EUKARYOT.CELL
immunological characterization of the extracellular galactomannan of As-
pergillus fumigatus. Infect. Immun. 62:5424–5433.
23. Leita˜o, E. A., V. C. Bittencourt, R. M. Haido, A. P. Valente, J. Peter-
Katalinic, M. Letzel, L. M. de Souza, and E. Barreto-Bergter. 2003. Beta-
galactofuranose-containing O-linked oligosaccharides present in the cell wall
peptidogalactomannan of Aspergillus fumigatus contain immunodominant
epitopes. Glycobiology 13:681–692.
24. Li, R. K., and M. G. Rinaldi. 1999. In vitro antifungal activity of nikkomycin
Z in combination with fluconazole or itraconazole. Antimicrob. Agents Che-
mother. 43:1401–1405.
25. Liebmann, B., T. W. Mu¨hleisen, M. Mu¨ller, M. Hecht, G. Weidner, A. Braun,
M. Brock, and A. A. Brakhage. 2004. Deletion of the Aspergillus fumigatus
lysine biosynthesis gene lysF encoding homoaconitase leads to attenuated
virulence in a low-dose mouse infection model of invasive aspergillosis. Arch.
Microbiol. 181:378–383.
26. Maras, M., I. van Die, R. Contreras, and C. A. van den Hondel. 1999.
Filamentous fungi as production organisms for glycoproteins of bio-medical
interest. Glycoconj. J. 16:99–107.
27. Mikusova´, K., M. Bela´nova´, J. Kordula´kova´, K. Honda, M. R. McNeil, S.
Mahapatra, D. C. Crick, and P. J. Brennan. 2006. Identification of a novel
galactosyl transferase involved in biosynthesis of the mycobacterial cell wall.
J. Bacteriol. 188:6592–6598.
28. Mohr, J., M. Johnson, T. Cooper, J. S. Lewis, and L. Ostrosky-Zeichner.
2008. Current options in antifungal pharmacotherapy. Pharmacotherapy 28:
614–645.
29. Morelle, W., M. Bernard, J. P. Debeaupuis, M. Buitrago, M. Tabouret, and
J. P. Latge´. 2005. Galactomannoproteins of Aspergillus fumigatus. Eukaryot.
Cell 4:1308–1316.
30. Nassau, P. M., S. L. Martin, R. E. Brown, A. Weston, D. Monsey, M. R.
McNeil, and K. Duncan. 1996. Galactofuranose biosynthesis in Escherichia
coli K-12: identification and cloning of UDP-galactopyranose mutase. J.
Bacteriol. 178:1047–1052.
31. Paisley, D., G. D. Robson, and D. W. Denning. 2005. Correlation between in
vitro growth rate and in vivo virulence in Aspergillus fumigatus. Med. Mycol.
43:397–401.
32. Pan, F., M. Jackson, Y. Ma, and M. McNeil. 2001. Cell wall core galactofu-
ran synthesis is essential for growth of mycobacteria. J. Bacteriol. 183:3991–
3998.
33. Patterson, T. F. 2006. Treatment of invasive aspergillosis: polyenes, echino-
candins, or azoles? Med. Mycol. 44(Suppl 1):357–362.
34. Pedersen, L. L., and S. J. Turco. 2003. Galactofuranose metabolism: a
potential target for antimicrobial chemotherapy. Cell. Mol. Life Sci. 60:259–
266.
35. Perea, S., G. Gonzalez, A. W. Fothergill, W. R. Kirkpatrick, M. G. Rinaldi,
and T. F. Patterson. 2002. In vitro interaction of caspofungin acetate with
voriconazole against clinical isolates of Aspergillus spp. Antimicrob. Agents
Chemother. 46:3039–3041.
36. Pontecorvo, G., J. A. Roper, L. M. Hemmons, K. D. MacDonald, and A. W.
Bufton. 1953. The genetics of Aspergillus nidulans. Adv. Genet. 5:141–238.
37. Punt, P. J., and C. A. van den Hondel. 1992. Transformation of filamentous
fungi based on hygromycin B and phleomycin resistance markers. Methods
Enzymol. 216:447–457.
38. Reichard, U., S. Bu¨ttner, H. Eiffert, F. Staib, and R. Ru¨chel. 1990. Purifica-
tion and characterisation of an extracellular serine proteinase from Aspergil-
lus fumigatus and its detection in tissue. J. Med. Microbiol. 33:243–251.
39. Reijula, K. E. 1991. Two common fungi associated with farmer’s lung: fine
structure of Aspergillus fumigatus and Aspergillus umbrosus. Mycopathologia
113:143–149.
40. Rementeria, A., N. Lo´pez-Molina, A. Ludwig, A. B. Vivanco, J. Bikandi, J.
Ponto´n, and J. Garaizar. 2005. Genes and molecules involved in Aspergillus
fumigatus virulence. Rev. Iberoam. Micol. 22:1–23.
41. Simenel, C., B. Coddeville, M. Delepierre, J. P. Latge´, and T. Fontaine. 2008.
Glycosylinositolphosphoceramides in Aspergillus fumigatus. Glycobiology 18:
84–96.
42. Stynen, D., J. Sarfati, A. Goris, M. C. Pre´vost, M. Lesourd, H. Kamphuis, V.
Darras, and J. P. Latge´. 1992. Rat monoclonal antibodies against Aspergillus
galactomannan. Infect. Immun. 60:2237–2245.
43. Suzuki, E., M. S. Toledo, H. K. Takahashi, and A. H. Straus. 1997. A
monoclonal antibody directed to terminal residue of beta-galactofuranose of
a glycolipid antigen isolated from Paracoccidioides brasiliensis: cross-reactiv-
ity with Leishmania major and Trypanosoma cruzi. Glycobiology 7:463–468.
44. Takayanagi, T., A. Kimura, S. Chiba, and K. Ajisaka. 1994. Novel structures
of N-linked high-mannose type oligosaccharides containing alpha-D-galacto-
furanosyl linkages in Aspergillus niger alpha-D-glucosidase. Carbohydr. Res.
256:149–158.
45. Takayanagi, T., K. Kushida, K. Idonuma, and K. Ajisaka. 1992. Novel
N-linked oligo-mannose type oligosaccharides containing an alpha-D-ga-
lactofuranosyl linkage found in alpha-D-galactosidase from Aspergillus niger.
Glycoconj. J. 9:229–234.
46. Tanaka, N., M. Konomi, M. Osumi, and K. Takegawa. 2001. Characteriza-
tion of a Schizosaccharomyces pombe mutant deficient in UDP-galactose
transport activity. Yeast 18:903–914.
47. Toledo, M. S., S. B. Levery, B. Bennion, L. L. Guimaraes, S. A. Castle, R.
Lindsey, M. Momany, C. Park, A. H. Straus, and H. K. Takahashi. 2007.
Analysis of glycosylinositol phosphorylceramides expressed by the opportu-
nistic mycopathogen Aspergillus fumigatus. J. Lipid Res. 48:1801–1824.
48. Trejo, A. G., J. W. Haddock, G. J. Chittenden, and J. Baddiley. 1971. The
biosynthesis of galactofuranosyl residues in galactocarolose. Biochem. J.
122:49–57.
49. Wallis, G. L., R. L. Easton, K. Jolly, F. W. Hemming, and J. F. Peberdy. 2001.
Galactofuranoic-oligomannose N-linked glycans of alpha-galactosidase A
from Aspergillus niger. Eur. J. Biochem. 268:4134–4143.
50. Weston, A., R. J. Stern, R. E. Lee, P. M. Nassau, D. Monsey, S. L. Martin,
M. S. Scherman, G. S. Besra, K. Duncan, and M. R. McNeil. 1997. Biosyn-
thetic origin of mycobacterial cell wall galactofuranosyl residues. Tuber.
Lung Dis. 78:123–131.
51. Wing, C., J. C. Errey, B. Mukhopadhyay, J. S. Blanchard, and R. A. Field.
2006. Expression and initial characterization of WbbI, a putative D-Galf:
alpha-D-Glc beta-1,6-galactofuranosyltransferase from Escherichia coli
K-12. Org. Biomol. Chem. 4:3945–3950.
52. Yoshida, T., Y. Kato, Y. Asada, and T. Nakajima. 2000. Filamentous fungus
Aspergillus oryzae has two types of alpha-1,2-mannosidases, one of which is a
microsomal enzyme that removes a single mannose residue from
Man9GlcNAc2. Glycoconj. J. 17:745–748.
53. Zhuo, H., H. Hu, L. Zhang, R. Li, H. Ouyang, J. Ming, and C. Jin. 2007.
O-Mannosyltransferase 1 in Aspergillus fumigatus (AfPmt1p) is crucial for
cell wall integrity and conidium morphology, especially at an elevated tem-
perature. Eukaryot. Cell 6:2260–2268.
VOL. 7, 2008 A. FUMIGATUS UDP-GALACTOPYRANOSE MUTASE 1277

Supplementary resource (1)

... Galactofuranose has been shown to be important for the growth and virulence of fungal pathogens (Tefsen et al. 2012;Senicar et al. 2020). Because galactofuranose is not found in mammalian cells, its biosynthetic pathway has been considered an excellent target for the development of antifungal agents (El-Ganiny, Sanders, and Kaminskyj 2008;Schmalhorst et al. 2008). Studies have shown that galactofuranose can be attached to N-linked glycans, O-linked glycans, GPIanchored galactomannans, and to free galactomannans (Senicar et al. 2020;Tefsen et al. 2012). ...
... The first step in the galactofuranose pathway is the conversion of UDG-galactopyranose to UDP-galactofuranose by the enzyme UDPgalactopyranose mutase (Fig. 1B). The gene encoding UDPgalactopyranose mutase has been identified and studied in Aspergillus fumigatus (Bakker et al. 2005;Oppenheimer et al. 2010;Schmalhorst et al. 2008), A. niger (Damveld et al. 2008), and A. nidulans (Alam et al. 2014;El-Ganiny, Sanders, and Kaminskyj 2008). Because of its relevance to human disease, the A. fumigatus UDP-galactopyranose mutase has been extensively studied. ...
... Previous studies have shown the importance of the galactofuranose pathway for growth and virulence in the Aspergilli (Schmalhorst et al. 2008;Komachi et al. 2013;Park et al. 2015;Afroz et al. 2011;Engel, Schmalhorst, and Routier 2012;Engel et al. 2009 show that the loss of galactofuranose in N. crassa not only affects vegetative growth and asexual development (conidiation) (Figs. 2, 3, and 4), but also dramatically affect female development and ascospore formation (Figs. 5 and 6). Our results lend credence to the idea that the galactofuranose pathway would be a good target for the development of antifungal agents. ...
Article
Full-text available
Galactofuranose is a constituent of the cell walls of filamentous fungi. The galactofuranose can be found as a component of N-linked oligosaccharides, in O-linked oligosaccharides, in GPI-anchored galactomannan, and in free galactomannan. The Neurospora genome contains a single UDP-galactose mutase gene (ugm-1/NCU01824) and two UDP-galactofuranose translocases used to import UDP-galactofuranose into the lumen of the Golgi apparatus (ugt-1/NCU01826 and ugt-2/NCU01456). Our results demonstrate that loss of galactofuranose synthesis or its translocation into the lumen of the secretory pathway affects the morphology and growth rate of the vegetative hyphae, the production of conidia (asexual spores), and dramatically affects the sexual stages of the life cycle. In mutants that are unable to make galactofuranose or transport it into the lumen of the Golgi apparatus, ascospore development is aborted soon after fertilization and perithecium maturation is aborted prior to the formation of the neck and ostiole. The Neurospora genome contains three genes encoding possible galactofuranosyltransferases from the GT31 family of glycosyltransferases (gfs-1/NCU05878, gfs-2/NCU07762, and gfs-3/NCU02213) which might be involved in generating galactofuranose-containing oligosaccharide structures. Analysis of triple KO mutants in GT31 glycosyltransferases shows that these mutants have normal morphology, suggesting that these genes do not encode vital galactofuranosyltransferases.
... The following strains were used in this study: A. fumigatus ATCC46645 [25], A. fumigatus CEA17 [26], A. fumigatus Af293 [27,28], A. fumigatus D141 [29], A. fumigatus AfS35 [30] and GFP-expressing AfS35 [31], A. fumigatus D141∆glfA (unable to produce galactofuranose) and the complemented strain D141∆glfA+glfA [32], A. fischeri DSM 3700 [33], A. terreus T9 [34], A. terreus SBUG 844 [17,19] Strains were routinely grown on Sabouraud agar or in Sabouraud liquid medium at 37 °C or 30 °C. When indicated, the following other media were used: Aspergillus minimal medium (AMM) (0.52 g KCl, 0.52 g MgSO4 × 7H2O, 1.52 g KH2PO4, 1 mL essential element solution (40 mg Na2B4O7 × 10H2O, 400 mg CuSO4 × 3H2O, 800 mg FePO4 × H2O, 800 mg MnSO4 × 4H2O, 800 mg Na2MoO4 × 2H2O, 8 g ZnSO4 × 7H2O; ad 1 l; autoclaved), 10 g glucose, sodium nitrate 6 g, pH 6.5; ad 1 l), yeast glucose (YG) medium (5 g/L yeast extract, 20 g/L glucose), RPMI-1640 (Gibco, Thermo Fisher Scientific) buffered with 20 mM HEPES, BSA medium (PBS supplemented with 2 g/L bovine serum albumin) and synthetic defined (SD) medium (1.7 g/L yeast nitrogen base with 5 g/L ammonium sulfate and 2 g/L D-glucose). ...
... At higher magnification, an overlay of both channels demonstrates that the spatial distributions of the two antigens are similar, but not identical ( Figure 4D). The partial co-localization of both antigens prompted us to stain a ∆glfA mutant that is unable to produce galactofuranose, an essential component of fungal galactomannan [32]. As shown in Figure 5, hyphae of the ∆glfA mutant showed a distribution of the AB90-E8 antigen that was clearly distinct from that observed for the complemented mutant and the corresponding wild type strain D141. ...
Article
Full-text available
In most cases, invasive aspergillosis (IA) is caused by A. fumigatus, though infections with other Aspergillus spp. with lower susceptibilities to amphotericin B (AmB) gain ground. A. terreus, for instance, is the second leading cause of IA in humans and of serious concern because of its high propensity to disseminate and its in vitro and in vivo resistance to AmB. An early differentiation between A. fumigatus and non-A. fumigatus infections could swiftly recognize a potentially ineffective treatment with AmB and lead to the lifesaving change to a more appropriate drug regime in high-risk patients. In this study, we present the characteristics of the monoclonal IgG1-antibody AB90-E8 that specifically recognizes a surface antigen of A. fumigatus and the closely related, but not human pathogenic A. fischeri. We show immunostainings on fresh frozen sections as well as on incipient mycelium picked from agar plates with tweezers or by using the expeditious tape mount technique. All three methods have a time advantage over the common procedures currently used in the routine diagnosis of IA and outline the potential of AB90-E8 as a rapid diagnostic tool.
... Great efforts have been made to establish the structure of GM (1, 3), but the analysis of GIPCs and their structural characterization using mass spectrometry remains highly challenging and time-consuming (5-7). Indeed, multiple enrichment steps such as organics solvent extractions, partitions and hydrophobic interaction chromatography are required purification of GIPCs prior to MALDI-TOF mass spectrometry (8)(9)(10)(11)(12). MALDI-TOF is a "soft-ionization" technique which ionizes large biomolecules while keeping the original structure intact (13). ...
Article
Full-text available
Glycosyl-Inositol-Phospho-Ceramides (GIPCs) or glycosylphosphatidylinositol-anchored fungal polysaccharide are major lipids in plant and fungal plasma membranes and play an important role in stress adaption. However, their analysis remains challenging due to...
... Galactomannan contributes to fungal cell wall architecture and function (Fontaine et al., 2000(Fontaine et al., , 2011Gravelat et al., 2013;Lamarre et al., 2009;Lee et al., 2014). In accordance with these major traits, the finding that uge3 and glfA/ugm1 mutants displayed attenuated virulence in a mouse model of invasive aspergillosis it was not unexpected (Gravelat et al., 2013;Schmalhorst et al., 2008). Collectively, in the human pathogen A. fumigatus, the roles of the galactosecontaining cell wall polymers in cell wall function and virulence are excellently documented . ...
Article
Full-text available
Fungal cell walls represent the frontline contact with the host and play a prime role in pathogenesis. While the roles of the cell wall polymers like chitin and branched β‐glucan are well understood in vegetative and pathogenic development, that of the most prominent galactose‐containing polymers galactosaminogalactan and fungal‐type galactomannan is unknown in plant pathogenic fungi. Mining the genome of the maize pathogen Colletotrichum graminicola identified the single‐copy key galactose metabolism genes UGE1 and UGM1, encoding a UDP‐glucose‐4‐epimerase and UDP‐galactopyranose mutase, respectively. UGE1 is thought to be required for biosynthesis of both polymers, whereas UGM1 is specifically required for fungal‐type galactomannan formation. Promoter:eGFP fusion strains revealed that both genes are expressed in vegetative and in pathogenic hyphae at all stages of pathogenesis. Targeted deletion of UGE1 and UGM1, and fluorescence‐labeling of galactosaminogalactan and fungal‐type galactomannan confirmed that Δuge1 mutants were unable to synthesize either of these polymers, and Δugm1 mutants did not exhibit fungal‐type galactomannan. Appressoria of Δuge1, but not of Δugm1 mutants, were defective in adhesion, highlighting a function of galactosaminogalactan in the establishment of these infection cells on hydrophobic surfaces. Both Δuge1 and Δugm1 mutants showed cell wall defects in older vegetative hyphae and severely reduced appressorial penetration competence. On intact leaves of Zea mays, both mutants showed strongly reduced disease symptom severity, indicating that UGE1 and UGM1 represent novel virulence factors of C. graminicola.
... In a paper disk assay, the fludioxonil-induced inhibition zones of the wild type and the Δgtb3 mutant were comparable ( Fig. 12A and B). We also tested a ∆glfA mutant, which lacks an essential gene for the biosynthesis of galactomannan [35]. The inhibition zones of this mutant were clearly larger than those of the wild type ( Fig. 12A and Since many results of this study indicate that the antifungal activity of fludioxonil is tightly linked to the cell wall architecture, we analyzed whether the cell wall targeting drugs caspofungin and nikkomycin Z interfere with the antifungal activity of fludioxonil. ...
Article
Full-text available
Background Aspergillus fumigatus is a major fungal pathogen that causes severe problems due to its increasing resistance to many therapeutic agents. Fludioxonil is a compound that triggers a lethal activation of the fungal-specific High Osmolarity Glycerol pathway. Its pronounced antifungal activity against A. fumigatus and other pathogenic molds renders this agent an attractive lead substance for the development of new therapeutics. The group III hydride histidine kinase TcsC and its downstream target Skn7 are key elements of the multistep phosphorelay that represents the initial section of the High Osmolarity Glycerol pathway. Loss of tcsC results in resistance to fludioxonil, whereas a Δskn7 mutant is partially, but not completely resistant. Results In this study, we compared the fludioxonil-induced transcriptional responses in the ΔtcsC and Δskn7 mutant and their parental A. fumigatus strain. The number of differentially expressed genes correlates well with the susceptibility level of the individual strains. The wild type and, to a lesser extend also the Δskn7 mutant, showed a multi-faceted stress response involving genes linked to ribosomal and peroxisomal function, iron homeostasis and oxidative stress. A marked difference between the sensitive wild type and the largely resistant Δskn7 mutant was evident for many cell wall-related genes and in particular those involved in the biosynthesis of chitin. Biochemical data corroborate this differential gene expression that does not occur in response to hyperosmotic stress. Conclusions Our data reveal that fludioxonil induces a strong and TcsC-dependent stress that affects many aspects of the cellular machinery. The data also demonstrate a link between Skn7 and the cell wall reorganizations that foster the characteristic ballooning and the subsequent lysis of fludioxonil-treated cells.
... In A. fumigatus, galactofuranose serves as a significant constituent of the cell wall and surface glycans. The deletion of the UGM gene, which encodes UDP-galactopyranose mutase, reduces virulence, promotes cell morphology defects and increases the fungus's susceptibility to antifungal agents [102][103][104]. ...
Article
Full-text available
Members of the Paracoccidioides complex are the causative agents of Paracoccidioidomycosis (PCM), a human systemic mycosis endemic in Latin America. Upon initial contact with the host, the pathogen needs to uptake micronutrients. Nitrogen is an essential source for biosynthetic pathways. Adaptation to nutritional stress is a key feature of fungi in host tissues. Fungi utilize nitrogen sources through Nitrogen Catabolite Repression (NCR). NCR ensures the scavenging, uptake and catabolism of alternative nitrogen sources, when preferential ones, such as glutamine or ammonium, are unavailable. The NanoUPLC-MSE proteomic approach was used to investigate the NCR response of Paracoccidioides lutzii after growth on proline or glutamine as a nitrogen source. A total of 338 differentially expressed proteins were identified. P. lutzii demonstrated that gluconeogenesis, β-oxidation, glyoxylate cycle, adhesin-like proteins, stress response and cell wall remodeling were triggered in NCR-proline conditions. In addition, within macrophages, yeast cells trained under NCR-proline conditions showed an increased ability to survive. In general, this study allows a comprehensive understanding of the NCR response employed by the fungus to overcome nutritional starvation, which in the human host is represented by nutritional immunity. In turn, the pathogen requires rapid adaptation to the changing microenvironment induced by macrophages to achieve successful infection.
Article
Full-text available
Aspergillus fumigatus represents a public health problem due to the high mortality rate in immunosuppressed patients and the emergence of antifungal-resistant isolates. Protein acetylation is a crucial post-translational modification that controls gene expression and biological processes. The strategic manipulation of enzymes involved in protein acetylation has emerged as a promising therapeutic approach for addressing fungal infections. Sirtuins, NAD⁺-dependent lysine deacetylases, regulate protein acetylation and gene expression in eukaryotes. However, their role in the human pathogenic fungus A. fumigatus remains unclear. This study constructs six single knockout strains of A. fumigatus and a strain lacking all predicted sirtuins (SIRTKO). The mutant strains are viable under laboratory conditions, indicating that sirtuins are not essential genes. Phenotypic assays suggest sirtuins’ involvement in cell wall integrity, secondary metabolite production, thermotolerance, and virulence. Deletion of sirE attenuates virulence in murine and Galleria mellonella infection models. The absence of SirE alters the acetylation status of proteins, including histones and non-histones, and triggers significant changes in the expression of genes associated with secondary metabolism, cell wall biosynthesis, and virulence factors. These findings encourage testing sirtuin inhibitors as potential therapeutic strategies to combat A. fumigatus infections or in combination therapy with available antifungals.
Article
Colletotrichum graminicola, the causal agent of maize leaf anthracnose and stalk rot, differentiates a pressurized infection cell called an appressorium in order to invade the epidermal cell, and subsequently forms biotrophic and necrotrophic hyphae to colonize the host tissue. While the role of force in appressorial penetration is established (Bechinger et al., 1999), the involvement of cell wall-degrading enzymes (CWDEs) in this process and in tissue colonization is poorly understood, due to the enormous number and functional redundancy of these enzymes. The serine/threonine protein kinase gene SNF1 identified in Sucrose Non-Fermenting yeast mutants mediates de-repression of catabolite-repressed genes, including many genes encoding CWDEs. In this study, we identified and functionally characterized the SNF1 homolog of C. graminicola. Δsnf1 mutants showed reduced vegetative growth and asexual sporulation rates on media containing polymeric carbon sources. Microscopy revealed reduced efficacies in appressorial penetration of cuticle and epidermal cell wall, and formation of unusual medusa-like biotrophic hyphae by Δsnf1 mutants. Severe and moderate virulence reductions were observed on intact and wounded leaves, respectively. Employing RNA-sequencing we show for the first time that more than 2,500 genes are directly or indirectly controlled by Snf1 in necrotrophic hyphae of a plant pathogenic fungus, many of which encode xylan- and cellulose-degrading enzymes. The data presented show that Snf1 is a global regulator of gene expression and is required for full virulence.
Article
The impact of the cell wall structure of Monascus purpureus M9 on the secretion of extracellular monascus pigments (exMPs) was investigated. To modify the cell wall structure, UDP-galactopyranose mutase (GlfA) was knocked out using Agrobacterium-mediated transformation method, leading to a significant reduction in the Galf-based polysaccharide within the cell wall. Changes in mycelium morphology, sporogenesis, and the expression of relevant genes in M9 were also observed following the mutation. Regarding MPs secretion, a notable increase was observed in six types of exMPs (R1, R2, Y1, Y2, O1 and O2). Specifically, these exMPs exhibited enhancement of 1.33, 1.59, 0.8, 2.45, 2.89 and 4.03 times, respectively, compared to the wild-type strain. These findings suggest that the alteration of the cell wall structure could selectively influence the secretion of MPs in M9. The underlying mechanisms were also discussed. This research contributes new insights into the regulation of the synthesis and secretion of MPs in Monascus spp..
Article
Analysis of glycans remains a difficult task due to their isomeric complexity. Despite recent progress, determining monosaccharide ring size, a type of isomerism, is still challenging due to the high flexibility of the five-membered ring (also called furanose). Galactose is a monosaccharide that can be naturally found in furanose configuration in plant and bacterial polysaccharides. In this study, we used the coupling of tandem mass spectrometry and infrared ion spectroscopy (MS/MS-IR) to investigate compounds containing galactofuranose and galactopyranose. We report the IR fingerprints of monosaccharide fragments and demonstrate for the first time galactose ring-size memory upon collision-induced dissociation (CID) conditions. The linkage of the galactose unit is further obtained by analyzing disaccharide fragments. These findings enable two possible applications. First, labeled oligosaccharide patterns can be analyzed by MS/MS-IR, yielding full sequence information, including the ring size of the galactose unit; second, MS/MS-IR can be readily applied to unlabeled oligosaccharides to rapidly identify the presence of a galactofuranose unit, as a standalone analysis or prior to further sequencing.
Article
Full-text available
The availability of new antifungal agents has multiplied the demand for in vitro antifungal susceptibility testing for Aspergillus spp. The European Committee on Antimicrobial Susceptibility Testing (EUCAST) has charged its Antifungal Susceptibility Testing Subcommittee (AFST-EUCAST) with the preparation of new guidelines for in vitro susceptibility testing of antifungals against Aspergillus spp (EUCAST-AST-ASPERGILLUS). This committee has modified the reference method for broth dilution antifungal susceptibility testing of filamentous fungi (M38-A) as follows: (i) RPMI 1640 supplemented with 2% glucose (RPMI 2%G) as assay medium, (ii) inoculum preparation by conidium counting in a haemocytometer and (iii) an inoculum size of 2×10(5)-5×10(5) CFU/ml. The incubation time is about 48 h at 35°C and MIC is read visually. The MIC value is a no-growth visual endpoint. The standard method described herein is intended to provide a valid and economic method for testing the susceptibility to antifungal agents of Aspergillus spp., to identify resistance, and to facilitate an acceptable degree of conformity, e.g. agreement within specified ranges and between laboratories in measuring the susceptibility.
Article
Full-text available
Aspergillus fumigatus causes a wide range of diseases that include mycotoxicosis, allergic reactions and systemic diseases (invasive aspergillosis) with high mortality rates. Pathogenicity depends on immune status of patients and fungal strain. There is no unique essential virulence factor for development of this fungus in the patient and its virulence appears to be under polygenetic control. The group of molecules and genes associated with the virulence of this fungus includes many cell wall components, such as ß(1-3)-glucan, galactomannan, galactomannanproteins (Afmp1 and Afmp2), and the chitin synthetases (Chs; chsE and chsG), as well as others. Some genes and molecules have been implicated in evasion from the immune response, such as the rodlets layer (rodA/hyp1 gene) and the conidial melanin-DHN (pksP/alb1 gene). The detoxifying systems for Reactive Oxygen Species (ROS) by catalases (Cat1p and Cat2p) and superoxide dismutases (MnSOD and Cu,ZnSOD), had also been pointed out as essential for virulence. In addition, this fungus produces toxins (14 kDa diffusible substance from conidia, fumigaclavin C, aurasperon C, gliotoxin, helvolic acid, fumagilin, Asp-hemolysin, and ribotoxin Asp fI/mitogilin F/restrictocin), allergens (Asp f1 to Asp f23), and enzymatic proteins as alkaline serin proteases (Alp and Alp2), metalloproteases (Mep), aspartic proteases (Pep and Pep2), dipeptidyl-peptidases (DppIV and DppV), phospholipase C and phospholipase B (Plb1 and Plb2). These toxic substances and enzymes seems to be additive and/or synergistic, decreasing the survival rates of the infected animals due to their direct action on cells or supporting microbial invasion during infection. Adaptation ability to different trophic situations is an essential attribute of most pathogens. To maintain its virulence attributes A. fumigatus requires iron obtaining by hydroxamate type siderophores (ornitin monooxigenase/SidA), phosphorous obtaining (fos1, fos2, and fos3), signal transductional falls that regulate morphogenesis and/or usage of nutrients as nitrogen (rasA, rasB, rhbA), mitogen activated kinases (sakA codified MAP-kinase), AMPc-Pka signal transductional route, as well as others. In addition, they seem to be essential in this field the amino acid biosynthesis (cpcA and homoaconitase/lysF), the activation and expression of some genes at 37 °C (Hsp1/Asp f12, cgrA), some molecules and genes that maintain cellular viability (smcA, Prp8, anexins), etc. Conversely, knowledge about relationship between pathogen and immune response of the host has been improved, opening new research possibilities. The involvement of non-professional cells (endothelial, and tracheal and alveolar epithelial cells) and professional cells (natural killer or NK, and dendritic cells) in infection has been also observed. Pathogen Associated Molecular Patterns (PAMP) and Patterns Recognizing Receptors (PRR; as Toll like receptors TLR-2 and TLR-4) could influence inflammatory response and dominant cytokine profile, and consequently Th response to infection. Superficial components of fungus and host cell surface receptors driving these phenomena are still unknown, although some molecules already associated with its virulence could also be involved. Sequencing of A. fumigatus genome and study of gene expression during their infective process by using DNA microarray and biochips, promises to improve the knowledge of virulence of this fungus.
Article
Full-text available
A mouse monoclonal antibody, MEST-1, was produced against Band 1 glycolipid antigen of Paracoccidioides brasiliensis. The glycan structure of Band 1 antigen was recently elucidated and the monosaccharides sequence was defined as: Gal/pi->6(Man/;al-»3)ManpBl->2Ins. The reactivity of MEST-1 MAb was determined by solid-phase radioimmunoassay and high performance thin layer chro- matography immunostaining. Selective oxidation of galac- tofuranose residues and inhibition assays with different methyl-glycosides, revealed that MAb MEST-1 is directed against the terminal residue of P-D-galactofu ranose of Band 1, a phosphoglyceroglycolipid antigen of P.brasilien- sis. By indirect immunofluorescence, it was observed that the epitope recognized by MEST-1 is accessible to the an- tibody in yeast forms of this fungus. Reactivity of MEST-1 with parasites known to express galactofuranose contain- ing glycoconjugates was also analyzed by indirect immu- nofluorescence. A positive fluorescence was observed with promastigotes of Leishmania major and epimastigotes of Trypanosoma cruzL GEPL-1 was identified as the antigen recognized by MEST-1 in Leishmania major, indicating that the MAb MEST-1 recognizes terminal galactofuranose residue in either 01—>6 or pi—>3 linkage to the mannose.
Article
The galactomannan (GM) produced extracellularly by Aspergillus fumigatus has been purified by a double sequential hydrazine-nitrous acid treatment of the ethanol precipitate of the culture filtrate. Nuclear magnetic resonance and gas-liquid chromatography-mass spectrometry analysis have been performed on intact GM, acid-hydrolyzed GM, and oligomers resulting from the acetolysis of the acid-hydrolyzed GM. Results show that A. fumigatus GM is composed of a linear mannan core with an alpha-(1-2)-linked mannotetraose repeating unit attached via alpha-(1-6) linkage. Side chains composed of an average of 4 to 5 beta-(1-5)-galactofuranose units are linked to C-6 and C-3 positions of alpha-(1-2)-linked mannose units of the mannan. The immunoreactivity of GM and HCl-hydrolyzed GM was studied by use of human sera from aspergillosis patients and an antigalactofuran monoclonal antibody. The alpha-(1-2) (1-6)-mannan core is not antigenic. The immunogenic galactofuran is found amongst several exocellular glycoproteins. According to a direct enzyme-linked immunosorbent assay with GM as the detector antigen, only 26% of the serum samples from aspergilloma patients (all positive by immunodiffusion assays) give optical density values superior to a cutoff estimated as the mean +/- 3 standard deviations of values obtained with control sera.
Article
Three classes of antifungals Á the polyenes, the echinocandins, and the extended spectrum azoles Á are now available for treating invasive aspergillosis (IA). New agents and formulations in these classes offer the possibility of decreased toxicity and improved outcomes. With the availability of newer antifungals, clinicians are challenged to understand the advantages and limitations of these new choices. Standard amphotericin B deoxycholate is associated with poor outcomes in addition to unacceptable toxicity and is no longer recommended as primary therapy for most patients. Lipid formulations of amphotericin reduce toxicity, but because of cost and toxicity concerns, they may not be used at optimal doses. Interestingly, a recent trial showed that initial use of higher doses of liposomal amphotericin at 10 mg kg (1 d (1 did not improve efficacy and was associated with more toxicity. Due to lack of complete killing or inhibition of Aspergillus, the echinocandins are not frequently used as primary therapy for aspergillosis, although their minimal toxicity and potential for combination therapy remains attractive. The newer triazoles, including voriconazole and posaconazole, offer fungicidal activity against Aspergillus and, for voriconazole, both intravenous as well as oral therapy. Voriconazole was compared with amphotericin B followed by other licensed therapy in a global trial that showed better outcomes and improved survival so that voriconazole is recommended as primary therapy for most patients with this disease. These studies also show that early, effective therapy is a key factor for a successful outcome. Consideration of risk for IA and early initiation of therapy may improve outcomes in this often lethal infection.
Article
The fine structure of Aspergillus fumigatus and Aspergillus umbrosus by transmission electron microscopy (TEM) is described. The fine structure of the ascosporic and asexual stages of A. umbrosus is described for the first time. Dense, homogenous material and fibers were detected on the outer hyphal cell wall of the Aspergilli. Septal pores were found in the hypha of A. umbrosus. Two wall layers were detected in the cell wall of the conidia of the both Aspergilli. The ascospores of A. umbrosus contained thick cell wall and the surface of which was smoother than that of the conidia.
Article
-Mannosidase activities towards high-mannose oligosaccharides were examined with a detergent-solubilized microsomal preparation from a filamentous fungus, Aspergillus oryzae. In the enzymatic reaction, the pyridylaminated substrate Man9GlcNAc2-PA was trimmed to Man8GlcNAc2-PA which lacked one -1,2-mannose residue at the nonreducing terminus of the middle branch (Man8B isomer), and this mannooligosaccharide remained predominant through the overall reaction. Trimming was optimal at pH 7.0 in PIPES buffer in the presence of calcium ion and kifunensine was inhibitory with IC50 below 0.1[emsp4 ]M. These results suggest that the activity is the same type as was previously observed with human and yeast endoplasmic reticulum (ER) -mannosidases. Considering these results together with previous data on a fungal -1,2-mannosidase that trimmed Man9GlcNAc2 to Man5GlcNAc2 (Ichishima, E., et al. (1999) bit>Biochem J, 339: 589–597), the filamentous fungi appear to have two types of -1,2-mannosidases, each of which acts differently on N-linked mannooligosaccharides.
Article
Filamentous fungi are commonly used in the fermentation industry for large scale production of glycoproteins. Several of these proteins can be produced in concentrations up to 20-40 g per litre. The production of heterologous glycoproteins is at least one or two orders of magnitude lower but research is in progress to increase the production levels. In the past years the structure of protein-linked carbohydrates of a number of fungal proteins has been elucidated, showing the presence of oligo-mannosidic and high-mannose chains, sometimes with typical fungal modifications. A start has been made to engineer the glycosylation pathway in filamentous fungi to obtain strains that show a more mammalian-like type of glycosylation. This mini review aims to cover the current knowledge of glycosylation in filamentous fungi, and to show the possibilities to produce glycoproteins with these organisms with a more mammalian-like type of glycosylation for research purposes or pharmaceutical applications.
Article
Structures of oligosaccharides from Aspergillus niger alpha-D-galactosidase [EC 3.2.1.22] were studied. Purified alpha-D-galactosidase was treated with N-glycosidase F, and six kinds of oligosaccharides were isolated by gel chromatography and anion-exchange chromatography. The structures of the oligosaccharides were determined by 1H-NMR and compositional analysis to be Man5GlcNAc2, Man6GlcNAc2, Man9GlcNAc2, GlcMan9GlcNAc2, GalMan4GlcNAc2 and GalMan5GlcNAc2. From mild acid hydrolysis, methylation analysis and ROESY spectral analysis, it was ascertained that the galactosyl residue in two oligosaccharides was in the furanose form and was bound to mannose at the nonreducing end with an alpha 1-2 linkage (GalfMan4GlcNAc2 and GalfMan5GlcNAc2).