ArticlePDF AvailableLiterature Review

Antifungal agents of use in animal health - Chemical, biochemical and pharmacological aspects

Authors:
  • Janssen Pharmaceutica, Johnson & Johnson
  • Janssen Pharmaceutica Beerse Belgium

Abstract and Figures

A limited number of antifungal agents is licensed for use in animals, however, many of those available for the treatment of mycoses in humans are used by veterinary practitioners. This review includes chemical aspects, spectra of activity, mechanisms of action and resistance, adverse reactions and drug interactions of the antifungals in current use.
Content may be subject to copyright.
Antifungal agents of use in animal health – chemical, biochemical
and pharmacological aspects
H. VANDEN BOSSCHE*
M. ENGELEN
&
F. ROCHETTE
*Steenweg op Gierle, Turnhout, Belgium;
Janssen Animal Health B.V.B.A.,
Turnhoutsebaan, Beerse, Belgium
Vanden Bossche, H., Engelen, M., Rochette, F. Antifungal agents of use in
animal health – chemical, biochemical and pharmacological aspects. J vet.
Pharmacol. Therap. 26, 5–29.
A limited number of antifungal agents is licensed for use in animals, however,
many of those available for the treatment of mycoses in humans are used by
veterinary practitioners. This review includes chemical aspects, spectra of
activity, mechanisms of action and resistance, adverse reactions and drug
interactions of the antifungals in current use.
(Paper received 12 July 2002; accepted for publication 29 November 2002)
Hugo Vanden Bossche, Steenweg op Gierle, 68, B2300 Turnhout, Belgium. E-mail:
hugo.vdbossche@skynet.be
INTRODUCTION
For most mycoses veterinary practitioners have a variety of
antifungal agents at their disposal. An outlook of the different
classes is given in Table 1. The earliest known compound with
proven antifungal activity is griseofulvin – the product of
Penicillium species. The spectrum of activity of this fungicidal
agent is limited to dermatophytes (reviewed by Polak, 1990).
The polyene macrolides, natural products of Streptomyces
species, are a class of poorly absorbed, large macrocyclic polyke-
tides that interact with membrane sterols. The antifungal polyenes
were the earliest broad-spectrum antifungal agents introduced for
clinical use (Odds, 1996). Amphotericin B (AmB), candicidin,
nystatin and natamycin (pimaricin) have found use in the topical
treatment of superficial infections of men and or animals. AmB is
mainly formulated for intravenous administration.
An amazing number of antifungal agents are found within the
group of azole-containing compounds. A great number were
breakthroughs in antifungal therapy in both human and
veterinary medicine, and some of them show excellent activity
against phytopathogenic fungi. Examples are the imidazole
derivatives bifonazole, clotrimazole, econazole, enilconazole
(imazalil), ketoconazole, miconazole and parconazole, and the
triazole derivatives terconazole, itraconazole, fluconazole and
propiconazole. Most of the imidazole derivatives (e.g. miconazole,
clotrimazole and econazole) are available for topical application
only, enilconazole and ketoconazole are the exceptions. Enil-
conazole has been developed as a smoke generator and as a
spray formulation (Desplenter, 1988). The dioxolane imidazole
derivative ketoconazole is the first compound of the azole class
that is systemically absorbed after oral administration (Thien-
pont et al., 1979; Daneshmend & Warnock, 1988), and is used
both orally and topically. Fluconazole and itraconazole are used
orally and intravenously. The antifungal activity of the azole
derivatives arises from interaction with the sterol-14a-demethy-
lase, involved in the biosynthesis of ergosterol.
Other groups of antifungal agents, the allylamines (naftifine
and terbinafine) (Ryder, 1988; Ryder & Mieth, 1992) and the
thiocarbamates (e.g. tolnaftate) (Barrett-Bee et al., 1986; Ryder
et al., 1986) perturb fungal sterol synthesis by inhibiting the
squalene epoxidase. The use of the allylamines and the much older
tolnaftate is essentially restricted to dermatophyte infections.
The benzimidazoles, such as carbendazim, fuberidazole, and
thiabendazole are effective at relatively low doses for the
inhibition of a broad range of plant pathogenic fungi (literature
reviewed by Delp, 1995).
This review includes chemical, biochemical and pharmacolo-
gical aspects of the most interesting antifungal agents registered
for the treatment of fungal infections in animals as well as those
used off label. In the next issue of this journal the clinical aspects
of the most interesting antifungal agents and their applications
in animals are reviewed.
GRISEOFULVIN
Although griseofulvin (Fig. 1) was isolated from Penicillium
griseofulvum as early as 1939 (Oxford et al., 1939), its action
against experimental ringworm in guinea-pigs was first des-
cribed in 1956 (Gentles, 1956). The use of griseofulvin as an
orally active antifungal against ringworm in man was first
described by Williams et al. (1958).
Griseofulvin is practically insoluble in water, slightly soluble in
ethanol and soluble in dimethylformamide (DMF). It is effective
after oral ingestion and reaches the stratum corneum. Fat and
reduction of the particle size enhances the absorption of
griseofulvin (Lin et al., 1973). The compound is metabolized by
the liver. Chang et al. (1975) compared the metabolism of
14
C-
griseofulvin in rat liver microsomes, isolated perfused rat livers,
and rats with bile duct cannulas. In all three models,
4-desmethylgriseofulvin and 6-desmethylgriseofulvin were the
major metabolites. In isolated perfused liver, most of the drug
J. vet. Pharmacol. Therap. 26, 5–29, 2003. REVIEW
2003 Blackwell Publishing Ltd 5
(59% of dose) was excreted into bile within 4 h, primarily as
4-desmethylgriseofulvin and 6-desmethylgriseofulvin. In animals
with bile duct cannulas, 65% of the dose was excreted into bile
and 18% into urine.
Mode of action
Griseofulvin inhibits the growth of dermatophytes by preventing
microtubules from sliding and thus inhibits mitosis (Crackower,
1972; Gull & Trinci, 1973; Langcake et al., 1983). Griseofulvin
also affects the cytoplasmic microtubules. These are involved in
intracellular transport of endogenous compounds. Polak (1990)
hypothesized that this interaction results in disturbance of the
transport of newly synthesized cell wall constituents and may
lead to the distortion of the hyphae, the ‘curling’ effect. In a more
recent study, Chaudhuri and Luduena (1996) proved that
griseofulvin interacts with tubulin and changes its conformation,
but this mycotoxin ‘binds’ to another site than colchicine, a well-
known inhibitor of tubulin polymerization.
Spectrum of activity
The spectrum of activity of griseofulvin is limited to Trichophyton,
Epidermophyton and Microsporum species. This is due to the fact
that dermatophytes possess a prolonged energy-dependent
transport system for this antibiotic, whereas in insensitive cells,
such as Candida albicans, this is replaced by a short energy-
independent transport system (Polak, 1990).
Differences, depending on the test method used, have to be
taken into account when comparing minimum inhibitory
concentration (MIC) data in the literature. For example,
Niewerth et al. (1998) compared the susceptibility of dermato-
phyte strains to griseofulvin, itraconazole, sertaconazole, terbin-
afine and cyclopiroxolamine by using an agar macrodilution and
broth microdilution test. For all five antimycotics, MIC data were
three- to seven-fold lower in the microdilution test system.
Korting et al. (1995) used the microdilution test to compare
the susceptibility of T. rubrum and T. mentagrophytes. The
O
CH3CH3
NH2
CH3
CH3
CH3
CH3
CH3
CH3
CH3
H3C
H3C
H3C
H3C
H2N
H2N
O
Cl
O
O
O
OGriseofulvin Natamycin
O
HO OH
O
O
O
OHO
OH
HO
COOH
H
O
O
HO
HO
O
OHOH OH
OH
HO
OH
OH
OH
COOH
OH
Nystatin A1
O
O
HO
HO
O
OHOH
CH3
OH
HO
O
O
H
H
OH
OH
OH
OH
COOH
Amphotericin B
Fig. 1. Chemical structures of griseofulvin
and the polyenes.
Table 1. Antifungal drugs
Chemical classes Drugs* Targets
Griseofulvin (o) Inhibits microtubule sliding
Polyenes Amphotericin B (iv, t, o) Membrane barrier function
Natamycin (t)
Nystatin (t, o)
Azoles Clotrimazole (t) Ergosterol biosynthesis
Imidazoles Econazole (t) 14a-demethylase
Enilconazole (t) CYP51
Ketoconazole (t, o)
Miconazole (t)
Parconazole (o)
Triazoles Fluconazole (o, iv)
Itraconazole (o, iv)
Allylamines Terbinafine (t, o) Squalene epoxidase
Thiocarbamates Tolnaftate (t)
Benzimidazoles Thiabendazole (o, t) Microtubule assembly
Pyrimidine Flucytosine (o, iv) DNA and RNA synthesis
*o, oral; t, topical; iv, intravenous.
6H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
concentration of griseofulvin that inhibited 90% of isolates of
T. mentagrophytes was 10 mg L, whereas for T. rubrum this was
achieved at 3 mg L. More recently Jessup et al. (2000a) used
oatmeal cereal agar medium to determine the antifungal
susceptibilities of T. rubrum to fluconazole, itraconazole, terbin-
afine and griseofulvin. The mean ± SE of the mean MICs
were 2.07 ± 0.29, 0.13 ± 0.01, 0.002 ± 0.0003 and 0.71 ±
0.05 mg L, respectively. Using a broth macrodilution method,
Perea et al. (2001a) found for M. canis and M. gypseum an
MIC50 of 0.5 mg L, whereas for T. equinum and T. verrucosum
MIC50s of 8 and 5.66 mg L were found.
Resistance of dermatophytes to griseofulvin is rare. However,
in the study of Korting et al. (1995) 11 of the 32 T. rubrum
strains tested showed an MIC of 3 mg L or greater. According to
Artis et al. (1981) such MIC values indicate resistance to
griseofulvin.
Drug interactions adverse reactions
Griseofulvin has been shown to induce drug-metabolizing
enzymes (Okey et al., 1986). For example, it increases the rates
of warfarin metabolism. Serum levels of griseofulvin are
decreased by coadministration of phenobarbital. It has been
reported that administration of this antibiotic markedly stimu-
lates the activity of 5-aminolevulinate synthase, the rate-limiting
enzyme of porphyrin synthesis, and causes a rapid inhibition of
mitochondrial porphyrin-metal chelatase and may thereby
inhibit the synthesis of haem in the liver (De Matteis, 1982).
Embryotoxic and teratogenic effects of griseofulvin have been
demonstrated in pregnant rats exposed during organogenesis
(De Carli & Larizza, 1988).
POLYENES
Polyene macrolides are a class of poorly absorbed, large
macrocyclic polyketides that interact with membrane sterols.
They are produced by species of Streptomyces, and characterized
by large lactone rings, containing three to eight conjugated
double bonds, which are generally combined with one sugar
moiety (Hamilton-Miller, 1973; Ryley et al., 1981). The polyenes
can be subdivided into trienes (rapamycin), tetraenes (natamy-
cin), pentaenes (pentamycin), heptaenes (AmB and candicidin)
and a pseudo-heptaene tetraene, nystatin. They are effective
against eukaryotic organisms and those prokaryotic organisms
containing sterols in their membranes (Kerridge & Whelan,
1984). The antifungal polyenes AmB, candicidin, nystatin and
natamycin (pimaricin) have found use in the topical treatment of
superficial infections of men and or animals.
Mechanism of action
The toxicity of the polyenes to fungal cells is thought to occur as a
result of its binding to sterols incorporated into cell membranes
changing as such the physical state of the membrane, causing an
impairment of membrane functions, resulting in an enhanced
permeability to protons and leakage of internal constituents such
as K
+
,Ca
2+
and PO
4
3–
(Bolard, 1986; Vanden Bossche et al.,
1987a; Brajtburg et al., 1990; Vanden Bossche, 1995). The
addition of AmB to yeast induced an increased permeability for
protons resembling the increased proton influx with the proton
conductor dinitrophenol. This collapses the proton gradient and
therefore the uptake of nutrients. The effects of polyenes on the
proton gradient may result from an interaction with the proton-
translocating adenosine triphosphatase (H
+
-ATPase) found to be
present in the plasma membrane of, for example, Saccharomyces
cerevisiae (Goffeau & Slayman, 1981), C. albicans (Hubbard et al.,
1986), Neurospora crassa (Rao et al., 1988), Fusarium oxysporum
(Brandao et al., 1992), C. neoformans (Soteropoulos et al., 2000)
and Aspergillus fumigates (Burghoorn et al., 2002). After solubi-
lization and purification, the H
+
-ATPase requires phospholipids
to provide sufficient fluidity for maximal activity (Goffeau &
Slayman, 1981). Therefore, it is tempting to speculate that the
effects of polyenes on the proton gradient may result from a
specific interaction with components that determine the fluidity
of the membrane and thus the activity of the H
+
-ATPase.
At concentrations >0.3 mg L AmB is lethal to, for example,
C. albicans and C. neoformans and is lytic to erythrocytes and
protoplasts (Brajtburg et al., 1990).
The polyenes have been divided into groups: those provoking
potassium leakage and cell death in S. cerevisiae and haemolysis
in erythrocytes at comparable concentrations, and a second
group in which haemolysis occurs at much higher concentra-
tions than yeast potassium leakage. The topically active
natamycin belongs to the first group, whereas AmB and nystatin
belong to the second group (Ryley et al., 1981). This selectivity
may be due to the greater interaction of, for example, AmB with
ergosterol as compared with cholesterol (Readio & Bittman,
1982; Fournier et al., 1998). It should be noted that lethality is
not a simple consequence of changes in permeability of cell
membranes (Brajtburg et al., 1990). Evidence for the role of
reactive forms of oxygen in the lethal or lytic action of AmB was
obtained from studies which showed that AmB injury to cells can
be modulated by extracellular scavengers, such as 1-phenyl-3-
pyrazolidone, and prooxidants (e.g. ascorbic acid). Exposure of
erythrocytes or protoplasts of C. albicans to AmB under hypoxic
conditions reduced the polyene-induced lysis, and cells with
higher catalase activity are less sensitive to lysis by AmB
(reviewed by Brajtburg et al., 1990, 1992).
Thus, the mechanism of action of polyenes is much more
complex than just binding to sterols.
Resistance
Resistance to polyenes is unusual, but known (for reviews see
Vanden Bossche et al., 1994a, 1998; White et al., 1998; Espinel-
Ingroff et al., 2000; Ellis, 2002). Resistance of C. albicans,
C. tropicalis and C. glabrata strains to AmB and to nystatin was
associated with loss or marked depression of ergosterol in the cell
membranes (Dick et al., 1980). Uchida et al. (1994) studied the
membrane composition in polyene antibiotic-sensitive and
resistant strains of Malassezia pachydermatis. The wild-strains
Antifungal agents 7
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
contained more than 0.8 mg of sterols (79% ergosterol) per
gram freeze-dried mycelium, whereas the resistant mutants
contained less than 0.35 mg g (76% ergosterol).
Ziogas et al. (1983) analysed the sterol content of pimaricin-
resistant A. nidulans mutants and found a correlation between
the resistance level and the ergosterol content of 17 mutants,
with low ergosterol contents leading to elevated levels of
resistance. However, in a number of mutants, resistance was
not related to the ergosterol content. Diverse factors influence
the sensitivity of fungi to polyene antibiotics. Therefore, it is not
surprising that a diversity of mechanisms may cause resistance.
Intrinsic resistance to AmB is common for C. lusitaniae (Pfaller
et al., 1994; Peyron et al., 2001), C. guilliermondii (Dick et al.,
1985) and Trichosporum beigelii (Walsh et al., 1990) and some
isolates of Pseudallescheria boydii (Scedosporium apiospermum) are
resistant to AmB in vitro with MICs well in excess of 2 mg L
(Vanden Bossche et al., 1994b; Ellis, 2002). The pseudofungus
Pythium insidiosum differs from the true fungi in many aspects,
for example the plasma membrane is free from sterols. Therefore,
polyenes (and also ergosterol biosynthesis inhibitors) cannot be
used in the treatment of pythiosis (Pier et al., 2000).
Polyenes of interest
Natamycin
Natamycin (pimaricin, Fig. 1) is a polyene antifungal antibiotic
produced by Streptomyces natalensis. A complex multienzyme
system encoded by five polyketide synthase genes is involved in
the biosynthesis of this 26-membered tetraene macrolide (Apar-
icio et al., 1999, 2000).
The natamycin molecule contains a large lactone ring of 25
C-atoms which is linked to the deoxysugar mycosamine (Raab,
1972). Natamycin is insoluble in water and most organic
solvents. It is not absorbed from the gastrointestinal tract.
Spectrum of activity
Natamycin is active against fungal organisms pathogenic for
animals and plants, and is important in the food industry as a
preservative to prevent mould contamination of foods. Natamy-
cin is effective against C. albicans,A. niger,Trichophyton spp.,
Epidermophyton floccosum,Acremonium spp., Cunninghamella spp.,
Fusarium spp. and Pseudallescheria boydii (Raab, 1972, 1978;
Reuben et al., 1989; Rotowa et al., 1990). Although it has been
advocated for use as an aerosol in pulmonary candidosis and
aspergillosis, Aspergillus spp., and also dermatophytes are
frequently resistant to natamycin. Its most important place as
an antifungal agent is in mycotic keratitis. Although natamycin
was found to control experimental A. niger keratitis, some
investigators did not find subconjunctival pimaricin therapy to
be of use in the treatment of experimental keratitis, in rabbits,
caused by A. fumigatus (reviewed by Thomas & Rajasekaran,
1988).
Nystatin
Nystatin (Fig. 1) was discovered in the early 1950s (Hazen &
Brown, 1950). This polyene macrolide antibiotic is produced by
S. noursei (ATCC 11455). The nystatin molecule contains a
polyketide moiety represented by a 38-membered macrolactone
ring to which the deoxysugar mycosamine is attached.
Molecular cloning and characterization of the genes governing
the nystatin biosynthesis resulted in the identification of genes
encoding the polyketide synthase, genes for thioesterase,
deoxysugar biosynthesis, modification, transport and regulatory
proteins (Brautaset et al., 2000). At pH 6–7, nystatin is slightly
soluble in water (4 mg mL). Nystatin is not absorbed appreci-
ably from the gastrointestinal tract.
Spectrum of activity
Nystatin is active against a broad spectrum of fungi in vitro and
in vivo, including Candida spp., A. fumigatus,C. neoformans and
T. beigelii (Raab, 1978; Johnson et al., 1998). It has poor activity
against dermatophytes. For C. albicans the MIC90 and the
minimum lethal concentration (MLC90) ¼2mgL, the MIC90
and MLC90 values for A. fumigatus were both 8 mg L (Johnson
et al., 1998). The use of nystatin is limited to superficial
cutaneous and mucosal mycoses.
Adverse reactions
Parenteral administration of nystatin resulted in dose-limiting
toxicities and infusion-related reactions, this precluded for a
long time its development as a parenteral therapeutic agent
(Hamilton-Miller, 1973). A less toxic parenteral preparation of
nystatin has been formulated. It is a multilamellar liposomal
formulation containing dimyristoylphosphatidylcholine and
dimyristoylphosphatidylglycerol in a 7:3 ratio (Mehta et al.,
1987; Groll et al., 1999). In vitro, liposomal nystatin appeared to
be as active as free nystatin (Johnson et al., 1998).
Amphotericin B
Amphotericin B (Fig. 1) contains eight free hydroxyl groups in
the macrocyclic lactone ring and possess a glycosidically linked
carbohydrate moiety: 3-amino-3,6-dideoxy-
D
-mannopyranose
(mycosamine). This polyene is produced by S. nodosus. The
molecule is rod-shaped with an opposing hydrophobic and
hydrophilic face, and a length of 2.4 nm or about half the
thickness of a gel-phase phospholipid bilayer (Andreoli, 1974).
At pH 6–7 AmB is almost insoluble in water.
Due to its extremely poor intestinal absorption, for systemic
infections AmB is administered intravenously. Normally AmB is
formulated for intravenous administration as a complex with
sodium deoxycholate. Topical preparations may be used for the
treatment of superficial Candida infections; they are of no value in
dermatophytoses (Odds, 1996).
Spectrum of activity
Amphotericin B-deoxycholate has, depending on the concen-
tration used, fungistatic and fungicidal properties and is
active in vitro against Candida species including C. albicans
(MIC90 ¼0.25 mg L; MLC90 ¼0.5 mg L), C. glabrata,C. kefyr,
C. krusei,C. tropicalis and C. parapsilosis (Johnson et al., 1998). It
is active against C. neoformans (MIC90 ¼0.5 mg L; MLC90 ¼
0.5 mg L), H. capsulatum,Sporothrix schenckii,Blastomyces
8H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
dermatitidis,Coccidioides immitis,Phialophora sp. and strains of
A. fumigatus (MIC90 ¼1mgL; MLC90 ¼4mgL) and A. flavus
(MIC90 ¼1mgL; MLC90 ¼1mgL) (see for example, Shad-
omy & Espinel-Ingroff, 1974; Troke, 1993; Johnson et al., 1998;
Goldberg et al., 2000; Moore et al., 2001). An MIC value of
0.09 mg L has been reported for Absidia corymbifera (Troke,
1993). Recent data indicate that the sensitivity of
S. schenckii to AmB is strain dependent; the MIC and minimal
fungicidal concentrations (MFC) were 0.125–2 mg L
(McGinnis et al., 2001). Troke (1993) determined MIC values
(mg L) for Microsporum canis (0.04), M. gypseum (0.3),
T. quinckeanum (0.6), T. mentagrophytes (0.6), T. rubrum (0.3),
T. soudanense (0.09), T. tonsurans (0.05) and T. verrucosum (0.1).
AmB is also active against F. solani and other Fusarium species
(Reuben et al., 1989). Using the microdilution method recom-
mended by the National Committee for Clinical Laboratory
Standards (1998) (NCCLS) the MICs found for Scedosporium
apiospermum,S. prolificans,F. oxysporum and F. solani were:
2–4 mg L, 0.5–16 mg L, 0.25–2 mg L and 0.5–2 mg L,
respectively (Meletiadis et al., 2000).
Adverse reactions
The multiplicity of interactions with both fungal and mamma-
lian cells may be the reason why AmB is such an effective, but
not a well tolerated, antifungal agent (for reviews see Armstrong
& Schmitt, 1990; Sarosi, 1990; Bennett, 1992). Nephrotoxicity
is the most common toxic effect, but according to Bennett
(1992), ‘the litany of toxicity would eventually include anaemia,
phlebitis, renal tubular acidosis, hypokalaemia, hypomagnesae-
mia, nausea, vomiting, weight loss and headache’. To produce
less toxic formulations AmB has been formulated in liposomes
and other lipid complexes. Examples are Ambisome (unilamellar
liposome), Amphocyl (ABCD, lipid disc), Abelcet (ABCL, lipid
ribbon) (for an overview of the lipid formulations see Dupont,
2002).
AZOLE ANTIFUNGAL AGENTS
The class of the antifungal imidazoles and triazoles is generally
the largest and most widely used. Examples of imidazole
derivatives are bifonazole, butoconazole, clotrimazole, econazole,
enilconazole, fenticonazole, isoconazole, miconazole, parcona-
zole and ketoconazole. Triazole derivatives are fluconazole,
itraconazole, terconazole and voriconazole. According to Burt
(2001), there are 24 new triazoles that have been patented in
the past several years.
Imidazole derivatives
Clotrimazole
Clotrimazole (Fig. 2) was one of the first imidazole-based
compounds to be developed as a broad spectrum agent for the
topical treatment of fungal infections (Plempel et al., 1969;
Bu
¨chel et al., 1972). Clotrimazole is a derivative of 1-methylim-
idazole with lipophilic substituents at the methyl group. It is
slightly soluble in water and soluble in DMF. Pharmacokinetic
investigations of clotrimazole in mice, rats and men revealed that
the compound had an inductive effect on microsomal liver
enzymes, which led to rapid metabolism and inactivation of the
active substance. Therefore, this imidazole derivative was
unsuitable as an oral antimycotic (Plempel et al., 1988).
Spectrum of activity
In vitro clotrimazole is active against dermatophytes, Candida
species, Aspergillus spp., C. immitis and C. neoformans.Allescheria
boydii and Phialophora spp. are less sensitive.
Using both the macrodilution and microdilution methods,
Pelletier et al. (2000) analysed 87 isolates of C. albicans for their
susceptibility to clotrimazole. The MIC for 90% of the strains
tested was £0.5 mg L. Among the isolates used in this study, 10
were recovered from patients who fulfilled the definition of
clinical resistance. For six of these 10, the MIC was 0.5 mg L.
Pelletier et al. (2000) suggest that a breakpoint of 0.5 mg L may
be useful in defining clotrimazole resistance in C. albicans.
Using a modified NCCLS method Ferna
´ndez-Torres et al.
(2000) determined the MIC value for 100 isolates of T. rubrum,
the MIC50 and MIC90 were 0.03 and 0.125 mg L, respectively.
When cultured on Sabouraud’s dextrose agar medium an MIC
value of 0.267 mg L was found (Fukushiro et al., 1992). Using
the same medium, the MIC values for T. mentagrophytes and
M. canis were 0.255 and 0.266 mg L, respectively (Fukushiro
et al., 1992). For M. pachydermatis the MICs ranged between
<0.06 and 32 mg L (mean ¼4mgL) (Schmidt, 1997).
Miconazole, econazole and enilconazole
Imidazoles with an ether group form an important class of
antifungal agents. Miconazole (Fig. 2), the first member of the
group, was introduced around the same time as clotrimazole
(Godefroi et al., 1969; Van Cutsem & Thienpont, 1972). Several
other ether compounds have been synthesized; examples are
econazole, isoconazole and orconazole (Godefroi et al., 1969).
These compounds form the basis for further synthesis in many
pharmaceutical companies. Examples are tioconazole, sulcona-
zole, butoconazole and cloconazole. Another ether compound is
imazalil. This imidazole derivative is used in plant protection
(imazalil) and as a veterinary antimycotic (enilconazole).
Enilconazole has also been developed as an antimycotic disin-
fectant (Van Cutsem et al., 1985). In this review we will focus on
miconazole, econazole and enilconazole.
Miconazole nitrate is a white microcrystalline or amorphous
powder, very slightly soluble in water and moderately soluble in
most organic solvents. Its log partition coefficient in a system of
n-octanol and an aqueous buffer is 5.9. The absorption of topical
applied miconazole is extremely low. When given orally to man,
even substantial doses only produce relatively low serum levels
(Boelaert et al., 1976). After intravenous infusion miconazole is
well distributed to the various body tissues, with high drug
concentrations in the lungs, liver, adrenals and kidneys. Poor
levels are observed in the cerebrospinal fluid. Miconazole is
metabolized in the liver, less than 1% of the dose is excreted in
the urine (reviewed in Borgers et al., 1986).
Antifungal agents 9
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
Econazole (Fig. 2) is very slightly soluble in water and most
common organic solvents (for a review of the older literature see
Raab, 1978).
Pharmacologically, enilconazole (Fig. 2) has virtually the
same characteristics as the other imidazole derivatives. Like the
other imidazole antifungal agents, enilconazole is largely broken
down and at least 35 metabolites have been identified. Enilcon-
azole has a low vapour pressure but evaporates easily in closed
environments to an antifungally active gas phase (Van Gestel
et al., 1981; Van Gestel, 1986).
Spectrum of activity
(1). Miconazole: Fungistatic concentrations [geometric mean
(GM) of MICs] of miconazole in casitone medium were tested
against C. albicans (3.9 mg L), C. tropicalis (1 mg L), C. parapsi-
losis (5 mg L), C. glabrata (5 mg L), Trichophyton spp. (1 mg L),
Aspergillus spp. (1 mg L) and H. capsulatum (mycelial phase,
1mgL) (Polak & Dixon, 1987). Using the microdilution method
recommended by NCCLS the MICs found for A. fumigatus,
A. flavus,S. apiospermum,S. prolificans,F. oxysporum and F. solani
were: 4–16, 1–16, 1–8, 16 to >64, 1 to >64 and 4 to
>64 mg L, respectively (Meletiadis et al., 2000). Using an other
NCCLS (1992) method Hennequin et al. (1997) found for 13
isolates of S. apiospermum a mean MIC90 value of 0.12 mg L.
Miconazole was tested for its activity against 23 isolates of
M. furfur by agar dilution. A GM of MIC of 14 mg L was found
(Van Gerven & Odds, 1995). A similar sensitivity of M. furfur was
found by Hammer et al. (2000). For T. beigelii an MIC90 value of
0.78 mg L was measured (Perparim et al., 1996). Microsporum
canis (in Sabouraud broth) is also sensitive to miconazole, the
lowest active antifungal concentration found by Van Cutsem
et al. (1986) was 1 mg L. Using the NCCLS (M38-P) guidelines
for testing filamentous fungi, Ferna
´ndez-Torres et al. (2001)
found an MIC90 of 0.25 mg L.
(2) Econazole: Thienpont et al. (1975) determined the fungi-
static activity against human, animal and plant pathogenic
fungi. Complete growth inhibition of Candida spp. was obtained
at 10–100 mg L (for C. albicans 100 mg L was needed). More
recently, Arias et al. (1996), using the broth dilution micro-
method, found complete inhibition of C. glabrata growth at
12.5 mg L. For M. canis (0.1–1 mg L), M. gypseum (1 mg L),
T. mentagrophytes (0.01 mg L), T. rubrum (0.1–1 mg L) and
N
NCl
Clotrimazole
N
N
Cl
Cl
O
Cl Cl
Miconazole
N
N
Cl
Cl
O
Cl
Econazole
N
N
Cl
Cl
O
Enilconazole
N
N
Cl
Cl
O
O
(±) Ketoconazole
ONN
CH3
O
N
N
Cl
Cl
O
O
O
(±) Parconazole
N
N
N
Cl
Cl
O
O
(±) Itraconazole
ONNN
N
N
ON
N
N
OH
N
N
N
F
F
Fluconazole
Fig. 2. Chemical structures of azole anti-
fungal agents.
10 H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
E. floccosum (0.01 mg L) much lower concentrations were
needed (Thienpont et al., 1975). Econazole was also found to
be a potent inhibitor of the growth of B. dermatitidis (0.1 mg L),
H. capsulatum (0.01 mg L), A. fumigatus and A. flavus (1 mg L)
and S. schenckii (1 mg L). Whereas, for Absidia corymbifera
(100 mg L) and Allescheria boydii (10 mg L) much higher
concentrations were needed (Thienpont et al., 1975). Hammer
et al. (2000) determined the in vitro susceptibilities of M. furfur
and M. sympodialis by a broth dilution assay. MIC90 values of
64 and 1 mg L were found, respectively. However, using the
agar dilution assay MIC90 values of 16 and 0.5 mg L were
reported.
(3) Enilconazole: Enilconazole base and different salts were
tested in vitro on various human, animal and plant pathogenic
fungi; as with econazole, dermatophytes were the most sensitive
(Thienpont et al., 1981). Its antimycotic activity is excellent
against M. canis,M. gypseum,T. mentagrophytes,T. verrucosum,
M. pachydermatis and Aspergillus spp. (Thienpont et al., 1981;
de Jaham et al., 2000). For example, growth of A. fumigatus in
Sabouraud broth was inhibited at 0.1–1 mg L (Van Cutsem &
Janssen, 1988). The in vivo antifungal activity was high for the
dermatophytes and rather low for Candida (Thienpont et al.,
1981).
Ketoconazole and parconazole
Ketoconazole (Fig. 2) belongs to the class of the imidazoles with
ketal groups. Its synthesis has been reported by Heeres et al.
(1979). Ketoconazole is a lipophilic molecule. Its log partition
coefficient (Log P) is 3.84. The pKa of the imidazole ring of
ketoconazole is 6.51. That means that at, for example, pH 5.6
the imidazole ring is more than 90% protonated (Vanden
Bossche et al., 1988). As the antifungal activity of the imidazole
derivatives depends, at least partly, on the binding of the
unprotonated imidazole nitrogen to cytochrome P450 (see
mechanisms of action of azole antifungals), protonation will
affect the in vitro antifungal activity of ketoconazole.
After oral administration, ketoconazole is metabolized in the
liver and is converted into many inactive metabolites. Absorp-
tion of ketoconazole from the gastrointestinal tract is more rapid
in rats and guinea-pigs than in rabbits and dogs. Following a
dose of 10 mg kg peak plasma levels were 1.0, 3.7, 8.9 and
16.5 mg L in rabbits, guinea-pigs, dogs and rats, respectively
(Heel, 1982). Limited pharmacokinetic studies in pigeons and
cockatoos are performed. The peak concentration varied
between 0.5 and 5 h and the elimination half-life time was
2–2.8 h in pigeons and 3.8 h in cockatoos (Kollias et al., 1986).
For dissolution and subsequent absorption of ketoconazole the
acidity of the stomach plays an essential role. Therefore,
coadministration of compounds such as cimetidine should be
avoided. After topical application ketoconazole is found in the
stratum corneum (Pershing et al., 1994). Both ketoconazole and
parconazole (Fig. 2) have a dioxalan group.
After a single oral dose of 13 mg parconazole per kg body
weight a blood peak level of 2.28 mL L was reached after 2 h in
guinea-fowls; the unchanged drug is specially localized in crop
and skin (Levron et al., 1985).
Spectrum of activity
(1) Ketoconazole: The spectrum of antifungal activity of ketocon-
azole resembles qualitatively that of miconazole. Quantitative
results for MICs vary among laboratories, depending on
inoculum size, culture medium, pH, duration and time of
incubation and growth phase of the fungus (see for example,
Galgiani & Stevens, 1976; Thienpont et al., 1979; Granade &
Artis, 1980). Using the NCCLS method (M38-P) Ferna
´ndez-
Torres et al. (2001) determined the in vitro activities of 10
antifungal drugs against 508 strains of dermatophytes. In this
study ketoconazole displayed excellent activity against, for
example, E. floccosum,M. canis,M. gypseum,T. mentagrophytes
and T. verrucosum. The GM of MICs were 0.08, 0.12, 0.23, 0.46
and 0.5 mg L, respectively. Trichophyton simii was less sensitive
with a GM of 4 mg L. When all 508 strains were considered
together the GM was 0.5 mg L. The tests were performed in
round-bottomed 96 microplates. Using tubes (final volume
10 mL) instead of microplates, Perea et al. (2001a) found
somewhat higher GMs. For example, the GMs for E. floccosum,
M. canis,M. gypseum,T. mentagrophytes and T. verrucosum were
0.61, 1.12, 1.78, 1.55 and 6.72 mg L, respectively. For
T. equinum these investigators found a GM of 6.35 mg L.
The MICs of ketoconazole were determined by the macrodi-
lution test for isolates of A. fumigatus,A. flavus,P. boydii,
R. arrhizus and S. schenckii (Espinel-Ingroff et al., 1995). The
MIC ranges were: 2–4 mg L, 0.5–16, 0.03–2, 1–16 and
0.25–4 mg L, respectively.
A broth microdilution method was used to determine the
in vitro activities of ketoconazole against Cladosporium spp.,
Exophiala dermatitidis,Cladophialophora spp., Scopulariopsis chart-
arum,Paecilomyces spp., and F. solani (Llop et al., 2000). The
mean MICs were: 1.32, 0.25, 0.09, 2, 0.17 and 32 mg L,
respectively.
Ketoconazole was tested for its activity against 23 isolates of
M. furfur by agar dilution. A GM of 0.51 mg L was found (Van
Gerven & Odds, 1995). It was much more potent than
bifonazole (GM ¼8.1 mg L), miconazole (GM ¼14 mg L) and
clotrimazole (GM ¼15 mg L). That ketoconazole is much more
active than miconazole against M. furfur was also found by
Hammer et al. (2000). Of interest is that these investigators
found that the MICs and MFCs of ketoconazole were equival-
ent. Microsporum sympodialis,M. slooffiae,M. pachydermatis,
M. globosa,M. obtusa and M. restricta were also susceptible to
ketoconazole with MIC ranging from £0.03 to 0.125 mg L
(Gupta et al., 2000).
Morace et al. (1995) used a microdilution broth method to
determine the susceptibility of Candida spp. isolates to ketocon-
azole. The MIC ranges for C. albicans,C. krusei,C. glabrata,
C. kefyr,C. tropicalis,C. rugosa and C. norvegensis were 0.04–50,
0.09–12.5, 0.09–12.5, 0.04, 12.5, 0.39 and 0.19 mg L,
respectively. With the NCCLS M27-A broth microdilution
method lower MIC values were found for C. albicans (0.03–
16 mg L), C. krusei (0.25–1 mg L), C. glabrata (£0.03–4 mg L)
and C. tropicalis (0.03–1 mg L) (Arthington-Skaggs et al.,
2000). For C. parapsilosis,C. dubliniensis and C. neoformans the
MIC ranges were: 0.03–2, 0.03–8 and 0.06–1, respectively.
Antifungal agents 11
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
Ketoconazole is also active against B. dermatitidis (MIC
range: 0.1–04 mg L) (Chapman et al., 1998), H. capsulatum
(0.39 mg L) and some isolates of C. immitis (0.39 mg L)
(Shadomy et al., 1985). However, for other C. immitis isolates
more than 100 mg L was needed (Shadomy et al., 1985).
(2) Parconazole: This orally active imidazole derivative has
been developed for veterinary applications. It has a spectrum of
antifungal activity in vitro which quantitatively resembles that of
the other azole antifungal agents. As with the other imidazole
derivatives, the inhibitory concentration depends on the culture
medium used. For example, in casein hydrolysate-yeast extract-
glucose (CYG) medium 50% inhibition of C. albicans growth was
achieved at 30 n
M
, in yeast-nitrogen base-glucose medium
2000 n
M
was needed (F. Cornelissen & H. Vanden Bossche,
unpublished result).
The in vitro antifungal activity of parconazole was tested in
Sabouraud liquid medium against dermatophytes, yeasts,
dimorphic fungi, mycetoma agents and phycomycetes (Desplen-
ter & Van Cutsem, 1980). Growth of dermatophytes is
completely inhibited at 1–10 mg L, Candida spp. at 10–
100 mg L and the yeast form of B. dermatitidis at 0.01 mg L.
Growth of mycetoma agents and actinomycetales is inhibited at
0.1–100 mg L. Absidia ramosa,Mucor spp., Rhizopus spp.,
Aspergillus spp. (except A. nidulans), Fusarium spp., S. schenckii,
Alternaria spp. and S. brevicaulis are less sensitive.
Triazole derivatives
Itraconazole and fluconazole
The orally active 1,2,4-triazole derivative, itraconazole (Fig. 2),
is a highly lipophilic molecule. Its log P (n-octanol aqueous
buffer of pH 8.1) is 5.66. Itraconazole is a weak base (pKa of the
triazole is <2 and of the piperazine is 3.7) and therefore is only
protonated at low pH, such as in gastric juice. Itraconazole is
soluble in N,N-dimethylformamide (3.5 g 100 mL) and dimethyl
sulphoxide (1.1 g 100 mL), slightly soluble in polyethylene glycol
400 (0.19 g 100 mL), but almost insoluble in water (at pH 7.0
<0.0001 g 100 mL) (for more information see Vanden Bossche
et al., 1993a).
Pharmacokinetics (Heykants et al., 1987, 1989): The distri-
bution of itraconazole was studied in male and female Wistar rats
after an oral 10 mg kg dose of
3
H-itraconazole. In most tissues,
peak radioactivity concentrations were achieved within 2–4 h
after dosing. The levels obtained in the adrenals, liver, kidney,
lung, skin and perirenal fat were consistently higher than in
plasma. Only in the brain were levels lower than in plasma.
Tissue levels in dogs were measured 24 h after the last dose of
a 12-month chronic toxicity experiment at daily doses of 5, 20 or
80 mg kg. Except for the brain, itraconazole tissue levels were
higher than the corresponding plasma levels. Brain tissue levels
were similar to those measured in plasma. Distribution studies
have shown that therapeutic active levels of itraconazole in
humans are maintained much longer in some infected tissues
than in plasma. For instance, active levels persist for 4 days in
the vaginal epithelium after a 1-day treatment and for 3 weeks
in the stratum corneum of the skin after treatment has been
stopped. After oral administration, itraconazole can be detected
in nails 1–2 weeks after the start of therapy (Debruyne &
Coquerel, 2001). Itraconazole is still detectable in nails 27 weeks
after stopping administration.
The recently developed oral hydroxypropyl-b-cyclodextrin
solution of itraconazole showed improved gastric absorption
and high topical drug concentrations in the oral cavity and the
oesophagus (Stevens, 1999). More information on the pharma-
cokinetics and pharmacodynamics of cyclodextrin itraconazole
in patients can be found in Groll et al. (2002).
Itraconazole is extensively metabolized. Main metabolic path-
ways are oxidative scission of the dioxolane ring, oxidative
degradation of the piperazine ring and N-dealkylation of the
1-methylpropyl substituent. One metabolite deserves special
attention. Hydroxy-itraconazole is the antifungally active meta-
bolite of itraconazole formed in the liver by oxidation of
1-methylpropyl substituent of itraconazole (Heykants et al.,
1989). In vitro tests showed the same IC
50
values for itraconazole
and hydroxy-itraconazole, within a mode ± 1 dilution range of
experimental error, for 90% of 1481 isolates of pathogenic fungi
(Odds & Vanden Bossche, 2000). Some 10–15% of C. glabrata
and T. mentagrophytes isolates were more susceptible to itracon-
azole than hydroxy-itraconazole.
In common with other 1-substituted 1,2,4-triazole derivatives,
the bis-triazole propanol derivative, fluconazole (Fig. 2) is a weak
base. In contrast with most of the available azole antifungal
agents, fluconazole is soluble in water (8.71 mg at 37 C;
2.20 mg Lat4C) (Hitchcock & Whittle, 1993). Its relatively
high polarity contributes to fluconazole’s low protein binding
and even distribution throughout the body (Troke, 1993).
Fluconazole is rapidly and virtually completely absorbed by the
oral route in mouse, rats, dogs and man (pharmacokinetics are
reviewed by Brammer & Tarbit, 1987). It is extensively
distributed throughout the body and penetrates readily into
cerebrospinal fluid and saliva. Following administration of
50 mg fluconazole per day for 8 days to male cats, fluconazole
concentrations in cerebrospinal fluid exceeded reported MICsof
this triazole for susceptible pathogenic fungi (Vaden et al., 1997).
Marked species differences exist in the elimination half-life of
fluconazole, values ranging from 4 h in rodents, almost 15 h in
dogs to about 25 h in human volunteers. The major route of
excretion of the drug is as unchanged fluconazole in the urine in
all species studied. Excretion ranges between 70% of the dose in
dogs and 80% in rodents and humans. As shown further,
fluconazole has only modest in vitro activity against several
pathogenic fungi. Nevertheless its clinical efficacy in many
systemic and superficially fungal infections is remarkable (Troke,
1993). Fluconazole pharmacokinetics certainly contribute to this
in vivo antifungal activity.
Spectrum of activity
(1) Itraconazole: With the NCCLS M27-A broth microdilution
method (Arthington-Skaggs et al., 2000) the following MIC
ranges were found: C. albicans (0.03 to >16 mg L), C. krusei
(0.06–0.25 mg L), C. glabrata (0.25 to 16 mg L), C. tropicalis
(0.125–0.5 mg L). For C. parapsilosis,C. dubliniensis and
12 H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
C. neoformans the MIC ranges were 0.03–0.125, 0.03–4 and
0.03–0.125 mg L, respectively. Itraconazole had MICsof
0.06–2, 0.03–0.5 and 0.03–1 mg L for C. famata,C. lusitaniae
and C. guilliermondii, respectively (Pfaller et al., 2001). Using the
same method, Li et al. (2000) tested the activity of itraconazole
against the mould forms of three dimorphic fungi. Itraconazole
had MIC90s (MIC at which 90% of the isolates are inhibited) of
0.06 mg L for H. capsulatum, 0.125 mg L for B. dermatitidis and
1mgmL for C. immitis. Using the macrobroth dilution suscep-
tibility test to evaluate the activity of itraconazole against
B. dermatitidis isolates, Chapman et al. (1998) found an MIC
range of £0.018–0.07 mg L. The range for the minimal lethal
concentration was the same. For T. beigelii an MIC90 value of
0.19 mg L was measured (Perparim et al., 1996). Using the
agar dilution method M. furfur,M. sympodialis,M. slooffiae,
M. pachydermatis,M. globosa,M. obtusa and M. restricta were
found susceptible to itraconazole at low concentrations (MICs
ranging from £0.03 to 0.125 mg L; Gupta et al., 2000).
Itraconazole displayed excellent activity against, for example,
E. floccosum,M. canis,M. gypseum,T. mentagrophytes,T. rubrum,
T. simii and T. verrucosum. The GM of MICs were 0.03, 0.08,
0.04, 0.17, 0.09, 0.12 and 0.01 mg L, respectively (Ferna
´ndez-
Torres et al., 2001). The tests were performed in round-bottomed
96 microplates. Using tubes (final volume 10 mL) instead of
microplates, Perea et al. (2001a) found for E. floccosum,M. canis,
M. gypseum,T. equinum,T. mentagrophytes and T. verrucosum the
following GMs: 0.11, 0.10, 0.04, 0.04, 0.04 and 0.35 mg L,
respectively. Thus with both methods itraconazole was found to
be more active than ketoconazole (vide supra) and, as shown
below, much more active than fluconazole.
Moore et al. (2000) determined the in vitro activity of
itraconazole against clinical isolates of Aspergillus. The GM of
MICs (mg L) for A. fumigatus,A. terreus,A. flavus and A. niger
were 0.77, 0.33, 0.5 and 0.93 mg L, respectively.
For isolates of P. boydii,R. arrhizus and S. schenckii the MIC
ranges were 0.03–1, 0.12–16 and 0.03–16 mg L, respectively
(Espinel-Ingroff et al., 1995).
A broth microdilution method was used to determine the
in vitro activity of itraconazole against Cladosporium spp.,
E. dermatitidis,Cladophialophora spp., S. chartarum,Paecilomyces
spp. and F. solani (Llop et al., 2000). The mean MICs were 0.76,
0.25, 0.04, 32, 0.19, 0.17 and 32 mg L, respectively.
(2) Fluconazole: As already mentioned, the clinical efficacy of
fluconazole in many systemic and superficial fungal infections is
remarkable, although it has only modest in vitro activity against
several pathogenic fungi (Troke, 1993).
With the NCCLS M27-A broth microdilution method the fol-
lowing MIC ranges were found: C. albicans (0.25 to 64 mg L),
C. krusei (16–64 mg L), C. glabrata (1 to 64 mg L), C. tropicalis
(0.25–64 mg L), C. parapsilosis (0.25–1 mg L), C. dubliniensis
(0.125 to 64 mg L) and C. neoformans (0.25–64 mg L) (Arth-
ington-Skaggs et al., 2000). Pfaller et al. (2001) found MICsof
£8mgL for 78% of the C. neoformans isolates tested; MICsof
16–32 mg L were found for 21% of the isolates. Using the same
method, Perparim et al. (1996) found for T. beigelii an MIC90 of
6.25 mg L. High MIC90 values were also found for E. floccosum
(4.39 mg L), M. canis (16 mg L), M. gypseum (>64 mg L),
T. mentagrophytes (>64 mg L) and T. rubrum (16 mg L) (Fern-
a
´ndez-Torres et al., 2001). For T. equinum the MIC was
64 mg L (Perea et al., 2001a). An MIC of 64 mg L was found
for A. fumigatus,A. flavus,R. arrhizus and S. schenckii. The
P. boydii isolates tested were more susceptible; MIC range was
2–32 mg L (Espinel-Ingroff et al., 1995).
Chapman et al. (1998) found an MIC range of 2.5–4 mg L for
B. dermatitidis, whereas the MIC range was £0.018–0.07 mg L
for itraconazole.
Mechanisms of action of azole antifungal agents
The antifungal activity of azole derivatives such as miconazole
(Vanden Bossche et al., 1978), econazole, bifonazole, clotrima-
zole (Vanden Bossche, 1985, 1988), ketoconazole (Vanden
Bossche et al., 1980, 1988), enilconazole (Siegel & Ragsdale,
1978; Vanden Bossche et al., 1987b), parconazole (Pye &
Marriott, 1982), terconazole (Vanden Bossche & Marichal,
1991), itraconazole (Vanden Bossche et al., 1986, 1987a,b,
1988) and fluconazole (Hitchcock, 1991; Hitchcock & Whittle,
1993) arises from a complex multimechanistic process initiated
by the inhibition of two cytochromes P450 involved in the
biosynthesis of ergosterol, namely the P450 that catalyses
the 14a-demethylation step encoded by ERG11 (CYP51) and the
D22-desaturase, encoded by ERG5 (CYP61) (Kelly et al., 1997a).
The sterol-14a-demethylase is the major fungal target for all
azole derivatives studied so far. Interaction with CYP51 results in
a decreased availability of ergosterol and an accumulation of
14-methylsterols, such as 14a-methyl-ergosta-8,24(28)-dien-
3b,6a-diol (3,6-diol) (Vanden Bossche et al., 1993b; Vanden
Bossche, 1995; Marichal et al., 1999a). In C. neoformans
(Vanden Bossche et al., 1993b) and H. capsulatum (Vanden
Bossche et al., 1990) azoles induce the accumulation of
14-methylated 3-ketosteroids such as obtusifolione and or
14-methylfecosterone, suggesting that azoles affect directly or
indirectly the 3-ketostereductase, a product of the ERG27 gene.
Ergosterol is an essential component of fungal plasma mem-
branes. It regulates membrane permeability and the activity of
membrane-bound enzymes; this sterol is a major component of
secretory vescicles and has an important role in mitochondrial
respiration and oxidative phosphorylation (for reviews see Vanden
Bossche, 1985, 1988, 1990; Daum et al., 1998). Thus, it can be
expected that changes in ergosterol levels and sterol structure
influence the activity of several metabolic pathways. For example,
in C. albicans grown for 16 h in the presence of 10 n
M
miconazole,
a shift from oleic acid to palmitic acid was observed suggesting an
inhibition of the D-9 desaturase (Vanden Bossche, 1985). An
increase in palmitate was also found in C. albicans grown in the
presence of 1 l
M
clotrimazole, econazole, miconazole or ketocon-
azole (Georgopapadakou et al., 1987). As expected from the
important role of ergosterol in mitochondrial activities, inhibition
of ergosterol synthesis by, for example, miconazole (De Nollin
et al., 1977) and econazole (Wilm & Stahl, 1983) affects
cytochrome c oxidase in the mitochondrial membranes of
C. albicans and S. cerevisiae. Chitin is a structural component of
Antifungal agents 13
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
fungal cell walls (see Kollar et al., 1995). Chitin synthesis is inhibited
at a high and stimulated at a low ergosterol content (Pesti et al.,
1981; Chiew et al., 1982). Thus, ergosterol biosynthesis inhibitors
increase chitin synthesis resulting in an irregular distribution of
patches of chitin in, for example, C. albicans (Barug et al., 1983;
Vanden Bossche, 1985) and A. fumigatus (Marichal et al., 1985).
Recently DNA chip technology has been used to study cellular
responses to perturbations of ergosterol biosynthesis caused by
itraconazole. Simultaneous examination of over 6600 C. albicans
gene transcript levels, representing the entire genome, upon
treatment of cells with itraconazole, revealed 296 genes to be
responsive: 116 gene transcript levels were at least 2.5-fold
decreased while 180 were increased (De Backer et al., 2001).
These results further prove that the antifungal activity of azole
derivatives arises from a complex multimechanistic process
initiated by the inhibition of ergosterol biosynthesis.
In addition to their effects on ergosterol biosynthesis high
concentrations of miconazole, econazole and clotrimazole disturb
the lipid organization in membranes (Brasseur et al., 1983;
Vanden Bossche, 1985, 1988). For example, using differential
scanning calorimetry, it has been shown that miconazole shifts the
transition temperature of dipalmitoylphosphatidylcholine (DPPC)
multilamellar vesicles to low values without affecting the enthalpy
of melting (Vanden Bossche et al., 1982, 1984). The conformation
of miconazole inserted in a DPPC monolayer was calculated by
using a procedure of conformational analysis (Brasseur et al.,
1983). The area occupied per miconazole molecule was 9 nm
2
.
This is higher than the area occupied per DPCC molecule: 6 nm
2
.
Such a conformation should result in a destabilizing effect of high
concentrations of miconazole. This effect may explain, at least
partly, its activity against Gram-positive bacteria. In Gram-
positive bacteria miconazole may also cause an impairment of
cell wall function by inhibiting carrier lipid function (phosphor-
ylated derivative of C-55 isoprenoid alcohol) and peptidoglucan
synthesis (Vanden Bossche et al., 1984). The miconazole induced
impairment of cell wall integrity in Gram-positive bacteria may be
at the origin of the synergism between miconazole and polymyxin
B; miconazole may increase the accessibility of the cell membrane
to polymyxin B (Cornelissen & Vanden Bossche, 1983).
Miconazole alone is inactive against Escherichia coli; but it
enhances the sensitivity of this Gram-negative bacterium to
polymyxin B. This synergism might be due to a direct interaction of
miconazole with membrane components. Miconazole, polymyxin
B together with prednisolone are the components of Surolan
(Janssen Animal Health, Beerse, Belgium), an otic preparation for
treatment of otitis externa in dogs.
Resistance
A comprehensive analysis of the clinical outcome of Candida
infections and of the MIC results obtained with the NCCLS M27-T
broth macrodilution method on Candida isolates suggested that
isolates with MIC >32mgL are regarded as fluconazole
resistant (Rex et al., 1997). Isolates for which the MICs are
£8mgL are susceptible to fluconazole, isolates with fluconazole
MICs of 16–32 mg L are considered susceptible on the basis of
data indicating a clinical response when >100 mg fluconazole
per day is given. In the same study, isolates with an MIC 1
mg L were considered to be resistant to itraconazole. Examples of
fluconazole and or itraconazole resistant C. albicans isolates are
shown in Table 2. It should be noted that large-scale surveys of
yeasts isolated from blood cultures, based on standardized
methodology and resistance definitions, do not support the view
that antifungal resistance in pathogenic yeasts constitutes a
significant or growing problem (Sanglard & Odds, 2002).
A significant reason for clinical treatment failure is severe
immunosuppression.
Until 1986, resistance of pathogens to azole antifungals was
only sporadically reported and invariably in immunocompro-
mized patients who had received long-term oral treatment with
ketoconazole for chronic mucocutaneous candidosis (Horsburgh
& Kirkpatrick, 1983). An increase in the number of immuno-
compromised patients, partially as a result of the AIDS
pandemic, preceded increasing reports of clinical failure to
fluconazole therapy (see for example Vanden Bossche et al.,
1994a, 1998; Johnson et al., 1995; Klepser et al., 1997; Odds,
1998; White et al., 1998; Stevens & Holmberg, 1999; Espinel-
Ingroff et al., 2000; Sanglard & Odds, 2002). The importance of
a normal immune system can be related to the fungistatic mode
of action of the azoles, which implies that part of the clearance of
the fungal infection must be accomplished by host-related factors
(Marichal, 1999).
Some C. albicans isolates from HIV-positive patients with
oropharyngeal candidosis and from some HIV-negative patients
with recurrent vaginal candidosis, have been reported to be
susceptible to other azole derivatives, such as ketoconazole or
itraconazole (Ruhnke et al., 1994; Johnson et al., 1995; Stevens
& Stevens, 1996). In a recent study on safety, pharmacokinetics
and pharmacodynamics of cyclodextrin itraconazole in HIV-
infected paediatric patients with oropharyngeal candidosis Groll
et al. (2002) showed that all patients with fluconazole-resistant
isolates responded to treatment with cyclodextrin itraconazole;
however, they also found that there was no clear correlation
between the MIC of itraconazole and response to therapy. In
many studies on C. albicans a progressive increase in the MICsof
itraconazole (and also of terbinafine) was found to parallel the
increase in fluconazole MICs (Barchiesi et al., 1994; Sanglard
et al., 1996; Lo
´pez-Ribot et al., 1999).
Like many other moulds, A. fumigatus is intrinsically resistant
to fluconazole (Edlind et al., 2001).
The incidence of resistance to itraconazole in Aspergillus
isolates is still rare (see Denning et al., 1997; Dupont et al.,
2000; Mosquera & Denning, 2002).
As a result of the widespread use of fluconazole Candida species
have been selected that are intrinsically resistant to fluconazole,
such as C. krusei (Marichal et al., 1995), or whose resistance is
easily inducible, such as C. glabrata (Rex et al., 1995; Sanglard
et al., 1999, 2001) and C. tropicalis (Barchiesi et al., 2000).
Fungal pathogens could use several general mechanisms to
become less susceptible to azole antifungals (see for example
Vanden Bossche et al., 1994a,b, 1998; Vanden Bossche, 1997;
White et al., 1998; Marichal, 1999; Sanglard et al., 1999, 2001).
14 H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
These mechanisms include: failure to accumulate azole
antifungals, changes in the interaction of the drug with the
target enzyme, the 14a-demethylase and changes in other
enzymes of the ergosterol pathway.
Failure to accumulate azole antifungals
Failure to accumulate azole antifungals has been identified as a
major cause of resistance in several post-treatment fungal isolates
and species that are less sensitive to azole antifungal agents and
other ergosterol biosynthesis inhibitors. These isolates include
C. albicans (Ryley et al., 1984; Venkateswarlu et al., 1995;
Sanglard et al., 1995), C. krusei (Marichal et al., 1995), C. glabrata
(Vanden Bossche et al., 1992; Sanglard et al., 1999, 2001),
C. dubliniensis (Moran et al., 1998), C. neoformans (Venkateswarlu
et al., 1997), A. flavus and A. fumigatus (Tobin et al., 1997).
Failure to accumulate antifungal agents can be the conse-
quence of impaired drug influx or enhanced drug efflux. Efflux
pumps are now recognized as a common cause of decreased
intracellular content of ergosterol biosynthesis inhibitors.
Two types of efflux pumps have been described: membrane
transport proteins belonging either to the major facilitator
superfamily (MFS) or to the ATP-binding cassette (ABC)
superfamily. The MFS are energized by the proton-motive force
(Goffeau et al., 1997). The ABC-type transporters utilize ATP as
a source of energy (Decottignies & Goffeau, 1997).
In C. albicans more than 10 different ABC-type transporters
have been reported (Marr et al., 1998), but only the pumps
encoded by CDR1 and CDR2 are known to be involved in the
mechanisms of azole resistance (Prasad et al., 1995; Sanglard
et al., 1997; Vanden Bossche et al., 1998). CDR1 or CDR2 over-
expression in S. cerevisiae was found to confer cross-resistance to
several antifungal agents, such as fluconazole, ketoconazole,
itraconazole, terbinafine and amorolfine (Sanglard et al., 1995,
1996, 1997, 1998a,b; Vanden Bossche et al., 1998).
In fluconazole-resistant C. albicans isolates a multidrug efflux
transporter of the MFS has been shown to be responsible for the
low level of accumulation of fluconazole (Sanglard et al., 1995).
The MFS gene, CaMDR1, was found to be overexpressed in
resistant isolates (Sanglard et al., 1995). The product of the
CaMDR1 gene, CaMdr1p, expels terbinafine and fluconazole, but
not itraconazole. This finding correlates with the observation by
Albertson et al. (1996) that a C. albicans mutant in which the
Table 2. Examples of reported amino acid
polymorphisms in sensitive and resistant
Candida albicans isolates Investigator Strain
Amino acid point
mutations*
GrowthMIC (lgmL)
Fluconazole Itraconazole
White (1997) 1 0.25 0.03
17 R467K >64.00 >2.00
Lo¨ffler et al. (1997) l4 F105L 128.00 0.064
17 V488I >256.00 >32.00
18 E266D, G464S >256.00 >32.00
112 F105L, E266D 8.00 0.25
TU3 E266D, G448E >256.00 4.00
TU6 F105L, E266D >256.00 6.00
TU9 K287R, G464S >256.00 >32.00
P2 F105L, G464S >256.00 >32.00
P3 F105L, G450E >256.00 0.75
P4 G450E, V488I >256.00 >32.00
Sanglard et al. (1998a,b) C26 D116E, Y132H, S405F >128 4.00
C34 D116E, E405F 2.00 0.062
C37 G464S, R467K 8.00 0.062
C39 S405F 32.00 0.125
C56 D116E, G129A, G464S 128.00 >2.00
Marichal et al. (1999b) NCPF3363 Y132H 32.00 2.00
B59626àD116E, K128T, A149V >64.00 1.00
B59630àD116E, K128T, A149V 64.00 0.5
J913004 D116E, K128T, V452A, 64.00 1.00
G464S
ATCC44858 D153E, E266D 0.13 0.016
Kakeya et al. (2000) Darlington Y132H, I471T >64.00nd
NCPF3310
*A, alanine; D, aspartic acid; E, glutamic acid; F, phenylalanine; G, glycine; H, histidine; I, isoleu-
cine; K, lysine; L, leucine; R, arginine; S, serine; T, threonine; V, valine; Y, tyrosine.
MICs determined by a broth microdilution method (Odds et al., 1995) based on the NCCLS M27A
protocol. Results taken from Marichal et al. (1999b).
àB59626 and B59630 were sequentially isolated from an AIDS patient in Germany (provided by
Prof. G. Just).
§J913004 1 was isolated from an AIDS patient in France (provided by Prof. B. Dupont).
Result taken from Kakeya et al. (2000).
Antifungal agents 15
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
CaMDR1 gene is overexpressed is specifically resistant to
fluconazole, but not to the other azole antifungal agents tested.
Lyons and White (2000) showed that CaMDR1 and CDR
mRNAs are transcriptionally overexpressed in a resistant
C. albicans isolate, suggesting that the antifungal drug resistance
in this isolate is associated with the promoter and trans-acting
factors of the CDR1,CDR2 and CaMDR1 genes.
The ATP-binding cassette-type transporters have also been
found in C. krusei (Katiyar & Edlind, 2001), C. dubliniensis
(Moran et al., 1998), C. glabrata (Miyazaki et al., 1998; Sanglard
et al., 1999, 2001), C. tropicalis (Barchiesi et al., 2000),
C. neoformans (Thornewell et al., 1997), A. nidulans (Del Sorbo
et al., 1997), A. fumigatus and A. flavus (Tobin et al., 1997). In
A. nidulans the ABC transporters AtrA and AtrB are involved in
resistance to enilconazole (Del Sorbo et al., 1997).
Bouchara et al. (2000) studied a C. glabrata isolate that
exhibited cross-resistance to ketoconazole and fluconazole. The
results suggested that the resistant isolate had a respiratory
deficiency. Further studies showed a partial deletion of mito-
chondrial DNA analogous to that described for the rho
0
petite
mutants of S. cerevisiae. Sanglard et al. (2000, 2001) observed
up-regulation of the ABC-type transporters CgCDR1 and CgCDR2
in such rho
0
mutants of C. glabrata. Thus mitochondrial loss may
be one of the pathways leading to the up-regulation of ABC-
transporters.
Members of the MFS have been found in C. dubliniensis (Moran
et al., 1998), C. glabrata (Sanglard et al., 1999), C. tropicalis
(Barchiesi et al., 2000) and C. neoformans (Spitzer & Spitzer, 1997).
Changes in the interaction of the drug with the target enzyme
In an azole-resistant C. albicans strain (NCPF 3363) isolated from
a patient with chronic mucocutaneous candidosis (Smith et al.,
1986) we observed a diminished affinity of the cytochromes
P450 for azole antifungals and a red shift in the maximum of the
CO absorption spectrum from 448 to 450 nm (Vanden Bossche
et al., 1990). We suggest that these changes resulted from a
mutation in the gene coding for the 14a-demethylase. More
recently, it was found that strain NCPF 3363 contained a
Y132H mutation (replacement of tyrosine with histidine at
amino acid 132) on both of its CYP51 alleles (Marichal et al.,
1999b) (Table 2).
An alteration of tyrosine 132 was found commonly among
fluconazole-resistant C. albicans isolates (in Table 2 examples of
reported amino acid polymorphisms are listed). The importance
of this Y132H mutation for azole susceptibility has been
demonstrated by Sanglard et al. (1998b), Asai et al. (1999)
and Kelly et al. (1999a). The results obtained by Kelly et al.
(1999a) and Asai et al. (1999) show that the Y132H substitu-
tion occurred without drastic perturbation of the 14a-demethy-
lase activity allowing mutants to produce ergosterol and retain
viability, an efficient strategy for resistance. Other mutations not
only reduce substantially the inhibitory effect of fluconazole but
also result in a reduced catalytic activity of the sterol 14a-deme-
thylase. Examples are the T315A (Lamb et al., 1997), G464S
(Kelly et al., 1999b) and the R467K (White, 1997; Lamb et al.,
2000b) amino acid substitutions. In the A. fumigatus CYP51
sequence isoleucine 301(I301) corresponds to C. albicans
T315 (Edlind et al., 2001). Edlind et al. (2001) propose that
the presence of I301 contributes to A. fumigatus fluconazole
resistance.
As shown in Table 2, a number of mutations affect only the
inhibitory effect of fluconazole and not the MICs of itraconazole.
These results suggest that the affinity of itraconazole is much less
affected by these mutations. The latter was further proved by
measuring the effects of these mutations on subcellular ergos-
terol biosynthesis. For example, C. albicans isolate C26 (Table 2)
contained two mutations S405F and Y132H, next to a mutation
also found in some azole-sensitive isolates (D116E). For subcel-
lular fractions of sensitive C. albicans isolates 50% inhibition of
ergosterol synthesis was found at about 40 n
M
fluconazole,
whereas for the C26 isolate 4880 n
M
fluconazole was needed
(Marichal et al., 1999b). With itraconazole, 50% inhibition of
ergosterol synthesis by subcellular fractions of azole-sensitive
C. albicans isolates was 33–44 n
M
, whereas only 20 n
M
was
required to reach 50% inhibition of ergosterol synthesis by the
subcellular fraction of the C26 isolate (Marichal et al., 1999b).
The smaller, hydrophilic fluconazole molecule has fewer stabil-
ization sites in the active pocket compared with the lipophilic
itraconazole molecule which may explain this difference in
activity (Vanden Bossche & Koymans, 1998).
A common mechanism of resistance on eukaryotic cells is
gene amplification (White et al., 1998). In a subcellular fraction
of a fluconazole-resistant clinical isolate of C. glabrata, ergosterol
synthesis was 8.2-fold higher than that from the pretreatment
isolate; the increased synthesis coincided with an increase in
microsomal P450 content (Vanden Bossche et al., 1992). This
increased level of the protein was associated with CYP51
(ERG11) gene amplification (Vanden Bossche et al., 1994b;
Marichal et al., 1997). The CYP51 mRNA transcript in the
resistant isolate was eight times greater than it was in the
susceptible isolate (Marichal et al., 1997). This increase in copy
number is due to duplication of the chromosome containing the
CYP51 gene. As expected, the amplification of an entire
chromosome has tremendous effects on the protein expression
in the cells: from 1377 proteins which were identified, 25 were
up-regulated by more than a factor of three and 76 were down-
regulated by the same factor (Marichal et al., 1997). Further
studies are needed to identify the up- and down-regulated
products. For example, it would be of interest to know whether
some of the products are involved in drug efflux. In fact, changes
in drug accumulation were observed in this fluconazole-resistant
C. glabrata isolate (Vanden Bossche et al., 1992).
Changes in other enzymes of the ergosterol pathway
Fungi achieve effective resistance to azoles by circumventing or
compensating the toxic consequences of the changes in sterol
composition. As already mentioned azole-induced growth inhi-
bition results from the depletion of ergosterol and the coinci-
dental accumulation of membrane-disturbing 14-methylated
sterols and 3-ketosteroids. However, not all 14-methylated
sterols disturb membranes; for example, 14-methylfecosterol
can fulfil some of the functions of ergosterol (Watson et al.,
16 H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
1989). Accumulation of 14-methylfecosterol is achieved when
C. albicans clinical isolates deficient in D
8)7
desaturase (product
of the ERG2 gene) and D
5)6
desaturase (product of the ERG3
gene) are incubated in the presence of, for example, fluconazole
(Kelly et al., 1997b). Azole antifungal agents induce the
accumulation of the membrane disturbing 14amethyl-ergosta-
8,24(28)-dien-3b,6a-diol (3,6-diol). Alterations in ERG3 prevent
the synthesis of this 3,6-diol and thus protect C. albicans against
the toxic effects of azole antifungals.
Summary of resistance to azole antifungal agents
Resistance seems to evolve over time as the result of several
alterations due to continuous selective pressure from the
antifungal used. The accumulation of alterations has been
documented by White et al. (1998) for a series of 17 clinical
C. albicans isolates. After the initial selection of a substrain of
C. albicans, the CaMDR1 efflux pump was overexpressed.
Subsequently, three simultaneous changes occurred at ERG11:
a point mutation, a gene conversion or mitotic recombination of
the ERG11 point mutation converting both alleles to the mutant
version, and overexpression of ERG11. Finally the CDR efflux
pumps are overexpressed. Perea et al. (2001b) investigated the
molecular mechanisms of resistance to azoles in C. albicans
strains displaying high-level fluconazole resistance
(MICs64 mg L). Their analysis confirmed the multifactorial
nature of fluconazole resistance and the prevalence of these
mechanisms of resistance in C. albicans isolates, with a predom-
inance of overexpression of genes encoding efflux pumps,
detected in 85% of all resistant isolates, being found. Alterations
in the target enzyme, including functional amino acid substitu-
tion and overexpression of the gene that encodes the enzyme,
were detected in 65 and 35% of the isolates, respectively.
Overall, multiple mechanisms were combined in 75% of these
C. albicans isolates.
Selectivity and adverse reactions
Ketoconazole affects, at higher doses than those needed to inhibit
ergosterol synthesis, other cytochrome P450s. For example, at
concentrations >100 n
M
it inhibits not only the mammalian
14a-demethylase (CYP51) (involved in the biosynthesis of
cholesterol), but also 17-hydroxylase-17,20-lyase (CYP17), the
cholesterol side-chain cleavage (CYP11A1) and the
11b-hydroxylase (CYP11B1) (Vanden Bossche, 1992; Vanden
Bossche et al., 1995).
Enilconazole shows high selectivity; it has 79, 98 and 108
times more affinity for the C. albicans P450s than for those of
piglet testes microsomes, bovine adrenal mitochondria and rat
liver microsomes, respectively. Parconazole is also a selective
inhibitor; this imidazole derivative has 11, 122 and >200 times
more affinity for C. albicans P450 than for those of piglet testes
microsomes, bovine adrenal mitochondria and rat liver micro-
somes (Vanden Bossche, 1991). Itraconazole inhibits at 0.01 m
M
the aromatase by 18% only (Vanden Bossche et al., 1989),
whereas at this concentration fluconazole inhibits this enzyme
by 70% (Latrille et al., 1989).
Ketoconazole is an inhibitor of CYP3A4, a major drug-
metabolizing P450 isoform. Co-administration of ketoconazole
with CYP3A4 substrates such as cyclosporine, tacrolimus,
lovastatin, terfenadine and astemisole can result in clinically
significant drug interactions (Venkatakrishnan et al., 2000).
Itraconazole is almost devoid of effects on P450-dependent
steroid biosynthesis and catabolism (Vanden Bossche et al.,
1986, 1989). However, similar to ketoconazole, itraconazole
inhibits CYP3A4 (Venkatakrishnan et al., 2000). Lamb et al.
(2000a) compared the inhibition by ketoconazole and itracon-
azole of human CYP3A4 and C. albicans CYP51 following
heterologous expression in S. cerevisiae.IC
50
-values for ketocon-
azole and itraconazole CYP3A4 inhibition were 250 n
M
and
200 n
M
, respectively. These values compared with much lower
concentrations required to obtain 50% inhibition of CYP51,
where IC
50
-values of 8 n
M
and 7.6 n
M
were observed for
ketoconazole and itraconazole, respectively. Chronic administra-
tion of itraconazole influences the metabolism of single dose
astemisole in normal volunteers; the reduction in astemisole
clearance may result in a marked increase in astemisole plasma
concentrations and QTc alterations during chronic combined
intake of astemisole with itraconazole (Lefebvre et al., 1997). A
prolonged QT-interval in ECG and symptomatic torsades de
pointes ventricular tachycardia has been reported as a conse-
quence of the interaction of itraconazole and terfenadine
(Pohjola-Sintonen et al., 1993). Ketoconazole, miconazole, hyd-
roxy-itraconazole, itraconazole and fluconazole inhibit the in vitro
metabolism of cisapride, a prokinetic drug that is predominantly
metabolized by CYP3A4 (Bohets et al., 2000).
Fluconazole and miconazole are potent inhibitors of CYP2C9.
Co-administration of phenytoin, warfarin, sulphamethoxazole
and losartan with fluconazole results in clinically significant drug
interactions (Venkatakrishnan et al., 2000). For example, flucon-
azole inhibits the metabolism of losartan to the active metabolite
E-3174, a CYP2C9-dependent reaction (Kaukonen et al., 1998).
In rats, the induction of certain cytochrome P450 isozymes
was observed 24 h after azole administration (Suzuki et al.,
2000). For example, CYP1A, CYP2B and CYP3A2, but not
CYP2E, were observed 24 h after clotrimazole, ketoconazole, or
miconazole treatment. Clotrimazole was the most potent inducer
of CYP3A and miconazole was a more potent inducer of CYP1A
and CYP2B. Ketoconazole induced CYP1A and CYP2B, but the
inducibility of ketoconazole was less than that of clotrimazole or
miconazole. Enhanced expression of CYP1A, CYP2B, CYP2C and
CYP3A has been demonstrated in small intestinal mucosa and
liver of mice treated for 3 days with 10 mg enilconazole kg day
(Muto et al., 1997).
In rat and dogs treated for periods of 3–12 months with orally
administered ketoconazole no signs of toxicity occurred with doses
up to 10 mg kg (Heel, 1982). In dogs treated daily with 40 mg kg
for 1 year appetite was reduced with emesis and decreased weight
gain. Liver weight was increased. This coincided with increased
levels of serum transaminases and alkaline phosphatase. Ketoc-
onazole is contraindicated in patients with known liver disease.
Gynaecomastia was reported in two patients during treatment
with 200 mg ketoconazole daily and in some patients receiving
Antifungal agents 17
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
high multiple doses (600–1200 mg day) (reviewed in Vanden
Bossche, 1992). This rare side-effect suggested that in such
patients ketoconazole affected androgen synthesis. Measuring
the testosterone serum levels, a dose related reversible decrease
was observed. This was related to the effect of ketoconazole on
the 17,20-lyase, a key enzyme in the androgen biosynthesis
pathway. This inhibition was observed in subcellular fractions of
rat, pig, dog, bovine and human testes, and bovine adrenals
(Vanden Bossche, 1992).
In laboratory animals the toxicity profile of itraconazole is
qualitatively similar to, but quantitatively clearly different from,
that of ketoconazole (Lampo et al., 1993). At normal doses
itraconazole causes in man only minor side-effects, most of
which are related to the gastrointestinal tract. Prophylactic
administration of itraconazole oral solution to patients with
haematological malignancy and profound neutropenia induced
drug-related adverse events in 5% of the patients (Harousseau
et al., 2000). The main drug-related events were local intoler-
ability (1.8%), nausea (0.7%) and vomiting (1.1%). Diarrhoea
induced by treatment with the oral solution is most likely due to
the hydroxypropyl-b-cyclodextrin carrier; nausea and vomiting
belong to the adverse event profile of both itraconazole capsules
and oral solution (Harousseau et al., 2000).
Fluconazole treatment is associated with very few side-effects.
The most common adverse effects are gastrointestinal, cutaneous
eruption and headache (Gupta & Summerbell, 2000).
Haematological reactions have occurred in seriously ill patients.
In a patient with renal transplant and cryptococcosis, adminis-
tration of fluconazole for 16 weeks may have contributed to a
reversible thrombocytopenia (Osterloh & Pluck, 1993).
ALLYLAMINES AND THIOCARBAMATES
The allylamines (naftifine and terbinafine) (Ryder & Mieth,
1992) and the thiocarbamates (e.g. tolnaftate) (Barrett-Bee et al.,
1986; Ryder et al., 1986) inhibit the fungal squalene epoxidase.
The prototype of the allylamines, naftifine, proved to be an
efficacious antimycotic for topical application in dermatomycoses
(Stu
¨tz, 1988).
The discovery of the antifungal activity of naftifine was the
starting point for studies on structure–activity relationships
resulting in the development of terbinafine (Fig. 3), an allylamine
with considerably enhanced antimycotic properties (Stu
¨tz, 1988).
Terbinafine has oral and topical formulations. It is strongly
lipophilic and is well distributed in the skin, fat and nails. It binds
strongly to plasma proteins and has high affinity for tissues. Oral
terbinafine is well absorbed, with 70–80% of the ingested dose
being absorbed. The peak plasma levels are reached within about
2 h. Terbinafine undergoes extensive hepatic metabolism, with
excretion being primarily (>70%) in the urine. Maximal levels in
plasma after a single oral dose of 250 or 500 mg of terbinafine are
0.9 and 1.7–2.0 mg L, respectively. Steady-state levels are
attained after 10–14 days of treatment, and low levels of
terbinafine in plasma can be measured for as long as 6 weeks after
cessation of therapy (Elewski, 1998; Gupta & Summerbell, 2000).
Terbinafine is rapidly delivered to the stratum corneum, nails
and hair – both through sebum and by direct diffusion through
dermis–epidermis (Faergemann et al., 1994).
The much older squalene epoxidase inhibitor, tolnaftate
(naphthiomate T; Fig. 3) belongs to the class of the thiocarba-
mates. It is almost insoluble in water, soluble in acetone,
chloroform and polyethylene glycol.
Terbinafine spectrum of activity
Mieth (1993) collected the MIC values obtained at different
laboratories. All dermatophytes tested were significantly more
susceptible to terbinafine than to naftifine. The MIC values
against T. rubrum,T mentagrophytes,T. verrucosum,M. canis and
E. floccosum were 0.001–0.038, 0.001–0.006, 0.001–0.006,
0.006–0.08 and 0.001–0.047 mg L, respectively. Similar results
were obtained more recently by using the NCCLS broth micro-
dilution (Jessup et al., 2000b; Ferna
´ndez-Torres et al., 2001) and
Fig. 3. Chemical structures of terbinafine,
tolnaftate, thiabendazole and flucytosine.
18 H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
macrodilution methodologies (Perea et al., 2001a). Jessup et al.
(2000b) found excellent in vitro activity against Rhizopus spp.,
Alternaria spp., Phialophora spp., Chrysosporium spp. and Exophiala
spp. Good antifungal activity is also found against
A. flavus,A. niger,A. terreus, but some A. fumigatus strains are
much less sensitive (MIC range: 0.05–5 mg L; Mieth, 1993).
Moore et al. (2001) compared the in vitro activities of terbinafine,
itraconazole and AmB against Aspergillus species. The in vitro
activity of terbinafine was inferior against A. fumigatus and
superior against A. flavus,A. terreus and A. niger. However, this
comparison should be made in vivo, as trailing endpoints are
problematic. Fusarium spp. varied greatly in their susceptibility to
terbinafine; MIC range: 0.5 to >64 mg L (Mieth, 1993).
Excellent in vitro activity was obtained against the dimorphic
fungi B. dermatitidis,C. immitis,H. capsulatum and S. schenckii
(Mieth, 1993).
For yeasts, most susceptibility studies with terbinafine showed
poor activity. Jessup et al. (2000b) found a mean MIC of
10.56 ± 0.133 mg L (0.06 to >16 mg L) for C. albicans and
Mieth (1993) found in the literature that, for the yeast form, MIC
ranged from 6.25 to >128 mg L. A number of non albicans
Candida species exhibited susceptibility to terbinafine with a mean
MIC of 4.91 ± 0.79 mg L. Candida parapsilosis and C. lusitaniae
had mean MICs of 1.16 ± 0.38 and 0.38 ± 0.4 mg L, respect-
ively. For C. guilliermondii and C. tropicalis MICs ranged from
0.25 to 2 mg L and 0.06–16 mg L. But, for C. glabrata,C. krusei
and C. kefyr the MICs were 2–16, 2–16 and >16 mg L (Jessup
et al., 2000b).
The in vitro susceptibility of C. neoformans highly depends on
the strains tested, MICs: from 0.06 to >16 mg L. Trichosporum
beigelii strains are more sensitive with MIC ranges of 0.13–
1mgL (Jessup et al., 2000b).
Minimum inhibitory concentrations of terbinafine against
M. furfur,M. sympodialis,M. slooffiae,M. pachydermatis,M. glo-
bosa,M. obtusa and M. restricta ranged from £0.03 to 64 mg L
(Gupta et al., 2000).
Tolnaftate
The use of tolnaftate is essentially restricted to dermatophyte
infections. When Sabouraud broth was used for testing the
antifungal activity of tolnaftate, the following MICs (mg L)
were found: M. canis (10–100 mg L), M. audouini (1000 mg L),
M. gypseum (1000 mg L), T. mentagrophytes (10 mg L), T. rubrum
(0.1 mg L) and T. verrucosum (0.1 mg L) (Thienpont et al.,
1975). When cultured on Sabouraud’s dextrose agar medium,
much lower MICs were found for T. mentagrophytes (0.05–
0.2 mg L), T. rubrum (0.006–0.39 mg L) and M. canis (0.05–
0.39 mg L) (Fukushiro et al., 1992).
Adverse reactions
In some patients (108 of 1583) treated with naftifine local side-
effects such as irritation, burning sensation, or dry skin were
observed (Bra¨ utigam & Weidinger, 1993).
A postmarketing surveillance study showed that gastrointest-
inal, skin and taste disturbances are the most common adverse
events related to the use of oral terbinafine in man (Hall et al.,
1997). A rare adverse reaction is hepatobiliary dysfunction
(one of 45 000–54 000 patients); this may be related to
7,7-dimethylhept-2-ene-4-ynal, a reactive metabolite of terbin-
afine (Iverson & Uetrecht, 2001).
Mechanism of action
The membrane-bound squalene epoxidase (product of the ERG1
gene) plays a key role in the biosynthetic pathway from acetate
to sterols (see Ryder, 1988; Mercer, 1991; Ryder et al., 1992;
Vanden Bossche, 1995; Favre & Ryder, 1997). The blockade of
this enzyme results in the accumulation of squalene and
deficiency of ergosterol.
Measuring the effects of naftifine, terbinafine and tolnaftate on
fungal and mammalian microsomal squalene epoxidation
revealed a high degree of selectivity for the fungal enzyme
system (Ryder et al., 1992; Nozawa & Morita, 1992). Naftifine
and terbinafine are reversible noncompetitive inhibitors of the
microsomal squalene epoxidase of C. albicans with IC
50
-values of
1.1 and 0.03 l
M
, respectively. Tolnaftate is less active with an
IC
50
-value of 5 l
M
. Terbinafine is a reversible competitive
inhibitor of the rat liver microsomal enzyme. Fifty per cent
inhibition of rat liver squalene epoxidase was achieved at
0.077 m
M
terbinafine (Ryder et al., 1992) and 50% inhibition of
the squalene epoxidase of guinea-pig liver microsomes was
achieved at 0.004 mM terbinafine (Ryder et al., 1992).
Favre and Ryder (1996) characterized the squalene epoxidase
activity from T. rubrum and its inhibition by terbinafine. Terbinafine
was a potent inhibitor, with a 50% inhibitory concentration of
15.8 n
M
. Inhibition was also demonstrated by tolnaftate (IC
50
:
51.5 n
M
). Trichophyton rubrum squaleneepoxidase was only slightly
more sensitive (approximately twofold) to terbinafine inhibition
than was the C. albicans enzyme. Therefore, this difference cannot
fully explain the much higher susceptibility 100-fold of dermat-
ophytes than of C. albicans and other yeasts to this drug.
Resistance
Resistance to allylamines has been reported only rarely. However,
the potential to develop resistance by the action of multidrug
efflux transporters does exist. Over-expression in S. cerevisiae of
the C. albicans CDR1 and CDR2 genes and of the CaMDR1 gene
confer resistance to terbinafine (Sanglard et al., 1995), suggesting
that terbinafine is a substrate for these transporters (see
discussion on resistance to azole antifungal agents).
BENZIMIDAZOLES
Benzimidazole derivatives were introduced commercially as
anthelmintics and or antifungals. They are effective at relatively
low doses for the inhibition of a broad range of fungi (literature
reviewed by Delp, 1995).
Mechanism of action
Biochemical, genetic and cytological studies provide evidence for
an effect of benzimidazoles on microtubule assembly (Davidse &
Antifungal agents 19
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
Flach, 1978). Microtubules are found in all eukaryotic cells.
They form a component of the cellular cytoskeleton and the
mitotic spindle. Microtubules are thus important for the
structural integrity of the cell and cell division. They also
function in organelle positioning and movement (Oakley &
Morris, 1980; Burland & Gull, 1984). The major component of
microtubules is tubulin which exists as a heterodimer composed
of the subunits a- and b-tubulin. Microtubules are never static,
and their constant assembly and disassembly involving tubulin
filaments and various accessory proteins is an integral part of
their function (Hollomon et al., 1998). This essential microtu-
bule property of dynamic instability is based on the binding,
hydrolysis and exchange of GTP in each tubulin dimer (Nogales,
1999). Benzimidazole derivatives interfere with microtubule
assembly by binding to b-tubulin (Davidse & Flach, 1978;
Burland & Gull, 1984; Hollomon et al., 1998). This results not
only in blockage of nuclear division but, for example, in hyphal
tip cells of F. acuminatum, also in displacement of mitochondria
from hyphal apices, disappearance of Spitzenko¨ rpers (vesicle
supply centres) and reduction of linear growth rate (Howard &
Aist, 1977, 1980).
Resistance
In plant pathogenic fungi resistance to the benzimidazole
derivatives is a major problem. Resistance is caused by point
mutations in b-tubulin (for a review see Davidse & Ishii, 1995).
In resistant fungi changes in the b-tubulin genes are restricted
to codons 198 and 200: mutations included replacement of the
glutamine at position 198 with alanine, lysine, valine or
glycine, and of the phenylalanine at position 200 with
tyrosine. The same mutations are also seen in veterinary
medicine where parasitic worms such as Haemonchus contortus
have become resistant to treatment with benzimidazoles (Kwa
et al., 1995).
Thiabendazole
Thiabendazole (Fig. 3), first sold as an anthelmintic was then
developed for postharvest fruit treatment and was later used in
veterinary medicine for the treatment of dermatoses. In some
countries thiabendazole is available as a smoke pellet or smoke
generator formulation for application in chicken hatcheries and
poultry houses (Desplenter, 1988). This benzimidazole derivative
is soluble in DMF and dimethyl sulphoxide; maximum solubility
in water at pH 2.2 is 3.84%.
Spectrum of activity
In vitro thiabendazole is active against Trichophyton spp.,
Microsporum spp., less active against A. fumigatus and inactive
against Candida spp. For example, it proved to be fungicidal on
specimens of skin scrapings infected with arthroconidia and
mycelial elements of T. verrucosum. The killing effect was
achieved using dilutions of 1:10 000 and 1:20 000 at 4 and
7 days, respectively (Gabal, 1986).
Thiabendazole inhibited A. fumigatus isolates from avian
species at concentrations between 25 and 50 mg L (Redig &
Duke, 1985).
Thiabendazole is active against some isolates of Fusarium spp.;
complete inhibition was achieved at 1.5–5 mg L (Hide et al.,
1992). In vitro tests did not show activity against M. furfur and
M. pachydermatis (Lorenzini et al., 1985). This indicates that
there is no relationship between in vitro activity and clinical
outcome. Indeed, thiabendazole in combination with other drugs
is used in topic otic preparations (de Jaham et al., 2000).
This antimicrotubule benzimidazole is also active against
Pneumocystis carinii at 10 mg L (Bartlett et al., 1992).
FLUCYTOSINE
5-Fluorocytosine (5-FC) (Fig. 3) is soluble in water
(1.5 g 100 mL at 25 C). 5-FC is well absorbed after oral
administration, penetrates well into body tissues and is excreted
mainly by the kidneys. Drugs that impair glomerular filtration
will decrease the elimination of 5-FC and thus prolong its half-life
(Vermes et al., 2000).
Spectrum of activity
5-Fluorocytosine has a narrow spectrum of activity. Using the
NCCLS M27A broth microdilution method the following MIC
ranges (mg L) were found: C. albicans (£0.125 to 64), C. glabrata
(£0.125 to 64), C. tropicalis (0.125–2), C. dubliniensis (0.125–
0.5) and C. parapsilosis (0.125–0.25) (Arthington-Skaggs et al.,
2000). Candida krusei was less sensitive (MIC range: 4–32 mg L).
Using the same NCCLS method an MIC range of 1 to >64 mg L
was found for C. neoformans (Arthington-Skaggs et al., 2000;
Barry et al., 2000). Flucytosine showed generally poor activity
against most T. beigelii isolates (Perparim et al., 1996).
A broth microdilution method was also used to determine the
in vitro activities of 5-FC against Cladosporium spp., E. dermatit-
idis,Cladophialophora spp., Paecilomyces spp. and Aspergillus spp.
(A. fumigatus and A. niger) (Llop et al., 2000). The mean MICs
were: 4.59, 0.25, 0.59, 3.17, and 9.75 mg L, respectively. 5-FC
has no activity against S. chartarum and F. solani (MICsof
256 mg L). Fusarium moniliforme,F. oxysporum and F. semitec-
tum isolates were also not sensitive to this pyrimidine derivative
(Reuben et al., 1989).
Although 5-FC inhibits growth of many Aspergillus isolates its
clinical efficacy in aspergillosis is equivocal. This difference in
sensitivity may result from the fact that in Aspergillus 5-FC exerts
only fungistatic activity whereas a fungicidal effect is seen in
yeast after prolonged contact (Polak, 1992).
Adverse effects
The severe side-effects of 5-FC include hepatotoxicity and bone
marrow depression. In most patients, these side-effects are
concentration dependent, predictable, and mostly avoidable with
close monitoring to maintain 5-FC concentrations at
20 H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
<100 mg L. These side-effects are reversible with drug dis-
continuation or reduction of the dose (Vermes et al., 2000).
Mechanism of action
5-Fluorocytosine is taken up into the cell by a cytosine permease,
an enzyme responsible for the uptake of adenine, guanine,
hypoxanthine and cytosine. Inside the cell 5-FC is rapidly
deaminated into 5-fluorouracil (5-FU) by a cytosine deaminase.
This is a critical enzyme: without conversion into 5-FU no
antifungal activity is observed. As mammalian cells have little or
no cytosine deaminase, 5-FC is fungus-specific (Polak, 1990).
After the formation of 5-FU two pathways are possible:
5-Fluorouracil can be converted by pyrimidine-processing
enzymes into 5-fluoro-dUMP which is a specific inhibitor of the
thymidylate synthase, an essential enzyme of DNA synthesis.
This inhibition is at the origin of 5-FC’s fungicidal effect. Via the
pyrimidine salvage pathway, 5-FU can be metabolized into
5-fluoroUTP, which is incorporated into RNA, thus disrupting
protein synthesis. This leads to a cytostatic effect in C. albicans,
Cryptococcus and Aspergilli (Polak, 1990).
Resistance
Monotherapy with 5-FC is limited because of the frequent
development of resistance (Vermes et al., 2000). Indeed the
incidence of primary and the selection of secondary resistance
during monotherapy with 5-FC is a major problem (Polak &
Scholer, 1980). However, in a more recent study on the in vitro
susceptibility of 8803 Candida spp. isolates a very low level of
primary resistance among virtually all Candida spp. was found
(Pfaller et al., 2002). Only for C. krusei a high level of primary
resistance was found: 28% of the 184 isolates tested were
resistant (Pfaller et al., 2002). In C. neoformans 1.8% of strains
tested were found to be naturally resistant, but during treatment
resistant mutants are easily selected (Polak & Scholer, 1980).
The lowest 5-FC resistance frequency is seen in C. albicans,
followed by C. neoformans, and the highest is seen in Aspergillus
(Polak & Hartman, 1991).
Intrinsic resistance to 5-FC is usually the result of a defect in
the cytosine deaminase. The most common enzyme associated
with secondary resistance to 5-FC in clinical isolates of C.
albicans is a decrease in the activity of the uracyl phosphor-
ibosyl transferase (UPRTase). This enzyme is involved in the
synthesis of 5-FUMP and 5-FdUMP (Whelan, 1987; Vanden
Bossche et al., 1987a, 1994a). Resistance in C. neoformans
usually results from the loss of cytosine deaminase or UPRTase
(Whelan, 1987).
ACKNOWLEDGEMENTS
We would like to thank Dr Lieven Meerpoel for providing us with
the chemical structures of the antifungal agents. The assistance
of Christel Herman in preparing this manuscript is gratefully
acknowledged.
REFERENCES
Albertson, G.D., Niimi, M., Cannon, R.D. & Jenkinson, H.F. (1996)
Multiple efflux mechanisms are involved in Candida albicans
fluconazole resistance. Antimicrobial Agents and Chemotherapy,40,
2300–2305.
Andreoli, T.E. (1974) The structure and function of amphotericin
B-cholesterol pores in lipid bilayer membranes. Annals of the New York
Academy of Sciences,235, 448–468.
Aparicio, J.F., Colina, A.J., Ceballos, E. & Martin, J.F. (1999) The bio-
synthetic gene cluster for the 26-membered ring polyene macrolide
pimaricin. A new polyketide synthase organisation encoded by two
subclusters separated by functionalization genes. Journal of Biological
Chemistry,274, 10133–10139.
Aparicio, J.F., Fouces, R., Mendes, M.V., Olivera, N. & Martin, J.F. (2000)
A complex multienzyme system encoded by five poliketide synthase
genes is involved in the biosynthesis of the 26-membered polyene
macrolide pimaricin in Streptomyces natalensis.Chemistry and Biology,
7, 895–905.
Arias, A., Arevalo, M.P., Andreu, A., Rodriguez, C. & Sierra, A. (1996)
Candida glabrata: in vitro susceptibility of 84 isolates to eight antifungal
agents. Chemotherapy,42, 107–111.
Armstrong, D. & Schmitt, H.J. (1990) Older drugs. In Chemotherapy of
Fungal Diseases. Ed. Ryley, J.F. pp. 439–454. Springer Verlag, Berlin.
Arthington-Skaggs, B.A., Motley, M., Warnock, D.W. & Morrison, C.J.
(2000) Comparative evaluation of PASCO and national committee for
clinical laboratory standards M27A broth microdilution methods for
antifungal drug susceptibility testing of yeasts. Journal of Clinical
Microbiology,38, 2254–2260.
Artis, W.M., Odle, B.M. & Jones, H.E. (1981) Griseofulvin resistant der-
matophytosis correlates with in vitro resistance. Archives of Dermatol-
ogy,117, 16–19.
Asai, K., Tsuchimori, N., Okonogi, K., Perfect, J.R., Gotoh, O. & Yoshida,
Y. (1999) Formation of azole-resistant Candida albicans by mutation of
sterol 14-demethylase P450. Antimicrobial Agents and Chemotherapy,
43, 1163–1169.
Barchiesi, F., Calabrese, D., Sanglard, D., Falconi di Francesco, L.,
Caselli, F., Giannini, D., Giacometti, S., Gavaudan, S. & Scalise, G.
(2000) Experimental induction of fluconazole resistance in Candida
tropicalis ATCC750. Antimicrobial Agents and Chemotherapy,44,
1578–1584.
Barchiesi, F., Colombo, A.L., McGough, D.A., Fothergill, A.W. & Rinaldi,
M.G. (1994) In vitro activity of itraconazole against fluconazole-sus-
ceptible and – resistant Candida albicans isolates from oral cavities of
patients with human immunodeficiency virus. Antimicrobial Agents and
Chemotherapy,38, 1530–1533.
Barrett-Bee, K.J., Lane, A.C. & Turner, R.W. (1986) The mode of action of
tolnaftate. Journal of Medical and Veterinary Mycology,24, 155–160.
Barry, A.L., Pfaller, M.A., Brown, S.D., Espinel-Ingroff, A., Ghannoum,
M.A., Knapp, C., Rennie, R.P., Rex, J.H. & Rinaldi, M.G. (2000)
Quality control limits for broth microdilution susceptibility tests of
ten antifungal agents. Journal of Clinical Microbiology,38, 3457–
3459.
Bartlett, M.S., Edlind, T.D., Durkin, M.M., Shaw, M.M., Queener, S.F. &
Smith, J.W. (1992) Antimicrotubule benzimidazoles inhibit in vitro
growth of Pneumocystis carinii.Antimicrobial Agents and Chemotherapy,
36, 779–782.
Barug, D., Samson, R.A. & Kerkenaar, A. (1983) Microscopic studies of
Candida albicans and Torulopsis glabrata after in vitro treatment with
bifonazole. Arzneimittel-Forschung,33, 528–537.
Bennett, J.E. (1992) Developing drugs for the deep mycoses: a short
history. In New Strategies in Fungal Disease. Eds Bennett, J.E., Hay, R.J.
& Peterson, P.K. pp. 3–12. Churchill Livingstone, Edinburgh.
Antifungal agents 21
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
Boelaert, J., Daneels, R., Van Landuyt, H. & Symoens, J. (1976) Mico-
nazole plasma levels in healthy subjects and patients with impaired
renal function. Chemotherapy,6, 165–169.
Bohets, H., Lavrijsen, K., Hendrickx, J., van Houdt, J., van Genechten, V.,
Verboven, P., Meuldermans, W. & Heykants, J. (2000) Identification of
the cytochrome P450 enzymes involved in the metabolism of cisapride:
in vitro studies of potential co–medication interactions. British Journal
of Pharmacology,129, 1655–1667.
Bolard, J. (1986) How do the polyene macrolide antibiotics affect the cel-
lular membrane properties? Biochimica Biophysica Acta,864, 257–304.
Borgers, M., Vanden Bossche, H. & Cauwenbergh, G. (1986) The phar-
macology of agents used in the treatment of pulmonary mycoses.
Clinics in Chest Medicine,7, 439–452.
Bouchara, J.P., Zouhair, R., Le Boudouil, S., Renier, G., Filmon, R.,
Chabasse, D., Hallet, J.B. & Defontaine, A. (2000) In-vivo selection of
an azole-resistant petite mutant of Candida glabrata.Journal of Medical
Microbiology,49, 977–984.
Brajtburg, J., Gruda, I. & Kobayashi, G.S. (1992) Amphotericin B: studies
focused on improving its therapeutic efficacy. In Recent Progress in
Antifungal Chemotherapy. Eds Yamaguchi, H., Kobayashi, G.S. &
Takahashi, H. pp. 65–75. Marcel Dekker, Inc., New York.
Brajtburg, J., Powderly, W.G., Kobayashi, G.S. & Medoff, G. (1990)
Amphotericin B: current understanding of mechanisms of action.
Antimicrobial Agents and Chemotherapy,34, 183–188.
Brammer, K.W. & Tarbit, M.H. (1987) A review of the pharmacokinetics
of fluconazole (UK-49,858) in laboratory animals and man. In Recent
Trends in the Discovery and Evaluation of Antifungal Agents. Ed. Frontling,
R.A. pp. 141–149. J.R. Prous Science Publishers S.A., Barcelona.
Brandao, R.L., Castro, I.M., Passos, J.R., Nicoli, J.R. & Thevelein, J.M.
(1992) Glucose-induced activation of the plasma membrane
H
+
-ATPase in Fusarium oxysporum.Journal of General Microbiology,
138, 1579–1586.
Brasseur, R., Vandenbosch, C., Vanden Bossche, H. & Ruysschaert, J.M.
(1983) Mode of insertion of miconazole, ketoconazole and deacylated
ketoconazole in lipid layers. Biochemical Pharmacology,32, 2175–
2180.
Brautaset, T., Sekurova, O.N., Sletta, H., Ellingsen, T.E., StrLm, A.M.,
Valla, S. & Zotchev, S.B. (2000) Biosynthesis of the polyene antifungal
antibiotic nystatin in Streptomyces noursei ATCC 11455: analysis of the
gene cluster and deduction of the biosynthetic pathway. Chemistry and
Biology,7, 395–403.
Bra¨ utigam, M. & Weidinger, G. (1993) Clinical experience with naftifine.
In Cutaneous Antifungal Agents. Eds Rippon, J.W. & Fromtling, R.A. pp.
99–115. Marcel Dekker, Inc., New York.
Bu
¨chel, K.W., Draber, W., Regel, E. & Plempel, E. (1972) Synthesen und
Eigenschaften von Clotrimazol und weiteren Antimykotischen 1-Tri-
phenylmethylimidazolen. Arzneimittel-Forschung,22, 1260–1272.
Burghoorn, H.P., Soteropoulos, P., Paderu, P., Kashiwazaki, R. & Perlin,
D.S. (2002) Molecular evaluation of the plasma membrane proton
pump from Aspergillus fumigatus.Antimicrobial Agents and Chemother-
apy,46, 615–624.
Burland, T.G. & Gull, K. (1984) Molecular and cellular aspects of the
interaction of benzimidazole fungicides with tubulin and microtubules.
In Mode of Action of Antifungal Agents. Eds Trinci, A.P.J. & Ryley, J.F. pp.
299–320. Cambridge University Press, Cambridge.
Burt, E.T. (2001) Current approaches to the treatment of Candida albicans
infections. Expert Opinion on Therapeutic Patents,11, 269–282.
Chang, R.L., Zampaglione, N. & Lin, C. (1975) Correlation of
14
C-gri-
seofulvin metabolism in rat liver microsomes, isolated perfused rat
livers, and in rats with bile duct cannulas. Drug Metabolism and
Disposition,3, 487–493.
Chapman, S.W., Rogers, P.D., Rinaldi, M.G. & Sullivan, D.C. (1998)
Susceptibilities of clinical and laboratory isolates of Blastomyces
dermatitidis to ketoconazole, itraconazole, and fluconazole. Anti-
microbial Agents and Chemotherapy,42, 978–980.
Chaudhuri, A.R. & Luduena, R.F. (1996) Griseofulvin: a novel interac-
tion with bovine brain tubulin. Biochemical Pharmacology,51, 903–
909.
Chiew, Y.Y., Sullivan, P.A. & Shepherd, M.G. (1982) The effect of
ergosterol and alcohols on germ-tube formation and chitin synthase in
Candida albicans.Canadian Journal of Biochemistry,60, 15–20.
Cornelissen, F. & Vanden Bossche, H. (1983) Synergism of the anti-
microbial agents miconazole, bacitracin and polymyxin B. Che-
motherapy,29, 419–427.
Crackower, S.H.B. (1972) The effect of griseofulvin on mitosis in
Aspergillus nidulans.Canadian Journal of Microbiology,18, 683–687.
Daneshmend, T.K. & Warnock, D.W. (1988) Clinical pharmacokinetics of
ketoconazole. Clinical Pharmacokinetics,14, 13–34.
Daum, G., Lees, N.D., Bard, M. & Dickson, R. (1998) Biochemistry, cell
biology and molecular biology of lipids of Saccharomyces cerevisiae.
Yeast,14, 1471–1510.
Davidse, L.C. & Flach, W. (1978) Interaction of thiabendazole with
fungal tubulin. Biochimica and Biophysica Acta,543, 82–90.
Davidse, L.C. & Ishii, H. (1995) Biochemical and molecular aspects of the
mechanisms of action of benzimidazoles, N-phenylcarbamates and
N-phenylformamidoximes and the mechanism of resistance to these
compounds in fungi. In Modern Selective Fungicides. Ed. Lyr, H. pp.
305–322. Gustav Fischer Verlag, Jena.
De Backer, M.D., Ilyina, T., Ma, X.-J., Vandoninck, S., Luyten, W.H.M.L.
& Vanden Bossche, H. (2001) Genomic profiling of the response of
Candida albicans to itraconazole treatment using a DNA micro-array.
Antimicrobial Agents and Chemotherapy,45, 1660–1670.
De Carli, L. & Larizza, L. (1988) Griseofulvin. Mutation Research,195,
91–126.
De Matteis, F. (1982) Loss of microsomal components in drug-induced
liver damage, in cholestasis and after administration of chemicals
which stimulate heme catabolism. In Hepatic Cytochrome P450
Monooxyganese System. Eds Schenkman, J.B. & Kupfer, D. pp. 307–
340. Pergamon Press, Oxford.
De Nollin, S., Van Belle, H., Goosens, F., Thone
´, F. & Borgers, M. (1977)
Cytochemical and biochemical studies of yeasts after in vitro exposure
to miconazole. Antimicrobial Agents and Chemotherapy,11, 500–513.
Debruyne, D. & Coquerel, A. (2001) Pharmacokinetics of antifungal
agents in onychomycoses. Clinical Pharmacokinetics,40, 441–472.
Decottignies, A. & Goffeau, A. (1997) Complete inventory of the yeast
ABC proteins. Nature Genetics,15, 137–145.
Del Sorbo, G., Andrade, A.C., Van Nistelrooy, J.G.M., Van Fan, J.A.L.,
Balzi, E. & De Waard, M.A. (1997) Multidrug resistance in Aspergillus
nidulans involves novel ATP-binding cassette transporters. Molecular
and General Genetics,254, 417–426.
Delp, C.J. (1995) Benzimidazole and related fungicides. In Modern Selective
Fungicides. Ed. Lyr, H. pp. 291–303. Gustav Fischer Verlag, Jena.
Denning, D.W., Venkateswarlu, K., Oakley, K.L., Anderson, M.J., Man-
ning, N.J., Stevens, D.A., Warnock, D.W. & Kelly, S.L. (1997) Itraco-
nazole resistance in Aspergillus fumigatus.Antimicrobial Agents and
Chemotherapy,41, 1364–1368.
Desplenter, L. (1988) Aspergillosis in animal health: environmental con-
trol. In Aspergillus and Aspergillosis. Eds Vanden Bossche, H., Mackenzie,
D.W.R. & Cauwenbergh, G. pp. 289–297. Plenum Press, New York.
Desplenter, L. & Van Cutsem, J. (1980) In vitro antifungal activity of
parconazole (R39500). In Preclinical Research Report. Janssen Research
Products Information Service. Ed. Cauwenbergh, G. pp. 289–297. Ple-
num Press, New York.
Dick, J.D., Merz, W.G. & Saral, R. (1980) Incidence of polyene-resistant
yeasts recovered from clinical specimens. Antimicrobial Agents and
Chemotherapy,18, 158–163.
22 H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
Dick, J.D., Rosengard, B.R., Merz, W.G., Stuart, R.K., Hutchins, G.M. &
Saral, R. (1985) Fatal disseminated candidiasis due to amphotericin
B-resistant Candida guilliermondii.Annals of Internal Medicine,102, 67–
68.
Dupont, B. (2002) Overview of the lipid formulations of amphotericin B.
Journal of Antimicrobial Chemotherapy,49 (Suppl. S1),31–36.
Dupont, B., Richardson, M., Verweij, P.E. & Meis, J.F.G.M. (2000)
Invasive aspergillosis. Medical Mycology,38 (Suppl. 1),215–224.
Edlind, T.D., Henry, K.W., Metera, K.A. & Katiyar, S.K. (2001) Aspergillus
fumigatus CYP51 sequence: potential basis for fluconazole resistance.
Medical Mycology,39, 299–302.
Elewski, B.E. (1998) Onychomycosis: pathogenesis, diagnosis, and
management. Clinical Microbiology Reviews,11, 415–429.
Ellis, D. (2002) Amphotericin B: spectrum and resistance. Journal of
Antimicrobial Chemotherapy,49 (Suppl. S1),7–10.
Espinel-Ingroff, A., Dawson, K., Pfaller, M. et al. (1995) Comparative and
collaborative evaluation of standardization of antifungal susceptibility
testing for filamentous fungi. Antimicrobial Agents and Chemotherapy,
39, 314–319.
Espinel-Ingroff, A., Warnock, D.W., Vaquez, J.A. & Arthington-Skaggs,
B.A. (2000) In vitro antifungal susceptibility methods and clinical
implications of antifungal resistance. Medical Mycology,38 (Suppl. 1),
293–304.
Faergemann, J., Zehender, H. & Millerioux, L. (1994) Levels of terbinafine
in plasma, stratum corneum, dermis-epidermis (without stratum cor-
neum), sebum, hair and nails during and after 250 mg terbinafine
orally once daily for 7 and 14 days. Clinical and Experimental Derma-
tology,19, 121–126.
Favre, B. & Ryder, N.S. (1996) Characterization of squalene epoxidase
activity from the dermatophyte Trichophyton rubrum and its inhibition
by terbinafine and other antimycotic agents. Antimicrobial Agents and
Chemotherapy,40, 443–447.
Favre, B. & Ryder, N.S. (1997) Characterization of squalene epoxidase
activity from the pathogenic yeast Candida albicans.Gene,189, 119–126.
Ferna
´ndez-Torres, B., Carrillo, A.J., Martin, E., Del Palacio, A., Moore,
M.K., Valverde, A., Serrano, M. & Guarro, J. (2001) In vitro activities
of 10 antifungal drugs against 508 dermatophyte strains. Antimicrobial
Agents and Chemotherapy,45, 2524–2528.
Ferna
´ndez-Torres, B., Va
´zquez-Veiga, H., Llovo, X., Pereiro, M., Jr &
Guarro, J. (2000) In vitro susceptibility to itraconazole, clotrimazole,
ketoconazole and terbinafine of 100 isolates of Trichophyton rubrum.
Chemotherapy,46, 390–394.
Fournier, I., Barwicz, J. & Tancredo, P. (1998) The structuring effects of
amphotericin B on pure and ergosterol- or cholesterol-containing
dipalmitoylphosphatidylcholine bilayers: a differential scanning
calorimetry study. Biochimica Biophysica Acta,1373, 76–86.
Fukushiro, R., Urabe, H., Kagawa, S., Watanabe, S., Takahashi, H.,
Takahashi, S. & Nakajima, H. (1992) Butenafine hydrochloride, a new
antifungal agent: clinical and experimental study. In Recent Progress in
Antifungal Chemotherapy. Eds Yamaguchi, H. & Kobayashi, G.S &
Takahashi, H. pp. 147–157. Marcel Dekker, Inc., New York.
Gabal, M.A. (1986) Study on the evaluation of the use of thiabendazole
in the treatment and control of bovine dermatophytosis. Mycopatho-
logia,93, 163–168.
Galgiani, J.N. & Stevens, D.A. (1976) Antimicrobial susceptibility testing
of yeasts: a turbidimetric technique independent of inoculum size.
Antimicrobial Agents and Chemotherapy,10, 721–726.
Gentles, J.C. (1956) Experimental ringworm in guinea pig: oral treatment
with griseofulvin. Nature,182, 476–477.
Georgopapadakou, N.H., Dix, B.A., Smith, S.A., Freudenberger, J. &
Funke, P.T. (1987) Effect of antifungal agents on lipid biosynthesis and
membrane integrity in Candida albicans.Antimicrobial Agents and Che-
motherapy,31, 46–51.
Godefroi, E.F., Heeres, J., Van Cutsem, J. & Janssen, P.A.J. (1969) The
preparation and antimycotic properties of derivatives of 1-phenethy-
limidazole. Journal of Medicinal Chemistry,12, 784–791.
Goffeau, A., Park, J., Paulsen, I.T., Jonniaux, J.L., Dinh, T., Mordant, P. &
Saier, M.H. (1997) Multidrug-resistant transport proteins in yeast:
complete inventory and phylogenetic characterization of yeast open
reading frames within the major facilitator superfamily. Yeast,13, 43–
54.
Goffeau, A. & Slayman, C.W. (1981) The proton translocating ATPase of
the fungal plasma membrane. Biochimica Biophysica Acta,639, 197–
223.
Goldberg, J., Connolly, P., Schnizlein, C., Durkin, M., Kohler, S., Sme-
dema, M., Brizendine, E., Hector, R. & Wheat, J. (2000) Comparison of
nikkomycin Z with amphotericin B and itraconazole for treatment of
histoplasmosis in a murine model. Antimicrobial Agents and Che-
motherapy,44, 1624–1629.
Granade, T.C. & Artis, W.M. (1980) Antimycotic susceptibility testing of
dermatophytes in microcultures with a standardized fragmented my-
celial inoculum. Antimicrobial Agents and Chemotherapy,17, 725–729.
Groll, A.H., Petraitis, V., Petraitiene, R., Field-Ridley, A., Calendario, M.,
Bacher, J., Piscitelli, S.C. & Walsh, T.J. (1999) Safety and efficacy of
multilamellar liposomal nystatin against disseminated candidiasis in
persistently neutropenic rabbits. Antimicrobial Agents and Chemother-
apy,43, 2463–2467.
Groll, A.H., Wood, L., Roden, M., Mickiene, D., Chiou, C.C., Townley, E.,
Dad, L., Piscitelli, S.C. & Walsh, T.J. (2002) Safety, pharmacokinetics,
and pharmacodynamics of cyclodextrin itraconazole in pediatric
patients with oropharyngeal candidiasis. Antimicrobial Agents and
Chemotherapy,46, 2554–2563.
Gull, K. & Trinci, A.J.P. (1973) Griseofulvin inhibits fungal mitosis.
Nature,244, 292–294.
Gupta, A.K., Kohli, Y., Faergemann, J. & Summerbell, R.C. (2000) In
vitro susceptibility of the seven Malassezia species to ketoconazole,
voriconazole, itraconazole and terbinafine. British Journal of Derma-
tology,142, 758–765.
Gupta, A.K. & Summerbell, R.C. (2000) Tinea capitis. Medical Mycology,
38, 255–287.
Hall, M., Monka, C., Krupp, P. & O’Sullivan, D. (1997) Safety of oral
terbinafine. Results of a postmarketing surveillance study in 25884
patients. Archives of Dermatology,133, 1213–1219.
Hamilton-Miller, J.M.T. (1973) Chemistry and biology of the polyene
macrolide antibiotics. Bacteriological Reviews,37, 166–196.
Hammer, K.A., Carson, C.F. & Riley, T.V. (2000) In vitro activities of
ketoconazole, econazole, miconazole, and Melaleuca alterifolia (tea tree)
oil against Malassezia species. Antimicrobial Agents and Chemotherapy,
44, 467–469.
Harousseau, J.L., Dekker, A.W., Stamatoullas-Bastard, A. et al. (2000)
Itraconazole oral solution for primary prophylaxis of fungal infections
in patients with hematological malignancy and profound neutropenia:
a randomized, double-blind, double-placebo, multicenter trial com-
paring itraconazole and amphotericin B. Antimicrobial Agents and
Chemotherapy,44, 1887–1893.
Hazen, E.L. & Brown, R. (1950) Two antifungal agents produced by a soil
actinomycete. Science,112, 423.
Heel, R.C. (1982) Ketoconazole: pharmacological profile. In Ketoconazole
in the Management of Fungal Disease. Eds Drouhet, E., Hay, R.J., Jones,
H.E. & Restrepo, A. pp. 55–77. ADIS Press, New York.
Heeres, J., Backx, L.J.J., Mostmans, J.H. & Van Cutsem, J. (1979) Anti-
mycotic imidazoles. Part 4. Synthesis and antifungal activity of keto-
conazole, a new potent orally active broad-spectrum antifungal agent.
Journal of Medicinal Chemistry,22, 1003–1005.
Hennequin, C., Benailly, N., Silly, C., Sorin, M., Scheinmann, P., Lenoir,
G., Gaillard, J.L. & Berche, P. (1997) In vitro susceptibility to
Antifungal agents 23
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
amphotericin B, itraconazole, and miconazole in filamentous fungi
isolated from patients with cystic fibrosis. Antimicrobial Agents and
Chemotherapy,41, 2064–2066.
Heykants, J., Michiels, M., Meuldermans, W., Monbaliu, J., Lavrijsen, K.,
Van Peer, A., Levron, J.C., Woestenborghs, R. & Cauwenbergh, G.
(1987) The pharmacokinetics of itraconazole in animals and men: an
overview. In Recent Trends in the Discovery and Evaluation of Antifungal
Agents. Ed. Fromtling, R.A. pp. 223–249. J.R. Prous Science Publishers
S.A., Barcelona.
Heykants, J., Van Peer, A., Van de Velde, V., Van Rooy, P., Meuldermans,
W., Lavrijsen, K., Woestenborghs, J., Van Cutsem, J. & Cauwenbergh,
G. (1989) The clinical pharmacokinetics of itraconazole: an overview.
Mycoses,32 (Suppl. 1),67–87.
Hide, C.A., Read, P.J. & Hall, S.M. (1992) Resistance to thiabendazole in
Fusarium species isolated from potato tubers affected by dry rot. Plant
Pathology,41, 745–748.
Hitchcock, C.A. (1991) Cytochrome P450-dependent 14a-sterol deme-
thylase of Candida albicans and its interaction with azole antifungals.
Biochemical Society Transactions,19, 782–787.
Hitchcock, C.A. & Whittle, P.J. (1993) Chemistry and mode of action of
fluconazole. In Cutaneous Antifungal Agents. Eds Rippon, J.W. &
Fromtling, R.A. pp. 183–197. Marcel Dekker, Inc., New York.
Hollomon, D.W., Butters, J.A., Barker, H. & Hall, L. (1998) Fungal
b-tubulin as a fusion protein, binds benzimidazole and phenylcarbamate
fungicides. Antimicrobial Agents and Chemotherapy,42, 2171–2173.
Horsburgh, C.R., Jr & Kirkpatrick, C.H. (1983) Long-term therapy of
chronic mucocutaneous candidiasis with ketoconazole: experience
with twenty-one patients. American Journal of Medicine,74 (Suppl. 1B),
23–29.
Howard, R.J. & Aist, J.R. (1977) Effects of MBC on hyphal tip organiza-
tion, growth and mitosis of Fusarium acuminatum, and their antag-
onism by D
2
O. Protoplasma,92, 195–210.
Howard, R.J. & Aist, J.R. (1980) Cytoplasmic microtubules and fungal
morphogenesis: ultrastructural effects of methyl-benzimidazol-2yl
carbamate determined by freeze-substitution of hyphal tip cells. Journal
of Cellular Biology,87, 55–64.
Hubbard, M.J., Surarit, T.R., Sullivan, P.A. & Shepherd, M.G. (1986) The
isolation of plasma membrane and characterisation of the plasma
membrane ATPase from the yeast Candida albicans.European Journal of
Biochemistry,154, 375–381.
Iverson, S.L. & Uetrecht, J.P. (2001) Identification of a reactive meta-
bolite of terbinafine: insights into terbinafine-induced hepatotoxicity.
Chemical Research in Toxicology,14, 175–181.
de Jaham, C., Paradis, M. & Papich, M.G. (2000) Antifungal dermatologic
agents: azoles and allylamines. Compendium on Continuing Education for
the Practicing Veterinarian,22, 548–559.
Jessup, C.J., Warner, J., Isham, N., Hasan, I. & Ghannoum, M.A. (2000a)
Antifungal susceptibility testing of dermatophytes: establishing a
medium for inducing conidial growth and evaluation of susceptibility
of clinical isolates. Journal of Clinical Microbiology,38, 341–344.
Jessup, C.J., Ryder, N.S. & Ghannoum, M.A. (2000b) An evaluation of
the in vitro activity of terbinafine. Medical Mycology,38, 155–159.
Johnson, E.M., Ojwang, J.O., Szekely, A., Wallace, T.L. & Warnock, D.W.
(1998) Comparison of in vitro antifungal activities of free and lipo-
some-encapsulated nystatin with those of four amphotericin B for-
mulations. Antimicrobial Agents and Chemotherapy,42, 1412–1416.
Johnson, E.M., Warnock, D.W., Luker, J., Porter, S.R. & Scully, C. (1995)
Emergence of azole drug resistance in Candida species from HIV-
infected patients receiving prolonged fluconazole therapy for oral
candidosis. Journal of Antimicrobial Chemotherapy,35, 103–114.
Kakeya, H., Miyazaki, Y., Miyazaki, H., Nyswaner, K., Grimberg, B. &
Bennett, J.E. (2000) Genetic analysis of azole resistance in the
Darlington strain of Candida albicans.Antimicrobial Agents and
Chemotherapy,44, 2985–2990.
Katiyar, S.K. & Edlind, T.D. (2001) Identification and expression of
multidrug resistance-related ABC transporter genes in Candida krusei.
Medical Mycology,39, 109–116.
Kaukonen, K.M., Olkkola, K.T. & Neuvonen, P.J. (1998) Fluconazole but
not itraconazole decreases the metabolism of losartan to E-3174.
European Journal of Clinical Pharmacology,53, 445–449.
Kelly, S.L., Lamb, D.C., Baldwin, B.C., Corran, A.J. & Kelly, D.E. (1997a)
Characterization of Saccharomyces cerevisiae CYP61, sterol D22-
desaturase, and inhibition by azole antifungal agents. Journal of Bio-
logical Chemistry,272, 9986–9988.
Kelly, S.L., Lamb, D.C. & Kelly, D.E. (1999a) Y132H substitution in Candida
albicans sterol 14a-demethylase confers fluconazole resistance by pre-
venting binding to haem. FEMS Microbiology Letters,180, 171–175.
Kelly, S.L., Lamb, D.C., Kelly, D.E., Mannung, N.J., Lo¨ ffler, J., Hebart, H.,
Schumacher, U. & Einsele, H. (1997b) Resistance to fluconazole and
cross-resistance to amphotericin B in Candida albicans from AIDs
patients caused by defective sterol delta 5,6-desaturation. FEBS Letters,
400, 80–82.
Kelly, S.L., Lamb, D.C., Loeffler, J., Einsele, H. & Kelly, D.E. (1999b) The
G464S amino acid substitution in Candida albicans sterol 14a-demethy-
lase causes fluconazole resistance in the clinic through reduced affinity.
Biochemical Biophysical Research Communications,262, 174–179.
Kerridge, D. & Whelan, W.L. (1984) The polyene macrolide antibiotics
and 5-fluorocytosine: molecular actions and interactions. In Mode of
Action of Antifungal Agents. Eds Trinci, A.P.J. & Ryley, J.F. pp. 343–375.
Cambridge University Press, Cambridge.
Klepser, M.E., Ernst, E.J. & Pfaller, M.A. (1997) Update on antifungal
resistance. Trends in Microbiology,5, 372–375.
Kollar, R., Petrakova, E., Ashwell, G., Robbins, P.W. & Cabib, E. (1995)
Architecture of the yeast cell wall. The linkage between chitin and b
(1–3) -glucan. Journal of Biological Chemistry,270, 1170–1178.
Kollias, G.V., Palgut, J. & Ross, J. (1986) The use of ketoconazole in birds:
prelimimirary pharmacokinetics and clinical applications. Proceedings
of the Annual Meeeting of the Association of Avian Veterinarians, Miami,
USA, pp. 103–104.
Korting, H.C., Ollert, M. & Abeck, D. and the German collaborative der-
mathophyte drug susceptibility Study Group (1995) Results of German
multicenter study of antimicrobial susceptibilities of Trichophyton
rubrum and Trichophyton mentagrophytes strains causing tinea
unguium. Antimicrobial Agents and Chemotherapy,39, 1206–1208.
Kwa, M.S.G., Veenstra, J.G., van Dujk, M. & Roos, M.H. (1995) b-Tubulin
genes from the parasitic nematode Haemonchus contortus modulate
drug resistance in Caenorhabditis elegans.Journal of Molecular Biology,
246, 500–510.
Lamb, D.C., Kelly, D.E., Baldwin, B.C. & Kelly, S.L. (2000a) Differential
inhibition of human CYP3A4 and Candida albicans CYP51 with azole
antifungal agents. Chemical Biological Interactions,125, 165–175.
Lamb, D.C., Kelly, D., Schunck, W.H., Shyadehi, A.Z., Akhtar, M., Lowe,
D.J., Baldwin, B.C. & Kelly, S.L. (1997) The mutation T315A in Can-
dida albicans sterol 14a-demethylase causes reduced enzyme activity
and fluconazole resistance through reduced affinity. Journal of Biolo-
gical Chemistry,272, 5682–5688.
Lamb, D.C., Kelly, D.E., White, T.C. & Kelly, S.L. (2000b) The R464K
amino acid substitution in Candida albicans sterol 14a-demethylase
causes drug resistance through reduced affinity. Antimicrobial Agents
and Chemotherapy,44, 63–67.
Lampo, A., De Beule, K., Cauwenbergh, G. & Van Cauteren, H. (1993)
Safety of itraconazole in animals and humans. In Cutaneous Antifungal
Agents. Eds Rippon, J.W. & Fromtling, R.A. pp. 285–293. Marcel
Dekker, Inc., New York.
24 H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
Langcake, P., Kuhn, P.L. & Wade, M. (1983) The mode of action of
systemic fungicides. In Progress in Pesticide Biochemistry and Toxicology,
Vol. 3. Eds Hutson,D.H. & Roberts, T.R. pp. 1–109. Wiley, Chichester.
Latrille, F., Charuel, C. & Lodola, A. (1989) A comparative study of the
effects of ketoconazole and fluconazole on 17b-estradiol production by
rat ovaries in vitro. Research Communications in Chemical Pathology and
Pharmacology,64, 173–176.
Lefebvre, R.A., Van Peer, A. & Woestenborghs, R. (1997) Influence of
itraconazole on the pharmacokinetics and electrocardiographic effects
of astemizole. British Journal of Clinical Pharmacology,43, 319–322.
Levron, J.C., Stephan, P. & Razanamiraka, G. (1985) Pharmacokinetic
study of parconazole, antifungal drug, in guinea fowl. Third EAVPT
Congress, Ghent, Belgium, Session 9, p. 120.
Li, R.K., Ciblak, M.A., Nordoff, N., Pasarell, L., Warnock, D.W. &
McGinnis, M.R. (2000) In vitro activities of voriconazole, itraconazole,
and amphotericin B against Blastomyces dermatitidis,Coccidioides
immitis, and Histoplasma capsulatum.Antimicrobial Agents and Che-
motherapy,44, 1734–1736.
Lin. C., Magat, R., Chang, J., McGlotten, J. & Symchowicz, S. (1973)
Absorption, metabolism and excretion of
14
C-griseofulvin in man.
Journal of Pharmacology and Experimental Therapeutics,187, 415–422.
Llop, C., Pujol, I., Aguilar, C., Sala, J., Riba, D. & Guarro, J. (2000)
Comparison of three methods of determining MICs for filamentous
fungi using different end point criteria and incubation methods.
Antimicrobial Agents and Chemotherapy,44, 239–242.
Lo¨ ffler, J., Kelly, S.L., Hebart, H., Schumacher, U., Lass-Flo¨ rl, C. & Einsele,
H. (1997) Molecular analysis of cyp51 from fluconazole-resistant
Candida albicans strains. FEMS Microbiology Letters,151, 263–268.
Lo
´pez-Ribot, J.L., McAtee, R.K., Perea, S., Kirkpatrick, W.R., Rinaldi, M.G.
& Patterson, T.F. (1999) Multiple resistant phenotypes of Candida
albicans coexist during episodes of oropharyngeal candidiasis in human
immunodeficiency virus-infected patients. Antimicrobial Agents and
Chemotherapy,43, 1621–1630.
Lorenzini, R., Mercantini, R. & De Bernardis, F. (1985) In vitro sensitivity
of Malassezia spp. to various antimycotics. Drugs and Experimental
Clinical Research,11, 393–395.
Lyons, C.N. & White, T.C. (2000) Transcriptional analyses of antifungal
drug resistance in Candida albicans.Antimicrobial Agents and
Chemotherapy,44, 2296–2303.
Marichal, P. (1999) Mechanisms of resistance to azole antifungal com-
pounds. Current Opinion in Anti-Infective Investigational Drugs,1, 318–
333.
Marichal, P., Gorrens, J., Coene, M.-C., Le Jeune, L. & Vanden Bossche, H.
(1995) Origin of differences in susceptibility of Candida krusei to azole
antifungal agents. Mycoses,38, 111–117.
Marichal, P., Gorrens, J., Laurijssens, L., Vermuyten, K., Van Hove, C., Le
Jeune, L., Verhasselt, P., Sanglard, D., Borgers, M., Ramaekers, F.C.,
Odds, F. & Vanden Bossche, H. (1999a) Accumulation of 3-ketosteroids
induced by itraconazole in azole resistant clinical Candida albicans iso-
lates. Antimicrobial Agents and Chemotherapy,43, 2663–2670.
Marichal, P., Gorrens, J. & Vanden Bossche, H. (1985) The action of
itraconazole and ketoconazole on growth and sterol synthesis in
Aspergillus fumigatus and Aspergillus niger.Sabouraudia: Journal of Med-
ical and Veterinary Mycology,23, 13–21.
Marichal, P., Koymans, L., Willemsens, S., Bellens, D., Verhasselt, P.,
Luyten, W., Borgers, M., Ramaekers, F.C.S., Odds, F.C. & Vanden
Bossche, H. (1999b) Contribution of mutations in the cytochrome
P450 14a-demethylase (Erg11p, Cyp51p) to azole resistance in
Candida albicans.Microbiology,145, 2701–2713.
Marichal, P., Vanden Bossche, H., Odds, F.C., Nobels, G., Warnock, D.W.,
Timmermans, V., Van Broekhoven, C., Fay, S. & Mose-Larsen, P. (1997)
Molecular biological characterization of an azole-resistant Candida
glabrata isolate. Antimicrobial Agents and Chemotherapy,41, 2229–2237.
Marr, K., Lyons, C.N., Rustad, T., Bowden, R.A. & White, T. (1998)
Rapid, transient fluconazole resistance in Candida albicans is associated
with increased mRNA levels of CDR. Antimicrobial Agents and
Chemotherapy,42, 2584–2589.
McGinnis, M.R., Nordoff, N., Li, R.-K., Pasarell, L. & Warnock, D.W.
(2001) Sporothrix schenckii sensitivity to voriconazole, itraconazole and
amphotericin B. Medical Mycology,39, 369–371.
Mehta, R.T., Hopfer, R.L., McQueen, T., Juliano, R.L. & Lopez-Berestein,
G. (1987) Toxicity and therapeutic effects in mice of liposome-
encapsulated nystatin for systemic fungal infections. Antimicrobial
Agents and Chemotherapy,31, 1901–1903.
Meletiadis, J., Meis, J.F.G.M., Mouton, J.W., Donnelly, J.P. & Verweij, P.E.
(2000) Comparison of NCCLS and 3-(4,5-dimethyl-2-thiazyl)-2,5-
diphenyl-2H-tetrazolium bromide (MTT) methods of in vitro suscept-
ibility testing of filamentous fungi and development of a new simplified
method. Journal of Clinical Microbiologyy,38, 2949–2954.
Mercer, E.I. (1991) Sterol biosynthesis inhibitors: their current status and
modes of action. Lipids,26, 584–597.
Mieth, H. (1993) In vitro and in vivo activities of terbinafine. In Cuta-
neous Antifungal Agents. Eds Rippon, J.W. & Fromtling, R.A. pp. 137–
149. Marcel Dekker, Inc., New York.
Miyazaki, H., Miyazaki, Y., Geber, A., Parkinson, T., Hitchcock, C.,
Falconer, D.J., Ward, J., Marsden, K. & Bennett, J.E. (1998) Flucona-
zole resistance associated with drug efflux and increased transcription
of a drug transporter gene, PDH1 in Candida glabrata.Antimicrobial
Agents and Chemotherapy,42, 1695–1701.
Moore, C.B., Walls, C.M. & Denning, D.W. (2000) In vitro activity of the
new triazole BMS-207147 against Aspergillus species in comparison
with itraconazole and amphotericin B. Antimicrobial Agents and
Chemotherapy,44, 441–443.
Moore, C.B., Walls, C.M. & Denning, D.W. (2001) In vitro activities of
terbinafine against Aspergillus species in comparison with those of
itraconazole and amphotericin B. Antimicrobial Agents and Chemother-
apy,45, 1882–1885.
Morace, G., Sanguinetti, M., Posteraro, B. & Pagano, L. (1995) In vitro
susceptibility of Candida species isolated from patients with haemato-
logical malignancies. Mycoses,38, 59–67.
Moran, G.P., Sanglard, D., Donelly, S.M., Shanley, D.B., Sullivan, D.J. &
Coleman, D.C. (1998) Identification and expression of multidrug
transporters responsible for fluconazole resistance in Candida
dubliniensis.Antimicrobial Agents and Chemotherapy,42, 1819–1830.
Mosquera, J. & Denning, D.W. (2002) Azole cross-resistance in Asper-
gillus fumigatus.Antimicrobial Agents and Chemotherapy,46, 556–557.
Muto, N., Hirai, H., Tanaka, T., Itoh, N. & Tanaka, K. (1997) Induction
and inhibition of cytochrome P450 isoforms by imazalil, a food con-
taminant, in mouse small intestine and liver. Xenobiotica,27, 1215–
1223.
National Committee for Clinical Laboratory Standards (1992) Reference
Method for Broth Dilution Antifungal Susceptibility Testing for Yeast.
Proposed standard. Document M27-P. National Committee for
Laboratory Standards, Wayne, PA.
National Committee for Clinical Laboratory Standards (1998) Reference
Method for Broth Dilution Antifungal Susceptibility Testing of Conidium –
Forming Filamentous Fungi. Proposed standard M38-P. National
Committee for Clinical Laboratory Standards, Wayne, PA.
Niewerth, M., Splanemann, C., Korting, H.C., Ring, J. & Abeck, D. (1998)
Antimicrobial susceptibility testing of dermatophytes-comparison of
the agar macrodilution and broth microdilution tests. Chemotherapy,
44, 31–35.
Nogales, E. (1999) A structural view of microtubule dynamics. Cellular
and Molecular Life Sciences,56, 133–142.
Nozawa, Y. & Morita, T. (1992) Biochemical aspects of squalene epoxidase
inhibition by a thiocarbamate derivative, naphthiomate T. In Recent
Antifungal agents 25
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
Progress in Antifungal Chemotherapy, Eds Yamaguchi, H., Kobayashi, G.S
& Takahashi, H. pp. 53–64. Marcel Dekker, Inc., New York.
Oakley, B.R. & Morris, N.R. (1980) Nuclear movement is b-tubulin
dependent in Aspergillus nidulans.Cell,19, 255–262.
Odds, F.C. (1998) Should resistance to azole antifungals in vitro be
interpreted as predicting clinical non-response? Drug Resistance
Updates,1, 11–15.
Odds, F.C. & Vanden Bossche, H. (2000) Antifungal activity of itraco-
nazole compared with hydroxy-itraconazole in vitro. Journal of Anti-
microbial Chemotherapy,45, 371–373.
Odds, F.C., Vranckx, L. & Woestenborghs, F. (1995) Antifungal sus-
ceptibility testing of yeasts: evaluation of technical variables for test
automation. Antimicrobial Agents and Chemotherapy,39, 2051–2060.
Odds. F.C. (1996) Antifungal therapy. In Principles and Practice of Clinical
Mycology. Eds Kibbler, C.C., Mackenzie, D.W.R & Odds, F.C. pp. 35–48.
John Wiley & Sons, Chichester.
Okey, A.B., Roberts, E.A., Harper, P.A. & Denison, M.S. (1986) Induction
of drug matabolising enzymes: mechanisms and consequences. Clinical
Biochemistry,19, 132–141.
Osterloh, I.H. & Pluck, N.D. (1993) Safety of fluconazole. In Cutaneous
Antifungal Agents. Eds Rippon, J.W. & Fromtling, R.A. pp. 215–229.
Marcel Dekker, Inc., New York.
Oxford, A.E., Raistrick, M. & Simonart, P. (1939) Studies in biochemistry
of micro-organisms. LX Griseofulvin C17H16O6Cl metabolic product
of Penicillium griseofulvin, Dierckx. Biochemical Journal,33, 240–248.
Pelletier, R., Peter, J., Antin, C., Gonzalez, C., Wood, L. & Walsh, T.J.
(2000) Emergence of resistance of Candida albicans to clotrimazole in
human immunodeficiency virus-infected children: in vitro and clinical
correlations. Journal of Clinical Microbiology,38, 1563–1568.
Perea, S., Fothergill, A.W., Sutton, D.A. & Rinaldi, M.G. (2001a) Com-
parison of in vitro activities of voriconazole and five established anti-
fungal agents against different species of dermatophytes using a broth
macrodilution method. Journal of Clinical Microbiology,39, 385–388.
Perea, S., Lo
´pez-Ribot, J.L., Kirckpatrick, W.R., McAtee, R.K., Santilla
´n,
R.A., Martinez, M., Calabrese, D., Sanglard, D. & Patterson, T.F.
(2001b) Prevalence of molecular mechanisms of resistance to azole
antifungal agents in Candida albicans strains displaying high-level
fluconazole resistance isolated from human immunodeficiency virus-
infected patients. Antimicrobial Agents and Chemotherapy,45, 2676–
2684.
Perparim, K., Nagai, H., Hashimoto, A., Goto, Y., Tashiro, T. & Nasu, M.
(1996) In vitro susceptibility of Trichosporon beigelii to antifungal
agents. Journal of Chemotherapy,8, 445–448.
Pershing, L.K., Corlett, J. & Jorgensen, C. (1994) In vivo pharmacoki-
netics and pharmacodynamics of topical ketoconazole and miconazole
in human stratum corneum. Antimicrobial Agents and Chemotherapy,
38, 90–95.
Pesti, M., Campbell, J.M. & Peberdy, J.F. (1981) Alteration of ergosterol
content and chitin synthase activity in Candida albicans.Current
Microbiology,5, 187–190.
Peyron, F., Favel, A., Michel-Nguyen, A., Gilly, M., Regli, P. & Bolmstrom,
A. (2001) Improved detection of amphotericin B-resistant isolates of
Candida lusitaniae by Etest. Journal of Clinical Microbiology,39, 339–342.
Pfaller, M.A., Messer, S.A., Boyken, L., Huynh, H., Hollis, R.J. & Diekema,
D.J. (2002) In vitro activities of 5-fluorocytosine against 8,803 clinical
isolates of Candida spp. global assessment of primary resistance using
national committee for clinical laboratory standards susceptibility
testing methods. Antimicrobial Agents and Chemotherapy,46, 3518–
3521.
Pfaller, M.A., Messer, S.A. & Hollis, R.J. (1994) Strain delineation and
antifungal susceptibilities of epidemiologically related and unrelated
isolates of Candida lusitaniae.Diagnostic Microbiology and Infectious
Diseases,20, 127–133.
Pfaller, M.A., Messer, S.A., Hollis, R.J. & Jones, R.N. (2001) In vitro activiti-
es of posaconazole (Sch 56592) compared with those of itraconazole and
fluconazole against 3,685 clinical isolates of Candida spp. & Cryptococcus
neoformans.Antimicrobial Agents and Chemotherapy,45, 2862–2864.
Pier, A.C., Caban˜ es, F.J., Chermette, R., Ferreiros, L., Guillot, J., Jensens,
H.E. & Santurio, J.M. (2000) Prominent animal mycoses from various
regions of the world. Medical Mycology,38 (Suppl. 1),47–58.
Plempel, M., Berg, D., Bu
¨chel, K.H. & Ritter, W. (1988) Experimental
antimycotic activity of azole derivatives – experience, knowledge and
questions. In Sterol Biosynthesis Inhibitors. Eds Berg, D. & Plempel, M.
pp. 349–396. Ellis Horwood Ltd, Chichester.
Plempel. M., Bartmann, K., Bu
¨chel, K.H. & Regel, E. (1969) Experi-
mentelle Befunde u
¨ber eines neues oral wirksames Antimykoticum mit
breiten Wirkungsspecktrum. Deutsche Medizinische Wochenschrift,94,
1356–1364.
Pohjola-Sintonen, S., Viitasalo, M., Toivonen, L. & Neuvonen, P. (1993)
Itraconazole prevents terbinafine metabolism and increases risk of
torsades de pointes ventricular tachycardia. European Journal of Clinical
Pharmacology,45, 191–193.
Polak, A. (1990) Mode of action studies. In Chemotherapy of Fungal Dis-
eases. Ed. Ryley, J.F. pp. 153–182. Springer Verlag, Berlin.
Polak, A. (1992) 5-Fluorocytosine and its combination with other anti-
fungal agents. In Recent Progress in Antifungal Chemotherapy. Eds
Yamaguchi, H. & Kobayashi, G.S. & Takahashi, H. pp. 77–85. Marcel
Dekker, Inc., New York.
Polak, A. & Dixon, D.M. (1987) Antifungal activity of amorolfine (Ro 14–
4767 002) in vitro and in vivo. In Recent Trends in the Discovery and
Evaluation of Antifungal Agents. Ed. Fromtling, R.A. pp. 555–573. J.R.
Prous Science Publishers S.A., Barcelona.
Polak, A. & Hartman, P.G. (1991) Antifungal chemotherapy. Are we
winning? Progress in Drug Research,37, 191–269.
Polak, A. & Scholer, H.J. (1980) Mode of action of 5-fluorocytosine. Revue
Institut Pasteur Lyon,13, 233–244.
Prasad, R., De Wergifosse, P., Goffeau, A. & Balzi, E. (1995) Molecular
cloning and characterization of a novel gene of Candida albicans,CDR1,
conferring multiple resistance to drugs and antifungals. Current
Genetics,27, 320–329.
Pye, G.W. & Marriott, M.S. (1982) Inhibition of sterol C-14-demethyla-
tion by imidazole containing antifungals. Sabouraudia,20, 325–329.
Raab, W.P. (1972) Natamycin (Pimaricin) its Properties and Possibilities in
Medicine. George Thieme Publishers, Stuttgart.
Raab, W.P. (1978) Mykosebehandlung mit imidazole derivaten. p. 41.
Springer-Verlag, Berlin.
Rao, U.S., Hennessey, J.P. & Scarborough, G.A. (1988) Protein chemistry
of the Neurospora crassa plasma membrane H
+
-ATPase. Analytical
Biochemistry,173, 251–264.
Readio, J.D. & Bittman, R. (1982) Equilibrium binding of amphotericin B
and its methyl ester and borate complex to sterols. Biochimica
Biophysica Acta,685, 219–224.
Redig, P.T. & Duke, G.E. (1985) Comparative pharmacokinetics of anti-
fungal drugs in domestic turkeys, red tailed hawks, broad-winged
hawks, and great-horned owls. Avian Diseases,29, 649–661.
Reuben, A., Anaissie, E., Nelson, P.E., Hashem, R., Legrand, C., Ho, D.H.
& Bodey, G.P. (1989) Antifungal susceptibility of 44 clinical isolates of
Fusarium species determined by using a broth microdilution method.
Antimicrobial Agents and Chemotherapy,33, 1647–1649.
Rex, J.H., Phaller, M.A., Galgiani, J.N. et al. (1997) Development of
interpretative breakpoints for antifungal susceptibility testing: con-
ceptual framework and analysis of in vitro- in vivo correlation data for
fluconazole, itraconazole, and Candida infections. Clinical Infectious
Diseases,24, 235–247.
Rex, J.H., Rinaldi, M.G. & Pfaller, M.A. (1995) Resistance of Candida
species to fluconazole. Antimicrobial Agents and Chemotherapy,39, 1–8.
26 H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
Rotowa, N.A., Shadomy, H.J. & Shadomy, S. (1990) In vitro activities of
polyene and imidazole antifungal agents against unusual opportunistic
fungal pathogens. Mycoses,33, 203–211.
Ruhnke, M., Eigler, A., Tennagen, I., Geiseler, B., Engelmann, E. &
Trautmann. M. (1994) Emergence of fluconazole-resistant strains in
patients with recurrent oropharyngeal candidosis and human
immunodeficiency virus infection. Journal of Clinical Microbiology,32,
2092–2098.
Ryder, N.S. (1988) Mechanism of action and biochemical selectivity of
allylamine antimycotic agents. Annals of the New York Academy of
Sciences,544, 208–220.
Ryder, N.S., Frank, I. & Dupont, M.C. (1986) Ergosterol biosynthesis
inhibition by the thiocarbamate antifungal agents, tolnaftate and tol-
ciclate. Antimicrobial Agents and Chemotherapy,29, 858–860.
Ryder, N.S. & Mieth, H. (1992) Allylamine antifungal drugs. Current
Topics in Medical Mycology,4, 158–188.
Ryder, N.S., Stuetz, A. & Nussbaumer, P. (1992) Squalene epoxidase
inhibitors. Structural determinants for activity and selectivity of ally-
lamines and related compounds. In Regulation of Isopentenoid Metabo-
lism. Eds Nes, W.D., Parish, E.J. & Trzaskos, J.M. pp. 192–204.
American Chemical Society, Washington, DC.
Ryley, J.F., Wilson, R.G. & Barrett-Bee, K.J. (1984) Azole resistance in
Candida albicans.Sabouraudia,22, 53–93.
Ryley, J.F., Wilson, R.G., Gravestock, M.B. & Poyser, J.P. (1981)
Experimental approaches to antifungal chemotherapy. Advances in
Pharmacology and Chemotherapy,18, 49–176.
Sanglard, D., Ischer, F. & Bille, J. (2000) Mitochondrial loss is associated
with high frequency azole resistance (HFAR) and with upregulation of
the ABC-transporter genes CgCDR1 and CgCDR2 in Candida Glabrata.
Abstract, 40th ICAAC, Toronto. Presentation Number 1954.
Sanglard, D., Ischer, F. & Bille, J. (2001) Role of ATP-binding-cassette
transporter genes in high-frequency acquisition of resistance to azole
antifungals in Candida glabrata.Antimicrobial Agents and Chemotherapy,
45, 1174–1183.
Sanglard, D., Ischer, F., Calabrese, D., de Micheli, M. & Bille, J. (1998a)
Multiple resistance mechanisms to azole antifungals in yeast clinical
isolates. Drug Resistance Updates,1, 255–265.
Sanglard, D., Ischer, F., Koymans, L. & Bille, J. (1998b) Amino acid
substitutions in the cytochrome P450 lanosterol 14a-demethylase
(CYP51A1) from azole-resistant Candida albicans clinical isolates con-
tribute to resistance to azole antifungal agents. Antimicrobial Agents and
Chemotherapy,42, 241–253.
Sanglard, D., Ischer, F., Calabrese, D., Majcherczyk, P.A. & Bille, J. (1999)
The ATP binding cassette transporter gene CgCDR1 from Candida
glabrata is involved in the resistance of clinical isolates to azole anti-
fungal agents. Antimicrobial Agents and Chemotherapy,43, 2753–2765.
Sanglard, D., Ischer, F., Monod, M. & Bille, J. (1997) Cloning of Candida
albicans genes conferring resistance to azole antifungal agents: char-
acterisation of CDR2, a new multidrug ABC transporter gene. Micro-
biology,143, 405–416.
Sanglard, D., Isher, F., Monod, M. & Bille, J. (1996) Susceptibilities of
Candida albicans multidrug transporter mutants to various antifungal
agents and other metabolic inhibitors. Antimicrobial Agents and
Chemotherapy,40, 2300–2305.
Sanglard, D., Kuchler, K., Ischer, F., Pagani, L., Monod, M. & Bille, J.
(1995) Mechanisms of resistance to azole antifungal agents in Candida
albicans isolates from AIDS patients involve specific multidrug trans-
porters. Antimicrobial Agents and Chemotherapy,39, 2378–2386.
Sanglard, D. & Odds, F.C. (2002) Resistance of Candida species to anti-
fungal agents: molecular mechanisms and clinical consequences.
Lancet Infectious Diseases,2, 73–85.
Sarosi, G.A. (1990) Amphotericin B still the ‘gold standard’ for anti-
fungal therapy. Postgraduate Medicine,88, 151–166.
Schmidt, A. (1997) In vitro activity of climbazole, clotrimazole and silver-
sulphadiazine against isolates of Malassezia pachydermatis.Zentralblatt
fu¨r Veterinarmedizin,44, 193–197.
Shadomy, S. & Espinel-Ingroff, A. (1974) Susceptibility testing of anti-
fungal agents. In Manual of Clinical Microbiology, 2nd edn. Eds
Lennette, E.H., Spaulding, E.H. & Truant, J.P. pp. 569–574. American
Society for Microbiology, Washington, DC.
Shadomy, S., White, S.C. & Dismukes, W.E. (1985) Treatment of systemic
mycoses with ketoconazole: in vitro susceptibilities of clinical isolates of
systemic and pathogenic fungi to ketoconazole. Journal of Infectious
Diseases,152, 1249–1256.
Siegel, M.R. & Ragsdale, N.N. (1978) Antifungal mode of action of
imazalil. Pesticide Biochemistry and Physiology,9, 48–56.
Smith, K.J., Warnock, D.W., Kennedy, C.T.C., Johnson, E.M., Hopwood, V.,
Van Cutsem, J. & Vanden Bossche, H. (1986) Azole resistance in
Candida albicans.Journal of Medical and Veterinary Mycology,24, 131–
142.
Soteropoulos, P., Vaz, T., Santangelo, R., Paderu, P., Huang, D.Y.,
Tamas. M.J. & Perlin, D.S. (2000) Molecular characterization of the
plasma membrane H
+
-ATPase, an antifungal target in Cryptococcus
neoformans.Antimicrobial Agents and Chemotherapy,44, 2349–2355.
Spitzer, S.G. & Spitzer, E.D. (1997) Identification of a Cryptococcus neo-
formans sequence belonging to the major facilitator superfamily (MFS).
Program and Abstracts of the 97th General Meeting of ASM. Abstract
F26.
Stevens, D.A. (1999) Itraconazole in cyclodextrin solution. Pharma-
cotherapy,19, 603–611.
Stevens, D.A. & Holmberg, K. (1999) Resistance to antifungal drugs:
current status and clinical implications. Current Opinion in Anti-
Infective Investigational Drugs,1, 306–317.
Stevens, D.A. & Stevens, J.A. (1996) Cross-resistance phenotypes of flu-
conazole-resistant Candida species: results with 655 clinical isolates
with different methods. Diagnostic Microbiology and Infectious Diseases,
26, 145–148.
Stu
¨tz, A. (1988) Synthesis and structure-activity correlations within
allylamine antimycotics. Annals of the New York Academy of Sciences,
544, 46–62.
Suzuki, S., Kurata, N., Nishimura, Y., Yasuhara, H. & Satoh, T. (2000)
Effects of imidazole antimycotics on the liver microsomal cytochrome
P450 isoforms in rats: comparison of in vitro and ex vivo studies.
European Journal of Drug Metabolism and Pharmacokinetics,25, 121–
126.
Thienpont, D., Van Cutsem, J., Van Cauteren, H. & Marsboom, R. (1981)
The biological and toxicological properties of imazalil. Arzeimittel-
Forschung,31, 309–315.
Thienpont, D., Van Cutsem, J., Van Gerven, F., Heeres, J. & Janssen, P.A.
(1979) Ketoconazole – a new broad spectrum orally active anti-
mycotic. Experientia,35, 606–607.
Thienpont, D., Van Cutsem, J., Van Nueten, J.M., Niemegeers, C.J.E. &
Marsboom, R. (1975) Biological and toxicological properties of eco-
nazole, a broad spectrum antimycotic. Arzeimittel-Forschung,25, 224–
230.
Thomas, P.A. & Rajasekaran, J. (1988) Treatment of Aspergillus keratitis
with imidazole and related compounds. In Aspergillus and Aspergillosis.
Eds Vanden Bossche, H., Mackenzie, D.W.R. & Cauwenbergh, G.
pp. 267–279. Plenum Press, New York.
Thornewell, S.J., Peery, T.J. & Skatrud, P.L. (1997) Cloning and
characterization of CneMDR1:aCryptococcus neoformans gene
encoding a protein related to multidrug resistance proteins. Gene,
201, 21–29.
Tobin, M.B., Peery, R.B. & Skatrud, P.L. (1997) Genes encoding multiple
drug resistance like proteins in Aspergillus fumigatus and Aspergillus
flavus.Gene,200, 11–23.
Antifungal agents 27
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
Troke, P. (1993) In vitro and experimental in vivo activities of flucon-
azole against some fungi causing cutaneous mycoses. In Cutaneous
Antifungal Agents. Eds Rippon, J.W. & Fromtling, R.A. pp. 199–214.
Marcel Dekker, Inc., New York.
Uchida, Y., Onodera, S., Nakade, T. & Otomo, K. (1994) Sterol
composition in polyene antibiotic-sensitive and resistant strains of
Malassezia pachydermatis.Veterinary Research Communications,18,
183–187.
Vaden, S.L., Heit, M.C., Hawkins, E.C., Manough, C. & Riviere, J.E. (1997)
Fluconazole in cats: pharmacokinetics following intravenous and oral
administration and penetration into cerebrospinal fluid, aqeuous
humour and pulmonary epithelial lining fluid. Journal of Veterinary
Pharmacology and Therapeutics,20, 181–186.
Van Cutsem, J. & Janssen, P.A.J. (1988) In vitro and in vivo models to
study the activity of antifungals against Aspergillus.InAspergillus and
Aspergillosis. Eds Vanden Bossche, H., Mackenzie, D.W.R. &
Cauwenbergh, G. pp. 215–227. Plenum Press, New York.
Van Cutsem, J. & Thienpont, D. (1972) Miconazole, a broad-spectrum
antimycotic agent with antibacterial activity. Chemotherapy,17, 392–
404.
Van Cutsem, J., Van Gerven, F., Geerts, H. & Rochette, F. (1985)
Treatment with enilconazole spray of dermatophytosis in rabbit farms.
Mycosen,28, 400–407.
Van Cutsem, J., Van Gerven, F. & Janssen, P.A.J. (1986) The in vitro
activity of azoles. In In Vitro and in Vivo Evaluation of Antifungal Agents.
Eds Iwata, K. & Vanden Bossche, H. pp. 51–64. Elsevier Science
Publishers B.V., Amsterdam.
Van Gerven, F. & Odds, F.C. (1995) The anti- Malassezia furfur activity in
vitro and its experimental dermatitis of six imidazole antifungal agents:
bifonazole, clotrimazole, flutrimazole, ketoconazole, miconazole and
sertaconazole. Mycoses,38, 389–393.
Van Gestel, J. (1986) The vapour phase activity of antifungal com-
pounds: a neglected or a negligible phenomenon? In In Vitro and
in Vivo Evaluation of Antifungal Agents. Eds Iwata, K. & Vanden Bossche,
H. pp. 207–218. Elsevier Science Publishers B.V., Amsterdam.
Van Gestel, J., Van Cutsem, J. & Thienpont, D. (1981) Vapour phase
activity of imazalil. Chemotherapy,27, 270–276.
Vanden Bossche, H. (1985) Biochemical targets for antifungal azole
derivatives: hypothesis on the mode of action. Current Topics in Medical
Mycology,1, 313–351.
Vanden Bossche, H. (1988) Mode of action of pyridine, pyrimidine and
azole antifungals. In Sterol Biosynthesis Inhibitors. Eds Berg, D. &
Plempel, M. pp. 79–119. Ellis Horwood Ltd, Chichester.
Vanden Bossche, H. (1990) Importance and role of sterols in fungal
membranes. In Biochemistry of Cell Walls and Membranes in Fungi. Eds
Kuhn, P.J., Trinci, A.P.J., Jung, M.J., Goosey, M.W. & Copping, L.G.
pp. 135–157. Springer-Verlag, Berlin.
Vanden Bossche, H. (1991) Anti-candida drugs – mechanisms of action.
In Candida and Candidamycoses. Eds Tu
¨mbay, E., Seelinger, H.P.R. &
O
¨zdem, A. pp. 83–95. Plenum Press, New York.
Vanden Bossche, H. (1992) Inhibitors of P450-dependent steroid bio-
synthesis: from research to medical treatment. Journal of Steroid
Biochemistry and Molecular Biology,43, 1003–1021.
Vanden Bossche, H. (1995) Chemotherapy of human fungal infections.
In Modern Selective Fungicides. Ed. Lyr, H. pp. 432–484. Gustav Fischer
Verlag, Jena.
Vanden Bossche, H. (1997) Mechanisms of antifungal resistance. Revista
Iberoamericano Micologı
´a,14, 44–49.
Vanden Bossche, H., Bellens, D., Cools, W. et al. (1986) Cytochrome
P450: target for itraconazole. Drug Development Research,8, 287–298.
Vanden Bossche, H., Dromer, F., Improvisi, I., Lozano-Chiu, M., Rex, J.H.
& Sanglard, D. (1998) Antifungal drug resistance in pathogenic fungi.
Medical Mycology,36 (Suppl. 1),119–128.
Vanden Bossche, H., Heeres, J., Backx, L.J.J., Marichal, P. & Willemsens, G.
(1993a) Discovery, chemistry, mode of action, and selectivity of itraco-
nazole. In Cutaneous Antifungal Agents. Eds Rippon, J.W. & Fromtling,
R.A. pp. 263–283. Marcel Dekker, Inc., New York.
Vanden Bossche, H. & Koymans, L. (1998) Cytochromes P450 in fungi.
Mycoses,41 (Suppl. 1),32–38.
Vanden Bossche, H., Koymans, L. & Moereels, H. (1995) P450 inhibitors
of use in medical treatment: focus on mechanisms of action. Pharma-
cology and Therapeutics,67, 79–100.
Vanden Bossche, H., Lauwers, W., Willemsens, G., Marichal, P., Corn-
elissen, F. & Cools, W. (1984) Molecular basis for the antimycotic and
antibacterial activity of N-substituted imidazoles and triazoles: the
inhibition of isoprenoid biosynthesis. Pesticide Sciences,15, 188–198.
Vanden Bossche, H. & Marichal, P. (1991) Mode of action of anti-Candida
drugs: focus on terconazole and other ergosterol biosynthesis
inhibitors. American Journal of Obstetrics and Gynecology,165, 1193–
1199.
Vanden Bossche, H., Marichal, P., Gorrens, J., Bellens, D., Moereels, H. &
Janssen, P.A.J. (1990) Mutation in cytochrome P450-dependent 14a-
demethylase results in decreased affinity for azole antifungals. Bio-
chemical Society Transactions,18, 56–59.
Vanden Bossche, H., Marichal, P., Gorrens, J., Bellens, D., Verhoeven, H.,
Coene, M.-C., Lauwers, W. & Janssen, P.A.J. (1987b) Interaction of
azole derivatives with cytochrome P450 isozymes in yeast, fungi,
plants and mammalian cells. Pesticide Sciences,21, 289–306.
Vanden Bossche, H., Marichal, P., Gorrens, J., Coene, M.-C., Willemsens,
G., Bellens, D., Roels, I., Moereels, H. & Janssen. P.A.J. (1989) Bio-
chemical approaches to selective antifungal activity. Focus on azole
antifungals. Mycoses,32 (Suppl. 1),35–52.
Vanden Bossche, H., Marichal, P., Gorrens, J., Geerts, H. & Janssen, P.A.J.
(1988) Mode of action studies. Basis for the search of new antifungal
drugs. Annals of the New York Academy of Sciences,544, 191–207.
Vanden Bossche, H., Marichal, P., Le Jeune, L., Coene, M.-C., Gorrens, J.
& Cools, W. (1993b) Effects of itraconazole on cytochrome P450-
dependent sterol 14a-demethylation and reduction of 3-ketosteroids in
Cryptococcus neoformans.Antimicrobial Agents and Chemotherapy,37,
2101–2105.
Vanden Bossche, H., Marichal, P. & Odds, F.C. (1994a) Molecular
mechanisms of drug resistance in fungi. Trends in Microbiology,2, 393–
400.
Vanden Bossche, H., Marichal, P., Odds, F.C., Le Jeune, L. & Coene, M.-C.
(1992) Charaterization of an azole-resistant Candida glabrata isolate.
Antimicrobial Agents and Chemotherapy,36, 2602–2610.
Vanden Bossche, H., Ruysschaert, J.M., Defrise-quertain, F., Willemsens,
G., Cornelissen, F., Marichal, P., Cools, W. & Van Cutsem, J. (1982)
The interaction of miconazole and ketoconazole with lipids. Biochemical
Pharmacology,31, 2609–2617.
Vanden Bossche, H., Warnock, D.W., Dupont, B., Kerridge, D., Sen
Gupta, S., Improvisi, L., Marichal, P., Odds, F.C., Provost, F. & Ronin, O.
(1994b) Mechanisms and clinical impact of antifungal drug
resistance. Journal of Medical and Veterinary Mycology,32 (Suppl. 1),
189–202.
Vanden Bossche, H., Willemsens, G., Cools, W., Cornelissen, F., Lauwers, F.
& Van Cutsem, J. (1980) In vitro and in vivo effects of ketoconazole
on sterol synthesis. Antimicrobial Agents and Chemotherapy,17, 922–
928.
Vanden Bossche, H., Willemsens, G., Cools, W., Lauwers, W.F.J. & Le
Jeune, L. (1978) Biochemical effects of miconazole on fungi II:
Inhibition of ergosterol biosynthesis in Candida albicans.Chemical
Biological Interactions,21, 59–78.
Vanden Bossche, H., Willemsens, G. & Marichal, P. (1987a) Anti-candida
drugs – the biochemical basis for their activity. CRC Critical Reviews in
Microbiology,15, 57–72.
28 H. Vanden Bossche et al.
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
Venkatakrishnan, K., von Moltke, L.L. & Greenblatt, D.J. (2000) Effects of
the antifungal agents on oxidative drug metabolism: clinical relevance.
Clinical Pharmacokinetics,38, 111–180.
Venkateswarlu, K., Denning, D.W., Manning, N.J. & Kelly, S.L. (1995)
Resistance to fluconazole in Candida albicans from AIDS patients cor-
related with reduced intracellular accumulation of drug. FEMS
Microbiology Letters,131, 337–341.
Venkateswarlu, K., Taylor, M., Manning, N.J., Rinaldi, M.G. & Kelly, S.L.
(1997) Fluconazole tolerance in clinical isolates of Cryptococcus neo-
formans.Antimicrobial Agents and Chemotherapy,41, 748–751.
Vermes, A., Guchelaar, H.J. & Dankert, J. (2000) Flucytosine: a review of
its pharmacology, clinical indications, pharmacokinetics, toxicity and
drug interactions. Journal of Antimicrobial Chemotherapy,46, 171–179.
Walsh, T.J., Melcher, G.P., Rinaldi, M.G., Lecciones, J., McGough, D.A.,
Kelly, P., Lee, J., Callender, D., Rubin, M. & Pizzo, P.A. (1990) Tri-
chosporon beigelii, an emerging pathogen resistant to amphotericin B.
Journal of Clinical Microbiology,28, 1616–1622.
Watson, P.F., Rose, M.E., Ellis, S.W., England, H. & Kelly, S.L. (1989)
Defective sterol C5-6 desaturation and azole resistance: a new
hypothesis for the mode of action of azole antifungals. Biochemical
Biophysical Research Communications,164, 1170–1175.
Whelan, W.L. (1987) The genetic basis of resistance to 5-fluorocytosine
in Candida species and Cryptococcus neoformans.CRC Critical Reviews in
Microbiology,15, 45–56.
White, T.C. (1997) The presence of an R467K amino acid substitution
and loss of allelic variation correlate with an azole -resistant lanosterol
14a-demethylase in Candida albicans.Antimicrobial Agents and Chemo-
therapy,41, 1488–1494.
White, T.C., Marr, K.A. & Bowden, R.A. (1998) Clinical, cellular, an
molecular factors that contribute to antifungal drug resistance. Clinical
Microbiology Reviews,11, 382–402.
Williams, D.J., Marten, R.M. & Sarkany, I. (1958) Oral treatment of
ringworm with griseofulvin. Lancet,2, 1212.
Wilm, K. & Stahl, A.J.C. (1983) Effects of econazole nitrate on yeast cells
and mitochondria. Biochemical Pharmacology,32, 1825–1830.
Ziogas, B.N., Sisler, H.D. & Lucas, W.R. (1983) Sterol content and other
characteristics of pimaricin-resistant mutants of Aspergillus nidulans.
Pesticide Biochemistry and Physiology,20, 320–329.
Antifungal agents 29
2003 Blackwell Publishing Ltd, J. vet. Pharmacol. Therap. 26, 5–29
... Another series of BR biosynthesis inhibitors possessing a dioxolane ring, in addition to the triazole ring, were designed using ketoconazole ( Figure 8A) as a lead compound. Ketoconazole is used in medicine [225] and veterinary medicine [226,227] for treatment of fungal infections, mainly by application on the skin [228][229][230] but in some cases also administrated orally [231]. Ketoconazole inhibits fungal ergosterol biosynthesis by binding to CYP51 [232], a cytochrome P450 14αdemethylase catalyzing oxidative demethylation of lanosterol [233]. ...
... Another series of BR biosynthesis inhibitors possessing a dioxolane ring, in addition to the triazole ring, were designed using ketoconazole ( Figure 8A) as a lead compound. Ketoconazole is used in medicine [225] and veterinary medicine [226,227] for treatment of fungal infections, mainly by application on the skin [228][229][230] but in some cases also administrated orally [231]. Ketoconazole inhibits fungal ergosterol biosynthesis by binding to CYP51 [232], a cytochrome P450 14α-demethylase catalyzing oxidative demethylation of lanosterol [233]. ...
Article
Full-text available
Chemical inhibitors are invaluable tools for investigating protein function in reverse genetic approaches. Their application bears many advantages over mutant generation and characterization. Inhibitors can overcome functional redundancy, their application is not limited to species for which tools of molecular genetics are available and they can be applied to specific tissues or developmental stages, making them highly convenient for addressing biological questions. The use of inhibitors has helped to elucidate hormone biosynthesis and signaling pathways and here we review compounds that were developed for the plant hormones brassinosteroids (BRs). BRs are steroids that have strong growth-promoting capacities, are crucial for all stages of plant development and participate in adaptive growth processes and stress response reactions. In the last two decades, impressive progress has been made in BR inhibitor development and application, which has been instrumental for studying BR modes of activity and identifying and characterizing key players. Both, inhibitors that target biosynthesis, such as brassinazole, and inhibitors that target signaling, such as bikinin, exist and in a comprehensive overview we summarize knowledge and methodology that enabled their design and key findings of their use. In addition, the potential of BR inhibitors for commercial application in plant production is discussed.
... Diagnosis can be done through culture and PCR testing. 42 Treatment of C. albicans infections in calves typically involves the use of antifungal drugs, such as fluconazole and nystatin. 42 Prevention of the infections in dairy calves involves good management practices, providing adequate nutrition and hydration, and minimizing stress. ...
... 42 Treatment of C. albicans infections in calves typically involves the use of antifungal drugs, such as fluconazole and nystatin. 42 Prevention of the infections in dairy calves involves good management practices, providing adequate nutrition and hydration, and minimizing stress. 43,44 Cryptococcus neoformans (C. ...
Article
Full-text available
Eukaryotic infections are common among dairy calves and can have significant impacts on their health and growth rates. Fungal infections caused by Aspergillus fumigatus, Trichophyton verrucosum, and Candida albicans can cause respiratory diseases, dermatophytosis, and diarrhea, respectively. Protozoan parasites, including Cryptosporidium parvum, Giardia duodenalis, and Eimeria spp., are also common in dairy calves. C. parvum is highly contagious and can cause severe diarrhea and dehydration, while Giardia duodenalis can lead to poor growth and is transmissible to humans through contaminated food or water. Eimeria spp. can cause coccidiosis and lead to reduced growth rates, poor feed conversion, and death. The common helminthic infections in dairy calves include Ostertagia ostertagi, Cooperia spp., Fasciola hepatica, and Strongyloides papillosus. These parasitic infections significantly impact calf health, growth, and dairy industry productivity. Diagnosis of these infections can be made through fecal samples using microscopy or molecular methods. However, diagnosis of the infections can be challenging and requires a combination of clinical signs and laboratory tests such as culture and PCR. Preventing and controlling eukaryotic infections in dairy calves requires several measures. Good hygiene and sanitation practices, proper management strategies, and timely treatment of affected animals are important. It is also necessary to avoid overcrowding and consider vaccination against ringworm. Further research is needed to better understand the epidemiology and characterization of eukaryotic infections in dairy calves, which will help in the development of more effective prevention and control strategies. In general, good hygiene practices, appropriate management strategies, and timely treatment of affected animals are crucial in preventing and controlling the infections, ensuring the health and well-being of dairy calves.
... Fungal sphingolipids are known to participate in several crucial roles in fungal cell physiology, such as signal transduction, endocytosis of membrane protein, and stress responses, in addition to being structurally different from the sphingolipids found in mammalian cells or even completely absent altogether in mammalian systems [3,64]. Therefore, they are emerging as potential targets for future antifungal drugs. ...
Article
Full-text available
Multidrug-resistant fungal pathogens and antifungal drug toxicity have challenged our current ability to fight fungal infections. Therefore, there is a strong global demand for novel antifungal molecules with the distinct mode of action and specificity to service the medical and agricultural sectors. Polyenes are a class of antifungal drugs with the broadest spectrum of activity among the current antifungal drugs. Epipyrone A, a water-soluble antifungal molecule with a unique, linear polyene structure, was isolated from the fungus Epiccocum nigrum. Since small changes in a compound structure can significantly alter its cell target and mode of action, we present here a study on the antifungal mode of action of the disalt of epipyrone A (DEA) using chemical-genetic profiling, fluorescence microscopy, and metabolomics. Our results suggest the disruption of sphingolipid/fatty acid biosynthesis to be the primary mode of action of DEA, followed by the intracellular accumulation of toxic phenolic compounds, in particular p-toluic acid (4-methylbenzoic acid). Although membrane ergosterol is known to be the main cell target for polyene antifungal drugs, we found little evidence to support that is the case for DEA. Sphingolipids, on the other hand, are known for their important roles in fungal cell physiology, and their biosynthesis has been recognized as a potential fungal-specific cell target for the development of new antifungal drugs.
... The continual research of the prospective uses of antifungal drugs, especially as antifungal agents, is a highly active area in drug development. Several imidazole derivatives, such as ketoconazole, miconazole, clotrimazole, and tioconazole, have been commercially available as antifungal drugs [7][8][9]. Imidazole compounds are currently being studied as potential corrosion inhibitors because of their capacity to effectively manage corrosion in extremely corrosive settings at low concentrations [10,11]. Drugs containing imidazole see a significant enhancement in their pharmacokinetic properties due to the presence of the polar imidazole ring. ...
Article
This paper describes a comparative study concerning the molecule 1-methyl-2-[(E)-2-(4-methylphenyl) ethenyl]-4-nitro-1H-imidazole (EMENI). The study involved experimental analysis using high-resolution X-ray diffraction and DFT calculations using the B3LYP functional, along with the 6–311G(d,p) basis set. An accurate depiction of the molecular electron charge density distribution in the compound was achieved using Hansen and Coppens' multipolar model. The study also involved a topological analysis of the total electron density and the location of critical points of the bond, which helps in evaluating the molecular reactivity. Additionally, Hirshfeld surface (HS) and reduced density gradient (RDG) analyses were conducted to explore the specific intermolecular interactions responsible for the stiffness of the molecular arrangement in the compound's packaging. The analyses revealed that H...H interactions significantly influenced the intermolecular contacts, comprising 34.3% of the total interactions as indicated by the 2D fingerprint plots. Furthermore, the study included an exploration of charge transfer and establishment of the electronic characteristics of EMENI through Frontier molecular orbitals (FMO) analysis and HOMO-LUMO energies. Finally, the current study involved an in-silico investigation using molecular docking to evaluate the biological activity of the compound studied. This investigation showed that the title compound exhibits exceptional antiprotozoal activity against Trichomonas
... It can infect animals as well as humans and cause serious diseases such as candidiasis, which cause serious infection according to the infection site, cutaneous candidiasis, oral and gastrointestinal mucosal candidiasis, and vaginal canal [8]. Although there are several types of antifungal drugs such as clotrimazole, econazole, miconazole, terbinafine, fluconazole, ketoconazole, and amphotericin, were extreme irritants and might be lethal, many studies reported that azole antifungal resistance cases of C. albicans [56,57]. This study aims to find an alternative green approach to overcome candidal infection problems especially C. albicans resistance to current commercial drugs. ...
Article
Full-text available
Candida albicans is one of the most dangerous pathogenic fungi in the world, according to the classification of the World Health Organization, due to the continued development of its resistance to currently available anticandidal agents. To overcome this problem, the current work provided a simple, one-step, cost-effective, and safe technique for the biosynthesis of new functionalized anticandidal selenium nanoparticles (Se NPs) against C. albicans ATCC10231 using the cell-free supernatant of Limosilactobacillus fermentum (OR553490) strain. The bacterial strain was isolated from yogurt samples available in supermarkets, in Damietta, Egypt. The mixing ratio of 1:9 v/v% between cell-free bacterial metabolites and sodium selenite (5 mM) for 72 h at 37 °C were the optimum conditions for Se NPs biosynthesis. Ultraviolet–visible spectroscopy (UV–Vis), Fourier transform infrared spectroscopy (FT-IR), transmission electron microscopy (TEM), X-ray diffraction (XRD), Zeta analyses, and elemental analysis system (EDS) were used to evaluate the optimized Se NPs. The Se NPs absorption peak appeared at 254 nm. Physicochemical analysis of Se NPs revealed the crystalline-shaped and well-dispersed formation of NPs with an average particle size of 17–30 nm. Se NPs have − 11.8 mV, as seen by the zeta potential graph. FT-IR spectrum displayed bands of symmetric and asymmetric amines at 3279.36 cm ⁻¹ and 2928.38 cm ⁻¹ , aromatic and aliphatic (C–N) at 1393.32 cm ⁻¹ and 1237.11.37 cm ⁻¹ confirming the presence of proteins as stabilizing and capping agents. Se NPs acted as a superior inhibitor of C. albicans with an inhibition zone of 26 ± 0.03 mm and MIC value of 15 µg/mL compared to one of the traditional anticandidal agent, miconazole, which revealed 18 ± 0.14 mm and 75 µg/mL. The cytotoxicity test shows that Se NPs have a low toxic effect on the normal keratinocyte (IC 50 ≈ 41.5 μg/mL). The results indicate that this green synthesis of Se NPs may have a promising potential to provide a new strategy for drug therapy.
... The antifungal agents commonly used in dogs belong to various chemical classes with different mechanisms of action. Antifungal agents include azole derivatives, polyene macrolides, allylamines and benzimidazoles (Bossche et al., 2003;Peano et al., 2020). In veterinary medicine, itraconazole (ITZ), ketoconazole (KTZ) and miconazole (MCZ), which are azole antifungal agents, are most frequently used, along with terbinafine (TER) and nystatin if necessary (Guillot et al., 2003). ...
Article
Full-text available
Background Fungal and yeast infections, including those caused by Malassezia spp., are becoming increasingly difficult to treat, likely due to the occurrence of drug resistance. Objectives This study aimed to evaluate the antifungal effects of omeprazole (OME), a proton pump inhibitor, against antifungal‐resistant Malassezia pachydermatis and to investigate the potential synergistic effects between OME and other antifungal compounds. Methods In total, 15 samples of M. pachydermatis isolated from the skin of dogs were tested. The susceptibility of M. pachydermatis to itraconazole, ketoconazole, miconazole, terbinafine and OME was assessed using a modified broth microdilution (BM) method. The in vitro efficacy of OME alone and in combination with other antifungal compounds was evaluated for all isolates using the BM chequerboard method. The data obtained were analysed using the fractional inhibitory concentration index (FICI). Results The minimum inhibitory concentration (MIC) values of antifungal compounds and OME against quality control strain (M. pachydermatis CBS1879) were lower than the MIC90 values of same drugs against clinically collected strains. There was no significant difference in MIC values between drugs alone and combination. According to the analysis by the FICI method, no interaction was observed with OME and antifungal compounds. Conclusions Most M. pachydermatis strains were resistant to azole antifungal compounds. OME exerted antifungal effects against Malassezia spp. and even showed good effects on antifungal‐resistant strains. No synergistic effects were observed between the antifungal compounds and OME.
Article
Imidazole is a versatile heterocyclic moiety that has been found to be widely present in many natural and synthetic molecules. It is a ubiquitous moiety that has been widely explored by the scientific community owing to its broad spectrum of biological activities. To pave the way for future research, a comprehensive study on the imidazole moiety was carried out. This review paper will assist researchers and academics in developing more therapeutically active imidazole-based compounds. Herein, we provide a thorough overview of the current research status of imidazole-based compounds with a wide variety of biological activities, including antiviral, anti-cancer, anti-microbial, analgesic and anti-inflammatory, antidiabetic, antiprotozoal, antipsychotic, anti-aging, antioxidant, and anthelmintic. Through this review article, our focus is to contribute towards the recent advancements in the synthetic methodologies and SAR study of imidazoles that contribute as a useful class of heterocyclic compounds involved in the development of therapeutically active molecules. We sincerely hope that this review will serve as a pertinent guide for imminent research in the development of imidazoles and their derivatives.
Chapter
The yeasts of the genus Malassezia are part of the normal skin microbiota of a wide range of warm-blooded animals, including humans. Currently, this genus includes 18 species that have been only isolated from different wild and domestic animals and humans. A main characteristic of these yeasts is their lipid dependency, which is caused by the lack of the gene encoding the fatty acid synthase. Thus, Malassezia yeasts have a predilection for sebaceous skin sites. Despite being part of the normal skin microbiota, the population of Malassezia can overgrow under certain circumstances and these yeasts can become opportunistic pathogens causing Malassezia-related diseases such as dermatitis, otitis, and other skin diseases. In dogs, Malassezia pachydermatis is considered a common etiological agent causing otitis and dermatitis. Although azole antifungals are usually effective in resolving these diseases, some cases of treatment failure have recently been reported.
Chapter
During recent years considerable advances have been made in the identification of potential targets for antifungal agents (for a review see Kerridge and Vanden Bossche, in press). A number of them are in the cell wall, others are in the plasma membrane, endoplasmic reticulum, nucleus, mitochondria, cytoskeleton or cytosol (Fig. 1).
Article
A companion article discussed the pharmacology and clinical uses of the more traditional antifungal therapies: polyenes, griseofulvin, and iodides. The availability of newer antifungal drugs, which are often more efficacious with fewer side effects, has led to many safe and effective applications in the management of small animal cutaneous fungal infections. This article describes the pharmacokinetics, modes of action, principal adverse effects, and clinical uses of antifungal agents of the azole (triazoles, imidazoles) and allylamine (terbinafine) classes for treating cutaneous fungal diseases in small animals. Clinical experience gained with the newer antifungals will aid practitioners in choosing appropriate drugs from an expanded armamentarium.
Chapter
In this study conducted over a period of thirty months, 228 cases of mycotic keratitis were treated with ketoconazole, itraconazole or amphotericin B. This included 69 cases of Aspergillus keratitis. In general, severity of the keratitis influenced response to therapy with all three compounds. A positive correlation was made between age and response of Aspergillus keratitis to itraconazole. Aspergillus keratitis was found to respond more frequently to itraconazole than to ketoconazole or amphotericin B. Itraconazole was significantly superior to ketoconazole and apparently superior to amphotericin B in the treatment of severe Aspergillus keratitis. Although oral azole therapy was not associated with major drawbacks, topical azole therapy was frequently hampered by ocular irritation and inefficacy in treating severe keratitis. Itraconazole may emerge as the drug of choice in treating Aspergillus keratitis.