The need for safe and effective antifungal agents increases in parallel with the expanding number of immunocompromised patients at risk for invasive fungal infections. The emergence of fungal pathogens resistant to current therapies further compounds the dearth of antifungal agents. Currently available antifungal compounds act on targets also found in mammalian cells (34), which may result in toxicity or an adverse drug interaction. It is therefore imperative to find antifungal compounds that are not toxic to mammalian cells. The past decade has witnessed a dramatic growth in knowledge of natural peptides. Peptides such as the cecropins were shown to be antimicrobial but not lethal for mammalian cells (21, 141, 162, 182). Most data on antimicrobial peptides concern bacteria. This minireview presents a review of the current literature on antifungal peptides, including their in vitro and in vivo activities, mechanisms of action, and structure-function relationships, when known.
CLASSIFICATION OF PEPTIDES
Antifungal peptides are classified by their mode of action. The first group acts by lysis, which occurs via several mechanisms (158). Lytic peptides may be amphipathic, that is, molecules with two faces, with one being positively charged and the other being neutral and hydrophobic. Some amphipathic peptides bind only to the membrane surface and can disrupt the membrane structure without traversing the membrane. Others traverse membranes and interact specifically with certain molecules. Finally, other amphipathic peptides aggregate in a selective manner, forming aqueous pores of variable sizes, allowing passage of ions or other solutes. The second peptide group interferes with cell wall synthesis or the biosynthesis of essential cellular components such as glucan or chitin (34). An excellent review of lipopeptide antifungal agents affecting cell wall synthesis has been published previously (9).
MAMMALIAN PEPTIDES
Defensins.
α-Defensins (“classic defensins”) and β-defensins (Table 1), which are present in many organisms, are predominantly β-sheet structures stabilized by three disulfide bonds that distinguish them from other antimicrobial peptides that form amphipathic helices (185). They are small, variably cationic proteins whose three-dimensional folds form highly amphipathic molecules (55). Defensins electrostatically bond to membranes, causing the formation of multimeric pores and the leakage of essential minerals and metabolites (102, 105, 133, 185). Defensin A caused membrane depolarization, decreased cytoplasmic ATP levels, and inhibited cellular respiration (31). The entrance of defensins into cells has caused DNA damage (58, 105).
TABLE 1.
Peptide | Source | No. of amino acids | Mode of action | Typical target organism | In vitro MIC (μg/ml) |
---|---|---|---|---|---|
Defensins | |||||
NP-1 | Rabbit granulocytes | 33 | Lysis | C. neoformans | 3.75–15.0a |
NP-2 | Rabbit granulocytes | 33 | Lysis | A. fumigatus | 25.0 |
NP-3A | Rabbit granulocytes | 34 | Lysis | A. fumigatus | 100.0 |
NP-3B | Rabbit granulocytes | 33 | Lysis | A. fumigatus | 100.0 |
NP-4 | Rabbit granulocytes | 33 | Lysis | A. fumigatus | >100.0 |
NP-5 | Rabbit granulocytes | 33 | Lysis | A. fumigatus | Inactive alone |
HNP-1 | Human neutrophils | 30 | Lysis | C. albicans | 50.0 |
HNP-2 | Human neutrophils | 29 | Lysis | C. albicans | 50.0 |
HNP-3 | Human neutrophils | 30 | Lysis | C. neoformans | 50.0 (LD50b) |
Gallinacin-1 | Chicken | 39 | Lysis | C. albicans | 25.0 |
Lactoferricin-B | Human, bovine | 18 | Lysis | C. albicans | 0.8 |
Protegrins 1 to 3 | Human, porcine | 16–18 | Lysis | C. albicans | 3.0–60.0 |
Tracheal antimicrobial peptide | Human, bovine | 38 | Lysis | C. albicans | 6.0–12.0 |
Tritrptcin | Human, porcine | 13 | Lysis | A. flavus | 250.0 |
MICs based on assays with multiple isolates.
LD50, lethal dose for 50% of the population.
Rabbit, guinea pig, rat, and human neutrophils contained defensins within azurophilic granules (42, 55, 155–157). Rabbit granulocytes contained six α-defensins structurally homologous to human defensins (106). Three such peptides, NP-1, NP-2, and NP-3a, were highly effective against Candida albicans (157). Although NP-5 lacked candidacidal properties alone, at submicromolar concentrations it potentiates the anti-Candida effects of other rabbit defensins (106). This effect of NP-5, however, was not observed with NP-3b or NP-4. NP-1 had MICs ranging from 3.75 to 15 μg/ml for encapsulated strains of Cryptococcus neoformans, while the MICs for acapsular strains were much lower (0.93 μg/ml) (3). NP-1 and other rabbit defensins were also lethal for Coccidiodes immitis, as well as hyphae and germinating conidia, but not dormant conidia, of Rhizopus oryzae and Aspergillus fumigatus (107, 153). As measured by the yellow tetrazolium salt assay, NP-1, NP-2, and NP-3 killed all A. fumigatus hyphae at 25, 25, and 100 μg/ml, respectively (107). At 100 μg/ml, NP-4 killed only 11% of the hyphae, while NP-5 had no effect. Resting conidia of A. fumigatus were resistant to 100 μg of these peptides per ml. Purified chitin and its fragments chitobiose and chitotrose bound to NP-1 and prevented the death of A. fumigatus, suggesting that the lethality of NP-1 was through binding to cell wall chitin (107).
Human α-defensins, HNP-1 to HNP-3, are constituents of the microbicidal granules of neutrophils (104). At 50 μg/ml, HNP-1 and HNP-2, but not HNP-3, were lethal for C. albicans (103). On a concentration basis, rabbit NP-1 was 10- to 20-fold more active than HNP-1 against C. albicans (103). HNP-1 to HNP-3 at 50 μg/ml inhibited C. neoformans growth, with a reduction of >103 CFU/ml compared to the growth of the control after 4 h (56).
Bovine tracheal antimicrobial peptide, a cysteine-rich β-defensin produced by respiratory epithelial cells, was active (41) against the yeast forms of several C. albicans strains. The synthetic form at 400 μg/ml was active against the hyphal forms of A. fumigatus and C. albicans (98). In contrast, magainin II, α-defensin, and amphotericin B had lower MICs for A. fumigatus (250, 200, and 0.8 μg/ml, respectively) (98).
Protegrins and gallinacins.
The protegrins, which are related to the β-defensins, are produced by porcine leukocytes. They are cationic, cysteine-rich molecules with two intermolecular, parallel, disulfide bridges which stabilize an amphipathic β-sheet structure comprising two antiparallel strands (7, 70, 89). Protegrins formed weakly selective ionic channels that anions and small cations permeated, indicating that the cysteine bridges are a prerequisite for membrane permeability alteration but not for antimicrobial activity (112). In contrast, others reported that these intramolecular disulfide bonds enhance the antimicrobial and lytic actions of protegrins (71). Zone inhibition studies showed that protegrins 1, 2, and 3 inhibited C. albicans growth at 60, 8, and 3 μg/ml, respectively (89). Chicken leukocytes produce the gallinacin peptide family (69). Gallinacins have three intramolecular cystine disulfide bonds, are relatively cationic, and are rich in lysine and arginine. Gallinacin-1 and -1α inhibited C. albicans in a radial diffusion assay (69). However, gallinacin-2 showed no activity at up to 400 μg/ml in this assay.
Tritrpticin and lactoferricin.
Precursors of many antimicrobial peptides of porcine, bovine, and rabbit origin share highly conserved regions with antifungal properties (108, 163, 189). Tritrpticin corresponds to 13 amino acids of the C-terminal portion of cathelin, a putative proteinase inhibitor from porcine blood leukocytes. In vitro, it was weakly inhibitory for Aspergillus flavus and C. albicans (97). Bovine lactoferrin, an iron-binding protein, had broad antimicrobial properties (25, 143). Lactoferricin, an enzymatic product of lactoferrin, possessed greater antimicrobial properties than lactoferrin and corresponds to the 18 amino acid residues near the N terminus of lactoferrin in a region distinct from its iron-binding sites (16, 176). Lactoferricin was active against C. albicans; however, its antimicrobial properties were diminished by Ca2+, Mg2+, and Fe2+ (186). The optimum pH for this peptide was 6.0, and it bound to outer bacterial membranes, causing disruption of normal permeability functions of the cytoplasmic membrane and ultrastructural damage (17, 186).
BPI protein domain III analogs.
The bactericidal and permeability-increasing (BPI) protein is a cationic protein stored principally in the azurophilic granules of neutrophils (43). Several potent antifungal peptides with activity against Candida spp., C. neoformans, and A. fumigatus were derived from BPI protein functional domain III (109). These constructs produced significant, dose-dependent reductions in the numbers of C. albicans CFU in the kidney and significant protection from mortality in murine candidiasis models (5). Three small synthetic peptides (XMP.284, XMP.366, and XMP.391) based on BPI protein domain III were found to be fungicidal for several Candida species, while subinhibitory concentrations of these peptides enhanced the anti-Candida activities of fluconazole (78). XMP.391 was effective against murine disseminated aspergillosis and enhanced the effectiveness of amphotericin B (4).
INSECT-DERIVED ANTIMICROBIAL PEPTIDES
Cecropins.
Cecropins (Table 2), which form α-helices in solution, are linear peptides in the hemolymph of the giant silk moth (Hyalopora cecropia) (21, 162). They are positively charged and form time-variant and voltage-dependent ion channels in planar lipid membranes (29). Cecropins were not lethal for mammalian cells at microbicidal levels and have been administered safely to animals (21, 65, 122, 141, 162, 182). At between 25 and 100 μg/ml it is fungicidal for pathogenic Aspergillus species (37, 38). Fusarium moniliforme and Fusarium oxysporum were especially sensitive to cecropin A, with total killing attained at 12.4 μg/ml (37).
TABLE 2.
Peptide | Source | No. of amino acids | Mode of action | Typical target organism | In vitro MIC (μg/ml) |
---|---|---|---|---|---|
Antifungal peptide | S. peregrina | 67 | Lysis | C. albicans | 25.0 |
Cecropins | |||||
A | H. cecropia | 37 | Lysis | F. oxysporum | 12.0 |
B | H. cecropia | 35 | Lysis | A. fumigatus | 9.5 |
Dermaseptins | |||||
b | P. sauvagii | 27 | Lysis | C. neoformans | 60.0 |
s | P. sauvagii | 34 | Lysis | C. neoformans | 5.0 |
Drosomycin | D. melanogaster | 44 | Lysis | F. oxysporum | 5.9–12.3a |
Magainin 2 | X. laevis | 23 | Lysis | C. albicans | 80 |
Thanatin | P. maculiventris | 21 | Unknown | A. fumigatus | 24–48a |
MICs based on assays with multiple isolates.
Drosomycin.
Drosophila melanogaster produces drosomycin, an insect defensin with significant homology with plant antifungal peptides isolated from seeds of members of the family Brassicaceae (47). It was similar in structure to the radish antifungal peptide, Rs-AFP1, and was particularly effective against F. oxysporum isolates (118).
Antifungal peptide, holotricin 3, and thanatin.
Insect peptides which are antifungal include antifungal peptide, holotricin 3, and thanatin. Antifungal protein, a histidine-rich peptide that causes cellular leakage, was purified from the third instar larval hemolymph of Sacrophaga peregrina, and in vitro, it was lethal for C. albicans (79). Holotricin 3, a glycine- and histidine-rich peptide purified from the larval hemolymph of Holotrichia diomphalia, inhibited C. albicans growth (101). Thanatin, produced by Podisus maculiventris, is nonhemolytic and is active against F. oxysporum and A. fumigatus (46).
AMPHIBIAN-DERIVED PEPTIDES
Magainins.
The African clawed frog (Xenopus laevis) produces the magainins, which are α-helical ionophores that dissipate ion gradients in cell membranes, causing lysis (184). Their helical, amphiphilic structure was responsible for affinity to membranes (28). An increase in the magainin concentration caused the artificial lipid bilayer thickness to decrease, suggesting adsorption within the head-group region of the lipid bilayer (111). Magainin 2 was nonhemolytic and inhibited C. albicans growth (190). This nonhemolytic property may result from a peptide-cholesterol interaction in mammalian membranes that inhibits the formation of peptide structures capable of lysis (179).
Dermaseptin.
The South American arboreal frog (Phyllomedusa sauvagii) produces the dermaseptin family of nonhemolytic antifungal peptides (38, 125). Dermaseptins are linear cationic, lysine-rich peptides and are believed to lyse microorganisms by interacting with lipid bilayers, leading to alterations in membrane functions responsible for osmotic balance (67, 124, 139). Zone inhibition assays demonstrated that 10 μg/ml suppresses the growth of A. fumigatus (123). Dermaseptins s1 to s5 were potent antifungal agents that inhibited a wide range of fungi (124). Dermaseptin b inhibited the in vitro growth of yeasts and some filamentous fungi; however, the dermaseptin s group was more effective (123).
ANTIFUNGAL PEPTIDES PRODUCED BY BACTERIA AND FUNGI
Iturins.
Various strains of Bacillus subtilis produce the iturin peptide family. They are small cyclic peptidolipids characterized by a lipid-soluble β-amino acid linked to a peptide containing d and l amino acids (136). Iturins affected membrane surface tension, which caused pore formation and which resulted in the leakage of K+ and other vital ions, paralleling cell death (19, 95, 175). One family member, bacillomycin F (Table 3), inhibited the growth of fungi including Aspergillus niger, C. albicans, and F. oxysporum (94, 117). In a disc assay, iturin A inhibited A. flavus and F. moniliforme growth (88). Initial clinical trials involving humans and animals showed that iturin A was effective against dermatomycoses and had a wide spectrum of antifungal properties and low allergenic effects (20, 30). Unfortunately, bacillomycin L and iturin A have been found to be hemolytic, which may reduce their potential use as antifungal drugs (96).
TABLE 3.
Peptide | Source | Structure | Mode of action | Typical target organism | In vitro MIC (μg/ml) |
---|---|---|---|---|---|
1901-II | P. lilacinus | Amino-lipopeptide | Unknown | C. tropicalis | 12.5 |
1907-VIII | P. lilacinus | Amino-peptide | Unknown | C. tropicalis | 50.0 |
A12-C | B. licheniformis | Peptide | Hyphal proliferation | M. canis | Unknown |
Aculeacins | Aspergillus aculeatus | Lipopeptide | Glucan synthesis | C. albicans | 0.2–6.3a |
Aureobasidin A | A. pullulans | Cyclic depsipeptide | Actin assembly | C. neoformans | 0.63 |
Bacillomycin F | Bacillus subtilis | Lipopeptide | Lysis | Aspergillus niger | 40.0 |
CB-1 | B. licheniformis | Lipopeptide | Chitin binding | F. oxysporum | 50.0 (IC50b) |
Cepacidine A1 | B. cepacia | Cyclic glycopeptide | Unknown | A. niger | 0.098 |
Cepacidine A2 | B. cepacia | Cyclic glycopeptide | Unknown | A. niger | 0.096 |
Echinocandin B | A. nidulans | Lipopeptide | Glucan synthesis | C. albicans | 0.625 |
Fungicin M-4 | B. licheniformis | Cyclic peptide | Unknown | Mucor sp. | 8.0 |
FR900403 | Kernia sp. | Lipopeptide | Chitin synthesis | C. albicans | 0.4 |
Helioferin A | M. rosea | Lipopeptide | Unknown | C. albicans | 5.0 |
Helioferin B | M. rosea | Lipopeptide | Unknown | C. albicans | 5.0 |
Iturin A | B. subtilis | Lipopeptide | Lysis | S. cerevisiae | 22.0 |
Leucinostatin A | P. lilacinum | Amino-lipopeptide | Unknown | C. neoformans | 0.5 |
Leucinostatin H | P. marquandii | Amino-lipopeptide | Unknown | C. albicans | 10.0 |
Leucinostatin K | P. marquandii | Amino-lipopeptide | Unknown | C. albicans | 25.0 |
Mulundocandin | A. syndowi | Lipopeptide | Glycan synthesis | C. albicans, A. niger | 0.97 |
31.25 | |||||
Nikkomycin X | Streptomyces tendae | Peptide-nucleoside | Chitin synthesis | C. immitis | 0.125 |
Nikkomycin Z | S. tendae | Peptide-nucleoside | Chitin synthesis | C. immitis | 0.77 |
Pneumocandin A0 | Z. arboricola | Lipopeptide | Glucan synthesis | C. albicans isolates | 0.12–2.0a |
Polyoxin D | S. cacaoi | Trinucleoside peptide | Chitin synthesis | C. immitis | 0.125 |
Pseudomycin A | P. syringae | Lipodepsinonapeptide | Lysis | C. neoformans | 1.56 |
Schizotrin A | Schizotrix sp. | Cyclic undecapeptide | Unknown | C. albicans | 0.02 |
Syringomycin E | P. syringae | Lipodepsipeptide | Lysis | C. neoformans | 0.8–12.5a |
Syringostatin A | P. syringae | Lipodepsipeptide | Lysis (?) | A. fumigatus | 5.0–40.0a |
Syringotoxin B | P. syringae | Lipodepsinonapeptide | Lysis (?) | C. albicans | 3.2–50.0a |
Trichopolyn A | T. polysporum | Amino-lipopeptide | Unknown | C. neoformans | 0.78 |
Trichopolyn B | T. polysporum | Amino-lipopeptide | Unknown | C. neoformans | 0.78 |
WF11899 A | Coleophoma empetri | Lipopeptide | Glucan synthesis | C. albicans | 0.16 |
WF11899 B | C. empetri | Lipopeptide | Glucan synthesis | C. albicans | 0.008 (IC50) |
WF11899 C | C. empetri | Lipopeptide | Glucan synthesis | C. albicans | 0.008 (IC50) |
MICs based on assays with multiple isolates.
IC50, inhibitory concentration for 50% of the population.
Syringomycins and related peptides.
Members of the Pseudomonas syringae pv. syringae group produce small cyclic lipodepsipeptides known as syringomycins (154), the major form being syringomycin E (SE). SE increased transmembrane K+, H+, and Ca2+ fluxes and the membrane potential in plasma membranes of plants and yeasts (142, 167, 169, 192). SE formed voltage-sensitive ion channels, altered protein phosphorylation and H+-ATPase activity (48). Ergosterol was a binding site in yeast for the syringomycins (168). Sorenson et al. (159) published a thorough study of the potent fungicidal properties of several compounds produced by P. syringae, including SE, syringotoxin B, and syringostantin A. These compounds were fungicidal for Candida, Cryptococcus, and Aspergillus isolates (159). A 12% (wt/vol) ointment of SE was effective in controlling vaginal candidiasis in a murine model (160). P. syringae also produced the pseudomycins, another family of peptides with broad-spectrum antifungal activity (68).
CHITIN SYNTHASE INHIBITORS
Nikkomycins.
Nikkomycins, which are produced by Streptomyces tendae, enter target cells via dipeptide permeases and inhibit chitin biosynthesis in C. albicans both in vitro and in vivo (27, 75, 114, 115, 121, 172). Nikkomycins provided antifungal protection to infected kidneys, while other organs were unprotected (27). Nikkomycin Z at high dosages prolonged the survival of mice with disseminated candidiasis (15, 72). Nikkomycins X and Z were active against pathogenic dimorphic fungi but showed only modest to poor activity against yeast and filamentous fungi (73, 74). However, they were highly efficacious in murine models of coccidioidomycosis and blastomycosis, with moderate efficacy against histoplasmosis. Given orally, the nikkomycins prevented the deaths of mice infected with a 100% lethal challenge of C. immitis, with nikkomycin Z being more active than nikkomycin X.
Polyoxins.
Polyoxins, which are produced by Streptomyces cacaoi, were active against isolated chitin synthases but had variable activity against intact organisms (76, 77, 84, 164). Polyoxin D was fungistatic for C. albicans at concentrations of 500 to 2,000 μg/ml, depending on the strain, and inhibited C. neoformans growth (14). Notably, polyoxin D reduced the ability of C. albicans to bind to buccal epithelial cells by as much as 58% compared to the binding ability of controls (61).
FR-900403.
FR-900403 differs in structure from the polyoxins and nikkomycins in that its nucleoside is adenosine and the peptide is linked to the nucleoside at the C-3′ residue. It was active against C. albicans but not against filamentous fungi (86).
PEPTIDES AFFECTING GLUCAN SYNTHESIS
Echinocandins.
Echinocandins, which consist of a diverse family of lipopeptides, are noncompetitive inhibitors of (1,3)-β-d-glucan synthase (13, 119, 134, 150). Their mode of action is similar to that of the papulacandins, naturally occurring antifungal glycolipids (8, 11, 64). The name echinocandin was originally applied to a small family of cyclic lipopeptide antifungal natural products with the same cyclic peptide nucleus but different fatty acid side chains (178). However, the echinocandin peptide family now includes the echinocandins, cilofungin, pneumocandins, aculeacins, mulundocandin, and WF11899 (Table 3). Three excellent reviews describe this peptide family (57, 93, 178). Of the three types of echinocandins (types B, C, and D), type B is the major species produced by some members of the Aspergillus nidulans and Aspergillus rugulosus groups (18, 87, 177). Echinocandins possessed antimicrobial activity against Pneumocystis carnii and C. albicans (10, 152, 178). Since echinocandin B is hemolytic due to the acyl side chain, it has not been used clinically (32, 33, 178).
Echinocandin analogs.
The hemolytic property of the native echinocandins was greatly reduced by enzymatically creating analogs (designated LY compounds) of echinocandin B, listed in Table 4 (50). Cilofungin (LY121019), an analog of echinocandin B, was greater than 10-fold less lytic for erythrocytes than the parent compound and retained potent fungicidal activity (13, 59, 60, 165). Cilofungin also showed excellent in vitro and in vivo activities against Candida spp. and A. fumigatus (13, 40, 137, 146, 161, 165, 183, 187) but displayed only limited activity (151) against P. carinii pneumonia (PCP).
TABLE 4.
Peptide | Structure | Mode of action | Typical target organism | In vitro MIC (μg/ml) |
---|---|---|---|---|
Cilofungin (LY121019) | Lipopeptide | Glucan synthesis | C. albicans | 0.62 |
D4E1 | Linear peptide | Lysis (?) | A. flavus | 26.25 |
L731,373 | Lipopeptide | Glucan synthesis | C. albicans | ≤0.06 |
L733,560 | Lipopeptide | Glucan synthesis | C. albicans | 0.06 |
L743,872 (MK-0991) | Lipopeptide | Glucan synthesis | A. flavus | 0.09–3.12 |
L773,560 | Lipopeptide | Glucan synthesis | C. albicans | 0.5 |
LY303366 | Lipopeptide | Glucan synthesis | Candida krusei | 0.5 |
LY303366, a semisynthetic derivative that has potent in vitro candidacidal properties on the basis of its selective inhibition of β-(1,3)-glucan synthase, is effective against Candida species clinical isolates, with MICs at which 90% of isolates are inhibited (MIC90s) ranging from 0.5 to 4.0 μg/ml in RPMI 1640 (34, 138, 180). MIC90s were considerably lower in antibiotic medium 3, ranging from 0.003 to 2.0 μg/ml. In antibiotic medium 3, LY303366 was 16- to >2,000-fold more active than itraconazole, fluconazole, amphotericin B, and flucytosine against all Candida species except Candida parapsilosis (138). However, in RPMI 1640, the activity of LY303366 was comparable to those of amphotericin B and itraconazole, but it was more active than fluconazole and flucytosine. Against Aspergillus species, LY303366 had a minimum effective concentrations for 90% of isolates tested and an MIC90 of 0.02 and 10.24 μg/ml, respectively (191). It was inactive against C. neoformans and Blastomyces dermatitidis. In contrast, amphotericin B and itraconazole were more potent than LY303366 against Aspergillus isolates. Amphotericin B, flucytosine, fluconazole, and ketoconazole were also more effective against C. neoformans and B. dermatitidis than LY303366. Ernst et al. (45) indicated that the use of the current interpretive endpoint MIC in RPMI 1640 may underestimate the antifungal activity of LY303366 and suggested that alternative media be used to obtain a more accurate MIC endpoint for this peptide. This may also hold true for other antimicrobial peptides. For example, the fungicidal properties of cecropin B and dermaseptin were reduced by increasing the pH of the bioassay media from 6 to 7 (38). A pH increase may neutralize the positive charges on some amino acids near the C terminus, which, in turn, could reduce the ability of the C termini of these peptides to insert into the negatively charged outer membrane, thereby preventing lysis. LY303366 is being studied in phase II clinical trials.
Pneumocandins.
Zalerion arboricola produces the pneumocandins, which were effective against P. carinii infections in rats and which had greater potency and spectra of activity than the echinocandins (50, 152). Pneumocandin A0, the most important member of this group, has potent anti-Candida activity and was more active than echinocandin against experimental murine infections (50). Pneumocandin A0 was generally more active than the echinocandin derivatives tretrahydroechinocandin B and cilofungin (11). However, pneumocandin A0 has no activity against A. flavus, A. fumigatus, C. neoformans, or Candida guilliermondii (50). Pneumocandin A0 was hemolytic at a level (6.25 μg/ml) much higher than that required for activity (50).
Pneumocandin analogs.
L-693,989, a phosphate ester of pneumocandin A, had a 90% minimum effective dose of 0.15 mg/kg of body weight and a 99% minimum effective dose of 3.0 mg/kg in animal models of PCP and candidiasis, respectively (10). In contrast, cilofungin was at least 15 times less potent than L-693,989 in a PCP model. Importantly, L-693,989 produced hemolysis only at levels greater than 400 μg/ml, which was considerably greater than the concentration that inhibited fungal growth.
L-773,560, L-731,373, L-733,560, and L-743,872 are water-soluble, semisynthetic derivatives of pneumocandin B0 and are significantly more potent than the narrow-spectrum parent compound (12, 113). The MICs of these compounds were 0.06 to 4.0 μg/ml for clinical isolates of Candida species, 8 to 64 μg/ml for C. neoformans, and >128 μg/ml for A. flavus and A. fumigatus. These peptides were relatively nonhemolytic for human and mouse erythrocytes. In contrast, amphotericin B was much more hemolytic (12). They were effective against disseminated aspergillosis and candidiasis but not cryptococcosis in murine models and delayed mortality due to pulmonary aspergillosis at an effective dose (administered intraperitoneally) of 5 mg/kg in a rat model (1, 92). Against Candida isolates, the tricationic analogs of pneumocandin, L-731,373 and L-733,560, were more potent than the dicationic analogs, which, in turn, were more potent than the monocationic analogs (188).
The highly soluble compound L-743,872 (MK-0991) was effective against clinically important fungal isolates and was well tolerated by rodents (35, 116). The MICs of L-743,872 were between 0.06 and 4.0 μg/ml for A. flavus and A. fumigatus. It appeared to lack significant in vitro activity against F. oxysporum, Fusarium solani, Rhizopus arrhizus, and Paecilomyces lilacinus but enhanced the efficacies of fluconazole and amphotericin B against C. neoformans (49). It significantly reduced the C. albicans numbers in the mouse kidney compared to the numbers in the kidneys of the controls and enhanced the activities of amphotericin B and fluconazole in vitro against C. neoformans (2, 49). The administration route affected L-743,872, with administration by the oral route being 300-fold less active than administration by the parental route. It was efficacious in mouse target organ assays against Candida tropicalis and other Candida species. This peptide significantly prolonged the survival of DBA/2N mice with disseminated aspergillosis, with 50 and 90% effective doses of 0.03 and 0.12 mg/kg/dose, respectively, at 28 days postchallenge but was ineffective against disseminated C. neoformans infections (2). In animals, the pharmacokinetics of L-743,872 featured a long half-life, ranging from 5.2 to 7.6 h, and the compound slowly accumulated in tissues (66). No significant differences in the in vitro antifungal activity of either LY-303366 or L-743,872 was observed (90). L-743,872 is being investigated in phase II studies.
Aculeacins.
Aculeacins (A through G) are produced by Aspergillus aculeatus (120, 149). The inhibitory concentrations for 50% of the cultures (IC50s) for aculeacin A were 0.008 to 0.62 μg/ml for Candida species and 2.5 μg/ml for A. niger and A. fumigatus (85). Aculeacins A through D, F, and G have good in vitro activity against C. albicans and Saccharomyces cerevisiae but reduced the growth of only a few filamentous fungi (119, 120, 149).
Mulundocandins.
Aspergillus syndowi var. mulundenis produces the mulundocandins, whose structures differ from those of the echinocandins by the replacement of one of the threonines with a serine residue, and the lipophilic side chain is 12-methylmyristoyl rather than lineoyl (127, 147). Mulundocandin and the related compound deoxymulundocandin were found to be active against C. albicans and A. niger (128).
WF11899 group.
Cleophoma empetri F-11899 produces the water-soluble lipopeptides WF11899 A, B, and C. The IC50 for C. albicans ranged from 0.0004 to 0.03 μg/ml (85). These peptides demonstrated potent in vivo anti-Candida activities in a murine model of systemic infection and were superior to cliofungin and fluconazole (85). However, WF11899 A, B, and C lysed mouse erythrocytes in vitro at 62 μg/ml (85).
Aureobasidins.
Aureobasidins are produced by Aureobasidium pullulans (170). This group has 18 members whose structures have eight lipophilic amino acid residues and an α-hydroxyacid (80, 81). Their modes of action and structures differ from those of the echinocandins in that they are believed to alter actin assembly and delocalize chitin in cell walls, resulting in lysis by disruption of cell membranes (44). Another study indicated that sphingolipid synthesis is the target of aureobasidin A (129). Aureobasidins A, B, C, E, S2b, S3, and S4 were potent and had MICs of 0.05 to 3.12 μg/ml for Candida species and C. neoformans isolates. The MICs for Histoplasma capsulatum and Blastomyces dermatitidis were less than 0.63 μg/ml. Aureobasidin A at ≤2.5 μg/ml was also effective against dematiaceous fungi but was inactive against A. fumigatus, A. niger, and A. flavus (91, 171). Its activity was superior to those of fluconazole and amphotericin B against murine candidiasis (171). Synthetic aureobasidin A was highly fungicidal, with MICs of 0.01 to 1.6 μg/ml for Candida species and C. neoformans (91). Aureobasidin showed several desirable properties, including lethality for growing C. albicans cells, a low level of acute toxicity, and improved survival and sterilization of kidneys in a murine model. It was one of the few peptides that had appreciable oral bioavailability (171).
OTHER ANTIFUNGAL PEPTIDES DERIVED FROM BACTERIA AND FUNGI
Bacillus licheniformis peptides.
CB-1 is a chitin-binding peptide containing fatty acids bound to amino acids and has an IC50 for F. oxysporum of 50 μg/ml (130). A B. licheniformis isolate, M-4, produces fungicin M-4 (99). It is a hydrophilic, narrow-spectrum antifungal peptide that is resistant to proteolytic enzymes and lipase and that inhibited the growth of Microsporum canis, Mucor species, and Sporothrix schenckii. However, fungicin M-4 was ineffective against C. albicans, C. neoformans, A. niger, and Trichophyton mentagrophytes. B. licheniformis also produces A12-C, a fungal cell growth and hyphal proliferation inhibitor. A12-C inhibited S. schenckii, T. mentagrophytes, and M. canis growth, as observed in zone-of-inhibition studies (54).
Schizotrin A.
A cyanobacterium, Schizotrix (TAU strain IL-89-2), produces schizotrin A, a cyclic undecapeptide (135). Zone-of-inhibition assays demonstrated that it has activity against C. albicans and C. tropicalis. It also inhibited the radial growth of F. oxysporum at 0.05 μg/ml.
Cepacidines.
Cepacidines A1 and A2 are glycopeptides that have similar structures and that are produced by Burkholderia cepacia (100, 110). Together, they displayed potent antifungal properties superior to those of amphotericin B (100). In vitro, the MICs of cepacidine A ranged from 0.049 to 0.391 μg/ml for Candida species, C. neoformans, A. niger, T. mentagrophytes, Trichophyton rubrum, M. canis, and F. oxysporum (100). Its activity was diminished significantly against C. albicans and C. neoformans in the presence of 50% human serum, which may limit its clinical potential.
1907-II and 1907-VIII.
P. lilacinus produces two antifungal peptides, 1907-II and 1907-VIII, consisting of several amino acids, a methylamine, and a fatty acid (148). In vitro, both peptides have a MIC of 6.25 μg/ml for C. albicans, while C. neoformans was very susceptible (MICs, 0.78 and 1.56 μg/ml for 1907-II and 1907-VIII, respectively).
Leucinostatin-trichopolyn group.
The leucinostatin-trichopolyn group is structurally related to 1907-II and 1907-VIII. Leucinostatins A and B are produced by submerged cultures of Penicillium lilacinum (6, 53). Leucinostatin A and 1907-VIII have the same molecular weight (1,217), while leucinostatin B and 1907-II have a molecular weight of 1,203 (82, 83). Leucinostatin A and B acted as uncouplers on rat mitochondria (126). Leucinostatins D, H, and K were isolated from Paecilomyces marquandii (Massee) Hughes and had a wide spectrum of antimicrobial properties against Candida species, C. neoformans, and other clinically important fungi (140, 145). Unfortunately, it is rather cytotoxic, with the following 50% inhibitory doses: 850 ng/ml for HeLa cells, 0.95 ng/ml for KB cells, and 1.00 ng/ml for P388/S cells. Trichopolyns A and B are produced by Trichoderma polysporum (51, 52). The MICs of trichopolyns A and B for C. albicans, C. neoformans, A. niger, A. fumigatus, and T. mentagrophytes were 0.78 to 6.25 μg/ml.
Helioferins.
Mycogone rosea produces helioferins A and B, which are members of the leucinostatin-trichopolyn group that also may not have clinical utility (63). They inhibited C. albicans (MIC, 5.0 μg/ml) but were toxic to chicken embryos at levels greater than 0.5 mg/kg and caused hemolysis at concentrations greater than 100 μg/ml. They also displayed cytotoxic activities, with IC50s for the L-1210 leukemia cell line and the L0929 mouse fibroblast cell line of 0.01 to 0.4 μg/ml.
PLANT ANTIFUNGAL PEPTIDES
Plant defensins.
Plant defensins (Table 5), which are not related to either the mammalian or the insect defensins, have eight disulfide-linked cysteines comprising a triple-stranded antiparallel β-sheet structure with only one α helix (23, 24). Their mechanisms of action have not yet been elucidated, although the possibility of permeabilization through direct protein-lipid interactions has been eliminated (174). They reduced hyphal elongation without marked morphological distortions (23, 24). Hs-AFP1 and Rs-AFP2 were isolated from Heuchera sanginea and Raphanus sativus seeds, respectively (131, 173). They possess poor lethality for the clinical fungi studied to date. Hs-AFP1 and Rs-AFP2 at a concentration of 125 μg/ml reduced the viability of germinated conidia of A. flavus by only 20 and 35%, respectively (39). In contrast, Hs-AFP1 at 125 μg/ml reduced the viabilities of nongerminated and germinating conidia of F. moniliforme by 42 and 85%, respectively, while Rs-AFP2 reduced the viabilities of these conidial types by 25 and 95%, respectively. Hs-AFP1 and Rs-AFP2 bound at different rates to mannan, chitin, cholesterol, ergosterol, galactocerebrosides, and sphingomyelin (39).
TABLE 5.
Peptide | Source | No. of amino acids | Mode of action | Typical target organism | In vitro MIC (μg/ml) |
---|---|---|---|---|---|
ACE-AMP1 | A. cepa | 84 | Unknown | F. oxysporum | 0.3 (IC50a) |
Hs-AFP1 | H. sanginea | 54 | Unknown | F. moniliforme | 125.0 |
Ib-AMP3 | I. balsamina | 20 | Unknown | F. moniliforme | 50.0 |
Rs-AFP2 | R. sativus | 51 | Unknown | F. moniliforme | 125.0 |
Zeamatin | Z. mays | 27 | Lysis (?) | C. albicans | 0.5 |
IC50, inhibitory concentration for 50% of the population.
Impatiens balsamina produces a highly basic peptide, Ib-AMP3, with four cysteine residues that form two intramolecular disulfide bridges (166). Ib-AMP3 at 50 μg/ml reduced the viability of germinated conidia of A. flavus by 42%, but it did not affect the viability of nongerminated conidia (39). At 50.0 μg/ml it was highly effective against the nongerminated and germinated conidia of F. moniliforme, reducing their viabilities by 95 and 99.5%, respectively, and had a very high affinity for chitin (39).
Lipid transfer proteins.
Some plants produce lipid transfer proteins, a family of homologous peptides having eight disulfide-linked cysteines. Onion seeds (Allium cepa L.) produce the lipid transfer peptide ACE-AMP1, which inhibited F. oxysporum (26).
Zeamatin.
Zea mays seeds produce the peptide zeamatin, which belongs to a third class of plant antifungal compounds (144). Peptides in the zeamatin family are also present in Avena sativa, Sorghum bicolor, and Triticum aestivum seeds (181). Zeamatin caused the release of cytoplasmic material from C. albicans and Neurospora crassa, resulting in hyphal rupture. It appears to permeabilize the fungal plasma membrane and inhibited C. albicans. Zeamatin activity was reduced by increasing concentrations of NaCl. A flax seed antifungal peptide similar to zeamatin, in synergy with nikkomycin Z, inhibits C. albicans (22).
Cyclopeptides.
Members of the family Rhamnaceae and other plant families produce the basic cyclopeptides in which a 10- or 12-membered peptide-type bridge spans the 1,3 or 1,4 positions of a benzene ring (62). The antifungal properties of many family members have not yet been determined. Frangufoline, amphibine H, rugosanines A and B, and nummularines B, K, R, and S showed significant activity against A. niger but not C. albicans in zonal inhibition studies (132).
SYNTHETIC PEPTIDES
D4E1.
D4E1 is a synthetic peptide that is active against germinated conidia of Aspergillus species, producing 50% lethal doses of between 2.1 and 16.8 μg/ml for several Aspergillus species and a 50% lethal dose of 1.1 μg/ml for F. moniliforme and F. oxysporum (36). Since D4E1 complexes with ergosterol, its mode of action may be lytic. D4E1 was more resistant in vitro to degradation by A. flavus proteases than the insect peptide cecropin A.
CONCLUSIONS
In conclusion, there has been a marked expansion of our knowledge of new antifungal peptides. Some of these agents have reached clinical trials, while others are undergoing detailed preclinical testing. Discovery and elucidation of antimicrobial peptides expand our understanding of intrinsic host defenses and provide new approaches to antifungal chemotherapy. The membership of this group will expand as additional natural peptides are isolated and identified and analogs of natural peptides or totally synthetic ones are produced.
REFERENCES
- 1.Abruzzo G K, Flattery A M, Gill C J, Long L, Smith J G, Krupa D, Pikounis V B, Kroop H, Bartizal K. Evaluation of water-soluble pneumocandin analogs L-733560, L-705589, and L-731373 with mouse models of disseminated aspergillosis, candidiasis, and cryptococcosis. Antimicrob Agents Chemother. 1995;39:1077–1081. doi: 10.1128/aac.39.5.1077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Abruzzo G K, Flattery A M, Gill C J, Long L, Smith J G, Pikounis V B, Balkovec J M, Bouffard A F, Droponski J F, Rosen H, Kropp H, Bartizal K. Evaluation of the echinocandin antifungal MK-0991 (L-743,8872): efficacies in mouse models of disseminated aspergillosis, candidiasis, and cryptococcosis. Antimicrob Agents Chemother. 1997;41:2333–2338. doi: 10.1128/aac.41.11.2333. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Alcouloumbre M S, Gharinoum M A, Ibrahim A S, Selsted M E, Edwards J E. Fungicidal properties of defensin NP-1 and activity against Cryptococcus neoformans in vitro. Antimicrob Agents Chemother. 1993;37:2628–2632. doi: 10.1128/aac.37.12.2628. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Ammons S, Aardalen K, Froebel S, Little R. Program and abstracts of the 37th Interscience Conference on Antimicrobial Agents and Chemotherapy. Washington, D.C: American Society for Microbiology; 1997. Efficacy of domain III peptide from bactericidal/permeability-increasing protein (BPI) in murine disseminated aspergillosis. abstr. B-16; p. 29. [Google Scholar]
- 5.Appenzeller L, Lim E, Wong P, Fadem M, Motchinik P, Bakalinsky M, Little R. Program and abstracts of the 36th Interscience Conference on Antimicrobial Agents and Chemotherapy. Washington, D.C: American Society for Microbiology; 1996. In vivo fungicidal activity of optimized domain III peptides derived from bactericidal/permeability-increasing protein (BPI), abstr. F187; p. 132. [Google Scholar]
- 6.Arai T, Mikami Y, Fukushima T, Utsumi T, Yazawa K. A new antibiotic, leucinostatin, derived from Penicillium lilacinum. J Antibiot. 1973;26:1606–1612. doi: 10.7164/antibiotics.26.157. [DOI] [PubMed] [Google Scholar]
- 7.Aumelas A, Mangoni M, Roumestand C, Chiche L, Despaux E, Grassy G, Calas B, Chavanieu A. Synthesis and solution structure of the antimicrobial peptide protegrin-1. Eur J Biochem. 1996;237:575–583. doi: 10.1111/j.1432-1033.1996.0575p.x. [DOI] [PubMed] [Google Scholar]
- 8.Baguley B C, Rommele G, Gruner J, Wehrlli W. Papulacandin B: an inhibitor of glucan synthesis in yeast spheroplasts. Eur J Biochem. 1979;97:345–351. doi: 10.1111/j.1432-1033.1979.tb13120.x. [DOI] [PubMed] [Google Scholar]
- 9.Balkovec J. Lipopeptide antifungal agents. Expert Opin Invest Drugs. 1994;3:65–82. [Google Scholar]
- 10.Balkovec J M, Black R M, Hammond M L, Heck J V, Zambias R A, Abruzzo G, Bartizal K, Kroop H, Trainor C, Schwartz R E, McFadden D C, Nollstadt K H, Pittarelli L A, Powles M A, Schmatz D M. Synthesis, stability, and biological evaluation of a new echinocandin lipopeptide. Discovery of a potential clinical agent for the treatment of systemic candidiasis and Pneumocystis carinii pneumonia. J Med Chem. 1992;35:194–198. doi: 10.1021/jm00079a027. [DOI] [PubMed] [Google Scholar]
- 11.Bartizal K, Abruzzo G, Trainor C, Krupa D, Nollstadt K, Schmatz D, Schwartz R, Hammond M, Balkovec J, Vanmiddlesworth F. In vitro antifungal activities and in vivo efficacies of 1,3-β-d-glucan synthesis inhibitors L-671,329, L-646,991, tetrahydroechinocandin B, and L-687,781, a papulacandin. Antimicrob Agents Chemother. 1992;36:1648–1657. doi: 10.1128/aac.36.8.1648. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Bartizal K, Scott T, Abruzzo G K, Gill C J, Pacholok C, Lynch L, Kropp H. In vitro evaluation of the pneumocandin antifungal agent L-733560, a new water-soluble hybrid of L-705589 and L-731,373. Antimicrob Agents Chemother. 1995;39:1070–1076. doi: 10.1128/aac.39.5.1070. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Beaulieu D, Tang J, Zeckner D J, Parr T R. Correlation of cilofungin in vivo efficacy with its activity against Aspergillus fumigatus (1,3)-β-d-glucan synthase. FEMS Microbiol Lett. 1993;108:133–138. doi: 10.1111/j.1574-6968.1993.tb06088.x. [DOI] [PubMed] [Google Scholar]
- 14.Becker J M, Covert N L, Shenbagamurthi P, Steinfeld A S, Naider F. Polyoxin D inhibits growth of zoopathogenic fungi. Antimicrob Agents Chemother. 1983;23:926–929. doi: 10.1128/aac.23.6.926. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Becker J M, Marcus S, Tullek J, Miller D, Krainer E, Khare R K, Narder F. Use of the chitin synthesis inhibitor nikkomycin to treat disseminated candidiasis in mice. J Infect Dis. 1988;157:212–214. doi: 10.1093/infdis/157.1.212. [DOI] [PubMed] [Google Scholar]
- 16.Bellamy W, Takase M, Yamauchi K, Wakabayashi H, Kawase K, Tomita M. Identification of the bactericidal domain of lactoferrin. Biochem Biophys Acta. 1992;1121:130–136. doi: 10.1016/0167-4838(92)90346-f. [DOI] [PubMed] [Google Scholar]
- 17.Bellamy W, Wakabayashi H, Takase M, Shimamura S, Tomita M. Role of cell-binding in the antibacterial mechanism of lactoferricin. J Appl Bacteriol. 1993;75:478–484. [PubMed] [Google Scholar]
- 18.Benz F, Knuesel F, Nuesch J, Treicher H, Voser W, Nyfeler R, Keller-Schlerein W. Echinocandin B, ein neuartiges Polypetid Antibioticum aus Aspergillus nidulans var echinlatus: Isolierung and Baudsteine. Helv Chim Acta. 1985;57:2459–2477. doi: 10.1002/hlca.19740570818. [DOI] [PubMed] [Google Scholar]
- 19.Besson F, Peypoux M, Quentin J, Michel G. Action of antifungal peptolipids from Bacillus subtilis on the cell membrane of Saccharomyces cerevisiae. J Antibiot. 1984;37:172–177. doi: 10.7164/antibiotics.37.172. [DOI] [PubMed] [Google Scholar]
- 20.Bloquiaux S, Delcambre L. Essais de traitement de dermatomycoses par l’iturine. Arch Belg Derm Syph. 1956;12:224. [PubMed] [Google Scholar]
- 21.Boman H G, Hultmark D. Cell-free immunity in insects. Annu Rev Microbiol. 1987;41:103–126. doi: 10.1146/annurev.mi.41.100187.000535. [DOI] [PubMed] [Google Scholar]
- 22.Borgmeyer J R, Smith C E, Khai Hutnk Q. Isolation and characterization of a 25 kDa antifungal protein from flax seeds. Biochem Biophys Res Commun. 1992;187:480–487. doi: 10.1016/s0006-291x(05)81519-0. [DOI] [PubMed] [Google Scholar]
- 23.Bruix M, Gonzales C, Santoro J, Soriano F, Rocher A, Mendez E, Rico M. 1HNMR studies on the structure of a new thionin from barely endosperm. Biopolymers. 1995;36:751–763. doi: 10.1002/bip.360360608. [DOI] [PubMed] [Google Scholar]
- 24.Bruix M, Jimenez M A, Santora J, Gonzales C, Colilla F J, Mendez E, Rico M. Solution structure of γ-1-P thionins from barley and wheat endosperm determined by 1H-NMR: a structural motif common to toxic arthropod proteins. Biochemistry. 1993;132:715–724. doi: 10.1021/bi00053a041. [DOI] [PubMed] [Google Scholar]
- 25.Bullen J J. The significance of iron in infection. Rev Infect Dis. 1981;3:1127–1138. doi: 10.1093/clinids/3.6.1127. [DOI] [PubMed] [Google Scholar]
- 26.Cammue B P A, Thevissen K, Hendricks M, Eggermont K, Goderis I J, Proost P, Van Damme J, Osborn R W, Guerbette F, Kader J-C, Broekaert W F. A potent antimicrobial protein from onion seeds showing sequence homology to plant lipid transfer protein. Plant Pathol. 1995;109:445–455. doi: 10.1104/pp.109.2.445. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Chapman T, Kinsman O, Houston J. Chitin biosynthesis in Candida albicans grown in vitro and in vivo and its inhibition by nikkomycin Z. Antimicrob Agents Chemother. 1992;36:1909–1914. doi: 10.1128/aac.36.9.1909. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Chen H-C, Boman J H, Morell J L, Huang C M. Synthetic magainin analogues with improved antimicrobial activity. FEBS Lett. 1988;236:462–466. doi: 10.1016/0014-5793(88)80077-2. [DOI] [PubMed] [Google Scholar]
- 29.Christensen B, Fink J, Merrifield R B, Mauzerall D. Channel-forming properties of cecropins and related model compounds incorporated into planar lipid membranes. Proc Natl Acad Sci USA. 1988;85:5072–5076. doi: 10.1073/pnas.85.14.5072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Clairbois J P, Delcambre L. A propos d’essais cliniques et biologiques sur l’iturine, antifongique noveau. Arch Belg Derm Syph. 1958;14:63. [PubMed] [Google Scholar]
- 31.Cociancich S, Ghazi A, Hetru A, Hoffman J A, Letellier L. Insect defensin, an inducible antibacterial peptide, forms voltage-dependent channels in Micrococcus luteus. J Biol Chem. 1993;260:19239–19245. [PubMed] [Google Scholar]
- 32.Debono M, Abbott B J, Fukuda D, Barnhart M, Willard K E, Molloy R M, Michel K H, Turner J R, Bulter T F, Hunt A H. Synthesis of new analogs of echinocandin B by enzymatic deacylation and chemical reacylation of the echinocandin B peptide: synthesis of the antifungal agent cilofungin ( LY121019) J Antibiot. 1989;42:389–397. doi: 10.7164/antibiotics.42.389. [DOI] [PubMed] [Google Scholar]
- 33.Debono M, Abbott B J, Turner J R, Howard L C, Gordee R S, Hunt A S, Barnhart M, Molloy R M, Willard K E, Fukuda D, Butler T F, Zeckner D J. Synthesis and evaluation of LY12019, a member of a series of semisynthetic analogues of the antifungal lipopeptide echinocandin B. Annu Rev N Y Acad Sci. 1988;544:152–167. doi: 10.1111/j.1749-6632.1988.tb40398.x. [DOI] [PubMed] [Google Scholar]
- 34.Debono M, Gordee R S. Antibiotics that inhibit fungal cell wall development. Annu Rev Microbiol. 1994;48:471–497. doi: 10.1146/annurev.mi.48.100194.002351. [DOI] [PubMed] [Google Scholar]
- 35.Del Porta M, Schell W A, Perfect J R. In vitro antifungal activity of pneumocandin L-743,872 against a variety of clinically important molds. Antimicrob Agents Chemother. 1997;41:1835–1836. doi: 10.1128/aac.41.8.1835. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.De Lucca A J, Bland J M, Grimm C, Jacks T J, Cary J W, Jaynes J M, Cleveland T E, Walsh T J. Fungicidal properties, sterol binding, and proteolytic resistance of the synthetic peptide, D4E1. Can J Microbiol. 1998;44:514–520. doi: 10.1139/w98-032. [DOI] [PubMed] [Google Scholar]
- 37.De Lucca A J, Bland J M, Jacks T J, Grimm C, Cleveland T E, Walsh T J. Fungicidal activity of cecropin A. Antimicrob Agents Chemother. 1997;41:481–483. doi: 10.1128/aac.41.2.481. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.De Lucca A J, Bland J M, Jacks T J, Grimm C, Walsh T J. Fungicidal and binding properties of the natural peptides cecropin B and dermaseptin. Med Mycol. 1998;36:291–298. [PubMed] [Google Scholar]
- 39.De Lucca, A. J., T. J. Jacks, and W. F. Broekaert. Fungicidal and binding properties of three plant peptides. Submitted for publication. [DOI] [PubMed]
- 40.Denning D W, Stephans D A. Effiicacy of cilofungin alone and in combination with amphotericin B in a murine model of disseminated aspergillosis. Antimicrob Agents Chemother. 1991;35:1329–1333. doi: 10.1128/aac.35.7.1329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Diamond G, Zasloff M, Eck H, Brasseur M, Maloy W L, Bevins C L. Tracheal antimicrobial peptide, a cysteine-rich peptide from mammalian tracheal mucosa: peptide isolation and cloning of cDNA. Proc Natl Acad Sci USA. 1991;88:3952–3956. doi: 10.1073/pnas.88.9.3952. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Eisenhauer P, Harwig S, Szlarek D, Ganz T, Selsted M, Lehrer R. Purification and antimicrobial properties of three defensins from rat neutrophils. Infect Immun. 1989;57:2021–2027. doi: 10.1128/iai.57.7.2021-2027.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Elsbach P, Weiss J. The bactericidal/permiability-increasing protein (BPI), a potent element in host-defense against gram-negative bacteria and lipopolysaccharide. Immunobiology. 1997;187:417–429. doi: 10.1016/S0171-2985(11)80354-2. [DOI] [PubMed] [Google Scholar]
- 44.Endo M, Takesako K, Kato I, Yamaguchi H. Fungicidal action of aureobasidin A, a cyclic depsipeptide antifungal antibiotic, against Saccharomyces cerevisiae. Antimicrob Agents Chemother. 1997;41:672–676. doi: 10.1128/aac.41.3.672. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Ernst M E, Klepser M E, Wolfe E J, Pfaller M A. Antifungal dynamics of LY 303366, an investigational echinocandin B analog, against Candida spp. Diagn Microbiol Infect Dis. 1996;26:125–131. doi: 10.1016/s0732-8893(96)00202-7. [DOI] [PubMed] [Google Scholar]
- 46.Fehlbaum P, Bulet P, Chernych S, Briand J-P, Roussel J P, Letellier L, Hetru C, Hoffmman J A. Structure-activity analysis of thanatin, a 21-residue inducible insect defense peptide with sequence homology to frog skin antimicrobial peptides. Proc Natl Acad Sci USA. 1996;93:1221–1225. doi: 10.1073/pnas.93.3.1221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Fehlbaum P, Bulet P, Michaut L, Laguex N, Broekaert W F, Hetru C, Hoffmman J A. Insect immunity. Septic injury of Drosophila induces the synthesis of a potent antifungal peptide with sequence homology to plant antifungal peptides. J Biol Chem. 1994;269:33159–33163. [PubMed] [Google Scholar]
- 48.Feign A M, Takemoto J Y, Wangspa R, Teeter J H, Brand J G. Properties of voltage-gated ion channels formed by syringomycin-E in planar lipid bilayers. J Membr Biol. 1996;149:41–47. doi: 10.1007/s002329900005. [DOI] [PubMed] [Google Scholar]
- 49.Franzot S, Casadevall A. Pneumocandin L-743,872 enhances the activities of amphotericin B and fluconazole against Cryptococcus neoformans in vitro. Antimicrob Agents Chemother. 1997;41:331–336. doi: 10.1128/aac.41.2.331. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Fromtling R, Abruzzo G K. L-671,329, a new antifungal agent. III. In vitro activity, toxicology, and efficacy in comparison to aculeacin. J Antibiot. 1989;42:174–178. doi: 10.7164/antibiotics.42.174. [DOI] [PubMed] [Google Scholar]
- 51.Fuji K, Fujita E, Takaishi Y, Fujita T, Arita I, Komatsu M, Hirasuka N. New antibiotics, trichopolyns A and B: isolation and biological activity. Experientia. 1978;34:237–239. doi: 10.1007/BF01944702. [DOI] [PubMed] [Google Scholar]
- 52.Fujita T, Takaishi Y, Okamura A, Fujita E, Fuji K, Hirasuka N, Komatsu M, Arita I. New peptide antibiotics, tricopolyns I and II, from Trichoderma polysporum. J Chem Soc Chem Commun. 1981;1981:585–587. [Google Scholar]
- 53.Fukushima K, Arai T, Mori Y, Tsuboi M, Suzuki M. Studies on peptide antibiotics, leucinostatins. I. Separation, physico-chemical properties and biological activities of leucinostatins A and B. J Antibiot. 1983;36:1606–1612. doi: 10.7164/antibiotics.36.1606. [DOI] [PubMed] [Google Scholar]
- 54.Gàlvez A, Maqueda M, Martinez-Bueno M, Lebbadi M, Valdivia E. Isolation and physico-chemical characterization of an antifungal and antibacterial peptide produced by Bacillus licheniformis A 12. Appl Microbiol Biotechnol. 1993;38:438–442. doi: 10.1007/BF00205029. [DOI] [PubMed] [Google Scholar]
- 55.Ganz T, Selsted M E, Lehrer R I. Defensins. Eur J Haematol. 1990;44:1–8. doi: 10.1111/j.1600-0609.1990.tb00339.x. [DOI] [PubMed] [Google Scholar]
- 56.Ganz T, Selsted M E, Szklarek D, Harwig S S L, Daher K, Bainton D F, Lehrer R I. Defensins: natural peptide antibiotics of human neutrophils. J Clin Invest. 1985;76:1427–1435. doi: 10.1172/JCI112120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Georgopapadakou N H, Walsh T J. Antifungal agents: chemotherapeutic targets and immunologic strategies. Antimicrob Agents Chemother. 1996;40:279–291. doi: 10.1128/aac.40.2.279. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Gera J F, Lichenstein A. Human neutrophil peptide defensins induce single strand DNA breaks in target cells. Cell Immunol. 1991;138:108–120. doi: 10.1016/0008-8749(91)90136-y. [DOI] [PubMed] [Google Scholar]
- 59.Gordee R S, Zeckner D J, Ellis L F, Thakker A L, Howard L C. In vitro and in vivo anti-Candida activity and toxicology of LY121019. J Antibiot. 1984;37:1054–1065. doi: 10.7164/antibiotics.37.1054. [DOI] [PubMed] [Google Scholar]
- 60.Gordee R S, Zeckner D J, Howard L C, Alborn W E, Jr, Debono M. Anti-Candida activity and toxicology of LY121019, a novel polypeptide antifungal antibiotic. Ann N Y Acad Sci. 1988;544:294–301. doi: 10.1111/j.1749-6632.1988.tb40415.x. [DOI] [PubMed] [Google Scholar]
- 61.Gottlieb S, Altboum Z, Savage D C, Segal E. Adhesion of Candida albicans to epithelial cells effect of polyoxin D. Mycopathology. 1991;115:197–216. doi: 10.1007/BF00462227. [DOI] [PubMed] [Google Scholar]
- 62.Gournelis D C, Laskaris G G, Verpoorte R. Cyclopeptide alkaloids. Nat Prod Rep. 1997;14:75–82. doi: 10.1039/np9971400075. [DOI] [PubMed] [Google Scholar]
- 63.Gräfe U, Ihn W, Ritzau M, Schade W, Stengel C, Schlegel B, Fleck W F, Kunkel W, Hartle A, Gutsche W. Helioferins: novel antifungal lipopeptides from Mycogone rosea: screening, isolation, structures, and biological properties. J Antibiot. 1995;48:126–133. doi: 10.7164/antibiotics.48.126. [DOI] [PubMed] [Google Scholar]
- 64.Gruner J, Traxler P. Papulacandin, a new antibiotic, active especially against yeasts. Experientia. 1977;33:137. [Google Scholar]
- 65.Haggius S D, Reed W A, Fatemi M B, White K L, Enright F M, Elzer P H. Abstracts of the 96th General Meeting of the American Society for Microbiology 1996. Washington, D.C: American Society for Microbiology; 1996. The brucellacidal activity in transgenic mice expressing a synthetic cecropin-like peptide or in mice following exogenous peptide treatment, abstr. E-46; p. 274. [Google Scholar]
- 66.Hajdu R, Thompson R, Sundelof J G, Pelak B A, Bouffard F A, Dropinski J F, Kropp H. Preliminary animal pharmacokinetics of the parenteral antifungal agent MK-0991 (L-743,872) Antimicrob Agents Chemother. 1997;41:2339–2344. doi: 10.1128/aac.41.11.2339. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Hani K, Nicolas P, Mor A. Structure-function relationships of antimicrobial dermaseptins. In: Maia H L S, editor. Proceedings of the 23rd European Peptide Symposium. Leiden, The Netherlands: Escom; 1994. pp. 47–49. [Google Scholar]
- 68.Harrison L, Teplow D B, Rinaldi M, Strobel G. Pseudomycins, a family of novel peptides from Pseudomonas syringae possing broad-spectrum antifungal activity. J Gen Microbiol. 1991;137:2857–2865. doi: 10.1099/00221287-137-12-2857. [DOI] [PubMed] [Google Scholar]
- 69.Harwig S S L, Swiderek K M, Kokryakov V N, Tan L, Lee T D, Panyutich E A, Aleshina G M, Shamova O V, Lehrer R I. Gallinacins: cysteine-rich antimicrobial peptides of chicken leukocytes. FEBS Lett. 1994;342:281–285. doi: 10.1016/0014-5793(94)80517-2. [DOI] [PubMed] [Google Scholar]
- 70.Harwig S S L, Swiderek K M, Lee T D, Lehrer R I. Determination of disulfide bridges in PG-2, an antimicrobial peptide from porcine leukocytes. J Pept Res. 1995;3:207–215. doi: 10.1002/psc.310010308. [DOI] [PubMed] [Google Scholar]
- 71.Harwig S S L, Waring L A, Yang H J, Cho Y, Tan L, Lehrer R I. Intramolecular disulfide bonds enhance the antimicrobial and lytic activities of protegrins at physiological sodium chloride concentrations. Eur J Biochem. 1996;240:352–357. doi: 10.1111/j.1432-1033.1996.0352h.x. [DOI] [PubMed] [Google Scholar]
- 72.Hector R F, Schaller K. Positive interaction of nikkomycins and azoles against Candida albicans in vitro and in vivo. Antimicrob Agents Chemother. 1992;36:1284–1289. doi: 10.1128/aac.36.6.1284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Hector R F, Zimmer B L, Pappagianis D. Evaluation of nikkomycins X and Z in murine models of coccodomycosis, histoplasmosis, and blastomycosis. Antimicrob Agents Chemother. 1990;34:587–593. doi: 10.1128/aac.34.4.587. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Hector R F, Zimmer B L, Pappagianis D. Inhibition of cell wall synthesis: nikkomycins. In: Yamaguchi H, Kobayashi G S, Takahish H, editors. Recent progress in antifungal chemotherapy. New York, N.Y: Marcel Dekker, Inc.; 1991. pp. 341–353. [Google Scholar]
- 75.Hóhne H. Nikkomycin, ein neuer Hemmstoff der mikrobiellen Chitin Synthese. Ph.D. dissertation. Tubingen, Germany: Universität Tubingen; 1974. [Google Scholar]
- 76.Hori M, Eguchi J, Kakiki K, Misato T. Studies of the mode of action of polyoxins. VI. Effect of polyoxin B on chitin synthesis in polyoxin-resistant strains of Alternaria kikuchiana. J Antibiot. 1974;27:260–266. doi: 10.7164/antibiotics.27.260. [DOI] [PubMed] [Google Scholar]
- 77.Hori M, Kakiki K, Misato T. Interaction between polyoxin and active center of chitin synthetase. Agric Biol Chem. 1974;38:699–705. [Google Scholar]
- 78.Horwitz A, Motchink P, Nadell R. Abstracts of the 37th Interscience Conference on Antimicrobial Agents and Chemotherapy. Washington, D.C: American Society of Microbiology; 1997. Fungicidal properties from bactericidal/permiability-increasing protein (BPI) act synergistically with fluconazole on a variety of Candida strains, abstr. F-102; p. 163. [Google Scholar]
- 79.Iijima R, Kurata S, Natori S. Purification, characterization, and cDNA cloning of an antifungal protein from the hemolymph of Sarcophaga peregina (flesh fly) larvae. J Biol Chem. 1993;268:12055–12061. [PubMed] [Google Scholar]
- 80.Ikai K, Shiomi K, Takesako K, Mizutanis S, Yamamoto J, Ogawa Y, Ueno M, Kato I. Structures of the aureobasidins B to R. J Antibiot. 1991;44:1187–1198. doi: 10.7164/antibiotics.44.1187. [DOI] [PubMed] [Google Scholar]
- 81.Ikai K, Takesako K, Shiomi K, Moriguchi M, Yamamoto J, Kato I, Naganawa H. Structure of aureobasidin-A. J Antibiot. 1991;44:925–933. doi: 10.7164/antibiotics.44.925. [DOI] [PubMed] [Google Scholar]
- 82.Isogai A, Suzuki A, Higashikawa S, Kuyama S, Tamura S. Constituents of a peptidal antibiotic P168 produced by Paecilomyces lilacinus (Thom) Samson. Agric Biol Chem. 1980;44:3029–3031. [Google Scholar]
- 83.Isogai A, Suzuki A, Higashikawa S, Kuyama S, Tamura S. Structure of peptidal antibiotic P168 produced by Paecilomyces lilacinus (Thom) Samson. Agric Biol Chem. 1980;44:3033–3035. [Google Scholar]
- 84.Isono K, Asahi K, Suzuki S. Studies on polyoxins, antifungal antibiotics. XIII. The structure of polyoxins. J Am Chem Soc. 1969;91:7490–7505. doi: 10.1021/ja01054a045. [DOI] [PubMed] [Google Scholar]
- 85.Iwamoto T, Fuji A, Nitta K, Hashimoto S, Okuhara M, Kohsaka M. WF11899A, B, and C novel antifungal lipopeptides. II. Biological properties. J Antibiot. 1994;47:1092–1097. doi: 10.7164/antibiotics.47.1092. [DOI] [PubMed] [Google Scholar]
- 86.Iwamoto T, Fujiie A, Tsurumi Y, Nanbata K, Shibuya K. FR900403, a new antifungal produced by a Kernia sp. J Antibiot. 1990;43:1183–1185. doi: 10.7164/antibiotics.43.1183. [DOI] [PubMed] [Google Scholar]
- 87.Keller-Juslen C, Huhn M, Loosli H R, Petcher T P, Weber H P, Von Wartburg A. Strucktur des Cyclopeptid-Antibiotikums SL7810 (=echinocandin B) Tetrahedron Lett. 1976;46:4147–4150. [Google Scholar]
- 88.Klich M A, Lax A R, Bland J M. Inhibition of some mycotoxigenic fungi by iturin A, a peptidolipid produced by Bacillus subtilis. Mycotpathology. 1991;116:77–80. doi: 10.1007/BF00436368. [DOI] [PubMed] [Google Scholar]
- 89.Kokryakov V N, Harwig S S L, Panyutich E A, Shevchenko A A, Aleshina G M, Shanova O V, Korneva H A, Lehrer R I. Protegrins: leucocyte antimicrobial peptides that combine features of corticostatic defensins and tachyplesins. FEBS Lett. 1993;327:231–236. doi: 10.1016/0014-5793(93)80175-t. [DOI] [PubMed] [Google Scholar]
- 90.Krishnarao T V, Galgiani J N. Comparison of the in vitro activities of the echinocadin LY303366, the pneumocandin MK-0991, and fluconazole against Candida species and Cryptococcus neoformans. Antimicrob Agents Chemother. 1997;41:1957–1960. doi: 10.1128/aac.41.9.1957. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Kurmoe T, Inami K, Inoue T, Ikai K, Takesako K, Kato I, Shiba T. Total synthesis of an antifungal cyclic depsipeptide aureobasidin A. Tetrahedron. 1996;52:4327–4356. [Google Scholar]
- 92.Kurtz M B, Bernard E M, Edwards F F, Marrinan J A, Dropinski J, Douglas C M, Armstrong D. Aerosol and parenteral pneumocandins are effective in a rat model of pulmonary aspergillosis. Antimicrob Agents Chemother. 1995;39:1784–1789. doi: 10.1128/aac.39.8.1784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Kurtz M B, Douglas C M. Lipopeptide inhibitors of fungal glucan synthase. J Med Vet Mycol. 1997;35:79–86. doi: 10.1080/02681219780000961. [DOI] [PubMed] [Google Scholar]
- 94.Landy M, Warren G H, Roseman S B, Colio L G. Bacillomycin, an antibiotic from Bacillus subtilis active against pathogenic fungi. Proc Soc Exp Biol Med. 1948;67:539–541. doi: 10.3181/00379727-67-16367. [DOI] [PubMed] [Google Scholar]
- 95.Latoud C, Peypoux F, Michel G. Action of iturin A, an antifungal antibiotic from Bacillus subtilis on the yeast Saccharomyces cerevisiae. Modification of membrane permeability and lipid composition. J Antibiot. 1987;40:1588–1595. doi: 10.7164/antibiotics.40.1588. [DOI] [PubMed] [Google Scholar]
- 96.Latoud C, Peypoux F, Michel G, Genet R, Morgat J L. Interactions of antibiotics of the iturin group with human erythrocytes. Biochim Biophys Acta. 1986;856:526–535. doi: 10.1016/0005-2736(86)90144-6. [DOI] [PubMed] [Google Scholar]
- 97.Lawyer C, Pai S, Watebe M, Borgia P, Mashimo T, Eagleton L, Watebe K. Antimicrobial activity of a 13-amino acid tryptophan-rich peptide derived from a putative porcine precursor protein of a novel family of antibacterial peptides. FEBS Lett. 1996;390:95–98. doi: 10.1016/0014-5793(96)00637-0. [DOI] [PubMed] [Google Scholar]
- 98.Lawyer C S, Pal S, Watebe M, Bakir H, Eagleton L, Watebe K. Effects of synthetic form of tracheal antimicrobial peptide on respiratory pathogens. J Antimicrob Chemother. 1996;37:599–604. doi: 10.1093/jac/37.3.599. [DOI] [PubMed] [Google Scholar]
- 99.Lebbadi M, Galvez A, Maqueda M, Martinez-Bueno M, Valdivia E. Fungicin M-4: a narrow spectrum peptide antibiotic from Bacillus licheniformis M-4. J Appl Bacteriol. 1994;77:49–53. doi: 10.1111/j.1365-2672.1994.tb03043.x. [DOI] [PubMed] [Google Scholar]
- 100.Lee C H, Kim S H, Hyun B C, Suh J W, Yon C, Kim C O, Lim Y A, Kim C S. Cepacidine A, a novel antifungal antibiotic produced by Pseudomonas cepacia. I. Taxonomy, production, isolation, and biological activity. J Antibiot. 1994;47:1402–1405. doi: 10.7164/antibiotics.47.1402. [DOI] [PubMed] [Google Scholar]
- 101.Lee S Y, Moon H-J, Kurata S, Natori S, Lee B L. Purification and cDNA cloning of an antifungal protein from the hemolymph of Holotrichia diomphalia larvae. Biol Pharm Bull. 1995;18:1049–1052. doi: 10.1248/bpb.18.1049. [DOI] [PubMed] [Google Scholar]
- 102.Lehrer R I, Barton A, Daher K A, Harwig S S L, Ganz T, Selsted M E. Interaction of human defensins with Escherichia coli. Mechanism of activity. J Clin Invest. 1989;84:553–561. doi: 10.1172/JCI114198. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Lehrer R I, Ganz T, Szklarek D, Selsted M E. Modulation of the in situ candidacidal activity of human neutrophil defensins by target cell metabolism and divalent cations. J Clin Invest. 1988;81:1829–1835. doi: 10.1172/JCI113527. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Lehrer R I, Lichtenstein A K, Ganz T. Defensins: antimicrobial and cytotoxic peptides of mammalian cells. Annu Rev Immunol. 1993;11:105–128. doi: 10.1146/annurev.iy.11.040193.000541. [DOI] [PubMed] [Google Scholar]
- 105.Lehrer R I, Szklarek D, Ganz T, Selsted M E. Correlation of binding of rabbit granulocyte peptides to Candida albicans with candidacidal activity. Infect Immun. 1985;49:207–211. doi: 10.1128/iai.49.1.207-211.1985. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Lehrer R I, Szklarek D, Ganz T, Selsted M E. Synergistic activity of rabbit granulocyte peptides against Candida albicans. Infect Immun. 1986;52:902–904. doi: 10.1128/iai.52.3.902-904.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Levitz S M, Selsted M E, Ganz T, Lehrer R I, Diamond R D. In vitro killing of spores and hyphae of Aspergillus fumigatus and Rhizopus oryzae by rabbit neutrophil cationic peptides and bronchoalveolar macrophages. J Infect Dis. 1986;154:483–489. doi: 10.1093/infdis/154.3.483. [DOI] [PubMed] [Google Scholar]
- 108.Levy O, Weiss J, Zarember K, Ooi C E, Elsbach P. Antibacterial 15-kDa isoforms (p15s) are members of a novel family of leukocyte proteins. J Biol Chem. 1993;268:6058–6063. [PubMed] [Google Scholar]
- 109.Lim E, Wong P, Fadem M, Motchinik P, Bakalinsky M, Little R. Program and abstracts of the 36th Interscience Conference on Antimicrobial Agents and Chemotherapy. Washington, D.C: American Society for Microbiology; 1996. Fungicidal activity derived from bactericidal/permeability-increasing protein (BPI), abstr. F185; p. 132. [Google Scholar]
- 110.Lim Y, Suh J-W, Kim S, Hyun B, Kim C, Lee C H. Cepacidine A, a novel antifungal antibiotic produced by Pseudomonas cepacia. II. Physico-chemical properties and structure elucidation. J Antibiot. 1994;47:1406–1416. doi: 10.7164/antibiotics.47.1406. [DOI] [PubMed] [Google Scholar]
- 111.Ludtke S, He K, Huang H. Membrane thinning by magainin 2. Proc Natl Acad Sci USA. 1995;34:16764–16769. doi: 10.1021/bi00051a026. [DOI] [PubMed] [Google Scholar]
- 112.Magoni M E, Aumelas A, Chamet P, Roumestand C, Chiche L, Despaux E, Grassy G, Calas B, Chavanieu A. Change in membrane permeability induced by protegrin 1: implication of disulfide bridges for pore formation. FEBS Lett. 1996;383:93–98. doi: 10.1016/0014-5793(96)00236-0. [DOI] [PubMed] [Google Scholar]
- 113.Martinez-Suarez J V, Rodriguez-Tudela J L. In vitro activities of semisynthetic pneumocandin L-733,560 against fluconazole-resistant and -susceptible Candida albicans isolates. Antimicrob Agents Chemother. 1996;40:1277–1279. doi: 10.1128/aac.40.5.1277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.McCarthy P, Newman D J, Nisbet L J, Kingsbury W D. Relative rates of transport of peptidyl drugs by Candida albicans. Antimicrob Agents Chemother. 1985;28:494–499. doi: 10.1128/aac.28.4.494. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.McCarthy P J, Troke P F, Gull K. Mechanism of action of nikkomycin and the peptide transport system of Candida albicans. J Gen Microbiol. 1985;131:775–780. doi: 10.1099/00221287-131-4-775. [DOI] [PubMed] [Google Scholar]
- 116.Merck & Co. May 21, 1996. Press release, business wire. Data on file. Merck & Co., White House Station, N.J.
- 117.Mhammedi A, Peypoux F, Besson F, Michel G. Bacillomycin F, a new antibiotic of iturin group. Isolation and characterization. J Antibiot. 1982;35:306–311. doi: 10.7164/antibiotics.35.306. [DOI] [PubMed] [Google Scholar]
- 118.Michaut L, Fehlbaum P, Moniatte M, Van Dorsselaer A, Rechart J-M, Bulet P. Determination of the disulfide array of the first inducible antifungal peptide from insects: drosomycin from Drosophila melanogaster. FEBS Lett. 1996;395:6–10. doi: 10.1016/0014-5793(96)00992-1. [DOI] [PubMed] [Google Scholar]
- 119.Mizoguchi J, Saito T, Mizuno K, Hayano K. On the mode of action of a new antifungal antibiotic, aculeacin A: inhibition of cell wall synthesis in Saccharomyces cerevisiae. J Antibiot. 1977;30:308–313. doi: 10.7164/antibiotics.30.308. [DOI] [PubMed] [Google Scholar]
- 120.Mizuno K, Yagi A, Satoi S, Takada M, Hayashi M, Asano K, Matsuda T. Studies on aculeacin. I. Isolation and characterization of aculeacin A. J Antibiot. 1977;30:297–302. doi: 10.7164/antibiotics.30.297. [DOI] [PubMed] [Google Scholar]
- 121.Moneton P, Sarthow P, Le Gaffic F. Transport and hydrolysis of peptides in Saccharomyces cerevisiae. J Gen Microbiol. 1986;132:2147–2153. doi: 10.1099/00221287-132-8-2147. [DOI] [PubMed] [Google Scholar]
- 122.Moore A J, Devine D A, Bibby M C. Preliminary experimental anticancer activity of cecropins. Pept Res. 1994;7:265–269. [PubMed] [Google Scholar]
- 123.Mor A, Amiche M, Nicolas P. Structure, synthesis, and activity of dermaseptin b, a novel vertebrate defensive peptide from frog skin: relationship to adenoregulin. Biochemistry. 1994;33:6642–6650. doi: 10.1021/bi00187a034. [DOI] [PubMed] [Google Scholar]
- 124.Mor A, Hani K, Nicolas P. The vertebrate peptide antibiotics dermaseptins have overlapping structural features but target specific organisms. J Biol Chem. 1994;269:31635–31641. [PubMed] [Google Scholar]
- 125.Mor A, Nguyen V H, Delfour A, Migliore-Samour D, Nicolas P. Isolation, amino acid sequence of dermaseptin, a novel antimicrobial peptide of amphibian skin. Biochemistry. 1991;30:8824–8830. doi: 10.1021/bi00100a014. [DOI] [PubMed] [Google Scholar]
- 126.Mori Y, Suzuki M, Fukushima K, Arai T. Structure of leucinostatin B, an uncoupler on mitochondria. J Antibiot. 1983;36:1084–1086. doi: 10.7164/antibiotics.36.1084. [DOI] [PubMed] [Google Scholar]
- 127.Mukhopadhyay T, Ganguli B N. Mulundocandin, a new lipopeptide antibiotic. II. Structure elucidation. J Antibiot. 1986;40:281–289. doi: 10.7164/antibiotics.40.281. [DOI] [PubMed] [Google Scholar]
- 128.Mukhopadhyay T, Roy T K, Bhat R G, Sawant S N, Blumbach J, Ganguli B N, Fehlhaber H W, Kogler H. Deoxymulundocandin—a new echinocandin type antifungal antibiotic. J Antibiot. 1992;45:618–623. doi: 10.7164/antibiotics.45.618. [DOI] [PubMed] [Google Scholar]
- 129.Nagiec M N, Nagiec E E, Baltisburger J A, Well G R, Lester R L, Dickson R L. Sphingolipid synthesis as a target for antifungal drugs. J Biol Chem. 1997;272:9809–9817. doi: 10.1074/jbc.272.15.9809. [DOI] [PubMed] [Google Scholar]
- 130.Oita S, Horita M, Yanagi S O. Purification and properties of a new chitin-binding antifungal CB-1 from Bacillus licheniformis M-4. Biosci Biotech Biochem. 1996;60:481–483. doi: 10.1271/bbb.60.481. [DOI] [PubMed] [Google Scholar]
- 131.Osborn R W, De Samblax G W, Thevissen K, Goderis I, Torrekens S, Van Leuven F, Attenborough S, Rees S B, Broekaert W F. Isolation and characterization of plant defensins from seeds of Asteraceae, Fabaceae, Hippocastanceae, and Saxifragaceae. FEBS Lett. 1995;368:257–262. doi: 10.1016/0014-5793(95)00666-w. [DOI] [PubMed] [Google Scholar]
- 132.Panday V B, Devi S. Biologically active cyclopeptide alkaloids from Rhamnaceae plants. Planta Med. 1990;56:649–650. [Google Scholar]
- 133.Patterson-Delafield J, Szklarek D, Martinez R J, Lehrer R I. Microbicidal cationic proteins of rabbit alveolar macrophages: amino acid composition and functional attributes. Infect Immun. 1981;31:723–731. doi: 10.1128/iai.31.2.723-731.1981. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Perez P, Varona R, Garcia-Acha I, Duran A. Effect of papulacandin B and aculeacin A on beta-(1,3) glucan synthase from Geotrichum latis. FEBS Lett. 1981;129:249–252. [Google Scholar]
- 135.Pergament I, Carmelli S. Schizotrin A: novel antimicrobial cyclic peptide from a cyanobacterium. Tetrahedron Lett. 1994;35:8473–8476. [Google Scholar]
- 136.Peypoux F, Guinand M, Michel G, Delcambre L, Das B C, Varenne P, Lederer E. Isoelement de l’acide 3-amino 12-mèthyl tétradécanöique à partir de l’iturine, antibiotique de Bacillus subtilis. Tetrahedron. 1973;29:3455–3459. [Google Scholar]
- 137.Pfaller M, Gordee R, Gerarden T, Yu M, Wenzel R. Fungicidal activity of cilofungin ( LY121019) alone and in combination with anticapsin or other antifungal agents. Eur J Clin Microbiol Infect Dis. 1989;8:564–567. doi: 10.1007/BF01967483. [DOI] [PubMed] [Google Scholar]
- 138.Pfaller M A, Meisser S A, Coffman S. In vitro susceptibilities of clinical yeast isolates to a new echinocandin derivative, LY303366, and other antifungal agents. Antimicrob Agents Chemother. 1997;41:763–766. doi: 10.1128/aac.41.4.763. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Pouny Y, Rapaport D, Mor A, Nicolas P, Shai Y. Interaction of antimicrobial dermaseptin and its fluorescently labeled analogues with phospholipid membranes. Biochemistry. 1992;31:12416–12423. doi: 10.1021/bi00164a017. [DOI] [PubMed] [Google Scholar]
- 140.Radics L M, Katjar-Perady M, Casinovi C G, Rossi C, Ricci M, Tuttobelo L. Leucinostatins H and K, two novel peptide antibiotics with tertiary amino-oxide terminal group from Paecilomyces marquandii. Isolation, structure and biological activity. J Antibiot. 1987;40:714–716. doi: 10.7164/antibiotics.40.714. [DOI] [PubMed] [Google Scholar]
- 141.Reed W A, Elzer P H, Enright F M, Jaynes J M, Morrey J D, White K L. Interleukin 2 promoter/enhancer controlled expression of a synthetic cecropin-class lytic peptide in transgenic mice and subsequent resistance to Brucella abortus. Transgenic Res. 1997;6:337–347. doi: 10.1023/a:1018423015014. [DOI] [PubMed] [Google Scholar]
- 142.Reidl H H, Takemoto J Y. Mechanism of action of bacterial phytotoxin, syringomycin. Simultaneous measurement of early responses in yeasts and maize. Biochim Biophys Acta. 1987;898:56–59. [Google Scholar]
- 143.Reiter B. The biological significance of lactoferrin. Int J Tissue React. 1983;5:87–96. [PubMed] [Google Scholar]
- 144.Roberts W K, Selitrennikoff C P. Zeamatin, an antifungal protein from maize with membrane-permeabilizing activity. J Gen Microbiol. 1990;136:1771–1778. [Google Scholar]
- 145.Rossi C, Tuttobello L, Ricci M, Casinovi C G, Radics L. Leucinostatin D, a novel peptide from Paecilomyces marquandii. J Antibiot. 1987;40:130–132. doi: 10.7164/antibiotics.40.130. [DOI] [PubMed] [Google Scholar]
- 146.Rouse M S, Tallan B M, Steckelberg J M, Henry N K, Wilson W R. Efficacy of cilofungin therapy administered by continuous intravenous infusion for experimental disseminated candidiasis in rabbits. Antimicrob Agents Chemother. 1992;36:56–58. doi: 10.1128/aac.36.1.56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Roy K, Mukhopadyay T, Reddy G C S, Desikan K R, Ganguli B N. Mulundocandin, a new lipopeptide antibiotic. I. Taxonomy, fermentation, isolation, and characterization. J Antibiot. 1986;40:275–280. doi: 10.7164/antibiotics.40.275. [DOI] [PubMed] [Google Scholar]
- 148.Sato M, Beppu T, Arima K. Properties and structure of a peptide antibiotic no. 1907. Agric Biol Chem. 1980;44:3037–3040. [Google Scholar]
- 149.Satoi S, Yagi A, Asano K, Mizuno K, Watanabe T. Studies of aculeacin. II. Isolation and characterization of aculeacins B, C, D, E, F, and G. J Antibiot. 1977;30:303–307. doi: 10.7164/antibiotics.30.303. [DOI] [PubMed] [Google Scholar]
- 150.Sawistowska-Schroder E T, Kerridge D, Perry H. Echinocandin inhibition of (1,3)-beta-d-glucan synthase from Candida albicans. FEBS Lett. 1984;173:134–138. doi: 10.1016/0014-5793(84)81032-7. [DOI] [PubMed] [Google Scholar]
- 151.Schmatz D M, Powles M A, McFadden D C, Pittarelli L, Balkovec J, Hammond M, Zambias R, Liberator P, Anderson J. Antipneumocystis activity of water-solubilized lipopeptide. Antimicrob Agents Chemother. 1992;36:1964–1970. doi: 10.1128/aac.36.9.1964. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Schmatz D M, Romancheck M A, Pittarelli L A, Schwartz R E, Fromtling R. Treatment of Pneumocystis carinii pneumonia with 1,3-β-glucan synthesis inhibitors. Proc Natl Acad Sci USA. 1990;87:5950–5954. doi: 10.1073/pnas.87.15.5950. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Segal G P, Lehrer R I, Selsted M E. In vitro effect of phagocyte cationic peptides on Coccidioides immitis. J Infect Dis. 1985;151:890–894. doi: 10.1093/infdis/151.5.890. [DOI] [PubMed] [Google Scholar]
- 154.Segre A, Bachmann R C, Ballio A, Bossa F, Grgurina I, Iacobellis N S, Marino G, Pucci P, Simmaco M, Takemoto J Y. The structure of syringomycins A1, E, and G. FEBS Lett. 1989;255:27–31. doi: 10.1016/0014-5793(89)81054-3. [DOI] [PubMed] [Google Scholar]
- 155.Selsted M, Harwig S. Purification, primary structure, and antimicrobial activities of a guinea pig neutrophil defensin. Infect Immun. 1987;55:2281–2286. doi: 10.1128/iai.55.9.2281-2286.1987. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Selsted M, Harwig S, Ganz T, Schilling J, Lehrer R. Primary structures of three human neutrophil defensins. J Clin Invest. 1985;76:1436–1439. doi: 10.1172/JCI112121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Selsted M, Szklarek D, Ganz T, Lehrer R. Activity of rabbit leukocyte peptides against Candida albicans. Infect Immun. 1985;49:202–206. doi: 10.1128/iai.49.1.202-206.1985. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Shai Y. Molecular recognition between membrane-spanning polypeptides. TIBS. 1995;20:460–464. doi: 10.1016/s0968-0004(00)89101-x. [DOI] [PubMed] [Google Scholar]
- 159.Sorensen K N, Kim K-H, Takemoto J Y. In vitro antifungal and fungicidal activities and erythrocyte toxicities of cyclic lipodepsinonapeptides produced by Pseudomonas syringae pv. syringae. Antimicrob Agents Chemother. 1996;40:2710–2713. doi: 10.1128/aac.40.12.2710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Sorensen K N, Wangstrom A A, Allen S D, Takemoto J Y. Efficacy of syringomycin E in a murine model of vaginal candidiasis. J Antibiot. 1998;51:743–749. doi: 10.7164/antibiotics.51.743. [DOI] [PubMed] [Google Scholar]
- 161.Spitzer E D, Travis S J, Kobayashi G S. Comparitive in vitro activity of LY121019 and amphotericin B against isolates of Candida species. Eur J Clin Microbiol Infect Dis. 1988;7:80–81. doi: 10.1007/BF01962183. [DOI] [PubMed] [Google Scholar]
- 162.Steiner H, Hultmark D, Engstrom A, Bennich H, Boman H G. Sequence and specificity of two antibacterial proteins involved in insect immunity. Nature. 1981;292:246–248. doi: 10.1038/292246a0. [DOI] [PubMed] [Google Scholar]
- 163.Storici P, Del Sal G, Schneider C, Romeo D. cDNA sequence of an antibiotic dodecapeptide from neutrophils. FEBS Lett. 1992;314:187–190. doi: 10.1016/0014-5793(92)80971-i. [DOI] [PubMed] [Google Scholar]
- 164.Suzuki S, Isono K, Nagatsu J, Mizutani T, Kawashima Y, Mizuno T. A new antibiotic, polyoxin A. J Antibiot. 1965;18:131. [PubMed] [Google Scholar]
- 165.Taft C S, Stark T, Selitrennikoff C P. Cilofungin ( LY121019) inhibits Candida albicans (1,3)-β-d-glucan synthase activity. Antimicrob Agents Chemother. 1988;32:1901–1903. doi: 10.1128/aac.32.12.1901. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Tailor R, Acland D, Attenborough S, Cammue B P A, Evans I J, Osborn R W, Ray J, Rees S B, Broekaert W F. A novel family of small cysteine-rich antimicrobial peptides from seed in Impatiens balsamina. J Biol Chem. 1997;272:24480–24487. doi: 10.1074/jbc.272.39.24480. [DOI] [PubMed] [Google Scholar]
- 167.Takemoto J Y, Giannini J L, Vassey T, Briskin D P. Syringomycin effects on plasma membrane Ca2+ transport. In: Graniti A, Durbin R D, Ballio A, editors. Phytotoxins and plant pathogenesis. Berlin, Germany: Springer-Verlag; 1989. pp. 167–175. [Google Scholar]
- 168.Takemoto J Y, Yaxin Y, Stock S D, Miyakawa T. Yeast genes involved in growth inhibition by Pseudomonas syringae pv. syringae syringomycin family lipodepsipeptides. FEMS Microbiol Lett. 1993;114:339–342. doi: 10.1111/j.1574-6968.1993.tb06595.x. [DOI] [PubMed] [Google Scholar]
- 169.Takemoto J Y, Zhang L, Taguchi N, Tachikawa T, Miyakawa T. Mechanism of action of the syringomycin: a resistant mutant of Saccharomyces cerevisiae reveals an involvement of Ca2+ transport. J Gen Microbiol. 1991;137:653–659. [Google Scholar]
- 170.Takesako K, Ikai K, Haruna F, Endo M, Shimanaka K, Sono E, Nakamura T, Kato I, Yamaguchi J. Aureobasidins, new antifungal antibiotics: taxonomy, fermentation, isolation, and properties. J Antibiot. 1991;44:919–924. doi: 10.7164/antibiotics.44.919. [DOI] [PubMed] [Google Scholar]
- 171.Takesako K, Kuroda H, Inoue T, Haruna F, Yosikawa Y, Kato I, Uchida K, Hiratani T, Yamaguchi H. Biological properties of aureobasidin A, a cyclic depsipeptide antifungal antibiotic. J Antibiot. 1993;46:1414–1420. doi: 10.7164/antibiotics.46.1414. [DOI] [PubMed] [Google Scholar]
- 172.Tariq V N, Develin P L. Sensitivity of fungi to nikkomycin Z. Fungal Genet Biol. 1996;20:4–11. doi: 10.1006/fgbi.1996.0003. [DOI] [PubMed] [Google Scholar]
- 173.Terras F R G, Goderis I J, Van Leuven F, Vanderleyden J, Cammue B P A. Analysis of two novel classes of antifungal proteins from radish (Raphanus sativus L.) seeds. J Biol Chem. 1992;267:15301–15309. [PubMed] [Google Scholar]
- 174.Thevissen K, Ghazi A, De Samblanx G W, Brownlee C, Osborn R W, Broekaert W F. Fungal membrane responses induced by plant defensins and thionins. J Biol Chem. 1996;271:15018–15025. doi: 10.1074/jbc.271.25.15018. [DOI] [PubMed] [Google Scholar]
- 175.Thimon L, Peypoux F, Maget-Dana R, Michel G. Surface-active properties of antifungal lipopeptides produced by Bacillus subtilis. J Am Oil Chem Soc. 1992;69:92–93. [Google Scholar]
- 176.Tomita M, Bellamy W, Takase M, Tamauchi K, Wakabayashi H, Kawase K. Potent antimicrobial peptides generated by pepsin digest of lactoferrin. J Dairy Sci. 1991;74:4137–4142. doi: 10.3168/jds.S0022-0302(91)78608-6. [DOI] [PubMed] [Google Scholar]
- 177.Traber R, Keller-Juslen C, Loosli H R, Huhn M, Von Wartburg A. Cyclopeptide-antibiotika aus Aspergillus arten. Structur der echinocandine C und D. Helv Chim Acta. 1979;62:1252–1267. [Google Scholar]
- 178.Turner W W, Current W L. Echinocandin antifungal agents. In: Strohl W R, editor. Biotechnology of antibiotic. 2nd ed. New York, N.Y: Marcel Dekker, Inc.; 1997. pp. 315–334. [Google Scholar]
- 179.Tyler E M, Anatharamaiah G M, Walker D E, Mishra V K, Palgunachan M N, Segrest J P. Molecular basis for prokaryotic specificity of magainin-induced lysis. Biochemistry. 1995;34:4393–4401. doi: 10.1021/bi00013a031. [DOI] [PubMed] [Google Scholar]
- 180.Uzun O, Kocagöz S, Çetinkaya Y, Arikan S, Ünal S. In vitro activity of a new echinocandin, LY303366, compared with those of amphotericin B and fluconazole against clinical yeast isolates. Antimicrob Agents Chemother. 1997;41:1156–1157. doi: 10.1128/aac.41.5.1156. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Vigers A J, Roberts W K, Selitrennikoff C P. A new family of antifungal proteins. Mol Plant-Microbe Interact. 1991;4:315–323. doi: 10.1094/mpmi-4-315. [DOI] [PubMed] [Google Scholar]
- 182.Wade D, Merrifield R B, Boman H G. Effects of cecropin and melittin analogs and hybrids on pro- and eukaryotic cells. In: River J E, Marshall G R, editors. Peptides: chemistry, structure, and biology. Proceedings of the 11th Peptide Symposium. Leiden, The Netherlands: Escom; 1989. pp. 120–121. [Google Scholar]
- 183.Walsh T J, Lee J W, Kelly P, Bacher J, Lecciones J, Thomas V, Lyman C, Coleman D, Gordee R, Pizzo P A. Antifungal effects of the nonlinear pharmacokinetics of cilofungin, a 1,3-β-glucan synthetase inhibitor, during continuous and intermittent intravenous infusions in treatment of experimental disseminated candidiasis. Antimicrob Agents Chemother. 1991;35:1321–1328. doi: 10.1128/aac.35.7.1321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184.Westerhoff H V, Juretic V D, Hendler R W, Zasloff M. Magainins and the disruption of membrane-linked free-energy transduction. Proc Natl Acad Sci USA. 1989;86:6597–6601. doi: 10.1073/pnas.86.17.6597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.White S H, Wimley W C, Selsted M E. Structure, function, and membrane integration of defensins. Curr Opin Struct Biol. 1995;5:521–527. doi: 10.1016/0959-440x(95)80038-7. [DOI] [PubMed] [Google Scholar]
- 186.Yamauchi K, Tomita M, Giehl T J, Ellison R T. Antibacterial activity of lactoferrin and a pepsin-derived lactoferrin peptide fragment. Infect Immun. 1993;61:719–728. doi: 10.1128/iai.61.2.719-728.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Zambias R A, Hammond M L, Heck J V, Bartizal K, Trainor C, Abruzzo G, Schmatz D M, Nollstadt K M. Preparation and structure-activity relationships of simplified analogues of the antifungal agent cilofungin: total synthesis approach. J Med Chem. 1992;35:2843–2855. doi: 10.1021/jm00093a018. [DOI] [PubMed] [Google Scholar]
- 188.Zambias R A, James C, Abruzzo G K, Bartizal K F, Hajdu R, Thompson R, Nollstadt K H, Marrinan J, Balkovec J M. Lipopeptide antifungal agents: amine conjugates of the semi-synthetic pneumocandins L-731,373 and L-733,560. Bioorg Med Chem Lett. 1997;7:2021–2026. [Google Scholar]
- 189.Zanetti M, Del Sal G, Storici P, Schneider C, Romeo D. The cDNA of the neutrophil antibiotic Bac5 predicts a pro-sequence homologous to a cysteine proteinase inhibitor that is common to other neutrophil antibiotics. J Biol Chem. 1993;268:522–526. [PubMed] [Google Scholar]
- 190.Zasloff M. Magainins, a class of antimicrobial peptides from Xenopus skin: isolation, characterization of two active forms and partial cDNA sequence of a precursor. Proc Natl Acad Sci USA. 1987;84:5449–5453. doi: 10.1073/pnas.84.15.5449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Zhanel G C, Karlowsky J A, Harding G A, Balko T V, Zelenitsky S A, Freisen M, Kabani A, Turik M, Hoban D J. In vitro activity of a new semisynthetic echinocandin, LY-303366, against systemic isolates of Candida species, Cryptococcus neoformans, Blastomyces dermatidis, and Aspergillus species. Antimicrob Agents Chemother. 1997;41:863–865. doi: 10.1128/aac.41.4.863. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Zhang L, Takemoto J Y. Effects of Pseudomonas syringae phytotoxin, syringomycin, on plasma membrane fractions of Rhodotorula pilimanae. Phytopathology. 1987;77:297–303. [Google Scholar]