Skip to main content
Antimicrobial Agents and Chemotherapy logoLink to Antimicrobial Agents and Chemotherapy
. 2007 Sep 4;51(11):4214–4216. doi: 10.1128/AAC.01589-06

Effects of Serum on In Vitro Susceptibility Testing of Echinocandins

Zekaver Odabasi 1,*, Victor Paetznick 2, John H Rex 2,, Luis Ostrosky-Zeichner 2
PMCID: PMC2151461  PMID: 17785512

Abstract

The effects of protein binding on the activities of caspofungin, anidulafungin, and micafungin were evaluated against Candida and Aspergillus species. Adding human serum sharply increased the MICs of micafungin and anidulafungin and modestly affected the MIC of caspofungin. The increase in MICs does not appear consistent with the rate of protein binding for the three compounds.


The echinocandins are a new class of lipopeptide antifungal agents that act by inhibiting the synthesis of (1,3)-β-d-glucan. Compounds of this class are relatively highly protein bound, with rates of 96% reported for caspofungin (5), 99.8% for micafungin (15), and ∼99% for anidulafungin (Eraxis package insert; Pfizer). Protein binding may change the in vitro and in vivo activities of antimicrobial agents (17). The available data on the effects of protein binding on the antifungal activities of echinocandins are limited to reports of increased MICs for anidulafungin when tested against Candida albicans (Eraxis package insert; Pfizer), of increased MICs for micafungin when tested against Candida spp. and Aspergillus fumigatus (4, 16), of no effect on caspofungin MICs for one isolate of C. albicans (2), and of a potentiation of the effect of caspofungin versus A. fumigatus (3). As none of these reports have provided comparative data and the total number of isolates evaluated has been small, we now report the effect of 50% human serum on the activities of the echinocandins against a collection of Candida and Aspergillus isolates.

A total of 16 Candida isolates and 8 Aspergillus isolates were tested. The isolates included were C. albicans (two isolates), C. parapsilosis (five isolates), C. krusei (three isolates), C. glabrata (two isolates), C. tropicalis (two isolates), C. lusitaniae (two isolates), A. fumigatus (four isolates), A. flavus (two isolates), A. terreus (one isolate), and A. niger (one isolate). The two quality control isolates specified in the Clinical and Laboratory Standards Institute (CLSI) M27-A2 procedure (10), ATCC 6258 (C. krusei) and ATCC 22019 (C. parapsilosis), were included in each set of the test and the results compared with published control limits (1).

Caspofungin, micafungin, and anidulafungin were supplied by their respective manufacturers. Stock solutions were prepared by dissolving the compounds in dimethyl sulfoxide (anidulafungin) or water (caspofungin and micafungin). Following the principles of CLSI M27-A2, serial dilutions at twice the desired final concentration were prepared in double-strength test medium (RPMI 1640 medium buffered with 0.165 M morpholinepropanesulfonic acid [MOPS] to pH 7.0). Test trays were prepared in advance by dispensing 100 μl of serially diluted drug into 96-well microdilution plates and freezing the plates at −70°C. All three compounds were tested over a 20-fold dilution range from 64 to 0.00012 μg/ml.

The MICs of test drugs with Candida species and Aspergillus species were determined using the broth microdilution variants of CLSI M27-A2 and CLSI M38-A (9), respectively. Testing was performed both in standard medium (RPMI 1640, 0.165 M MOPS, pH 7.0) and in standard medium containing 50% pooled human serum (Sigma, St. Louis, MO). To achieve this, inocula were standardized spectrophotometrically and diluted either with sterile water or with 100% pooled human serum to final concentrations of 1 × 103 to 5 × 103 yeast or 0.4 × 104 to 5 × 104 conidia. Then, 100-μl volumes of these double-strength inocula were added to the 100-μl volumes of serially diluted drug in the microdilution trays. MICs of Candida and Aspergillus species were read after 24 and 48 h at 35°C as MIC-0, i.e., complete growth inhibition, and MIC-2, i.e., at least 50% growth inhibition. For Aspergillus, a minimum effective concentration was also determined microscopically (7) and found to be equivalent to the macroscopically determined MIC-2 reading. The MIC-2 readings are thus used for both fungal genera.

A line listing of the results seen after 24 h is shown in Table 1. The qualitative effects of serum on echinocandin susceptibility were similar for Candida species at 24 and 48 h and using endpoints of MIC-0 and MIC-2 (not shown); thus, the 24-h MIC-2 results are shown as being characteristic of the entire data set. For Aspergillus species, all MIC-0 readings were off-scale, and thus only the MIC-2 readings provide useful information; qualitative effects of serum on echinocandin susceptibility were similar at 24 and 48 h, as seen for Candida spp. A summary of the ratios of the MIC-2 values by time, genus, and drug is shown in Table 2. The addition of 50% serum consistently elevated the observed MICs for anidulafungin and micafungin. For caspofungin, on the other hand, an increased MIC was observed for some (but not all) isolates, and this effect was never as frequent as or of the magnitude of that seen for the other two agents. An analysis of the effect of serum for Candida and Aspergillus species (Table 3) showed that the effect of serum was similar for each drug across the range of tested species.

TABLE 1.

Comparison of MIC-2 values observed for RPMI and 50% serum

Species Isolate code Anidulafungin
Caspofungin
Micafungin
RPMI MIC-2 (μg/ml) 50% serum MIC-2 (μg/ml) Ratioa RPMI MIC-2 (μg/ml) 50% serum MIC-2 (μg/ml) Ratio RPMI MIC-2 (μg/ml) 50% serum MIC-2 (μg/ml) Ratio
C. albicans 0012-025 0.0156 0.25 16 0.25 0.25 1 0.0156 1 64
0021-076 0.5 128 256 0.5 4 8 0.5 64 128
C. glabrata 0022-012 0.0313 1 32 0.5 0.5 1 0.0156 1 64
0049-038 0.0625 2 32 0.5 0.5 1 0.004 1 256
C. krusei 0001-058 0.0313 1 32 0.5 0.5 1 0.0625 2 32
0013-054 0.125 4 32 0.5 2 4 0.125 8 64
ATCC 6258 0.125 2 16 0.5 4 8 0.0625 8 128
C. lusitaniae 0013-002 0.002 1 512 0.25 0.25 1 0.008 1 128
0015-049 0.125 1 8 0.5 1 2 0.0156 4 256
C. parapsilosis 0014-074 0.004 0.25 64 0.25 0.5 2 0.0156 0.5 32
0011-013 0.0313 2 64 0.25 1 4 0.0156 4 256
0026-003 2 128 64 1 8 8 0.5 64 128
0036-030 1 128 128 1 8 8 0.5 128 256
ATCC 22019 0.5 64 128 0.5 8 16 0.5 32 64
C. tropicalis 0022-021 0.0156 2 128 0.5 0.5 1 0.0156 2 128
0050-028 1 128 128 0.5 2 4 0.25 32 128
A. terreusb 0031-087 0.0156 0.25 16 0.5 0.5 1 0.002 0.25 128
A. niger 0001-073 0.001 0.5 512 0.0004 0.25 512 0.002 0.25 128
A. fumigatus 0002-0016 2 128 64 8 128 16 0.0313 2 64
0020-018 0.5 128 256 16 128 8 0.0156 0.5 32
0002-020 0.008 0.5 64 0.5 128 256 0.004 0.5 128
0002-021 0.002 128 64,000 1 128 128 0.001 1 1,024
A. flavus 0001-033 4 128 32 32 128 4 0.125 128 1,024
0001-061 4 128 32 128 128 1 0.125 128 1,024
a

Shown is the ratio of the MIC-2 values with/without serum.

b

Growth rate was insufficient for A. terreus at 24 h; MIC-2 results were detected at 48 h.

TABLE 2.

Ratios of MIC-2 values with and without serum

Genus Time (h) Drug Ratio of MIC-2 in standard medium with/without 50% seruma
0.5 1 2 4 8 16 32 64 128 256 512+
Candida 24 Anidulafungin 1 2 4 3 4 1 1
Caspofungin 6 2 3 4 1
Micafungin 2 4 6 4
48 Anidulafungin 1 5 4 4 1 1
Caspofungin 3 2 3 6 2
Micafungin 2 1 9 2 2
Aspergillus 24 Anidulafungin 1 2 2 1 2
Caspofungin 2 1 1 1 1 1 1
Micafungin 1 1 3 3
48 Anidulafungin 2 4 2
Caspofungin NA
Micafungin NA
a

NA, not applicable; for those drugs, most of the MIC-2 results with and without serum were out of range, and so ratio determinations were not meaningful.

TABLE 3.

Relationship of MIC-2 values with/without serum to Candida and Aspergillus species

Drug Species Ratio of MIC-2 in standard medium with/without 50% serum
0.5 1 2 4 8 16 32 64 128 256 512
Anidulafungin C. albicans 1 1
C. glabrata 2
C. krusei 1 2
C. lusitaniae 1 1
C. parapsilosis 3 2
C. tropicalis 2
A. fumigatus 2 1 1
A. flavus, others 1 2 1
Caspofungin C. albicans 1 1
C. glabrata 2
C. krusei 1 1 1
C. lusitaniae 1 1
C. parapsilosis 1 1 2 1
C. tropicalis 1 1
A. fumigatus 1 1 1 1
A. flavus, others 2 1 1
Micafungin C. albicans 1 1
C. glabrata 1 1
C. krusei 1 1 1
C. lusitaniae 1 1
C. parapsilosis 1 1 1 2
C. tropicalis 2
A. fumigatus 1 1 1 1
A. flavus, others 2 2

In this study, we have extended prior work by testing a broad range of Candida and Aspergillus spp. and by directly comparing the effects of serum on all three echinocandins (2, 3, 4, 16; Eraxis package insert [Pfizer]). The effect of serum was greatest for anidulafungin and micafungin, and this is entirely consistent with the results of prior studies on these two drugs (4, 16; Eraxis package insert [Pfizer]). Prior work with caspofungin against single isolates of C. albicans (2) and A. fumigatus (3) found that serum had no effect against the Candida isolate and enhanced activity against the Aspergillus isolate. These results are consistent with the absence of effect of serum seen for some isolates in our survey, and we did not see any enhancing activity with the addition of serum.

In summary, we have observed that serum strongly affects the observed MICs of anidulafungin and micafungin and modestly affects the MIC of caspofungin. This rank order of effects does not appear consistent with the rank order of protein binding for the three compounds and is thus a further demonstration of the relatively unpredictable nature of the effect of serum on antifungal effect. As related examples, terbinafine (>99% protein bound [13]) has a reduced in vitro potency in the presence of serum that has been correlated with reduced in vivo potency (8, 11, 14). On the other hand, although the potency of itraconazole (99.8% protein bound [6]) does appear affected in vitro or in vivo by the presence of serum, protein binding does not impede the drug's activity (12, 13, 18). For the echinocandin antifungal agents, the impact of these in vitro effects on in vivo activity is not yet understood and is likely not of clinical significance. However, we think that the different serum effects with these three echinocandins may explain the different dosage regimens of these drugs in clinical practice.

Footnotes

Published ahead of print on 4 September 2007.

REFERENCES

  • 1.Barry, A. L., M. A. Pfaller, S. D. Brown, A. Espinel-Ingroff, M. A. Ghannoum, C. Knapp, R. P. Rennie, J. H. Rex, and M. G. Rinaldi. 2000. Quality control limits for broth microdilution susceptibility tests of ten antifungal agents. J. Clin. Microbiol. 38:3457-3459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Bartizal, K., C. J. Gill, G. K. Abruzzo, A. M. Flattery, L. Kong, P. M. Scott, J. G. Smith, C. E. Leighton, A. Bouffard, J. F. Dropinski, and J. Balkovec. 1997. In vitro preclinical evaluation studies with the echinocandin antifungal MK-0991 (L-743,872). Antimicrob. Agents Chemother. 41:2326-2332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Chiller, T., K. Farrokhshad, E. Brummer, and D. A. Stevens. 2000. Influence of human sera on the in vitro activity of the echinocandin caspofungin (MK-0991) against Aspergillus fumigatus. Antimicrob. Agents Chemother. 44:3302-3305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Ernst, E. J., E. E. Roling, C. R. Petzold, D. J. Keele, and M. E. Klepser. 2002. In vitro activity of micafungin (FK-463) against Candida spp.: microdilution, time-kill, and postantifungal-effect studies. Antimicrob. Agents Chemother. 46:3846-3853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Hajdu, R., R. Thompson, J. G. Sundelof, B. A. Pelak, F. A. Bouffard, J. F. Dropinski, and H. Kropp. 1997. Preliminary animal pharmacokinetics of the parenteral antifungal agent MK-0991 (L-743,872). Antimicrob. Agents Chemother. 41:2339-2344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Heykants, J., M. Michiels, W. Meuldermans, J. Monbaliu, K. Lavrijsen, A. Van Peer, J. C. Levron, R. Woestenborghs, and G. Cauwenbergh. 1987. The pharmacokinetics of itraconazole in animals and man: an overview, p. 1-29. In R. A. Fromtling (ed.), Recent trends in the discovery, development and evaluation of antifungal agents. J. R. Prous Science Publishers, Barcelona, Spain.
  • 7.Kurtz, M. B., I. B. Heath, J. Marrinan, S. Dreikorn, J. Onishi, and C. Douglas. 1994. Morphological effects of lipopeptides against Aspergillus fumigatus correlate with activities against (1,3)-β-d-glucan synthase. Antimicrob. Agents Chemother. 38:1480-1489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Machard, B., P. Misslin, and M. Lemaire. 1989. Influence of plasma protein binding on the brain uptake of an antifungal agent, terbinafine, in rats. J. Pharm. Pharmacol. 41:700-704. [DOI] [PubMed] [Google Scholar]
  • 9.National Committee for Clinical Laboratory Standards. 2002. Reference method for broth dilution antifungal susceptibility testing of filamentous fungi; approved standard NCCLS document M38-A. National Committee for Clinical Laboratory Standards, Wayne, PA.
  • 10.National Committee for Clinical Laboratory Standards. 2002. Reference method for broth dilution antifungal susceptibility testing of yeasts; approved standard NCCLS document M27-A2. National Committee for Clinical Laboratory Standards, Wayne, PA.
  • 11.Ryder, N. S., and I. Frank. 1992. Interaction of terbinafine with human serum and serum proteins. J. Med. Vet. Mycol. 30:451-460. [DOI] [PubMed] [Google Scholar]
  • 12.Schafer-Korting, M., H. C. Korting, F. Amann, R. Peuser, and A. Lukacs. 1991. Influence of albumin on itraconazole and ketoconazole antifungal activity: results of a dynamic in vitro study. Antimicrob. Agents Chemother. 35:2053-2056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Schafer-Korting, M., H. C. Korting, W. Rittler, and W. Obermuller. 1995. Influence of serum protein binding on the in vitro activity of anti-fungal agents. Infection 23:292-297. [DOI] [PubMed] [Google Scholar]
  • 14.Schmitt, H. J., J. Andrade, F. Edwards, Y. Niki, E. Bernard, and D. Armstrong. 1990. Inactivity of terbinafine in a rat model of pulmonary aspergillosis. Eur. J. Clin. Microbiol. Infect. Dis. 9:832-835. [DOI] [PubMed] [Google Scholar]
  • 15.Suzuki, S., M. Terakawa, F. Yokobayashi, F. Fujiwara, and T. Hata. 1998. Pharmacokinetics of FK463, a novel water-soluble echinocandin-like lipopeptide, in animals, abstr. F-144, p. 269. Abstr. 38th Intersci. Conf. Antimicrob. Agents Chemother., San Diego, CA.
  • 16.Tawara, S., F. Ikeda, K. Maki, Y. Morishita, K. Otomo, N. Teratani, T. Goto, M. Tomishima, H. Ohki, A. Yamada, K. Kawabata, H. Takasugi, K. Sakane, H. Tanaka, F. Matsumo, and S. Kuwahara. 2000. In vitro activities of a new lipopeptide antifungal agent, FK463, against a variety of clinically important fungi. Antimicrob. Agents Chemother. 44:57-62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Wise, R. 1985. The relevance of pharmacokinetics to in-vitro models: protein binding—does it matter? J. Antimicrob. Chemother. 15(Suppl. A):77-83. [DOI] [PubMed] [Google Scholar]
  • 18.Wright, K. A., J. R. Perfect, and W. Ritter. 1990. The pharmacokinetics of BAY R3783 and its efficacy in the treatment of experimental cryptococcal meningitis. J. Antimicrob. Chemother. 26:387-397. [DOI] [PubMed] [Google Scholar]

Articles from Antimicrobial Agents and Chemotherapy are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES