Skip directly to site content Skip directly to page options Skip directly to A-Z link Skip directly to A-Z link Skip directly to A-Z link
Volume 15, Number 7—July 2009
Research

Frequency and Evolution of Azole Resistance in Aspergillus fumigatus Associated with Treatment Failure1

Author affiliations: Regional Mycology Laboratory, Manchester, UK (S.J. Howard, D.W. Denning); University of Manchester, Manchester Academic Health Science Centre, Manchester (S.J. Howard, M.J. Anderson, A. Albarrag, D.W. Denning); University Medical Centre, Ljubljana, Slovenia (D. Cerar); Imperial College, London, UK (M.C. Fisher); Universidade Federal do Rio Grande do Sul, Porto Alegre, Brazil (A.C. Pasqualotto); Hôpital Maisonneuve-Rosemont, Montreal, Québec, Canada (M. Laverdiere); Statens Serum Institut, Copenhagen, Denmark (M.C. Ardendrup); Public Health Research Institute, Newark, NJ, USA (D.S. Perlin)

Cite This Article

Abstract

Azoles are the mainstay of oral therapy for aspergillosis. Azole resistance in Aspergillus has been reported infrequently. The first resistant isolate was detected in 1999 in Manchester, UK. In a clinical collection of 519 A. fumigatus isolates, the frequency of itraconazole resistance was 5%, a significant increase since 2004 (p<0.001). Of the 34 itraconazole-resistant isolates we studied, 65% (22) were cross-resistant to voriconazole and 74% (25) were cross-resistant to posaconazole. Thirteen of 14 evaluable patients in our study had prior azole exposure; 8 infections failed therapy (progressed), and 5 failed to improve (remained stable). Eighteen amino acid alterations were found in the target enzyme, Cyp51A, 4 of which were novel. A population genetic analysis of microsatellites showed the existence of resistant mutants that evolved from originally susceptible strains, different cyp51A mutations in the same strain, and microalterations in microsatellite repeat number. Azole resistance in A. fumigatus is an emerging problem and may develop during azole therapy.

Invasive aspergillosis in immunosuppressed patients is difficult to diagnose, is problematic to treat, and results in a high mortality rate (1). Chronic and allergic pulmonary and sinus aspergillosis are increasingly recognized in numerous clinical settings. Treatment with itraconazole, voriconazole, and, recently, posaconazole is the backbone of therapy for these conditions because azoles are the only licensed class of oral drugs for treatment of aspergillosis (2,3). Amphotericin B and caspofungin are licensed intravenous agents for invasive aspergillosis but have limited utility for chronic and allergic aspergillosis.

Itraconazole resistance in Aspergillus spp. was first reported in 1997 in 3 clinical isolates obtained from California in the late 1980s (4); since then, only a few clinical cases have been published (59). The emergence of itraconazole resistance alone is of concern, but widespread azole cross-resistance would be devastating because oral treatment would not be effective.

The primary mechanism of resistance described for A. fumigatus clinical isolates is mutation in the target protein. The cyp51A gene encodes the target of azoles, lanosterol 14α-demethylase, and this enzyme catalyzes a step in the biosynthetic pathway of ergosterol (an essential cell membrane component of filamentous fungi). Mutations in the open reading frame of the cyp51A gene can result in structural alterations to the enzyme, which in turn may inhibit binding of drugs. Mutational hotspots confirmed to cause resistance have been characterized in the gene at codons 54 (6,1013), 220 (6,14,15), and 98 (1618). Other mutations in the cyp51A gene have been reported, and additional resistance mechanisms have been postulated (11,19,20). The environmental or antifungal pressures driving azole resistance are unclear because few clinical azole-resistant Aspergillus strains have been studied in any detail; many reports simply describe individual patient cases. In this study, we investigated the frequency of A. fumigatus itraconazole resistance in a referral laboratory collection, defined the resulting azole cross-resistance pattern, identified mutations in the cyp51A gene, and investigated any epidemiologic links between resistant isolates.

Materials and Methods

Isolates

Isolates deposited in the Regional Mycology Laboratory Manchester (RMLM) culture collection (between 1992 and 2007) were identified as A. fumigatus by macro- and micromorphologic characteristics. All isolates were screened for growth at 50oC, thus confirming A. fumigatus and excluding A. lentulus. Aspergilli were subcultured onto Sabouraud glucose agar (Oxoid, Basingstoke, UK) for 48 h at 37°C. Thirty-four azole-resistant and 5 susceptible isolates from 17 patients were studied from the RMLM collection (prefixed F); 36 isolates were respiratory specimens, 1 was cerebral, and 2 were from unknown sites. In addition, 18 azole-resistant isolates from a single aspergilloma case-patient (prefixed A, patient 3) collected at autopsy were also investigated.

Patients

Pertinent details from patients were extracted from the clinical records. All but 6 were under the care of 1 investigator (D.W.D.). Information was collected on underlying disease(s), type of aspergillosis, antifungal treatment, azole plasma levels, and characteristics of therapeutic failure.

Susceptibility Testing

Susceptibilities were determined by a modified European Committee on Antimicrobial Susceptibility Testing (EUCAST) method (21). The modification was a lower final inoculum concentration (0.5 × 105 as opposed to 1–2.5 × 105 CFU/mL). Isolates were tested at a final drug concentration range of 8, 4, 2, 1, 0.5, 0.25, 0.125, 0.06, 0.03, 0.015 mg/L against itraconazole (Research Diagnostics Inc, Concord, MA, USA), voriconazole (Pfizer Ltd, Sandwich, UK), posaconazole (Schering-Plough, Kenilworth, NJ, USA), and amphotericin B (Sigma, Poole, UK). RPMI-1640 (Sigma) was supplemented to 2% glucose (Sigma). Inocula were prepared in phosphate-buffered saline with 0.05% Tween 80 (Sigma); Aspergillus spores were counted on a hemacytometer and adjusted to a final concentration of 5 × 104 CFU/mL. Inocula were loaded into flat-bottomed microtiter plates (Costar Corning, Lowell, MA, USA) and incubated at 37oC for 48 h. A no-growth end point was determined by eye. MIC testing was performed on RMLM isolates in triplicate, and a consensus mean was derived (median or mode). Susceptibilities of the aspergilloma isolates were determined once, except for 6 that were tested 3 times. Values of >8 mg/L were classed as 16.

Clinical or epidemiologic breakpoints/cutoffs have not been declared by the Clinical and Laboratory Standards Institute (CLSI) or EUCAST for azoles and Aspergillus spp. However, proposed epidemiologic cutoff values have been mooted for the latter (22), and we have recently proposed clinical breakpoints (23). Cutoffs used in this study were itraconazole and voriconazole >2 mg/L and posaconazole >0.5 mg/L (we have not defined an intermediate zone of susceptibility).

Sequencing

DNA was extracted by using commercially available kits (FastDNA Kit, Q-biogene, Cambridge, UK; Ultraclean Soil DNA Isolation Kit, MO BIO Laboratories Inc., Cambridge; and DNeasy plant tissue kit, QIAGEN, Crawley, UK). The entire coding region of the cyp51A gene was amplified as previously described (7), except 3 mmol/L MgCl2 was used and both strands were sequenced using 8 primers (7). Twelve of the aspergilloma (A) isolates were sequenced with only 1 primer, covering the region of interest in this case. Sequences were aligned against the sequence from an azole-susceptible strain (GenBank accession no. AF338659), and mismatches were identified by using AlignX (VectorNTI; Invitrogen, Paisley, UK). Mutations were confirmed by repeating the PCR and sequencing both strands by using the closest 2 primers. Isolates with an alteration in the cyp51A gene at codon 98 were also investigated for promoter modifications by sequencing this region (17). GenBank accession numbers for the cyp51A sequences determined in this study are EU807919−EU807922 and FJ548859−FJ548890.

Microsatellite Typing

Six microsatellite loci (3A, 3B, 3C, 4A, 4B, 4C) were amplified as previously described (24). Initially some amplicons were sequenced, whereas later ones were sized by using capillary electrophoresis on an ABI PRISM 3130×l Genetic Analyzer (Applied Biosystems, Warrington, UK). Electrophoresis data were analyzed by using Peak Scanner Software version 1.0 (Applied Biosystems); amplicon sizes were adjusted by using a correction factor derived from sequenced alleles to determine the actual sizes of alleles (25). Concatenated multilocus microsatellite genotypes were created for each isolate and used to generate allele-sharing genetic distance matrices, DAS. Here, DAS = 1 – (the total number of shared alleles at all loci / n), where n is the total number of loci compared (26). Subsequently, phylogenetic comparisons using 5 of the loci (not 3B) were performed with the software PAUP* 4.0 (www.paup.csit.fsu.edu) by using the neighbor-joining algorithm with the minimum-evolution option active. The strength of support for relationships was assessed by using 1,000 bootstrap resamples of the dataset.

Results

Susceptibility

The susceptibility of 519 A. fumigatus RMLM culture collection isolates was determined. All isolates were tested for susceptibility against itraconazole and amphotericin B; 456 and 118 isolates were also tested against voriconazole and posaconazole, respectively. Subsequently, all itraconazole-resistant isolates were tested against voriconazole and posaconazole. Geometric means, ranges, MIC50 (median MIC), and MIC90 (90% of the isolates tested had a MIC at or below this level) values are shown in Table 1. Amphotericin B susceptibility was retained in the 34 itraconazole-resistant isolates tested. Of these, 65% (22) were cross-resistant to voriconazole and 74% (25) were cross-resistant to posaconazole. We did not identify any isolates that were resistant to voriconazole or posaconazole while remaining susceptible to itraconazole.

Figure 1

Thumbnail of Azole resistance in clinical Aspergillus fumigatus isolates received in the Regional Mycology Laboratory Manchester, UK, 1997–2007. Overall azole resistance for each year is shown above each column as a percentage. Data do not include sequential isolates from the same patient.

Figure 1. Azole resistance in clinical Aspergillus fumigatus isolates received in the Regional Mycology Laboratory Manchester, UK, 1997–2007. Overall azole resistance for each year is shown above each column as a percentage. Data...

Five percent of 400 isolates were resistant to itraconazole (when duplicate isolates from the same patient with similar susceptibility profiles were removed from the analysis). The overall frequency of itraconazole resistance in this collection (with repeat specimens included) was 7% (n = 519). The first case of azole resistance in this collection was seen in 1999. The frequency of resistance since 2004 (8%) has increased significantly (Fisher exact test, p<0.001), compared with the period prior to 2004 (Figure 1).

Azole Exposure in Patients with Azole-Resistant Isolates and Response to Therapy

Of the 17 patients identified for respective review, limited data were available for 3 patients. Of the remaining 14 patients with antifungal data (Table 2), azole exposure of 1–30 months before the identification of the first resistant isolate was evident for all except patient 7. Thirteen patients received itraconazole as initial therapy, and 12 of these were evaluable. Infections failed to respond to therapy in 7 itraconazole-treated patients (i.e., their disease progressed), although 3 patients appeared to improve before their conditions began to deteriorate. Infections failed to improve with azole therapy in 5 patients (i.e, their disease remained stable), and no patient had a sustained response to therapy. Nine of the 12 patients had at least 1 therapeutic concentration of itraconazole (>5.0 mg/L) documented at steady state during their treatment course (online expanded version of Table 2, available from www.cdc.gov/EID/content/15/7/1068-T2.htm). The infections in patient 1 (treated with voriconazole only for 18 months) failed therapy, and the 1 isolate identified had MICs of >8 mg/L for both itraconazole and voriconazole.

Of the 14 patients with available data, 2 had invasive disease; 9 had chronic disease with >1 aspergillomas; 2 had allergic bronchopulmonary aspergillosis; and 1 had Aspergillus bronchitis. At least 5 of the patients died of progressive infection, despite alternative therapies for some.

Mutations in the cyp51A Gene

A summary of Cyp51A amino acid substitutions and azole cross-resistance patterns identified in 34 resistant isolates from our clinical culture collection is shown in Table 3 and listed by line in the Appendix Table. The sequences of all 5 azole-susceptible isolates examined were identical to that of a previously published cyp51A gene sequence from an azole-susceptible isolate (AF338659). No cyp51A mutations were found in 3 itraconazole-resistant isolates (from 2 patients). In addition to the L98H substitution, 2 isolates from 2 patients had a 34-bp sequence that was duplicated in the promoter region (16,17) of the cyp51A gene. One isolate had 2 amino acid substitutions, H147Y and G448S. Three isolates from 2 patients had the same 6 mutations, 3 nonsynonymous ones (F46Y, M172V, E427K), along with 3 synonymous (silent) alterations at codons 89, 358, and 454 (data not shown), and an isolate from a third patient had additional mutations (N248T, D255E) as well as these 6. Four novel mutations were found (H147Y, P216L, Y431C, and G434C). The isolate bearing the P216L mutation was resistant to itraconazole and posaconazole, whereas the isolates with Y431C and G434C showed pan-azole resistance phenotypes.

Patient 3 had 2 respiratory samples taken while she was alive, in addition to 18 aspergilloma isolates sampled at autopsy. All isolates were resistant to itraconazole (>8 mg/L), and 1 of 2 different mutations at codon 220 was detected in the cyp51A gene. Isolates with a methionine-to-lysine substitution were highly cross-resistant to voriconazole (4 mg/L) and posaconazole (>8 mg/L), whereas those with an alteration to threonine had variable voriconazole (0.5–4 mg/L) and posaconazole (0.125–1 mg/L) MICs.

Microsatellite Typing

Figure 2

Thumbnail of Unrooted phylogenetic tree showing the genetic relationship of isolates from 7 patients.The genetic relationship of these isolates is shown in relation to each other and to 18 other isolates. AF numbers belong to a collection of &gt;200 isolates, held in Manchester, UK. ATCC, American Type Culture Collection; CBS, Centraalbureau voor Schimmelcultures; FGSC, Fungal Genetics Stock Center. Bootstrap values &gt;90 only are shown. Scale bar indicates nucleotide substitutions per site.

Figure 2. Unrooted phylogenetic tree showing the genetic relationship of isolates from 7 patients.The genetic relationship of these isolates is shown in relation to each other and to 18 other isolates. AF numbers...

The relatedness of isolates obtained from patients 3, 4, 5, 6, 8, 9, and 13 were compared by microsatellite typing (Figure 2). The isolates from 5 patients consisted of a susceptible/resistant pair, whereas an overlapping group of 4 patients had more than 1 cyp51A mutation. All isolates were from the lower respiratory tract, except the resistant isolate from patient 5, which was from a cerebral lesion.

Multiple isolates from 5 of 7 patients had identical or nearly identical genotypes. The isolates from 2 of these 5 patients (3 and 6) differed by 1 and 2 trinucleotide repeat units, respectively, at the most polymorphic locus (3A). Three matched sets (isolates pre- and postdevelopment of resistance) were identified, where resistance almost certainly evolved from an originally susceptible strain.

Figure 2 shows an unrooted tree of the phylogenetic relationships, derived from 5 of the 6 microsatellite markers, for the isolates from these 7 patients plus 18 A. fumigatus isolate controls. Only bootstrap values >90 are shown. Strains from these 7 patients are distributed among other clinical isolates; statistically supported clustering is not evident. Therefore, none of the azole-resistant isolates have been transmitted from patient to patient, indicating that they have all evolved independently from different original strains. The only statistically supported clades contain isolates that only differ from each other by 1 of the 5 markers.

Discussion

Itraconazole resistance and azole cross-resistance in Aspergillus spp. have been reported infrequently, which suggests that they are infrequent events to date. A contributing factor to this low prevalence has been variability in testing between laboratories. Since the initial report of resistance in isolates collected before the licensure of itraconazole, substantial improvements in susceptibility testing methods that allow confidence in reported azole MICs have been implemented. Recommended methods are now promulgated by the CLSI method M38-A2 (27) and EUCAST (21), and work is ongoing to establish internationally agreed interpretative cutoffs (22) and clinical breakpoints (23).

By using such methods, some researchers have documented and published the frequency of itraconazole resistance in clinical A. fumigatus isolates (8,2832); frequency ranged between 2% and 6%. However, most of these studies included fewer isolates (<200) than our study (519) and covered the pre-2004 era. The frequency of itraconazole resistance in our collection before 2004 was 1%; since 2004, however, it has been remarkably high at 8%. The high frequency probably reflects, at least in part, the specialized referral base for patients with chronic and allergic aspergillosis at our center, although there has been no material change in catchment area in the past decade. Referral numbers are rising, however, and susceptibility testing of isolates of patients receiving therapy has been more frequent since 2003.

Another remarkable aspect of this study is the diversity of cyp51A mutations. Both previously published and novel alterations were identified in our resistant isolates (Table 3). In contrast, a recent series of 32 itraconazole-resistant isolates from the Netherlands was published; 94% had the same 2 alterations: an L98H-aa substitution in Cyp51A, in combination with a duplication in the promoter region (32). This combination of mutations was found in 2 of our isolates from 2 patients.

Several authors have identified hot-spot regions associated with resistance in clinical isolates at codons 54 (6,1013,22), 98 (1618,22,33), and 220 (6,14,15,22,32) in the cyp51A gene. We previously reported an alteration at codon 138 (G138C) in multiple isolates from 1 patient (7). A single clinical isolate with a mutation at codon 448 (G to S) has also been previously reported (34). In addition, G138R and G448S mutants have been generated in the laboratory and were azole resistant (35). Mutations in codons 46, 172, 248, 255, and 427 have been found in azole-susceptible strains by us (A. Albarrag, unpub. data) and others (22) and so are not associated with resistance. The resistant isolates with these mutations must therefore have another resistance mechanism. Four novel cyp51A mutations, 3 of which were unassociated with any other mutations (in codons 147, 216, 431, and 434), were identified in this series, although their association with resistance remains to be confirmed experimentally. The H147Y substitution is probably unimportant for resistance because it was found with G448S in 1 isolate and the cross-resistance profile of this isolate was identical to an isolate that had only G448S. We did not find any examples of previously reported mutations at codons 297 and 495 (17,32) or 22, 394, 491, and 440 (14) in our collection. Three of our resistant isolates had no mutations in their cyp51A gene, indicating the presence of other resistance mechanisms.

The position and type of amino acid substitution within the Cyp51A protein determines the pattern of azole cross-resistance (Table 3), which is consistent with predicted structural properties of the demethylase enzyme and its interaction with chemically different azole drugs (36). Resistance to itraconazole is usually associated with a reduction in posaconazole susceptibility, predictably because the 2 drugs are structurally similar; they have variably elevated posaconazole MICs compared with wild type isolates (22,30). Many of the isolates reported here reflect this MIC shift. Isolates with alterations at codons 98 (including the duplication in the promoter region), 138, 431, and 434 demonstrated cross-resistance to voriconazole and posaconazole. All isolates with substitutions at codons 54 and 216 remained susceptible to voriconazole. Some isolates in this study showed cross-resistance between itraconazole and voriconazole and not posaconazole, unlike the results in previous reports (22,32). However, this cross-resistance could be because of differing breakpoints; therefore, determination of an internationally agreed cutoff for posaconazole will be necessary to guide clinicians. No difference in amphotericin B MICs was seen in our azole-resistant isolates compared with susceptible ones, although the clinical utility of MIC testing of amphotericin B in Aspergillus spp. is suboptimal.

More than 1 azole-resistant A. fumigatus isolate was obtained from 6 of the 17 patients described. Microsatellite typing demonstrated that the isolates from each patient had evolved from a single original strain, because they were either identical at 6 markers or differed only in the most polymorphic marker. The resistant isolates from 4 patients had different cyp51A mutations. Given that a single colony is picked from the primary isolation plate and referred for susceptibility testing, additional mutants may have been found had multiple colonies been tested. Within this dataset, the chance of 2 isolates being identical by chance alone within a recombining population are infinitesimal given the high allelic variability that we observed (mean number of alleles per locus = 14; p recovering the same multilocus genotype twice ≈145). The existence of susceptible and resistant isolates that are genetically identical from 3 patients, and the phylogeny performed on the multiple-resistant isolates from an additional 4 patients, almost certainly indicates that the evolution of azole resistance has occurred in these patients independently and repeatedly from unrelated strains. The presence of genetically identical isolates with different cyp51A codon mutations in 3 patients (and 1 almost identical) suggests that they must have evolved independently from the same original strain, because the resistance mutations are not being accumulated sequentially as has been shown to happen in Candida albicans (37). The isolates from 2 patients had differing numbers of repeats of microsatellite marker 3A, which is further proof that strains are evolving in the lung. In contrast, Snelders et al. (32) suggested that many of their patients were infected with a primary resistant strain from the environment.

The referral base for these isolates includes a specialized clinical service for the management of aspergillosis. Many of our resistant isolates came from this group, in particular from 9 patients with chronic cavitary pulmonary aspergillosis with >1 aspergillomas, which may explain the high frequency of resistance in our center. Because surgery is not an option for most patients with chronic cavitary pulmonary aspergillosis, these patients usually require long-term (if not lifelong) antifungal therapy, under which conditions as we have shown, strains of A. fumigatus may evolve resistance. Another contributory explanation could be our systematic application of susceptibility testing of Aspergillus spp. isolates in all cases in which treatment is to be given.

In 6 of 10 patients, steady state itraconazole plasma level data were at or above minimum therapeutic levels (i.e., <5 mg/L), as determined by bioassay (38,39). Low plasma levels of itraconazole were attributable to limited bioavailability in some patients, low doses (i.e., 200 mg daily, the standard UK registered dose), drug interactions in patients with concomitant atypical mycobacterial infection, and use of generic itraconazole (40). Low plasma levels of itraconazole, in combination with the high proportion of patients in this study with prior azole exposure (13 out of 14), indicates that resistance primarily emerged during or after azole therapy.

Our observations are of concern on several fronts. We found a sudden rise in the frequency of azole resistance in A. fumigatus since 2004, and many isolates showed cross-resistance between all the currently licensed azole options. Clinical data indicate that resistance has occurred during and after azole therapy in all but 1 of these cases. The infections caused by azole-resistant isolates fail therapy or at best do not respond. The molecular epidemiology shows that resistance evolved in infecting strains within the lung, rather than by superinfection with a resistant stain from the environment. Because azoles are the only useful class of oral drugs for aspergillosis (and many other serious filamentous fungal infections), clinical management of these chronically infected cases is therefore problematic. Vigilance is called for to identify azole-resistant aspergilli, and novel classes of oral antifungal would be welcome for those infected with azole-resistant strains.

Ms Howard is a senior clinical scientist at the RMLM and a research associate at The University of Manchester. Her research interests focus on resistance in Aspergillus spp. and in vitro and in vivo models of invasive aspergillosis.

Top

Acknowledgments

We thank the University of Manchester for processing the sequencing samples, and The University of Warwick for performing the capillary electrophoresis. We also thank Marianne Skov for providing information on the Danish patient and isolate. Grateful thanks also go to Steve Park and Rebecca Gardiner for performing the sequencing of some isolates.

This work was funded by the hospital infectious diseases endowment fund, a scholarship from the Saudi Arabian Ministry of Education to A.A., and a travel scholarship from Schering Plough to D.C.

Top

References

  1. Denning  DW. Invasive aspergillosis. Clin Infect Dis. 1998;26:781803. DOIPubMedGoogle Scholar
  2. Walsh  TJ, Anaissie  EJ, Denning  DW, Herbrecht  R, Kontoyiannis  DP, Marr  KA, Treatment of aspergillosis: clinical practice guidelines of the Infectious Diseases Society of America. Clin Infect Dis. 2008;46:32760. DOIPubMedGoogle Scholar
  3. Maschmeyer  G, Haas  A, Cornely  OA. Invasive aspergillosis: epidemiology, diagnosis and management in immunocompromised patients. Drugs. 2007;67:1567601. DOIPubMedGoogle Scholar
  4. Denning  DW, Venkateswarlu  K, Oakley  KL, Anderson  MJ, Manning  NJ, Stevens  DA, Itraconazole resistance in Aspergillus fumigatus. Antimicrob Agents Chemother. 1997;41:13648.PubMedGoogle Scholar
  5. Dannaoui  E, Borel  E, Monier  MF, Piens  MA, Picot  S, Persat  F. Acquired itraconazole resistance in Aspergillus fumigatus. J Antimicrob Chemother. 2001;47:33340. DOIPubMedGoogle Scholar
  6. Chen  J, Li  H, Li  R, Bu  D, Wan  Z. Mutations in the cyp51A gene and susceptibility to itraconazole in Aspergillus fumigatus serially isolated from a patient with lung aspergilloma. J Antimicrob Chemother. 2004;55:317. DOIPubMedGoogle Scholar
  7. Howard  SJ, Webster  I, Moore  CB, Gardiner  RE, Park  S, Perlin  DS, Multi-azole resistance in Aspergillus fumigatus. Int J Antimicrob Agents. 2006;28:4503. DOIPubMedGoogle Scholar
  8. Chryssanthou  E. In vitro susceptibility of respiratory isolates of Aspergillus species to itraconazole and amphotericin B acquired resistance to itraconazole. Scand J Infect Dis. 1997;29:50912.PubMedGoogle Scholar
  9. Warris  A, Weemaes  CM, Verweij  PE. Multidrug resistance in Aspergillus fumigatus. N Engl J Med. 2002;347:2173a4. DOIGoogle Scholar
  10. Diaz-Guerra  TM, Mellado  E, Cuenca-Estrella  M, Rodriguez-Tudela  JL. A point mutation in the 14alpha-sterol demethylase gene cyp51A contributes to itraconazole resistance in Aspergillus fumigatus. Antimicrob Agents Chemother. 2003;47:11204. DOIPubMedGoogle Scholar
  11. Nascimento  AM, Goldman  GH, Park  S, Marras  SA, Delmas  G, Oza  U, Multiple resistance mechanisms among Aspergillus fumigatus mutants with high-level resistance to itraconazole. Antimicrob Agents Chemother. 2003;47:171926. DOIPubMedGoogle Scholar
  12. Balashov  SV, Gardiner  R, Park  S, Perlin  DS. Rapid, high-throughput, multiplex, real-time PCR for identification of mutations in the cyp51A gene of Aspergillus fumigatus that confer resistance to itraconazole. J Clin Microbiol. 2005;43:21422. DOIPubMedGoogle Scholar
  13. Mann  PA, Parmegiani  RM, Wei  SQ, Mendrick  CA, Li  X, Loebenberg  D, Mutations in Aspergillus fumigatus resulting in reduced susceptibility to posaconazole appear to be restricted to a single amino acid in the cytochrome P450 14alpha-demethylase. Antimicrob Agents Chemother. 2003;47:57781. DOIPubMedGoogle Scholar
  14. da Silva Ferreira  ME, Capellaro  JL, dos Reis Marques  E, Malavazi  I, Perlin  D, Park  S, In vitro evolution of itraconazole resistance in Aspergillus fumigatus involves multiple mechanisms of resistance. Antimicrob Agents Chemother. 2004;48:440513. DOIPubMedGoogle Scholar
  15. Mellado  E, Garcia-Effron  G, Alcazar-Fuoli  L, Cuenca-Estrella  M, Rodriguez-Tudela  JL. Substitutions at methionine 220 in the 14alpha-sterol demethylase (Cyp51A) of Aspergillus fumigatus are responsible for resistance in vitro to azole antifungal drugs. Antimicrob Agents Chemother. 2004;48:274750. DOIPubMedGoogle Scholar
  16. Mellado  E, Alcazar-Fuoli  L, Garcia-Effron  G, Alastruey-Izquierdo  A, Cuenca-Estrella  M, Rodriguez-Tudela  JL. New resistance mechanisms to azole drugs in Aspergillus fumigatus and emergence of antifungal drugs−resistant A. fumigatus atypical strains. Med Mycol. 2006;44(Suppl):36771. DOIGoogle Scholar
  17. Mellado  E, Garcia-Effron  G, Alcazar-Fuoli  L, Melchers  WJ, Verweij  PE, Cuenca-Estrella  M, A new Aspergillus fumigatus resistance mechanism conferring in vitro cross-resistance to azole antifungals involves a combination of cyp51A alterations. Antimicrob Agents Chemother. 2007;51:1897904. DOIPubMedGoogle Scholar
  18. Verweij  PE, Mellado  E, Melchers  WJ. Multiple-triazole-resistant aspergillosis. N Engl J Med. 2007;356:14813. DOIPubMedGoogle Scholar
  19. Moore  CB, Sayers  N, Mosquera  J, Slaven  J, Denning  DW. Antifungal drug resistance in Aspergillus. J Infect. 2000;41:20320. DOIPubMedGoogle Scholar
  20. Mowat  E, Warn  P, Denning  D, Fitzpatrick  W, Patterson  T, Jones  B, The potential role of Afumdr4 in azole resistance during Aspergillus fumigatus multicellular growth. Presented at: 48th Interscience Conference on Antimicrobial Agents and Chemotherapy; 2008 Oct 25–28; Washington, DC. Abstract M–2190.
  21. Subcommittee on Antifungal Susceptibility Testing (AFST) of the ESCMID European Committee on Antimicrobial Susceptibility Testing EUCAST. EUCAST technical note on method for the determination of broth dilution minimum inhibitory concentrations of antifungal agents for conidia-forming moulds. Clin Microbiol Infect. 2008;14:9824. DOIPubMedGoogle Scholar
  22. Rodriguez-Tudela  JL, Alcazar-Fuoli  L, Mellado  E, Alastruey-Izquierdo  A, Monzon  A, Cuenca-Estrella  M. Epidemiological cutoffs and cross-resistance to azole drugs in Aspergillus fumigatus. Antimicrob Agents Chemother. 2008;52:246872. DOIPubMedGoogle Scholar
  23. Verweij  PE, Howard  SJ, Melchers  WJG, Denning  DW. Azole resistance in Aspergillus: proposed nomenclature and breakpoints. Drug Resist Rev. In press.
  24. de Valk  HA, Meis  JF, Curfs  IM, Muehlethaler  K, Mouton  JW, Klaassen  CH. Use of a novel panel of nine short tandem repeats for exact and high-resolution fingerprinting of Aspergillus fumigatus isolates. J Clin Microbiol. 2005;43:411220. DOIPubMedGoogle Scholar
  25. Pasqualotto  AC, Denning  DW, Anderson  MJ. A cautionary tale: lack of consistency in allele sizes between two laboratories for a published multilocus microsatellite typing system. J Clin Microbiol. 2007;45:5228. DOIPubMedGoogle Scholar
  26. Bowcock  AM, Ruiz-Linares  A, Tomfohrde  J, Minch  E, Kidd  JR, Cavalli-Sforza  LL. High resolution of human evolutionary trees with polymorphic microsatellites. Nature. 1994;368:4557. DOIPubMedGoogle Scholar
  27. National Committee for Clinical Laboratory Standards. Reference method for broth dilution antifungal susceptibility testing of filamentous fungi; approved standard, 2nd ed. Document M38–A2. 2002.
  28. Gomez-Lopez  A, Garcia-Effron  G, Mellado  E, Monzon  A, Rodriguez-Tudela  JL, Cuenca-Estrella  M. In vitro activities of three licensed antifungal agents against Spanish clinical isolates of Aspergillus spp. Antimicrob Agents Chemother. 2003;47:30858. DOIPubMedGoogle Scholar
  29. Dannaoui  E, Persat  F, Monier  MF, Borel  E, Piens  MA, Picot  S. In vitro susceptibility of Aspergillus spp. isolates to amphotericin B and itraconazole. J Antimicrob Chemother. 1999;44:5535. DOIPubMedGoogle Scholar
  30. Pfaller  MA, Messer  SA, Boyken  L, Rice  C, Tendolkar  S, Hollis  RJ, In vitro survey of triazole cross-resistance among more than 700 clinical isolates of Aspergillus species. J Clin Microbiol. 2008;46:256872. DOIPubMedGoogle Scholar
  31. Verweij  PE, Te Dorsthorst  DT, Rijs  AJ, De Vries-Hospers  HG, Meis  JF. Nationwide survey of in vitro activities of itraconazole and voriconazole against clinical Aspergillus fumigatus isolates cultured between 1945 and 1998. J Clin Microbiol. 2002;40:264850. DOIPubMedGoogle Scholar
  32. Snelders  E, van der Lee  HA, Kuijpers  J, Rijs  AJ, Varga  J, Samson  RA, Emergence of azole resistance in Aspergillus fumigatus and spread of a single resistance mechanism. PLoS Med. 2008;5:e219. DOIPubMedGoogle Scholar
  33. van der Linden  JW, Jansen  RR, Bresters  D, Visser  CE, Geerlings  SE, Kuijper  EJ, Azole-resistant central nervous system aspergillosis. Clin Infect Dis. 2009;48:11113. DOIPubMedGoogle Scholar
  34. Manavathu  EK, Espinel-Ingroff  A, Alangaden  G, Chandrasekar  PH. Molecular studies on voriconazole-resistance in a clinical isolate of Aspergillus fumigatus. In: Proceedings of 43rd Interscience Conference on Antimicrobial Agents and Chemotherapy; 2003 Sep 14–17; Chicago, IL, USA. Abstract M–392.
  35. Manavathu  EK, Baskaran  I, Krishnan  S, Alangaden  G, Chandrasekar  PH. Cytochrome P450 14-alpha-sterol demethylase mutation dependent triazole cross-resistance in Aspergillus fumigatus. In: Proceedings of 43rd Interscience Conference on Antimicrobial Agents and Chemotherapy; Chicago, IL, USA; 2003 Sep 14–17. Abstract M–471.
  36. Xiao  L, Madison  V, Chau  AS, Loebenberg  D, Palermo  RE, McNicholas  PM. Three-dimensional models of wild-type and mutated forms of cytochrome P450 14alpha-sterol demethylases from Aspergillus fumigatus and Candida albicans provide insights into posaconazole binding. Antimicrob Agents Chemother. 2004;48:56874. DOIPubMedGoogle Scholar
  37. Coste  A, Selmecki  A, Forche  A, Diogo  D, Bougnoux  ME, d'Enfert  C, Genotypic evolution of azole resistance mechanisms in sequential Candida albicans isolates. Eukaryot Cell. 2007;6:1889904. DOIPubMedGoogle Scholar
  38. Tucker  RM, Williams  PL, Arathoon  EG, Stevens  DA. Treatment of mycoses with itraconazole. Ann N Y Acad Sci. 1988;544:45170. DOIPubMedGoogle Scholar
  39. Law  D, Moore  CB, Denning  DW. Bioassay for serum itraconazole concentrations using hydroxyitraconazole standards. Antimicrob Agents Chemother. 1994;38:15616.PubMedGoogle Scholar
  40. Pasqualotto  AC, Denning  DW. Generic substitution of itraconazole resulting in sub-therapeutic levels and resistance. Int J Antimicrob Agents. 2007;30:934. DOIPubMedGoogle Scholar

Top

Figures
Tables

Top

Cite This Article

DOI: 10.3201/eid1507.090043

1These data were presented in part at the 2nd Advances Against Aspergillosis meeting, February 22–26, 2006, Athens, Greece; and 46th Interscience Conference on Antimicrobial Agents and Chemotherapy, September 27–30, 2006, San Francisco, CA, USA.

Table of Contents – Volume 15, Number 7—July 2009

EID Search Options
presentation_01 Advanced Article Search – Search articles by author and/or keyword.
presentation_01 Articles by Country Search – Search articles by the topic country.
presentation_01 Article Type Search – Search articles by article type and issue.

Top

Comments

Please use the form below to submit correspondence to the authors or contact them at the following address:

David W. Denning, 2nd Floor Education and Research Centre, University Hospital of South Manchester, Southmoor Road, Manchester M23 9LT, UK

Send To

10000 character(s) remaining.

Top

Page created: November 09, 2010
Page updated: November 09, 2010
Page reviewed: November 09, 2010
The conclusions, findings, and opinions expressed by authors contributing to this journal do not necessarily reflect the official position of the U.S. Department of Health and Human Services, the Public Health Service, the Centers for Disease Control and Prevention, or the authors' affiliated institutions. Use of trade names is for identification only and does not imply endorsement by any of the groups named above.
file_external